Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Journal of The Electrochemical

Society

OPEN ACCESS You may also like


- Application of quantitative XRD on the
Bench-Scale Demonstration and Thermodynamic precipitation of struvite from Brine Water
E Heraldy, F Rahmawati, Heryanto et al.
Simulations of Electrochemical Nutrient Reduction - The Effect of Mixing Rate on Struvite
Recovery from The Fertilizer Industry
in Wastewater via Recovery as Struvite Warmadewanthi, A Rodlia, N Ikhlas et al.

- Phosphorus recovery from septage


To cite this article: Zineb Belarbi et al 2020 J. Electrochem. Soc. 167 155524 treatment plant sludge by struvite
formation with alkaline hydrolysis as pre-
treatment
C K L Nochefranca, Z K C Oyson, A B San
Pedro et al.

View the article online for updates and enhancements.

This content was downloaded from IP address 197.3.11.36 on 08/03/2024 at 16:21


Journal of The Electrochemical Society, 2020 167 155524

Bench-Scale Demonstration and Thermodynamic Simulations of


Electrochemical Nutrient Reduction in Wastewater via Recovery
as Struvite
Zineb Belarbi,1,=,z Damilola A. Daramola,1,2,=,*,z and Jason P. Trembly1,2,3
1
Institute for Sustainable Energy and the Environment, Department of Chemical and Biomolecular Engineering, Ohio
University, Athens Ohio 45701, United States of America
2
Department of Chemical and Biomolecular Engineering, Ohio University, Athens Ohio 45701, United States of America
3
Department of Mechanical Engineering, Ohio University, Athens Ohio 45701, United States of America

Land application of manure can be a sustainable supply chain practice that improves soil quality by recycling important nutrients
contained in animal waste. Yet, runoff of phosphorus and nitrogen nutrients contained in the animal waste has contributed to
significant watershed eutrophication. Recovery of the dissolved nutrient species as a condensed solid fertilizer product would increase
sustainability of the agricultural supply chain, while reducing watershed pollution. This study was conducted to evaluate the recovery
of phosphorus (primarily) as struvite using an electrochemical process while varying temperature, applied cathodic potential,
turbulence and Ca2+ concentration. High phosphorus recovery with high current efficiency and low specific energy consumption was
possible at 20 °C, −1.1 V vs Ag/AgCl at the cathode, and a Reynolds number of 9150 in the absence of Ca2+ when the Mg:N:P ratio
was 1.37:1:1. Further, a thermodynamic model of the waste solution indicated an increase in Ca2+ concentration, which impedes
struvite recovery, can be negated by increasing dissolved Mg2+ concentration and operating at a pH below NH3 volatilization.
© 2020 The Author(s). Published on behalf of The Electrochemical Society by IOP Publishing Limited. This is an open access
article distributed under the terms of the Creative Commons Attribution 4.0 License (CC BY, http://creativecommons.org/licenses/
by/4.0/), which permits unrestricted reuse of the work in any medium, provided the original work is properly cited. [DOI: 10.1149/
1945-7111/abc58f]

Manuscript submitted July 8, 2020; revised manuscript received September 17, 2020. Published November 6, 2020. This paper is
part of the JES Focus Issue on Organic and Inorganic Molecular Electrochemistry.
Supplementary material for this article is available online

Excess phosphorus, due to anthropogenic activity, is one of the cause phosphate formation within the solution, inducing struvite
driving forces for harmful algae bloom formation and eutrophication of precipitation due to aqueous phase magnesium and ammonium
inland and coastal waters in the U.S. and worldwide.1 Associated contained in the animal waste.
anthropogenic activity includes agriculture, treated wastewater dis-
charge and stormwater runoff, resulting in nutrient pollution impacting 1
Cathodic Reaction O 2 (g) + H2 O + 2e-  2OH- [1]
human health, food production, tourism and recreation coupled with 2
adverse economic effects in these sectors.2 Agricultural activity is
especially impactful due to the growth of Concentrated Animal Feeding Solution Reactions H3PO4 + xOH-  H3 - xPO x4 - +xH2 O [2]
Operations (CAFOs), which directly correlates to excess nutrient
availability due to manure application for increased crop yield.3 A
variety of techniques have been considered for phosphorus removal and Struvite Deposition Mg 2 + + NH + 3-
4 + PO 4 + 6H2 O
[3]
some of these techniques have been deployed at scale in centralized  MgNH4 PO4 · 6H2 O
wastewater treatment facilities.4 However, agricultural sources are
typically geographically distributed and thus require technologies that Furthermore, an applied potential provides an additional driving
are feasible without economies of scale. Furthermore, P is a key force for electromigration, potentially improving transport condi-
macronutrient and finite resource and therefore P discharge control tions for struvite recovery. This electrochemical approach could
should combine reduction with recovery for re-use. reduce the costs and improve product quality, which are typical
These discharge control conditions can be met if P is recovered as issues associated with chemical-based pH control.19,20 However,
struvite—MgNH4PO4·6H2O—as aqueous phase nitrogen can be simul- some of the previous electrochemical studies did not consider the
taneously captured forming the slow-release fertilizer product.5,6 A effect of pH conditions and competing ions e.g. calcium, on struvite
modular approach that includes N and P removal as struvite provides availability due to more favorable equilibrium thermodynamics for
multiple advantages for effective nutrient control and agricultural other phosphatic species.21–24
supply chain sustainability. Electrochemical approaches when com- A prior physicochemical study from this group7 used a rotating
bined with distributed low-cost renewable power sources, offer a cylinder electrode (RCE) configuration to evaluate parameters for P
potentially attractive solution for addressing distributed nutrient pollu- recovery as struvite: turbulent flow, applied voltage, temperature and
tion. A variety of electrochemical reactor configurations have been competing ion concentration, and provided key parameters to evaluate
considered in literature varying from purely physicochemical7–13 to in a bench-scale demonstration. Specifically, the prior study showed
bio-based14–16 electrochemical techniques. These studies differ from that higher hydrodynamic flow affected O2 transport, with increased
the capacitive deionization configuration,17,18 where phosphorus reduc- struvite precipitation at higher rotational speeds. In addition, tempera-
tion occurs without solid recovery. In summary, electrochemical tures above 20 °C did not appear to influence struvite recovery. This
techniques provide localized pH control through oxygen reduction at approach differed from prior and recent literature studies that ignored
the cathode (Eq. 1) resulting in ion speciation conditions (Eq. 2), where struvite formation9,11 due to the low concentration of nitrogen and the
x could be 1, 2 or 3 depending on initial state of phosphate species) that absence of magnesium species. In addition, the variability of ionic
favor the recombination of Mg2+, PO43− and NH4+ ionic species species due to the waste source requires an evaluation of solubility
(Eq. 3). In brief, oxygen reduction provides the pH control necessary to thermodynamics, which will govern operating conditions required to
control non-struvite solid phase formations.
= Within this context, this study was performed to quantify the
These authors contributed equally to this work.
*Electrochemical Society Member. efficiency and energetics through electrochemically induced P
z
E-mail: belarbi_zineb@hotmail.fr; daramola@ohio.edu recovery as struvite from synthetic animal wastewater. Stainless
Journal of The Electrochemical Society, 2020 167 155524

steel was selected as the working electrode material of choice in this


study (in comparison to engineered electrocatalytic materials), which
if successful could provide an efficiently scaled low-cost modular
approach to simultaneously address watershed eutrophication and
agriculture supply chain sustainability. The effect of temperature,
applied potential, turbulence, and competing divalent ion (Ca2+)
concentration on the system’s responses were evaluated under
similar mass transfer conditions exhibited by the RCE study.7
Further, these findings were coupled with a thermodynamic simula-
tion correlating ionic speciation with expected thermodynamically-
favorable conditions, thus providing insight into reduced struvite
recovery in the presence of calcium ions.

Experimental
Bench-scale demonstration.— Materials and setup.—Experiments
were performed with an aqueous solution (250 ml) containing 7.84 ×
10−3 M of phosphate and ammonium as NH4H2PO4 and 10.71 ×
10−3 M of magnesium as MgCl2·6H2O, with the solutions prepared
using the same procedure described in a prior RCE configuration.7
Experiments were carried out using a jacketed electrochemical cell
with a three-electrode set up (Fig. 1).
The working electrode was a stainless-steel flat electrode with an
area of 29.39 cm2, the counter electrode was carbon graphite with a
large surface area, while the reference electrode was saturated Ag/
Figure 1. Typical configuration of a jacketed stirred tank reactor for the
AgCl. Before each experiment, the working electrode surface was electrochemical recovery of phosphorus used in this study. The working
polished with silicon carbide paper (240, 400 and 600 grit), electrode was stainless steel.
sonicated in isopropanol for 5 min to remove any residual deposits
and finally air-dried before immersion in the cell. Tests were in Fig. 1, used a flat Ni working electrode (area = 5.42 cm2). In
performed under potentiostatic conditions at the cathode for a period both configurations, linear voltammetry was conducted from 0.2 V vs
of 100 min (80 min when varying temperature), while varying Ag/AgCl to −1.3 V vs Ag/AgCl at a scan rate of 5 mV s−1. Flow was
Reynolds number, applied cathodic potential, temperature and controlled between 50 and 1000 rpm at 20 °C and 35 °C.
Ca2+ ion concentration (Table I). These potentiostatic experiments After the current vs potential measurements, the average of the
were conducted using a Gamry instrument and the bulk pH was current density at the plateau portion of the cathodic curves (the
checked at the beginning and end of each experiment (Table SI in limiting current) was used to calculate the mass transfer coefficient
Supplementary Information is available online at stacks.iop.org/JES/ using Eq. 5, which simplifies to Eq. 6 due to the excess NaOH
167/155524/mmedia). The current measured at each condition was significantly decreasing the ferri-ferrocyanide transference numbers.25
reported as residual current by taking the ratio between the
instantaneous current (I) and initial current (I0) measured during Ilim
(1 - ti ) = Cb k m [5]
the experiments. nF
In order to conduct the electrochemical tests using the setup
shown in Fig. 1, the mass transport characteristics of the system Ilim
were initially characterized in order to ensure that transport observed » Cb k m [6]
in the impeller configuration did not significantly differ from nF
observations in the RCE configuration.7 Where, Ilim is the limiting current in A m−2, n is the number of
moles of electrons transferred between reductant and oxidant in mol
Mass transport characterization.—This characterization was electrons mol−1 (1 as shown in Eq. 4), F is Faraday’s constant in
conducted on the ferri-ferrocyanide coupled reaction (Eq. 4) to C mol−1 electrons, Cb is the concentration of the active species in
obtain a mass transfer coefficient that was characteristic of both the mol m−3 and km is the mass transfer coefficient in m s−1.
RCE configuration and the impeller configuration in Fig. 1. The mass transfer coefficient and the geometry of the system can
be used to calculate relevant dimensionless numbers (Sherwood,
[Fe (CN)6 ]3 - + e-  [Fe (CN)6 ]4 - [4]
Reynolds and Schmidt) that characterize the flow hydrodynamics of
both the RCE and impeller configurations using Eqs. 7–9 below
This set of coupled reactions are typically used to study mass
transfer for various types of geometries and hydrodynamics due to km L
several advantages including good reproducibility, rapid achieve- Sh = [7]
ment of steady state and high accuracy.25 Deoxygenated equimolar D
aqueous solutions of 0.01 M K3(Fe(CN)6 and K4Fe(CN)6·6H2O in
0.5 M of NaOH was used as the electrolyte. A large excess of uL
Re = [8]
sodium hydroxide eliminates the contribution of ionic migration to n
mass transfer and the solutions were kept in the dark to mitigate slow
decomposition in light.25 n
The RCE configuration for this mass characterization is a described Sc = [9]
D
in prior literature, with a cylindrical nickel RCE electrode (12.68 mm
diameter and 12.27 mm height), a saturated Ag/AgCl reference Where, Sh is the Sherwood number, km is the mass transfer
electrode and carbon graphite counter electrode in a three-electrode coefficient in m s−1, L is the characteristic length in m, D is the
electrochemical cell setup.7 The impeller configuration (34.56 mm temperature-adjusted diffusion coefficient in m2 s−1, Re is the
impeller diameter) for this mass transport characterization, as illustrated Reynolds number, u is the characteristic velocity in m s−1, ν is
Journal of The Electrochemical Society, 2020 167 155524

Table I. List of independent variables and values evaluated in this study for phosphorus recovery of struvite using the setup in Fig. 1.

Stimulus Valuesc)

Applied Cathodic Potential (V vs (Ag/AgCl)) −0.8, −1.1, −1.3 and −1.5


Reynolds numbera) 0, 3660, 9150 and 12811
Temperature (°C) 5, 20 and 35
Ca2+ concentration (mg l−1)b) 0, 72 and 210

a) These values correspond to impeller angular speeds of 0, 200, 500 and 700 rpm at 20 °C. b) Calcium is added as CaCl2•2H2O. c) The baseline conditions
were −1.1 V vs Ag/AgCl, Reynolds number of 9150, 20 °C and 0 mg l−1 of Ca2+.

the temperature-adjusted kinematic viscosity in m2 s−1 and Sc is the


Schmidt number. Temperature adjusted values for both D and ν were
Es =
ò E (t ) I (t ) dt [12]
obtained from literature,26 while the characteristic length and
V (C0 - C (t ))
velocity for both the RCE and impeller are the diameter and
peripheral velocity, respectively.27 Where, C0 is the initial phosphorus concentration in g l−1 from ICP-
OES, C(t) is the phosphorus concentration at time t in g l−1 from
Chemical analyses.—Liquid samples were tested for elemental ICP-OES, z is the number of electrons transferred (2 as described
compositions (Ca, Mg and P) using Inductively Coupled Plasma below with Eq. 13), F is Faraday’s constant in C mol−1 electrons, V
Optical Emission Spectrometry (ICP-OES). The amount of phos- is the electrolyte volume within the reactor in L, M is the atomic
phorus removed by the electrochemical process was measured by weight of phosphorus (31 g mol−1), I(t) is the total current at time t
monitoring the respective phosphorus concentrations in solution in A and E(t) is the cell voltage at time t in V.
before and after each experiment. These values were used to
characterize the electrolyzer based on figures of merit. 2 mol e- 96485 C 2 mol OH- 1 mol PO43 -
´ ´ ´
Figures of merit.—The performance of this electrochemical setup 2 mol OH- 1 mol e- 1 mol PO43 - 1 mol P
for phosphorus recovery was characterized based on the phosphorus 1 mol P ⎛⎜ gP gP ⎞
´ (0) - (t ) ⎟ ´ L
removal efficiency in % (Eq. 10), the current efficiency for 31 g P ⎝ L L ⎠
phosphorus recovery in % (Eq. 11) and the specific energy
= charge for solid P removal [13]
consumption in kWh kg−1 P (Eq. 12)
C0 - C (t ) As shown in Eq. 13 above, 2 mole electrons are transferred in
hP Removal = ´ 100% [10] Eqs. 1 and 2 moles OH─ stoichiometrically converts 1 mole H2PO4─
C0 to 1 mole PO43− in Eq. 2. This efficiency evaluation assumes all the
hydroxide ions generated by the electrons are used for phosphate
zFV (C0 - C (t )) formation as PO43− and would neglect electrons used in converting
b= [11]
H2PO4─ to other aqueous species (highlighted in Table II). This
ò
M I (t ) dt
neglect will result in an underestimate of the current efficiency.

Table II. Equilibrium expressions and related constants at 25 °C for the solid, ionic and molecular species studied in the thermodynamic
simulations.

Equation Equilibrium Expression pK (25 °C)a)

(3) Mg 2 + + NH4+ + PO43 - + 6H2 O  MgNH4 PO4 · 6H2 O(s) −13.26


(14) H2 PO4- + OH-  HPO42 - + H2 O −8.54
(15) HPO42 - + OH-  PO43 - + H2 O −3.47
(16) NH4+ + OH-  NH3 + H2 O −6.50
(17) Mg 2 + + H2 PO4-  MgH2 PO4( +
aq)
−1.44
(18) + 2-
Mg + HPO4  MgHPO4(aq)
2 −3.11
(19) Mg 2 + + PO43 -  MgPO4( -
aq)
−6.47
(20) 3Mg 2 + + 2PO43 -  Mg3 (PO4 )2 (s) −26.01
(21) Mg 2 + + OH-  MgOH+ (aq) −2.58
(22) Mg 2 + + 2OH-  Mg (OH )2 (s) −11.21
(23) Ca 2 + + H2 PO4-  CaH2 PO4( +
aq)
−1.44
(24) + 2-
Ca + HPO4  CaHPO4(aq)
2 −2.60
(25) Ca 2 + + PO43 -  CaPO4-(aq) −6.47
(26) 3Ca 2 + + 2PO43 -  Ca3 (PO4 )2 (s) −32.69
(27) Ca 2 + + OH-  CaOH+ (aq) −1.23
(28) Ca 2 + + 2OH-  Ca (OH )2 (s) −5.37

a) These values are based on the format of the equilibrium expressions and the references for the constants are as follows: (3) from Ohlinger et al.,28 (14) to
(16) from the Aspen Plus Database29 and (17) to (28) from Fritz.30
Journal of The Electrochemical Society, 2020 167 155524

Figure 2. Aspen Plus process flowsheet for evaluating pH effect on ion speciation based on reactant mix. The Mg Inlet, N Inlet, P Inlet and Ca Inlet streams are
MgCl2, NH4Cl, NaH2PO4, CaCl2 respectively, while the caustic soda stream is NaOH solution. All streams are at 25 °C.

Thermodynamic simulations.—The effect of solution pH on


ionic speciation and solid deposition was modeled in ASPEN Plus29
using the electrolyte nonrandom two-liquid (eNRTL) model31,32 for
activity coefficients. The individual ionic species of interest were
Mg2+, H2PO4−, NH4+, Ca2+ and OH– and these were present as
MgCl2, NH4Cl, NaH2PO4, CaCl2 and NaOH. The multi-ion species
studied are a combination of aqueous molecules, aqueous ions and
solids listed in Table II along with equilibrium constants of
formation (for molecules and ions) or solubility product (for solids)
at 25 °C from literature.
The simulation was conducted based on the flow diagram
provided in Fig. 2, where MgCl2, NH4Cl, NaH2PO4, CaCl2 were
mixed in a tank under isothermal conditions and the mixture was
sent to another isothermal tank that was dosed with NaOH solution.
The concentrations of MgCl2 (1.37 molal), NH4Cl (1 molal),
NaH2PO4 (1 molal) were mixed such that Mg:NH4+:H2PO4− was
similar to the molar ratios in the experimental analyses. In addition,
NaOH concentration was varied from 0.05 molal and 3.00 molal,
while CaCl2 concentration was varied between 0 and 0.68 molal (for
a maximum Ca:Mg ratio of 0.5:1). This variation captures both the
effect of pH and calcium on the driving forces for struvite deposition
and competing reactions described in Table II.
All figures associated with experimental and simulation data have
been plotted using Matplotlib (version 3.1.1).33

Results and Discussion


Mass Transport Characterization.—Voltammograms collected
for both the RCE and impeller at 20 °C in equimolar 0.01 M
K3(Fe(CN)6 and K4Fe(CN)6·6H2O in 0.5 M of NaOH are shown in
Fig. 3.
From Fig. 3 it can be seen the limiting current increased as the
controlled speed increased. It should also be noted that the RCE
activity at 100 rpm is close to the impeller activity at 50 rpm. The Figure 3. Cathodic polarization curves of ferricyanide ion reduction using
variation in the limiting current portion of the voltammograms for the (a) RCE experimental setup described in prior literature7 and (b) impeller
experimental setup described in Fig. 1.
the impeller could be explained by the vibration from a slightly off-
balance impeller blade and possibly due to turbulent vortex of the
created flow. Consequently, the test was repeated to ensure repeat- phosphorus removal. Additional regression analyses (described in
ability of the measurement technique and the experimental appa- the Supplementary Information) yielded an empirical mass transfer
ratus. Additional impeller tests conducted at 35 °C showed higher correlation (29) for the impeller setup that combines the Sherwood,
temperatures resulted in a higher limiting current at each controlled Reynolds and Schmidt numbers. This expression is in the form
speed (Fig. S1 available in Supplementary Information). found in other literature evaluating three dimensional rotating
Using the Levich equation for the RCE and the Sherwood (Eq. 7) electrodes configurations.27,34
and Reynolds numbers (Eq. 8) for the impeller, the relationship
between the flow and mass transport conditions for both configura- Sh = 0.22Re0.568 Sc 0.47 [29]
tions are as described in Fig. 4.
The similarity in slopes suggest the baseline experiments values
are practically similar in both experimental configurations and Correlation between operating variables and phosphorus re-
therefore the performance in the RCE configuration can be replicated moval metrics.— Effect of applied potential.—The current gener-
by the impeller configurations under the flow conditions evaluated. ated, phosphorus removal efficiency, current efficiency and specific
Therefore, the impeller setup in Fig. 1 could be used to evaluate the energy consumption due to the applied potential are provided in
effects of the parameters previously studied: applied voltage, fluid Fig. 5. Cathodic current at the lower applied potentials (−0.8 V vs
turbulence, temperature and Ca2+ concentration, on quantifiable Ag/AgCl) initially decreases then reaches a relatively stable
Journal of The Electrochemical Society, 2020 167 155524

electrode surface coverage by struvite and therefore less surface area


for oxygen reduction and a quicker decrease in residual current. The
steepest decrease in current can be seen with an applied potential of
−1.3 V (vs Ag/AgCl), suggesting this potential had the fastest
kinetics of the range evaluated. Overall, applied potential did not
seem to affect the nature of the crystallized phase that formed on the
electrode surface, which were thin and elongated in nature (Fig. S2
in Supplementary Information). EDS analysis of the electrode
surface shows the presence of phosphorus, magnesium and oxygen,
which are the main struvite components (Fig. S2).
For a given potential, the largest increase in removal efficiency
(steepest slope) occurs in the first 10 min of the experiments
(Fig. 5b). After 30 min, the efficiency approaches a constant value
of 30, 65, 70 and 80% at −0.8 V, −1.5 V, −1.3 V and −1.1 V (all vs
Ag/AgCl), respectively. When the applied potential decreases from
−0.8 to −1.1 V vs Ag/AgCl, the removal efficiency increases;
however, in the potential range from −1.1 V to −1.5 V vs Ag/AgCl,
the removal efficiency decreased. This behavior can be explained by
considering that for potentials more negative than −1.1 V vs
Ag/AgCl, the limiting current is achieved. Further, at potentials
Figure 4. Momentum (Reynolds number) and mass (mass transfer coeffi- lower than −1.3 V vs Ag/AgCl, hydrogen evolution becomes
cient) transport correlations for the impeller (squares) and RCE (triangles) important and the deposited struvite can become detached from
modes at 20 °C. the electrode and re-dissolved. Therefore, the potential of −1.1 V vs
Ag/AgCl could be the threshold potential of phosphorus removal in
asymptotic value at long immersion time, without an I/I0 ap- this reactor configuration. The current efficiency is close to 23% at
proaching zero (Fig. 5a). This is due to available electrode surface the onset of the experiment for −0.8 V and −1.1 V vs Ag/AgCl
area for oxygen reduction as there is less struvite precipitation on the (Fig. 5c) but decreases with time since the phosphorus concentration
electrode surface. This can be seen in the SEM observations of the in solution decreases. As such, oxygen reduction predominates as the
precipitated struvite (Fig. S2 in Supplementary Information) at this cathodic reaction. For more negative potentials, hydrogen evolution
applied potential. As the cathodic potential increases, there is greater is more important as a cathodic side reaction and lower current

Figure 5. Effect of applied potential (vs Ag/AgCl) on (a) residual current, (b) phosphorus removal efficiency, (c) current efficiency and (d) specific energy
consumption during phosphorus removal. Initial conditions: 250 ml of 7.84 × 10−3 M of NH4H2PO4 and 10.71 × 10−3 M of MgCl2·6H2O neutralized to a pH of
∼7.5, Reynolds number of 9150, 20 °C and 0 mg l−1 of Ca2+.
Journal of The Electrochemical Society, 2020 167 155524

Figure 6. Effect of Reynolds number on (a) residual current, (b) phosphorus removal efficiency, (c) current efficiency and (d) specific energy consumption
during phosphorus removal. Initial conditions: 250 ml of 7.84 × 10−3 M of NH4H2PO4 and 10.71 × 10−3 M of MgCl2·6H2O neutralized to a pH of ∼7.3,
−1.1 V (vs Ag/AgCl), 20 °C and 0 mg l−1 of Ca2+.

efficiency is observed due to reduced struvite deposition due to gas smaller and more numerous crystals with the same morphology as
formation on the electrode surface. The specific energy consumption struvite. These smaller crystals could be due to the increasing shear
as a function of time for different cathodic potentials for stainless stress at the electrode surface causing larger particles to shear off
steel electrode is shown in Fig. 5d. The increase of Es when the after formation or could be due to a higher number of nucleation sites
potential becomes more negative can be attributed to decrease in the on the electrode due to the faster transport of ionic species for
current efficiency due to the onset of hydrogen evolution. Minimum deposition. Figure 6b shows phosphorus removal efficiency as a
energy consumption was observed at −1.1 V that corresponds to the function of time with turbulence, with the greatest rate of P removal
maximum phosphorus removal. Therefore, the potential of −1.1 V in the first 10 min. Furthermore, this initial rate was directly
vs Ag/AgCl is the potential threshold of phosphorus removal in this proportional to the Reynolds number up to 9150. After 20 min, the
reactor configuration. efficiency approached a constant value of 30%, 35%, 80% and 50%
at Reynolds numbers of 0, 3360, 9150 and 12811, respectively, with
Effect of fluid flow.—The current generated, phosphorus removal the optimum removal at 500 rpm (Reynolds number of 9150). This
efficiency, current efficiency and specific energy consumption with result shows that hydrodynamic conditions are favorable for phos-
respect to electrolyte turbulence are provided in Fig. 6. An increase phorus removal as the rate of struvite electrodeposition is governed
in turbulence accelerates the electrodeposition process and suggest by mass transport at these varying momentum transport conditions.
hydrodynamic conditions are favorable for the struvite formation. This effect of turbulence on struvite deposition i.e. higher turbulence
The residual current is higher at stagnant conditions (Re = 0) in favoring struvite deposition has also been observed under other
comparison to hydrodynamic conditions, indicating oxygen reduc- physicochemical conditions. Le Corre et al. noted deposition at
tion is mass transport limited. The turbulent flow at the electrode wastewater treatment facilities where fluid turbulence was high and
transports both oxygen gas for reduction and ionic species for attributed this to increase in pH due to carbon dioxide desorption
struvite deposition. When turbulence increases, the convection factor from solution.5 This increased pH favors ion speciation for struvite
of total mass transfer increases; thus, the mass transfer of dissolved deposition. Other authors36–38 found similar correlations with in-
oxygen increases. The results shown in Fig. 6a are in agreement with creased mixing energy resulting in increased crystal size albeit pH
those obtained by Kear et al.35 increase was not infered. It should be noted that this correlation
Microscopic observations (Fig. S3 in Supplementary Information) between turbulence and struvite deposition reaches a point of
indicated that under stagnant conditions, the struvite layer which diminishing returns, as higher mixing rates result in shearing of the
results from a vertical crystal growth frequently presents a less struvite crystals and thus reduce deposition/growth rate.7,39
compact structure. Above a certain value of the Reynolds number Figure 6c shows typical curves of cumulative current efficiency
(9150), the scale deposit was more compact, and composed of as a function of time for different Reynolds number, with
Journal of The Electrochemical Society, 2020 167 155524

Figure 7. Effect of temperature on (a) residual current, (b) phosphorus removal efficiency, (c) current efficiency and (d) specific energy consumption during
phosphorus removal. Initial conditions: 250 ml of 7.84 × 10−3 M of NH4H2PO4 and 10.71 × 10−3 M of MgCl2·6H2O neutralized to a pH of ∼7.3, −1.1 V (vs
Ag/AgCl), Reynolds number of 9150 and 0 mg l−1 of Ca2+.

efficiencies close to 16% for Reynolds numbers of 0, 3360 and coefficients. At 5 °C, 20 °C and 35 °C, oxygen’s solubility values in
12811 in the first 10 min and decreasing to ∼4% after 100 min. pure water are 8.93, 6.35 and 4.85 ml l−1, respectively and these
However, for Reynolds number of 9150, the current efficiency is solubilities are indirectly proportional to total dissolved solids.40
close to 23% early in the experiment and decreases to 9% after This would mean as the electrolyte temperature increases, oxygen’s
100 min. The decrease in current efficiency for all the Reynolds concentration in water decreases and there is less oxygen available at
numbers is likely due to an asymptotic phosphorus removal the cathode regardless of turbulent conditions due to a lower
efficiency associated with an increasing charge transfer during the availability in the bulk solution. On the other hand, O2 diffusivity
cathodic oxidation. The specific energy consumption as a function of in water is directly proportional to temperature41 with values of
time for different Reynolds number for stainless steel electrode is 1.2 × 10−5 cm2 s−1 at 0 °C, 2.0 × 10−5 cm2 at 20 °C and between
shown in Fig. 6d. Minimum specific energy consumption was 2.6 × 10−5 cm2 s−1 and 2.9 × 10−5 cm2 s−1 at 35 °C. This would
observed at the Reynolds number of 9150, since this energy indicate faster transport of the available oxygen through the
corresponds to the maximum phosphorus removal. The increase of electrolyte to the cathode. Finally, the solubility product equilibrium
Es when the Reynolds increases (0, 3360 and 12811) can be constant for struvite increases with increasing temperature42 from
attributed to the lower current efficiency associated with these pKsp (15 °C) = 13.27 to pKsp (40 °C) = 12.52. Other studies43,44
flow conditions. have suggested that the Ksp increases between 15 °C and 30 °C and
slightly decreases between 30 °C and 35 °C with Ksp (20 °C) lower
Effect of temperature.—The current generated, phosphorus re- than Ksp (35 °C). Under both prescribed temperature effects, struvite
moval efficiency, current efficiency and specific energy consumption is most likely to deposit at 5 °C and least likely to form at 35 °C due
due to different temperature conditions are provided in Fig. 7. As to process thermodynamics.
observed in Fig. 7a, a residual current of 0.10, 0.25 and 0.60 was Overall, the combination of transport and thermodynamic con-
reached after 80 min of immersion for 20 °C, 35 °C and 5 °C, straints provide insight into the temperature and phosphorus
respectively. A similar phenomenon was observed in the RCE recovery observed. At 5 °C, struvite deposition is thermodynami-
configuration,7 although the residual current profile at 20 °C and cally favored, and bulk oxygen concentration is maximum but
35 °C were identical in comparison to this study. In addition, oxygen reduction (and therefore struvite deposition) is transport-
maximum phosphorus recovery (Fig. 7b) and current efficiency limited due to low temperatures that inhibit both oxygen and ionic
(Fig. 7c) were observed at 20 °C. Therefore, the rate of struvite diffusivity. Therefore, there is limited struvite deposition on the
deposition increased when the temperature increased from 5 °C to 20 ° cathode and the residual current is high because the electrode area
C but decreased when the temperature increased from 20 °C to 35 °C. remains exposed for oxygen reduction (Fig. 7a). This lower
These observed phenomena are likely due to a combination of temperature would also limit crystal growth kinetics.5 At 35 °C,
oxygen solubility, struvite deposition thermodynamics and diffusion struvite deposition is thermodynamically inhibited, and the bulk
Journal of The Electrochemical Society, 2020 167 155524

Figure 8. Effect of calcium concentration on (a) residual current, (b) phosphorus removal efficiency, (c) current efficiency and (d) specific energy consumption
during phosphorus removal. Initial conditions: 250 ml of 7.84 × 10−3 M of NH4H2PO4 and 10.71 × 10−3 M of MgCl2·6H2O neutralized to a pH of ∼7.5,
−1.1 V (vs Ag/AgCl), Reynolds number of 9150 and 20 °C.

oxygen concentration is significantly reduced in comparison to 5 °C, generated when the temperature is 35 °C as opposed to 5 °C. The
however oxygen and ionic transport to the cathode are favored. energy consumption (Fig. 7d) illustrates the impact of the signifi-
Therefore, at this temperature, although Mg2+, NH4+ and PO43− are cantly low phosphorus recovery at 35 °C as this results in the highest
more likely to remain in solution, fast transport of ions and oxygen energy consumption across all variables in this study, meanwhile at
as well as favored kinetics for crystal growth allow for a faster rate 20 °C this specific energy is low since the most significant amount of
of struvite deposition. This deposition causes a smaller residual phosphorus is recovered.
current than at 5 °C since more of the surface is covered by struvite. SEM and EDS analysis of the deposits generated at different
Over time, this higher temperature and higher oxygen reduction temperatures are shown in Fig. S4 in Supplementary Information
(a high localized pH) will result in NH3 formation and volatilization38 and indicate the electrode surface was covered with a porous struvite
leading to reduced phosphorus removal from solution as observed in layer precipitated in orthorhombic crystalline form.7 There were no
Fig. 7b. At 20 °C, both thermodynamic conditions and oxygen significant morphological differences between the different tempera-
solubility are intermediate between 5 °C and 35 °C, with favored ture conditions, suggesting minimal temperature effect on mor-
transport and kinetics conditions allowing for significant struvite phology. The EDS analysis of the surface indicates the presence of
deposition at the surface that both reduces residual current (due to phosphorus, magnesium and oxygen, elements typical of struvite
cathode coverage) and maximizes phosphorus recovery as struvite. formation.
These results suggest a complex relationship between temperature
conditions and phosphorus recovery as struvite. In fact, Le Corre Effect of Ca2+.—The current generated, phosphorus removal
et al. suggest 35 °C as the upper limit for precipitating struvite from efficiency, current efficiency and specific energy consumption due
synthetic/real liquors based on their literature review,5 while a more to different calcium concentrations are provided in Fig. 8. The
recent review45 highlights conflicting temperature effects from prior chronoamperometric curves (CA) in Fig. 8 indicate a residual CA
studies and prescribes improved temperature control as a funda- value of 0.10 at 0 mg l−1 Ca2+, 0.5 at 72 mg l−1 Ca2+ and 0.9 at
mental requirement for high fidelity experiments. In this study, a 210 mg l−1 Ca2+ after 60 min. In addition, the phosphorus removal
jacketed reactor (Fig. 1) was used to ensure isothermal conditions at (Fig. 8b) and current efficiency (Fig. 8c) decrease with the increased
the specified temperature conditions and provides the recommended presence of calcium. However, the specific energy consumption
fidelity. In addition, a multi-factor analyses could also delineate the (Fig. 8d) as a function of time increases with the presence of
effects of thermodynamics and transport conditions, since it is calcium, due to the decrease in the current efficiency.
plausible that a higher potential could accelerate struvite deposition The decrease in residual current in the absence of calcium is
at 35 °C and mimic the behavior at 20 °C and −1.1 V vs Ag/AgCl attributed to solid struvite buildup on the electrode surface inhibiting
(Fig. 7a). The current efficiency behavior (Fig. 7c) also reflects the oxygen diffusion and reduction. The presence of Ca2+ ions was
reduced impact of phosphorus recovery in comparison to the current expected to reduce struvite precipitation, due to the potential
Journal of The Electrochemical Society, 2020 167 155524

Figure 9. Effect of bulk solution pH and calcium concentration on solid deposition and ion speciation based on a ratio of Mg:N:P ratio of 1.37:1:1 and using the
flowsheet illustrated in Fig. 2. Expected solid deposition based on thermodynamics are shown with respect to (a) 0 Ca:Mg, (b) 0.17 Ca:Mg and (c) 0.49 Ca:Mg,
with neither Ca(OH)2 or Mg(OH)2 deposition possible under these conditions. Expected phosphorus speciation in the liquid phase based on moles of H2PO4−,
HPO42−, PO43−, MgH2PO4+, MgHPO4, MgPO4−, CaH2PO4+, CaHPO4 and CaPO4− and illustrated as (d) total liquid phosphorus species, (e) Mg-based liquid
phosphorus species and (f) Ca-based liquid phosphorus species.

formation of solid calcium-based species (e.g. amorphous calcium phenomenon: (1) the presence of solvated calcium-phosphate
phosphate, hydroxyapatite, dicalcium phosphate etc.).46 This re- species—CaH2PO4+, CaHPO4 and CaPO4−—that reduce available
duced struvite formation and increase calcium deposition would still phosphorus for struvite formation and (2) the formation of amor-
result in sufficient electrode coverage to decrease residual current. phous calcium phosphate (ACP or Ca9(PO4)6•xH2O) that initially
However, the residual current increased with increasing [Ca2+] at co-precipitates with struvite then re-dissolves as the reaction
the conditions evaluated. It is plausible the solid Ca-based species proceeds.47 The former hypothesis will be explained within the
may not have adsorbed on the electrode surface and instead settled to context of the developed thermodynamic model, so the latter
the bottom of the reactor, however the phosphorus removal hypothesis will be discussed.
efficiency did not support this (Fig. 8b). As [Ca2+] increased, the ACP is an unstable calcium phosphate phase that serves as a
phosphorus removal efficiency decreased, suggesting that more P- precursor for the least water-soluble calcium orthophosphate
based species—H2PO4−, HPO42− and PO43−—remained in solution phase46: hydroxyapatite (HAP) or Ca10(PO4)6(OH)2. This is espe-
during the experimental trial. Therefore, the combination of an cially true at pH > 7, as is this case in this current study. However,
increase in residual current coupled with a decrease in phosphorus HAP crystal growth is significantly inhibited by the presence of
recovery, indicate a decrease in solid struvite deposition with [Mg2+] with concentrations as low as 1 mM Mg2+ reducing HAP
increasing [Ca2+]. There are two possible reasons for this crystal size by 80% in comparison to magnesium-free HAP
Journal of The Electrochemical Society, 2020 167 155524

1:1:1) qualitatively agree with the pH range of 8–10 providing the


maximum amount of solid P deposited as struvite (Fig. 9a), with
other solid Mg-P phases available outside of this range (due to
decreased NH4+ activity).
Once calcium is added to the system, the pH window for potential
struvite deposition shrinks, especially at pH values ∼11. When
[Ca2+] = 72 ppm is added, struvite will only deposit at 6.7 ⩽ pH ⩽
10.6 (Fig. 9b), while when [Ca2+] = 210 ppm is added the deposition
window decreases to 6.7 ⩽ pH ⩽ 10.0 (Fig. 9c). In other words,
calcium will further limit struvite deposition at higher pH values
(in addition to decreased NH4+ activity) under the conditions tested.
This occurs because the supersaturation index for calcium phosphate is
higher than the index for struvite under these conditions, therefore it
will precipitate at lower pH values as seen in Figs. 9b and 9c and
limits the phosphate available for struvite formation at high pH values.
The low concentration of solvated calcium-based phosphate species in
solution—CaH2PO4+, CaHPO4 and CaPO4−—in the region where
struvite deposition is maximized (Fig. 9f) suggests there is no
competition for phosphate species from these species. This negates
Figure 10. Contour plot depicting percentage of phosphorus recovered as the alternate hypothesis previously proposed for reduced struvite
solid struvite based on solution pH and Ca:Mg ratio. The predicted deposition when [Ca2+] increases. In fact, it is more likely that
percentage is based on the supersaturation index of struvite based on the phosphate species in solution bind to Mg2+—as MgH2PO4+,
respective activities of Mg2+, NH4+ and PO43−. MgHPO4 and MgPO4−—due to the activity of the [Mg2+] in this
system (Fig. 9e). Overall, these findings seem to support the
crystallization.48 Therefore, under the conditions of this study, hypothesis that Ca3(PO4)2 as ACP competes with struvite deposition,
calcium is likely to form ACP which does not result in a significant as seen in Fig. 8, and the high activity of Mg2+ limits the conversion
amount of solid calcium phosphate. Instead, it is hypothesized that of ACP to HAP: the most stable solid calcium-based phosphate.
this ACP phase is formed and initially stabilized by the presence of Therefore, high calcium concentrations reduce struvite deposition,
Mg2+ and then rather than transforming to stable HAP, the ACP re- while also retaining phosphorus in solution due to ACP instability.
dissolves as observed in a prior literature study where ACP dissolved These thermodynamic simulations therefore predict regions that
as early as the first 30 min.47 This could explain the reason for the potentially maximize struvite recovery based on the pH and the Ca:
observed high concentration of P retained in solution (Fig. 8b) at Mg ratio in the system (Fig. 10) and provide a useful heuristic for
high [Ca2+], while minimum struvite deposition occurs. Some of the optimum design variables for maximum struvite recovery via
ACP that did not re-dissolve may also be trapped in the struvite electrochemical oxygen reduction.
matrix since Ca is observed in the SEM EDS analyses seen in
Fig. S5 in Supplementary Information. ACP formation and dissolu- Conclusions
tion kinetics were not evaluated in this study but could be captured, Overall, under the conditions tested, optimum struvite recovery
using in situ ultraviolet-visible spectroscopy for example, since ACP defined by high phosphorus removal efficiency, high current
gives off a milky iridescence.46 efficiency and low specific energy consumption was possible at
20 °C, potential of −1.1 V vs Ag/AgCl, a Reynolds number of 9150,
Thermodynamic Simulations.—Initial simulations reproduced in the absence of calcium ions. Cathodic potentials lower than
Bjerrum plots for orthophosphate species and ion speciation plots −1.1 V vs Ag/AgCl will result in competing reactions that could lead
for ammonium species (Fig. S6 in Supplementary Information) to struvite re-dissolution. Higher turbulence i.e. Reynolds number, in
similar to those observed in literature.22,23,49 When the overall N:P the system will decrease phosphorus recovery due to shearing
molar ratio was 1:1, the equimolar amounts of the buffer systems (and likely re-dissolving) of surface deposited solid species.
occurred at the following pH values: [NH4+] = [NH3] at pH = Temperatures higher than 20 °C result in competing phenomena
10.05, [H2PO4−] = [HPO42−] at pH = 8.32 and [HPO42−] = associated with oxygen solubility, oxygen transport and struvite
[PO43−] at pH = 12.79 as shown in Fig. S6. In the presence of deposition thermodynamics.
magnesium and calcium ions, other species highlighted in Table II Finally, the presence of calcium in the system leads to the
begin to form (Fig. 9). thermodynamically favorable formation of Ca3(PO4)2 as unstable
When magnesium is added into the ammonium-phosphate mix, ACP, which rather than form stable HAP re-dissolves in the solution.
solid magnesium species can potentially form, with struvite deposi- However, since calcium must be present in the system, due to its
tion between 6.6 and 12.1, magnesium phosphate precipitation ubiquity in animal waste, an increase in magnesium concentration
between 6.7 and 14.0 and no magnesium hydroxide generated coupled with a pH below NH3 volatilization will decrease the Ca:Mg
(Fig. 9a). As the [OH−] increases, NH4+ is converted to volatile ratio and retain NH4+ in solution. These conditions will thermo-
NH3 and struvite deposition becomes thermodynamically limited, dynamically favor struvite deposition and limit calcium deposition
leaving excess magnesium for magnesium phosphate precipitation. as a stable phase. Other alternatives to optimum struvite generation
On the other hand, Mg(OH)2 does not precipitate within this system under these conditions will require multi-factor analyses to delineate
since the available Mg2+ will form either magnesium phosphate or interacting variables for electrochemical phosphorus removal from
struvite. Overall, solid phosphorus recovery is possible between pH wastewater. Furthermore, the effect of additional solvated and solid
values of 6.6 and 14.0, with most of this being primarily due to species including magnesium ammine complexes and other magne-
struvite deposition. Under the conditions of the experiment per- sium and calcium orthophosphates will need to be evaluated to
formed (Mg2+:NH4+:PO43− is 1.37:1:1) and simulating swine further improve the fidelity of the thermodynamic simulations.
wastewater, the maximum potential phosphorus removal as solid is
100% between pH = 9.0 and 11.0 (Fig. 9d). This value corresponds
Acknowledgments
to a combined supersaturation index for both struvite and magne-
sium phosphate based on the solubility products in Table II. The authors would like to thank the Ohio Water Development
Experimental results50,51 (with Mg2+:NH4+:PO43− present as Authority for funding this work under a 2016 R & D grant.
Journal of The Electrochemical Society, 2020 167 155524

ORCID 24. S. Daneshgar, A. Buttafava, D. Capsoni, A. Callegari, and A. G. Capodaglio,


“Impact of PH and ionic molar ratios on phosphorous forms precipitation and
Zineb Belarbi https://orcid.org/0000-0001-8241-3156 recovery from different wastewater sludges.” Resources, 7, 71 (2018).
Damilola A. Daramola https://orcid.org/0000-0002-6737-415X 25. M. Eisenberg, C. W. Tobias, and C. R. Wilke, “Ionic mass transfer and
Jason P. Trembly https://orcid.org/0000-0002-9851-2914 concentration polarization at rotating electrodes.” J. Electrochem. Soc., 101, 306
(1954).
26. H. Saraç, M. A. Patrick, and A. A. Wragg, “Physical properties of the ternary
References electrolyte potassium ferri-ferrocyanide in aqueous sodium hydroxide solution in
the range 10–90 °C.” J. Appl. Electrochem., 23, 51 (1993).
1. J. Pelley, “Taming toxic algae blooms.” ACS Cent. Sci., 2, 270 (2016). 27. D. R. Gabe, G. D. Wilcox, J. Gonzalez-Garcia, and F. C. Walsh, “The rotating
2. U.S. Environmental Protection Agency, A Compilation of Cost Data Associated cylinder electrode: its continued development and application.” J. Appl.
with the Impacts and Control of Nutrient Pollution; EPA 820-F-15-096 (Office of Electrochem., 28, 759 (1998).
Water, US EPA) 110 (2015). 28. K. N. Ohlinger, T. M. Young, and E. D. Schroeder, “Predicting struvite formation
3. C. M. Long, R. L. Muenich, M. M. Kalcic, and D. Scavia, “Use of manure nutrients in digestion.” Water Res., 32, 3607 (1998).
from concentrated animal feeding operations.” J. Gt. Lakes Res., 44, 245 (2018). 29. Aspen Technology Inc, Aspen Plus V10.0 (36.0.0.249) (2017).
4. J. T. Bunce, E. Ndam, I. D. Ofiteru, A. Moore, and D. W. Graham, “A review of 30. B. Fritz, Etude thermodynamique et modélisation des réactions hydrothermales et
phosphorus removal technologies and their applicability to small-scale domestic diagénétiques (Doctoral Thesis), Université Louis-Pasteur (now University of
wastewater treatment systems.” Front. Environ. Sci., 6, 8 (2018). Strasbourg) (1981), https://www.persee.fr/doc/sgeol_0302-2684_1981_mon_65_1.
5. K. S. Le Corre, E. Valsami-Jones, P. Hobbs, and S. A. Parsons, “Phosphorus 31. C.-C. Chen, H. I. Britt, J. F. Boston, and L. B. Evans, “Local composition model for
recovery from wastewater by struvite crystallization: a review.” Crit. Rev. Environ. excess Gibbs energy of electrolyte systems. Part I: Single solvent, single completely
Sci. Technol., 39, 433 (2009). dissociated electrolyte systems.” AIChE Journal, 28, 588 (1982).
6. M. M. Rahman, M. A. M. Salleh, U. Rashid, A. Ahsan, M. M. Hossain, and 32. H. Renon and J. M. Prausnitz, “Local compositions in thermodynamic excess
C. S. Ra, “Production of slow release crystal fertilizer from wastewaters through functions for liquid mixtures.” AIChE J., 14, 135 (1968).
struvite crystallization—a review.” Arab. J. Chem., 7, 139 (2014). 33. J. D. Hunter, “Matplotlib: A 2D graphics environment.” Comput. Sci. Eng., 9, 90
7. Z. Belarbi and J. P. Trembly, “Electrochemical processing to capture phosphorus (2007).
from simulated concentrated animal feeding operations waste.” J. Electrochem. 34. J. M. Grau and J. M. Bisang, “Electrochemical removal of cadmium from dilute
Soc., 165, E685 (2018). aqueous solutions using a rotating cylinder electrode of wedge wire screens.”
8. Y. Lei, B. Song, M. Saakes, R. D. van der Weijden, and C. J. N. Buisman, J. Appl. Electrochem., 37, 275 (2007).
“Interaction of calcium, phosphorus and natural organic matter in electrochemical 35. G. Kear, K. Bremhorst, S. Coles, and S.-H. Huáng, “Rates of mass transfer to patch
recovery of phosphate.” Water Res., 142, 10 (2018). electrodes in disturbed flows using oxygen reduction under limiting current
9. Y. Lei, J. C. Remmers, M. Saakes, R. D. van der Weijden, and C. J. N. Buisman, “Is conditions.” Corros. Sci., 50, 1789 (2008).
there a precipitation sequence in municipal wastewater induced by electrolysis ? ” 36. K. N.;. P. E.,. N. Ohlinger, T. M. Young, and E. D. Schroeder, “Kinetics effects on
Environ. Sci. Technol., 52, 8399 (2018). preferential struvite accumulation in wastewater.” J. Environ. Eng., 125, 730 (1999).
10. Y. Lei, J. C. Remmers, M. Saakes, R. D. van der Weijden, and C. J. N. Buisman, 37. J. D. Doyle and S. A. Parsons, “Struvite formation, control and recovery.” Water
“Influence of cell configuration and long-term operation on electrochemical Res., 36, 3925 (2002).
phosphorus recovery from domestic wastewater.” ACS Sustain. Chem. Eng., 7, 38. A. Capdevielle, E. Sýkorová, B. Biscans, F. Béline, and M.-L. Daumer,
7362 (2019). “Optimization of struvite precipitation in synthetic biologically treated swine
11. Y. Lei, I. Hidayat, M. Saakes, R. van der Weijden, and C. J. N. Buisman, “Fate of wastewater—determination of the optimal process parameters.” J. Hazard.
calcium, magnesium and inorganic carbon in electrochemical phosphorus recovery Mater., 244–245, 357 (2013).
from domestic wastewater.” Chem. Eng. J., 362, 453 (2019). 39. A. E. Durrant, M. D. Scrimshaw, I. Stratful, and J. N. Lester, “Review of the
12. I. Wu, A. Teymouri, R. Park, L. F. Greenlee, and A. M. Herring, “Simultaneous feasibility of recovering phosphate from wastewater for use as a raw material by the
electrochemical nutrient recovery and hydrogen generation from model wastewater phosphate industry.” Environ. Technol., 20, 749 (1999).
using a sacrificial magnesium anode.” J. Electrochem. Soc., 166, E576 (2019). 40. R. F. Weiss, “The solubility of nitrogen, oxygen and argon in water and seawater.”
13. L. Kékedy-Nagy, A. Teymouri, A. M. Herring, and L. F. Greenlee, “Electrochemical Deep Sea Res. Oceanogr. Abstr, 17, 721 (1970).
removal and recovery of phosphorus as struvite in an acidic environment using pure 41. W. Xing, M. Yin, Q. Lv, Y. Hu, C. Liu, and J. Zhang, “Oxygen solubility, diffusion
magnesium vs the AZ31 magnesium alloy as the anode.” Chem. Eng. J., 380, 122480 coefficient, and solution viscosity.” Rotating Electrode Methods and Oxygen
(2020). Reduction Electrocatalysts, ed. W. Xing, G. Yin, and J. Zhang (Elsevier,
14. O. Modin and D. J. I. Gustavsson, “Opportunities for microbial electrochemistry in Amsterdam) 1 (2014).
municipal wastewater treatment—an overview.” Water Sci. Technol., 69, 1359 42. H. K. Aage, B. L. Andersen, A. Blom, and I. Jensen, “The solubility of struvite.”
(2014). J. Radioanal. Nucl. Chem., 223, 213 (1997).
15. P. T. Kelly and Z. He, “Nutrients removal and recovery in bioelectrochemical 43. M. I. H. Bhuiyan, D. S. Mavinic, and R. D. Beckie, “A solubility and
systems: a review.” Bioresour. Technol., 153, 351 (2014). thermodynamic study of struvite.” Environ. Technol., 28, 1015 (2007).
16. Y. V. Nancharaiah, S. Venkata Mohan, and P. N. L. Lens, “Recent advances in 44. M. Hanhoun, L. Montastruc, C. Azzaro-Pantel, B. Biscans, M. Frèche, and
nutrient removal and recovery in biological and bioelectrochemical systems.” L. Pibouleau, “Temperature impact assessment on struvite solubility product: a
Bioresour. Technol., 215, 173 (2016). thermodynamic modeling approach.” Chem. Eng. J., 167, 50 (2011).
17. G.-H. Huang, T.-C. Chen, S.-F. Hsu, Y.-H. Huang, and S.-H. Chuang, “Capacitive 45. B. Li, H. M. Huang, I. Boiarkina, W. Yu, Y. F. Huang, G. Q. Wang, and
deionization (CDI) for removal of phosphate from aqueous solution.” Desalination B. R. Young, “Phosphorus recovery through struvite crystallisation: recent
Water Treat., 52, 759 (2014). developments in the understanding of operational factors.” J. Environ. Manage.,
18. J. Choi, P. Dorji, H. K. Shon, and S. Hong, “Applications of Capacitive Deionization: 248, 109254 (2019).
Desalination, Softening, Selective Removal, and Energy Efficiency..” Desalination, 46. M. S. Johnsson and G. H. Nancollas, “The role of brushite and octacalcium
449, 118 (2019). phosphate in apatite formation.” Crit. Rev. Oral Biol. Med., 3, 61 (1992).
19. B. Li, I. Boiarkina, W. Yu, H. M. Huang, T. Munir, G. Q. Wang, and B. R. Young, 47. A. Capdevielle, E. Sýkorová, F. Béline, and M.-L. Daumer, “Kinetics of struvite
“Phosphorous recovery through struvite crystallization: challenges for future precipitation in synthetic biologically treated swine wastewaters.” Environ.
design.” Sci. Total Environ., 648, 1244 (2019). Technol., 35, 1250 (2014).
20. R. D. Cusick, M. L. Ullery, B. A. Dempsey, and B. E. Logan, “Electrochemical 48. M. H. Salimi, J. C. Heughebaert, and G. H. Nancollas, “Crystal growth of calcium
struvite precipitation from digestate with a fluidized bed cathode microbial phosphates in the presence of magnesium ions.” Langmuir, 1, 119 (1985).
electrolysis cell.” Water Res., 54, 297 (2014). 49. G. Hanrahan, Chapter 3 - Aqueous Chemistry. In Key Concepts in Environmental
21. T. Michałowski and A. Pietrzyk, “A thermodynamic study of struvite+water Chemistry, ed. G. Hanrahan (Academic, New York: Boston) p. 73 (2012).
system.” Talanta, 68, 594 (2006). 50. N. Ma, A. A. Rouff, and B. L. Phillips, “A 31P NMR and TG/DSC-FTIR
22. B. Tansel, G. Lunn, and O. Monje, “Struvite formation and decomposition investigation of the influence of initial PH on phosphorus recovery as struvite.” ACS
characteristics for ammonia and phosphorus recovery: a review of magnesium- Sustain. Chem. Eng., 2, 816 (2014).
ammonia-phosphate interactions.” Chemosphere, 194, 504 (2018). 51. V. Babić-Ivančić, J. Kontrec, L. Brečević, and D. Kralj, “Kinetics of struvite
23. M. I. Ali, “Struvite crystallization in fed-batch pilot scale and description of to newberyite transformation in the precipitation system MgCl2–NH4H2PO4–
solution chemistry of struvite.” Chem. Eng. Res. Des., 85, 344 (2007). NaOH–H2O.” Water Res., 40, 3447 (2006).

You might also like