Energy Recovery Through Reverse Electrodialysis Harnessing The Salinity Gradient From The Flushing of Human Urine

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Water Research 186 (2020) 116320

Contents lists available at ScienceDirect

Water Research
journal homepage: www.elsevier.com/locate/watres

Energy recovery through reverse electrodialysis: Harnessing the


salinity gradient from the flushing of human urine
Federico Volpin a,f, Yun Chul Woo b, Hanki Kim c, Stefano Freguia d, Namjo Jeong c,
June-Seok Choi b, Jaeweon Cho e, Sherub Phuntsho a, Ho Kyong Shon a,∗
a
School of Civil and Environmental Engineering, University of Technology, Sydney (UTS), City Campus, Broadway, NSW 2007, Australia
b
Department of Land, Water and Environment Research, Korea Institute of Civil Engineering and Building Technology, 283 Goyang-Daero, Ilsanseo-Gu,
Goyang-Si,Gyeonggi-Do, 10223, Republic of Korea
c
Jeju Global Research Center, Korea Institute of Energy Research, 200 Haemajihaean-ro, Gujwa-eup, Jeju, 63359, Republic of Korea
d
Department of Chemical Engineering, The University of Melbourne, Victoria 3010, Australia
e
School of Urban and Environmental Engineering, Ulsan Institute of Science and Technology (UNIST), UNIST-gil 50, Ulsan 689-798, Republic of Korea
f
City Water Technology, 2072 Sydney, Australia

a r t i c l e i n f o a b s t r a c t

Article history: Urine dilution is often performed to avoid clogging or scaling of pipes, which occurs due to urine’s Ca2+
Received 17 April 2020 and Mg2+ precipitating at the alkaline conditions created by ureolysis. The large salinity gradient between
Revised 6 July 2020
urine and flushing water is, theoretically, a source of potential energy which is currently unexploited. As
Accepted 19 August 2020
Available online 19 August 2020
such, this work explored the use of a compact reverse electrodialysis (RED) system to convert the chem-
ical potential energy of urine dilution into electric energy. Urine’ composition and ureolysis state as well
Keywords: as solution pumping costs were all taken into account. Despite having almost double its electric con-
Human urine ductivity, real hydrolysed urine obtained net energy recoveries ENet of 0.053–0.039 kWh/m3 , which is
Reverse electrodialysis similar to energy recovered from real fresh urine. The reduced performances of hydrolysed urine were
Nutrients recovery linked to its higher organic fouling potential and possible volatilisation of NH3 due to its high pH. How-
Power generation
ever, the higher-than-expected performance achieved by fresh urine is possibly due to the fast diffusion
Oxidation
of uncharged urea to the freshwater side. Real urine was also tested as a novel electrolyte solution and
its performance compared with a conventional K4 Fe(CN)6 /K3 Fe(CN)6 couple. While K4 Fe(CN)6 /K3 Fe(CN)6
outperformed urine in terms of power densities and energy recoveries, net chemical reactions seemed to
have occurred in urine when used as an electrolyte solution, leading to TOC, ammonia and urea removal
of up to 13%, 6% and 4.4%, respectively. Finally, due to the migration of K+ , NH4 + and PO4 3− , the low
concentration solution could be utilised for fertigation. Overall, this process has the potential of provid-
ing off-grid urine treatment or energy production at a household or building level.
© 2020 Elsevier Ltd. All rights reserved.

1. Introduction rus (P) in the remaining WW, thereby making it easier to treat
(Maurer et al., 2006; Putnam, 1971a; Rose et al., 2015). The re-
The latest WHO/UNICEF joint monitoring program for water ported concentration of nitrogenous compounds in typical urine
supply, sanitation and hygiene shows that more than 30% of the can exceed 80 0 0 mg-N/L, while total phosphorus levels can add up
world population still does not have access to essential sanita- to 500 mg/L (Larsen et al., 2016, 2013; Udert et al., 2006). That is
tion services (WHO and UNICEF, 2017). However, in areas like ur- because urine contains over 60 0 0 mgN /L of urea (which hydrolyses
ban slums, centralised wastewater (WW) treatment is not prac- into NH3 during storage), 300 mg/L of uric acid, 250 mg/L of cre-
tical as sewage pipelines cannot be laid. As such, on-site sanita- atine and 1400 mg/L creatinine (Putnam, 1971b). This makes urine
tion is the only option. One commonly proposed approach for im- almost as saline as seawater (Volpin et al., 2019). As urine is gen-
proving decentralised sanitation is to treat human urine separately erally mixed with flushing water, which has an average total dis-
from the remaining WW. This is because urine separation would solved solids (TDS) between 0.1 and 1 g/L, there is a large salinity
reduce up to 81% of the nitrogen (N) and 50% of the phospho- gradient of mixing, which could be converted into chemical poten-
tial energy.
∗ While several articles looked at the energy recovery from the
Corresponding author.
E-mail address: Hokyong.Shon-1@uts.edu.au (H.K. Shon). mixing of saline industrial wastewaters, mainly using processes

https://doi.org/10.1016/j.watres.2020.116320
0043-1354/© 2020 Elsevier Ltd. All rights reserved.
2 F. Volpin, Y.C. Woo and H. Kim et al. / Water Research 186 (2020) 116320

like pressure retarded osmosis (PRO) and reverse electrodialysis In this work, we refer to the urine before ureolysis as FU and after
(RED), so far only a few articles looked into the recovery of urine’s as HU.
chemical potential energy (Han et al., 2020; Ma et al., 2020; Also, each test was firstly run with synthetic urine solution (Ta-
Mehdizadeh et al., 2020). ble S1) and afterwards, a real solution. The model solutions were
For example, in one case, urine was used as a draw solu- used to provide an upper boundary in the maximum energy ex-
tion for forward osmosis (FO) to dewater a microalgae culture tractable from human urine. That is because synthetic urine re-
(Volpin et al., 2019). There, urine was used to draw water from sembles the ionic strength and composition of real urine without
the algae solution as its osmotic pressure was higher than the its complex, and high fouling, organic matrix. All the experimental
one of the algae solutions. However, this approach is only use- investigations were run were in triplicates.
ful for dewatering applications. To extract water and energy from
urine, Mercer et al. (2019) used a combination of membrane distil- 2.1.2. HC and LC solutions
lation (MD) and RED. Here, urine is firstly concentrated urine MD, The collection and storage of urine were done as described in
and the MD concentrate is then used to power the RED process. our previous work (Volpin et al., 2019). Before the feeding in the
In RED, cation and anion exchange membranes (CEMs and AEMs) module, HU was pre-filtered using a 0.1 μm hollow fibre microfil-
were used to allow only for the transport of ions. By separating the tration module to remove the carbonates and phosphates salts that
higher concentration (HC) urine solution from the lower concen- typically precipitate during urine hydrolysis. This was done to pre-
tration (LC) freshwater through alternated CEMs and AEMs mem- vent clogging of the RED flow channels. The average composition
branes, an ionic flux from HC to LC is created (Post et al., 2007; of FU and HU is displayed in Table 1.
Veerman et al., 2011; Vermaas et al., 2013). The ionic flux is then Before running the experiments with real urine, synthetic FU
converted into an electrical current thanks to redox reactions oc- and HU solutions were prepared (see composition in Table S1) and
curring at the anode and cathode of the electrode compartment fed to the RED stack (Volpin et al., 2019; Zhang et al., 2014). The
(Veerman et al., 2010). However, the hybrid RED-MD process was results from these tests were used to compare open circuit volt-
designed only for scenarios where waste heat is available. age (OCV) and gross power densities (PDgross ) of synthetic and real
However, if used as a stand-alone process, RED is possibly more urine, as well as to analyse the transport of monovalent and mul-
competitive compared to other salinity gradient-driven energy re- tivalent anions and cations from the HC to the LC solution. Each
covery processes such as PRO in the recovery of energy from urine. stack was also tested with synthetic seawater (0.6 M NaCl) before
That is because RED here has the advantage of being able to oper- and after the urine experiments to ensure similar stack baseline
ate at low flow rates, of requiring lower capital investment (given performances and measure any loss in performance after the urine
that no turbine or pressure exchangers are necessary) and is eas- tests. All the standard tests were performed using 0.015 M NaCl
ily scalable (Post et al., 2007; Yip and Elimelech, 2014). As reduced as LC solution. However, to have consistent results, real freshwa-
process complexity and cost is paramount if energy is to be recov- ter (Table 1) was also examined with real FU and HU before the
ered from urine at a household/building level, RED seems to be the analysis of the fouled membrane. This was done to see if any of
best candidate for this job. the Ca2+ or Mg2+ in the freshwater would precipitate on the IEMs.
That is why the goal of this work is to explore the use of stand- However, the results might differ if harder water is used.
alone RED to recover energy from the dilution of real fresh and Finally, both the LC and HC solutions were fed to the RED cell,
hydrolysed urine. in single-pass and counter-current flow, using peristaltic pumps
Also, for the first time, real urine was used both as a high- (Longer BT100 2 J). Flow rates of 5, 10, 15, 25 and 35 mL/min were
concentration solution and as an electrolyte solution. This would used and kept the same for both HC and LC solutions at all condi-
reduce the costs of the process, as no additional chemicals e.g., tions. According to Eq. (1), where Acs is the channels’ cross-section
electrolytes would be required. Additionally, the net chemical reac- area (as spacer thickness multiplied by channel width) and ϕ is the
tions occurring at the electrode compartment when urine is used spacer’ open area, the above flowrates translate in linear velocities
as electrolyte solution could promote organics and nitrogen re- of 2.82, 5.65, 8.47, 14.1 and 19.8 cm/s.
moval. Finally, the transport of plant macronutrients (N, P, K) from  cm   mL 
urine to the freshwater was also investigated to understand if the Q min
v = (1)
low concentration solution could then be used as fertigation solu- s 60 · Acs [cm2 ] · ϕ [%]
tion.
2.1.3. Electrolyte solutions
2. Methodology The experiments for System A (Fig. 1B) were con-
ducted recirculating a mixture of the reversible redox couple
2.1. Experimental plan K3 Fe(CN)6 /K4 Fe(CN)6 (0.05 M) combined with 0.5 M NaCl, while
in System B real FU or HU were used both as HC and ES solution
2.1.1. Process description (Fig. 1B). In both cases, the ES solutions were recirculated with a
Two different process operation modes were adopted and com- flow rate of 5 mL/min independently of the HC and LC solution
pared. While the first approach is expected to maximise the energy flow rates.
recovery (i.e., System A), the second does not need any additional
chemical (i.e., System B) (see Fig. 1): 2.1.4. Reverse electrodialysis cell
The RED cell consisted of a 20 cell pair stack, each with an
Ø System A: A mixture of K4 Fe(CN)6 /K3 Fe(CN)6 and NaCl was re-
area of 19.63 cm2 , with an overall active membrane area of 0.039
circulated as electrolyte solution (ES)
m2 . Figure S1, in the supporting information, shows the picture
Ø System B: Urine was flowed, in single-pass, as ES
of the RED setup used. Each membrane pair was separated by
Both systems were tested with fresh and hydrolysed human a reinforced glass fibre gasket and a woven mesh PTFE spacer
urine as HC solutions. That is because fresh and hydrolysed urine having a thickness of 100 μm to create an adequate flow chan-
are chemically very different as, when stored without any pH ad- nel. The lab-scale RED stack was provided by Jeju Global Research
justment, the urea in the fresh urine hydrolyses into ammonia and Center (JGRC) of Korea Institute of Energy Research (KIER). The
bicarbonate causing a significant increase in urine’s ionic strength spacer had an open area of 51% with a porosity of 70%. The elec-
and pH (Randall et al., 2016; Ray et al., 2018; Volpin et al., 2019). trode compartments (HCs), sitting at the anode and cathode of the
F. Volpin, Y.C. Woo and H. Kim et al. / Water Research 186 (2020) 116320 3

Fig. 1. (A) Schematic drawing of the single-pass RED experiments using fresh and hydrolysed urine. (B) Schematics of the two different configurations used for the exper-
iments. In the first configuration, (System A) K4 Fe(CN)6 /K3 Fe(CN)6 are used as ES and recirculated in the anode and cathode chamber. In the second configuration (System
B), real fresh or hydrolysed urine is used as ES.

Table 1
Average composition of the fresh and hydrolysed urine solutions, as well as the freshwater used as LC solution.

Urea N-NH4 + PO4 3− -P Cl− SO4 2− K+ Mg2+ Ca2+ Na+ EC pH


[mg/L] [mS/cm] [-]

Fresh Urine 9016 ± 1381 341 ± 155 332 ± 49 1320 ± 210 693 ± 79 529 ± 15 50 ± 18 47 ± 11 2533 ± 177 17.0 ± 3.2 6.6 ± 0.5
Hydrolysed Urine <1 4200 ± 1200 174 ± 11.5 1320 ± 210 693 ± 79 1611 ± 291 <1 <1 2211 ± 371 33.1 ± 4.4 9.1 ± 0.2
Drinking water a n.d. n.d. n.d. 14.2 ± 0.9 0.76 ± 0.06 1.6 ± 0.4 3.6 ± 2.5 13.5 ± 3 4.7 ± 0.3 [-] 7.5 ± 0.3
a
Based on lab analysis and Sydney Water potable water quality report (Sydney Water, 2019). n.d. = not detected.
4 F. Volpin, Y.C. Woo and H. Kim et al. / Water Research 186 (2020) 116320

Table 2 10 min was used to obtain stable OCV measurements. Afterwards,


Characteristic and proprieties of Fujifilm Type I membranes.
the slope of the I-V curve gave the value of Rstack .
Proprieties AEM-Type I CEM-Type I Unit W 
Membrane type Homogeneous Homogeneous –
P DGross =V ·I (4)
m2
Thickness 125 135 μm
Permeselectivity: The net power density (PDNet ) was then calculated by sub-
KCl (0.05 - 0.5 M) 92 93 % tracting the energy of pumping the HC and LC solutions (i.e.,
NaCl (0.05 - 0.5 M) 92 92 % 2∗ Ppumping ) to the PDGross value (Eq. (5)). After measuring the pres-
Electrical Resistance:
2 M NaCl 0.8 1.3 .cm−2
sure drop via a high resolution (±0.1 kPa) pressure sensor (Keller,
0.5 M NaCl 1.3 2.7 .cm−2 Reinacherstrasse, Basel, Switzerland), the pumping energy was cal-
Water permeation 6 10 kg.cm−2 culated as assuming 70% pump efficiency.
pH Stability 2 to 10 4 to 12 – W  2 · PPumping
P DNet = P DGross − (5)
m2 At
cell, comprised a Pt/Ti mesh (coating thickness, 3 μm, Wesco Elec- Finally, ENet was calculated as the produced power per urine
trode, South Korea) welded to a titanium rod which was then con- flow rate pumped through the stack (Eq. (6)).
nected to the source meter (Keithley 2400, Ohio, USA) (Kim et al.,  
kW h P DNet · At
2019). Galvanostatic measurements were conducted using Keithley ENet = (6)
m3 QUrine
KickStart software (Tektronix, Oregon, USA) following the protocol
described in the literature (Laktionov et al., 2003; Mercer et al.,
2019). Type I Fujifilm AEMs and CEMs (TYPE-I, FUJIFILM Manufac- 2.3. Nutrients recovery and removal
turing Europe BV, Tilburg, Netherlands) were used for the experi-
ments. Prior to use, IEMs were soaked in 0.5 M NaCl for swelling 2.3.1. Chemical analysis
for 24 h. The first characteristics are displayed in the table below While anions concentration was measured via Ion Chromatog-
(Table 2). New pristine membranes were used after each test with raphy (IC, Thermo Fisher Scientific, USA), cations were mea-
real human urine. sured using Microwave Plasma Atomic Emission Spectroscopy (MP-
AES 4100, Agilent, USA). Chloride was measured using Chlo-
2.2. Theoretical and experimental electrochemical calculations ride test kit (Kit 114,897, Merck KGaA, Darmstadt, Germany).
Also, the urea/ammonium Megazime kit (Megazyme, Australia)
2.2.1. Specific extractable energy from human urine was used to measure the concentration of NH3 and urea. This
The theoretical energy, as maximum Gibbs free energy was done by measuring the NH3 concentration, through a UV-
(Gmax ), from the dilution of human urine, was calculated as in Spectrophotometer (UV-1700, Shimadzu, Japan) at 340 nm wave-
Lin et al. (2014). This was done to calculate the energy efficiency length, before and after inducing urea hydrolysis via the addition
(η) of the process. η was calculated via Eq. (2), where ENet is the of urease enzyme. Finally, TOC was measured using a Multi N/C
measured recovered energy (Eq. (6)). 20 0 0 (Analytikjena, Sydney, Australia).
 kWh 
ENet 2.3.2. Mass balance calculations for nutrients recovery
η [% ] = mkWh  · 100
3
(2)
Gmax m3
The transport of ions (and urea) from the HC to the LC solu-
tion was measured for each feed solution, operating mode and lin-
To define Gmax , some initial assumption had to be made. ear velocities. As both HC and LC solutions were pumped with the
Firstly, the osmotic pressure of the solutions was approximated us- same flow-rate, and assuming a negligible water transport due to
ing the van’t Hoff equation π (c) = ν RTc, where ν , R and T are the osmosis, mass balance equations were used to calculate the recov-
van’t Hoff factor for strong electrolytes. An activity coefficient of ery of each targeted species (Eq. (7)).
one was used for the calculations. Using an activity coefficient of  mg   
one, Lin et al. (2014) showed less than 10% error on the G of RO Ci(HC,E f f ) Ci(LC,E f f ) mg
Ci, recovery [%] =1−  mg  ≈
L
 mgL  (7)
brine/river water mixing. As urine is even more dilute than sea- Ci(Urine) Ci(Urine)
L L
water, it is expected that the error would be even lower. Finally, a
maximum HC/LC mixing fraction of φ = 50% was used for the cal- Here, Ci (HC,Eff) and Ci (LC,Eff) are the concentration of the tar-
culations as both solutions were pumped at the same rate (i.e., at geted species (i), in HC and LC solutions, after passing through the
equilibrium 50% of the solutes in the HC solution is moved to the stack. Ci (Urine) is the initial concentration of the species (i) in the
LC solution). Once these assumptions were made, the Gmax for urine. If the difference between the mass balance in the HC and
synthetic FU and HU was calculated using Eq. (3) (Lin et al., 2014; LC solution was less than 5%, an average between the two values
Straub et al., 2016; Volpin et al., 2018; Yip and Elimelech, 2012). was taken. HC and LC effluents were sampled after operating the
Table S1 shows the osmotic pressure of FU and HU, calculated us- stack for at least 15 min at the amperage that maximises PDgross .
ing OLI Analyser (Version 9.5, Oli Systems Inc., USA). During this period, voltage and pressure losses were continuously
measured to ensure that the system was operated in steady-state
GMax = G(φ = 0.5) conditions.
= πM ln (πM ) − φπLC ln (πLC ) − (1 − φ )πHC ln (πHC ) (3) Finally, the difference in the total dissolved solids (TDS) be-
tween the inlet urine and the HC and LC effluents was also mea-
2.2.2. Power density and energy recovery sured to calculate the overall TDS recovery using Eq. (7).
Gross power densities measurements were conducted using a
source meter as galvanostat (Keithley 2400, Ohio, USA). A linear 2.3.3. TOC, nitrogen and chloride removal calculations
current from 0 to 100 mA and back was swept, with 5 mA step and Only for Scenario B (i.e., urine as electrolyte solution) the
30 s of delay, and Keithley KickStart software (Tektronix, Oregon, composition of urine before and after flowing through the elec-
USA) was used to record the voltage (V) continuously. Eq. (4) was trolyte compartments was measured to assess the any change in
used to calculate the gross power density (PDGross ) (Nam et al., the NH3 /NH4 + , urea, TN, TOC and Cl− concentrations. The re-
2018). For each experiment, an initial stabilisation period of at least sults were then discussed by comparing the mass balance of
F. Volpin, Y.C. Woo and H. Kim et al. / Water Research 186 (2020) 116320 5

such compounds, as Ci,removal [%] = 1 – (Ci (Urine,Effluent) [mg/L] / higher slopes than the seawater benchmark, despite having similar
Ci (Urine,Influent) [mg/L]), with theoretical pathways of electrooxida- or higher ionic conductivities. This again leads to the hypothesis of
tion. Urine samples was taken only after reaching stable voltage organic fouling introducing additional resistance.
readings. NO2 − and NO3 − concentrations at the effluent were also
measured as NH4 + could be directly oxidised to those species 3.1.1.2. Real FU vs HU performances. Fig. 3 shows that the per-
(Tarpeh et al., 2018). formances of real HU are just 13 ± 4% higher than FU, even
though HU has 50% higher salinity compared to FU (i.e., 33.1 vs
2.4. Membrane characterisation 17.1 mS/cm) and a much lower concentration of multivalent ions
(as almost half of the PO4 3− and all the Ca2+ and Mg2+ precipi-
Pristine and used CEMs and AEMs were characterised by scan- tates during storage). Ureolysis of FU during the experiments was
ning electron microscopy coupled with electron – dispersive X-Ray also excluded as the pH of FU in the influent and effluent was the
(EDX), using a SUPRA 55-VP instrument (Zeiss, Germany), and by same, indicating no increase in alkalinity. The explanation for this
Fourier transform infrared (FTIR) Spectroscopy using the IRAffinity- could be twofold.
1 from Shimadzu (Japan). The IEMs were characterised before and The most plausible reason is that HU has higher fouling ten-
after the tests with real FU and real HU. Before the characterisa- dency compared to FU due to the bacterial-aided transformation of
tion, a picture of the membranes was taken for visual analysis. the organics in urine during storage. The hydrolysis/fermentation
that occurs during storage causes the transformation of the long-
3. Results and discussion chain organic acids, creatinine, amino acids and carbohydrates into
acetate and other smaller molecules (Barbosa et al., 2019). In fact,
3.1. Performance using reversible redox couple as ES Fig. 9 shows a significantly higher amount of coloured compounds
deposited on the AEMs tested with real HU (see Section 3.3 for
In this section, the outcomes from the experiments that em- more details), though fouling did not cause any steep increase in
ployed K3 Fe(CN)6 / K4 Fe(CN)6 as ES are discussed. The change in the urine channel’s pressure drop, which increased linearly at in-
PDNet , ENet and N/P/K recovery at increasing urine flow rate was creasing flow velocity (Fig. 3A).
investigated. Firstly, baseline tests using synthetic seawater were Additionally, we here propose another mechanism that could
conducted and the results shown in Fig. 2A and Figure S4. The have enhanced FU performances. The hypothesis is that the high
goals of the section were: flux of urea from the HC to LC solution in FU could accelerate the
ionic transport to the LC solution, as urea molecules could have
Ø To describe and discuss the differences in the outcomes using helped to push the ions through the IEMs. In fact, the uncharged
fresh and hydrolysed urine, nature, high diffusivity coefficient (i.e., 1.38 • 109 m2 /s) and high
Ø To provide an optimal stack operating condition to balance the concentrations of urea in FU would mean that a significant amount
trade-off between power density and energy/nutrients recovery. would diffuse through the IEMs following a pure diffusion-driven
model (Ma et al., 2018). This can be observed by looking at the
3.1.1. Energy recovery and energy efficiency high TN in the LC effluent when using FU (Fig. 6B). However, at
3.1.1.1. Synthetic vs real urine. The results in Fig. 2 and Figure S2 this stage, this hypothesis still has to be experimentally validated.
confirm the hypothesised drop in the performances when real Overall, at the tested flow rates, real FU achieved OCV = 1.6
urine is used. Precisely, a 10–12% and a 90% drop in performances - 3.3 V, PDNet = 0.32 - 0.75 W.m−2 , ENet = 0.006 -
were measured with real and synthetic FU and HU respectively. 0.039 kWh.m−3 and overall energy efficiency η = 5 – 22% (as-
The two main explanation for the reduction in voltage and PD are: suming GFU,Max = 0.140 kWh.m−3 as per Eq. (3) and Table S1).
For real HU, OCV = 1.53 - 2.75 V, PDNet = 0.40 - 0.70 W.m−2 ,
1 The effect of organic fouling on the AEMs and spacers, which
ENet = 0.018 - 0.053 kWh.m−3 and η = 7 – 20% were obtained
was also observed by Rijnaarts et al. (2019), and
(GSU,Max = 0.262 kWh.m−3 ).
2 The higher ohmic resistance in real HU caused by its lower EC
The only other available benchmark to these results is from a
compared to synthetic HU. That is because the synthetic recipe
recent article from Mercer et al. (2019). Their experimental condi-
used to simulate HU overestimated its ionic strength as it dis-
tions, however, were quite different. Firstly, they operated in co-
regarded the NH3 losses due to volatilisation.
current mode and used an 0.25 M NaCl as the electrode rinse so-
Regarding the effect of organic fouling on the RED performance, lution. Additionally, the authors concentrated real HU twice (i.e.,
Rijnaarts et al. (2019) found a 15–20% decrease in RED perfor- from EC of 12.4 mS/cm to 24.1 mS/cm) using MD. As MD is
mances after operating it with natural river water for over 12 days. a thermal process and HU has over 30% of its nitrogen in the
Of that, only 2–4% was due to an increase in AEMs resistance and form of volatile NH3 , it is expected that some of the ammonia
decrease in its permeselectivity due to sorption of humic-like sub- would be lost in the MD permeate during the concentration pro-
stances. The remaining was due to spacer fouling. Even though our cess. The combination of all the above could explain why the au-
experiments were run only for a maximum of 8 h, the concentra- thors achieved lower performances (i.e., OCV ≈ 0.79 – 0.9 V and
tion of low molecular weight organic (LMW) and humic-like sub- PDGross ≈ 0.2 – 0.3 W.m−2 ).
stances in real urine is up to 50 times higher compared to river Finally, given the trade-off between power density and energy
water (Jacquin et al., 2018). As such, AEMs performance drop is recovery, this section aimed at suggesting an optimal operating
highly likely when using real urine. This was also confirmed by range to achieve relatively high PDNet and ENet . Overall, the re-
the results from the membrane autopsy, which is detailed in Sec- sults showed that low flow velocities, i.e. 2.5 – 7.5 cm/s are rec-
tion 3.3. ommended to obtain reasonable PD and ENet values.
Finally, the performances of real FU and HU were benchmarked
with the one of synthetic seawater (a standard RED benchmark. 3.1.2. Ionic transport
While, it is clear that urine achieved rather low PD compared to 3.1.2.3. Synthetic urine. The transport of the main species through
a 0.6 M NaCl solution (i.e., 60–70% less PDGross ), Fig. 2 also reveals the IEMs was firstly analysed at different linear velocities. Fig. 4
that, except for real HU, its OCV was similar. This could be ascribed shows a characterisation of the transport of different species when
to organic fouling, changing the surface charge of the membranes. synthetic FU (Fig. 4A) and synthetic HU (Fig. 4B) were used. Syn-
Moreover, the I-V curves for HU (both synthetic and real) display thetic solutions were used to eliminate the effect of membrane
6 F. Volpin, Y.C. Woo and H. Kim et al. / Water Research 186 (2020) 116320

Fig. 2. Performances of the RED stack operated with an 0.6 M NaCl solution (A), Synthetic FU and HU (B, C) and real FU and HU (D,E). HC and LC solutions were flowed
inside the stack in single-pass and with flow rates of 5, 10, 15, 25 and 35 mL.min−1 and K3 Fe(CN)6 / K4 Fe(CN)6 was used as ES.

Fig. 3. (A) shows the net power densities achieved (PDNet ), (B) displays the energy recovered ENet and energy efficiency η are displayed. HC and LC solutions were flowed
inside the stack in single-pass and with flow rates of 5, 10, 15, 25 and 35 mL.min−1 and K3 Fe(CN)6 / K4 Fe(CN)6 was used as ES.
F. Volpin, Y.C. Woo and H. Kim et al. / Water Research 186 (2020) 116320 7

Fig. 4. Single ion species transport from HC to LC at different flow velocities. In Fig. 4(A) synthetic FU was used while synthetic HU was used in Fig. 4(B).

Fig. 5. (A) shows the overall ions transport, based on electric conductivity measurements, from the HC to the LC solution at increasing urine flow rate. The overall nutrients
recovery, measured as the transport of nitrogen/phosphorus/potassium from urine to the LC solution, is displayed in Fig. 5(B).

fouling. As expected, the transport rate of multivalent ions (e.g., shows the overall transport of TN, PO4 3− and K+ (i.e., N/P/K). N, P,
SO4 2− , PO4 3− ) is 2 to 5 times lower than that of monovalent ions K analyses were chosen as it is hypothesised that the LC effluent
(e.g., Na+ , K+ , Cl− ). Interestingly, HU has higher Cl− and slightly could be used as fertigation.
lower Na+ and K+ transport compared to FU. This could be bal- Overall, by operating at 5.61 cm/s, which has been identified
anced by the more significant flow of NH4 + ions in HU, as they as the velocity that maximises the stack’s power densities in our
account for over 70% of all the cations in HU, compared to only experiments, about 30 ± 5% of nitrogen, 5 ± 2.1% for phosphorus
10% in FU. and 31 ± 4% for potassium were recovered in the LC solution.
By focusing on the total nitrogen transport, FU shows a lower
decrease in the TN transport at increasing flow velocities. This 3.2. Performance using urine as electrolyte solution
could be due to the high concentration of urea, which is un-
charged, whose transport is regulated solely by the concentration In this section, the outcomes from the experiments that em-
difference in the HC-LC channels as it follows solely the solution- ployed urine as ES are discussed. The principal goal of this config-
diffusion transport mechanism (Ma et al., 2018). As such, the flow uration was to compare the power densities achievable with urine
rate would have an effect on the urea transport only when it in- as ES and whether TOC or TN removal can occur at the electrode
creases the hydraulic retention time (HRT) enough to allow for a chambers.
higher CUrea inside the stack.
3.2.1. Energy recovery, energy efficiency and ionic transport
3.1.2.4. Real urine. Fig. 5(A) shows that the relative ionic transport Fig. 6 and Figure S5 show that despite only having a decrease
from the real urine to the LC solution. The widening gap between in OCV of just 10 - 20% compared to synthetic urine, the reduction
FU and HU results at increasing flow velocities could also be the in PDGross /PDNet and ENet in real urine was between 80 – 110%. Ad-
results of lower AEMs permeselectivity due to more severe foul- ditionally, a 10 - 50% decrease in ionic transport was also observed
ing deposition in AEMs fouled with HU (see Section 3.3). These (Fig. 7A).
EC data, however, provide only a piece of information on the elec- While the reduction of the OCV and ionic transport could
trolyte transport, while neglecting NH3 and urea. As such, Fig. 5(B) be explained by urine’s higher ohmic resistance, compared to
8 F. Volpin, Y.C. Woo and H. Kim et al. / Water Research 186 (2020) 116320

Fig. 6. Baseline performances of the RED stack tested with synthetic and real FU and HU solutions both as HC and ES solutions. HC and LC solutions were flowed inside
the stack in single-pass and with flow rates of 5, 10, 15, 25 and 35 mL.min−1 while ES was pumped with a flow rater of 5 mL.min−1 . Fig. 6(A) shows the OCV and stack
resistance while (B) and (C) PDGross and PDNet respectively. In Fig. 6(D), the energy recovered ENet and energy efficiency η are displayed.

Fig. 7. (A) shows the overall ions transport, based on electric conductivity measurements, from the HC to the LC solution at increasing urine flow rate. The overall nutrients
recovery, measured as the transport of nitrogen/phosphorus/potassium from urine to the LC solution, is displayed in Fig. 7(B).

K3 Fe(CN)6 /K4 Fe(CN)6 + NaCl, the decrease in power and energy re- probably because the transport of uncharged urea and NH3 from
covery was also due to the electromotive force and overpotentials the HC to the LC solution would not be affected by the increase in
of the oxidation and reduction reactions occurring at the ECs. This ohmic resistances as this is purely regulated by their concentration
is also confirmed by the reduction in NH4 + , urea and TOC concen- profile.
trations in Fig. 8. Overall, utilising urine as ES drastically decreases the energy ef-
As the relative ionic transport decreases compared to System A, ficiency and nutrients recovery of the process as it increases the
the recovery of K+ and PO4 3− from real urine was also found to stacks resistance and the voltage losses due to net chemical reac-
be about 2 – 5 times lower. Interestingly, though, the reduction in tions at the electrodes.
TN was only 1.4 and 1.7 times in FU and HU, respectively. That is
F. Volpin, Y.C. Woo and H. Kim et al. / Water Research 186 (2020) 116320 9

Fig. 8. Mass balance of the Cl− , TOC, NH4 + and urea in the urine used as electrolyte solution at different stack’s voltage and current. The mass balance was calculated as
per Section 2.3.3 and the ES was pumped with a flow rate of 5 mL/min. Figure A shows the results for real FU while real HU is shown in Figure B.

3.2.2. Change in the Cl− , NH4 + , urea and TOC in the electrolyte NH4 + is directly decomposed to N2 as per reaction below (Eq. 13)
solution (Candido and Gomes, 2011). In the case of urea, Eq. 14 was sug-
Here, urine’s mass balance results, before and after flowing gested as a possible degradation pathway (Ye et al., 2018).
through the ECs of the RED stack, are presented. The stack was
E θ =−0.517 V
always operated at the current densities that maximise the power 2NH4+ → N2 + 8H + + 8e− (9)
densities (Figure S5). The initial hypothesis was that the sharp re-
duction in the stack’s performances was a result of net chemical E θ =0.34 V
CO(N H2 )2 + H2 O → N2 + C O2 + 3H2 (10)
reactions occurring at the ECs.
The open circuit voltages as shown in Fig. 6A are in most Finally, TOC removal of up to 13% was measured in HU. At this
cases higher than the minimum voltage to achieve water split- stage, though, it is still unknown whether the mineralisation has
ting (i.e., 1.229 V) or anodic chloride formation coupled to cathodic occurred through direct electrochemical oxidation, indirect oxida-
hydrogen evolution (i.e., 1.358 V). However, that does not neces- tion due to water splitting and the formation of reactive oxygen
sarily mean that reactive oxygen species or free chlorine species species (ROS) or a combination of both. Also, the TOC removal in
were produced (Katsounaros et al., 2012; Tarpeh et al., 2018), as HU is over 6 times higher than FU. That might be because of the
activation overpotentials could have played a major role. Indeed, high concentration of LMW acids and neutrals (MW < 350 Da) in
Fig. 8 shows that the Cl− mass balance in the urine before and urine (62 ± 16% of the TOC) which are generally more efficient in
after the flowing through the anode and cathode is closed, indicat- alkaline media as it occurs with lower overpotential (Jacquin et al.,
ing no loss to chlorine gas. If true, this would be a positive result 2018; Munoz et al., 2015).
as Zöllig et al. (2015) demonstrated that the formation of Cl2 (gas) To conclude, the mass balance of NH4 + , urea and TOC showed
in urine can lead to the formation of dangerous chlorination by- that some degree of removal was achieved, which would also ex-
products which, if able to transfer to the LC solution, might ham- plain the lower stack’s performances when urine was used as ES.
per its reuse for fertigation. However, given the high reactivity of However, the redox reactions occurring at the anode and cathode
free chlorine and significant mass transfer in the system, Cl− is also compartments are still quite speculative and should be investigated
unlikely to be present in its free state but rather in a reacted state further. Additionally, it should be noted that, with the stack’s scale-
which would not have been detected by our analysis. up and optimisation, higher voltages are expected which could
Fig. 8 shows that NH4 + removal ranged from 0% - 5% and lead to the formation of chlorine, peroxides, persulfate species.
it was the highest when using HU. On the other hand, up to This, together with the effect of electrode material, HRT and cur-
4.4% of the initial urea was removed (i.e., 397 mg/L of urea) in rent density, should be optimised to achieve a higher degree of nu-
FU. In the case of ammonia, if no free Cl2 was produced, in- trients removal.
direct oxidation through monochloramine should not have oc-
curred (Candido and Gomes, 2011). One other hypothesis is In- 3.3. Characterisation of used AEMs and CEMs
stead the direct ammonia oxidation through NO2 − /NO3 − (Eq. (12))
(Candido and Gomes, 2011; Tarpeh et al., 2018). In fact, platinum- The high concentration of LMW organic acids, such as oxalic,
based electrodes, like the one used in this study, have often proven uric, citric, hippuric, lactic and acetic acid (Putnam, 1971b), in
to be an effective electro-catalyst for the oxidation of NH4 + to urine, makes it a solution with high fouling potential (Nam et al.,
NO3 − or N2 gas (Bunce and Bejan, 2011). 2018). This is particularly true for HU as during storage anaero-
pka =9.2 N O− ↔ N O− bic microorganisms can use the creatinine, fatty acids and other
N H4+ ⇔ N H3 ↔ N O ↔ 2 3 (8) organic acids in urine as an electron donor to grow and possibly
N2 or N2 O → N2
release high fouling potential polymeric substances (Udert et al.,
Also, it was found that the ammonia removal for HU was about 2006). The results of the interaction between organics and the
3 times higher compared to FU, and NO3 − and NO2 − concentrations AEMs are evident when looking at Fig. 9. There, the change in
of 4.5 ± 0.4 mg/L and 0.1 ± 0.04 mg/L were measured in HU ef- AEMs colour before and after testing the stack with real urine is
fluents but not in FU effluents. That is probably because of the quite evident (especially for HU). Given the negative charge density
high concentration of NH3 in HU, whose lone electron pair makes of LMW acids, it is expected that the positive fix charge in AEMs
it easier to be oxidised compared to NH4 + ion (Cheng et al., 2005). would allow them to migrate through and possibly accumulate on
While NH3 adsorbs onto the surface of the electrode before be- the membrane’s surface (Rijnaarts et al., 2019). This was confirmed
ing oxidised to N2 , at pH < 7 no adsorption/oxidation occurs and both by analysing the TOC transport to the LC effluent (Figure S3)
10 F. Volpin, Y.C. Woo and H. Kim et al. / Water Research 186 (2020) 116320

Fig. 9. Characterisation of the AEMs (top) and CEMs (bottom) before and after the tests with real fresh and hydrolysed urine. Each picture shows the pristine membrane at
the bottom, the membrane fouled with HU at the top and the membrane fouled with FU in the middle. Energy dispersive X-Ray spectra are displayed on the left-hand side
while FTIR spectra are shown on the right-hand side.

and by analysis FTIR analysis (Fig. 9). FTIR analysis, in particular, or the uphill transport of Mg2+ and Ca2+ to the urine, could
shows that in the AEMs after being in contact with FU and HU trigger carbonates and phosphates precipitation, especially in the
shows a new peak appears between 1040 and 1240 cm−1 wave- alkaline environment of HU.
length. This is generally attributed to the presence of C–O bonds
due to ethers, carboxylic acids and polysaccharides (Cho et al., 4. Conclusions
1998). On the other hand, there is no difference between the CEMs’
FTIR analyses. This suggests that only negatively charged low MW This study proved that RED could effectively be used to convert
organics were able to pass through the non-porous IMEs. the ionic gradient between urine and flushing water into poten-
Also, the EDX spectra for both CEMs and AEMs show no tial energy. The results showed that (i) a maximum ENet of 0.053
significant difference between pristine and fouled membrane, - 0.039 kWh per m3 of real urine can be achieved when utilising
indicating that phosphate scaling has not occurred during the homogenous redox couples as ES, while (ii) a TOC, ammonia and
short experimental period. This was expected as no significant urea removal of up to 13%, 6% and 4.4% can be obtained when real
increase in pressure drop was observed during the experiments. FU and HU are used as ES. On the other hand, in the experiments
Over time, however, the transport of PO4 3− to the tap water side, with K3 Fe(CN)6 /K4 Fe(CN)6 , relatively high OCV and power densities
F. Volpin, Y.C. Woo and H. Kim et al. / Water Research 186 (2020) 116320 11

were achieved. In fact, η ≈ 5 - 22% and PDNet ≈ 0.3 - 0.7 W.m−2 Lin, S., Straub, A.P., Elimelech, M., 2014. Thermodynamic limits of extractable energy
were obtained using real solutions. Interestingly, though, despite by pressure retarded osmosis. Energy Environ. Sci. 7 (8), 2706–2714.
Ma, L., Gutierrez, L., Vanoppen, M., Lorenz, D.N., Aubry, C., Verliefde, A., 2018. Trans-
its higher EC, HU performed similarly to FU. The principal expla- port of uncharged organics in ion-exchange membranes: experimental valida-
nation for that was the higher AEMs fouling with real HU, prob- tion of the solution-diffusion model. J. Memb. Sci. 564, 773–781.
ably due to sorption of LMW acids and humic substances on the Ma, P., Hao, X., Galia, A., Scialdone, O., 2020. Development of a process for the treat-
ment of synthetic wastewater without energy inputs using the salinity gradient
membrane surface. This was also confirmed by visually observing of wastewaters and a reverse electrodialysis stack. Chemosphere 248, 125994.
the darker colour of AEMs after flowing HU. Though, despite the Maurer, M., Pronk, W., Larsen, T.A., 2006. Treatment processes for source-separated
sorption of organics on the AEMs, no significant pressure drop was urine. Water Res. 40 (17), 3151–3166.
Mehdizadeh, S., Yasukawa, M., Suzuki, T., Higa, M., 2020. Reverse electrodialysis for
measured during the experiments. power generation using seawater/municipal wastewater: effect of coagulation
Overall, an important outcome from this work is that organ- pretreatment. Desalination 481, 114356.
ics and nitrogen removal can be achieved when urine is used as Mercer, E., Davey, C.J., Azzini, D., Eusebi, A.L., Tierney, R., Williams, L., Jiang, Y.,
Parker, A., Kolios, A., Tyrrel, S., Cartmell, E., Pidou, M., McAdam, E.J., 2019. Hy-
ES and RED driving force simultaneously. This opens the possibil- brid membrane distillation reverse electrodialysis configuration for water and
ity of designing a gravity-driven and decentralised urine treatment energy recovery from human urine: an opportunity for off-grid decentralised
unit to, partially, sanitise urine in-situ before it is discharged in the sanitation. J. Memb. Sci. 584, 343–352.
Munoz, F., Hua, C., Kwong, T., Tran, L., Nguyen, T.Q., Haan, J.L., 2015. Palladium—
combined sewage network.
copper electrocatalyst for the promotion of the electrochemical oxidation of
polyalcohol fuels in the alkaline direct alcohol fuel cell. Appl. Catal. B 174-175,
Declaration of Competing Interest 323–328.
Nam, J.-.Y., Hwang, K.-.S., Kim, H.-.C., Jeong, H., Kim, H., Jwa, E., Yang, S., Choi, J.,
Kim, C.-.S., Han, J., Jeong, N., 2018. Assessing the behavior of the feed-water
The authors declare that they have no known competing finan- constituents of a pilot-scale 10 0 0-cell-pair reverse electrodialysis with seawater
cial interests or personal relationships that could have appeared to and municipal wastewater effluent. Water Res..
influence the work reported in this paper. Post, J.W., Veerman, J., Hamelers, H.V.M., Euverink, G.J.W., Metz, S.J., Nymeijer, K.,
Buisman, C.J.N., 2007. Salinity-gradient power: evaluation of pressure-retarded
osmosis and reverse electrodialysis. J. Memb. Sci. 288 (1), 218–230.
Acknowledgements Putnam, D., 1971a. Composition and Concentrative Proprieties of Human Urine.
NASA, Huntington Beach, CA.
Putnam, D., 1971b. Composition and Concentrative Proprieties of Human Urine.
This research was supported by a grant (code NRF-
NASA, Washington, United States.
2017M1A2A2047369) from Climate Change Response Technology Randall, D.G., Krähenbühl, M., Köpping, I., Larsen, T.A., Udert, K.M., 2016. A novel
Development Program funded by National Research Foundation approach for stabilizing fresh urine by calcium hydroxide addition. Water Res.
of Korea, Republic of Korea, the National Research Foundation 95, 361–369.
Ray, H., Saetta, D., Boyer, T.H., 2018. Characterization of urea hydrolysis in fresh hu-
of Korean Grant funded by the Korean Government (MSIP) (No. man urine and inhibition by chemical addition. Environ. Sci. 4 (1), 87–98.
NRF-2015R1A5A7037825), and the Australian Research Council Rijnaarts, T., Moreno, J., Saakes, M., de Vos, W.M., Nijmeijer, K., 2019. Role of anion
(ARC) through Future Fellowship (FT140101208). exchange membrane fouling in reverse electrodialysis using natural feed waters.
Colloids Surf. A 560, 198–204.
Rose, C., Parker, A., Jefferson, B., Cartmell, E., 2015. The characterization of feces and
Supplementary materials urine: a review of the literature to inform advanced treatment technology. Crit.
Rev. Environ. Sci. Technol. 45 (17), 1827–1879.
Straub, A.P., Deshmukh, A., Elimelech, M., 2016. Pressure-retarded osmosis for power
Supplementary material associated with this article can be generation from salinity gradients: is it viable? Energy Environ. Sci. 9 (1), 31–48.
found, in the online version, at doi:10.1016/j.watres.2020.116320. Sydney Water, 2019. Water Analysis: Potts Hill Water Supply System.
Tarpeh, W.A., Barazesh, J.M., Cath, T.Y., Nelson, K.L., 2018. Electrochemical stripping
References to recover nitrogen from source-separated urine. Environ. Sci. Technol. 52 (3),
1453–1460.
Barbosa, S.G., Rodrigues, T., Peixoto, L., Kuntke, P., Alves, M.M., Pereira, M.A., Ter Udert, K.M., Larsen, T.A. and Gujer, W. 2006 Fate of major compounds in source-
Heijne, A, 2019. Anaerobic biological fermentation of urine as a strategy to en- separated urine, pp. 413–420.
hance the performance of a microbial electrolysis cell (MEC). Renew. Energy Veerman, J., Saakes, M., Metz, S.J., Harmsen, G.J., 2010. Electrical power from sea
139, 936–943. and river water by reverse electrodialysis: a first step from the laboratory to a
Bunce, N.J., Bejan, D., 2011. Mechanism of electrochemical oxidation of ammonia. real power plant. Environ. Sci. Technol. 44 (23), 9207–9212.
Electrochim. Acta 56 (24), 8085–8093. Veerman, J., Saakes, M., Metz, S.J., Harmsen, G.J., 2011. Reverse electrodialysis: a
Candido, L., Gomes, J.A.C.P., 2011. Evaluation of anode materials for the electro-oxi- validated process model for design and optimization. Chem. Eng. J. 166 (1),
dation of ammonia and ammonium ions. Mater. Chem. Phys. 129 (3), 1146–1151. 256–268.
Cheng, H., Scott, K., Christensen, P.A., 2005. Paired electrolysis in a solid polymer Vermaas, D.A., Kunteng, D., Saakes, M., Nijmeijer, K., 2013. Fouling in reverse elec-
electrolyte reactor—simultaneously reduction of nitrate and oxidation of ammo- trodialysis under natural conditions. Water Res. 47 (3), 1289–1298.
nia. Chem. Eng. J. 108 (3), 257–268. Volpin, F., Gonzales, R.R., Lim, S., Pathak, N., Phuntsho, S., Shon, H.K., 2018. Green-
Cho, J., Amy, G., Pellegrino, J., Yoon, Y., 1998. Characterization of clean and natural PRO: a novel fertiliser-driven osmotic power generation process for fertigation.
organic matter (NOM) fouled NF and UF membranes, and foulants characteriza- Desalination 447, 158–166.
tion. Desalination 118 (1), 101–108. Volpin, F., Yu, H., Cho, J., Lee, C., Phuntsho, S., Ghaffour, N., Vrouwenvelder, J.S.,
Han, J.-.H., Jeong, H., Hwang, K.S., Kim, C.-.S., Jeong, N., Yang, S., 2020. Asymmetrical Shon, H.K., 2019. Human urine as a forward osmosis draw solution for the ap-
Electrode System for Stable Operation of a Large-Scale Reverse Electrodialysis plication of microalgae dewatering. J. Hazard. Mater. 378, 120724.
(RED) System. Environmental Science: Water Research & Technology. WHO and UNICEF, 2017. Joint Monitoring Programme for Water Supply, Sanitation
Jacquin, C., Monnot, M., Hamza, R., Kouadio, Y., Zaviska, F., Merle, T., Lesage, G., and Hygiene - Annual Report. WHO/UNICEF.
Héran, M., 2018. Link between dissolved organic matter transformation and pro- Ye, K., Wang, G., Cao, D., Wang, G., 2018. Recent Advances in the Electro-Oxidation
cess performance in a membrane bioreactor for urinary nitrogen stabilization. of Urea for Direct Urea Fuel Cell and Urea. Electrolysis 376 (6), 42.
Environ. Sci. 4 (6), 806–819. Yip, N.Y., Elimelech, M., 2012. Thermodynamic and energy efficiency analysis of
Katsounaros, I., Schneider, W.B., Meier, J.C., Benedikt, U., Biedermann, P.U., Auer, A.A., power generation from natural salinity gradients by pressure retarded osmosis.
Mayrhofer, K.J.J., 2012. Hydrogen peroxide electrochemistry on platinum: to- Environ. Sci. Technol. 46 (9), 5230–5239.
wards understanding the oxygen reduction reaction mechanism. PCCP 14 (20), Yip, N.Y., Elimelech, M., 2014. Comparison of energy efficiency and power density in
7384–7391. pressure retarded osmosis and reverse electrodialysis. Environ. Sci. Technol. 48
Kim, H., Jeong, N., Yang, S., Choi, J., Lee, M.-.S., Nam, J.-.Y., Jwa, E., Kim, B., Ryu, K.-s, (18), 11002–11012.
Choi, Y.-.W., 2019. Nernst–Planck analysis of reverse-electrodialysis with the Zhang, J., She, Q., Chang, V.W.C., Tang, C.Y., Webster, R.D., 2014. Mining nutrients
thin-composite pore-filling membranes and its upscaling potential. Water Res. (N, K, P) from urban source-separated urine by forward osmosis dewatering.
165, 114970. Environ. Sci. Technol. 48 (6), 3386–3394.
Laktionov, E.V., Pismenskaya, N.D., Nikonenko, V.V., Zabolotsky, V.I., 2003. Method Zöllig, H., Remmele, A., Fritzsche, C., Morgenroth, E., Udert, K.M., 2015. Formation of
of electrodialysis stack testing with the feed solution concentration regulation. chlorination byproducts and their emission pathways in chlorine mediated elec-
Desalination 151 (2), 101–116. tro-oxidation of urine on active and nonactive type anodes. Environ. Sci. Tech-
Larsen, T.A., Hoffmann, S., Lüthi, C., Truffer, B., Maurer, M., 2016. Emerging solutions nol. 49 (18), 11062–11069.
to the water challenges of an urbanizing world. Science 352 (6288), 928–933.
Larsen, T.A., Udert, K.M., Lienert, J., 2013. Source Separation and Decentralization for
Wastewater Management. IWA Publishing.

You might also like