You are on page 1of 11

Journal of Water Process Engineering 49 (2022) 102949

Contents lists available at ScienceDirect

Journal of Water Process Engineering


journal homepage: www.elsevier.com/locate/jwpe

Water and green ammonia recovery from anaerobic digestion effluent by


two-stage membrane distillation
Mingfei Shi, Man Xiao, Liang Feng, Te Tu, Qingyao He *, Shuiping Yan *
College of Engineering, Huazhong Agricultural University, No.1, Shizishan Street, Hongshan District, Wuhan 430070, PR China
Technology & Equipment Center for Carbon Neutrality in Agriculture, Huazhong Agricultural University, Wuhan 430070, PR China

A R T I C L E I N F O A B S T R A C T

Keywords: A water and green ammonia recovery strategy for anaerobic digestion (AD) effluent treatment through a two-
Ammonia removal stage membrane distillation (2s-MD) was put forward in this study. In the 1st-stage, the MAP (mono­
Water recovery ammonium phosphate solution, NH4H2PO4)/DAP (diammonium phosphate solution, (NH4)2HPO4) loop, was
Recyclable solvent
used as the recyclable ammonia absorbent to recover ammonia from AD effluent. The multicycle ammonia
Membrane distillation
Anaerobic digestion effluent
absorption-regeneration experiments illustrated that the ammonia removal efficiency was stable (~81 %) and
the recovered ammonia solution was 1.05 mol/L. In the 2nd-stage, the water recovery experiments from AD
effluent with different ammonia concentrations were conducted. The results indicated that the water recovered
from ammonia-removed AD effluent could be adopted as the general industrial or agricultural water. Finally, the
economic evaluation demonstrated that the application of MAP solution as the recyclable ammonia absorbent
was more economically efficient than the sulfuric acid solution. The treatment cost and net profit of AD effluent
treatment by 2s-MD were $ 1.73/m3 and $ 0.83/m3, respectively. The 2s-MD can be acted as a valuable
candidate strategy for sustainable green ammonia and water recovery from AD effluent.

1. Introduction concentrate the nutrients such as ammonia nitrogen and phosphate


simultaneously, followed by the separation of ammonia nitrogen from
Anaerobic digestion (AD) has been extensively applied for livestock the concentrated AD effluent after pH adjustment by adding the alkali.
and poultry waste treatment by converting the biodegradable organic This procedure may be chemical-intensive because of the good acid-base
pollutants into organic fertilizers and biogas [1]. The liquid effluent buffering performance of AD effluent [7]. Another procedure is
discharged from the AD digester has a high content of nutrients, such as ammonia nitrogen recovery from the AD effluent firstly and then water
ammonium nitrogen, phosphate, and potassium [2]. Especially, the total recovery from the treated AD effluent ensued. Ideally, ammonia transfer
ammonia nitrogen (TAN) concentration may be higher than 5000 mg-N/ from the AD effluent into the reclaimed water can be inhibited for
L [3]. Generally, the amount of liquid effluent accounts for >80 % of the achieving an improved water quality in this procedure [8], which makes
total AD residues [4]. Given the high organic loading and nutrient it a promising approach for AD effluent treatment.
concentration of AD effluent [5], it is vital to seek the appropriate Known as a typical thermally-driven process, membrane distillation
treatment method for AD effluent [6,7]. Predictably, if 6.2 billion m3 of (MD) is a promising membrane technology for ammonia and water re­
AD effluent produced annually in China can be treated efficiently [5], covery from AD effluent because of the high volatility of ammonia and
about 3.1–31 million tons of nitrogen and about 5.89 billion m3 of water [9–11]. Compared with these membrane processes including
reclaimed water may be recovered. Otherwise, the potential environ­ forward osmosis (FO), reverse osmosis (RO), nanofiltration (NF), and
mental pollution may be caused by the direct application of these AD microfiltration (MF), MD operates at a lower hydrostatic pressure,
effluents [3]. Obviously, recovering useful resources like ammonia ni­ enabling fouling resistance. The water recovery process via MD is not
trogen and water from the AD effluent is essential. limited by the concentration degree of treated wastewater. In addition,
Two possible procedures have been developed for recovering the it is especially beneficial when the low-grade or waste heat is available
ammonia nitrogen and water efficiently from the AD effluent. The first in the MD process [7,12–15]. As for ammonia recovery, plenty of
procedure is to recover water from the acidified AD effluent and technologies such as ion exchange [16], nitrification-denitrification

* Corresponding authors at: College of Engineering, Huazhong Agricultural University, No.1, Shizishan Street, Hongshan District, Wuhan 430070, PR China.
E-mail addresses: qingyao.he@mail.hzau.edu.cn (Q. He), yanshp@mail.hzau.edu.cn (S. Yan).

https://doi.org/10.1016/j.jwpe.2022.102949
Received 22 February 2022; Received in revised form 26 May 2022; Accepted 14 June 2022
Available online 1 July 2022
2214-7144/© 2022 Published by Elsevier Ltd.
M. Shi et al. Journal of Water Process Engineering 49 (2022) 102949

[16], chemical precipitation [17], and air stripping [18] have been multicycle experiments were conducted to explore the stability of
investigated. However, the ammonia recovery efficiency was not satis­ ammonia recovery using the MAP/DAP loop. After ammonia recovery,
fied when treating AD effluent with the technologies mentioned above in extensive investigations were performed in 2nd-stage MD to recover
terms of a combination of economic, environmental, and energy con­ water from the treated AD effluent with different ammonia concentra­
siderations [19–21]. Comparatively, MD seems a promising approach tions in terms of the water quality. Finally, a techno-economic evalua­
for fast ammonia recovery from AD effluent [9], wherein the direct tion was performed to provide a deep insight into the sustainable green
contact membrane distillation (DCMD) [22,23], vacuum membrane ammonia and water recovery from AD effluent by the 2s-MD process.
distillation (VMD) [24,25], and sweeping gas membrane distillation
(SGMD) processes [26] have been extensively experimented. Typically, 2. Materials and methods
due to the low free ammonia content, the pH increment of feed AD
effluent greatly relies on the consumption of alkali [9,11]. On the 2.1. Materials
permeate side, acid solutions are always used for ammonia recovery
[27]. To reduce the alkali consumption, vacuum or aeration was applied AD effluent was collected from a large-scale biogas plant using pig
on the feed side [9,28]. However, the ammonia removal performance manure as the substrate (digestion temperature: ~35 ◦ C) located at
needs to be improved. To minimize the acid solutions consumption, Wuhan City, Hubei Province, China, and it was stored anaerobically at
VMD is put forward to recover ammonia, but the separation factor is too the ambient temperature until no biogas was produced. The AD effluent
low to be applied in industry [9]. During the ammonia recovery, the was centrifuged at 4000 rpm for about 20 min to remove particles and
chemical consumption cost accounts for >48 % of the operating cost [9], suspended solids before the experiments. Characteristics of AD effluent
making the development of new technology featured with a low are listed in Table 1. Each liquid sample was measured at least three
chemical consumption and a satisfied ammonia recovery performance times to determine its average values and standard deviations. The
important. specific test methods for the samples can be found in the Supporting
To solve these problems as mentioned above, a novel two-stage Information file. The monoammonium phosphate solution (MAP,
membrane distillation (2s-MD) process is put forward in this study. As NH4H2PO4) and diammonium phosphate solution (DAP, (NH4)2HPO4),
shown in Fig. 1, ammonia is recovered from AD effluent in the 1st-stage with a purity content of >99 %, were purchased from Sinopharm Group
MD, and the water is recovered from the ammonia-removed AD effluent Chemical Reagent Co., Ltd., China.
in the 2nd-stage MD. Especially, a DCMD-based green ammonia recovery
strategy is developed in the 1st-stage MD, wherein recyclable ammonia 2.2. Experimental setup
receiving solution is adopted to get rid of the acid consumption required
in the conventional DCMD process [29], and aeration is also combined Fig. 2 shows the experimental setup for water and green ammonia
to improve the ammonia recovery efficiency other than the exogenous recovery from AD effluent based on the 2s-MD process. In Fig. 2, the
alkali addition [30]. In this strategy, the monoammonium phosphate green part represents the green ammonia recovery process (i.e., the 1st-
(MAP, NH4H2PO4) solution is used as a recyclable ammonia receiving stage MD), and the blue part represents the water recovery process (i.e.,
solution. According to Eq. (1), MAP has an affinity toward free ammonia the 2nd-stage MD). During the single MD process, the experimental setup
from AD effluent to form the ammonia-rich MAP, i.e., DAP. Through the is composed of a circulation system of AD effluent on the feed side, a
thermal regeneration of DAP, as reported by Rice and Busa [31], the circulation system of receiving solution on the permeate side, and a
aqueous ammonia solution with a concentration of 15–25 wt% can be hollow fiber membrane contactor. For ammonia recovery in the 1st-stage
achieved. Ideally, the MAP/DAP loop may be the good ammonia MD, the real lab-scale diagram of aeration-assisted DCMD process is
absorbent for recovering ammonia from AD effluent by DCMD. To shown in Fig. A.1. An air pump (Chongqing Yanlong Machinery
achieve a sustainable AD effluent treatment by the novel 2s-MD process, Equipment Co., Ltd., China) is placed in front of membrane contactor,
2 key problems should be investigated including the feasibility of
adopting MAP solution as the stable ammonia absorbent for high-
Table 1
performance ammonia recovery in the 1st-stage MD and the technical Characteristics of the centrifuged anaerobic digestion effluent used in this study.
feasibility of clean water recovery from the alkaline AD effluent after
Parameter Values Units
ammonia removal in the 2nd-stage MD.
pH 8.105 ± 0.007 –
absorbtion Electrical conductivity (EC) 7.185 ± 0.007 mS/cm
NH4 H2 PO4 + NH3 (NH4 )2 HPO4 (1)
desorption Chemical oxygen demand (COD) 770.84 ± 4.49 mg/L
CO2 loading 0.058 ± 0.0044 mol-CO2/L
In this study, ammonia recovery performance was experimentally Total ammonia nitrogen (TAN) 3.86 ± 0.12 g-N/L
investigated in the 1st-stage MD by using MAP solution, and then Total solid (TS) 3.396 ± 0.32 g/L
ammonia desorption performance was also evaluated. Furthermore, Volatile fatty acids (VFA) 76 ± 13 mg/L

1st-stage MD for
Ammonia recovery 2nd-stage MD for Water
recovery

High TAN-containing Ammonia absorption Low TAN-containing AD Concentrated AD


AD effluent effluent H2O effluent
MAP DAP
(NH4H2PO4) ((NH4)2HPO4)

Ammonia desorption

NH3·H2O

Fig. 1. Schematic diagram of green ammonia and water recovery from anaerobic digestion (AD) effluent through two-stage membrane distillation (2s-MD).

2
M. Shi et al. Journal of Water Process Engineering 49 (2022) 102949

Fig. 2. Experimental setup for water and green ammonia recovery from anaerobic digestion (AD) effluent by two-stage membrane distillation (2s-MD). (For
interpretation of the references to color in this figure, the reader is referred to the web version of this article.)

and the air flow rate is fixed at 0.6 L/min monitored by a rotameter 2.3. Data analysis
(LZB-3, Nanjing Shunlaida Co., Ltd., China). To reduce the membrane
plugging risks and promote the aeration on the feed side, the AD effluent 2.3.1. Ammonia absorption performance at the ammonia absorption stage
flows through the shell side of membrane contactor while the receiving The overall mass transfer coefficient of ammonia (Kov, m/s) and
solution passes through the lumen side. If MAP is used as the receiving ammonia flux (JA, g/m2h) are the important factors to measure the
solution for ammonia recovery, the green aqueous ammonia solution ammonia separation performance, which can be determined in accor­
can be obtained after the ammonia regeneration from the saturated dance with the following equations [9,32]:
ammonia absorbent. For water recovery in the 2nd-stage MD, the AD ( A)
VA C
effluent flows through the lumen side of the membrane contactor while KOV = 0 ln 0A (2)
the receiving solution, i.e., DI water, flows through the shell side. The AΔt Ct
hydrophobic microporous polypropylene (PP) membrane contactors
C0A V0A − CtA VtA
(Hangzhou Jiefu Membrane Technology Co., Ltd., China) are used for JA = (3)
water and green ammonia recovery. The properties of membrane con­ AΔt
tactors are listed in Table A.1 in the Supporting Information file. where CA A
0 and Ct (g/L) are the TAN concentrations of AD effluent at the
To explore the feasibility of using MAP as the recyclable ammonia initial and any time t (h) in the ammonia recovery process, respectively;
absorbent, the ammonia absorption performance of ammonium phos­ VA A
0 and Vt (L) are the volumes of AD effluent at the initial and any time t
phate solution with different ammonia loadings (α, mol/mol, defined by (h) in the ammonia recovery process, respectively; A is the effective
the ratio of total ammonia nitrogen concentration to phosphate con­ membrane area (m2).
centration) was investigated. The ammonia mass transfer coefficient was The experimental TAN concentration profile of AD effluent during
also tested using AD effluent with and without pH adjustment. Impor­ ammonia removal is fitted to an exponential decay curve as shown in Eq.
tantly, ammonia regeneration performance of ammonia-rich solution (4) for calculating the pseudo-first-order kinetics constant of TAN
(DAP) was conducted, including the effects of key operating parameters removal (k, h− 1) [33]. As shown in Eq. (5), the time constant (τ, h)
like temperature, system pressure, and regeneration time. Finally, the ( )
represents the time required to reduce the TAN concentration by 1 − 1e
multicycle experiments of ammonia absorption-regeneration were per­
(~63 % of the initial TAN concentration), allowing a quantitative
formed to explore the feasibility of the MAP/DAP loop as the recyclable
comparison between experimental results at different initial and final
ammonia absorbent.
conditions [33].
In the 2nd stage MD section, the technical feasibility of clean water
recovery from the alkaline AD effluent after ammonia removal was CtA = C0A e− kt
(4)
investigated. AD effluents discharged from the 1st-stage MD with
different ammonia concentrations were used as the feed solution in the 1
k= (5)
DCMD process. TAN concentrations of AD effluents were set at 4.3 g-N/ τ
L, 1.5 g-N/L, 0.6 g-N/L, or 0.003 g-N/L for mimicking the treated AD During ammonia absorption from AD effluent using MAP and H2SO4
effluents with different ammonia removal performance in the 1st-stage in the 1st-stage MD, ammonia removal efficiency (Rm, %), ammonia
MD. The evaluation parameters of the recovered water, including the recovery efficiency (Rc, %), and ammonia loss ratio (η, %) are adopted,
TAN concentration, EC value, and ions concentrations, were detected. which can be determined using the following Eqs. (6)–(8). The ammonia
The detailed experimental conditions were presented in the Supporting removal efficiency is used to represent how much ammonia is released
Information file. from the AD effluent during ammonia absorption, while ammonia re­
covery efficiency represents how much ammonia is captured by
ammonia absorbent. The ammonia loss ratio then represents how much
ammonia released from the AD effluent is not captured by ammonia

3
M. Shi et al. Journal of Water Process Engineering 49 (2022) 102949

absorbent. Clearly, a high ammonia absorption ratio leads to a low η ammonia absorbents to capture ammonia from the pH adjusted AD
value. effluent in the 1st-stage MD. The results indicate that whatever ammonia
absorbent is adopted, the ammonia removal efficiency (Rm, %) in­
C0A V0A − CtA VtA
Rm = × 100% (6) creases, while the ammonia flux (JA, g/m2h) decreases with the increase
C0A V0A
of operation time (Fig. 3). The ammonia removal efficiency can reach
nNH3 ,p above 98 % after 12 h of operation by using both MAP and sulfuric acid
Rc = × 100% (7) as the absorbent. However, the total TAN concentration in the AD
C0A V0A
effluent decreases with the increase of operation time, resulting in the
( ) reduction of free ammonia content (i.e., reduction in ammonia transfer
Rc
η = 1− × 100% (8) driving force), and consequently the drop in ammonia flux [19].
Rm
Notably, there are no significant differences in the ammonia removal
where nNH3, p is the amount of ammonia absorbed by ammonia absorbent efficiency and ammonia flux between the MAP and H2SO4 cases, illus­
on the permeate side. trating the comparable ammonia absorption performance between MAP
and H2SO4. As shown in Fig. 3b, when H2SO4 is adopted as the ammonia
2.3.2. Ammonia regeneration performance at the ammonia desorption absorbent, no ammonia loss is observed during ammonia recovery.
stage Comparatively, the ammonia loss ratio is ranged from about 9 % to 2.3
With the proceeding of ammonia absorption by MAP, the receiving % in MAP case in the 12-h experiment process, implying that some
solution gradually transforms from the initial MAP to DAP, showing that ammonia released from the AD effluent may not be completely captured
the ammonia loading (α, mol/mol) increases. In this study, the ammonia by MAP. Hence, MAP has a relatively lower ammonia absorption rate
loading is defined by the molar concentration ratio of total ammonia than that of H2SO4. Fortunately, the overall ammonia loss ratio can be
nitrogen (mol/L) to the total phosphate concentration (mol/L). Clearly, controlled below 2 %, implying a good ammonia recovery performance.
the ammonia loadings of MAP and DAP are 1.0 and 2.0 mol/mol, The ammonia recovery performance of MAP from AD effluents
respectively. So according to the change of ammonia loading, the without pH adjustment and with pH adjustment by aeration or alkali
ammonia regeneration efficiency (ε, %) from the ammonia-rich phos­ addition are shown in Fig. 4. The mass transfer coefficient of ammonia
phate solution can be evaluated by the following equation [29]: (Kov, m/s) in the alkali addition case is the highest, followed by the
aeration case, and then the case without pH adjustment (Fig. 4a).
α0 − αt
ε= × 100% (9) Furthermore, the change of kinetics constant (k, h− 1) of TAN removal
α0 − 1
versus pH adjustment method shows the same trend with Kov value [33].
where α0 is the initial ammonia loading (mol/mol) of ammonia-rich These results can be expected because the AD effluent adjusted by
phosphate solution; αt is the ammonia loading (mol/mol) of ammo­ adding alkali can obtain the highest initial pH value, meaning the
nium phosphate solution obtained experimentally at any time t (h); 1 is highest free ammonia content in the AD effluent at the same tempera­
the ammonia loading (mol/mol) of MAP solution. ture, and consequently highest ammonia transfer driving force [34,35].
Additionally, the pH value of AD effluent assisted by aeration is higher
2.3.3. Water recovery performance than that of AD effluent without adjustment. This is attributed to the fact
Water flux (JW, kg/m2h) is used to assess the water recovery per­ that CO2 can be easily released from AD effluent under an aeration
formance from the treated AD effluent by the 2nd-stage MD, which is condition, and the chemical equilibrium between NH+ 4 and HCO3 is

expressed according to the following equation: broken to improve the pH value and free ammonia content [36,37].
Compared with the alkali addition, the aeration assistance can not only
JW =
ΔmW
(10) reduce the chemical consumption but also mitigate the membrane
AΔt fouling for AD effluent treatment by the MD process [38]. Hence,
aeration may be a sustainable approach to enhance ammonia recovery
where A (m2) is the membrane area; ΔmW (kg) is the quantity of water
performance.
transferred across the membrane during the time Δt (h).
As shown in Fig. 4(b), in 12 h of operation, the average ammonia loss
The ammonia rejection efficiency (β, %) is adopted for showing how
ratio is observed to be about 4.2 %, 11.5 %, and 0.5 % for alkali addition,
much ammonia (i.e., TAN) can be retained in the AD effluent without
aeration, and no pH adjustment groups, respectively. The lowest
being transferred into the recovered water. The ammonia rejection ef­
ammonia loss ratio for the no pH adjustment group (0.5 %) might be
ficiency is calculated according to the following equation:
mainly due to the lower free ammonia content in AD effluent, and
CtW VtW therefore the least ammonia volatilization. The high ammonia loss ratio
β= × 100% (11)
C0W V0W in the aeration case (11.5 %) is mainly due to the ammonia escaping into
the air [28]. Certainly, this ammonia loss ratio can be further decreased
where CW W
0 and Ct are the TAN concentrations (g/L) of AD effluent at the through operating parameters optimization. If the TAN removal per­
initial and time t (h) in the water recovery process, respectively; VW
0 and formance from AD effluent and ammonia loss ratio are both considered,
VWt (L) are the volumes of AD effluent at the initial and any time t (h) in aeration can also be accepted during ammonia removal process to get rid
water recovery process, respectively. of the chemical's consumption [22].
The total ammonia nitrogen (TAN) concentration (Cp, g-N/L) of the The effects of ammonia loading of receiving solution (α, mol/mol) on
recovered water is calculated according to the following equation: ammonia recovery performance are shown in Fig. 5. When the α value is
<2.0 mol/mol, ammonia can be successfully removed from AD effluents,
C0W V0W − CtW VtW
Cp = (12) and the ammonia removal efficiency decreases with the increase of the α
C0W V0W
value as shown in Fig. 5(a). Especially when α = 2.0 mol/mol meaning
DAP is adopted, no ammonia removal happens from AD effluent,
3. Results and discussion
implying the extremely low ammonia absorption rate by DAP. More­
over, when the α value increases from 1.0 mol/mol to 1.8 mol/mol, the
3.1. Green ammonia recovery performance
corresponding time constant increases sharply from 3.94 h to 12.34 h by
about 213.2 %, and the corresponding mass transfer coefficient of
3.1.1. Ammonia absorption performance of MAP
ammonia (Kov) decreases sharply from 6 × 10− 7 m/s to 1.91 × 10− 7 m/s
Aqueous MAP and sulfuric acid (H2SO4) solutions were used as the
by about 68 % (Fig. 5b). Obviously, the free active MAP molecules

4
M. Shi et al. Journal of Water Process Engineering 49 (2022) 102949

(a) 100 10
(b) 12

Ammonia removal efficiency

Ammonia loss ratio (η, %)


80 8

NH3 flux (JA, g/m2h)


MAP
MAP
9
MAP
MAP H2SO4
H₂SO₄
60 6
(Rm, %)
H2SO4
H2SO4
6
MAP
MAP
40 4
H2SO4
H2SO4
3
20 2

0 0 0
0 2 4 6 8 10 12 0 2 4 6 8 10 12
Time (h) Time (h)

Fig. 3. Ammonia absorption performance including (a) ammonia removal efficiency and transfer flux, (b) ammonia loss ratio in MAP and H2SO4 cases (Experimental
conditions: AD effluent: 3.86 g-N/L initial TAN concentration, 11.0 initial pH value adjusted by adding NaOH, 60 ◦ C, 112 mL/min flow rate; Ammonia absorbent:
30 ◦ C, 10 mL/min flow rate, 400 mL volume, 1 mol/L, and 1.0 mol/mol ammonia loading for MAP, 0.8 mol/L for H2SO4).

(a) 8 0.4 (b) 15


Kov
Kov 11.5%

Ammonia loss ratio (η, %)


Ammonia transfer coefficient

kkvalue
value 12
6 0.3
k value (h-1)
(Kov, h10-7 m/s)

9
4 0.2
6
4.1%
2 0.1
3
0.7%
0 0 0
No adjustment Alkali addition Aeration No adjustment Alkali addition Aeration
pH adjustment methods pH adjustment methods
Fig. 4. Effects of pH adjustment methods including no adjustment, NaOH addition and aeration on (a) ammonia transfer coefficient (Kov, m/s) and k value (h− 1), (b)
ammonia loss ratio (Experimental conditions: AD effluent: 3.86 g-N/L TAN concentration, 8.1 initial pH value for no pH adjustment case and aeration case with 0.6
L/min air flow rate, 11.0 pH value for NaOH addition case, 60 ◦ C, 112 mL/min flow rate; Ammonia absorbent: MAP with 400 mL volume, 1 mol-PO3− 4 /L con­
centration and 1.0 mol/mol ammonia loading, 30 ◦ C, 10 mL/min flow rate).

(a) 4 (b) 7 Kov


15
Kov ττ
TAN content of AD effluent (g/L)

Ammonia transfer coefficient

6
12
3 5
(Kov, h10-7 m/s)

4 9
τ (h)

2 3
α=1 mol/mol 6
α=1.2 mol/mol
2
α=1.4 mol/mol
1 3
α=1.6 mol/mol 1
α=1.8 mol/mol
α=2 mol/mol 0 0
0 1 1.2 1.4 1.6 1.8
0 1 2 3 4 5 6
Time (h) Ammonia loading of absorbents (α, mol/mol)

Fig. 5. Effect of different Ammonia loadings of ammonium phosphate solution on ammonia removal performance from AD effluent in MD process. (a) TAN content
profile of AD effluent. (b) Ammonia transfer coefficient (Kov, m/s) and the time constant (τ, h) (Experimental conditions: AD effluent: 3.86 g-N/L TAN concentration,
11.0 initial pH value adjusted by adding NaOH, 60 ◦ C, 112 mL/min flow rate; Ammonia absorbent: 30 ◦ C, 10 mL/min flow rate, 1 mol-PO3− 4 /L concentration).

available to react with ammonia in the ammonium phosphate solution Additionally, the increased α value also results in the increase of
decrease with the increased α value, causing the decrease in ammonia ammonia partial pressure over the ammonium phosphate solution and
absorption rate, thus leading to the increase in time constant. then the decrease of ammonia transfer driving force, consequently the

5
M. Shi et al. Journal of Water Process Engineering 49 (2022) 102949

decrease of Kov and the increase of time constant. When the α value of explained by fact that the increment of water vapor flux is higher than
receiving solution is above 1.6 mol/mol, the ammonia removal perfor­ that of ammonia flux [32]. Obviously, a high regeneration temperature
mance reduces greatly (Fig. 5). The results suggest that the α value of should be recommended for achieving a high ammonia regeneration
ammonium phosphate solution should be <1.6 mol/mol in the real performance and a high concentration of green ammonia formed.
applications to achieve good ammonia removal performance. This can A reduction in vacuum distillation pressure contributes to a drop of
provide technical guidance for the subsequent ammonia recovery pro­ ammonia partial pressure in the gas phase, showing an increase in the
cess using ammonium phosphate solution as the ammonia absorbent. mass transfer driving force of ammonia regeneration at a fixed tem­
perature [39], and thus a rise in ammonia regeneration efficiency as
3.1.2. Ammonia regeneration performance of NH3-rich absorbent shown in Fig. 6(a) and (b). For example, when DAP is regenerated at
Practically, after ammonia absorption from the AD effluent, the 90 ◦ C for 10 min, the ammonia regeneration efficiency increases greatly
formed ammonia-rich absorbent will be regenerated to generate the from 4.14 % for 60 kPa case to 27.73 % for 10 kPa case by about 569.81
green ammonia and ammonia-lean absorbent for capturing ammonia %. As shown in Fig. 6(b), the TAN concentration of obtained aqueous
again. So, it is necessary to investigate the ammonia regeneration per­ ammonia solution increases from 0.65 mol/L to 12.94 mol/L when the
formance of ammonia-rich absorbents. In this study, the ammonia- vacuum distillation pressure increases from 10 kPa to 60 kPa. This is
saturated absorbent, i.e., DAP, was investigated preliminarily by the mainly because an increase in vacuum distillation pressure can cause a
lab-scale vacuum distillation in terms of the ammonia desorption reduction degree of water vapor transfer flux higher than that of
performance. ammonia flux [24]. Notably, although the TAN concentration of ob­
As shown in Fig. 6, ammonia can be successfully regenerated from tained aqueous ammonia solution can reach up to above 12 mol/L when
DAP to form the green aqueous ammonia solution by vacuum distilla­ DAP is regenerated at >50 kPa and 90 ◦ C, the ammonia loading of
tion, and the ammonia regeneration performance changes with the ammonia-lean absorbent is still too high (>1.9 mol/mol) to capture
operating conditions. With an increase in regeneration temperature at ammonia again. If both a high ammonia regeneration efficiency and a
fixed system pressure, the ammonia regeneration performance of DAP high TAN concentration of obtained aqueous ammonia solution are
rises represented by the elevated ammonia regeneration efficiency targeted, 90 ◦ C and 40 kPa may be the better conditions for DAP
(Fig. 6a and b). This can be expected because a high reaction tempera­ regeneration in this study.
ture can directly cause an increase in the ammonia regeneration rate, As shown in Fig. 6(c), extending the regeneration time is beneficial to
and then a rise in the ammonia partial pressure over the DAP solution (i. achieving a high ammonia regeneration performance, while it is not
e., an elevation in ammonia regeneration driving force), consequently conducive to obtaining a high concentration of obtained aqueous
an improved ammonia regeneration efficiency [9]. However, the TAN ammonia solution. Under the conditions of 90 ◦ C, 40 kPa, and 90 min,
concentration of obtained aqueous ammonia solution decreases with the the ammonia regeneration efficiency of DAP is 81.78 %, and the TAN
increase of temperature as shown in Fig. 6(a) and (b), which can be concentration of obtained aqueous ammonia solution is about 1.03 mol/

(a) 30 (b) 30
80℃ 10 min 12 90℃ 10 min
Ammonia regeneration efficiency

12
Ammonia regeneration efficiency

TAN concentration (mol/L)


TAN concentration (mol/L)

25 25
10 10
20 20
8 8
(ε, %)

(ε, %)

15 15
6 6

10 4 10 4

5 2 5 2

0 0 0 0
10 20 30 10 20 30 40 50 60
System pressure (kPa) System pressure (kPa)

(c) 100
2.5
90℃ 40 kPa
Ammonia regeneration efficiency

TAN concentration (mol/L)

75 2

1.5
(ε, %)

50

25
0.5

0 0
10 30 60 90
Desorption time (min)

Fig. 6. Effects of temperature, pressure, operation time on the ammonia regeneration efficiency of DAP (ε, %) and the TAN concentration of the recovered aqueous
ammonia solution (mol/L) (Experimental conditions: DAP solution: 180 mL volume, 1 mol/L and 2.0 mol/mol ammonia loading; (a) 80 ◦ C desorption temperature,
10 min desorption time; (b) 90 ◦ C desorption temperature, 10 min desorption time; (c) 90 ◦ C desorption temperature, 40 kPa desorption pressure).

6
M. Shi et al. Journal of Water Process Engineering 49 (2022) 102949

L. Certainly, there is still room for further optimizing the ammonia 1st reg. 2nd reg.
(a) 3rd reg. 4th reg. 5th reg.
regeneration process by various means, such as increasing the DAP
concentration and selecting a low-energy consumption reactor.
2

3.1.3. Ammonia absorption-regeneration performance


Five sequential ammonia absorption-regeneration cycles were 1.8

distillate (α, mol/mol)


investigated to confirm the stability of green ammonia recovery using

Ammonia loading of
the MAP/DAP loop. In the 1st cycle, DAP was regenerated to generate
5th abs.
the ammonia-lean solvent in the first ammonia regeneration, followed 1.6 1st abs. 2nd abs. 3rd abs.
by ammonia absorption from the AD effluent using this ammonia-lean 4th abs.
solvent. After ammonia absorption from AD effluent, the ammonia-
rich solvent was regenerated to release the ammonia (i.e., the recov­ 1.4
ered green ammonia), and the lean solvent was recycled again for Average α=1.52 mol/mol
ammonia recovery, as shown in Fig. 7.
As shown in Fig. 7(a), at every regeneration cycle, the ammonia 1.2
loading of ammonia-rich solvent can be reduced to about 1.52 mol/mol,
showing a relatively stable ammonia regeneration efficiency (36.45 %–
49.24 %) as shown in Fig. 7(b) and good stability of ammonia regen­ 1
eration stage. Similarly, at every absorption cycle, the ammonia loading 100

Ammonia removal efficiency


increases to about 1.9–2.0 mol/mol, indicating a stable ammonia ab­ Average Rm=81%
sorption performance, which can be seen from the stable ammonia
removal efficiency (~81 %) from the AD effluent (Fig. 7a). In addition, 80
as shown in Fig. 7(b), the TAN concentration of the recovered ammonia
solution varied slightly from 0.94 mol/L to 1.23 mol/L (averaged at

(Rm, %)
60
1.05 mol/L) during the five ammonia absorption-regeneration cycles.
The renewable aqueous ammonia can be used as a biomass pretreatment
40
additive for energy production [40], industrial feedstock for chemical
production [41], or chemical absorbent for biogas upgrading [42],
increasing the prevalence of applications and reducing the carbon 20
footprint [43].
Cycle 1 Cycle 2 Cycle 3 Cycle 4 Cycle 5
Clearly, the MAP/DAP loop is suitable as the recyclable ammonia
0
absorbent in the 1st-stage MD process to recover the green ammonia
0 6 12 18 24 30
with a stable concentration from the AD effluent in a long-term opera­
tion. Moreover, compared with the typical method for recovering the Time (h)
green aqueous ammonia solution from the AD effluent, i.e., the vacuum
membrane distillation [9], the 1st-stage MD process adopting the recy­
100
(b)

TAN content of recovered aqueous


clable ammonia absorbent (MAP/DAP loop) provided in this study can Average 1.05 mol/L 1.2

achieve a more stable yield of green ammonia with a higher

ammonia solution(mol/L)
Ammonia regeneration

80
concentration. 1
efficiency (ε, %)

3.2. Water recovery performance from ammonia-removed AD effluent 60 0.8

Average ε=44%
The AD effluents with different ammonia concentrations were 0.6
treated in the 2nd-stage MD to investigate the water recovery perfor­ 40
mance. The changes in water flux (JW, kg/(m2 h)) and ammonia rejec­ 0.4
tion efficiency (β, %) with the experimental time are shown in Fig. 8(a)
20
and (b). In all the groups, the water flux decreases almost linearly with 0.2
time. This is mainly due to the increase in the salt concentration caused
by the water loss in the AD effluent, leading to the decrease of water 0 0
vapor partial pressure over the AD effluent and then the drop in water 1 2 3 4 5
vapor transfer driving force [44]. Accordingly, with the continual mass Recycling times
transfer of water vapor, the AD effluent is concentrated, which causes an
increase in the TAN concentration of AD effluent [45], then an increase Fig. 7. Performance of ammonia absorption and desorption in multi-cycling
in the ammonia partial pressure over AD effluent, and thus a rise in the experiments. (a) Ammonia loading of the MAP/DAP loop (α, mol/mol) and
ammonia transfer driving force across the membrane from AD effluent ammonia removal efficiency (Rm, %). (b) Ammonia regeneration efficiency (ε,
into the recovered water. So regardless of the AD effluent with a high or %) of the MAP/DAP loop and TAN concentration of the recovered aqueous
low TAN concentration, the ammonia transfer increases inevitably with ammonia solution (mol/L) (Experimental conditions: in the ammonia absorp­
the experimental time [7], thus causing the ammonia rejection effi­ tion process, AD effluent: 3.86 g-N/L TAN concentration, 11.0 initial pH value
ciency decreases linearly from about 100 % to 6–18 % for all the groups. adjusted by adding NaOH, 60 ◦ C, 112 mL/min flow rate; Ammonia absorbent:
30 ◦ C, 10 mL/min flow rate; in the ammonia desorption process, 90 ◦ C
The EC value of recovered water decreases with the time as shown in
desorption temperature, 40 kPa desorption pressure, 90 min desorption time).
Fig. 8(d), which is highly in line with the trend of the TAN concentration
in the recovered water (Fig. 8c), indicating that the change of EC value is
mainly induced by the penetration of ammonia across membrane [7]. L, 0.6 g-N/L, and 0.003 g-N/L are adopted for water recovery in the 2nd-
Notably, the lower the ammonia concentration of AD effluent, the lower stage MD process, the corresponding TAN concentration/EC value of
the EC value and TAN concentration of recovered water. For example, recovered water after 2 h operation is about 22.09 g-N/L/201.61 ms/
when the AD effluents with the TAN concentration of 4.3 g-N/L, 1.5 g-N/ cm, 10.3 g-N/L/78.48 ms/cm, 2.54 g-N/L/20.12 ms/cm, and 0.02 g-N/

7
M. Shi et al. Journal of Water Process Engineering 49 (2022) 102949

(a)
0.2
Water flux (JA, kg/m2h)

0.15

0.1

4.3 g-N/L 1.5 g-N/L 0.6 g-N/L 0.003 g-N/L


0.05

0
0 5 10 15 20 25 30 0 5 10 15 20 25 30 0 5 10 15 20 25 30 0 5 10 15 20 25 30
Time (h)

(b) 100
Ammonia rejection
efficiency (β, %)

80
4.3 g-N/L 1.5 g-N/L 0.6 g-N/L 0.003 g-N/L
60

40

20

0
0 5 10 15 20 25 30 0 5 10 15 20 25 30 0 5 10 15 20 25 30 0 5 10 15 20 25 30

Time (h)

25 25
(c) (g-N/L)

TAN of recovered water


TAN of recovered water

20 20
(mg-N/L)
4.3 g-N/L 1.5 g-N/L 0.6 g-N/L 0.003 g-N/L

(mg-N/L)
(Cp, g-N/L)

15 15

10 10
(g-N/L)
5 (g-N/L) 5

0 0
0 5 10 15 20 25 30 0 5 10 15 20 25 30 0 5 10 15 20 25 30 0 5 10 15 20 25 30
Time (h)

(d) 200 (ms/cm) 200

EC of recovered water
EC of recovered water

150 150
4.3 g-N/L 1.5 g-N/L 0.6 g-N/L 0.003 g-N/L

(μs/cm)
(ms/cm)

100 (ms/cm) 100


(μs/cm)
50 (ms/cm) 50

0 0
0 5 10 15 20 25 30 0 5 10 15 20 25 30 0 5 10 15 20 25 30 0 5 10 15 20 25 30

Time (h)

Fig. 8. (a) Water flux (JW, kg/m2h), (b) ammonia rejection efficiency (β, %), (c) TAN concentration (Cp, g-N/L), and (d) electrical conductivity (EC, − ) of recovered
water from AD effluent with different ammonia concentrations through the 2nd-stage MD (Experimental conditions: AD effluent: 70 ◦ C, 70 mL/min flow rate;
deionized water: 20 ◦ C, 150 mL/min flow rate).

L/0.29 ms/cm, respectively. Thus, it is vital to remove ammonia nitro­ concentrations are below 0.09 mg/L (Table 1). Inferable, the sole im­
gen as much as possible before water recovery to achieve high-quality purity in the recovered water is ammonium ion. The primary volatile
reclaimed water. organic components like VFAs are not detected in the recovered water,
In addition to the TAN concentration and EC value, the ion con­ while other organic components are concentrated in the feed. This
centration of recovered water in the experimental group with the TAN quality of recovered water can meet the water quality requirements for
concentration of 0.003 g-N/L is also determined as listed in Table 2. reclaimed water [46], which can be used as general industrial or agri­
Clearly as shown in Fig. 8(c), the ammonium (NH+ 4 ) ion concentration cultural water.
can be kept to <1.5 mg/L after 18 h of operation, while all other ion

Table 2
Characteristics of the water recovered from the AD effluent with TAN concentration of 0.003 g-N/L by the 2nd-stage MD after 30 h of operation.
Parameters NH+
4 K+ Ca2+ Mg2+ F− Cl− NO−3 PO3−
4 SO2−
4

Value (mg/L) 1.5 0.001 0.083 0.003 0.019 0.090 0.046 0.051 0.026

8
M. Shi et al. Journal of Water Process Engineering 49 (2022) 102949

3.3. Economic evaluation of green ammonia and water recovery from AD 219.37 % reduction in the cost of chemical consumption. Furthermore,
effluent through 2s-MD the AD effluent treatment cost with pH adjusted by NaOH addition is the
highest, followed by that with pH adjusted by CaO addition, aeration,
The economic feasibility of green ammonia and water recovery from and waste alkali addition (Fig. 9). When the heat was provided by nat­
AD effluent by 2s-MD was evaluated. In the 1st-stage MD process, two ural gas and the MAP was used as the absorbent, the AD effluent treat­
kinds of ammonia absorbents, including MAP solution and H2SO4 so­ ment cost using 2s-MD system is about 1.73/m3-AD-effluent, which is
lution, were used for ammonia recovery with a targeted ammonia re­ only 2 % higher than that with pH adjusted by waste alkali (1.69/m3-
covery efficiency (>98 %) to generate the green ammonia and AD-effluent). This might be due to the extra cost of equipment and
(NH4)2SO4 aqueous solutions, respectively. It should be noted that the electricity consumption for aeration. Compared to pH adjustment by
formed aqueous ammonia solution in the MAP case was used to capture adding NaOH or CaO, the aeration method can be more economical,
CO2 for biogas upgrading to generate NH4HCO3. Different pH adjust­ which can also be confirmed by the reported results [34].
ment methods, such as NaOH, CaO, waste alkali additions, or aeration If the net profit is targeted, the combined utilization of the MAP/DAP
were used for different cases. In the 2nd-stage MD process, the water was loop and aeration technology in the 2s-MD process can provide a po­
recovered from the ammonia-removed AD effluent. Furthermore, the tential pathway to recover green ammonia and water from AD effluent.
heat required in the 2s-MD process was from the natural gas combustion Notably, the heat source has a significant effect on the net profit of AD
or waste heat recovery. The net profit of the 2s-MD system was calcu­ effluent treatment by 2s-MD process featured with MAP/DAP loop and
lated as the difference between the by-product profit and the cost of AD aeration. For example, the net profit of AD effluent is about $ 0.83/m3
effluent treatment. The main criteria for the economic evaluation were when heat is produced by natural gas combustion, while the net profit
presented in Tables A.5 and A.6, and the costs and profits for each increases greatly to $ 1.62/m3 when using the waste heat (Fig. 9).
treatment case were summarized in Table A.7 in the Supplement Comparisons of ammonia and water recovery costs from wastewater
Information. are presented in Table 3. At an initial TAN concentration of 2.5 g-N/L,
Fig. 9 shows the costs and profits of AD effluent treatment for the cost of ammonia recovery from leachate through ultrafiltration
recovering ammonia and water in different cases. In all the cases, the coupled with the MD is the highest ($ 7.52–9.11/m3-leachate), which
system using MAP solution as the recyclable absorbent can obtain a may be mainly due to the high cost of ultrafiltration pretreatment [19].
profit higher than that using the H2SO4 solution. This is mainly because The low cost of ammonia recovery through VMD without pH adjustment
a much less chemical is consumed in 2s-MD process, which results in a may be mainly attributed to the fact that the chemical costs of pH

NaOH-P
(a) NaOH-P (b) 3
2.5

2 AA-S 1 CaO-P
AA-S CaO-P
1.5 -1
1
-3
0.5

WA-S 0 WA-P WA-S -5 WA-P

Depreciation
Fixed O&M
Electricity-N Treatment-t
Heat-N
CaO-S AA-P CaO-S AA-P Profit-N
Alkali-N
Absorbent- N Profit-W
Electricity-W
NaOH-S Heat-W NaOH-S Profit-t

NaOH-P
(c) NaOH-P (d) 3
2.5

2 AA-S 1 CaO-P
AA-S CaO-P
1.5 -1
1
-3
0.5

WA-S 0 WA-P WA-S -5 WA-P

Depreciation
Fixed O&M Treatment-t
Electricity-N CaO-S AA-P
CaO-S AA-P Profit-N
Alkali-N
Profit-W
Absorbent- N
NaOH-S Electricity-W NaOH-S Profit-t

Fig. 9. Treatment cost profiles of water and green ammonia recovery by 2s-MD with pH adjusted by aeration (AA) and alkali additions (NaOH, CaO, and waste alkali
(WA)) with natural gas (a) and waste heat (b) as heat sources. The profit profiles of 2s-MD system with pH adjusted by aeration (AA) and alkali additions (NaOH,
CaO, and waste alkali (WA)) with natural gas (c) and waste heat (d) as heat sources. The investment depreciation of the system and the fixed operation and
maintenance cost (Fixed O&M) are shown in Table A.7. Electricity-W/N and heat-W/N refer to the cost of electricity and heat consumption in the water/ammonia
recovery process, respectively. The cost of ammonia absorbent (Absorbent-N) includes the costs of absorbent itself and its regeneration when MAP solution acts as
absorbent, or only the cost of absorbent itself when H2SO4 serves as absorbent. The cost of pH adjustment by alkali addition (Alkali-N) refers to alkali chemical costs
including NaOH, CaO, and waste alkali. Profit-W/N represents the profit of the products generated from the water/ammonia recovery process. The treatment-t and
profit-t represent the total treatment cost and profit of the water and ammonia recovery system.

9
M. Shi et al. Journal of Water Process Engineering 49 (2022) 102949

Table 3 Acknowledgments
Cost comparisons of water and ammonia recovery from the wastewater (initial
TAN concentration = 2 ± 0.5 g-N/L) by different methods. The authors thank the financial support from the Natural Science
Ammonia and water Products Cost Reference Foundation of Hubei Province of China (Nos. 2020CFA107,
recovery methods 2020CFB209) and the National Natural Science Foundation of China
Integrated ultrafiltration, ~50 % concentrated $ 4.62/m3- [18] (NSFC) (Nos. 52076101, 32002222).
reverse osmosis, and AD effluent; ~50 % AD-effluent
cold stripping process clean water; (NH4)2SO4 Appendix A. Supplementary data
Air stripping technique (NH4)2SO4; AD effluent $ 3.76/m3- [36]
with low TAN AD-effluent
concentration Supplementary data to this article can be found online at https://doi.
VMD without pH NH4HCO3; AD effluent $ 0.97–1.08/ [9] org/10.1016/j.jwpe.2022.102949.
adjustment with low TAN m3-AD-
concentration effluent
VMD with pH adjustment NH4HCO3; AD effluent $ 3.51–3.54/ [9] References
by NaOH with low TAN m3-AD-
concentration effluent [1] L.W. Deng, Y. Liu, D. Zheng, L. Wang, X.D. Pu, L. Song, Z.Y. Wang, Y.H. Lei, Z.
Ultrafiltration associated (NH4)2SO4; Landfill $ 9.11/m3- [19] A. Chen, Y. Long, Application and development of biogas technology for the
with MD-electrical leachate with low TAN leachate treatment of waste in China, Renew. Sustain. Energy Rev. 70 (2017) 845–851,
https://doi.org/10.1016/j.rser.2016.11.265.
energy concentration
[2] P.P. Wang, C. Xu, X. Zhang, Q.X. Yuan, Performance evaluation of a solar
Ultrafiltration associated (NH4)2SO4; Landfill $ 7.52/m3- [19]
evaporation system for liquid digestate concentration, Water Res. 211 (2022),
with MD-solar energy leachate with low TAN leachate 118056, https://doi.org/10.1016/j.watres.2022.118056.
concentration [3] J.P. Sheets, L. Yang, X. Ge, Z. Wang, Y. Li, Beyond land application: emerging
2s-MD process with pH ~30 % concentrated $ 3.48–4.25/ This technologies for the treatment and reuse of anaerobically digested agricultural and
adjusted by NaOH AD effluent; ~70 % m3-AD- study food waste, Waste Manag. 44 (2015) 94–115, https://doi.org/10.1016/j.
clean water; NH4HCO3 effluent wasman.2015.07.037.
2s-MD process with pH ~30 % concentrated $ 0.94–1.73/ This [4] A. Robles, D. Aguado, R. Barat, L. Borras, A. Bouzas, J.B. Gimenez, N. Marti,
adjusted by aeration AD effluent; ~70 % m3-AD- study J. Ribes, M.V. Ruano, J. Serralta, J. Ferrer, A. Seco, New frontiers from removal to
clean water; NH4HCO3 effluent recycling of nitrogen and phosphorus from wastewater in the circular economy,
2s-MD process with pH ~30 % concentrated $ 0.93–1.69/ This Bioresour. Technol. 300 (2020), 122673, https://doi.org/10.1016/j.
adjusted by waste AD effluent; ~70 % m3-AD- study biortech.2019.122673.
alkali clean water; NH4HCO3 effluent [5] B. Lauterbock, M. Ortner, R. Haider, W. Fuchs, Counteracting ammonia inhibition
in anaerobic digestion by removal with a hollow fiber membrane contactor, Water
Res. 46 (2012) 4861–4869, https://doi.org/10.1016/j.watres.2012.05.022.
[6] D. Hou, A. Iddya, X. Chen, M. Wang, W. Zhang, Y. Ding, D. Jassby, Z.J. Ren, Nickel-
adjustment and ammonia absorbents are not required [9]. As shown in based membrane electrodes enable high-rate electrochemical ammonia recovery,
Table 3, the cost of the 2s-MD combined with pH aeration adjustment is Environ. Sci. Technol. 52 (2018) 8930–8938, https://doi.org/10.1021/acs.
$ 0.94–1.73/m3-AD-effluent, which is lower than the cost reported by est.8b01349.
[7] Z.S. Yan, K. Liu, H.R. Yu, H. Liang, B.H. Xie, G.B. Li, F.S. Qu, B. van der Bruggen,
Ledda [18] ($ 4.62/m3-AD-wastewater) in which cost is mainly paid for Treatment of anaerobic digestion effluent using membrane distillation: effects of
(NH4)2SO4, concentrated AD effluent, and clean water productions. feed acidification on pollutant removal, nutrient concentration and membrane
Hence, our two-stage MD process has a certain economic advantage fouling, Desalination 449 (2019) 6–15, https://doi.org/10.1016/j.
desal.2018.10.011.
based on the simultaneous recovery of water and green ammonia. [8] Z.S. Yan, H.Y. Yang, F.S. Qu, H.R. Yu, H. Liang, G.B. Li, J. Ma, Reverse osmosis
brine treatment using direct contact membrane distillation: effects of feed
temperature and velocity, Desalination 423 (2017) 149–156, https://doi.org/
4. Conclusions 10.1016/j.desal.2017.09.010.
[9] M.F. Shi, Q.Y. He, L. Feng, L.L. Wu, S.P. Yan, Techno-economic evaluation of
In this study, the novel two-stage membrane distillation (2s-MD) ammonia recovery from biogas slurry by vacuum membrane distillation without
pH adjustment, J. Clean. Prod. 265 (2020), 121806, https://doi.org/10.1016/j.
process allows the continuous recovery of ammonia and water from the
jclepro.2020.121806.
AD effluent. In the 1st-stage MD, the monoammonium phosphate solu­ [10] Q.Y. He, J. Xi, M.F. Shi, L. Feng, S.P. Yan, L.W. Deng, Developing a vacuum-
tion (MAP, NH4H2PO4) acts as the recyclable ammonia absorbent for assisted gas-permeable membrane process for rapid ammonia recovery and CO2
ammonia recovery, where no significant difference in ammonia flux and capture from biogas slurry, ACS Sustain. Chem. Eng. 8 (2020) 154–162, https://
doi.org/10.1021/acssuschemeng.9b04873.
ammonia removal efficiency between MAP and H2SO4 absorbents was [11] S.N. McCartney, N.A. Williams, C. Boo, X. Chen, N.Y. Yip, Novel isothermal
observed for the ammonia recovery process. Then, the green ammonia is membrane distillation with acidic collector for selective and energy-efficient
recovered from the ammonia-saturated MAP, i.e., diammonium phos­ recovery of ammonia from urine, ACS Sustain. Chem. Eng. 8 (2020) 7324–7334,
https://doi.org/10.1021/acssuschemeng.0c00643.
phate solution (DAP, (NH4)2HPO4) by vacuum distillation. The MAP/ [12] R.D. Silva, H. Ramlow, C.D.K. Cavalcanti, R.D.S.C. Valle, R.A.F. Machado,
DAP loop is recycled and the renewable aqueous ammonia concentra­ C. Marangoni, Steady state evaluation with different operating times in the direct
tion is maintained at about 1.05 mol/L in the multicycle ammonia contact membrane distillation process applied to water recovery from dyeing
wastewater, Sep. Purif. Technol. 230 (2020), https://doi.org/10.1016/j.
absorption-desorption experiments. In the 2nd-stage MD, water is seppur.2019.115892.
recovered from the AD effluent with different ammonia concentrations. [13] A. Alkhudhiri, N. Darwish, N. Hilal, Membrane distillation: a comprehensive
The recovered water can be used as the general industrial or agricultural review, Desalination 287 (2012) 2–18, https://doi.org/10.1016/j.
desal.2011.08.027.
water with good and stable quality. The economic analysis reveals the [14] M. Darestani, V. Haigh, S.J. Couperthwaite, G.J. Millar, L.D. Nghiem, Hollow fibre
economic feasibility of water and green ammonia recovery from AD membrane contactors for ammonia recovery: current status and future
effluent through the 2s-MD process. Our results indicate that it is developments, J. Environ. Chem. Eng. 5 (2017) 1349–1359, https://doi.org/
10.1016/j.jece.2017.02.016.
essential to optimize the technical parameters and treatment methods
[15] C.J. Davey, P. Liu, F. Kamranvand, L. Williams, Y. Jiang, A. Parker, S. Tyrrel, E.
for the ammonia and water recovery process to minimize the chemical J. McAdam, Membrane distillation for concentrated blackwater: influence of
consumption and treatment cost. configuration (air gap, direct contact, vacuum) on selectivity and water
productivity, Sep. Purif. Technol. 263 (2021), https://doi.org/10.1016/j.
seppur.2021.118390.
[16] Y. Lin, M. Guo, N. Shah, D.C. Stuckey, Economic and environmental evaluation of
Declaration of competing interest nitrogen removal and recovery methods from wastewater, Bioresour. Technol. 215
(2016) 227–238, https://doi.org/10.1016/j.biortech.2016.03.064.
The authors declare that they have no known competing financial [17] A. Mavhungu, S. Foteinis, R. Mbaya, V. Masindi, I. Kortidis, L. Mpenyana-
Monyatsi, E. Chatzisymeon, Environmental sustainability of municipal wastewater
interests or personal relationships that could have appeared to influence treatment through struvite precipitation: influence of operational parameters,
the work reported in this paper. J. Clean. Prod. 285 (2021), https://doi.org/10.1016/j.jclepro.2020.124856.

10
M. Shi et al. Journal of Water Process Engineering 49 (2022) 102949

[18] C. Ledda, A. Schievano, S. Salati, F. Adani, Nitrogen and water recovery from distillation, Sep. Purif. Technol. 191 (2018) 182–191, https://doi.org/10.1016/j.
animal slurries by a new integrated ultrafiltration, reverse osmosis and cold seppur.2017.09.030.
stripping process: a case study, Water Res. 47 (2013) 6157–6166, https://doi.org/ [33] M. Walker, K. Iyer, S. Heaven, C.J. Banks, Ammonia removal in anaerobic digestion
10.1016/j.watres.2013.07.037. by biogas stripping: an evaluation of process alternatives using a first order rate
[19] M.M. Zico, B.C. Ricci, B.G. Reis, N.C. Magalhaes, M.C.S. Amaral, Sustainable model based on experimental findings, Chem. Eng. J. 178 (2011) 138–145, https://
ammonia resource recovery from landfill leachate by solar-driven modified direct doi.org/10.1016/j.cej.2011.10.027.
contact membrane distillation, Sep. Purif. Technol. 264 (2021), https://doi.org/ [34] M.C. Garcia-Gonzalez, M.B. Vanotti, Recovery of ammonia from swine manure
10.1016/j.seppur.2021.118356. using gas-permeable membranes: effect of waste strength and pH, Waste Manag. 38
[20] B. Pandey, L. Chen, Technologies to recover nitrogen from livestock manure-a (2015) 455–461, https://doi.org/10.1016/j.wasman.2015.01.021.
review, Sci. Total Environ. 784 (2021), https://doi.org/10.1016/j. [35] L. Zhang, D. Jahng, Enhanced anaerobic digestion of piggery wastewater by
scitotenv.2021.147098. ammonia stripping: effects of alkali types, J. Hazard. Mater. 182 (2010) 536–543,
[21] J. Zhang, M. Xie, X. Tong, S. Liu, D. Qu, S. Xiao, Recovery of ammonium nitrogen https://doi.org/10.1016/j.jhazmat.2010.06.065.
from human urine by an open-loop hollow fiber membrane contactor, Sep. Purif. [36] G. Provolo, F. Perazzolo, G. Mattachini, A. Finzi, E. Naldi, E. Riva, Nitrogen
Technol. 239 (2020), https://doi.org/10.1016/j.seppur.2020.116579. removal from digested slurries using a simplified ammonia stripping technique,
[22] P.J. Dube, M.B. Vanotti, A.A. Szogi, M.C. Garcia-Gonzalez, Enhancing recovery of Waste Manag. 69 (2017) 154–161, https://doi.org/10.1016/j.
ammonia from swine manure anaerobic digester effluent using gas-permeable wasman.2017.07.047.
membrane technology, Waste Manag. 49 (2016) 372–377, https://doi.org/ [37] S. Daguerre-Martini, M.B. Vanotti, M. Rodriguez-Pastor, A. Rosal, R. Moral,
10.1016/j.wasman.2015.12.011. Nitrogen recovery from wastewater using gas-permeable membranes: impact of
[23] P. Jacob, P. Phungsai, K. Fukushi, C. Visvanathan, Direct contact membrane inorganic carbon content and natural organic matter, Water Res. 137 (2018)
distillation for anaerobic effluent treatment, J. Membr. Sci. 475 (2015) 330–339, 201–210, https://doi.org/10.1016/j.watres.2018.03.013.
https://doi.org/10.1016/j.memsci.2014.10.021. [38] Y. Ye, S. Yu, L. Hou, B. Liu, Q. Xia, G. Liu, P. Li, Microbubble aeration enhances
[24] M.S. EL-Bourawi, M. Khayet, R. Ma, Z. Ding, Z. Li, X. Zhang, Application of vacuum performance of vacuum membrane distillation desalination by alleviating
membrane distillation for ammonia removal, J. Membr. Sci. 301 (2007) 200–209, membrane scaling, Water Res. 149 (2019) 588–595, https://doi.org/10.1016/j.
https://doi.org/10.1016/j.memsci.2007.06.021. watres.2018.11.048.
[25] M. Sheikh, M. Reig, X. Vecino, J. Lopez, M. Rezakazemi, C.A. Valderrama, J. [39] M.R. Conde, Properties of aqueous solutions of lithium and calcium chlorides:
L. Cortina, liquid-liquid membrane contactors incorporating surface skin formulations for use in air conditioning equipment design, Int. J. Therm. Sci. 43
asymmetric hollow fibres of poly(4-methyl-1-pentene) for ammonium recovery as (2004) 367–382, https://doi.org/10.1016/j.ijthermalsci.2003.09.003.
liquid fertilisers, Sep. Purif. Technol. 283 (2022), https://doi.org/10.1016/j. [40] J. Wang, D. Xin, X. Hou, J. Wu, X. Fan, K. Li, J. Zhang, Structural properties and
seppur.2021.120212. hydrolysabilities of Chinese pennisetum and hybrid pennisetum: effect of aqueous
[26] Z. Xie, T. Duong, M. Hoang, C. Nguyen, B. Bolto, Ammonia removal by sweep gas ammonia pretreatment, Bioresour. Technol. 199 (2016) 211–219, https://doi.org/
membrane distillation, Water Res. 43 (2009) 1693–1699, https://doi.org/ 10.1016/j.biortech.2015.08.046.
10.1016/j.watres.2008.12.052. [41] C.R. Wu, H.H. Yan, Z.G. Li, X.L. Lu, Ammonia recovery from high concentration
[27] I. Sancho, E. Licon, C. Valderrama, N. de Arespacochaga, S. Lopez-Palau, J. wastewater of soda ash industry with membrane distillation process, Desalin.
L. Cortina, Recovery of ammonia from domestic wastewater effluents as liquid Water Treat. 57 (2016) 6792–6800, https://doi.org/10.1080/
fertilizers by integration of natural zeolites and hollow fibre membrane contactors, 19443994.2015.1010233.
Sci. Total Environ. 584–585 (2017) 244–251, https://doi.org/10.1016/j. [42] Q.Y. He, G. Yu, W.C. Wang, S.P. Yan, Y.L. Zhang, S.F. Zhao, Once-through CO2
scitotenv.2017.01.123. absorption for simultaneous biogas upgrading and fertilizer production, Fuel
[28] M.C. Garcia-Gonzalez, M.B. Vanotti, A.A. Szogi, Recovery of ammonia from swine Process. Technol. 166 (2017) 50–58, https://doi.org/10.1016/j.
manure using gas-permeable membranes: effect of aeration, J. Environ. Manag. fuproc.2017.05.027.
152 (2015) 19–26, https://doi.org/10.1016/j.jenvman.2015.01.013. [43] Q.Y. He, M.F. Shi, F.H. Liang, L. Ji, X. Cheng, S.P. Yan, BEEF: a sustainable process
[29] H. Huang, P. Zhang, J. Xiao, D. Xiao, F. Gao, Repeatedly using the decomposition concerning negative CO2 emission and profit increase of anaerobic digestion, ACS
product of struvite by ultrasound stripping to remove ammonia nitrogen from Sustain. Chem. Eng. 7 (2019) 2276–2284, https://doi.org/10.1021/
landfill leachate, Ultrason. Sonochem. 38 (2017) 622–628, https://doi.org/ acssuschemeng.8b04963.
10.1016/j.ultsonch.2016.08.019. [44] S. Yarlagadda, V.G. Gude, L.M. Camacho, S. Pinappu, S. Deng, Potable water
[30] Q.B. Zhao, J.W. Ma, I. Zeb, L. Yu, S.L. Chen, Y.M. Zheng, C. Frear, Ammonia recovery from As, U, and F contaminated ground waters by direct contact
recovery from anaerobic digester effluent through direct aeration, Chem. Eng. J. membrane distillation process, J. Hazard. Mater. 192 (2011) 1388–1394, https://
279 (2015) 31–37, https://doi.org/10.1016/j.cej.2015.04.113. doi.org/10.1016/j.jhazmat.2011.06.056.
[31] R.D. Rice, J.V. Busa, Recovering NH3 by the phosam-W process, Chem. Eng. Prog. [45] E.U. Khan, A. Nordberg, Membrane distillation process for concentration of
80 (1984) 61–63. nutrients and water recovery from digestate reject water, Sep. Purif. Technol. 206
[32] Q.Y. He, T. Tu, S.P. Yan, X. Yang, M. Duke, Y.L. Zhang, S.F. Zhao, Relating water (2018) 90–98, https://doi.org/10.1016/j.seppur.2018.05.058.
vapor transfer to ammonia recovery from biogas slurry by vacuum membrane [46] A.N.S.I.I. United States of America, Standard Methods for the Examination of
Water And Wastewater, 16th edition, American National Standard, 1985.

11

You might also like