You are on page 1of 20

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/344799212

Adsorbent materials for ammonium and ammonia removal: A review

Article in Journal of Cleaner Production · October 2020


DOI: 10.1016/j.jclepro.2020.124611

CITATION READS

1 220

5 authors, including:

Bing Han Clayton Butterly


University of Melbourne University of Melbourne
15 PUBLICATIONS 180 CITATIONS 47 PUBLICATIONS 933 CITATIONS

SEE PROFILE SEE PROFILE

Ji-Zheng He Deli Chen


Chinese Academy of Sciences University of Melbourne
325 PUBLICATIONS 11,479 CITATIONS 304 PUBLICATIONS 7,879 CITATIONS

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

microbial View project

Sprayable polymers for agriculture View project

All content following this page was uploaded by Bing Han on 21 October 2020.

The user has requested enhancement of the downloaded file.


Journal of Cleaner Production xxx (xxxx) 124611

Contents lists available at ScienceDirect

Journal of Cleaner Production


journal homepage: http://ees.elsevier.com

F
OO
Review

Adsorbent materials for ammonium and ammonia removal: A review


Bing Han, Clayton Butterly, Wei Zhang, Ji-zheng He, Deli Chen ∗

PR
Faculty of Veterinary and Agricultural Sciences, The University of Melbourne, Parkville, VIC, 3010, Australia

ARTICLE INFO ABSTRACT

Article history: Ammonium (NH4+) and ammonia (NH3) are notorious hard-to-treat pollutants, leading to serious deterioration
Received 3 July 2020 of aquatic ecosystems and significant risks to human health. While adsorption is a promising method to tackle
Received in revised form 14 September 2020 this problem, finding suitable adsorbent materials which are abundant, low-cost and efficient remains a constant
Accepted 9 October 2020 challenge. Thus, this review summarizes recent development of important adsorbent materials implemented for

D
Available online xxx
NH3/NH4+ removal. Advantages and disadvantages of representative adsorbent materials including bentonite,
Handling editor: Prof. Jiri Jaromir Klemeš
zeolite, clay, biochar, activated carbon, metal organic framework and their modified forms are compared, and the
nature of their adsorption processes are discussed in context of adsorption sites, isotherm models (e.g. Langmuir
Keywords and Freundlich), kinetic equations (e.g. pseudo-first order, pseudo-second order and intra-particle diffusion) and
TE
Ammonium thermodynamic analysis. Future perspective on the utilization of inexpensive lignite is also conferred. Although
Ammonia both conventional and nanostructured materials face challenges regarding economic cost, energy consumption,
Adsorption secondary pollution and adsorption efficiency, these can be tackled by adopting various of advanced options.
Modification Current research on adsorption mechanisms forms a solid basis for the design and development of novel adsor-
Mechanism
bent materials. We speculate that the pursuit of strategies for effective surface modification of natural abundant
Kinetics
resources will lead to a bright future of removal processes suited to low NH3/NH4+ concentration conditions.
EC

© 2020

1. Introduction the atmospheric environment (Megaritis et al., 2013; Stokstad,


2014).
Over the past century, improper disposal of industrial and agricul- It has been estimated that >200 Mt of NH3/NH4+ is being gen-
RR

tural waste has led to significant social, economic and environmental erated across the world on an annual basis (DeCoste et al., 2016;
problems (Ravishankara et al., 2009; Stokstad, 2014). Ammonium Khabzina and Farrusseng, 2018). Major anthropogenic sources in-
(NH4+) is the most frequently encountered nitrogenous (N) compound clude N fertilizer synthesis, intensive animal feedlots, chemical fiber
in wastewaters. The elevated levels of N nutrients in water sources plants, agricultural wastewater, aquaculture industry and others (Busca
could lead to various issues such as toxic algal blooms, fish kills and and Pistarino, 2003; Hao and Yan, 2016; Maia et al., 2012).
poor drinking water quality via eutrophication (Adam et al., 2019a; For example, intensive animal feedlots are large hotspots of NH3 emis-
CO

Blazquez et al., 2017; Huang et al., 2018). In particular, it is toxic sions where ~60% of the N consumed in feedlot rations is wasted
to some fish even at a low level of 3 ppm (Boopathy et al., 2013); Ni- as NH3 via volatilization (Chen et al., 2015). Such a loss not only
trate (NO3−) and nitrite (NO2−) derived from NH4+ are involved in hu- bears an environmental consequence, but also is a huge waste of an
man diseases such as methemoglobinemia. This poses a significant risk economical resource. With a typical energy consumption of 12.1 kWh
to aquatic ecosystems as well as human health. On the other hand, a kg−1 NH3–N (Cruz et al., 2019), industrial Haber-Bosch process of
large proportion of NH4+ in aqueous streams is lost via NH3 volatiliza- NH3 production costs about 2% world's energy production and cause
tion. As a relatively strong base and corrosive pollutant, NH3 gas could 1% CO2 annual emission (Service, 2018). Domestic wastewater an-
UN

cause undesired odors and lead to severe burning of eyes, skin, and res- nually comprises ~20 Mt of NH3/NH4+, which is equivalent to ~19%
piratory systems at a concentration as low as 50–100 ppm (Katz et al., of the annual industrial NH3 production, and the amount is projected
2016; Katz and Leznoff, 2009; Vikrant et al., 2017). It can also to further increase to 35 Mt annually by the middle of the 21st cen-
react with other air pollutants such as SOx or NOx to form ammonium tury (Bodirsky et al., 2014; Cruz et al., 2019). Therefore, regula-
salt, contributing to fine particulate matter (PM2.5) concentration in tory agencies in many jurisdictions require the removal of NH3/NH4+
from emission hotspots before discharging to environment. According
to American Conference Governmental Industrial Hygienists (ACGIH),
only NH3 concentration < 25 ppm for time-weighted (8-h work day)
and <35 ppm for short-term (15 min) values are permitted (NIOSH,

Corresponding author.
E-mail address: delichen@unimelb.edu.au (D. Chen) 1994; Wang et al., 2017). The National Emission Ceilings

https://doi.org/10.1016/j.jclepro.2020.124611
0959-6526/© 2020.
2 B. Han et al. / Journal of Cleaner Production xxx (xxxx) 124611

Directive 2016/2284/EU have required European Union (EU) countries both advantages and environmental and technological limitations. A
to decrease the NH3 emissions, in order to reduce PM2.5 concentra- quantitative comparison of NH3/NH4+ removal between several es-
tions (Giannakis et al., 2019). In addition, the European Environment tablished and emerging approaches from wastewater technologies are
Agency (EEA) has urged researchers to explore novel approaches for briefly summarized in Table 1. Among these approaches, while bio-
NH3 removal (Karri et al., 2018). logical processes are very important for wastewater treatment (Ardern
To alleviate the negative impact of N pollution, various kinds of and Lockett, 1914), the transformation of NH4+ into its elemental N2

F
techniques were studied, and several excellent review articles are avail- is incompatible with 21st century's goal to develop a circular economy
able (Adam et al., 2019a; Cruz et al., 2019; Gupta et al., 2015; (Cruz et al., 2019; Guest et al., 2009; Verstraete et al., 2009).
Karri et al., 2018). The current routes for NH3/NH4+ removal have Other approaches include air stripping, oxidation, precipitation, photo

OO
Table 1
Summary of several established and emerging approaches for removal of NH3/NH4 + from wastewater.

Outlet
concentration Removal
Technology Working conditions Advantage Challenges (mg/L NH4–N) efficiency Ref.

PR
Adsorption • Wide ranges of temperature and • Easy operation • Different adsorbents have different 1 43–100% This work
pH • Effectively remove removal efficiencies in Table
NH4 + 2
• Able to work in
low NH4 + concen-
trations
Air stripping • pH range: 10.8–11.5 • Widely used • High chemical demands 500–1000 50–99% Adam et
• High NH4 + concentration process for waste- • Hindered by low temperature al.

D
• >15 °C water pre-treat- • Time consuming (2019)
ment Taşdemir
• High energy consuming
• Simple equipment et al.
• Scaling and fouling
• Not sensitive to (2020)
toxic substances Yin et al.
TE
(2018)
Zhu et
al.
(2017)
Reverse • Liquid-liquid membrane is • Low field and • Produces dilute concentrate stream, 1 60–99% Liu et al.
osmosis (RO) needed space only concentrating max 1 order of (2007)
• Easy and continu- magnitude. Talalaj
EC

ous operation, • Membrane fouling (2015)


• simplicity • Clean the membrane regularly Guo and
Jin
(2014)
Struvite • Require certain pH and tempera- • A valuable slow-re- • Needs phosphate to proceed 29–100 20–99% Karri et
precipitation ture lease fertilizer • Many competitors for phosphate al.
• Medium cost (calcium, magnesium phosphates) (2018)
Hakimi
RR

• Introduce new pollutants


et al.
(2020)
Zhao et
al.
(2019)
Peng et
al.
(2017)
CO

Biological • Growth of heterotrophic or pho- • Free of chemical • Produces biomass, not NH4 + – not <5 70–99.9% Adam et
treatment tosynthetic (or phototrophic) al- reagents readily useable. al.
gae or bacteria • No need for com- • Only operates at low input/output (2019)
• Sensitive to temperature plicated configura- concentrations Lin et al.
tion • Long start-up time (2020)
Moondra
et al.
(2020)
UN

Miao et
al.
(2020)
Photocatalysis • In both liquid and gas phase • Solar energy is free • Efficient photocatalysts are still un- 35–100% Yao et
• No secondary pol- der development al.
lution • Efficiency depends on light source (2020)
Vikrant
et al.
(2020)
Bahmani
et al.
(2020)
Ren et
al.
(2020)
B. Han et al. / Journal of Cleaner Production xxx (xxxx) 124611 3

catalysis and chemical coagulation (Dong et al., 2019a; Gupta et lished during the period 2016–2020 were included to this review. More-
al., 2015; Kinidi et al., 2018; Ma et al., 2020; Rada-Ariza et al., over, relevant studies found during screening other studies were in-
2017; Tao et al., 2020; Ye et al., 2018)Ma et al., 2020. These cluded to the list during literature identification. The literature search
processes, however, required complicated configuration, high NH4+ lev- was limited to articles published in peer-reviewed journals in English
els (>2 g L−1 NH4–N) and pose difficulties for maintenance (Boopa- language. Reports published in other languages as well as books were
thy et al., 2013; Cruz et al., 2019). Adsorption to solid media has excluded from the literature search.

F
many favourable characteristics including a high removal efficiency, Prepared list of scientific articles was critically analysed through ex-
ease of operation and low energy consumption. It is suitable especially tracting relevant information using the list of questions shown below:
for dilute NH4+ concentration in domestic wastewater (40–60 mg L−1

OO
NH4–N) (Cruz et al., 2019; Henze et al., 2008), landfill leachate • Which type of adsorbent material was used?
(100–1000 mg L−1 NH4–N) (Karadag et al., 2008), rainfall runoff • Was the regeneration of NH3/NH4+ materials studied?
(20–49 mg L−1 NH4–N) (Wang et al., 2016), municipal secondary ef- • Which type of modification approach was used?
fluent (~50 mg L−1 NH4–N) (Zhang et al., 2016) and fish pond water • Was the recycling or reuse of the adsorbent materials studied?
(~10 mg L−1 NH4–N) (Soetardji et al., 2015). These significant advan- • Which type of wastewater was used?
tages make it a promising method to be applied on a commercial scale. • What was the scale of performed experiments?
Nevertheless, the success of the technique greatly relies on finding suit- • Was the modification approach realistic, applicable in real world

PR
able adsorbent materials. So far, the representative adsorbent materials cases?
widely used in NH3/NH4+ removal in both lab- and pilot-scale studies • Were the adsorption efficiencies of adsorbent materials compared?
include bentonite, zeolite, clay, biochar, activated carbon, metal organic • Was the concentration of NH4+ in the solution identified?
frameworks (MOFs) and other nanostructured materials, which will be • Were the mechanisms underlying adsorption process addressed?
discussed below.
Recent review articles about NH4+ adsorption have emphasised the 3. Adsorbent materials
importance of operation parameters like adsorbent material dosage, pH,
contact time and coexisting ions (Wang and Peng, 2010). Huang

D
Due to the impact of NH3/NH4+ emission, numerous adsorbent ma-
et al. (2018) compared a range of cost-effective adsorbent materials terials have been studied. Those materials were roughly categorized
in context of cost, efficiency and preparation. However, very little at- into conventional (e.g. clay, carbonaceous materials) and nanostruc-
tention has been paid to mechanistic/kinetic pathways and thermody- tured (e.g. MOFs) materials based on their structures and properties.
TE
namic insights. Furthermore, despite intensive efforts to enhance NH4+ Both have advantages and disadvantages (Fig. 1). Bentonite, zeolite,
removal by surface modifications, comprehensive reviews on this are other clay materials and biochar were mainly used for NH4+ removal
scarce. Therefore, the goal of the present review is to summarize our from aqueous phase, whereas activated carbon and MOFs were mainly
current understanding of representative adsorbent materials as well as employed for NH3 removal from gas phase. Generally, conventional ma-
their modified forms for the removal of NH4+ (aqueous) and NH3 (in terials are relatively abundant and low-cost, but their binding affinities
the gas phase), by using published studies mainly in last 3 years as ex- with NH3/NH4+ are usually weak. On the other hand, nanostructured
EC

amples. Specifically, this review will (i) compare different classes of ad- materials could show high adsorption capacity and rapid adsorption ki-
sorbent materials, (ii) illustrate the promising modification approaches, netics in many cases. However, most of those materials were only pre-
and (iii) discuss the adsorption mechanisms, kinetics and thermodynam- pared from laboratory with tedious preparation procedures using expen-
ics involved. sive precursors. Hence the yield is low and material properties vary a
lot. In this sense, modifying conventional materials to enhance their in-
2. Methods teraction with NH3/NH4+ is a promising approach for real world ap-
RR

plication. The modification methods can be roughly divided into phys-


The methods applied in this literature review included identification ical modification including sonication, calcination (to increase surface
of the relevant studies and preparing set of questions to be addressed area) and pelletization (to avoid pressure drop), and chemical modifica-
to selected literature relevant to the scope of this review. Identifica- tion (e.g. acid/base treatment, oxidation and coating) to promote bind-
tion of relevant literature was performed by searching in Science Di- ing affinity (Fig. 2). Following the order of cost of adsorbent materi-
rect, Scopus and Google Scholar databases using following keywords: als, here we firstly discuss natural abundant bentonite, zeolite and other
“ammonia”, “ammonium”, “removal”, “adsorption”. After that the gen- clay materials both in their natural and modified form, and then move
CO

erated literature list was checked manually (reading materials and meth- to biochar and activated carbon which need manufacturing but are po-
ods and results) in order to exclude studies, in which (i) adsorbent ma- tentially low-cost. In the end, we briefly summarize novel materials that
terials were not used; (ii) removal efficiency of NH3/NH4+ before and are still under laboratory testing including MOFs and other nanostruc-
after adsorption was not conducted. Moreover, literature reviews were tured materials.
not considered. It should be noted that mainly relevant articles pub
UN

Fig. 1. The advantages and disadvantages of conventional and nanostructured adsorbent materials.
4 B. Han et al. / Journal of Cleaner Production xxx (xxxx) 124611

40 min at pH of 6, which is 2 times greater than the that of nat-


ural zeolite reported by Martins et al. (2017). Based on the review
by Wang and Peng, 2010, NH4+ adsorption on natural zeolite is
2.7–30.6 mg g−1.
Modification to zeolite was done to enhance NH4+ adsorption by i)
acid/base treatment, ii) physical modification through increasing sur-

F
face areas, iii) incorporation of functional coatings that decreases con-
Fig. 2. Common modification methods to enhance NH3/NH4+ removal.
centrations of competing adsorbates and strengthen binding affinity of
NH4+, and iv) embedding zeolite particles in a framework to reduce dif-

OO
fusion resistance. Generally acid treatment of zeolite decreased removal
3.1. Bentonite efficiency of NH4+, while base treatment improves it by e.g. convert-
ing zeolite's exchange sites to the Na form, where NH4+ has high affin-
Bentonite is frequently utilized as an adsorbent material because of ity over Na+ (Wang and Peng, 2010). Physical modification by ultra-
its cost-effectiveness and availability throughout the world. It is an alu- sound with a power density of ~5 W cm2 was employed by Zielinski
minosilicate material composed of one layer of aluminum oxygen octa- et al., 2016 who increased NH4+ removal rate by ca. 30%. Zhang et
hedron and two layers of silicon oxygen tetrahedron inclusions clamped al. (2016) showed that coating Na-modified zeolite with an anion ex-
together in a ratio of 1:2 (Cheng et al., 2019). Strong NH4+ adsorp- change resin improved NH4+ removal efficiency from 78% to 95%, due

PR
tion capacity of bentonite originated from the high negative charge on to the resin's ability to reduce dissolved organic matter in effluent. In
surface, and high cation exchange capacity (CEC), which is generally 2015, Kang et al. (2015) showed that coating of hydrophilic zeolite
40–130 cmol kg−1. In 2017, Angar et al. (2017) found that raw ben- with microporous organic polymer materials improves NH3 adsorption
tonite exhibited an NH4+ adsorption capacity of 19.01 mg g−1 corre- capacity from 1.8 to 5.6 mg g−1 under 43% relative humidity. Putra
sponding to a removal efficiency of 53.4% (pH = 7, 30 °C) and this per- and Lee, 2020 proved that the entrapment of zeolite particles in porous
centage could be improved to 81.2% by increasing adsorbent material hydrogels had a high NH4+ adsorption capacity of 28.17 mg g−1, which
dosage to 40 g L−1. was 4.3 times higher than those of zeolite beads. Additional benefits of

D
Organic and metal oxide materials have been combined with ben- this strategy include minimal agglomeration and good settle ability. Pu-
tonite by mixing in solution and heating at elevated temperature, in or- tra and Lee, 2020 also summarized the maximum NH4+ adsorption
der to enhance adsorption performance of NH4+. On one hand, coat- capacities of both natural and synthesized zeolite-type adsorbent mate-
ing bentonite with functional materials can increase the number of ac- rials as a function of particle sizes in the literatures, which are presented
TE
tive sites for chemical adsorption. As a result, the hybrid samples usu- in Fig. 3. Nevertheless, although zeolite is a low-cost material, some
ally show superior adsorption capacity. On the other hand, bentonite modification processes could be energy intensive and requires expensive
can act as a binder to create composites in a pelletized form with con- chemical reagent, which may prohibit its large-scale application.
trolled particle size, and thus be suitable for dynamic conditions in
comparison with powder adsorbent materials considering pressure drop 3.3. Other clay materials
(Seredych et al., 2016). In 2015, Seredych et al. (2016) reported
EC

that on phenolic-formaldehyde resin/bentonite the adsorption capacity Abundant natural clay materials (NCM), geopolymers and discarded
of NH3 reached 38.0 mg g−1 and 12.4 mg g−1 in dry and moist condi- concrete, which are inexpensive and non-toxic are also promising adsor-
tions, respectively. Yadi et al. (2016) found that a bentonite/chitosan bent materials for NH4+. It was reported that NCM is about 20 times
composite shows a maximum NH4+ adsorption capacity of 11.6 mg g−1, cheaper than activated carbon (Alshameri et al., 2018). Therefore,
whereas natural bentonite showed NH4+ adsorption capacities of only investigation on adsorption capacities of different types of NCM is of
0.75 mg g−1, respectively. Recently, Cheng et al. (2019) first modi- high importance. Alshameri et al. (2018) compared six natural clay
RR

fied bentonite with aluminum via cross-linking reaction and then adding minerals for NH4+ removal from aqueous solution including kaolin-
concentrated tannic acid to further enhance its functional capacity. The ite, halloysite, montmorillonite, vermiculite, palygorskite and sepiolite
product showed a high removal rate of >70% while that of natural ben-
tonite is only ~30%.

3.2. Zeolite
CO

Up to 4 Mt of natural zeolites are mined every year across the


world. Different from bentonite which has 2D layered structure, zeo-
lites composed Al3+ and Si4+ form tetrahedral with O bridge to form
3D porous framework with negatively charged lattice. These negative
charges, balanced by exchangeable cations, make zeolite one of the most
common ion-exchangers and have been widely applied to NH4+ ad-
sorption since 1970's (Liu et al., 2018). However, the removal effi-
UN

ciency depends on the physiochemical properties of zeolite. In 2017,


Martins et al. (2017) first evaluated NH4+ adsorption capacity of
natural zeolite, which was 10.8 mg g−1 and then biological regenera-
tion was carried out with nitrifying bacteria suspension. It was found
that the adsorption capacity decreased by only 4.55% after regenera-
tion. Zeolites can also be produced synthetically. Compared with nat-
ural zeolite, synthetic zeolites possess favourable properties towards
Fig. 3. Comparison of the maximum NH4+ adsorption capacities of zeolite-type adsor-
adsorption like uniform porosity and high surface area. In 2018, Liu
bent materials as a function of particle sizes (ZP: zeolite particles, PAZ: polyvinyl alcohol
et al. (2018) synthesized zeolite from fluidized bed fly ash, which (PVA)-alginate-ZPs, ZB: zeolite bead) (Putra and Lee, 2020).
showed a maximum NH4+ adsorption capacity of 22.9 mg g−1 in
B. Han et al. / Journal of Cleaner Production xxx (xxxx) 124611 5

(Fig. 4). The Langmuir isotherms indicated that the highest NH4+ ad- at high pyrolysis temperature can't be neglected. Zheng et al. (2018)
sorption capacities were achieved by vermiculite (50.06 mg g−1) and prepared biochar from distillers' grains residue at 300–800 °C and found
montmorillonite (40.84 mg g−1). Jing et al. (2017) also found that hal- biochar produced at 700 °C exhibited the highest NH4+ adsorption ca-
loysite showed a NH4+ adsorption capacity of 1.66 mg g−1 (pH = 5.6, pacity of 5.86 mg g−1. Furthermore, the pH value of NH4+ solution also
303 K) with initial NH4+ concentration of 600 mg L−1, and it increased influenced NH4+ adsorption on biochar. Fan et al. (2019) found that
with increasing initial concentration of NH4+, pH values and tempera- adsorption capacity of bamboo biochar increased with increasing pH

F
tures (288 K–313 K). values at pH range of 2–7 and then decreased at pH range of 7–8, ex-
Artificial clay materials have also been produced and tested, but hibiting an optimal 114.84 mg g−1 at pH = 7.
many of their NH4+ adsorption capacities are comparable or even Biochar has been modified by subsequent treatment in order to pro-

OO
lower than that of NCM. Cheng et al. (2017) synthesized Na-rich mote NH4+ adsorption. A range of oxidative reagents such as HNO3
birnessite, which showed NH4+ adsorption capacity of 22.61 mg g−1. and H2O2 have been employed to increase amount of oxygen-containing
In 2016, Luukkonen et al. (2016) synthesized a geopolymer from functional groups. Furthermore, NH4+ adsorption capacity could also
metakaolin and found that the NH4+ adsorption capacity of the geopoly- be enhanced by neutralizing surface acidity, because affinity of alka-
mer (21.07 mg g−1) was 46% greater than that of clinoptilolite-heulan- line metal ions to functional groups are generally lower than that of
dite zeolite (14.42 mg g−1) as a reference. Later on, they found that the H+, and thus can be replaced by NH4+ more easily. In 2015, Wang et
optimal performance was achieved, when hydroxide and silicate maxi- al. (2015) showed that NH4+ adsorption of modified wood biochar in-
mized, metakaolin minimized, and Na-based reagents used (Luukkonen

PR
creased to 5.44 mg g−1 by 3 folds after neutralizing biochar pH from 3.7
et al., 2017). Reutilization of discarded concrete as a kind of industrial to 7. However, others have shown that coating metal oxide on biochar
construction wastes was also investigated. Zhang et al. (2019) found surface can be used for simultaneous removal of other pollutants in ad-
that tricalcium aluminate, a mineral phase in Portland cement, showed dition to NH4+ capture. For example, Li et al. (2017) showed that 20%
maximum removal capacity of 155.4 mg g−1 at 298 K with dosage of MgO-coated biochar materials produced by an integrated adsorption-py-
6.5 g L−1. Additional advantages of such approach include saving land rolysis approach can adsorb 22 mg g−1, 398 mg g−1 and 247 mg g−1 for
resources by reducing secondary pollution to the landfill. NH4+, PO43+, and humate through electrostatic attraction, struvite pre-
cipitation, and π–π interactions mechanisms, respectively. Xu et al.

D
3.4. Biochar (2018) also demonstrated Mg-biochar has high removal capacity of
47.5 mg g−1 for NH4+ and 116.4 mg g−1 for PO43+ and the product ob-
Biochar obtained from carbonization of biomass feedstocks under tained after adsorption can be used as a nutrient-enriched fertilizer. The
oxygen-limited conditions has the potential to be a cost-effective and last but not the least, incorporation of additives during pyrolysis process
TE
highly efficient adsorbent material but in general exhibits relatively can further improve the product structures. A bentonite hydrochar com-
lower NH4+ adsorption capacity. Fidel et al. (2018) found that only posite prepared by Ismadji et al. (2016) at 500 °C under CO2 shows
0.84 mg g−1 NH4+ was adsorbed at around pH = 7 on biochar pre- NH4+ adsorption capacity of 23.67 mg g−1, which was greater than ei-
pared at low pyrolysis temperature (400 °C). However, Yu et al. (2016) ther sole bentonite (12.37 mg g−1) or biochar (9.49 mg g−1). This is as-
prepared two types of biochar using pig manure and straw biogas cribed to the higher surface area and pore volume in the composite
residues, which both exhibited good NH4+ adsorption capacities of which can capture more NH4+ from aqueous phase via van der Waals
EC

13.66 mg g−1 and 11.36 mg g−1, respectively. Yang et al., 2018 syn- force.
thesized biochar by pyrolyzing pine sawdust at 300 °C which exhib-
ited higher NH4+ adsorption capacity (5.38 mg g−1) than that at 550 °C 3.5. Activated carbon
(3.37 mg g−1), and ascribe it to higher H/C and O/C ratios of former
one (0.78 and 0.32) than the latter one (0.35 and 0.10). This suggests Activated carbon is usually produced from coal or polymer at rela-
that existence of oxygen-containing functional groups on biochar sur- tively high temperature (700–1000 °C) under atmospheric control with
RR

face is important. However, contribution of high surface area generated addition of chemical reagents (Quach et al., 2017). It is well known as
a versatile adsorbent materials because of the high surface area, porous
structure and rich surface chemistry (Bandosz, 2006). Mochizuki et
al. (2016) prepared activated carbon from petroleum coke and KOH
and ascribed the adsorbed NH3 to surface hydroxyl and carboxyl groups.
However, most common activated carbon lacks sufficient surface acid-
CO

ity, and thus only adsorbs NH3 through H bonding or Lewis acid-base in-
teractions, which were relatively weak (Seredych et al., 2009). More-
over, given that the average pore size of common activated carbon
(1–2 nm) is much larger than that of NH3 molecules (0.3 nm), the mol-
ecular trapping of NH3 within porous activated carbon is suppressed.
Therefore, activated carbon is generally regarded as a less efficient NH3
adsorptive material.
UN

Much effort has been dedicated to modifying or functionalizing the


activated carbon surface to enhance interaction between surface and
NH3 molecules. Main approaches include i) chemical or aerobic oxida-
tion to generate acidic functional groups that can capture alkaline NH3
and ii) combination of activated carbon with metals, oxides and chlo-
rides, which facilitates complexation reactions. In 2015, work of Qajar
et al. (2015) showed that adsorbed NH3 of synthesized nonporous car-
Fig. 4. Influence of agitation time on the NH4+ adsorption by kaolinite (Kaol), halloysite
bon material increased from 180 to 306 mg g−1 under 1 bar at 25 °C af-
(Hal), montmorillonite (Mt), vermiculite (Ver), palygorskite (Pal) and sepiolite (Sep) (Al- ter HNO3 treatment. Zheng et al. (2016) prepared carbon fiber com-
shameri et al., 2018). posites from phenolic precursor, which showed increased NH3 adsorp-
tion capacity from 10 to 50 mg g−1 after oxidation by concentrated
6 B. Han et al. / Journal of Cleaner Production xxx (xxxx) 124611

HNO3. In 2017, Travlou and Bandosz, 2017 prepared N-containing tively. Khabzina and Farrusseng, 2018 have summarized NH3 ad-
polymeric resin-derived carbons substrate that adsorb 73 mg g−1 NH3 at sorption capacity of many MOF materials as shown in Fig. 5.
1000 ppm after oxidation in air at 350 °C, which is about 45% higher
than that before oxidation. In the same year, Wang et al. (2017) 3.7. Other nanostructured materials
prepared mesoporous carbon beads for NH3 capture from atmosphere.
They found that the type and loading amount of transition metal chlo- Recently, there are emerging efforts on synthesizing novel nanos-

F
ride play important role. For example, impregnated mesoporous carbon tructured materials towards adsorption of NH3, which can be roughly
beads displayed increased breakthrough capacity of 111.4 mg g−1 from divided into inorganic and organic materials. Inorganic nanostructured
6.1 mg g−1 after NiCl2 loading for NH3 adsorption at 30 °C with a rela- adsorbent materials mainly include metal particles, metal complex and

OO
tive humidity of 80%. graphene-like nanosheets. Li et al. (2019) showed that the density of
NH3 molecules adsorbed on the inner wall surface of AgI-coated hol-
3.6. Metal organic frameworks (MOF) low waveguides was 1015 molecules cm−2. In 2016, Takahashi et al.
(2016) found that Prussian blue, first synthesized at the beginning of
MOFs are novel porous materials that provide immense possibilities 18th century, showed 212.5 mg g−1 of NH3 capacity under 0.1 MPa.
for design of NH3 adsorbent materials, but current study was mostly Recently, Silva et al. (2019) showed the dinuclear FeIII metallacy-
limited at laboratory scale due to the high cost (Rieth et al., 2019a; cle complex had excellent capacity of 192.1 mg g−1 at 25 °C to selec-
Vikrant et al., 2017). Jasuja et al. (2015) investigated six func- tively adsorb NH3 in comparison with amines, alcohols and hydrocar-

PR
tionalized variations of UiO-66 (a type of Zr-based MOF) towards NH3 bons. In 2017, Yang et al. (2017) reported boron doped graphene ox-
removal from air and found that UiO-66-OH has the highest adsorp- ide showed an increase of NH3 adsorption with increasing concentration
tion capacity of ~96.9 mg g−1 under dry conditions. DeCoste et al. of boron to as high as 16.30 mol kg−1. In 2018, Yang et al. (2018) pre-
(2016) prepared HKUST-1 membranes from polyvinylidene difluoride pared boron nitride nanosheets that has NH3 adsorption up to 91 mg g−1
for NH3 gas removal and found they exhibited outstanding structural at 1 bar.
stability in comparison with original powdered HKUST-1. Recently, Organic nanostructured adsorbent materials include polymer, ionic
Moribe et al. (2019) compared three isoreticular porphyrin-based liquid and hydrogel. (i) NH3 removal: Lee et al. (2017) compared

D
MOFs containing Brønsted acidic bridging hydroxyl groups and FTIR NH3 removal performance of porous polycyclic imide (PI) and porous
analysis suggested that the strength of the Brønsted acidic –OH sites poly(amic acid) (PAA) and found that PAA (27.2 mg g−1) is higher than
was determined by the identity of the metal center. In 2018, Chen PI (6.8 mg g−1) under 1 mbar, which is ascribed to the presence of
et al. (2018) synthesized a series of M(NA)2 (M = Zn, Co, Cu, Cd; amic acid functionality on PAA. In 2016, Ruckart et al. (2016) syn-
TE
NA = nicotinate) flexible MOFs and among them Co(NA)2 had the high- thesized a series of ionic liquid composites which exhibit NH3 load-
est NH3 uptake of 297.5 mg g−1. Rieth et al. (2016) designed a se- ing as high as 26.01 mg g−1 at a low pressure of 0.01 kPa and outper-
ries MOFs prepared from extended bisbenzenetriazolate linkers, which form many known materials such as MOF and activated materials. (ii)
possess coordinatively unsaturated metal sites that contribute to high NH4+ removal: NH4+ removal from aqueous phase has also been in-
and reversible uptake of NH3 and found that isostructural Mn, Co, vestigated. Wang et al. (2016) showed that amphoteric straw cellu-
and Ni materials adsorbed 263.0, 204.0, and 204.3 mg g−1 NH3 respec lose had a high NH4+ adsorption capacity of 68.4 mg g−1 at pH = ~7.0.
EC

Recently Cruz et al. (2018) reported on the rapid removal of NH4+


RR
CO
UN

Fig. 5. NH3 uptake vs water uptake at 40% relative humidity measured and 20 °C by breakthrough experiments for a NH3 concentration of 1200 ppm (yellow), 1294 ppm (purple) and
1431 ppm (blue). The yellow line indicates dissolved NH3 based on Henry's law with k°H = 70 mol kg−1bar−1 at 25 °C (Khabzina and Farrusseng, 2018). (For interpretation of the
references to colour in this figure legend, the reader is referred to the Web version of this article.)
B. Han et al. / Journal of Cleaner Production xxx (xxxx) 124611 7

on polymer hydrogels from domestic wastewater. The maximized ad- cation approach depends on factors like the initial concentration, waste-
sorption capacity is 32.2 mg g−1, which corresponded to removal effi- water component, capital investment and operational cost.
ciency of 80%. It is worth mentioning that porous 3D framework of hy- Despite above mentioned challenges, all those well-investigated ma-
drogels with high water content can decrease mass transfer resistance terials and modification methods form a solid basis for the development
of diffusion process and efficient regeneration can be obtained by mild of novel technologies which are inexpensive, environmentally friendly
acid washing. and highly efficient.

F
3.8. Comparison of adsorbent materials and modification methods 4. Mechanisms

OO
Considering the perspective of economy, naturally abundant re- Mechanism studies reveal common principles underlying adsorption
sources such as clay rather than artificial product (e.g. activated carbon, process, which would provide guidelines for systematic design of novel
nanostructured materials) were preferred for low concentration NH4+ efficient adsorbent materials. In this section we firstly try to identify
wastewater treatment. For example, the reported price of MOFs (~70 k adsorption sites on representative substrates based on published litera-
USD kg−1), biochar (0.35–1.2 USD kg−1) and activated carbon (1.8–2.1 ture. For gas phase NH3 adsorption, it is worth mentioning that the pres-
USD kg−1) are much higher than that of clay materials (0.04 USD kg−1) ence of water plays an important role in adsorption sites and mechanism
(Burakov et al., 2018; Thompson et al., 2016; Vikrant et al., pathways. For NH4+ adsorption in the aqueous phase, isotherm and ki-
2017). Hence application of MOFs is hindered by production cost that netic model fitting provide insights into chemical/physical adsorption

PR
is higher than other commercial products, although MOFs display out- pathways. At last, the effect of temperature on adsorption was explained
standing performance for the removal of NH3 (Chen et al., 2018; De- by thermodynamic analysis.
Coste et al., 2016; Mounfield et al., 2016; Rieth et al., 2016). Fur-
thermore, since nanostructured materials are generally expensive, they 4.1. Adsorption sites
must be regenerated by chemicals such as acid once exhausted, which
is a source of secondary pollution (Cruz et al., 2019). The poor sta- NH4+ adsorption sites: The NH4+ adsorption sites and underly-
bility also holds back deployment of MOFs because the empty voids ing mechanisms depend on the nature of the adsorbent materials and

D
of porous materials tend to collapse during adsorption process (Rieth the interactive processes. Generally, since NH3/NH4+ are small mole-
et al., 2019a). Those facts are detrimental to practical applications of cules/ions, their physical interactions with surface of adsorbent materi-
MOFs and many other nanomaterials. als are expected to be very weak. As to chemical adsorption, electrostatic
NH4+ adsorption capacities of biochar and activated carbon are com- interaction induces the transfer of target ion from liquid phase to solid
TE
parable or even lower than that of clay materials in many cases (Table surface, while ion exchange involves a reversible replacement of ions of
1). This is because performance of carbonaceous materials varies dra- the same charge on solid surface. So far intensive efforts have focussed
matically depending on their feedstock and treatment conditions. It is on novel modification methods to enhance chemical adsorption, which
reported that lower treatment temperature (<300 °C) leads to higher is preferable due to strong adsorption affinity.
density of oxygen-containing functionalities with strong bonding with It is generally agreed that biochar and activated carbon's capaci-
NH3/NH4+ (Yang et al., 2018). Nevertheless, the use of clay materi- ties to sorb NH4+ depend on their physical (surface area) and chemical
EC

als such as zeolite have some serious drawbacks that limited the adsorp- (functional groups) properties. However, Yu et al. (2016) found that
tion of clay as a full-scale technology. For example, powdered forms of the specific surface area, pore volume and micromorphology in biochar
clay materials require conduction in conventional packed-bed columns materials had no direct correlation with the NH4+ adsorption capacity.
which is not suitable for continuous operation within existing waste- Furthermore, Yang et al., 2018 found that the dominant mechanism
water treatment plant (WWTP) (Cooney et al., 1999). Furthermore, was chemical bonding and electrostatic interaction between NH4+ and
biofouling which causes a significant drop in adsorption capacity over surface functional groups. In this case, surface functional groups rather
RR

time (Damayanti et al., 2011; Lu et al., 2010) and high chemical re- than surface area of carbonaceous materials seem to play the domi-
quirement for regeneration (Wang and Peng, 2010; Weatherley and nant role for NH4+ adsorption. Furthermore, a fast adsorption kinetic
Miladinovic, 2004) were also considered as major drawbacks for ze- of NH4+ is preferred to obtain a high selectivity over other pollutants.
olite. While seawater can be potentially used as a cheap regeneration It was suggested that surface functional groups and porous structure ex-
chemical, it is only cost-effective at coastal area and scaling could be a hibit faster adsorption kinetics for small NH4+ than larger and nonpolar
concern (Cruz et al., 2019). molecules (Tu et al., 2019).
Various modification methods have been demonstrated effective to In contrast, clay materials exhibit versatile mechanism pathways.
CO

enhance capacity of adsorbent materials, but the cost of chemical ad- Similar to functionalized carbonaceous surface, electrostatic interac-
ditives and consumed energy must be considered for practical applica- tion could cause NH4+ adsorption to bentonite, which has negatively
tion. Functional coating could be a promising approach for modifica- charged surface due to the isomorphous substitution of Si4+ and Al3+.
tion of natural abundant clay materials if coating materials are low-cost The occurrence of Ca2+, Na+ and Mg2+ that balance electrons con-
and eco-friendly. Potential candidates include resin (Seredych et al., tributes to high ion-exchangeability of bentonite. However, Alshameri
2016; Zhang et al., 2016), chitosan (Yadi et al., 2016) and tannic et al. (2018) found that halloysite represent more negative surface
acid (Cheng et al., 2019). By contrast, acid/base treatment, calcina- charge than other natural clay materials but lower NH4+ adsorption,
UN

tion or sonication either consume energy or generate secondary pollu- indicating that the electrostatic interaction was not the only adsorp-
tion (Cheng et al., 2017; Luukkonen et al., 2017). In those cases, tion pathway. Instead, cation exchange was believed to be the dominant
chemical modification is not economically viable at a large scale. Car- mechanism for natural clay such as zeolite (Wang and Peng, 2010).
bonaceous materials are amenable to modification for varying surface This is further supported by observation of NH4+ deformation by FTIR
chemistry, but utilization of high temperature (Ismadji et al., 2016) analysis at 1430 cm−1 (Alshameri et al., 2018). This mechanism also
or chemical reagent such as HNO3 and NaOH (Qajar et al., 2015; allows NH4+-saturated clay material to be regenerated by Na or K ion
Zheng et al., 2016) should be avoided. In this sense, aerobic oxida- exchange besides heat treatment for NH3 devolatilization (Wang and
tion at relatively low temperature should be investigated as a potential Peng, 2010). Recently, Zhang et al. (2019) also demonstrated a dif-
approach to enhance adsorption behaviour (Jin et al., 2020; Travlou ferent removal pathway by monitoring Ca2+ and Al(OH)4- concentra-
and Bandosz, 2017). In addition, the selection of appropriate modifi tions in aqueous solution: Al(OH)4- from tricalcium aluminate can react
8 B. Han et al. / Journal of Cleaner Production xxx (xxxx) 124611

with NH4+ to produce Al(OH)3(H2O) and NH3 and contributes to the re- The pKa of the reaction is 9.3 at room temperature, which indicates
moval of NH4+. that the concentration of the ionized (NH4+) and molecular (NH3) forms
NH3 adsorption sites: Regarding to NH3 gas, different adsorption are equal at this pH value. At pH > 9.3, NH4+ tended to be hydrolysed
mechanisms have been reported. Thanks to the lone pair of electrons in to NH3·H2O, making it uncharged and resulting in the decreased adsorp-
NH3, Lewis and Brønsted acid sites can facilitate adsorption. Ye et al. tion capacity. Given that NH3 in solution can easily volatilate into the
(2016) investigated the atomic positions and interactions between ad- atmosphere, increasing the pH of the aqueous solution will further ac-
celerate the release of NH3 from the solution. Therefore, low NH4+ ad-

F
sorbed NH3 in zeolite by in situ synchrotron powder X-ray diffraction.
It was found that NH3 can interact strongly with the Brønsted acid site sorption capacity is expected for solution with pH > 9.3. On the other
to form NH4+ species (O … +HN), while another two NH3 molecules hand, exceedingly acidic pH will cause enrichment of proton in the so-

OO
can only form typical hydrogen bonding (N–H…N) along the sinusoidal lution, and the ionization of acidic functional groups on adsorbents are
channel. Silva et al. (2019) studied adsorption mechanism of FeIII met- restrained, leading to the decrease of ion-exchange positions for the ad-
allacycle complex. Thermogravimetric-mass spectrometry (TG-MS) and sorption of NH4+. For example, -COO- groups are converted into –COOH
NH3 isotherm experiments showed that 8 NH3 molecules were phys- groups at pH < 3, which decreases ion-exchange positions and electro-
iosorbed by the complex cavity and 4 NH3 molecules were chemisorbed static force between NH4+ and adsorbent surface, reducing the effective-
per dinuclear metallacycle unit. Nijem et al. (2015) showed that NH3 ness of adsorption process. Karri et al. (2018) summarized investiga-
adsorbed to the Cu2+ in HKUST-1 through Lewis base interactions, tions on the multiple ranges of pH (2–13) in a review, which concluded
while for Cu1+ the interaction is electrostatic in nature. Generally, in- that the efficiency of NH4+ removal is reduced for c.a. pH < 3 and pH

PR
creasing the amount of acidic functional groups can enhance NH3 cap- >7, whereas removal of NH4+ increases gradually with increase of pH
ture capacity (Lee et al., 2017; Moribe et al., 2019). Specifically, for a range of 3–8. Specifically, Eturki et al. (2012) investigated the ef-
Van Humbeck et al., 2014 prepared a functionalized organic poly- fect of solution pH on the NH4+ removal by clay materials and it was
mer which exhibited NH3 loadings of >289 mg g−1 under 1 bar. The shown that the best performance was achieved at an optimum pH be-
low-pressure uptake is also as high as ~14 mg g−1 at 0.01 kPa which tween 6 and 8 (Eturki et al., 2012). Wang et al. (2016) found that
is important for practical application. However, Jasuja et al. (2015) with pH increasing from 2 to 4, the adsorption capacity of amphoteric
found that bulky surface groups such as COOH and HSO3 can block straw cellulose (ASC) for NH4+ increased and then kept almost constant

D
small pores in framework, resulting in reduction of NH3 adsorption ca- in the pH range from 4 to 9. Therefore, both acidic and alkaline condi-
pacity. tions were not favourable for ion-exchange adsorption of NH4+ on ad-
sorbents, while the maximum equilibrium adsorption amount was usu-
4.2. Effect of water ally obtained at the pH of 6–9 (Dong et al., 2019a; Huang et al.,
TE
2018; Tu et al., 2019).
NH3 gas adsorption experiments are performed in both dry and hu-
mid air to investigate how water modulates the removal efficiency as 4.4. Effect of pore structure
shown in Table 3. Khabzina and Farrusseng, 2018 investigated NH3
uptake by functionalized UiO-66 and 5 different other MOF type ma- The effects of the pore structure of the adsorbent materials can be
terials using dynamic breakthrough experiments in both dry and hu-
EC

discussed from prospective of pore size distribution, pore stability and


mid condition. It was found that in the presence of water, the NH3 up- pore blockage by functional groups. Pore size distribution and void vol-
take by most MOFs increased with increasing water adsorption capac- ume are crucial parameters governing the sorption of NH3/NH4+. In
ity except Cu-based MOF. In these cases, the H-bonding between the general, adsorbents for NH3/NH4+ prefer a well-developed pore struc-
water and NH3 enhances the metal-NH3 interaction through coopera- ture, with the pore sizes comparable to the size of the NH3/NH4+ (3 Å)
tive interactions and water can promote acidic group dissociation and (Bashkova and Bandosz, 2014). This is imperative as the small sizes
NH3 protonation. In addition, water can also alter pathway of chemical of NH3/NH4+ lead to exceedingly weak retention forces for conven-
RR

reaction between NH3 and adsorbent materials. For example, the reac- tional adsorbents such as zeolites and activated carbon (Vikrant et al.,
tion between anhydride functional groups and NH3 were investigated 2017). Pore modulation has been employed for optimization of NH3 ad-
(Zheng et al., 2016). Whether the final products are in amide or imide sorption by MOFs. In this case the ability to finely tune the pore size
form depends on the presence of water and is critical for NH3 regen- and shape of MOFs to microporous dimensions is advantageous in re-
eration. In these cases, the presence of moisture in air was expected to moving NH3/NH4+ (Vikrant et al., 2017). However, stability remains
be beneficial for the NH3 adsorption process. Nevertheless, it is worth important, as pore collapse during use can affect the adsorption perfor-
CO

mentioning that water can deteriorate performance of zeolites (Grant mance (Rieth et al., 2019a). Some applications may require exten-
Glover et al., 2011). Seredych et al. (2016) also reported that on sive NH3 cycling stability. For instance, NH3 is a common impurity in
phenolic-formaldehyde resin/bentonite the adsorption capacity of NH3 feed gas streams and can poison catalysts and membranes, necessitat-
reached in dry conditions (38.0 mg g−1) is greater than that in moist ing the use of sorbents to capture NH3 prior to the desired chemical
conditions (12.4 mg g−1), and ascribe that to blocking of surface pores process. On a thermodynamic basis, NH3 is an excellent working fluid
by reaction products such as reversal of the exfoliation process. There- for heat transfer in adsorption heat pumps, which require many thou-
fore, the presence of water can either enhance or deteriorate NH3 gas sands of adsorption cycles and materials with extreme stability to this
adsorption depending on the mechanism pathway.
UN

corrosive gas (de Lange et al., 2015). The above-mentioned factors


are important for physical interactions, whereas in some cases chemical
4.3. Effect of water pH interactions play dominant role. Recently, we found that after dewater-
ing at 200 °C NH4+ adsorption by lignite increased from 2.25 mg g−1
Water pH is an important parameter that influences sorption to 3.73 mg g−1, while the BET surface area reduces from 2.92 m2 g−1
processes at water/adsorbent interfaces (Fu et al., 2020). In aqueous to 2.41 m2 g−1 (Han et al., 2020). Furthermore, the linear correlation
solutions, NH3 is a Brønsted base forming a conjugated pair with the between adsorbed NH4+ and the total concentration of acidic groups
NH4+ ion, according to the reversible reaction: on the lignite surface was high (R2 = 0.93). This is consistent with ob-
NH3+H+↔NH4+ (1) servations for NH3 adsorption on activated carbon prepared from pe-
troleum coke by KOH chemical activation (Mochizuki et al., 2016).
B. Han et al. / Journal of Cleaner Production xxx (xxxx) 124611 9

Table 2
Adsorption capacity and other parameters of adsorbent materials for NH4 + removal.

Capacity (mg Concentration range Optimized removal Contact Temperature


a
Adsorbent materials g −1) (mg L −1) efficiency (%) time (h) Size (μm) pH (°C) Ref.

Natural bentonite 19.01 10–1000 81.2 80 7.0 30 Angar et


Langmuir al. (2017)

F
Modified bentonite 5.85 0–350 71.7 1.0 7.5 25 Cheng et
Langmuir al. (2019)
Bentonite/Chitosan 15.90 10–40 100 3.0 6.0 25 Yadi et al.

OO
Langmuir (2016)
Bentonite/Chitin 16.16 10–40 100 3.0 6.0 25 Yadi et al.
Langmuir (2016)
Natural Na-bentonite 33.33 20–480 24.0 25 Zhou et al.
Langmuir (2015)
Natural Ca-bentonite 0.5 10.5 48 3.75 75 7.6 18 Sahin
(2018)
Natural Ca-bentonite 26.63 3–11 1.5 7.0 20 Sun et al.
Langmuir (2015)

PR
Purified bentonite 43.22 3–11 1.5 7.0 20 Sun et al.
Langmuir (2015)
Modified bentonite 46.90 3–11 1.5 7.0 20 Sun et al.
Langmuir (2015)
Chinese zeolite 13.18 5–500 10.0 6.8 26 Wang, X.L.
Langmuir et al.
(2016)
Synthesized zeolite 22.90 100 0.67 6.0 25 Liu et al.
(2018)
Zeolite/hydrogel 60.61 20–400 3.0 4040 ± 130 5.7 20 Putra and

D
Langmuir Lee (2020)
Modified zeolite 21.31 0–70 95.0 25 25 Zhang et
Langmuir al. (2016)
Iron oxide/zeolite 3.47 5–100 43.3 1.0 6.4 30 Xu et al.
TE
Langmuir (2020)
Natural vermiculite 22.61 10–1000 2.0 7.0 25 Alshameri
Langmuir et al.
(2018)
Natural montmorillonite 50.06 10–1000 2.0 7.0 25 Alshameri
Langmuir et al.
(2018)
Natural Palygorskite 40.40 10–1000 2.0 7.0 25 Alshameri
EC

Langmuir et al.
(2018)
Natural Sepiolite 21.51 10–1000 2.0 7.0 25 Alshameri
Langmuir et al.
(2018)
Natural Kaolinite 17.27 10–1000 2.0 7.0 25 Alshameri
Langmuir et al.
RR

(2018)
Natural Halloysite 15.58 10–1000 2.0 7.0 25 Alshameri
Langmuir et al.
(2018)
Modified birnessite 22.61 2–50 5.0 5.8 10 Cheng et
Langmuir al. (2017)
Natural halloysite 1.66 600 5.6 30 Jing et al.
(2017)
CO

Geopolymer 21.07 1000 80 24.0 63–125 6.0 22 Luukkonen


et al.
(2016)
Clinoptilolite-heulandite zeolite 14.42 1000 80 24.0 6.0 22 Luukkonen
et al.
(2016)
Tricalcium aluminate 155.40 92.5 4.0 25 Zhang et
al. (2019)
Polymer hydrogel 32.20 180 80 0.5 7.0 23 Cruz et al.
UN

(2018)
Biochar 0.84 300 24.0 7.0 Fidel et al.
(2018)
Biochar 13.66 100 4.0 5.0 30 Yu et al.
(2016)
10 B. Han et al. / Journal of Cleaner Production xxx (xxxx) 124611

Capacity (mg Concentration range Optimized removal Contact Temperature


a
Adsorbent materials g −1) (mg L −1) efficiency (%) time (h) Size (μm) pH (°C) Ref.

Biochar 5.38 100 16.0 7.0 20 Yang et


al., 2018
Biochar 5.86 250 120.0 25 Zheng et
al. (2018)

F
Biochar 114.84 3200 24.0 88–917 7.0 36 Fan et al.
(2019)
Biochar 2.86 12 43 1.5 7.5 23 Khalil et

OO
al. (2018)
Biochar 43.29 60–480 24.0 <500 7.0 20 Li et al.
Langmuir (2018)
Modified biochar 5.44 0–100 16.0 Wang et
Langmuir al. (2015)
Modified biochar 22 Li et al.
(2017)
Modified biochar 23.67 200 6.0 149–177 6.0 30 Ismadji et
al. (2016)

PR
Modified biochar 47.50 75 24.0 25 Xu et al.
(2018)
Straw cellulose 64.80 50 2.0 7.0 25 Wang et
al., 2016
Fe3O4 NPs 140 93.12 0.67 10.0 25 Zare et al.
(2016)
Lignite 3.43 50–450 2.0 25 Tu et al.
(2019)
Lignite 0.45 2–60 72.0 500–2000 6.0 25 ± 2 Nazari et
Langmuir al. (2018)

D
Base-washed lignite 0.67 2–60 72.0 500–2000 6.0 25 ± 2 Nazari et
Langmuir al. (2018)
Aged coconut shell-based activated 5.47 5–500 24.0 2000 25 Sumaraj et
carbon Langmuir al. (2020)
TE
Ni foam-supported activated carbon 5–7 20–500 1.3 7.0 Shih et al.
(2020)
Commercial activated carbon 0.5 660 24.0 600–2000 25 Carey et
al. (2015)
NaOH treated corncob activated 17.03 10–105 2.0 500–2000 7.0 20 Vu et al.
carbon Langmuir (2018)
Avocado seed-activated carbon 5.4 Langmuir 50–450 6.0 5000 5.0 25 Zhu et al.
(2016)
EC

Canola meal-based KOH activated 148 260 4.0 100–2500 9.0 20 Rambabu
carbon et al.
(2015)
Canola meal-based ammonia-treated 56.7 260 1.5 100–2500 9.0 20 Rambabu
activated carbon et al.
(2015)
Canola meal-based steam activated 55.4 260 2.0 100–2500 9.0 20 Rambabu
RR

carbon et al.
(2015)
Canola meal-based CO2 activated 17.9 260 0.5 100–2500 9.0 20 Rambabu
carbon et al.
(2015)

a The maximum adsorption capacities calculated from Langmuir model were shown as “number + Langmuir” and the adsorption capacities obtained by fixed NH + concentrations
4
were only displayed with “number”.
CO

From these results, it is suggested that the chemical interaction be- 4.5. Adsorption isotherms
tween NH3/NH4+ and surface acidic functional groups rather than the
pore structure would be dominant in adsorption processes. Nonethe- The most widely used adsorption isotherms are Langmuir (eq. (2))
less, those acidic functional groups sometimes have adverse effect. In and Freundlich (eq. (3)) models. The former assumes that adsorption
spite of the acidic chemical nature of functional groups such as –SO3H takes place at specific homogeneous sites on adsorbent material surface,
and –COOH that are favourable for the adsorption of alkaline NH3, also known as monolayer adsorption (Alshameri et al., 2018). The lat-
UN

UiO-66-SO3H and UiO-66-(COOH)2 showed lower NH3 sorption capaci- ter one recognizes surface heterogeneity and assumes that the adsorp-
ties than UiO-66-OH and UiO-66-NH2 (Jasuja et al., 2015). This is as- tion sites have various adsorption energy (Alshameri et al., 2018).
cribed to a significant decrease in the surface area and pore volume of Their linear equations are:
the MOF by the addition of excessively bulky groups such as –SO3H and
–COOH. These examples illustrate the importance of rigorous design of (2)
pore structure for improvement of adsorption property.
(3)
B. Han et al. / Journal of Cleaner Production xxx (xxxx) 124611 11

Table 3
Summary of NH3 gas adsorption conditions, capacities and methods for adsorbent materials.

Dry NH3 capacity (mg Wet NH3 capacity (mg NH3 gas concentration/
Adsorbent materials g −1) g −1) pressure Method Ref.

Carbon beads 111.4 200 ppm NH3 breakthrough Wang et al. (2017)
Bentonite/resin 38.0 1000 ppm NH3 breakthrough Seredych et al. (2016)

F
Modified zeolite 6.4 4.1 4500 ppm NH3 breakthrough Kang et al. (2015)
Nano porous carbon 306.0 1.0 bar Cyclic NH3 adsorption Qajar et al. (2015)
Activated carbon 50.0 22 500 ppm NH3 breakthrough Zheng et al. (2016)

OO
Activated carbon 73.0 1000 ppm NH3 breakthrough Travlou and Bandosz
(2017)
Boron doped graphene oxide 293.4 2.0 bar Isothermal NH3 Yang et al. (2017)
adsorption
Boron nitride nanosheets 91 1.0 bar Dynamic NH3 Yang et al., 2018
adsorption
Modified ionic liquid 27.5 39.6 0.01 kPa NH3 breakthrough Ruckart et al. (2016)
MOF 62.6 73.1 1500 ppm NH3 breakthrough Mounfield et al. (2016)
Multi-walled carbon nanotubes 22.69–90.05 0.368–0.744 MPa Isothermal NH3 Yan et al. (2015)

PR
adsorption
Nano TiO2-activated carbon 3.22–7.73 0.051 kPa NH3 breakthrough Rezaei et al. (2017)
composites

where is the maximum adsorption capacity (mg g−1), the concen- sorbed at any time (min) and equilibrium time, respectively; (min−1)
tration of NH4+ at equilibrium (mg L−1), the adsorption capacity (mg and (g mg−1 min−1) and (mg g−1 min−0.5) are the rate constants of
g−1) of NH4+ at equilibrium and the Langmuir adsorption constant; the three kinetics equations, respectively. is proportional to the ad-

D
is Freundlich adsorption constant (L g−1), the characteristic con- sorption rate and is inversely proportional to the adsorption rate.
stant related to adsorption capacity and adsorption strength. It is gener- (mg g−1) is the characteristic parameter representing the diffusion film
ally known that 1/n = 0.1–1 is favourable for adsorption while n = 1 thickness. For eq. (6), indicates that the adsorption is dominated
indicates a linear adsorption relationship, and the reaction is difficult to by the intra-particle diffusion and otherwise controlled by both film dif-
TE
occur when 1/n is greater than 2 (Tu et al., 2019). fusion and intra-particle diffusion. Furthermore, higher value of rep-
In many studies, the Langmuir equation represents experimental data resents significant film diffusion.
better than Freundlich (Table 4). Work by Ismadji et al. (2016) Generally, pseudo-second order kinetics represented a better correla-
showed that R2 of Langmuir model is closer to 1 than that of Fre- tion with the adsorption process of NH4+ (Table 5). According to Al-
undlich model for bentonite, biochar and bentonite hydrochar compos- shameri et al. (2018), well fitted pseudo-second-order indicates three
ite. However, Freundlich model also provides insights into the adsorp- steps involved: (i) the NH4+ diffusion from liquid phase to liquid-solid
EC

tion process. Cheng et al. (2019) found that the value of 1/n in Fre- interface; (ii) the NH4+ movement from liquid-solid interface to solid
undlich equation was ~0.2, indicating that the adsorption of NH4+ on surfaces; and (iii) the NH4+ diffusion into the particle pores. Xu et
modified bentonite was favourable. As temperature is increased, the al. (2020) found simultaneous removal of NH4+ and PO43+ by mod-
value of 1/n becomes lower, which suggests that the adsorption process ified zeolite is more consistent with the pseudo-second order kinetic
is an exothermic process. Furthermore, Xu et al. (2020) observed the and ascribed it to either electrostatic adsorption or ligand exchange
value of 1/n < 1 for NH4+ adsorption onto modified zeolite and as- between NH4+-PO43+ and modified zeolite, but individual NH4+ re-
RR

cribed it to chemical adsorption. Other mathematic models were also moval was due to electrostatic physical adoption. Wang et al. (2016)
explored. Recently, Putra and Lee, 2020 used the Dubinin-Radushke- found NH4+ adsorption onto amphoteric straw cellulose (ASC) fitted
vitch isotherm model to describe NH4+ adsorption process on zeo- better with pseudo-second-order model, which implies the existence of
lite-entrapped hydrogel and found that the obtained mean free energy both chemisorption and physisorption. The chemisorption might be the
was higher than 8 kJ mol−1, indicating the adsorption process is mainly rate-determining step, where valent or coordinate bonds were intro-
through chemical ion exchange mechanism. duced by sharing or exchange of electrons between adsorbate ions and
adsorbent material.
CO

4.6. Adsorption kinetics The intraparticle diffusion model was also employed when both
boundary layer diffusion and intera-particle diffusion are rate-limiting
Three kinetic models have been proposed, pseudo-first order by steps. Wang et al. (2016) found intraparticle diffusion plots onto
Lagergren (eq. (4)) (Lagergren, 1898), pseudo-second order by Ho and amphoteric straw cellulose (ASC) adsorbent material exhibit multilin-
Mckay (eq. (5)) (Ho and McKay, 1999), and intra-particle diffusion earity. Putra and Lee, 2020 compared zeolite particles (ZPs), zeo-
by Weber and Morris (eq. (6)) (Weber and Morris, 1963). The linear lite beads (ZBs) and polyvinyl alcohol alginate-ZPs (PAZ) and found
form of these three model equations can be expressed as: that the overall adsorption process is divided into one linear segment
UN

for ZBs and two linear segments for ZPs and PAZ (Fig. 6). Therefore,
(4) NH4+ adsorption onto ZBs is mainly controlled by intra-particle dif-
fusion, while the rate-determining steps using ZPs and PZA are both
(5) boundary layer diffusion (the first linear stage) and the intraparticle
diffusion (the second linear stage). Tu et al. (2019) compared diffu-
(6) sion kinetics of NH4+ and naphthalene sulfonic acid sodium formalde-
hyde condensate (NSF) which is a considerably larger molecule. Both
where (mg g−1) and (mg g−1) represent the amount of NH4+ ad
obtained values of by eq. (6) are not 0, indicating the adsorption
process were controlled by both liquid film diffusion and internal diffu
12 B. Han et al. / Journal of Cleaner Production xxx (xxxx) 124611

Table 4
Isotherm parameters of Freundlich and Langmuir models for NH4 + adsorption.

Adsorbent materials Temperature (°C) Freundlich Langmuir Ref.

1/n kf (mg g −1) R2 qs (mg g −1) k1 (L mg −1) R2

Modified bentonite 30 0.24 1.79 0.882 5.85 0.077 0.999 Cheng et al. (2019)

F
Natural bentonite 30 0.67 0.61 0.980 50.00 0.006 0.990 Angar et al. (2017)
Bentonite/Chitosan 25 2.33 1.96 0.990 15.90 0.050 0.971 Yadi et al. (2016)
Bentonite/Chitin 25 2.38 2.05 0.990 16.16 0.050 0.973 Yadi et al. (2016)

OO
Natural Ca-bentonite 20 0.76 2.84 0.897 26.63 0.111 0.936 Sun et al. (2015)
Purified bentonite 20 0.82 3.37 0.979 43.22 0.081 0.993 Sun et al. (2015)
Modified bentonite 20 0.77 6.86 0.973 46.90 0.111 0.996 Sun et al. (2015)
Natural zeolite 26 0.42 1.21 0.878 13.18 0.021 0.981 (Wang, X.L. et al., 2016)
Synthesized zeolite 25 0.44 3.00 0.832 25.13 0.006 0.997 Liu et al. (2018)
Modified zeolite 30 0.42 1.05 0.708 5.44 0.100 0.955 Xu et al. (2020)
Modified zeolite 25 0.40 4.24 0.996 21.31 0.130 0.936 Zhang et al. (2016)
Modified birnessite 10 0.51 3.17 0.904 22.61 0.147 0.982 Cheng et al. (2017)
Natural vermiculite 0.41 8.81 0.965 50.06 0.301 0.967 Alshameri et al. (2018)

PR
Natural montmorillonite 0.51 6.81 0.957 40.40 0.227 0.976 Alshameri et al. (2018)
Natural Palygorskite 0.70 5.21 0.945 21.51 0.223 0.968 Alshameri et al. (2018)
Natural Sepiolite 0.55 5.05 0.878 17.27 0.156 0.954 Alshameri et al. (2018)
Natural Kaolinite 0.50 4.00 0.964 15.58 0.143 0.926 Alshameri et al. (2018)
Natural Halloysite 0.97 2.14 0.906 9.97 0.092 0.939 Alshameri et al. (2018)
Natural Halloysite 15 0.42 0.06 0.981 0.51 0.043 0.944 Jing et al. (2017)
Geopolymer 22 0.27 4.45 0.833 22.82 0.051 0.950 Luukkonen et al. (2016)
Biochar 20 0.51 0.40 0.970 5.38 0.026 0.940 (Yang, H.I. et al., 2018)
Modified biochar 0.41 1.05 0.987 5.44 0.140 0.988 Wang et al. (2015)
Modified biochar 0.17 20.13 0.885 22.10 0.018 0.989 Li et al. (2017)

D
Modified biochar 30 0.34 2.13 0.941 12.37 0.050 0.983 Ismadji et al. (2016)
Modified biochar 25 3.13 11.17 0.870 51.48 0.093 0.963 Xu et al. (2018)
Straw cellulose 25 0.63 6.63 0.831 94.90 0.048 0.909 (Wang, X.G. et al., 2016)
Lignite 25 0.47 0.15 0.964 3.43 0.007 0.992 Tu et al. (2019)
TE
Lignite 25 ± 2 0.73 0.02 0.992 0.45 0.034 0.975 Nazari et al. (2018)
Base-washed lignite 25 ± 2 0.43 0.17 0.957 0.67 0.392 0.997 Nazari et al. (2018)
Avocado seed-activated carbon 25 0.45 0.26 0.967 5.4 0.006 0.995 Zhu et al. (2016)
NaOH treated corncob activated carbon 20 2.05 1.61 0.965 17.03 0.039 0.995 Vu et al. (2018)
Aged coconut shell-based activated carbon 25 0.75 0.04 0.980 5.47 0.003 0.98 Sumaraj et al. (2020)
EC

Table 5
Pseudo-first-order and pseudo-second-order kinetic model constants for the removal of NH4 +.

Adsorbent Temperature
materials (°C) Pseudo-first order Pseudo-second order Ref.

qe qe
RR

(exp.) (cal.)
k1 qe (exp.) (mg qe (cal.) (mg k2 (g mg −1 (mg (mg
(min −1) R2 g −1) g −1) min −1) R2 g −1) g −1)

Lignite 25 0.533 0.708 1.31 0.33 5.949 0.999 1.31 1.27 Tu et al. (2019)
Modified bentonite 5 0.016 0.797 5.00 4.90 0.194 0.999 4.98 4.90 Cheng et al.
(2019)
Bentonite/Chitosan 25 1.29 1.29 Yadi et al. (2016)
CO

Bentonite/Chitin 25 1.04 1.01 Yadi et al. (2016)


Modified zeolite 25 0.028 0.957 2.80 0.012 0.993 Zhang et al.
(2016)
Modified zeolite 20 0.378 0.927 37.99 0.002 0.977 3.93 Xu et al. (2020)
Tricalcium 25 0.358 0.778 37.30 0.067 0.999 136.90 Zhang et al.
aluminate (2019)
Natural zeolite 0.016 0.898 2.46 2.759 0.944 2.80 (Wang, X.L. et al.,
2016)
Straw cellulose 25 0.247 0.944 18.4 0.015 0.994 21.70 (Wang, X.G. et al.,
UN

2016)

sion. Furthermore, the calculated value of NH4+ was greater than


that of NSF. This is ascribed to (i) the rich pore structure of lignite
facilitates easy transport of the small NH4+ ion to pores; (ii) ion ex-
change of NH4+ with H+ ionized from surface acidic functional groups
B. Han et al. / Journal of Cleaner Production xxx (xxxx) 124611 13

Table 6
Thermodynamic parameters for NH4 + adsorption.

ΔH° ΔS° (J
(kJ mol −1
Sample mol −1) K −1) ΔG° (kJ mol −1) Ref.

Lignite −6.16 −6.39 −14.33 −14.04 −13.76 Tu et

F
(298 K) (308 K) (318 K) al.
(2019)
Modified −80.56 −0.79 −13.91 −8.87 −7.37 Cheng

OO
bentonite (278 K) (298 K) (308 K) et al.
(2019)
Synthesized −11.61 −30.00 −2.60 −2.37 −1.99 Liu et
zeolite (298 K) (308 K) (318 K) al.
(2018)
Modified −3.40 9.60 −0.60 −0.48 −0.41 Xu et
zeolite (293 K) (303 K) (313 K) al.
(2020)
Modified −8.12 35.19 −17.95 −18.50 −19.05 Cheng

PR
birnessite (283 K) (293 K) (303 K) et al.
(2017)
Tricalcium 21.96 5.89 4.36 3.92 3.17 Zhang
aluminate (298 K) (308 K) (318 K) et al.
(2019)
Biochar −8.07 41.6 −3.48 −3.79 −4.11 Fan et
(277 K) (285 K) (293 K) al.
(2019)
Straw −38.53 154.8 −7.59 −9.14 −10.68 (Wang,
cellulose (298 K) (308 K) (318 K) X.G. et

D
al.,
2016)
TE
ture is favourable. In other words, the dissociation of ions (i.e., Na+
in natural zeolites) from the adsorbent materials requires more energy
than that released by combining with NH4+ (Fu et al., 2020). Cheng
et al. (2019) found that the obtained values of were consistent
with the Freundlich isotherm analysis where the value of 1/n decreases
EC

Fig. 6. NH4+ adsorption data fitted by Weber and Morris’ intra-particle diffusion model with increasing temperature (Cheng et al., 2019). This is in contrast
(ZP: zeolite particles, PAZ: polyvinyl alcohol (PVA)-alginate-ZPs, ZB: zeolite bead) (Putra
with the removal of Zn2+ and Cr3+ by Al-pillared bentonite, which are
and Lee, 2020).
endothermic reactions and not favourable at low temperature (Ghnimi
and Frini-Srasra, 2018). The positive ΔS implies that the adsorption
(e.g. –OH and –COOH) is faster than adsorption of NSF at hydrophobic
process creates randomness. Wang et al. (2016) observed ΔS > 0 indi-
sites driven by intermolecular forces on the lignite surface.
cating that the affinity of the adsorbent materials toward NH4+ and also
RR

suggest the randomness at the solid/liquid interface increased as a re-


4.7. Thermodynamic analysis
sult of adsorption. Tu et al. (2019) compared adsorptive behaviour of
naphthalene sulfonic acid sodium formaldehyde condensate (NSF) and
Thermodynamic analysis can provide valuable information to the in-
NH4+ on lignite and it was found that the adsorption of NH4+ was
ternal energy changes during the adsorption process, which indicates
ΔH < 0 and ΔS < 0, whereas the adsorption of NSF was ΔH > 0 and
the effect of temperature on NH4+ adsorption. The parameters are de-
ΔS > 0, but was observed for both adsorbates. The adsorption
termined from (the thermodynamic distribution coefficient) and the
CO

onto NaOH-treated activated carbon was a spontaneous process (Vu et


standard Gibbs free energy (kJ mol−1), standard enthalpy change
al., 2018a). Furthermore, the positive ΔS suggests that the organiza-
(kJ mol−1), and standard entropy change (J mol−1 K−1) are cal-
tion of NH4+ ions at the solid/solution interface during the adsorption
culated using Eqs.
process becomes more random. Additionally, the negative ΔH° reflects
(7) the exothermic nature of the adsorption process demonstrated by the
decrease in the adsorption capacity. However, the thermodynamic para-
(8) meters show that the adsorption process on clinoptilolite was non-spon-
UN

taneous and adsorption rate decreased with an increase in the temper-


where (K) was the absolute temperature, and R was molar gas con- ature, thus showing the exothermic nature of the adsorption (Tosun,
stant. The value of , and were calculated from the slope 2012). Therefore, both adsorbate and adsorbent materials determined
and intercept of the plot of versus 1/T. the thermodynamic behaviour of adsorption process.
Table 6 listed thermodynamic parameters for NH4+ adsorption
process. The negative values of in most studies indicate that the ad- 4.8. Field trials
sorption process was spontaneous. The reaction is generally considered
to proceed via physical sorption (ΔG value between 0 and -20 kJ mol−1) Laboratory experiments alone are unable to give a comprehensive
and chemical sorption (ΔG value between −80 and −400 kJ mol−1) picture of the dynamic removal of NH4+. Therefore, field trials should
(Bhatt et al., 2012). observed by many studies indicate that be carried out in order to validate the practical application. Wang et
adsorption of NH4+ is exothermic and low tempera al. (2016) explored the dynamic adsorption efficiency of NH4+ in
14 B. Han et al. / Journal of Cleaner Production xxx (xxxx) 124611

wastewater from a paddy rice field using natural Chinese zeolite during crease of NaCl ratio from 0 to 10 wt%, whereas there was no significant
rainfall runoff, which indicates that the adsorption rate was the highest change above 10 wt%. However, NaCl regeneration requires periodical
when the base angle of adsorption barrier is 45°. Moreover, a high initial replacement due to the accumulation of NH4+ in the regenerant solu-
removal rate was obtained at low flow rates while high adsorption was tion, which resulted in secondary pollution that needs further treatment
achieved at high flow rates. Ismadji et al. (2016) demonstrated that (Zhang et al., 2017). Alkaline regeneration by NaOH solution success-
5 kg bentonite-hydrochar composite could almost completely remove fully avoided NH4+ accumulation by air stripping but was restrained by
the NH4+ from the Koi fish tank system (500 L of water and 30 Japanese

F
its high cost in regenerant and gaseous pollution control (Sancho et al.,
Koi) within 60 min. This meets of healthy aquaculture requirement of 2017; Zhang et al., 2017). Similarly, regeneration with HCl did not
NH4+ content <0.1 ppm. Zhang et al. (2019) applied tricalcium alu- prove to be more beneficiary than regeneration with NaCl as the low pH

OO
minate (a mineral phase of discarded concrete) to remove NH4+ from of spent regenerants had implications in consequent treatments (Huang
digested piggery wastewater and demonstrated a maximum water treat- et al., 2018). Therefore, the high chemical requirements for regener-
ment capacity of 200 L water kg−1 that meets the discharge standard of ation was also considered a major drawback. Generally, NH4+ loading
<80 mg L−1 for livestock and poultry breeding (Fig. 7). Furthermore, capacity decreased after each sorption-desorption cycle, probably due to
the recovery rate of NH3 was 90.5%. Zhang et al. (2016) used modi- an incomplete desorption of NH4+ which results in less available sites
fied zeolite to treat secondary effluent from WWTP in Nanjing. The coat- on the adsorbent surface after each cycle (You et al., 2017). Recently,
ing of an anion exchange resin onto the Na-form zeolite increased the Fu et al. (2020) showed that the NH4+ removal rate of natural zeolite
NH4+ removal efficiency from 78% to 95% and increased the treatment is 81.7%, and that of the recycled zeolite fluctuates within 79.4–84.8%

PR
volume of the Na-form zeolite from 51 to 76 bed volume. Those field tri- over six cycles of regeneration. Nonetheless, the presence of alkaline or
als would provide precious experience for NH4+ removal in real world oxidative components could lead to more favoured sorption conditions
applications, but more field trials are needed for comprehensive study. (You et al., 2017). Zhang et al. (2017) found that zeolites regen-
erated by mixed NaClO–NaCl solution showed higher NH4+ adsorption
4.9. Regeneration rate and lower capacity than unused zeolite.
Compared to chemical washing, the biological regeneration method
Regeneration and reusability of adsorbents are pivotal to the ability was developed to remove adsorbed NH4+ from zeolite through nitri-

D
to recycle NH3/NH4+ during successive rounds of adsorption and des- fication and/or denitrification process. It avoids the cost of chemical
orption processing, because one-time use of adsorbent materials is not reagents or complicated configuration and thus shows advantages in an
economically feasible. The key is to develop cost-effective adsorbents environmental-friendly and economical manner. Hence, biological re-
that are viable for regeneration that would yield a concentrated NH4+ generation was the relatively expected choice, even though the regen-
TE
stream and completely desorb the NH4+ without affecting the material's eration rate is generally not as good and fast as other methods. Han et
structural integrity and adsorption performance over a guaranteed ser- al. (2019) treated anaerobically digested dispersed swine wastewater
vice life. Well-demonstrated options include chemical washing, biologi- with a pilot-scale zoning tidal flow constructed wetland (TFCW) with a
cal regeneration, electrochemical regeneration, mild heating and use of bottom wastewater saturation layer, and found that nitrification-denitri-
an electric field (; Han et al., 2019; Rahmani et al., 2009; Shelp and fication which accounted for 80.32% of total N removal was the main N
Seed, 2007). removal pathway.
EC

Chemical washing involves displacement of the target ions into the Other cost-effective regeneration methods that do not rely heavily on
acid/base/salt solution and has been the most commonly used method chemical dosing were also developed for regeneration of NH4+ adsor-
for regenerating adsorbents, where low regenerant volume and high re- bents. Thermal regeneration of adsorption materials is an interesting op-
generation rate were preferred (Malovanyy et al., 2013). A single tion among them. An increase in temperature or vacuum treatment may
component or mixture of NaCl, NaClO, KOH, and NaOH in solution were account for the desorption of absorbed NH3 to prepare the MOFs for the
selected as the regenerant (Cruz et al., 2019; You et al., 2017; Zhang next round of NH3 uptake (Vikrant et al., 2017). Depending on the
RR

et al., 2016, 2017bib_Zhang_et_al_2016bib_Zhang_et_al_2017). binding strength of NH4+ ions onto the functional groups in the adsor-
Hermassi et al. (2017) showed that NH4+ desorption efficiency in- bent surface, the removal of NH4+ ions can be initiated by heat through
creased with in the thermal transformation of NH4+ into NH3. In particular, the fact that
heat can be used for regeneration purposes offers opportunities to ben-
eficially reuse low value heat released during electricity generation us-
ing combined heat power (CHP) systems at Wastewater Treatment Plant
CO

(WWTPs) that implement anaerobic digestion. This requires the struc-


tural integrity of the adsorbent materials to remain intact after multiple
thermal regeneration cycles, which is a major concern that needs to be
taken into consideration during material development.

5. Outlook

Finding suitable adsorbent materials constantly remains a challenge.


UN

Lignite (brown coal) which has rich pore structure and high density of
surface functionalities has emerged as a promising adsorbent material
for cations. Although there is no academic reference, the market price
of lignite is about 16.5 times lower than that of natural zeolite (Nazari
et al., 2018). Particularly in Australia, the Latrobe Valley in Victoria
has around 25% of the world's known lignite reserves, which is mined
in huge quantities mainly for burning to generate electricity (Barton et
al., 1993). Recently, Tu et al. (2019) reported that NH4+ adsorption
Fig. 7. Treatment capacity of tricalcium aluminate for NH4+ contaminated digested pig- capacity of lignite is 3.43 mg g−1 using Langmuir analysis and the ad
gery wastewater (Zhang et al., 2019).
B. Han et al. / Journal of Cleaner Production xxx (xxxx) 124611 15

sorption process is driven by ion-exchange interactions with H+ of –OH ter with one another in terms of removal efficiency, due to the com-
and –COOH on the lignite surface. In 2017, Nazari et al. (2018) ex- plexity of the wastewater treatment process. Moreover, once laboratory
plored the influence of pH, NH4+ concentration and lignite dosage for experiments are accomplished, work should be directed to the design
removal of NH4+ from synthetic wastewater. It was found that lignite and implementation of pilot-scale studies to validate their industrial fea-
treated with base solution has greater NH4+ removal efficiency in com- sibility. Importantly, real wastewater rather than synthetic wastewater
parison with raw material. In a field trial, Chen et al. (2015) found should be tested, where other components in solution such as heavy

F
that application of 4.5 kg m−2 lignite significantly reduced NH3 gas loss metal ions and organics will reduce ammonium exchange capacity due
from cattle pen by about 66% (Fig. 8). Nevertheless, lignite's optimal to competitive adsorption (Wang and Peng, 2010).
adsorption for NH3 and detailed mechanisms of lignite-NH3 interaction The reutilization and disposal of adsorbed NH4+ together with ad-

OO
remains elusive. In addition, regeneration of lignite using simple and sorbent materials is another challenge. While in some regions food pro-
cost-effective methods for adsorbent material recycling should also be duction leads to vast losses of N to the environment, in other regions
investigated in future research and whether lignite has toxicity to the not enough N is available for sustainable agricultural crop production.
subsequent biological treatment must be clarified. Hence there is a need to develop alternative NH4+ management ap-
Both chemical and physical modifications are important for lignite to proaches that center around recovery of NH4+ rather than deal with
be used as an adsorbent material. It was reported that treatment using its destruction into elemental dinitrogen. The NH3/NH4+-saturated ad-
chemical reagents can alter the surface chemistry of lignite (Cihlář et sorbent materials have been widely proposed to be used as fertilizer
al., 2014; Tu et al., 2019), potentially enhancing their NH4+ and NH3 which have advantages over traditional treatments. Such approaches not

PR
capture capacity (Fig. 9). However, modifying a large quantity of lig- only to reduce the capital and energy costs required in the adsorption
nite via addition of chemical reagents during the manufacturing process processes by supplementing industrial Haber-Bosch process, but also re-
is not always ideal because this can produce secondary pollution such as duce the environmental footprint by reducing N2O production which
formaldehyde and surfactants (Cihlář et al., 2014; Tu et al., 2019). is a potent greenhouse gas comprising the equivalent of 14–26% of
On the other hand, physical modifications of lignite such as dewatering, the overall carbon footprint of wastewater treatment plants (WWTPs)
pelletisation, granulation are even crucial. For example, by decreasing (Cruz et al., 2019; Service, 2018). For example, after the last sorp-
water content (~60% dry weight of lignite), the transportation cost can tion desorption working cycle, loaded zeolites can be used as fertilizer

D
be dramatically decreased; Turning lignite powder into porous pellets, after a separation process by filtration (Hermassi et al., 2017). Zeo-
granules or concrete bricks can not only prevent humic acid dissolving lite, biochar and other adsorbent materials were reported to be capa-
into water which darkens the colour and increases the dissolved organic ble for improving soil physicochemical properties, microbial capacity
carbon (DOC) content of water, but also improves animal well-being and soil fertility (Ding et al., 2016). When adsorbent materials exhibit
TE
when used in animal feedlot by reducing the dust content. Thus, more strong binding to NH3/NH4+, the loaded adsorbent materials could be
efforts should focus on developing effective “green” modification meth- used as a carrier slow-release fertilizer (Kotoulas et al., 2019). Note
ods to enhance the functions but minimize the drawbacks of lignite. that the concentrated effluent is also a promising precursor for fertilizer
In the past, most tests for NH4+ removal of lignite were conducted production after the concentration stage, i.e., using liquid-liquid mem-
in short-term batch experiments using synthetic solutions (Nazari et brane contactors (Sancho et al., 2017). However, only a few stud-
al., 2018; Tu et al., 2019), focusing on adsorption capacity and re- ies have investigated recycling of NH4+ for nutrients release (Malekian
EC

moval efficacy. However, dynamic breakthrough analysis is lacking, and et al., 2011). Wang et al. (2016) prepared a novel multifunctional
more efforts should be dedicated to the interaction of each parame slow-release compound fertilizer using recovered NH4+ and H2PO4−
from aqueous solutions. There is a bright future for marketing recycled
N. In Australia, 1.678 Mt urea was applied in 2017 which is an equiva
RR
CO
UN

Fig. 8. Hourly NH3 emissions and air temperature from 13.11.4 to 13.12.13 where cattle moved in pens at 9–11 a.m. on 4th November and moved out at 1–3 pm (Chen et al., 2015).
16 B. Han et al. / Journal of Cleaner Production xxx (xxxx) 124611

Angar, Y., Djelali, N.E., Kebbouche-Gana, S., 2017. Investigation of ammonium adsorption
on algerian natural bentonite. Environ. Sci. Pollut. Res. 24 (12), 11078–11089.
Ardern, E., Lockett, W.T., 1914. Experiments on the oxidation of sewage without the aid
of filters. 33 (10), 523–539.
Bahmani, M., Dashtian, K., Mowla, D., Esmaeilzadeh, F., Ghaedi, M., 2020.
UIO-66(Ti)-Fe3O4-WO3 photocatalyst for efficient ammonia degradation from
wastewater into continuous flow-loop thin film slurry flat-plate photoreactor. J.
Hazard Mater. 393, 122360.

F
Bandosz, T.J., 2006. Activated Carbon Surfaces in Environmental Remediation. Elsevier.
Barton, C.M., Gloe, C.S., Holdgate, G.R., 1993. Latrobe valley, Victoria, Australia - a
world-class brown-coal deposit. Int. J. Coal Geol. 23 (1–4), 193–213.
Bashkova, S., Bandosz, T.J., 2014. Effect of surface chemical and structural heterogeneity

OO
of copper-based mof/graphite oxide composites on the adsorption of ammonia. J.
Colloid Interface Sci. 417, 109–114.
Bhatt, A.S., Sakaria, P.L., Vasudevan, M., Pawar, R.R., Sudheesh, N., Bajaj, H.C., Mody,
H.M., 2012. Adsorption of an anionic dye from aqueous medium by organoclays:
equilibrium modeling, kinetic and thermodynamic exploration. RSC Adv. 2 (23),
Fig. 9. Modification of lignite to enhance NH4+ adsorption. 8663–8671.
Blazquez, E., Bezerra, T., Lafuente, J., Gabriel, D., 2017. Performance, limitations and
microbial diversity of a biotrickling filter for the treatment of high loads of ammonia.
lent amount of N (0.772 Mt) as that contained in 15.44 Mt manure dry Chem. Eng. J. 311, 91–99.
solid (5% N), but feedlot herds are only estimated to produce 3.29 Mt Bodirsky, B.L., Popp, A., Lotze-Campen, H., Dietrich, J.P., Rolinski, S., Weindl, I., Schmitz,

PR
per annum. Nevertheless, in this case the recycled N from feedlots alone C., Muller, C., Bonsch, M., Humpenoder, F., Biewald, A., Stevanovic, M., 2014.
Reactive nitrogen requirements to feed the world in 2050 and potential to mitigate
has the potential to supply about 21% of the N fertilizer market share nitrogen pollution. Nat. Commun. 5, 7.
in Australia. Therefore, the potential of NH4+ adsorption to generate Boopathy, R., Karthikeyan, S., Mandal, A.B., Sekaran, G., 2013. Adsorption of ammonium
value-added product is very evident, despite challenges that remain to ion by coconut shell-activated carbon from aqueous solution: kinetic, isotherm, and
thermodynamic studies. Environ. Sci. Pollut. Res. 20 (1), 533–542.
be addressed. Burakov, A.E., Galunin, E.V., Burakova, I.V., Kucherova, A.E., Agarwal, S., Tkachev, A.G.,
Gupta, V.K., 2018. Adsorption of heavy metals on conventional and nanostructured
6. Conclusion materials for wastewater treatment purposes: a review. Ecotoxicol. Environ. Saf. 148,
702–712.

D
Busca, G., Pistarino, C., 2003. Abatement of ammonia and amines from waste gases: a
Effective removal of NH4+ and mitigation of NH3 emissions from summary. J. Loss Prev. Process. Ind. 16 (2), 157–163.
various hotspots are critical for not only protection of water resources Carey, D.E., McNamara, P.J., Zitomer, D.H., 2015. Biochar from pyrolysis of biosolids
for nutrient adsorption and turfgrass cultivation. Water Environ. Res. 87 (12),
but also reduction of atmospheric PM2.5 pollution. Adsorption by
2098–2106.
TE
low-cost abundant materials is regarded as an effective approach, and Chen, D.L., Sun, J.L., Bai, M., Dassanayake, K.B., Denmead, O.T., Hill, J., 2015. A new
promising modification methods have been investigated to increase ad- cost-effective method to mitigate ammonia loss from intensive cattle feedlots:
application of lignite. Sci. Rep. 5, 5.
sorption capacity of NH3/NH4+. Although both conventional and nanos-
Chen, Y., Shan, B.H., Yang, C.Y., Yang, J.F., Li, J.P., Mu, B., 2018. Environmentally
tructured materials face challenges regarding economic cost, energy friendly synthesis of flexible MOFs M(NA)2 (M = Zn, Co, Cu, Cd) with large and
consumption, secondary pollution and adsorption efficiency, these can regenerable ammonia capacity. J. Mater. Chem. A 6 (21), 9922–9929.
be tackled by adopting a variety of advanced options. We envision that Cheng, Y., Huang, T.L., Shi, X.X., Wen, G., Sun, Y.K., 2017. Removal of ammonium ion
EC

from water by na-rich birnessite: performance and mechanisms. J. Environ. Sci. 57,
the pursuit of strategies for surface modification of natural abundant re- 402–410.
sources will lead to a bright future of adsorbent materials suited to low Cheng, H.M., Zhu, Q., Xing, Z.P., 2019. Adsorption of ammonia nitrogen in low
NH3/NH4+ concentration conditions. temperature domestic wastewater by modification bentonite. J. Clean. Prod. 233,
720–730.
Cihlář, Z., Vojtová, L., Conte, P., Nasir, S., Kučerík, J., 2014. Hydration and water holding
Uncited reference properties of cross-linked lignite humic acids. Geoderma 230–231, 151–160.
Cooney, E.L., Booker, N.A., Shallcross, D.C., Stevens, G.W., 1999. Ammonia removal from
wastewaters using natural australian zeolite. I. Characterization of the zeolite. Separ.
RR

,,. Sci. Technol. 34 (12), 2307–2327.


Cruz, H., Luckman, P., Seviour, T., Verstraete, W., Laycock, B., Pikaar, I., 2018. Rapid
Declaration of competing interest removal of ammonium from domestic wastewater using polymer hydrogels. Sci. Rep.
8, 6.
Cruz, H., Law, Y.Y., Gues, J.S., Rabaey, K., Batstone, D., Laycock, B., Verstraete, W.,
The authors declare that they have no known competing financial in- Pikaar, I., 2019. Mainstream ammonium recovery to advance sustainable urban
terests or personal relationships that could have appeared to influence wastewater management. Environ. Sci. Technol. 53 (19), 11066–11079.
Damayanti, A., Ujang, Z., Salim, M.R., 2011. The influenced of PAC, zeolite, and Moringa
the work reported in this paper.
CO

oleifera as biofouling reducer (BFR) on hybrid membrane bioreactor of palm oil mill
effluent (POME). Bioresour. Technol. 102 (6), 4341–4346.
Acknowledgements de Lange, M.F., Verouden, K.J.F.M., Vlugt, T.J.H., Gascon, J., Kapteijn, F., 2015.
Adsorption-driven heat pumps: the potential of metal-organic frameworks. Chem. Rev.
115 (22), 12205–12250.
This work was financially supported by the Australia-China Joint DeCoste, J.B., Denny, M.S., Peterson, G.W., Mahle, J.J., Cohen, S.M., 2016. Enhanced
Research Centre – Healthy Soils for Sustainable Food Production and aging properties of HKUST-1 in hydrophobic mixed-matrix membranes for ammonia
adsorption. Chem. Sci. 7 (4), 2711–2716.
Environmental Quality (ACSRF48165) and Cooperative Research Cen-
Ding, Y., Liu, Y., Liu, S., Li, Z., Tan, X., Huang, X., Zeng, G., Zhou, L., Zheng, B., 2016.
tres Projects (CRC-P) ‘Optimising Nitrogen Recovery from Livestock
UN

Biochar to improve soil fertility. A review. Agron. Sustain. Dev.: Official journal of the
Waste for Multiple Production and Environmental Benefits”, Natural Institut National de la Recherche Agronomique (INRA) 36 (2), 1.
Dong, Y., Yuan, H., Zhang, R., Zhu, N., 2019. Removal of ammonia nitrogen from
Science Foundation of China (22002032), Hebei Key R&D Program wastewater: a review. Trans. ASABE (Am. Soc. Agric. Biol. Eng.) 62 (6), 1767–1778.
(20327303D, 19223811D), Natural Science Foundation of Hebei Eturki, S., Ayari, F., Jedidi, N., Ben Dhia, H., 2012. Use of clay mineral to reduce
Province (B2019201064) and Advanced Talents Incubation Program of ammonium from wastewater. Effect of various parameters. Surf. Eng. Appl.
Electrochem. 48 (3), 276–283.
the Hebei University (1081/801260201284). Fan, R.M., Chen, C.L., Lin, J.Y., Tzeng, J.H., Huang, C.P., Dong, C.D., Huang, C.P., 2019.
Adsorption characteristics of ammonium ion onto hydrous biochars in dilute aqueous
References solutions. Bioresour. Technol. 272, 465–472.
Fidel, R.B., Laird, D.A., Spokas, K.A., 2018. Sorption of ammonium and nitrate to biochars
is electrostatic and pH-dependent. Sci. Rep. 8, 10.
Adam, M.R., Othman, M.H.D., Abu Samah, R., Puteh, M.H., Ismail, A.F., Mustafa, A.,
Fu, H., Li, Y., Yu, Z., Shen, J., Li, J., Zhang, M., Ding, T., Xu, L., Lee, S.S., 2020. Ammonium
Rahman, M.A., Jaafar, J., 2019. Current trends and future prospects of ammonia
removal using a calcined natural zeolite modified with sodium nitrate. J. Hazard
removal in wastewater: a comprehensive review on adsorptive membrane
Mater. 393.
development. Separ. Purif. Technol. 213, 114–132.
Ghnimi, S.M., Frini-Srasra, N., 2018. A comparison of single and mixed pillared clays for
Alshameri, A., He, H.P., Zhu, J.X., Xi, Y.F., Zhu, R.L., Ma, L.Y., Tao, Q., 2018. Adsorption
zinc and chromium cations removal. Appl. Clay Sci. 158, 150–157.
of ammonium by different natural clay minerals: characterization, kinetics and
adsorption isotherms. Appl. Clay Sci. 159, 83–93.
B. Han et al. / Journal of Cleaner Production xxx (xxxx) 124611 17

Giannakis, E., Kushta, J., Bruggeman, A., Lelieveld, J., 2019. Costs and benefits of Li, S.M., Barreto, V., Li, R.W., Chen, G., Hsieh, Y.P., 2018. Nitrogen retention of biochar
agricultural ammonia emission abatement options for compliance with european air derived from different feedstocks at variable pyrolysis temperatures. J. Anal. Appl.
quality regulations. Environ. Sci. Eur. 31 (1), 13. Pyrolysis 133, 136–146.
Grant Glover, T., Peterson, G.W., Schindler, B.J., Britt, D., Yaghi, O., 2011. MOF-74 Li, J.Y., Yang, S., Du, Z.H., Wang, R.X., Yuan, L.M., Wang, H.Y., Wu, Z.C., 2019.
building unit has a direct impact on toxic gas adsorption. Chem. Eng. Sci. 66 (2), Quantitative analysis of ammonia adsorption in Ag/AgI-coated hollow waveguide by
163–170. mid-infrared laser absorption spectroscopy. Optic Laser. Eng. 121, 80–86.
Guest, J.S., Skerlos, S.J., Barnard, J.L., Beck, M.B., Daigger, G.T., Hilger, H., Jackson, S.J., Lin, Z., Huang, W., Zhou, J., He, X., Wang, J., Wang, X., Zhou, J., 2020. The variation
Karvazy, K., Kelly, L., Macpherson, L., Mihelcic, J.R., Pramanik, A., Raskin, L., Van on nitrogen removal mechanisms and the succession of ammonia oxidizing archaea

F
Loosdrecht, M.C.M., Yeh, D., Love, N.G., 2009. A new planning and design paradigm and ammonia oxidizing bacteria with temperature in biofilm reactors treating saline
to achieve sustainable resource recovery from wastewater. Environ. Sci. Technol. 43 wastewater. Bioresour. Technol. 314, 123760.
(16), 6126–6130. Liu, C.C.K., Xia, W., Park, J.W., 2007. A wind-driven reverse osmosis system for
Guo, X., Jin, X., 2014. Purification of uf-treated anaerobically digested manure wastewater aquaculture wastewater reuse and nutrient recovery. Desalination 202 (1), 24–30.

OO
by two-pass reverse osmosis. Desalination Water Treat 52 (16–18), 3027–3034. Liu, Y., Yan, C.J., Zhao, J.J., Zhang, Z.H., Wang, H.Q., Zhou, S., Wu, L.M., 2018. Synthesis
Gupta, V.K., Sadegh, H., Yari, M., Ghoshekandi, R.S., Maazinejad, B., Chahardori, M., of zeolite P1 from fly ash under solvent-free conditions for ammonium removal from
2015. Removal of ammonium ions from wastewater a short review in development of water. J. Clean. Prod. 202, 11–22.
efficient methods. Glob. J. Environ. Sci. Manag. 1 (2), 149–158. Lu, J., Liu, N., Li, L.X., Lee, R., 2010. Organic fouling and regeneration of zeolite
Hakimi, M.H., Jegatheesan, V., Navaratna, D., 2020. The potential of adopting struvite membrane in wastewater treatment. Separ. Purif. Technol. 72 (2), 203–207.
precipitation as a strategy for the removal of nutrients from pre-anmbr treated abattoir Luukkonen, T., Sarkkinen, M., Kemppainen, K., Ramo, J., Lassi, U., 2016. Metakaolin
wastewater. J. Environ. Manag. 259, 109783. geopolymer characterization and application for ammonium removal from model
Han, Z., Dong, J., Shen, Z., Mou, R., Zhou, Y., Chen, X., Fu, X., Yang, C., 2019. Nitrogen solutions and landfill leachate. Appl. Clay Sci. 119, 266–276.
removal of anaerobically digested swine wastewater by pilot-scale tidal flow Luukkonen, T., Tolonen, E.T., Runtti, H., Kemppainen, K., Peramaki, P., Ramo, J., Lassi,
constructed wetland based on in-situ biological regeneration of zeolite. Chemosphere U., 2017. Optimization of the metakaolin geopolymer preparation for maximized

PR
217, 364–373. ammonium adsorption capacity. J. Mater. Sci. 52 (16), 9363–9376.
Han, B., Zhang, W., He, J.-Z., Chen, D., 2020. Lignite ammonia adsorption and surface Ma, Zhiling, Jia, Qunpeng, Shi, Yanzhe, Duan, Xiaofei, Han, Bing, 2020. Synergistic effect
chemistry after dewatering. Separ. Purif. Technol. 253. of aeration-assisted photocatalysis: Two parallel mechanism pathways for TXP-10
Hao, J.N., Yan, B., 2016. Simultaneous determination of indoor ammonia pollution and surfactant removal. Journal of Water Process Engineering 38, 101618. doi:https://
its biological metabolite in the human body with a recyclable nanocrystalline doi.org/10.1016/j.jwpe.2020.101618. Submitted for publication.
lanthanide-functionalized mof. Nanoscale 8 (5), 2881–2886. Ma, Z.L., Jia, Q.P., Tao, C., Han, B., 2020. Highlighting unique function of immobilized
Henze, M., VanLoosdrecht, M.C.M., Ekama, G.A., Brdjanovic, D., 2008. Biological superoxide on TiO2 for selective photocatalytic degradation. Separ. Purif. Technol.
Wastewater Treatment: Principles, Modelling and Design. Iwa Publishing, London. 238, 6.
Hermassi, M., Valderrama, C., Gibert, O., Moreno, N., Querol, X., Batis, N.H., Cortina, J.L., Maia, G.D.N., Day, G.B., Gates, R.S., Taraba, J.L., 2012. Ammonia biofiltration and nitrous
2017. Recovery of nutrients (N-P-K) from potassium-rich sludge anaerobic digestion oxide generation during the start-up of gas-phase compost biofilters. Atmos. Environ.

D
side-streams by integration of a hybrid sorption-membrane ultrafiltration process: 46, 659–664.
use of powder reactive sorbents as nutrient carriers. Sci. Total Environ. 599–600, Malekian, R., Abedi-Koupai, J., Eslamian, S.S., Mousavi, S.F., Abbaspour, K.C., Afyuni, M.,
422–430. 2011. Ion-exchange process for ammonium removal and release using natural iranian
Ho, Y.S., McKay, G., 1999. Pseudo-second order model for sorption processes. Process zeolite. Appl. Clay Sci. 51 (3), 323–329.
Biochem. 34 (5), 451–465. Malovanyy, A., Sakalova, H., Yatchyshyn, Y., Plaza, E., Malovanyy, M., 2013.
TE
Huang, J.Y., Kankanamge, N.R., Chow, C., Welsh, D.T., Li, T.L., Teasdale, P.R., 2018. Concentration of ammonium from municipal wastewater using ion exchange process.
Removing ammonium from water and wastewater using cost-effective adsorbents: a Desalination 329, 93–102.
review. J. Environ. Sci. 63, 174–197. Martins, T.H., Souza, T.S.O., Foresti, E., 2017. Ammonium removal from landfill leachate
Huang, J., Kankanamge, N.R., Chow, C., Welsh, D.T., Li, T., Teasdale, P.R., 2018. by clinoptilolite adsorption followed by bioregeneration. J. Environ. Chem. Eng. 5 (1),
Removing ammonium from water and wastewater using cost-effective adsorbents: a 63–68.
review. J. Environ. Sci. 63, 174–197. Megaritis, A.G., Fountoukis, C., Charalampidis, P.E., Pilinis, C., Pandis, S.N., 2013.
Ismadji, S., Tong, D.S., Soetaredjo, F.E., Ayucitra, A., Yu, W.H., Zhou, C.H., 2016. Response of fine particulate matter concentrations to changes of emissions and
EC

Bentonite hydrochar composite for removal of ammonium from Koi fish tank. Appl. temperature in Europe. Atmos. Chem. Phys. 13 (6), 3423–3443.
Clay Sci. 119, 146–154. Miao, J., Shi, Y., Zeng, D., Wu, G., 2020. Enhanced shortcut nitrogen removal and
Jasuja, H., Peterson, G.W., Decoste, J.B., Browe, M.A., Walton, K.S., 2015. Evaluation of metagenomic analysis of functional microbial communities in a double sludge system
MOFs for air purification and air quality control applications: ammonia removal from treating ammonium-rich wastewater. Environ. Technol. 41 (14), 1877–1887.
air. Chem. Eng. Sci. 124, 118–124. Mochizuki, T., Kubota, M., Matsuda, H., Camacho, L.F.D., 2016. Adsorption behaviors of
Jin, Z.H., Wang, B.D., Ma, L., Fu, P.B., Xie, L.L., Jiang, X., Jiang, W.J., 2020. Air ammonia and hydrogen sulfide on activated carbon prepared from petroleum coke by
pre-oxidation induced high yield N-doped porous biochar for improving toluene KOH chemical activation. Fuel Process. Technol. 144, 164–169.
adsorption. Chem. Eng. J. 385, 9. Moondra, N., Jariwala, N.D., Christian, R.A., 2020. Sustainable treatment of domestic
Jing, Q.X., Chai, L.Y., Huang, X.D., Tang, C.J., Guo, H., Wang, W., 2017. Behavior of wastewater through microalgae. Int. J. Phytoremediation 1–7.
RR

ammonium adsorption by clay mineral halloysite. Trans. Nonferrous Metals Soc. Moribe, S., Chen, Z., Alayoglu, S., Syed, Z.H., Islamoglu, T., Farha, O.K., 2019. Ammonia
China 27 (7), 1627–1635. capture within isoreticular metal–organic frameworks with rod secondary building
Kang, S., Chun, J., Park, N., Lee, S.M., Kim, H.J., Son, S.U., 2015. Hydrophobic zeolites units. ACS Materials Letters 1 (4), 476–480.
coated with microporous organic polymers: adsorption behavior of ammonia under Mounfield, W.P., Taborga Claure, M., Agrawal, P.K., Jones, C.W., Walton, K.S., 2016.
humid conditions. Chem. Commun. 51 (59), 11814–11817. Synergistic effect of mixed oxide on the adsorption of ammonia with metal–organic
Karadag, D., Tok, S., Akgul, E., Turan, M., Ozturk, M., Demir, A., 2008. Ammonium frameworks. Ind. Eng. Chem. Res. 55 (22), 6492–6500.
removal from sanitary landfill leachate using natural gordes clinoptilolite. J. Hazard Nazari, M.A., Mohaddes, F., Pramanik, B.K., Othman, M., Muster, T., Bhuiyan, M.A., 2018.
Mater. 153 (1–2), 60–66. Application of victorian brown coal for removal of ammonium and organics from
CO

Karri, R.R., Sahu, J.N., Chimmiri, V., 2018. Critical review of abatement of ammonia from wastewater. Environ. Technol. 39 (8), 1041–1051.
wastewater. J. Mol. Liq. 261, 21–31. Nijem, N., Fursich, K., Bluhrn, H., Leone, S.R., Gilles, M.K., 2015. Ammonia adsorption
Katz, M.J., Leznoff, D.B., 2009. Highly birefringent cyanoaurate coordination polymers: and co-adsorption with water in HKUST-1: spectroscopic evidence for cooperative
the effect of polarizable C-X bonds (X = Cl, Br). J. Am. Chem. Soc. 131 (51), interactions. J. Phys. Chem. C 119 (44), 24781–24788.
18435–18444. NIOSH, 1994. Immediately dangerous to life or health concentrations (IDLH): Ammonia.
Katz, M.J., Howarth, A.J., Moghadam, P.Z., DeCoste, J.B., Snurr, R.Q., Hupp, J.T., Farha, The National Institute for Occupational Safety and Health. https://www.cdc.gov/
O.K., 2016. High volumetric uptake of ammonia using Cu-MOF-74/Cu-CPO-27. Dalton niosh/idlh/7664417.html.
Trans. 45 (10), 4150–4153. Peng, C., Chai, L.-Y., Tang, C.-J., Min, X.-B., Ali, M., Song, Y.-X., Qi, W.-M., 2017.
Khabzina, Y., Farrusseng, D., 2018. Unravelling ammonia adsorption mechanisms of Feasibility and enhancement of copper and ammonia removal from wastewater using
UN

adsorbents in humid conditions. Microporous Mesoporous Mater. 265, 143–148. struvite formation: a comparative research. J. Chem. Technol. Biotechnol. 92 (2),
Khalil, A., Sergeevich, N., Borisova, V., 2018. Removal of ammonium from fish farms 325–333.
by biochar obtained from rice straw: isotherm and kinetic studies for ammonium Putra, R.N., Lee, Y.H., 2020. Entrapment of micro-sized zeolites in porous hydrogels:
adsorption. Adsorpt. Sci. Technol. 36 (5–6), 1294–1309. strategy to overcome drawbacks of zeolite particles and beads for adsorption of
Kinidi, L., Tan, I.A.W., Wahab, N.B.A., Bin Tamrin, K.F., Hipolito, C.N., Salleh, S.F., ammonium ions. Separ. Purif. Technol. 237, 8.
2018. Recent development in ammonia stripping process for industrial wastewater Qajar, A., Peer, M., Andalibi, M.R., Rajagopalan, R., Foley, H.C., 2015. Enhanced ammonia
treatment. Int. J. Chem. Eng. 14. adsorption on functionalized nanoporous carbons. Microporous Mesoporous Mater.
Kotoulas, A., Agathou, D., Akratos, C.S., Tatoulis, T.I., Tekerlekopoulou, A.G., 218, 15–23.
Triantaphyllidou, I.E., Vayenas, D.V., 2019. Zeolite as a potential medium for Quach, N.K.N., Yang, W.D., Chung, Z.J., Tran, H.L., 2017. The influence of the activation
ammonium recovery and second cheese whey treatment. Water 11 (1), 136. temperature on the structural properties of the activated carbon xerogels and their
Lagergren, S.K., 1898. About the theory of so-called adsorption of soluble substances. electrochemical performance. Ann. Mater. Sci. Eng. 9.
Sven. Vetenskapsakad. Handingarl 24, 1–39. Rada-Ariza, A.M., Lopez-Vazquez, C.M., van der Steen, N.P., Lens, P.N.L., 2017.
Lee, J.W., Barin, G., Peterson, G.W., Xu, J., Colwell, K.A., Long, J.R., 2017. A microporous Nitrification by microalgal-bacterial consortia for ammonium removal in flat panel
amic acid polymer for enhanced ammonia capture. ACS Appl. Mater. Interfaces 9 (39), sequencing batch photo-bioreactors. Bioresour. Technol. 245, 81–89.
33504–33510. Rahmani, A.R., Samadi, M.T., Ehsani, H.R., 2009. Investigation of clinoptilolite natural
Li, R.H., Wang, J.J., Zhou, B.Y., Zhang, Z.Q., Liu, S., Lei, S., Xiao, R., 2017. Simultaneous zeolite regeneration by air stripping followed by ion exchange for removal of
capture removal of phosphate, ammonium and organic substances by MgO
impregnated biochar and its potential use in swine wastewater treatment. J. Clean.
Prod. 147, 96–107.
18 B. Han et al. / Journal of Cleaner Production xxx (xxxx) 124611

ammonium from aqueous solutions. Iran. J. Environ. Health. Sci. Eng. 6 (3), 167–172. Vu, M.T., Chao, H.-P., Van Trinh, T., Le, T.T., Lin, C.-C., Tran, H.N., 2018. Removal
Rambabu, N., Rao, B.V.S.K., Surisetty, V.R., Das, U., Dalai, A.K., 2015. Production, of ammonium from groundwater using naoh-treated activated carbon derived from
characterization, and evaluation of activated carbons from de-oiled canola meal for corncob wastes: batch and column experiments. J. Clean. Prod. 180, 560–570.
environmental applications. Ind. Crop. Prod. 65, 572–581. Wang, S.B., Peng, Y.L., 2010. Natural zeolites as effective adsorbents in water and
Ravishankara, A.R., Daniel, J.S., Portmann, R.W., 2009. Nitrous oxide (N2O): the dominant wastewater treatment. Chem. Eng. J. 156 (1), 11–24.
ozone-depleting substance emitted in the 21st century. Science 326 (5949), 123–125. Wang, B., Lehmann, J., Hanley, K., Hestrin, R., Enders, A., 2015. Adsorption and
Ren, H.-T., Liang, Y., Han, X., Liu, Y., Wu, S.-H., Bai, H., Jia, S.-Y., 2020. Photocatalytic desorption of ammonium by maple wood biochar as a function of oxidation and pH.
oxidation of aqueous ammonia by Ag2O/TiO2 (p25): new insights into selectivity and Chemosphere 138, 120–126.

F
contributions of different oxidative species. Appl. Surf. Sci. 504, 144433. Wang, X.G., Lu, S.Y., Gao, C.M., Feng, C., Xu, X.B., Bai, X., Gao, N.N., Yang, J.L., Liu,
Rezaei, E., Azar, R., Nemati, M., Predicala, B., 2017. Gas phase adsorption of ammonia M.Z., Wu, L., 2016. Recovery of ammonium and phosphate from wastewater by wheat
using nano TiO2-activated carbon composites – effect of TiO2 loading and composite straw-based amphoteric adsorbent and reusing as a multifunctional slow-release
characterization. Journal of Environmental Chemical Engineering 5 (6), 5902–5911. compound fertilizer. ACS Sustain. Chem. Eng. 4 (4), 2068–2079.

OO
Rieth, A.J., Tulchinsky, Y., Dinca, M., 2016. High and reversible ammonia uptake in Wang, X.L., Qiao, B., Li, S.M., Li, J.S., 2016. Using natural Chinese zeolite to remove
mesoporous azolate metal organic frameworks with open Mn, Co, and Ni sites. J. Am. ammonium from rainfall runoff following urea fertilization of a paddy rice field.
Chem. Soc. 138 (30), 9401–9404. Environ. Sci. Pollut. Res. 23 (6), 5342–5351.
Rieth, A.J., Wright, A.M., Dinca, M., 2019. Kinetic stability of metal-organic frameworks Wang, J.T., Jiang, W.Y., Zhang, Z.X., Long, D.H., 2017. Mesoporous carbon beads
for corrosive and coordinating gas capture. Nat. Rev. Mater. 4 (11), 708–725. impregnated with transition metal chlorides for regenerative removal of ammonia in
Ruckart, K.N., Zhang, Y.C., Reichert, W.M., Peterson, G.W., Glover, T.G., 2016. Sorption of the atmosphere. Ind. Eng. Chem. Res. 56 (12), 3283–3290.
ammonia in mesoporous-silica ionic liquid composites. Ind. Eng. Chem. Res. 55 (47), Weatherley, L.R., Miladinovic, N.D., 2004. Comparison of the ion exchange uptake of
12191–12204. ammonium ion onto New Zealand clinoptilolite and mordenite. Water Res. 38 (20),
Sahin, D., 2018. Evaluation of natural minerals (zeolite and bentonite) for nitrogen 4305–4312.
compounds adsorption in different water temperatures suitable for aquaculture. Int. Weber, W.J., Morris, J.C., 1963. Kinetics of adsorption on carbon from solution. J. Sanit.

PR
Lett. Nat. Sci. 71. Eng. Div. 89 (2), 31–60.
Sancho, I., Licon, E., Valderrama, C., de Arespacochaga, N., López-Palau, S., Cortina, J.L., Xu, K.N., Lin, F.Y., Dou, X.M., Zheng, M., Tan, W., Wang, C.W., 2018. Recovery of
2017. Recovery of ammonia from domestic wastewater effluents as liquid fertilizers ammonium and phosphate from urine as value-added fertilizer using wood waste
by integration of natural zeolites and hollow fibre membrane contactors. Sci. Total biochar loaded with magnesium oxides. J. Clean. Prod. 187, 205–214.
Environ. 584–585, 244–251. Xu, Q.Y., Li, W.P., Ma, L., Cao, D., Owens, G., Chen, Z.L., 2020. Simultaneous removal of
Seredych, M., Petit, C., Tamashausky, A.V., Bandosz, T.J., 2009. Role of graphite precursor ammonia and phosphate using green synthesized iron oxide nanoparticles dispersed
in the performance of graphite oxides as ammonia adsorbents. Carbon 47 (2), onto zeolite. Sci. Total Environ. 703, 8.
445–456. Yadi, M.G., Benguella, B., Gaouar-Benyelles, N., Tizaoui, K., 2016. Adsorption of ammonia
Seredych, M., Ania, C., Bandosz, T.J., 2016. Moisture insensitive adsorption of ammonia from wastewater using low-cost bentonite/chitosan beads. Desalin. Water Treat. 57
on resorcinol-formaldehyde resins. J. Hazard Mater. 305, 96–104. (45), 21444–21454.

D
Service, R.F., 2018. Liquid sunshine. Science 361 (6398), 120–123. Yan, T., Li, T.X., Wang, R.Z., Jia, R., 2015. Experimental investigation on the ammonia
Shelp, G.S., Seed, L.P., 2007. Electrochemical Treatment of Ammonia in Waste-Water. adsorption and heat transfer characteristics of the packed multi-walled carbon
Google Patents.. nanotubes. Appl. Therm. Eng. 77, 20–29.
Shih, Y.J., Dong, C.D., Huang, Y.H., Huang, C.P., 2020. Loofah-derived activated carbon Yang, H., Zuttel, A., Kim, S., Ko, Y., Kim, W., 2017. Effect of boron doping on graphene
oxide for ammonia adsorption. ChemNanoMat 3 (11), 794–797.
TE
supported on nickel foam (Ac/Ni) electrodes for the electro-sorption of ammonium
ion from aqueous solutions. Chemosphere 242, 125259. Yang, C., Wang, J.F., Chen, Y., Liu, D., Huang, S.M., Lei, W.W., 2018. One-step
Silva, I.F., Teixeira, I.F., Barros, W.P., Pinheiro, C.B., Ardisson, J.D., do Nascimento, G.M., template-free synthesis of 3D functionalized flower-like boron nitride nanosheets for
Pradie, N.A., Teixeira, A.P.C., Stumpf, H.O., 2019. An Fe-III dinuclear metallacycle NH3 and CO2 adsorption. Nanoscale 10 (23), 10979–10985.
complex as a size-selective adsorbent for nitrogenous compounds and a potentially Yang, H.I., Lou, K., Rajapaksha, A.U., Ok, Y.S., Anyia, A.O., Chang, S.X., 2018. Adsorption
effective ammonia storage material. J. Mater. Chem. A 7 (25), 15225–15232. of ammonium in aqueous solutions by pine sawdust and wheat straw biochars.
Soetardji, J.P., Claudia, J.C., Ju, Y.H., Hriljac, J.A., Chen, T.Y., Soetaredjo, F.E., Santoso, Environ. Sci. Pollut. Res. 25 (26), 25638–25647.
S.P., Kurniawan, A., Ismadji, S., 2015. Ammonia removal from water using sodium Yao, F., Fu, W., Ge, X., Wang, L., Wang, J., Zhong, W., 2020. Preparation and
EC

hydroxide modified zeolite mordenite. RSC Adv. 5 (102), 83689–83699. characterization of a copper phosphotungstate/titanium dioxide
Stokstad, E., 2014. Air pollution ammonia pollution from farming may exact hefty health (Cu-H3PW12O40/TiO2) composite and the photocatalytic oxidation of
costs. Science 343 (6168) 238-238. high-concentration ammonia nitrogen. Sci. Total Environ. 727, 138425.
Sumaraj, Xiong, Z., Sarmah, A.K., Padhye, L.P., 2020. Acidic surface functional groups Ye, L., Lo, B.T.W., Qu, J., Wilkinson, I., Hughes, T., Murray, C.A., Tang, C.C., Tsang, S.C.E.,
control chemisorption of ammonium onto carbon materials in aqueous media. Sci. 2016. Probing atomic positions of adsorbed ammonia molecules in zeolite. Chem.
Total Environ. 698, 134193. Commun. 52 (16), 3422–3425.
Sun, Z., Qu, X., Wang, G., Zheng, S., Frost, R.L., 2015. Removal characteristics of Ye, Y.Y., Ngo, H.H., Guo, W.S., Liu, Y.W., Chang, S.W., Nguyen, D.D., Liang, H., Wang,
ammonium nitrogen from wastewater by modified Ca-bentonites. Appl. Clay Sci. 107, J., 2018. A critical review on ammonium recovery from wastewater for sustainable
wastewater management. Bioresour. Technol. 268, 749–758.
RR

46–51.
Takahashi, A., Tanaka, H., Parajuli, D., Nakamura, T., Minami, K., Sugiyama, Y., Hakuta, Yin, S., Chen, K., Srinivasakannan, C., Guo, S., Li, S., Peng, J., Zhang, L., 2018. Enhancing
Y., Ohkoshi, S., Kawamoto, T., 2016. Historical pigment exhibiting ammonia gas recovery of ammonia from rare earth wastewater by air stripping combination of
capture beyond standard adsorbents with adsorption sites of two kinds. J. Am. Chem. microwave heating and high gravity technology. Chem. Eng. 337, 515–521.
Soc. 138 (20), 6376–6379. You, X., Valderrama, C., Querol, X., Cortina, J.L., 2017. Recovery of ammonium by powder
Talalaj, I.A., 2015. Removal of nitrogen compounds from landfill leachate using reverse synthetic zeolites from wastewater effluents: optimization of the regeneration step.
osmosis with leachate stabilization in a buffer tank. Environ. Technol. 36 (9), Water Air Soil Pollut. 228 (10) 396-396.
1091–1097. Yu, Q.Q., Xia, D., Li, H., Ke, L.T., Wang, Y.P., Wang, H.T., Zheng, Y.M., Li, Q.B., 2016.
Tao, C., Jia, Q.P., Han, B., Ma, Z.L., 2020. Tunable selectivity of radical generation over Effectiveness and mechanisms of ammonium adsorption on biochars derived from
CO

TiO2 for photocatalysis. Chem. Eng. Sci. 214, 7. biogas residues. RSC Adv. 6 (91), 88373–88381.
Taşdemir, A., Cengiz, İ., Yildiz, E., Bayhan, Y.K., 2020. Investigation of ammonia stripping Zare, K., Sadegh, H., Shahryari-ghoshekandi, R., Asif, M., Tyagi, I., Agarwal, S., Gupta,
with a hydrodynamic cavitation reactor. Ultrason. Sonochem. 60, 104741. V.K., 2016. Equilibrium and kinetic study of ammonium ion adsorption by Fe3O4
Thompson, K.A., Shimabuku, K.K., Kearns, J.P., Knappe, D.R.U., Summers, R.S., Cook, nanoparticles from aqueous solutions. J. Mol. Liq. 213, 345–350.
S.M., 2016. Environmental comparison of biochar and activated carbon for tertiary Zhang, H.Y., Li, A.M., Zhang, W., Shuang, C.D., 2016. Combination of na-modified zeolite
wastewater treatment. Environ. Sci. Technol. 50 (20), 11253–11262. and anion exchange resin for advanced treatment of a high ammonia-nitrogen content
Tosun, I., 2012. Ammonium removal from aqueous solutions by clinoptilolite: municipal effluent. J. Colloid Interface Sci. 468, 128–135.
determination of isotherm and thermodynamic parameters and comparison of kinetics Zhang, H., Li, A., Zhang, W., Shuang, C., 2016. Combination of na-modified zeolite and
by the double exponential model and conventional kinetic models. Int. J. Environ. anion exchange resin for advanced treatment of a high ammonia–nitrogen content
UN

Res. Publ. Health 9 (3), 970–984. municipal effluent. J. Colloid Interface Sci. 468, 128–135.
Travlou, N.A., Bandosz, T.J., 2017. N-doped polymeric resin-derived porous carbons as Zhang, W., Zhou, Z., An, Y., Du, S., Ruan, D., Zhao, C., Ren, N., Tian, X., 2017.
efficient ammonia removal and detection media. Carbon 117, 228–239. Optimization for zeolite regeneration and nitrogen removal performance of a
Tu, Y.N., Feng, P., Ren, Y.G., Cao, Z.H., Wang, R., Xu, Z.Q., 2019. Adsorption of ammonia hypochlorite-chloride regenerant. Chemosphere 178, 565–572.
nitrogen on lignite and its influence on coal water slurry preparation. Fuel 238, 34–43. Zhang, P., Zeng, X.Z., Wen, X.H., Yang, C.K., Ouyang, S.D., Li, P., Gu, Z., Wu, D.S.,
Van Humbeck, J.F., McDonald, T.M., Jing, X.F., Wiers, B.M., Zhu, G.S., Long, J.R., 2014. Frost, R.L., 2019. Insights into efficient removal and mechanism for ammonium from
Ammonia capture in porous organic polymers densely functionalized with bronsted aqueous solution on tricalcium aluminate. Chem. Eng. J. 366, 11–20.
acid groups. J. Am. Chem. Soc. 136 (6), 2432–2440. Zhao, X., Tu, C., Zhou, Z., Zhang, W., Ma, X., Yang, J., 2019. Recovery of ammonia
Verstraete, W., de Caveye, P.V., Diamantis, V., 2009. Maximum use of resources present in nitrogen and magnesium as struvite from wastewaters in coal-fired power plant. Asia
domestic "used water. Bioresour. Technol. 100 (23), 5537–5545. Pac. J. Chem. Eng. 14 (5), e2355.
Vikrant, K., Kumar, V., Kim, K.H., Kukkar, D., 2017. Metal-organic frameworks (MOFs): Zheng, W.H., Hu, J.T., Rappeport, S., Zheng, Z., Wang, Z.X., Han, Z.S., Langer, J.,
potential and challenges for capture and abatement of ammonia. J. Mater. Chem. A 5 Economy, J., 2016. Activated carbon fiber composites for gas phase ammonia
(44), 22877–22896. adsorption. Microporous Mesoporous Mater. 234, 146–154.
Vikrant, K., Kim, K.-H., Dong, F., Giannakoudakis, D.A., 2020. Photocatalytic platforms for Zheng, X.B., Yang, Z.M., Xu, X.H., Dai, M., Guo, R.B., 2018. Characterization and ammonia
removal of ammonia from gaseous and aqueous matrixes: status and challenges. ACS adsorption of biochar prepared from distillers’ grains anaerobic digestion residue with
Catal. 10 (15), 8683–8716. different pyrolysis temperatures. J. Chem. Technol. Biotechnol. 93 (1), 198–206.
Zhou, C., Fan, X., Ning, Z., Li, P., Liu, C., Yang, P., Liu, Y., Shi, Z., Li, Y., 2015. Reducing
riverbed infiltration using mixtures of sodium bentonite and clay. Environmental
Earth Sciences 74 (4), 3089–3098.
B. Han et al. / Journal of Cleaner Production xxx (xxxx) 124611 19

Zhu, Y., Kolar, P., Shah, S.B., Cheng, J.J., Lim, P.K., 2016. Avocado seed-derived activated
carbon for mitigation of aqueous ammonium. Ind. Crop. Prod. 92, 34–41.
Zhu, L., Dong, D., Hua, X., Xu, Y., Guo, Z., Liang, D., 2017. Ammonia nitrogen removal and
recovery from acetylene purification wastewater by air stripping. Water Sci. Technol.
75 (11), 2538–2545.
Zielinski, M., Zielinska, M., Debowski, M., 2016. Ammonium removal on zeolite modified
by ultrasound. Desalin. Water Treat. 57 (19), 8748–8753.

F
OO
PR
D
TE
EC
RR
CO
UN

View publication stats

You might also like