Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Environ Sci Pollut Res (2018) 25:25638–25647

DOI 10.1007/s11356-017-8551-2

ENVIRONMENTAL FUNCTIONS OF BIOCHAR

Adsorption of ammonium in aqueous solutions by pine sawdust


and wheat straw biochars
Hye In Yang 1,2 & Kangyi Lou 1 & Anushka Upamali Rajapaksha 3 & Yong Sik Ok 4 &
Anthony O. Anyia 5 & Scott X. Chang 1

Received: 5 August 2016 / Accepted: 1 February 2017 / Published online: 22 February 2017
# Springer-Verlag Berlin Heidelberg 2017

Abstract Ammonium (NH4+) is a common form of reactive bonding and electrostatic interaction of NH4+ with the surface
nitrogen in wastewater, and its discharge to water bodies can functional groups. Lower pyrolysis temperature resulted in a
lead to eutrophication. This study was conducted to under- higher NH4+ adsorption capacity by the pine sawdust biochar.
stand NH4+ adsorption mechanisms of pine sawdust and At the same pyrolysis temperature (550 °C), the biochar made
wheat straw biochars in aqueous solutions and the factors with pine sawdust as the feedstock had a higher NH4+ adsorp-
affecting NH 4 + removal. Biochars were produced by tion capacity than biochar made from wheat straw. We con-
pyrolysing pine sawdust at 300 °C (PS300) and 550 °C clude that biochars can be efficient absorbents for NH4+ re-
(PS550) and wheat straw at 550 °C (WS550). Pseudo- moval from wastewater, and the removal efficiency can be
second-order and Redlich-Peterson models best fitted the ad- optimised by selecting different feedstocks or the pyrolysis
sorption data. The PS300 showed the highest NH4+ adsorption condition for biochar production.
capacity (5.38 mg g−1), followed by PS550 (3.37 mg g−1) and
WS550 (2.08 mg g−1). Higher H/C and O/C ratios of PS300
Keywords Black carbon . Charcoal . Ammonium
(0.78 and 0.32, respectively) indicated the greater presence of
adsorption . Kinetics . Isotherm . Wastewater treatment
functional groups on the biochar’s surface as compared to
PS550 (0.35 and 0.10, respectively) and WS550 (0.36 and
0.08, respectively), resulting in different NH4+ adsorption
through electrostatic interactions. The dominant mechanism Introduction
for NH4+ adsorption by the biochars was likely chemical
Ammonium (NH4+) is one of the common forms of reactive
nitrogen (N) in water and wastewater (Cui et al. 2016).
Responsible editor: Philippe Garrigues
Pollution of NH4+ mainly originates from the discharge of
NH4+ in municipal sewage, wastewater from fertiliser facto-
* Scott X. Chang
scott.chang@ualberta.ca
ries and agricultural wastewater. High levels of NH4+ in water
bodies accelerate eutrophication that depletes dissolved oxy-
1
Department of Renewable Resources, University of Alberta, 442
gen and substantially contribute to toxicity on aquatic organ-
Earth Sciences Building, Edmonton, AB T6G 2E3, Canada isms (Ismail and Hameed 2014). Ammonium nitrogen (NH4+-
2
Department of Rural and Biosystems Engineering, Chonnam
N) is the primary form of N pollution (reactive N), and remov-
National University, Gwangju 61186, Republic of Korea al of NH4+ in wastewater often involves biological systems
3
Department of Basic Sciences, Faculty of Health Sciences, The Open
using the nitrification-denitrification process. Ammonium is
University of Sri Lanka, Nawala, Nugegoda 10250, Sri Lanka converted to nitrate (NO3−) and the N gas is produced as the
4
Korea Biochar Research Center & School of Natural Resources and
final conversion product (Thornton et al. 2007). However,
Environmental Science, Kangwon National University, biological processes for removing NH4+ can be slow and dif-
Chuncheon 24341, Republic of Korea ficult to deal with when the ammonium concentration in
5
Aquatic and Crop Resource Development Portfolio, National wastewater suddenly increases (Jorgensen and Weatherly
Research Council of Canada, Ottawa, ON K1N 5A2, Canada 2003).
Environ Sci Pollut Res (2018) 25:25638–25647 25639

Adsorption is considered to be an effective water treatment practical implications for wastewater treatment by biochar
technique as it is cost-effective, straightforward and efficient with different properties.
in removing pollutants (Inyang et al. 2016; Rajapaksha et al.
2016). Different adsorbents such as activated carbon and ze-
olites were reported to be effective in adsorbing contaminants Materials and methods
in wastewater, but limitations exist because the cost of activat-
ed carbon is high, and the availability of zeolites is limited Preparation of biochar and aqueous solutions containing
(Ahmad et al. 2012; Wahab et al. 2010; Wang and Peng 2010). NH4+
Biochar has gained global interest in environmental sci-
ences due to its cost-effectiveness as an adsorbent, as well as The pine sawdust biochars were produced at the pyrolysis
the benefits in long-term carbon storage (Lehmann and Joseph temperature of 300 °C (PS300) and 550 °C (PS550) and the
2009; Mohan et al. 2014; Ok et al. 2015). Biochar production wheat straw biochar was pyrolysed at 550 °C (WS550). The
involves thermal decomposition of biomass rich in carbon production of the pine sawdust biochar was described in Lou
such as leaves, manure and agricultural residues under a et al. (2016a), and the production of the wheat straw biochar
low-oxygen condition (Lehmann and Joseph 2009; Yao was provided by Alberta Innovates Technology Futures.
et al. 2012). In this manner, waste materials can be converted Briefly, PS300 and PS550 were produced in a pyrolyser
into stabilised carbon as biochar, thereby reducing carbon (N11/H Nabertherm, Germany) at 300 and 550 °C, respective-
emission to the atmosphere. Also, it has been reported that ly. The WS550 was produced in a batch-type carboniser
biochar, with its high cation exchange capacity, is effective (Alberta Innovates Technology Futures, Prototype 1.0) at
in adsorbing cations (Cao et al. 2009). Studies have found that 500–550 °C.
biochar effectively removes organic and inorganic contami- The cation exchange capacity (CEC) of biochar was mea-
nants in water through sorption (Ahmad et al. 2014; Tan et al. sured using a modified NH4+-acetate compulsory displace-
2015; Lou et al. 2016b; Rajapaksha et al. 2016). Furthermore, ment method (Rajkovich et al. 2012). Briefly, 1.0 g of each
biochar produced from a range of feedstocks has been shown biochar sample was saturated with 50 mL of 1 N ammonium
to mitigate nutrient loss from agricultural soils by reducing acetate at pH 7 and shaken in a shaker for 16 h. Then, the
nitrogen leaching and increasing NH4+ sorption (Ding et al. sample was filtered and washed with a 50-mL ammonium
2010; Hollister et al. 2013; Zheng et al. 2013). In addition, acetate. The samples were then washed with 30 mL of ethanol
nitrogen-loaded biochar could also be added to the soil as a three times. Then, 50 mL of 2 N potassium chloride (KCl)
fertiliser (Hina et al. 2013; Wu et al. 2013). However, research solution was added to the biochar and the samples were shak-
on removing nutrients such as NH4+ from aqueous environ- en for 16 h to replace the adsorbed NH4+ with K+ ions. The
ments such as wastewater by biochar adsorption is limited. samples were filtered, and a second aliquot of 50 mL of 2 N
Plant-based waste materials such as sawdust, wood chips KCl solution was used to wash the biochar. The extracted
and agricultural wastes are produced from the forest products NH4+ concentration was determined colorimetrically using
and agriculture industries, and these wastes incur disposal the indophenol blue method with the absorbance read at
cost. Utilising waste materials has economic advantages of 636 nm (USEPA 1993).
obtaining low-cost feedstocks for biochar production and re- An NH4+-N stock solution of 100 mg N L−1 was prepared
ducing waste disposal costs. It also offers an environmental with ammonium sulphate ((NH4)2SO4, Certified A.C.S.,
advantage of reusing and minimising wastes. One example of Fisher Scientific). Then, the stock solution was used to pre-
such waste materials which can be utilised as a cost-effective pare NH4+-N solutions of varying concentrations between 10
feedstock for biochar production is pine sawdust. Wheat straw and 100 mg L−1 by dilution. The solution pH was adjusted
is an agricultural by-product that is abundantly available and using HCl (0.1 N) and NaOH (0.1 N) solutions. Adsorption
can be utilised as a biochar feedstock. Studies on adsorption batch experiments were conducted to study the effects of bio-
capacity and adsorption mechanism of NH4+ on sawdust and char application rate, pH, contact time and the presence of
straw biochars in aqueous solutions and the effects from op- competing ions on NH4+ adsorption. All preparations and ex-
eration conditions (e.g., pH, application rate and competing periments were conducted at room temperature (approximate-
ions) are limited (Chun et al. 2004; Hina et al. 2013). ly 20 °C).
The objectives of this study were to (i) investigate the
effects of pyrolysis temperature and feedstock types on Effects of biochar application rate, pH, initial NH4+
NH4+ adsorption from aqueous solutions by biochars, concentration and competing ions on NH4+ adsorption
(ii) evaluate the mechanisms of NH4+ adsorption by bio-
chars and (iii) examine how biochar application rate and The influence of biochar application rate on NH4+ adsorption
environmental conditions including pH and competing was investigated by treating 20 mL of NH4+-N solutions
ions affect the adsorption capacity of biochars to provide (40 mg N L−1) at different biochar application rates (1, 2, 3,
25640 Environ Sci Pollut Res (2018) 25:25638–25647

5 and 7 g L−1); the initial NH4+ concentration of 40 mg N L−1 where qe is the amount adsorbed (mg g−1), qt is the amount
is typical of municipal wastewaters (Bolan et al. 2004). The adsorbed at time t (h) and k1 is the pseudo-first-order rate
pH was adjusted to 7.0 for consistency and the samples were constant (h−1). The values of k1 and qe were obtained from
shaken for 24 h in a mechanical shaker. the slope and intercepts of the plots of log(qe − qt) against t.
To investigate the effect of initial NH4+ concentration, The pseudo-second-order equation and its linear form are
3 g L−1 of biochar was applied in NH4+ solutions of varying given as follows (Ho and McKay 1999; Rajapaksha et al.
concentrations (0, 10, 20, 30, 50, 70 and 100 mg N L−1) at 2011)
adjusted pH of 7.0. The samples were shaken in a mechanical
dq
shaker for 24 h. The effect of contact time on NH4+ removal ¼ k 2 ðqe −qt Þ2 ð3Þ
was studied by applying 3 g L−1 of biochar in 20 mL of dt
40 mg N L−1 solutions. The samples were then adjusted to t 1 t
¼ þ ð4Þ
pH 7 and shaken in a mechanical shaker for 0, 1, 2, 4, 6, 12, 24 qt k 2 q2e qe
and 48 h to determine at which time equilibrium was reached.
The effect of pH on NH4+ adsorption was studied over a pH where qe is the amount adsorbed (mg g−1), qt is the amount
range of 3.0–10.0 in 1.0 pH unit increments, with the biochar adsorbed at time t (h) and k2 is the pseudo-second-order rate
application rate of 3 g L−1 and NH4+ solution of 40 mg N L−1. constant (g mg−1 h−1). Plots of qt against t were presented, and
t

The samples were shaken for 24 h, and the pH values of the k2 and qe values were obtained from the slope and intercepts.
samples were readjusted to the initial pH values.
Competing ions effect on biochar NH4+ adsorption was Isotherm study
investigated by preparing solutions containing sodium (Na+),
potassium (K+), calcium (Ca2+) and magnesium (Mg2+) using Adsorption isotherms are used to describe how adsorbates
NaCl, KCl, CaCl2 and MgCl2, respectively. These ions were interact with adsorbents at equilibrium (Mittal et al. 2010).
selected because they are commonly present in various types Langmuir, Freundlich, Temkin and Redlich-Peterson isotherm
of wastewater including municipal and industrial wastewater models were used to analyse the experimental data. The
and compete with NH4+ for adsorption sites (Jorgensen and Langmuir isotherm is expressed as
Weatherly 2003). A 10-mL 80 mg L−1 NH4+-N solution and a
ðkqm C e Þ
10-mL solution containing one of the competing cations at qe ¼ ð5Þ
80 mg L−1 concentration were mixed and reacted with ð1 þ k  C e Þ
3 g L−1 of biochar. The initial concentrations of NH4+-N and
where qe is the amount of adsorbate adsorbed at equilibrium
each of the competing cations were thus 40 mg L−1. The
(mg g−1), qm is the maximum adsorption capacity (mg g−1), Ce
samples were adjusted to pH 7 and shaken in a mechanical
is the equilibrium concentration of the adsorbate (mg L−1) and
shaker (200 rpm) for 24 h. A 0.45-μm membrane was used to
k is the Langmuir constant (L mg−1).
filter the samples. The filtrate NH4+-N concentration was de-
A dimensionless separation factor (R) expresses Langmuir
termined colorimetrically using the indophenol blue method
isotherm and can be calculated as (Weber and Chakravorti
with the absorbance read at 636 nm (USEPA 1993). All the
1974)
samples in the above experiments were run in triplicates.
1
R¼ ð6Þ
Kinetic study 1 þ kC 0

Ammonium adsorption by biochar was observed over 48 h to where k is the Langmuir constant and C0 is the initial adsor-
establish the time taken to reach equilibrium and to determine bate concentration. The separation factor indicates the iso-
the adsorbed quantity of NH4+ at equilibrium. To understand therm as unfavourable (R > 1), linear (R = 1), favourable
the mechanisms of NH4+ adsorption by pine sawdust and (0 < R < 1) or irreversible (R = 0).
wheat straw biochars, pseudo-first-order and pseudo-second- The Freundlich model equation is stated as
order kinetic models were used to model the data. The pseudo- qe ¼ KC e n ð7Þ
first-order equation and its linearised form are given as fol-
lows (Ho and McKay 1998; Rajapaksha et al. 2011) where qe is the amount of adsorbate adsorbed at equilibrium
(mg g−1), Ce is the equilibrium concentration of the adsorbate
dq
¼ k 1 ðqe −qt Þ ð1Þ (mg L−1), K is the Freundlich constant indicating adsorption
dt capacity and n is adsorption intensity.
The Temkin model assumes that indirect interactions
k1t among adsorbates lead to a linear decrease in the heat
logðqe −qt Þ ¼ logqe − ð2Þ
2:303 of adsorption with coverage (Kavitha and Namasivayam
Environ Sci Pollut Res (2018) 25:25638–25647 25641

2006; Zhang et al. 2011). The equation is expressed as pyrolysis temperature tend to have lower CEC values because
follows they lose their surface functional groups (Silber et al. 2010).
The carbon (C) content was the highest for WS550 (86.97%),
qe ¼ B1 lnðK t C e Þ ð8Þ
followed by PS550 (78.0%) and PS300 (62.7%). As the py-
RT rolysis temperature increases, C content increases but the re-
B1 ¼ ð9Þ
b moval of oxygen-containing functional groups results in de-
creased hydrogen (H) and oxygen (O) contents (Chun et al.
where R is the gas constant (8.31 J mol−1 K−1), T is the abso-
2004). The H/C ratio for PS300, PS550 and WS550 was 0.78,
lute temperature (K), b is a constant related to the heat of
0.35 and 0.36, respectively. The highest H/C ratio of PS300
adsorption, qe is the amount of adsorbate adsorbed at equilib-
indicated that PS300 had a relatively lower aromatic structure
rium (mg g−1), Kt is the equilibrium constant (L mg−1), which
and may have more adsorption sites due to more organic C
corresponds to the maximum binding energy, and Ce is the
being retained (Cui et al. 2016; Lou et al. 2016b). The O/C
equilibrium concentration of the adsorbate (mg L−1).
ratio was also the highest for PS300 (0.32) as compared to
The Redlich-Peterson model is expressed as
PS550 (0.1) and WS550 (0.08), indicating that PS300 had
AC e greater oxygen-containing functional groups on its surface
qe ¼ ð10Þ
1 þ BC ge and, therefore, remained hydrophilic and polar (Wang et al.
2015a; Cui et al. 2016; Lou et al. 2016b). The surface area was
where qe is the amount of adsorbate adsorbed at equilibrium the largest for PS550 (189.2 m2 g−1), followed by WS550
(m g−1), A (L g−1) and B (L mg−1) are Redlich-Peterson iso- (55.24 m2 g−1) and PS300 (<1 m2 g−1). PS300 had the lowest
therm constants, Ce is the equilibrium concentration of the surface area, but with its higher CEC and H/C and O/C ratios,
adsorbate (mg L−1) and g (0 < g < 1) is the Redlich-Peterson it was more favourable for NH4+ adsorption than the other two
constant. When g is equal to 1, Eq. 10 converts to the biochars.
Langmuir form (Han et al. 2009). When g is equal to 0, it is
simplified to Henry’s law equation, and when the value of
BC ge is much larger than 1, the equation becomes Freundlich Effect of application rate, pH and competing ions on NH4+
equation (Yao et al. 2010). adsorption

As the application rate of biochars increased from 1 to 7 g L−1,


the NH4+-N removal efficiency increased from 10.2 to 19.5,
Results and discussion 6.9 to 28.6 and 1.0 to 7.7% for PS300, PS550 and WS550,
respectively (Fig. 1). Such increases in the NH4+-N removal
Biochar characteristics efficiency may be caused by an increase in the total adsorptive
surface resulting in an increased availability of adsorptive sites
The pH of PS300 was low (4.92), and the pH values of PS550 (Zhou et al. 2015). The total amount of NH4+-N removed or
and WS550 were higher (8.16 and 10.13, respectively) the removal efficiency can be expected to continuously in-
(Table 1). At low pyrolysis temperature, the biochar’s pH crease as the application rate increases until the adsorption
was low because of the production of organic acids and phe- sites are saturated (Ihsanullah et al. 2015). As the biochar
nolic substances, whereas at a high pyrolysis temperature application rate increased, the amount of NH4+ adsorbed per
(>300 °C), pH of the biochar increased because of alkali salts unit mass of biochar decreased from 4.07 to 1.11 and 2.74 to
that separated from the feedstock material (Chen et al. 2011). 1.63 mg N g−1 for PS300 and PS550, respectively (Fig. 1).
The biochars’ CEC values were 6.05 (PS550), 29.04 (PS300) However, NH4+ adsorption of WS550 increased from 0.39 to
and 30.46 cmol kg−1 (WS550). Biochar produced at high 0.73 mg N g−1 as the application rate increased from 1 to

Table 1 Physical and chemical


properties of pine sawdust Biochar pH CEC %C %H %O %N H/C O/C Surface area
biochar produced at 300 °C (cmol kg−1) (m2 g−1)
(PS300) and 550 °C (PS550) and
wheat straw biochar produced at PS300a 4.92 29.04 62.7 4.1 27.1 0.34 0.78 0.32 <1
550 °C (WS550) PS550a 8.16 6.05 78.0 2.3 10.2 0.57 0.35 0.10 189.2
WS550b 10.13 30.46 87.0 2.6 8.9 1.32 0.36 0.08 55.2
a
Data including pH, %C, %H, %O, %N, H/C, O/C and surface area obtained from Lou et al. (2016a)
b
Data including pH, %C, %H, %O, %N, H/C, O/C and surface area obtained from Dugdug et al. (personal
communication)
25642 Environ Sci Pollut Res (2018) 25:25638–25647

3 g L−1 and decreased to 0.44 mg N g−1 as the application rate adsorption capacity compared to the other two biochars. The
increased further to 7 g L−1 (Fig. 1). The decrease in NH4+ adsorbate to adsorption site ratio for WS550 may not have
adsorption per unit mass of biochar may be attributed to a decreased as much as PS300 or PS550. However, as more
lower NH4+ ion concentration to biochar adsorption sites ratio biochar was added, the number of adsorption site increased,
resulting from the increased amount of biochar added to the eventually resulting in less NH4+ ions diffusing onto the bio-
solution (Zhang et al. 2011). As a result, less NH4+ was dif- char surface, thereby decreasing NH4+ adsorption per unit
fused onto the surface and more adsorption sites on the bio- mass of biochar. Possible aggregation of biochar particles un-
char remained unsaturated on a per unit mass basis even der higher biochar application rates may have reduced the
though the total amount of adsorbate removed increased. For adsorption sites and the total surface area, decreasing the ad-
WS550, the slight increase in NH4+ adsorption with the initial sorption capacity (Limousin et al. 2007; Zhang et al. 2011).
increase in biochar application rate may be attributed to its low Kizito et al. (2015) also observed the same effect using wood
surface functional groups and thus low adsorption sites and and rice husk biochars.
The pH significantly affected the NH4+-N removal by bio-
5 30 char, as increasing pH increased the NH4+ adsorption capacity
4.5 qe (mg g-1) (Fig. 2). The adsorption capacity was the lowest at pH 3 for
4
% adsorbed
25
PS300, PS550 and WS550 (2.41, 0.01 and 0.51 mg g−1, re-
3.5
20 spectively) and the highest at pH 9 for PS300 and PS550 (6.26

% Removal
and 3.12 mg g−1, respectively). Ammonium adsorption did
qe (mg g-1)

3
2.5 15 not reach the maximum for WS550 in the pH range of 3–10.
2 Below pH 7, higher concentrations of H+ ions led to compe-
10
1.5 tition between NH4+ and H+ for adsorption sites on biochar
1
5 surface, thereby reducing the adsorption of NH4+. As pH in-
0.5 creased above 7, the H+ concentration decreased and more
0 0 NH4+ ions were adsorbed. When pH >9.5, NH4+ ions can
0 2 4 6 8
react with OH− ions to form NH3, causing the NH4+ concen-
4.5 35 tration in the aqueous solution to decrease thereby decreasing
4 qe (mg g-1)
30
the amount of NH4+ adsorption (Ismail and Hameed 2014).
3.5 % adsorbed Therefore, a slight alkaline solution will be the most suitable
3
25 for NH4+ adsorption.
The presence of competing ions including Na+, K+, Ca2+
% Removal
qe (mg g-1)

2.5 20
and Mg2+ decreased NH4+ adsorption capacity of PS300
2 15 (Fig. 3), with the NH4+ adsorption higher in the control solu-
1.5
10
tion (3.53 mg g−1) as opposed to those containing one of the
1 cations (3.04 to 3.26 mg g−1). Ammonium adsorption capacity
5 for PS300 was higher in solutions containing monovalent cat-
0.5
0 0 ions (Na+ and K+) than that containing divalent cations (Ca2+
0 2 4 6 8 and Mg2+), illustrating that the divalent cations were stronger
4.5 30
7
4 qe (mg g-1) 25 6
3.5
3 % adsorbed 20 5
% Removal
qe (mg g-1)

qe (mg g-1)

2.5 4
15
2 PS300
3 PS550
1.5 10
2 WS550
1
5
0.5 1
0 0 0
0 2 4 6 8 0 2 4 6 8 10 12
Application rate (g L-1) pH
Fig. 1 Effect of biochar application rate on NH4+ adsorption by pine Fig. 2 Sorption-edge experiment on the effect of pH on NH4+ adsorption
sawdust biochar produced at 300 °C (PS300) and 550 °C (PS550) and by pine sawdust biochar produced at 300 °C (PS300) and 550 °C (PS550)
wheat straw biochar produced at 550 °C (WS550) (pH = 7; contact and wheat straw biochar produced at 550 °C (WS550) (biochar
time = 24 h) application rate = 3 g L−1; contact time = 24 h)
Environ Sci Pollut Res (2018) 25:25638–25647 25643

4 5
4.5
3.5 4
3 3.5
PS300

qe (mg g-1)
3
2.5 PS550
qe (mg g-1)

2.5
2
2 PS300 WS550
1.5
1.5 PS550 1
WS550 0.5
1
0
0 10 20 30 40 50 60
0.5
Time (h)
0 Fig. 4 Effect of contact time on NH4+ adsorption by pine sawdust
Na+ K+ Ca2+ Mg2+ NH4+ biochar produced at 300 °C (PS300) and 550 °C (PS550) and wheat
Competing ions straw biochar produced at 550 °C (WS550) (pH = 7; NH 4 +
Fig. 3 The effect of competing ions, Na+, K+, Ca2+ and Mg2+, on NH4+ concentration = 40 mg L−1; biochar application rate = 3 g L−1)
adsorption by pine sawdust biochar produced at 300 °C (PS300) and
550 °C (PS550) and wheat straw biochar produced at 550 °C (WS550) (Liu et al. 2013a; Liu et al. 2013b; Gao et al. 2015). The
(pH = 7; biochar application rate = 3 g L−1; contact time = 24 h)
adsorption sites eventually became saturated, and equilibrium
was reached (Karadag et al. 2006). The correlation coeffi-
competitors for the adsorption sites and the divalent cations cients obtained from the pseudo-first-order model were low
occupied more adsorption sites due to its divalent charge. for PS300 and PS550 (0.66 and 0.52, respectively) and higher
Ammonium adsorption for PS550 and WS550 mostly non- for WS550 (0.95). However, the calculated values for NH4+
significantly decreased in the solution with competing ions. removal capacities (qe1, cal) for all three biochars were not
One of the mechanisms for NH4+ adsorption by biochar is consistent with the experimental (qe, exp) values (Table 2).
electrostatic attraction between the cations and the oppositely The pseudo-first-order model is applicable at the beginning
charged functional groups (Ahmad et al. 2014). Therefore, all of adsorption (20–30 min), and it provides a better fit at high
coexisting cations had a negative effect on NH4+ adsorption of initial adsorbate concentration (Azizan 2004; Ismail and
the biochar as they competed for adsorption sites. Other stud- Hameed 2014). Therefore, the pseudo-first-order model did
ies also have shown that the presence of cations in aqueous not represent the experimental data very well.
solutions reduced NH4+ adsorption capacity of biochar and The pseudo-second-order model resulted in a better fit for
other adsorbents such as zeolites (Zhang et al. 2011; Kizito the experimental data. The graph plotted between t/qt and time
et al. 2015). For instance, Kizito et al. (2015) reported that (Fig. 5) showed good linearity, and high R2 values were ob-
NH4+ adsorption rates by wood and rice husk biochar were tained (Table 2). The calculated adsorption capacities (qe, cal)
lower in digestate slurry as compared to those in a pure NH4+- of the three biochars all closely matched their corresponding
N solution, due to the competition of NH4+ with the cations in experimental adsorption capacities (qe, exp) (Table 2). The
the slurry. pseudo-second-order model is applicable over the entire peri-
od of the adsorption experiment (Ismail and Hameed 2014).
Adsorption kinetics and effect of contact time Therefore, the pseudo-second-order model best fitted the ex-
perimental data, suggesting that chemisorption was the rate-
The NH4+ adsorption process was time-dependent for all test- limiting step that involved ion exchange and valency forces
ed biochars. A rapid removal of NH4+ occurred during the through electron sharing between the adsorbent and the adsor-
initial 1 h. The adsorption rate then decreased gradually until bate (Ho 2006).
it almost reached equilibrium around 12 h and became stable
after 24 h (Fig. 4). A high number of unoccupied adsorption Adsorption isotherm
sites and a high-solute concentration gradient at the beginning
of adsorption process promoted the rapid initial uptake (Zhang The NH 4 + adsorption data were fitted to Langmuir,
et al. 2011). Further, the initial interaction of NH4+ ions with Freundlich, Temkin and Redlich-Peterson isotherm models.
negatively charged surface anion groups such as carboxylate Parameters and the coefficients of determination (R2) of each
(COO−) and hydroxyl (OH−) groups also contributed to the isotherm model are given in Table 3. The Redlich-Peterson
fast adsorption (Ismail and Hameed 2014; Wahab et al. 2010). model resulted in higher R2 values for the three biochars (0.97,
The slower adsorption rate that followed after the first 1 h 0.98 and 0.89 for BC300, BC550 and WS550, respectively)
indicated an ionic balance between the aqueous solution and compared to the other three models. The Langmuir and
the adsorbent (Kizito et al. 2015). Other studies have also Temkin model fittings of the adsorption data were not as close
found rapid NH4+ adsorption during the first 1 to 2 h which as that of the Redlich-Peterson model (Table 3). The Temkin
slowed and became stable after 10 to 24 h or even after 4 h model is suitable for modelling gas phase equilibrium and
25644 Environ Sci Pollut Res (2018) 25:25638–25647

Table 2 Kinetic parameters for


NH4+ adsorption by pine sawdust Experimental Pseudo-first-order model Pseudo-second-order model
biochar produced at 300 °C
(PS300) and 550 °C (PS550) and qexp (mg g−1) qe1, cal (mg g−1) k1 (h−1) R2 qe2, cal (mg g−1) k2 (g mg−1 h−1) R2
wheat straw biochar produced at
550 °C (WS550) (pH = 7; biochar PS300 4.66 1.04 0.12 0.66 4.67 0.68 1.00
application rate = 3 g L−1) PS550 1.27 0.43 0.15 0.52 1.29 1.62 1.00
WS550 0.82 0.56 0.14 0.96 0.84 0.72 1.00

may not appropriately represent liquid-phase adsorption (Foo The maximum NH4+ adsorption capacity given by the
and Hameed 2010). The Freundlich model exhibited the same Langmuir model was 5.38, 3.37 and 2.08 mg g−1 for PS300,
R2 values as the Redlich-Peterson model for BC300 and PS550 and WS550, respectively. For comparison, NH4+ ad-
BC550, but a lower R2 value for WS550 (0.75). The isotherm sorption capacities of adsorbents such as clinoptilolites, saw-
model fitting for WS550 was relatively poor because there dust, various biosorbents and biochars made from various
was little change in the amount of NH 4+ adsorbed by feedstocks reported in the literature are given in Table 4, indi-
WS550 at lower NH 4 + initial concentrations (10– cating that the maximum NH4+ adsorption capacities of those
30 mg L−1). Ammonium adsorption by the other two biochars absorbents were similar. However, a few studies reported sig-
increased with initial concentrations to fit the isotherm curve. nificantly higher NH4+ adsorption capacities of biochars com-
The Redlich-Peterson model is an improvement over pared to the NH4+ adsorption capacities found in this study
Langmuir and Freundlich models and has both features of and a few similar studies using biochars. For instance, Kizito
the two models (Behnamfard and Salarirad 2009; Han et al. et al. (2016) reported NH4+ adsorption capacity of hardwood
2009). The BC ge values at different equilibrium NH4+ concen- and corncob biochars to be over 114.7 and 104.8 mg g−1,
trations for the three biochars ranged from 7.48 to 96.03, respectively, in a flow system study using biochar beds. Gao
which were well over 1. Therefore, it approached Freundlich et al. (2015) also reported that NH4+ adsorption capacities of
model, indicating a multi-layer adsorption on a heterogeneous peanut shell, corncob and cotton stalk biochars to be over
surface (Foo and Hameed 2010; Yao et al. 2010). 200 mg g−1. The initial NH4+ concentration used in those
The k values given by the Langmuir model for PS300, studies were 500 mg L−1 which was much higher than that
PS550 and WS550 were 0.026, 0.1 and 0.025 L mg−1, respec- used in this study. The adsorption values obtained in this
tively. When the Langmuir isotherm was expressed by the study, especially that of PS300, were comparable to or higher
dimensionless separation factor R for each biochar, R was
between 0 and 1, suggesting that the adsorption was Table 3 Isotherm parameters for ammonium adsorption by pine
sawdust biochar produced at 300 °C (PS300) and 550 °C (PS550) and
favourable. In addition, the adsorption is deemed favourable
wheat straw biochar produced at 550 °C (WS550) (pH = 7; biochar
if the n value in the Freundlich model is between 0.1 and 1 application rate = 3 g L−1; contact time = 24 h)
(Zhang et al. 2011). The n values obtained were 0.51, 0.29 and
0.47 for PS300, PS550 and WS550, respectively. Thus, both PS300 PS550 WS550
the Langmuir and Freundlich isotherm models confirmed that Langmuir
NH4+ adsorption by the three biochars were effective. qm (mg g−1) 5.38 3.37 2.08
k (L mg−1) 0.026 0.1 0.025
70 R2 0.94 0.83 0.61
60
y = 1.192x + 1.984 Freundlich
R² = 0.997
K (mg g−1) 0.4 0.87 0.18
50
y = 0.777x + 0.373 n 0.51 0.09 0.47
t/qt (h g mg-1)

R² = 0.999
40 R2 0.97 0.98 0.75
PS300
30 Temkin
PS550
y = 0.214x + 0.067
B1 1.06 0.64 0.36
20 WS550
R² = 0.999 Kt (L mg−1) 0.33 1.56 0.47
10
R2 0.92 0.95 0.68
0 Redlich-Peterson
0 10 20 30 40 50 60
A (L g−1) 1.32 3.53 0.55
Time (h)
B (L mg−1) 2.82 4.00 4.93
Fig. 5 Pseudo-second-order plot for NH4+ adsorption by pine sawdust
biochar produced at 300 °C (PS300) and 550 °C (PS550) and wheat straw g 0.52 0.71 0.41
biochar produced at 550 °C (WS550) (pH = 7; biochar application R2 0.97 0.98 0.89
rate = 3 g L−1)
Environ Sci Pollut Res (2018) 25:25638–25647 25645

Table 4 Maximum adsorption


capacities (mg g−1) of various Adsorbents Adsorption capacities Source
adsorbents for NH4+ (mg g−1)

Clinoptilolite 5.2–8.1 Karadag et al.(2006)


Zeolite 5.22 Liang and Ni (2009)
Strawberry leaf powder 3.9–7.7 Liu et al. (2010a)
Strawberry stem 4.62 Liu et al. (2010b)
Boston ivy stem 5.01 Liu et al. (2010b)
Sawdust 1.7 Wahab et al. (2010)
Posidonia oceanica fibres 1.9 Jellali et al. (2011)
Biochar made from Salix rosthornii 3.3–7.4 Zeng et al. (2013)
Biochar made from Thalia dealbata 7.5–17.6 Zeng et al. (2013)
Biochar made from Vetiveria zizanioides 1.9–4.4 Zeng et al. (2013)
Biochar made from Phragmites spp. 2.2–5.4 Zeng et al. (2013)
Maple wood biochar 0.73–5.44 Wang et al. (2015b)
Oak wood sawdust biochar 5.3–32.0 Wang et al. (2015a)
Giant reed biochar 1.2–1.5 Hou et al. (2016)
Southern magnolia leaves 6.22 Liu et al. (2010b)
Maple wood biochar 0.46–0.87 Wang et al. (2016)

than biochars made from different feedstocks and other adsor- the main mechanism of adsorption (Jassal et al. 2015). Since
bents reported in the literature, indicating that the biochars NH4+ adsorption capacities of the biochars used in this study
have considerable potential to adsorb NH 4+ (Table 4). did not exceed the CEC values, physical entrapment is likely
Higher adsorption capacity of PS300 compared to PS550 not the dominant mechanism for NH4+ adsorption.
and WS550 is in agreement with a study by Wang et al. Also, PS300 had the highest adsorption capacity but had
(2015a), which reported that biochars produced at 300 °C the lowest surface area (<1 m2 g−1), while PS550 and WS550
yielded a higher NH4+ adsorption capacity than biochars pro- had higher surface areas (189.2 and 55.2 m2 g−1, respectively)
duced at higher pyrolysis temperature (600 °C). In another but had lower NH4+ adsorption capacities. This indicates that
study, Zeng et al. (2013) tested biochars derived from four large surface area did not necessarily result in more adsorption
herbaceous plants at 500, 600 and 700 °C, and half of the of NH4+ on the biochar surfaces. Therefore, physical adsorp-
biochars’ adsorption capacities were below 5 mg g−1. tion did not play a significant role in NH4+ adsorption by the
pine sawdust and wheat straw biochars, similar to the findings
Possible mechanisms for NH4+ adsorption by Kizito et al. (2015) on wood and rice husk biochars. Kizito
et al. (2015) also suggested that chemical interaction between
Ammonium adsorption by biochar may involve several NH4+ and biochar surface with various functional groups in-
adsorbent-adsorbate interactions such as electrostatic attrac- fluenced NH4+ adsorption.
tion, ion exchange, surface complexation and physical adsorp- The chemical properties of the functional groups on bio-
tion (Tan et al. 2015). Ion exchange contributes to the adsorp- chars’ surface influence NH4+ adsorption more than physical
tion, and biochars with high CEC resulted in relatively high properties such as surface area (Spokas et al. 2012). Higher H/
NH4+ adsorption capacity (Cui et al. 2016), with CEC as the C and O/C ratios suggest more presence of functional groups
potential dominant factor influencing the NH4+ adsorption and presents a possibility of chemical bonding of polar com-
capacity of biochars (Zeng et al. 2013, Gai et al. 2014). pounds such as NH4+ (Jassal et al. 2015). High O/C ratios also
However, WS550 and PS300 in this study had similar CECs indicate high biochar surface polarity, which can result in
(30.46 and 29.04 cmol kg−1, respectively) but different ad- higher NH4+ adsorption through electrostatic interaction (Gai
sorption capacities (2.08 and 5.38 mg g−1, respectively). et al. 2014; Takaya et al. 2016). Removal of hydrogen and
Jassal et al. (2015) reported that the NH4+ adsorption capaci- oxygen-containing functional groups at high pyrolysis tem-
ties of poultry litter biochar, spruce-pine-fir chip biochar and perature results in a higher surface area but low H/C and O/
the mix of the two did not correspond to their CEC values. C ratios for biochars (Gai et al. 2014). The highest H/C and O/
They found that NH4+ adsorption exceeded the CEC of the C ratios were found in PS300, which corresponded to its high
biochars and suggested physical entrapment of NH4+ ions as NH4+ adsorption capacity, while PS550 and WS550 produced
25646 Environ Sci Pollut Res (2018) 25:25638–25647

at higher pyrolysis temperatures exhibited lower H/C and O/C shell-derived biochar properties and TCE adsorption in water.
Bioresour Technol 118:536–544
ratios and correspondingly lower NH4+ adsorption capacities.
Ahmad M, Rajapaksha AU, Lim JE, Zhang M, Bolan N, Mohan D,
This illustrates that pyrolysis temperature and feedstock types Vithanage M, Lee SS, Ok YS (2014) Biochar as a sorbent for contam-
affect NH4+ adsorption. The PS300 biochar may have retained inant management in soil and water: a review. Chemosphere 99:19–33
more cellulose chains and lignin functional groups such as Azizan S (2004) Kinetic models of sorption: a theoretical analysis. J
Colloid Interf Sci 276:47–52
alcohols, acids and hydroxides from its wood-derived feed-
Behnamfard A, Salarirad MM (2009) Equilibrium and kinetic studies on
stock (Abdolali et al. 2014). Surface complexes formation free cyanide adsorption from aqueous solution by activated carbon. J
with oxygen-containing groups such as carboxyl and carbonyl Hazard Mater 170:127–133
groups may also have enhanced NH4+ adsorption by the bio- Bolan N, Wong L, Adriano D (2004) Nutrient removal from farm efflu-
ents. Bioresour Technol 94:251–260
chars (Cui et al. 2016). Further studies on the analysis of
Cao X, Ma L, Gao B, Harris W (2009) Dairy manure derived biochar effec-
different feedstocks’ characteristics and how they may affect tively sorbs lead and atrazine. Environ Sci Technol 42:3285–3291
NH4+ adsorption efficiency of biochars would be beneficial. Chen X, Chen G, Chen L, Chen Y, Lehmann J, McBride MB, Hay AG
The dominant mechanism for NH4+ adsorption by the three (2011) Adsorption of copper and zinc by biochars produced from
pyrolysis of hardwood and corn straw in aqueous solution.
biochars studied is likely chemical bonding and electrostatic
Bioresour Technol 102:8877–8884
interaction with functional groups rather than physical adsorp- Chun Y, Sheng G, Chiou CT, Xing B (2004) Compositions and sorptive
tion or ion exchange. properties of crop residue-derived chars. Environ Sci Technol 38:
2629–2655
Cui X, Hao H, Zhang C, He Z, Yang X (2016) Capacity and mechanisms
of ammonium and cadmium sorption on different wetland-plant
Conclusions derived biochars. Sci Total Environ 539:566–575
Ding Y, Liu Y, Wu W, Shi D, Yang M, Zhong Z (2010) Evaluation of
The three biochars (PS300, PS550 and WS550) were effective biochar effects on nitrogen retention and leaching in multi-layered
soil columns. Water Air Soil Poll 213:47–55
for NH4+ removal from aqueous solutions; their adsorption Foo KY, Hameed BH (2010) Insights into the modeling of adsorption
capacities were affected by pH, competing ions and biochar isotherm systems. Chem Eng J 156:2–10
application rate, showing a competitive adsorption capacity to Gai X, Wang H, Liu J, Zhai L, Liu S, Ren T, Liu H (2014) Effects of
existing adsorbents. The pseudo-second-order kinetic model feedstock and pyrolysis temperature on biochar adsorption of am-
monium and nitrate. PLoS One 9(12):e113888
and Redlich-Peterson isotherm model best fitted adsorption Gao F, Xue Y, Deng P, Cheng X, Yang K (2015) Removal of aqueous
data obtained for the biochars used in this study. Chemical ammonium by biochars derived from agricultural residuals at different
bonding and polar interaction of NH4+ and the surface func- pyrolysis temperatures (2015). Chem Speciat Bioavailb 27(2):92–97
tional groups are likely the primary influence on NH4+ adsorp- Han R, Zhang J, Han P, Wang Y, Zhao Z, Tang M (2009) Study of
equilibrium, kinetic and thermodynamic parameters about methy-
tion on the biochars. Pine sawdust biochar produced at the
lene blue adsorption onto natural zeolite. Chem Eng J 145:496–504
lower pyrolysis temperature (300 °C) had the highest H/C Hina K, Hedley M, Camps-Arbestain M, Hanly J (2013) Comparison of
and O/C ratios and NH4+ adsorption capacity. In contrast, pine bark, biochar, and zeolite as sorbents for NH4+-N removal from
wheat straw biochar pyrolysed at the higher temperature water. Clean Soil Air Water 43:86–91
Ho YS (2006) Review of second-order models for adsorption systems. J
(550 °C) had low H/C and the lowest O/C ratio and the lowest
Hazard Mater 136:681–689
NH4+ adsorption capacity. We conclude that biochar is an Ho YS, McKay G (1998) Sorption of dye from aqueous solution by peat.
effective absorbent for removing NH4+ from wastewater, Chem Eng J 70:115–124
and it is possible to design pyrolysis conditions or select dif- Ho YS, McKay G (1999) Pseudo-second order model for sorption pro-
cesses. Process Biochem 34:451–465
ferent feedstock types to optimise the adsorption of NH4+ by
Hollister CC, Bisogni JJ, Lehmann J (2013) Ammonium, nitrate, and
biochars. phosphate adsorption to solute leaching from biochar prepared from
corn stover (Zea mays L.) and oak wood (Quercus spp.). J Environ
Qual 42:137–144
Acknowledgements This study was conducted with financial support
Hou J, Huang L, Yang Z, Zhao Y, Deng C, Chen Y, Li X (2016)
from an NSERC CREATE Grant that supported the Land Reclamation
Adsorption of ammonium on biochar prepared from giant reed.
International Graduate School (LRIGS) and partially supported by an
Environ Sci Pollut Res 23:19107–19115
NSERC Discovery grant.
Ihsanullah A-KFA, Abusharkh B, Kjaled M, Atieh MA, Nasser MS,
Laoui T, Saleh TA, Agarwal S, Tyagi I, Gupta VK (2015)
Adsorptive removal of cadmium(II) ions from liquid phase using
References acid modified carbon-based adsorbents. J Mol Liq 204:255–263
Inyang MI, Gao B, Yao Y, Xue Y, Zimmerman A, Mosa A,
Pullammanappallil P, Ok YS, Cao X (2016) A review of biochar
Abdolali A, Guo WS, Ngo HH, Chen SS, Nguyen NC, Tung KL (2014) as a low-cost adsorbent for aqueous heavy metal removal. Crit Rev
Typical lignocellulosic wastes and by-products for biosorption pro- Environ Sci Technol 46(4):406–433
cess in water and wastewater treatment: a critical review. Bioresour Ismail ZZ, Hameed BB (2014) A new application of giant reed waste
Technol 160:57–66 material for ammonium removal. Int J Environ Stud 71:122–138
Ahmad M, Lee SS, Dou X, Mohan D, Sung J, Yang JE, Ok YS (2012) Jassal RS, Johnson MS, Molodovskaya M, Black TA, Jollymore A,
Effects of pyrolysis temperature on soybean stover- and peanut Sveinson K (2015) Nitrogen enrichment potential of biochar in
Environ Sci Pollut Res (2018) 25:25638–25647 25647

relation to pyrolysis temperature and feedstock quality. J Environ Rajapaksha AU, Chen SS, Tsang DCW, Zhang M, Vithanage M, Mandal
Manag 152:140–144 S, Gao B, Bolan NS, Ok YS (2016) Engineered/designer biochar for
Jellali S, Wahab MA, Anane M, Riahi K, Jedidi N (2011) Biosorption contaminant removal/immobilization from soil and water: potential
characteristics of ammonium from aqueous solutions onto Posidoni and implication of biochar modification. Chemosphere 148:276–291
oceanica (L.) fibers. Desalination 270:40–49 Rajkovich S, Enders A, Hanley K, Hyland C, Zimmerman AR, Lehmann J
Jorgensen TC, Weatherly LR (2003) Ammonia removal from wastewater (2012) Corn growth and nitrogen nutrition after additions of biochars
by ion exchange in presence of organic contaminants. Water Resour with varying properties to a temperate soil. Biol Fertil Soils 48:271–284
37:1723–1728 Silber A, Levkovitch I, Graber ER (2010) pH-dependent mineral release
Karadag D, Koc Y, Turan M, Armagan B (2006) Removal of ammonium and surface properties of cornstraw biochar: agronomic implica-
ion from aqueous solution using natural Turkish clinoptilolite. J tions. Environ Sci Technol 44:9318–9323
Hazard Mater 136:604–609 Spokas KA, Novak JM, Venterea RT (2012) Biochar’s role as an alterna-
Kavitha D, Namasivayam C (2006) Experimental and kinetic studies on tive N-fertilizer: ammonia capture. Plant Soil 350:35–42
methylene blue adsorption by coir pith carbon. Bioresour Technol Takaya CA, Fletcher LA, Singh S, Anyikude KU, Ross AB (2016)
98:14–21 Phosphate and ammonium sorption capacity of biochar and
Kizito S, Wu S, Kirui WK, Lei M, Lu Q, Bah H, Dong R (2015) hydrochar from different wastes. Chemosphere 145:518–527
Evaluation of slow pyrolyzed wood and rice husks biochar for ad- Tan X, Liu Y, Zeng G, Wang X, Hu X, Gu Y, Yang Z (2015) Application
sorption of ammonium nitrogen from piggery manure anaerobic of biochar for the removal of pollutants from aqueous solutions.
digestate slurry. Sci Total Environ 505:102–112 Chemosphere 125:70–85
Kizito S, Wu S, Wandera SM, Guo L, Dong R (2016) Evaluation of Thornton A, Pearce P, Parsons SA (2007) Ammonium removal from
ammonium adsorption in biochar-fixed beds for treatment of anaer- digested sludge liquors using ion exchange. Water Resour 41:433–
obically digested swine slurry: experimental optimization and 439
modeling. Sci Tot Environ 563-564:1095–1104 USEPA (1993) Method 350.1: determination of ammonia nitrogen by
Lehmann J, Joseph S (2009) Biochar for environmental management: an semi-automated colorimetry. Online at http://www.caslab.com/
introduction. In: Lehmann J, Joseph S (eds) Biochar for environmen- EPA-Methods/PDF/EPA-Method-3501.pdf Accessed 17 Oct 2015
tal management: science and technology. Earthscan, London, pp 1–12 Wahab MA, Jellali S, Jedidi N (2010) Ammonium biosorption onto saw-
Liang Z, Ni J (2009) Improving the ammonium ion uptake onto natural dust: FTIR analysis, kinetics and adsorption isotherms modeling.
zeolite by using an integrated modification process. J Hazard Mater Bioresour Technol 101:5070–5075
166:52–60 Wang S, Peng Y (2010) Natural zeolites as effective adsorbents in water
and wastewater treatment. Chem Eng J 156:11–24
Limousin G, Gaudet J-P, Charlet L, Szenknect S, Barthès V, Krimissa M
Wang Z, Guo H, Shen F, Yang G, Zhang Y, Zeng Y, Wang L, Xiao H,
(2007) Sorption isotherms: a review on physical bases, modeling
Deng S (2015a) Biochar produced from oak sawdust by lanthanum
and measurement. Appl Geochem 22:249–275
(La)-involved pyrolysis for adsorption of ammonium (NH4+), nitrate
Liu H, Dong Y, Liu Y, Wang H (2010a) Screening of novel low-cost
(NO3−), and phosphate (PO43−). Chemosphere 119:646–653
adsorbents from agricultural residues to remove ammonium nitro-
Wang B, Lehmann J, Hanley K, Hestrin R, Enders A (2015b) Adsorption
gen from aqueous solution. J Hazard Mater 178(1–3):1132–1136
and desorption of ammonium by maple wood biochar as a function
Liu H, Dong Y, Wang H, Liu Y (2010b) Ammonium adsorption from
of oxidation and pH. Chemosphere 138:120–126
aqueous solutions by strawberry leaf powder: equilibrium, kinetics
Wang B, Lehmann J, Hanley K, Hestrin R, Enders A (2016) Ammonium
and effects of coexisting ions. Desalination 263:70–75
retention by oxidized biochars produced at different pyrolysis tem-
Liu N, Sun Z, Wu Z, Zhan X, Zhang K, Zhao E, Han X (2013a) peratures and residence times. RSC Adv 6:41907–41913
Adsorption characteristics of ammonium nitrogen by biochar from Weber TW, Chakravorti RK (1974) Pore and solid diffusion models for
diverse origins in water. Adv Mater Res 664:305–312 fixed-bed adsorbers. AICHE J 20:228–238
Liu N, Zhou C, Fu S, Ashraf MI, Zhao E, Shi H, Han X, Hong ZB Wu FP, Jia ZK, Wang SG, Chang SX, Startsev A (2013) Contrasting
(2013b) Study on characteristics of ammonium nitrogen adsorption effects of wheat straw and its biochar on greenhouse gas emissions
by biochar prepared in different temperature. Adv Mater Res 724- and enzyme activities in a Chernozemic soil. Biol Fertil Soils 49:
725:452–456 555–565
Lou K, Rajapaksha AU, Ok YS, Chang SX (2016a) Pyrolysis tempera- Yao ZY, Qi JH, Wang LH (2010) Equilibrium, kinetic and thermodynam-
ture and steam activation effects on sorption of phosphate on pine ic studies on the biosorption of Cu(II) onto chestnut shell. J Hazard
sawdust biochars in aqueous solutions. Chem Speciat Bioavailab Mater 174:137–143
28(1–4):42–50 Yao Y, Gao B, Zhang M, Inyang M, Zimmerman AR (2012) Effect of
Lou K, Rajapaksha AU, Ok YS, Chang SX (2016b) Sorption of biochar amendment on sorption and leaching of nitrate, ammonium
copper(II) from synthetic oil sands process-affected water (OSPW) and phosphate in a sandy soil. Chemosphere 89:1467–1471
by pine sawdust biochars: effects of pyrolysis temperature and steam Zeng Z, Zhang S, Li T, Zhao F, He Z, Zhao H, Yang X, Wang H, Zhao J,
activation. J Soils Sediments 16(8):2081–2089 Rafiq MT (2013) Sorption of ammonium and phosphate from aque-
Mittal A, Mittal J, Malviya A, Kaur D, Gupta VK (2010) Adsorption of ous solution by biochar derived from phytoremediation plants. J
hazardous dye crystal violet from wastewater by waste materials. J Zhejiang Univ Sci B 14:1152–1161
Colloid Interf Sci 343:463–473 Zhang M, Zhang H, Xu D, Han L, Niu D, Tian B, Zhang J, Zhang L, Wu
Mohan D, Sarswat A, Ok YS, Pittman CU Jr (2014) Organic and inor- W (2011) Removal of ammonium from aqueous solutions using
ganic contaminants removal from water with biochar, a renewable, zeolite synthesized from fly ash by a fusion method. Desalination
low cost and sustainable adsorbent—a critical review. Bioresour 271:111–121
Technol 160:191–202 Zheng H, Wang Z, Deng X, Herbert S, Xing B (2013) Impacts of adding
Ok YS, Chang SX, Gao B, Chung H (2015) SMART biochar technolo- biochar on nitrogen retention and bioavailability in agricultural soil.
gy—a shifting paradigm towards advanced materials and healthcare Geoderma 206:32–39
research. Environ Technol Innov 4:206–209 Zhou Z, Yuan J, Hu M (2015) Adsorption of ammonium from aqueous
Rajapaksha AU, Vithanage M, Jayarathna L, Kumara CK (2011) Natural solutions on environmentally friendly barbecue bamboo charcoal:
red earth as a low cost material for arsenic removal: kinetics and the characteristics and kinetic and thermodynamics studies. Environ
effect of competing ions. Appl Geochem 26:648–654 Prog Sustain Energy 34:655–662

You might also like