You are on page 1of 34

Journal Pre-proofs

Adsorption of Sulfate Ion from Water by Zirconium Oxide-Modified Biochar


Derived from Pomelo Peel

Hanting Ao, Wei Cao, Yixia Hong, Jun Wu, Lin Wei

PII: S0048-9697(19)35084-3
DOI: https://doi.org/10.1016/j.scitotenv.2019.135092
Reference: STOTEN 135092

To appear in: Science of the Total Environment

Received Date: 4 July 2019


Revised Date: 15 October 2019
Accepted Date: 19 October 2019

Please cite this article as: H. Ao, W. Cao, Y. Hong, J. Wu, L. Wei, Adsorption of Sulfate Ion from Water by Zirconium
Oxide-Modified Biochar Derived from Pomelo Peel, Science of the Total Environment (2019), doi: https://doi.org/
10.1016/j.scitotenv.2019.135092

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition of a cover
page and metadata, and formatting for readability, but it is not yet the definitive version of record. This version will
undergo additional copyediting, typesetting and review before it is published in its final form, but we are providing
this version to give early visibility of the article. Please note that, during the production process, errors may be
discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

© 2019 Elsevier B.V. All rights reserved.


Adsorption of Sulfate Ion from Water by Zirconium Oxide-Modified Biochar Derived from

Pomelo Peel

Hanting Ao, Wei Cao*, Yixia Hong, Jun Wu, Lin Wei

(College of Civil Engineering, Huaqiao University, Xiamen 361021, China)

*Corresponding author: weicao@hqu.edu.cn


Abstract

Zirconium oxide-modified pomelo peel biochar (ZrBC) was synthesized for the adsorption of sulfate

ion from aqueous solution. Zirconyl chloride octahydrate (ZCO) was used to modify pomelo peel biochar

into ZrBC. The optimal dose of ZCO for modification is 0.5 mol/L, at which ZrBC shows the highest

adsorption of sulfate ion. The adsorbents were characterized by the field emission scanning electron

microscopy, X-ray photoelectron spectroscopy and surface area measurement. The results confirm that the

presence of zirconium oxides and hydroxide groups on the ZrBC surface, and ZrBC has a porous structure

and a higher specific surface area in comparison with pomelo peel biochar. ZrBC shows good affinity for

sulfate ion with a maximum sulfate adsorption capacity of 35.21 mg/g, which is much higher than that of

pomelo peel biochar (1.02 mg/g). The adsorption of sulfate on ZrBC is pH dependent, and acidic

conditions favor the adsorption. The adsorption can reach near-equilibrium in approximately 120 min. The

adsorption kinetics and isotherm follow the pseudo second-order equation and Langmuir adsorption model,

respectively. Furthermore, nitrate and fluoride anions exhibit little influence on the adsorption of sulfate by

ZrBC, whereas phosphate inhibits the adsorption under the same concentration conditions. ZrBC has the

potential to be used for removal of sulfate from aqueous solution.

Key words: sulfate ion, zirconium oxide, pomelo peel biochar, XPS characterization, adsorption

performance.

1 Introduction

Sulfate ions is widely distributed in surface water, groundwater and industrial effluents, such as acid

mining drainage and pharmaceutical, printing and dyeing wastewater. The main sources of sulfate in

natural water are chemical weathering and oxidation processes of sulfur-containing minerals (Fernando et
al., 2018). Although sulfate ion often is considered to be nontoxic, its potential harm to living organisms

and environment cannot be ignored. High concentrations of sulfate ion in water can lead to an imbalance in

the natural sulfur cycle in ecosystem, and endanger human health under long-term ingestion (Amaral Filho

et al., 2016; Pol et al., 1998). Therefore, it is necessary to remove sulfate ion from wastewater before

discharge into the surrounding environment.

The established methods for sulfate removal from water include chemical precipitation, ion exchange,

biological treatment, adsorption and membrane filtration (Gupta et al., 2012; Hong et al., 2014; Silva et al.,

2002; Silva et al., 2010). Adsorption separation is widely used because of its rapid and effective removal

of sulfate ion. The adsorbent plays an important role in determining the effectiveness of adsorption

technology, and the desired adsorbent needs to have a low cost, a high adsorption capacity and be

renewable. Granular activated carbon (GAC) and anion exchange resins have been widely studied to

remove sulfate from industrial wastewater (Fernando et al., 2018; Haghsheno et al., 2009). However, GAC

has low adsorption efficiency, while anion exchange resins are too expensive to be utilized in wastewater

treatment. Hence, a cost-effective sulfate adsorption material must be developed.

Zirconium oxide particles, which are nontoxic and stable in water, have been used to separate

sulfate ion from industrial brine water owing to their high affinity for sulfate (Cui et al., 2012). However,

their practical application is limited because zirconium oxides are difficult to immobilize, separate and

recycle (Yu et al., 2018). To overcome these drawbacks, researchers have attempted to load zirconium

oxides on porous materials such as cellulose, activated carbon, and zeolite (Mulinari and da Silva, 2008).

Recently, biochar has attracted great attention as an alternative desirable adsorbent and catalyst

support due to its porous structure, low cost and environmental compatibility (Jung and Ahn, 2016).

Biochar is a carbon-rich material obtained from the thermochemical or hydrothermal conversion of


biomass under oxygen-limited conditions (Jin et al., 2017; Nielsen et al., 2014). It is believed that biochar

has the potential to delay global climate change and to improve soil fertility and water retention (Jin et al.,

2017; Joseph et al., 2015; Spokas et al., 2009; Warnock et al., 2007; Yanai et al., 2007). Several reports

have confirmed that biochar can serve as an adsorbent to immobilize organic pollutants (Inyang et al.,

2015a; Rajapaksha et al., 2014) and heavy metals (Inyang et al., 2015b; Yang and Jiang, 2014) in soil and

water. Biochar is usually prepared from agricultural and forestry wastes, such as wood, crop straw and fruit

husks, and organic wastes produced in industrial applications and urban life. Pomelo (Citrus maximal) peel

is a kind of low-cost biomass waste, and has a sponge-like structure, which is suitable for preparing porous

biochar (Argun et al., 2014; Tasaso, 2014). However, ordinary biochar has shown poor adsorption capacity

toward anionic species due to its negative surface charge (Bian et al., 2014; Kameyama et al., 2016).

Therefore, metal oxides such as magnesia, lanthana, and iron oxides are applied to modify biochar to

improve its adsorption capacity for nitrate and phosphate (Wang et al., 2016; Zhang et al., 2012). However,

few studies have investigated sulfate adsorption by zirconium oxide-modified biochar.

Herein, this study aims to prepare zirconium oxide-modified pomelo peel biochar (ZrBC) and to

evaluate its potential to remove sulfate ion from water. Pomelo peel (PP) was selected to generate biochar

due to its sponge-like structure and high yield, especially in the main production area, such as Fujian,

China. Zirconyl chloride was used to modify the pomelo peel biochar. The generated adsorbents were

characterized by field emission scanning electron microscopy equipped with energy-dispersive X-ray

spectroscopy (FE-SEM/EDS), N2 adsorption-desorption isotherms and X-ray photoelectron spectroscopy

(XPS). Additionally, the adsorption isotherms and kinetics, as well as the effect of environmental factors

(i.e., pH and competitive ions) on the sulfate adsorption capacity of ZrBC, were investigated.

2 Materials and methods


2.1 Materials and chemicals

Fresh PP was collected from Xiamen, Fujian Province, China. It was washed with distilled water

several times and dried at 50°C for 48 h. Chemicals used in this work such as ammonia (NH3H2O),

potassium sulfate (K2SO4), potassium nitrate (KNO3), potassium phosphate monobasic (KH2PO4), sodium

fluoride (NaF), sodium hydroxide (NaOH) and concentrated hydrochloric acid (HCl) were purchased from

Sinopharm Chemical Reagent Co., Ltd., China. Zirconyl chloride octahydrate (ZCO, ZrOCl28H2O) was

obtained from Aladdin Co., Ltd., China. Deionized (DI) water (18.2 MΩ.cm) was used to prepare

experimental solutions.

2.2 Preparation of ZrBC

The dried PP was cut up with a knife to obtain particles with a diameter of 0.2~0.3cm, and heated at

10 °C/min up to 400 °C under N2 flow for 1.0 h in a tube furnace (NBD-O1200-50IT, Nobody Materials

Science and Technology Co.,Ltd, China). After carbonization, the sample was rinsed with deionized water

several times until the pH of filtrate did not change, and then it was dried in an oven at 80 °C overnight.

The resulting material was blank pomelo peel biochar, labeled as PBC. A solution of ZCO (0.5 mol/L) of

100 mL was prepared in a 250 mL Erlenmeyer flask. PBC (1.0 g) was mixed with the ZCO solution in

the flask, which was placed in a shaker for 2.0 h. Ammonia solution (1:3, v/v) was added dropwise into the

mixture under stirring to adjust pH value to approximately 4.5. After reaction, the sample was aged at

room temperature (25 °C ) for 24 h. Finally, the material was collected by filtration, washed with deionized

water until the conductivity of filtrate was less than 2.0 μS/cm, and then dried in a vacuum oven at 50 °C

for 12 h. The obtained sample, ZrBC, was sealed in a container. To study the effect of the initial ZCO

concentration on the modification process, different ZrBCs were prepared by varying the concentration

from 0.25 mol/L to 0.75 mol/L.


2.3 Characterizations

The surface morphology and microstructure of PP, PBC and ZrBC were characterized by

FE-SEM/EDS (ZEISS Sigm, Carl Zeiss, Germany), which is applied to determine the elemental

composition of materials The surface functional groups of the materials were probed by using XPS

(Escalab 250Xi, Thermo Fisher Scientific, USA). The monochromatized X-ray source was an Al K-alpha

radiation (anode target). The surface area of the adsorbents was measured by N2 adsorption-desorption

isotherms at 77 K using a NOVA 2000e surface area analyzer (Quantachrome, USA). The specific surface

area and average pore radius were calculated based on BET (Brunauer-Emmett-Teller) method. The total

pore volume was analyzed via single-point adsorption at a relative pressure of 0.99. The pore distribution

was determined from the adsorption curve by using BJH (Barret–Joyner–Halenda) method.

2.4 Batch adsorption experiments

Batch adsorption experiments were performed to study the sulfate adsorption performance on the

ZrBC material. A stock sulfate solution with a concentration of 1000 mg/L was prepared by dissolving

potassium sulfate (K2SO4) in deionized water. All sulfate solutions used in adsorption experiments were

diluted from this stock solution. A certain weight (0.1 g) of the adsorbent (PBC or ZrBC) was added into a

150 mL conical flask containing 50 mL sulfate solution. The flasks were placed in a shaker at 150 rpm and

298 K. After remaining in contact for 24 h, the solid adsorbent and solution were separated by filtration,

and the filtrate was collected to detect sulfate concentration. The sulfate ion concentrations in aqueous

solution were measured using ion chromatography method (Metrohm, Switzerland). The detection limit of

the ion chromatography method is 0.13 mg/L and the relative standard deviation is 1.25%. All adsorptoin

experimental samples were designed in duplicate.

To study the effect of pH on sulfate adsorption, adsorption experiments were conducted under
different pH conditions. The initial pH of the sulfate solution was changed from 2.0 to 12.0 by using 1.0

mol/L HCl and 1.0 mol/L NaOH. The adsorption isotherms were measured by varying the initial sulfate

concentration from 5 to 300 mg/L at the optimum pH. The equilibrium adsorption amount (qe) was

calculated by equation (1).

(C0  Ce )  V
qe  , (1)
m
where C0 (mg/L) and Ce (mg/L) are the initial and equilibrium concentrations of sulfate ion, respectively;

m (g) is the mass of adsorbent; and V (L) is the volume of sulfate solution. For the adsorption kinetics

study, the contact time intervals were 5, 10, 30, 60, 120, 240, 480, and 720 min. In addition, the adsorption

kinetics were measured at 288K, 298K, 308K and 318K to investigate the effect of temperature on the

adsorption. The effects of coexisting ions (NO3-, PO43-, and F-) on the adsorption of sulfate on ZrBC were

also studied in a binary adsorption system, and the concentrations of both sulfate and coexisting ions were

1.0 mmol/L. The concentrations of these ions were also determined by using ion chromatography method

(Metrohm, Switzerland).

3 Results and discussion

3.1 Effect of initial ZCO concentration on the preparation of ZrBC

As shown in Fig. 1, all the biochars modified with different initial ZCO concentrations exhibit sulfate

removal ability. The sulfate adsorption amount on ZrBC prepared with an initial ZCO concentration of 0.5

mol/L is higher than that of biochars obtained with other concentrations. Because the surface of most

untreated biochars are neutral or alkaline pH, such materials usually have good cation exchange capacity

and poor anion adsorption ability (Kameyama et al., 2016). Zirconium oxide/hydroxide compounds on the

surface of biochar can supply active adsorption sites for anions, such as phosphate, fluoride, and sulfate

(Cui et al., 2012; Su et al., 2013; Yu et al., 2018). A low initial concentration of ZCO resulted in the poor
sulfate adsorption (Fig. 1) which is due to the decrease in loading amount of zirconium oxides at a low

initial concentration (Mulinari and da Silva, 2008; Robson, 2004; Zhao et al., 2011). A high concentration

of initial ZCO also caused a decrease in sulfate adsorption (Fig. 1). Therefore, a moderate concentration

(0.5 mol/L) of ZCO was selected to prepare ZrBC, which was used in subsequent experiments on the

sulfate adsorption performance.

3.2 FE-SEM/EDS characterization

SEM images of PP, PBC and ZrBC are presented in Fig. 2. As shown in Fig. 2a and Fig. 2d, PP has a

honeycomb structure with many large pores inside, and probably consists of cellulose, hemicellulose, and

lignin (Torab-Mostaedi et al., 2013). It can be clearly seen in Fig. 2b that the PBC surface is rougher than

that of PP and many slice structures are apparent, resulting from pore collapse during the pyrolysis process

of PP (Han et al., 2013). There are many agglomerates deposited across the entire matrix of ZrBC, in

contrast to PBC (Fig. 2b-2c, Fig. 2e-2f), and the aggregated particles have diameters in the range of 0.2~5

μm. Furthermore, EDS element mapping analysis (Fig. S1) confirms that the aggregated particles contain

substantial amount zirconium and oxygen, indicating that zirconium oxides were formed on the surface of

PBC.

3.3 BET specific surface area and pore structure analysis

Pore size distribution and pore structure of PP, PBC, and ZrBC were determined by N2

adsorption-desorption isotherms (Fig. 3), and information on the specific surface area, total pore volume

and average pore radius are listed in Tab. 1. As shown in Fig. 3a, the shapes of adsorption-desorption

isotherms are consistent with type III isotherm according to the IUPAC classification (Sing et al., 1985).

The specific surface area and total pore volume of ZrBC are larger than those of the other materials,

whereas average pore radius of ZrBC is smaller than that of PBC. This is possibly because loading
zirconium oxides on the PBC resulted in a very rough surface, which is in agreement with the

FE-SEM/EDS analysis. ZrBC has more mesoporous than the other samples, as showed in Fig.3b, which

also enhances the specific surface area and total pore volume of ZrBC. In addition, the pore structure of PP

probably collapsed during pyrolysis process, which resulted in a decrease in surface area in PBC. The

experimental results also show that the pyrolysis and chemical precipitation process did not contribute to

the formation of micropores on biochar materials (Fig. 3b).

3.4 XPS characterization

The XPS characterization results of PBC, ZrBC and post-adsorption ZrBC are shown in Fig. 4 and

Tab. 2. Based on comparison with PBC, zirconium was successfully loaded on the surface of ZrBC (Fig.

4a), which is in accordance with the analysis results of FE-SEM/EDS. The weak peak found at 169.16 eV

(Fig. 4b) was assigned to S 2p spectrum (Hong et al., 2014), which confirms the adsorption of sulfate ions

on the ZrBC surface . The Zr 3d spectra of ZrBC and post-adsorption ZrBC were both fitted by two peaks,

corresponding to the Zr3d5/2 and Zr3d3/2 electronic orbitals, respectively (Wang et al., 2019). After sulfate

adsorption on ZrBC, the binding energies of the Zr3d peaks decreased by approximately 0.17 eV and 0.14

eV (Fig. 4c), which is attributed to the bonding between sulfate anion and zirconium (He et al., 2016; Yu et

al., 2018). The O1s XPS spectra are consistent with two overlapping O1s peaks of Zr-OH and O-C, and the

percentage of hydroxyl groups decreases from 21.19% to 3.24% after sulfate adsorption (Fig. 4d and Tab.

2), which indicates that hydroxyl groups (-OH) on the ZrBC surface participate in the adsorption of sulfate.

Furthermore, the increase in solution pH observed after adsorption confirms the release of -OH from ZrBC

during the removal process of sulfate. It can be inferred that the adsorption mechanism of sulfate anion on

ZrBC is probably ligand exchange between -OH and sulfate anion based on Zr4+(Yu et al., 2018).The

exchange formula can be described as equation (2).


ZrO(OH)2 + SO42- = ZrOSO4 + 2OH- (2)

3.5 Effect of pH on the adsorption of sulfate on ZrBC

The effect of the solution pH on sulfate adsorption by ZrBC is shown in Fig. 5. When the solution pH

value increased from 2.0 to 12.0, the adsorption capacity decreased gradually from 34.26 mg/g to 1.31

mg/g. The sulfate adsorption capacity was highly dependent on the initial solution pH, and the optimum

pH was approximately 2.0. The variation in sulfate adsorption capacity with pH can be attributed to the

surface charge of the adsorbent. The point of zero charge (pHPZC) of ZrBC was determined to be 6.19, and

detailed information is shown in Fig. S2. When the pH of the sulfate solution is lower than pHPZC, the

ZrBC surface will be protonated and positively charged (Wang et al., 2016). Sulfate anions are easily

captured by ZrBC through electrostatic attraction. Otherwise, the ZrBC surface is deprotonated at high pH

values, resulting in an increase in electrostatic repulsion and a decrease in sulfate uptake by ZrBC.

Moreover, the OH- ions that exist in solution in large numbers at high pH will compete with sulfate ions

for the active adsorption sites on the ZrBC surface, which also leads to a lower sulfate adsorption capacity.

3.6 Adsorption isotherms and modeling analysis

The isotherm behaviors of sulfate ions on PBC and ZrBC were examined, and the results are shown in

Fig. 6. The sulfate adsorption amount of PBC was as low as 1.02 mg/g, while the adsorption amount of

ZrBC reached 35.14 mg/g. The adsorption amount of ZrBC was obviously increased approximately 30-40

times compared with PBC under the same experimental conditions. In addition, five different isotherm

models were used to analyze the isotherms data. The related equations are respectively shown as follows

(Sun et al., 2015; Yao et al., 2011).

Qmax K l Ce
Langmuir model: qe  , (3)
1  K l Ce

Freundlich model: qe  K f (Ce ) ,


n
(4)
Qmax K lf (Ce ) n
Langmuir-Freundlich model: qe  , (5)
1  K lf (Ce ) n

K r Ce
Redlich-Peterson model: qe  , (6)
1  aCe
RT
Temkin model: qe  ln( ACe ) , (7)
b
where qe (mg/g) and Ce (mg/L) are the adsorption capacity and sulfate concentration at equilibrium,

respectively; Qmax (mg/g) represents the maximum adsorption capacity; Kl (L·mg-1), Kf (mg1-n·Ln·g-1), Klf

(Ln·mg-n), and Kr (L·g-1) are the Langmuir, Freundlich, Langmuir-Freundlich and Redlich-Peterson

constants, respectively; n (dimensionless) is the Freundlich linear constant; a (Ln·mg-n) is the

Redlich-Peterson isotherm constant; A (L·mol-1) is the Temkin equilibrium binding constant and b

(J·g·mg-1) is the Temkin constant (Sun et al., 2015; Yao et al., 2011). Except for Langmuir model, which is

a theoretical equation, all the other models use empirical or semiempirical equations describe

heterogeneous adsorption.

The resulting parameters are summarized in Tab. 3. The Langmuir model matched the experimental

data better than the other models from the comparison of correlation coefficient (R2) values, which

indicated that monolayer adsorption probably occurred (Yao et al., 2013). This result is consistent with

reported studies on sulfate adsorption in a rape straw biochar-soil system (Zhao et al., 2017) and on

polypyrrole-modified activated carbons (Hong et al., 2017). The Qmax of ZrBC calculated from the

Langmuir isotherm model is 35.21 mg/g. Moreover, the partition coefficient (PC), the ratio of the

maximum adsorption amount to the equilibrium concentration of sulfate, is also an important parameter to

evaluate the actual performance of an adsorbent in addition to the adsorption capacity (Gogoi et al., 2019;

Na et al., 2019; Vikrant and Kim, 2019). The PC value for sulfate adsorption on ZrBC is 0.014

mg·g-1·μM-1 mg, which is near the values reported in other sulfate adsorption studies. A detailed
comparison between ZrBC and other sulfate adsorbents is shown in Tab. 4. Furthermore, RL was calculated

according to equation (8) to confirm the favorability of sulfate adsorption on ZrBC (Sun et al., 2015).

1
RL  , (8)
1  K l C0

where Kl (L·mg-1) is the Langmuir constant and C0 (mg·L-1) is the initial solution concentration of sulfate.

The RL values calculated with equation (8) are in the range of 0 to 1, suggesting that sulfate adsorption on

ZrBC is favorable (Sun et al., 2015).

3.7 Adsorption kinetics

The adsorption kinetics of sulfate ions on ZrBC at different temperatures were investigated and are

shown in Fig.7. Fig.7a shows that the overall sulfate adsorption process can be divided into two stages: the

rapid-increase stage and the near-equilibrium stage. The equilibrium time for sulfate adsorption was

approximately 120 min. In addition, sulfate adsorption decreased with increasing of experimental

temperature, indicating that sulfate adsorption was exothermic and that a lower temperature was beneficial

to adsorption. This result agrees with previous reports on sulfate adsorption by Ni–Al and Ni–Fe

quartz-albitophire (Sadeghalvad et al., 2016). To further explore the kinetic characteristics, the pseudo

first- and second-order models were applied to fit the adsorption kinetic data (Azizian, 2004; Sadeghalvad

et al., 2016). The kinetic equations are shown as follows.

dqt
Pseudo first-order equation:  k1 ( qe  qt ) , (9)
dt
dq
Pseudo second-order equation: t  k2 ( qe  qt ) ,
2
(10)
dt
where qt (mg·g-1) and qe (mg·g-1) are the amounts of sulfate adsorbed at time t and at equilibrium,

respectively; and k1 (min-1) and k2 (g· mg-1·min-1) are the rate constants for the pseudo first- and

second-order models, respectively. The analysis results from pseudo first- and second-order models are

demonstrated in Fig. 7b and Fig. 7c, respectively. The related parameters are listed in Tab. 5. The kinetics
of sulfate adsorption on the ZrBC are well described by the pseudo second-order model, whose correlation

coefficient (R2) is closer to 1.0, and the calculated equilibrium adsorption amounts are much closer to the

experimental values. A previous reported study also claimed that sulfate adsorption on

polypyrrole-modified activated carbon follows a pseudo second-order models (Hong et al., 2017). As

shown in Tab. 4, the pseudo second-order rate constant (k2) decreases with increasing of temperature from

288 K to 318 K, suggesting that a decrease in temperature may accelerate adsorption.

3.8 Competitive adsorption

The effects of different coexisting anions on sulfate removal by ZrBC are shown in Fig. 8. The

presence of PO43- decreases the adsorption of sulfate under the same concentration conditions. A previous

study (Zhang et al., 2017) showed that zirconium oxide/hydroxide also has affinity toward phosphate,

which can compete with sulfate for the active adsorption sites on ZrBC, resulting in a decrease in sulfate

removal. For NO3- and F-, the decrease in sulfate adsorption is small, which is in accordance with a

reported study (Zhang et al., 2017). It can be concluded that NO3- and F- have little influence on sulfate

removal by ZrBC, whereas PO43- can impact it under the same concentration conditions. In the case of

practical sulfate wastewater, such as acid mine drainage, interference from coexisting anions to sulfate

removal will not be significant because the concentration of sulfate is much higher than that of other

anions.

4 Conclusion

In this study, a novel and low-cost adsorbent, ZrBC, was synthesized by a chemical precipitation

method for sulfate removal from aqueous solution. The adsorbent has a high specific surface area and total

pore volume and contains numerous zirconium oxides and hydroxyl groups. The sulfate adsorption

isotherm agrees well with the Langmuir adsorption model. The maximum sulfate adsorption capacity of
ZrBC was calculated as 35.21 mg/g, which is much higher than that of the original PBC (1.02 mg/g). The

adsorption of sulfate is pH dependent and better adsorption was achieved at acidic conditions. The

adsorption equilibrium time is approximately 120 min and the adsorption process can be well described by

the pseudo second-order equation. Sulfate adsorption is exothermic, and a lower temperature is beneficial

to adsorption. In addition, nitrate and fluoride have little effect on sulfate removal, whereas phosphate can

decrease adsorption under the same concentration conditions.

Acknowledgements

This work is financially supported by the National Natural Science Foundation of China (No. 51408239),

the Natural Science Foundation of Fujian Province, China (No. 2016J01193), and the Fundamental

Research Funds for the Central Universities (No. ZQN-712).

References

Amaral Filho J, Azevedo A, Etchepare R, Rubio J., 2016. Removal of sulfate ions by dissolved air flotation
(DAF) following precipitation and flocculation. International Journal of Mineral Processing. 149, 1-8.
Argun ME, Güclü D, Karatas M.,2014. Adsorption of Reactive Blue 114 dye by using a new adsorbent: Pomelo
peel. Journal of Industrial and Engineering Chemistry. 20, 1079-1084.
Azizian S., 2004. Kinetic models of sorption: a theoretical analysis. J Colloid Interface Sci. 276, 47-52.
Bian R, Joseph S, Cui L, Pan G, Li L, Liu X, Zhang A, Rutlidge H, Wong S, Chia C, Marjo C, Gong B, Munroe
P, Donne S., 2014. A three-year experiment confirms continuous immobilization of cadmium and lead
in contaminated paddy field with biochar amendment. J Hazard Mater. 272, 121-128.
Chen W, Liu H-c., 2014. Adsorption of sulfate in aqueous solutions by organo-nano-clay: Adsorption
equilibrium and kinetic studies. Journal of Central South University. 21, 1974-1981.
Cui H, Li Q, Gao S, Shang JK., 2012. Strong adsorption of arsenic species by amorphous zirconium oxide
nanoparticles. Journal of Industrial and Engineering Chemistry. 18, 1418-1427.
Fernando WAM, Ilankoon IMSK, Syed TH, Yellishetty M., 2018. Challenges and opportunities in the removal
of sulphate ions in contaminated mine water: A review. Minerals Engineering. 117, 74-90.
Gogoi H, Leiviska T, Ramo J, Tanskanen J., 2019. Production of aminated peat from branched
polyethylenimine and glycidyltrimethylammonium chloride for sulphate removal from mining water.
Environ Res, 175, 323-334.
Gupta VK, Ali I, Saleh TA, Nayak A, Agarwal S., 2012. Chemical treatment technologies for waste-water
recycling—an overview. RSC Advances. 2, 6380-6388.
Haghsheno R, Mohebbi A, Hashemipour H, Sarrafi A.,2009. Study of kinetic and fixed bed operation of
removal of sulfate anions from an industrial wastewater by an anion exchange resin. J Hazard Mater.
166, 961-966.
Han Y, Boateng AA, Qi PX, Lima IM, Chang J.,2013. Heavy metal and phenol adsorptive properties of
biochars from pyrolyzed switchgrass and woody biomass in correlation with surface properties. J
Environ Manage. 118, 196-204.
He JG, Li Y, Xue XX, Ru HQ, Huang XW, Yang H., 2016. Separation of fluorine/cerium from fluorine-bearing
rare earth sulfate solution by selective adsorption using hydrous zirconium oxide. Rsc Advances.
6,43814-43822.
Hong S, Cannon FS, Hou P, Byrne T, Nieto-Delgado C., 2014. Sulfate removal from acid mine drainage using
polypyrrole-grafted granular activated carbon. Carbon. 73, 51-60.
Hong S, Cannon FS, Hou P, Byrne T, Nieto-Delgado C., 2017. Adsorptive removal of sulfate from acid mine
drainage by polypyrrole modified activated carbons: Effects of polypyrrole deposition protocols and
activated carbon source. Chemosphere. 184, 429-437.
Inyang M, Gao B, Zimmerman A, Zhou Y, Cao X., 2015a. Sorption and cosorption of lead and sulfapyridine on
carbon nanotube-modified biochars. Environ Sci Pollut Res Int. 22, 1868-1876.
Inyang MI, Gao B, Yao Y, Xue Y, Zimmerman A, Mosa A, Pullammanappallil P, Ok YS, Cao X., 2015b. A
review of biochar as a low-cost adsorbent for aqueous heavy metal removal. Critical Reviews in
Environmental Science and Technology. 46, 406-433.
Jin J, Sun K, Wang Z, Han L, Du P, Wang X, Xing B., 2017. Effects of chemical oxidation on phenanthrene
sorption by grass- and manure-derived biochars. Sci Total Environ. 598, 789-796.
Joseph S, Anawar HM, Storer P, Blackwell P, Chia C, Lin Y, Munroe P, Donne S, Horvat J, Wang J, Solaiman
ZM., 2015. Effects of Enriched Biochars Containing Magnetic Iron Nanoparticles on Mycorrhizal
Colonisation, Plant Growth, Nutrient Uptake and Soil Quality Improvement. Pedosphere. 25, 749-760.
Jung KW, Ahn KH., 2016. Fabrication of porosity-enhanced MgO/biochar for removal of phosphate from
aqueous solution: Application of a novel combined electrochemical modification method. Bioresour
Technol. 200, 1029-1032.
Kameyama K, Miyamoto T, Iwata Y, Shiono T., 2016. Influences of feedstock and pyrolysis temperature on the
nitrate adsorption of biochar. Soil Science and Plant Nutrition. 62, 180-184.
Mulinari DR, da Silva MLCP., 2008. Adsorption of sulphate ions by modification of sugarcane bagasse
cellulose. Carbohydrate Polymers. 74, 617-620.
Na C-J, Yoo M-J, Tsang DCW, Kim HW, Kim K-H., 2019. High-performance materials for effective sorptive
removal of formaldehyde in air. Journal of Hazardous Materials. 366, 452-465.
Nielsen S, Minchin T, Kimber S, van Zwieten L, Gilbert J, Munroe P, Joseph S, Thomas T., 2014. Comparative
analysis of the microbial communities in agricultural soil amended with enhanced biochars or
traditional fertilisers. Agriculture, Ecosystems & Environment. 191, 73-82.
Pol LWH, Lens PNL, Stams AJM, Lettinga G., 1998. Anaerobic treatment of sulphate-rich wastewaters.
Biodegradation. 9, 213-224.
Rajapaksha AU, Vithanage M, Zhang M, Ahmad M, Mohan D, Chang SX, Ok, YS., 2014. Pyrolysis condition
affected sulfamethazine sorption by tea waste biochars. Bioresour Technol. 166, 303-308.
Robson JD., 2014. A new model for prediction of dispersoid precipitation in aluminium alloys containing
zirconium and scandium. Acta Materialia. 52, 1409-1421.
Runtti H, Luukkonen T, Niskanen M, Tuomikoski S, Kangas T, Tynjälä P, Tolonen E, Sarkkinen M,
Kemppainen K, Rämö J, Lassi, U., 2016 Sulphate removal over barium-modified blast-furnace-slag
geopolymer. Journal of Hazardous Materials. 317, 373-384.
Sadeghalvad B, Azadmehr A, Hezarkhani A., 2016. Enhancing adsorptive removal of sulfate by metal layered
double hydroxide functionalized Quartz-Albitophire iron ore waste: preparation, characterization and
properties. RSC Advances. 6, 67630-67642.
Silva AJ, Varesche MB, Foresti E, Zaiat M., 2002. Sulphate removal from industrial wastewater using a
packed-bed anaerobic reactor. Process Biochemistry. 37, 927-935.
Silva R, Cadorin L, Rubio J., 2010. Sulphate ions removal from an aqueous solution: I. Co-precipitation with
hydrolysed aluminum-bearing salts. Minerals Engineering. 23, 1220-1226.
Sing KSW., 1985. Reporting physisorption data for gas/solid systems with special reference to the
determination of surface area and porosity. Pure and Applied Chemistry. 57, 603–619.
Spokas KA, Koskinen WC, Baker JM, Reicosky DC., 2009. Impacts of woodchip biochar additions on
greenhouse gas production and sorption/degradation of two herbicides in a Minnesota soil.
Chemosphere. 77, 574-581.
Su Y, Cui H, Li Q, Gao S, Shang JK., 2013. Strong adsorption of phosphate by amorphous zirconium oxide
nanoparticles. Water Res. 47, 5018-5026.
Sun P, Hui C, Azim Khan R, Du J, Zhang Q, Zhao YH., 2015. Efficient removal of crystal violet using
Fe3O4-coated biochar: the role of the Fe3O4 nanoparticles and modeling study their adsorption
behavior. Sci Rep. 5, 12638.
Tasaso P., 2014. Adsorption of Copper Using Pomelo Peel and Depectinated Pomelo Peel. Journal of Clean
Energy Technologies. 154-157.
Torab-Mostaedi M, Asadollahzadeh M, Hemmati A, Khosravi A., 2013. Equilibrium, kinetic, and
thermodynamic studies for biosorption of cadmium and nickel on grapefruit peel. Journal of the
Taiwan Institute of Chemical Engineers. 44, 295-302.
Vikrant K, Kim K-H., 2019. Nanomaterials for the adsorptive treatment of Hg(II) ions from water. Chemical
Engineering Journal. 358, 264-282.
Wang X, Pan S, Zhang M, Qi J, Sun X, Gu C, Wang L, Li J., 2019. Modified hydrous zirconium oxide/PAN
nanofibers for efficient defluoridation from groundwater. Sci Total Environ. 685, 401-409.
Wang Z, Shen D, Shen F, Li T., 2016. Phosphate adsorption on lanthanum loaded biochar. Chemosphere. 150,
1-7.
Warnock DD, Lehmann J, Kuyper TW, Rillig MC., 2007. Mycorrhizal responses to biochar in soil – concepts
and mechanisms. Plant and Soil. 300, 9-20.
Yanai Y, Toyota K, Okazaki M., 2007. Effects of charcoal addition on N2O emissions from soil resulting from
rewetting air-dried soil in short-term laboratory experiments. Soil Science and Plant Nutrition. 53,
181-188.
Yang GX, Jiang H, 2014. Amino modification of biochar for enhanced adsorption of copper ions from synthetic
wastewater. Water Res. 48, 396-405.
Yao Y, Gao B, Chen J, Yang L., 2013. Engineered biochar reclaiming phosphate from aqueous solutions:
mechanisms and potential application as a slow-release fertilizer. Environ Sci Technol. 47, 8700-8708.
Yao Y, Gao B, Inyang M, Zimmerman AR, Cao X, Pullammanappallil P, Yang L., 2011. Removal of phosphate
from aqueous solution by biochar derived from anaerobically digested sugar beet tailings. J Hazard
Mater. 190, 501-507.
Yu Z, Xu C, Yuan K, Gan X, Feng C, Wang X, Zhu L, Zhang G, Xu D., 2018. Characterization and adsorption
mechanism of ZrO2 mesoporous fibers for health-hazardous fluoride removal. J Hazard Mater. 346,
82-92.
Zhang C, Li Y, Wang F, Yu Z, Wei J, Yang Z, Ma C, Li Z, Xu Z, Zeng G., 2017. Performance of magnetic
zirconium-iron oxide nanoparticle in the removal of phosphate from aqueous solution. Applied Surface
Science. 396, 1783-1792.
Zhang M, Gao B, Yao Y, Xue Y, Inyang M., 2012. Synthesis of porous MgO-biochar nanocomposites for
removal of phosphate and nitrate from aqueous solutions. Chemical Engineering Journal. 210, 26-32.
Zhao B, Nan X, Xu H, Zhang T, Ma F., 2017. Sulfate sorption on rape (Brassica campestris L.) straw biochar,
loess soil and a biochar-soil mixture. J Environ Manage. 201, 309-314.
Zhao B, Tian C, Zhang Y, Tang T, Wang F., 2011. Size control of monodisperse nonporous silica particles by
seed particle growth. Particuology. 9, 314-317.
Table and Figure List

Tables

Tab. 1 BET-N2 specific surface area (SA), total pore volume (TPV), and average pore radius (APR) of PP,

PBC, and ZrBC.

Tab. 2 XPS full-spectrum scanning data for PBC, ZrBC, and post-sorption ZrBC.

Tab. 3 Adsorption isotherm model parameters and correlation coefficients (R2) for sulfate adsorption by

ZrBC.

Tab. 4 Comparison of ZrBC with other sulfate adsorbents

Tab. 5 Adsorption kinetic model parameters and correlation coefficients (R2) for sulfate adsorption by

ZrBC.

Figures

Fig. 1 Effect of initial ZrOCl2 concentration on ZrBC for the adsorption of sulfate ion.

Fig. 2 FE-SEM/EDS images of PP (a, d), PBC (b, e), and ZrBC (c, f).

Fig. 3. N2 adsorption-desorption isotherms (a) and the pore size distributions (b) of PP, PBC, and ZrBC.

Fig. 4 XPS spectra of PBC, ZrBC, and post-sorption ZrBC: (a) full-spectrum scanning; (b) S2p orbital

spectrum for post-sorption ZrBC ; (c) Zr3d orbital and (d) O1s orbital spectra for ZrBC and post-sorption

ZrBC.

Fig. 5 Effect of pH on the sulfate adsorption capacity of ZrBC.

Fig. 6 Adsorption isotherm of sulfate by PBC and ZrBC (a) and isotherm models of sulfate adsorption on

ZrBC (b).

Fig. 7 Adsorption kinetics of sulfate adsorption on ZrBC at different temperature: (a) raw data, (b) linear
fitting of the pseudo first-order kinetic model, and (c) linear fitting of the pseudo second-order kinetic

model.

Fig. 8 Effect of coexisting anions on sulfate removal by ZrBC. (The initial concentrations of sulfate ion

and the added coexisting anion are both 1.0 mmol/L).


Tab. 1 BET-N2 specific surface area (SA), total pore volume (TPV), and average pore radius (APR) of PP,

PBC, and ZrBC

Adsorbent SA(m2/g) TPV(cm3/g) APR(nm)

PP 10.3551 0.007772 3.00221

PBC 6.7238 0.010585 6.29709

ZrBC 21.7626 0.034591 4.8632


Tab. 2 XPS full-spectrum scanning data for PBC, ZrBC, and post-sorption ZrBC

Atomic (%)
biochar
C O Zr Cl N S

PBC 82.44 13.65 - 0.44 1.86 -

ZrBC 62.10 22.90 4.57 3.93 2.62 0.37

post-sorption ZrBC 69.14 22.08 2.64 1.27 2.37 0.78


Tab. 3 Adsorption isotherm model parameters and correlation coefficients (R2) for sulfate adsorption by

ZrBC

Parameter 1 Parameter 2 Parameter 3 R2

Langmuir Kl=0.135 Qmax=35.208 0.923

Freundlich Kf=8.494 n=0.277 0.789

Langmuir- Freundlich Klf=0.077 Qmax=32.534 n=-0.557 0.922

Redlich-Peterson Kr=4.692 a=0.129 n=1.006 0.904

Temkin b=391.591 A=1.640 0.892


Tab. 4 Comparison of ZrBC with other sulfate adsorbents

Initial Equilibrium Adsorption Partition


pH
absorbents concentration concentration capacity coefficient Reference
value
(mg·L-1) (mg·L-1) (mg·g-1) (mg·g-1·μM-1)

Barium modified blast-furnace slag geopolymer 7-8 - 750 90.0 0.012 (Runtti et al., 2016)

Organo-nanoclay 7.0 - 300 20.0 0.0064 (Chen and Liu, 2014)

PG-Peat 2.4 1835 1050 189.5 0.017 (Gogoi et al., 2019)

Polypyrrole-grafted granular activated carbon - 250 140 44.7 0.031 (Hong et al., 2014)

ZrBC 2 300 240 35.2 0.014 Present study


Tab. 5 Adsorption kinetic model parameters and correlation coefficients (R2) for sulfate adsorption by

ZrBC

qe,exp pseudo first-order equation pseudo second-order equation


T(K)
(mg·g-1) qe,Cal(mg·g-1) k1(1·min-1) R2 qe,Cal(mg·g-1) k2(g· mg-1·min-1) R2

288K 29.967 12.210 0.0177 0.949 30.026 0.0050 0.999

298K 28.016 8.488 0.0046 0.897 28.111 0.0026 0.998

308K 27.226 10.798 0.0049 0.789 27.393 0.0021 0.998

318K 18.766 8.767 0.0040 0.516 17.052 0.0024 0.939


Fig. 1 Effect of initial ZrOCl2 concentration on ZrBC for the adsorption of sulfate ion
a b c

PP (500×) PBC (500×) ZrBC (500×)

d e f

PP (5000×) PBC (5000×) ZrBC (5000×)

Fig. 2 FE-SEM/EDS images of PP (a, d), PBC (b, e), and ZrBC (c, f)
a
b

Fig. 3 N2 adsorption-desorption isotherms (a) and the pore size distributions (b) of the PP, PBC, and ZrBC
a b

c d

Fig. 4 XPS spectra of PBC, ZrBC, and post-sorption ZrBC: (a) full-spectrum scanning; (b) S2p orbital

spectrum for post-sorption ZrBC ; (c) Zr3d orbital and (d) O1s orbital spectra for ZrBC and post-sorption

ZrBC
Fig. 5 Effect of pH on the sulfate adsorption capacity of ZrBC
Fig. 6 Adsorption isotherm of sulfate by PBC and ZrBC (a) and isotherm models of sulfate adsorption on

ZrBC (b)
a b

Fig. 7 Adsorption kinetics of sulfate adsorption on ZrBC at different temperature: (a) raw data, (b) linear

fitting of the pseudo first-order kinetic model, and (c) linear fitting of the pseudo second-order kinetic

model
Fig. 8 Effect of coexisting anions on sulfate removal by ZrBC. (The initial concentrations of sulfate ion

and the added coexisting anion are both 1.0 mmol/L)


Highlights

1) Zr oxide-modification enhanced sulphate adsorption on PBC significantly.

2) XPS confirmed presence of Zr oxides and hydroxyl groups on ZrBC.

3) Acidic conditions favor sulphate removal by ZrBC.

4) The adsorption follows Langmuir model and pseudo second-order rate equation.

You might also like