You are on page 1of 17

Bioresource Technology 246 (2017) 176–192

Contents lists available at ScienceDirect

Bioresource Technology
journal homepage: www.elsevier.com/locate/biortech

Review

Challenges and recent advances in biochar as low-cost biosorbent: From


batch assays to continuous-flow systems
Emilio Rosales, Jessica Meijide, Marta Pazos, María Angeles Sanromán ⇑
Department of Chemical Engineering, University of Vigo, Campus Universitario As Lagoas – Marcosende, E-36310 Vigo, Spain

h i g h l i g h t s g r a p h i c a l a b s t r a c t

 It is proved that biochar can be


Alkali
considered as low-cost adsorbent. Nanoparcles Acid
 Advances on biochar’ synthesis to
Steam Coang
enhance their adsorptive capacity
Biochar Organic Heavy
have been examined. Magnec
modificaon metals
compounds
 Recent findings on organic and
inorganic pollutants by biochars are
summarized.
 Identification of knowledge gaps
Nitrogen and
before scale-up in continuous-flow phosphorus
system.

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

Article history: Over the past few years, the increasing amount of pollutants and their diversity demand to develop ver-
Received 31 May 2017 satile low-cost adsorption systems. The use of biomass feedstock such as agricultural residues, wood
Received in revised form 14 June 2017 chips, manure or municipal solid wastes as source to produce low-cost biosorbent, and the new advances
Accepted 15 June 2017
in their synthesis have encouraged remarkable efforts towards the development of biochar ‘‘on demand”
Available online 19 June 2017
in which their characteristics can be improved. This new trend opens the potential of biochar application
in the removal of pollutants from wastewater, however, its use in environmental management requires
Keywords:
the development of full-scale biosorption in engineered systems. Thus, this paper provides a brief review
Adsorption
Biochar
of recent progress in the research and practical application of biochar with a special emphasis on its
Fixed-bed potential to reduce the pollutants present in wastewater or to render them harmless. Furthermore,
Water pollution research gaps and uncertainties detected in their scale-up in continuous-flow systems are highlighted.
Regeneration Ó 2017 Elsevier Ltd. All rights reserved.
Modelling

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 177
2. Inorganic ions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 178
2.1. Nitrogenated ions: ammonium and nitrate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 180
2.2. Phosphorous ions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 180
3. Heavy metals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 180
3.1. Mercury . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 181
3.2. Arsenic . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 181
3.3. Lead . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 181

⇑ Corresponding author.
E-mail address: sanroman@uvigo.es (M.A. Sanromán).

http://dx.doi.org/10.1016/j.biortech.2017.06.084
0960-8524/Ó 2017 Elsevier Ltd. All rights reserved.
E. Rosales et al. / Bioresource Technology 246 (2017) 176–192 177

3.4. Zinc . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 182


3.5. Copper . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 182
3.6. Cadmium . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 182
3.7. Chromium . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 183
4. Organic pollutants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 183
4.1. Adsorption mechanisms for organic pollutant removal . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 183
4.2. Influence of biochar aging and solution chemistry on organic pollutant adsorption . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 185
4.3. Biochar enhacements for organic pollutant removal . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 186
4.3.1. Coating biochar. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 186
4.3.2. Physical and chemical pretreatment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 186
5. Continuous flow system for adsorption: Fixed-bed columns . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 186
5.1. Operational management . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 186
5.1.1. Particle size . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 186
5.1.2. Porosity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 187
5.2. Regeneration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 188
5.3. Modelling fixed bed adsorption . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 189
5.3.1. Influence of bed height . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 189
5.3.2. Effect of flow rate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 189
5.3.3. Effect of inlet concentration. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 189
5.3.4. Modelling of the column data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 189
6. Conclusions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 190
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 190
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 190

1. Introduction cultures would use biochar obtained by burning agricultural


wastes in order to increase soil productivity (Aller, 2016). This
Nowadays, in the management of aqueous waste streams is term defines a low-density carbon-rich solid with a stable porous
essential to comply the stringent environmental regulations of structure that is produced from a variety of biomass feedstock by
the different countries (EPA, 2017). One option is the development different thermal decomposition methods including slow and fast
of alternative systems for the treatment and disposal of wastewa- pyrolysis, gasification, carbonization, torrefaction and hydrother-
ter, such as evaporation to reduce the volume of effluents gener- mal treatment (Jung et al., 2015a). As is described in the next sec-
ated or pre-treatment prior to disposal or referral to the tions, numerous studies have demonstrated that biochar is a low-
management company including simple treatments such as pre- cost adsorbent for the removal of different kind of pollutants. The
cipitation or adsorption. Useful technical innovations are sought production cost of biochar is in relation to biomass feedstock and
to save energy and reduce the wastewater treatment cost (Cobas procedure used in its production, being estimated a cost from a
et al., 2015). Among these numerous techniques for removing inor- typical biochar around 0.076 $/kg which is approximately 3–6%
ganic and organic pollutants (lead, arsenic, cadmium, phosphates, the price of other commercial carbon-based adsorbents (Yoder
trichloroethylene, furfural, diclofenac, carbamazepine, etc.), et al., 2011).
adsorption is considered to be one of the most promising By revision of a basedata as ScopusÒ, a search for articles,
techniques, especially for effluents containing moderate or low reviews, books, data studies and reports including the keyword
concentrations of these pollutants, and when their rapid removal ‘‘biochar” reveals that the number of studies published per year
or immobilization are required (Ahmed et al., 2017). has grown from 1 in 2000 to over 1332 in 2016 (Fig. 1). The num-
Many materials can be used as adsorbents, and the selection of ber of relevant publications increased over the 10-year period.
one or the other will depend primarily on its application. In recent Only 33 papers in relation to adsorption were published between
years, the efforts of the scientific community have focused on the 2005 and 2010, but the number of published papers had increased
search for new adsorbents having low price, good mechanical to 989 from 2011 until 2016. In addition, it has been detected a
and chemical resistance, abundance, easy handling, easy regenera- higher increase of studies focused on the metal adsorption which
tion, high adsorption capacity or high specific surface area (Crini, represent around 62% of the papers published in relation to pollu-
2006). It is not easy that a material meets all these characteristic tants adsorption.
simultaneously, but accomplish the largest number of these prop- Initially, the studies have been focused on different biochars’
erties is required (Abdolali et al., 2014). Among them, cost is an capacity for uptake of nutrients such as ammonium, nitrate, and
important parameter and depends on the degree of processing phosphate for its application as a soil amendment with the intent
required and local availability of the material based to manufac- of improving soil or promoting other environmental services
ture the adsorbent. (Lehmann et al., 2006). However, in recent years, an emerging sub-
Normally, so-called ‘‘low cost adsorbents” are characterised for set of biochar research pertains to its application as an environ-
requiring scarce processing, are abundant in nature, or are by- mentally sustainable and cost-effective adsorbent for greenhouse
products or waste materials from another industry (Rosales et al., gases, heavy metals, and organic compounds (Li et al., 2017a). In
2015). The economic crisis of the 2000s led researchers to turn these studies, the pollutant uptake under appropriately related
their interest in this kind of adsorbent materials with lower cost. experimental conditions by different biochars were determined.
In addition, recently an environmentally friendly waste manage- However, our efforts in this review are also focused specifically
ment of agricultural and animal wastes, which can contaminate on biochar adsorption of pollutants from aqueous solutions in
surface water as well as aquifers, focuses its attention on the pro- engineered systems. Until date, most of the adsorption studies
duction of biochar with several applications pointing up in adsorp- for the removal of pollutants are based on batch kinetic and batch
tion process (Mohan et al., 2014). Although the origin of the term equilibrium studies. However, the data obtained during batch
‘‘biochar” is unknown, it is believed that ancient South American adsorption are not enough and in the practical operation of full-
178 E. Rosales et al. / Bioresource Technology 246 (2017) 176–192

1400

1200 Adsorption
All publications
Articles published per year
1000

800

600

400

200

0
2008 2009 2010 2011 2012 2013 2014 2015 2016

Fig. 1. Graph depicting the increase in the number of publications using the search term ‘biochar’ (Source: ScopusÒ).

scale biosorption processes, continuous-flow systems are ulate plant and algal growth and promote eutrophication of aqua-
requested to understand the real application potential. tic ecosystems. Among the different available processes,
adsorption represents an appropriate technology, and several
2. Inorganic ions authors have proposed the use of biochar as low-cost adsorbent
to prevent eutrophication and water quality deterioration (Wang
The widespread application of agrochemicals on agriculture has et al., 2015a; Vu et al., 2017). The negative surface charge of
increased the release of nitrogenated and phosphorous compounds unmodified biochar provides non-attraction/weak interactions
to aquatic ecosystems (Wang et al., 2015a). According to EPA with anionic ions and for this reason, several modifications on bio-
(2017), nitrate concentration in freshwater is generally ranging char surface and structure are required to increase the maximum
from 0.1 to 4 mg/L NO3-N whereas unpolluted waters usually con- adsorption capacity (Micháleková-Richveisová et al., 2017). The
tain nitrate levels below 1 mg/L NO3-N. Nevertheless, effluents presence of acidic functional groups such as phenolic and
from sewage treatment plant may have concentrations above carboxylic groups on surface probably promotes ammonium
10 mg/L NO3-N, which may cause health problems. Phosphate con- adsorption whereas electrostatic interactions and in a minor
centration in unpolluted lakes and streams is typically below to extent, ionic exchange mechanism are responsible for nitrate
0.03 mg/L whereas phosphate concentration above 1.0 mg/L stim- adsorption (Fig. 2).

Fig. 2. Overview of proposed mechanisms for inorganic and organic pollutant adsorption on biochars.
Table 1
Adsorption characteristics of various inorganic contaminants with biochars.

Biochar Pyrolysis Pre-treatment (PR) Post-treatment (PO) Pollutant qm Sorption mechanism Reference
(mg/g)
Oak sawdust 300 °C (PR) Dried at 105 °C for 6 h (<0.2 mm sieved) NH+4 5.31* Electrostatic interactions (acidic Wang et al. (2015a)
functional groups)
(PO) Immersed in deionized water
La-oak sawdust (PR) Dried at 105 °C for 6 h (<0.2 mm sieved) 10.10*
(PO) Immersed LaCl3 solution 0.1 M for 6 h
Giant reed 500 °C (PR) Dried at 100 °C for 48 h NH+4 1.805** Ion exchange Hou et al. (2016)
(PO) < 0.2 mm sieved and dried at 80 °C for 24 h
Corn biochar 400 °C (PR) Dried at 100 °C for 2 h (0.5–2 mm sieved) NH+4 3.93* Ion exchange Vu et al. (2017)
Acid/alkali modified corn (PR) Dried at 100 °C for 2 h (0.5–2 mm sieved) 22.63* Ion exchange/electrostatic interactions
(PO) Immersed HNO3 6 M for 8 h, NaOH 0.3 M for 24 h (1:5 ratio)
Oak sawdust 600 °C (PR) Dried at 105 °C for 6 h (<0.2 mm sieved) NO
3 8.94* Electrostatic interactions Wang et al. (2015a)
(PO) Immersed in deionized water
*
La-oak sawdust (PR) Dried at 105 °C for 6 h (<0.2 mm sieved) 100.0
(PO) Immersed LaCl3 solution 0.1 M for 6 h
MgFe-LHD modified wheat-straw 600 °C under (PR) Dried (<2 mm sieved) NO
3 24.8* Ion exchange Xue et al. (2016)

E. Rosales et al. / Bioresource Technology 246 (2017) 176–192


N2 flow
(PO)Immersed MgCl2 0.3 M and FeCl3 0.1 M (1:100 ratio)
Peanut shell 700 °C n.a. PO3
4 7.56* Electrostatic interactions Jung et al. (2015b)
Corncob 500 °C (PR) 0.5–1.0 mm sieved PO3
4 0.036* Ion exchange Micháleková-Richveisová
et al. (2017)
Garden wood 0.132*
Wood chip 0.269*
Fe-corncob (PR) 0.5–1.0 mm sieved 1.988*
(PO) Immersed Fe(NO3)3, dried at 105 °C overnight
Fe-garden wood 2.754*
Fe-wood chip 3.201*
Oak sawdust 500 °C (PR) Dried at 105 °C for 6 h (<0.2 mm sieved) PO3
4 32.0* Electrostatic interactions (basic Wang et al. (2015a)
functional groups)
(PO) Immersed in deionized water
La-oak sawdust (PR) Dried at 105 °C for 6 h (<0.2 mm sieved) 142.7*
(PO)Immersed LaCl3 solution 0.1 M for 6 h
Malt spent rootlets 850 °C (PR) Dried at 50 °C overnight (0.15–0.18 mm sieved) Hg2+ 103* Physical adsorption Boutsika et al. (2014)
Bi impregnation- wheat straw 500 °C (PR) 0.6–0.8 mm sieved Immersed bismuth sol., HCl (1:100 ratio) As3+ 0.273* Ligand exchange Zhu et al. (2016)
(PO) Washed NaHCO3 0.01 M and distilled water
Algae Slow pyrolysis (PR) Dried at 75 °C for 24 h (<0.5 mm sieved) PO3
4 98.20* Electrostatic interactions Wang et al. (2016a)
500 °C
Magnetic gelatin-modified 500 °C (PO) 0.07–0.2 mm sieved, immersed mixture gelatin, FeCl3 0.02 M and As5+ 49.15* Electrostatic interactions and Zhou et al. (2017)
chestnut shell FeCl2 0.01 M innersphere complexation
Nut shield 600 °C (PO) Dried at 60 °C for 24 h Pb2+ 4.61* Ion exchange and surface complexation Trakal et al. (2016)
FeSO4- nut shield (PO) 60 °C (24 h), immersed iron oxide sol. 50.6*
2+
Sugarcane bagasse <500 °C (PR) Dried at 80 °C for 2 days Pb 86.96* Electrostatic interactions, precipitation Abdelhafez and Li (2016)
and ion exchange
Orange peel 27.86*
Algae-dairy-manure 400 °C 500 °C (PR) Anaerobic co-digestion for 92 days and dried at 65 °C (<2 mm Cu2+ 21.12* Surface complexation Jin et al. (2016)
sieved)
KOH modified-algae-dairy-manure (PR) Anaerobic co-digestion for 92 days and dried at 65 °C (<2 mm 50.71*
sieved)
(PO) Immersed KOH 2 M (1:4 ratio) Dried at 105 °C overnight
HNO3-modified cactus fibres 600 °C (PR) Hand-peeled, washed and dried at 70 °C Cu2+ 3.5* Innersphere complexation Hadjittofi et al. (2014)
(PO) Immersed HNO3 12 M at room temperature for 24 h and at 80 °C
for 3 h. Dried at 100 °C
Steam-activated Porphyra tenera 500 °C (PR) 1.0–5.7 mm sieved Cu2+ 75.1* Ion exchange and physical adsorption Park et al. (2016)
(PO) Steam activation (40% water vapour at 700 °C for 1 h)
2+ *
Composted swine manures 400 °C (PR) Immersed HCl 0.1 M (<2 mm sieved) Cu 21.88 Surface precipitation Meng et al. (2014)
700 °C 9.15*

179
(continued on next page)
180 E. Rosales et al. / Bioresource Technology 246 (2017) 176–192

2.1. Nitrogenated ions: ammonium and nitrate

Goswami et al. (2016)


Several researches have demonstrated the influence of pyroly-

Zhang et al. (2015a)


Trakal et al. (2016)
tic temperature on surface chemical composition. Low-
Kim et al. (2016)
Gao et al. (2015)

Li et al. (2017b)
temperature biochars exhibit high content of oxygen-containing
groups, which contribute to increase the cation exchange capac-
Reference

ity (Gao et al., 2015). Wang et al. (2015a) have analyzed the
influence of pyrolytic temperature ranging from 300 to 600 °C
on ammonium and nitrate removal by biochar prepared from
Innersphere complexes and ion exchange oak sawdust with/without LaCl3. The results exhibited ammo-
Ion exchange and surface complexation
Ion exchange and surface precipitation

nium and nitrate adsorption capacity around 10.1 and


100.0 mg/g after modification, which significantly enhances the
Electrostatic interactions and ion

adsorption ability compared to raw material. As can be seen in


Table 1, other studies have evaluated ammonium sorption on
Electrostatic interactions giant reed-biochar (Hou et al., 2016) as well as a mixture of
Electrostatic interaction
Surface complexation

brewer spent grains and sewage sludge enriched with


Sorption mechanism

magnesium and phosphorous (Zhang and Wang, 2016). However,


the obtained results showed sorption capacity below 5 mg
NH4-N/g. The combined treatment by strong acid and alkali on
exchange

corncob-biochar was analyzed by Vu et al. (2017), the highest


uptake (22.6 mg NH4-N/g) was achieved after immersion into
HNO3 6M and NaOH 0.3 M for 8 h and 24 h, respectively. This
67.363*
72.369*

167.31*
81.096*

can be explained by the formation of sodium organic salts which


71.43**
72.43**
(mg/g)
96.18*

30.0*
137*

273*
311*

increase the segregation of modified biochar. Moreover, previous


20*
qm

studies demonstrated relatively high selectivity for nitrate ions


Pollutant

on Mg/Fe layered double hydroxide (LDH) modified wheat-


straw biochar achieving 5.62 mg NO3-N/g. The modification of
Cu2+

Cd2+

Cd2+

Cd2+
Cr6+

Cr6+

metals on LDH has influence on pollutant selectivity with


Ni/Fe-LDH or Mg/Fe-LDH reported as appropriated for nitrate
(PR) Immersed Zn (NO3)2 (20%) under 130 rpm, 30 °C and 24 h. Dried

(PO) Immersed FeCl3 1 M 80 °C for 3 h, dried at 105 °C and 600 °C for

ions adsorption (Xue et al., 2016).


(PO)Immersed KOH 1 M (1:1 ratio) and dried at 100 °C for 24 h

(PO) Immersed KOH solution for 2 h (1:3 ratio), dried at 105 °C

(PO) Immersed NaOH 2 M 100 °C 12 h, rinsed NaHCO3 0.01 M

(PO) Immersed chitosan/acetic acid (2%) (1:6 ratio), 150 mL

2.2. Phosphorous ions


(PO)Steam activation (40% water vapour at 700 °C for 1 h)

Several studies exhibited poor adsorption capacity for


(PO) Dried at 60 °C for 24 h, immersed iron oxide

phosphate ions from aqueous solution with different biomass


(PR) Dried at 105 °C for 12 h (2–5 mm sieved).
(PR) Dried at 105 °C for 12 h (2–5 mm sieved)

(PR) Dried at 105 °C for 12 h (2–5 mm sieved)

(PR) Dried at 105 °C for 12 h (2–5 mm sieved)

feedstock including peanut shell, bamboo and maize waste


(Table 1) (Jung et al., 2015b). For this reason, biochar production
Pre-treatment (PR) Post-treatment (PO)

from biomass with high content of calcium, iron, magnesium or


aluminium is considered and a focus of attention on scientific com-
munity. Several authors have reported biochar modifications
through functional additives such as metals, metals oxides or
(PO) Dried at 60 °C for 24 h

1 h, rinsed NaHCO3 0.01 M

metal chlorides to enhance the maximum adsorption capacity (M


(PR) 1.0–5.7 mm sieved

(PR) 1.0–5.7 mm sieved

icháleková-Richveisová et al., 2017). Based on the literature,


(PR) < 0.2 mm sieved

glutaraldehyde (1%)

Wang et al. (2015a) produced lanthanum-modified biochar from


oak-sawdust; phosphate sorption capacity was improved
(142.7 mg/g) compared to raw material (32.0 mg/g). However, this
at 60 °C

method involves high economic costs and operational problems.


Therefore, the development of simple and cost-effective methods
to modify biochar with natural minerals such as vermiculite
represents an appropriate alternative to enhance the phosphate
adsorption onto biochar (Wang et al., 2016a). They have
Pyrolysis

successfully evaluated novel SiO2–biochar nanocomposites


450 °C

500 °C

600 °C

400 °C

600 °C

600 °C

prepared by thermal pyrolysis of vermiculite treated algal biomass.


This material showed high effectiveness on phosphate removal
(159.42 mg/g) compared to unmodified biochar because the SiO2
Zinc-modified sugarcane bagasse

Chitosan- Eichhornia crassipes/c-

particles on the surface corresponds to activated sites through


KOH-modified Ipomoea fistulosa
Steam activation E. compressa

electrostatic interactions.
KOH activation E. compressa

c-Fe2O3 Eichhornia crassipes


NaOH-modified rape straw
MnOx-modified rape straw

FeOx-modified rape straw

3. Heavy metals
Fe2O3 composite
FeSO4-wheat straw
Ipomoea fistulosa
Table 1 (continued)

n.a. not available.

The presence of heavy metals in drinking water represents an


Wheat straw

Langmuir.

important threat to the environment and human health. These


Biochar

metallic elements have been recognised as mutagenic and carcino-


genic agents, and may be toxic for aquatic life (EPA, 2017). They are
*

considered systemic toxicants, which are well known to cause


E. Rosales et al. / Bioresource Technology 246 (2017) 176–192 181

multiple organ damage, even at low exposure level. The applica- 3.2. Arsenic
tion of adsorption methods using biochar-based materials is con-
sidered a suitable via for heavy metals remediation. Based on Arsenic, well-known environmental hazard, has been classified
literature, several mechanisms have been proposed to describe as human carcinogenic at lower exposure levels by International
accurately the metal adsorption on biochar including electrostatic Agency for Research on Cancer (IARC). Although inorganic arsenic
interactions between adsorbent surface and heavy metals, cation is an element in the Earth’s crust, human activities such as agricul-
exchange between metals and protons or alkaline metals on sur- ture, mining or medicine contribute to the arsenic environmental
face, metal complexation with functional groups, metal precipita- burden. Arsenic concentration in surface water is usually below
tion to form no soluble compounds, and metal reduction with 10 ppb, which has been established as maximum arsenic concen-
further reduced metal species adsorption (Fig. 2). The main mech- tration in drinking water. The removal of this metallic element
anism depends on biochar properties (surface area, porosity, pH, from aqueous solution is considered an important challenge on sci-
surface charge, functional groups, and mineral components) and entific community owing to crucial influence of the pH on arsenic
target metals (Li et al., 2017a). The modifications on the biochar speciation (Zhou et al., 2017). Despite iron oxy-hydroxide powders
surface allow enhancing sorption capacity of heavy metals onto are considered the most efficient adsorbent for arsenic removal
biochar (Table 1). The impregnation of biomass feedstock with from aqueous solution; additional costs corresponding to further
metal ions such as AlCl3, MgCl2, KMnO4 and ZnCl2 enhances the adsorbent separation may hinder their application at industrial
pore properties and oxygen-containing functional groups and, scale. Several researches have reported the use of modified-
therefore, major interactions occur between biochar and heavy biochar as alternative low-cost adsorbent with excellent results.
metals through the formation of surface complexes, cation-p bond- Zhou et al. (2017) synthetized a novel chestnut shell-biochar mod-
ing, electrostatic attraction and ion exchange (Fig. 2). On the other ified with magnetic gelatin and ferric ion, the results showed the
hand, the base activation promotes high surface areas and addi- biochar modification enhanced significantly the physiochemical
tional hydroxyl radicals, which increase the sorption capacity of properties and increased the oxygen-containing functional groups.
heavy metals. Regarding to magnetic treatments, obtained modi- Therefore, maximum adsorption capacity has increased after
fied biochars have an expanded average pore diameter, which chemical treatment achieving 45.8 mg/g under optimal conditions.
enhance adsorption properties for anionic pollutants and ensure On the other hand, Zhu et al. (2016) reported bismuth oxide as an
an optimal separation of biochar particles on further treatments appropriate treatment to impregnate wheat straw biochar
(Li et al., 2017b). (16.21 mg/g for As3+); the mechanism may be attributed to the
Lewis acid-base (ligand exchange) reaction between bismuth atom
3.1. Mercury and arsenite. Moreover, two manganese oxide-modified biochar
composites were proposed by Wang et al. (2015b) to remove
Mercury, natural element found in the Earth’s crust, has been arsenic from wastewater; pine wood feedstock was pyrolyzed in
widely released into the aquatic environments owing to anthro- the presence of MnCl24H2O and other, was impregnated with bir-
pogenic activities such as agriculture, mining, incineration and nessite via precipitation. The birnessite presents high oxidation
municipal and industrial wastewater discharges (Tan et al., potential able to transform arsenite to arsenate. Although the
2016). The US Environmental Protection Agency has established results showed higher adsorption capacities compared to unmodi-
maximum mercury concentration around 10 lg/L for wastewater fied biochar, low arsenic removal percentages were achieved.
discharge and 2 lg/L for drinking water (EPA, 2017). Few studies Recent studies have evaluated the generation of LDH with different
were found related to mercury adsorption using biochar over last metal combinations for As5+ removal (Wang et al., 2016b). For this
years. Boutsika et al. (2014) have evaluated the potential use of purpose, two LDH-biochar were produced by pyrolysis of Ni/Fe-
biochar produced from malt-spent rootlets to remove mercury LDH-modified pine feedstock or precipitation of LDH onto pristine
ion from aqueous solution. The results exhibited excellent adsorp- biochar. The maximum arsenate adsorption capacities under neu-
tion ability, 103 mg/g under optimal conditions, onto brewing tral conditions were 1.56 and 4.38 mg/g, respectively higher than
industry by-product. The speciation diagram shows mercury ion unmodified material (0.20 mg/g). The obtained results suggested
corresponds to its main form in the solution at pH below 6 while multiple mechanisms (anion exchange and surface complexation)
mercury precipitates as hydroxide at pH above 6 (Boutsika et al., takes place on adsorption process.
2014). Further studies were carried out by Manariotis et al.
(2015) who reported the influence of pyrolytic temperature on 3.3. Lead
mercury adsorption using malt spend rootlets as biomass feed-
stock. The adsorption capacity increased by factor around 4 and Lead compounds have been widely used as industrial raw mate-
6 for low- and high-temperature biochars compared to the raw rial in lead-acid batteries, rolled extrusions such as lead sheet, wire
material. These results corroborated that high surface area and less and pipe, pigments, cable sheathing, munitions and lead alloys. The
functional groups promote the mercury species adsorption owing maximum allowable concentration for lead in drinking water is
to non-polar adsorbent with the concomitant high adsorption for 15 lg/L (EPA, 2017). Several remediation technologies including
organic hydrophobic compounds and heavy metals. On the other chemical precipitation, ions exchange, filtration, biosorption,
hand, modifications of biochar with chemical reagents provide dif- biodegradation and phytoremediation have been developed to
ferent physicochemical characteristics (surface area, pore size, remove lead compounds from aqueous ecosystems. However,
molecular weight, hydrophobicity, polarity and functional groups) energy requirements and sludge generation promote the develop-
leading to adsorption capacity changes (Tan et al., 2016). The ment of other technologies such as adsorption on biochar derived
chemical modification by 2 M Na2S and 2 M KOH on corn-straw biomass feedstock over last years (Abdelhafez and Li, 2016). The
biochar which was obtained via slow pyrolysis at 500 °C was eval- adsorption characteristics of celery-biochar, rich in alkaline miner-
uated by Tan et al. (2016). After alkali treatment, O-containing als, for lead removal from aqueous solution have been investigated
groups raised suggesting higher sorption ability through electro- by Zhang et al. (2017a). The alkali earth metals (Na, K, Ca and Mg)
static interactions. The results showed potassium and sulphur were converted into carbonates throughout high-temperature
impregnation were efficient treatments to enhance aqueous mer- pyrolysis leading to alkaline medium and heavy metal cation pre-
cury removal increasing by 31.12% and 76.95%, respectively, com- cipitation. The high biochars’ effectiveness on lead removal from
pared to pristine material. aqueous solution was reflected (around 300 mg/g). The contribu-
182 E. Rosales et al. / Bioresource Technology 246 (2017) 176–192

tion to lead removal by insoluble organic carbon was below 4% biochar derived from red macroalgae Porphyra tenera, even though
whereas inorganic minerals played a dominant role on lead the increase of specific surface area and pore volume. On the
adsorption, these results are in accordance to previous studies contrary, steam activation enhances sorption ability owing to the
reported on literature (Wang et al., 2015c). To increase lead ion formation of mesopores and macropores along decomposition of
adsorption ability on biochar derived from different biomass feed- volatile matter. The pH effect was evaluated by Kim et al. (2016)
stock, Trakal et al. (2016) evaluated the use of microwave- who analyzed fast-pyrolysis biochar derived from green
synthesised magnetic iron oxide particles. The magnetic modifica- macroalgae Enteromorpha compressa reaching 137 mg/g at
tion often decreases specific surface area owing to the formation of optimal conditions (pH around 5.5). According to the speciation
secondary iron (hydro) oxides on the surface. This might improve diagram, Cu2+ mostly exists at pH below 5.5 whereas the formation
the sorption efficiency of metals and metalloids, which are fixed of Cu(OH)+ complex and Cu(OH)2 precipitation may occur at pH
on the surface of these precipitated (hydro) oxides. Although mag- above 6, respectively. Novel biochar was synthesized through
netization enhances more adsorption effectiveness with well- slow pyrolysis of hickory wood pretreated with KMnO4 for Cu2+
developed structures and mobile metals, lead adsorption was removal from aqueous solution owing to other studies have
increased significantly after modification achieving around reported the conversion from KMnO4 to MnO2 ultrafine particles
300 mg/g for nut shield-biochar. As can be seen in Table 1, lead (Wang et al., 2015d). The maximum adsorption capacity of
ion showed higher affinity to biochar adsorption than other tested engineered biochar for Cu2+ removal was above 34 mg/g, which
metals. This fact could be explained as a result of its low hydrated was about 2.8 times greater than those corresponding to pristine
radius, high electronegativity and hydrolysis constant compared to biochar.
other metal cations.

3.4. Zinc 3.6. Cadmium

Zinc has been released into the environment from natural and Cadmium has been used for metal surface treatment as stabiliz-
anthropogenic sources; however, human activities such as refining, ing agent in polyvinyl chloride products and alloys. However, the
electroplating, steel production, carbon combustion and solid most important cadmium source comes from cigarette smoke
waste incineration are mostly responsible for zinc pollution. EPA owing to the average cigarette cadmium content exceeds
has set the threshold limit of zinc in wastewater and drinking 1.60 lg/g. The EPA has established maximum cadmium concentra-
water as 5 and 3 mg/L, respectively. Alkali modification has been tion as 10 lg/L for drinking water (EPA, 2017). Scarce ability of
commonly used to increase the sorption ability of biochars by high unmodified biochars to remove heavy metals from wastewater
specific surface area and functional groups modification. For this has been reported on literature. For this reason, to evaluate the
reason, Ding et al. (2016) used NaOH to modify biochar derived influence of three modifications (NaOH base treatment, MnOx
from hickory wood chips and evaluate zinc adsorption ability. impregnation with KMnO4 and magnetic modification by thermal
The modified biochar showed around 2.5 times zinc adsorption pyrolysis of FeCl3), Li et al. (2017b) have investigated the use of
capacities than pristine biochar (0.71 mg/g). Furthermore, Frišták modified-biochars from rape straw. According to them, the main
et al. (2015) evaluated the pH effect on adsorption process by mechanisms responsible for Cd2+ removal using MnOx-biochar
two different woody-derived biochars. Zinc is present as Zn2+ ions were cation exchange and cation-p bonding achieving the greatest
at pH from 3.0 to 8.5 whereas an enhancement of soluble hydrox- adsorption capacity (81.10 mg/g) as a result of the increase of the
ide forms concentrations such as Zn(OH)2 and ZnOH+ are generated specific surface area, number of oxygen-containing functional
at pH above 7.5. The authors suggested the precipitation and ion groups and the pore size and structure. On the contrary, magnetic
exchange as main processes involved into adsorption of metals biochar prepared by chemical precipitation showed reduced sur-
cation. face area and more acidic groups such as carboxylic acid and phe-
nolic groups on the surface, which were occupied by protons
3.5. Copper leading to competitive interactions with cadmium ions. As was
commented above in lead section, magnetite/maghemite modified
Copper and its alloys (brass and bronze) have been employed in biochar are more efficient with well-developed structure and
common household wiring, photovoltaic cells and phytosanitary mobile metals such as Cd2+ achieving 50.6 mg/g for nut shield
products as fungicides, fertilizers, algicides and insecticides. The and 75.3 mg/g for wheat straw, which represents 10.97 and 3.17
EPA has set the acceptable limit of copper in drinking water as times better than those reported for unmodified biochar (Trakal
1.3 mg/L. The removal of Cu2+ from aqueous solution via adsorp- et al., 2016). On the other hand, the use of aquatic plants such as
tion has been widely used on wastewater treatment plant (Jin Ipomoea fistulosa to produce biochar able to remove Cd2+ ions from
et al., 2016). In recent times, the use of biochar as low-cost adsor- aqueous solution was investigated by Goswami et al. (2016). The
bent has been increased because of its high efficiency for copper effect of KOH activation and pyrolytic temperature on sorption
removal. Among them, Meng et al. (2014) removed copper by using process was evaluated resulting in higher adsorption capacity for
biochar derived from swine manure, and the results suggested sil- biochar pyrolyzed at temperatures below 400 °C and after alkali
icate and phosphate particles contained on biochar were the main modification (41.67–72.43 mg/g). The authors suggested surface
sorption sites. Nitric acid modified-biochar produced from cactus complexation as main mechanism responsible for Cd2+ removal.
fibres was also applied for Cu2+ removal identifying carboxyl moi- Biochar derived from biophysical dried sewage sludge through
eties as main binding site (Hadjittofi et al., 2014). On the other pyrolysis at 900 °C was evaluated by Chen et al. (2015a). The pH
hand, marine macroalgae are considered an appropriate alternative solution and adsorbent dosage represented the dominant factors
biomass material to obtain biochar (Park et al., 2016). Jin et al. on adsorption process whereas the influence of temperature was
(2016) analyzed biochars pyrolyzed from anaerobically digested considered as negligible. The results showed a minimal
algae-dairy-manure slurry before and after activation by 2 M contribution of organic matters on cadmium removal whereas
KOH solution. The alkali modification enhanced sorption capacity alkaline earth metals such as calcium are responsible for adsorp-
(50.71 mg/g) compared to unmodified material (21.12 mg/g). tion process owing to cation exchange and surface precipitation
These results are not in agreement with Park et al. (2016) who by the formation of insoluble cadmium salts under alkaline
reported KOH activation reduced the copper removal on pyrolytic conditions.
E. Rosales et al. / Bioresource Technology 246 (2017) 176–192 183

3.7. Chromium 4.1. Adsorption mechanisms for organic pollutant removal

Chromium, well known carcinogenic and mutagenic, has been The adsorption mechanisms for organic removal depend on the
identified as hazardous pollutant by EPA, which set 0.05 mg/L as pollutant properties as well as biochar surface chemistry, which
maximum permissible concentration in drinking water. This varies greatly with raw material and proposed pyrolysis condi-
metal has been widely used in many industrial processes such tions. In general, high treatment temperatures increase the aro-
as photography, metallurgy, electroplating and tanneries (Shang maticity, hydrophobicity, and surface area of biochars making
et al., 2017). Strong electrostatic repulsion forces take place them more accessible to hydrophobic organic pollutants. Thus,
between biochar surface and inorganic anions HCrO 2
4 or CrO4 . Chen et al. (2017) perceived that the increase in pyrolytic temper-
As a consequence, surface modifications have been evaluated ature resulted in the enlarged pore volume of biochar and
over past years to enhance adsorption ability. Among them, enhanced affinity of carbamazepine on peanut shell-derived bio-
the use of nanoscale FeS particles showed higher reactivity com- chars. The relationship between the H/C and O/C molar ratios rep-
pared to bulk particles and natural minerals owing to small par- resents an interesting parameter which shows the effect of the
ticle size with the concomitant larger specific surface area. pyrolysis conditions on the biochars. Both molar ratios decreased
Recently, Lyu et al. (2017) have developed novel FeS-biochar by increasing heat treatment temperatures mainly due to the
composite using carboxymethyl cellulose as stabilizing being demethylation and decarboxylation reactions (Devi and Saroha,
the FeS particles soldered onto oxygen-containing functional 2015; Sun et al., 2016). The decrease of H/C is indicative of the for-
groups. Higher chromium adsorption capacity (130.5 mg/g) was mation of highly carbonized and aromatic structures. Chen et al.
obtained compared to FeS (38.6 mg/g) and biochar (25.4 mg/g), (2015b) determined H/C molar ratio for bamboo char was in the
and the main mechanisms to describe metal removal were range of 0.15–0.45 and the ratio decreased by increasing pyrolytic
adsorption, precipitation and reduction. Several authors have temperature, indicating that the biochar was highly carbonized
proposed the use of biochars as support materials for nanoscale and had a highly aromatic structure. The reduction of O/C molar
zero-valent iron nanoparticles (nZVI) stabilization (Shang et al., ratio is also indicative that the surfaces of the high-temperature
2017). The carboxyl groups and silicon minerals were responsi- biochars became less hydrophilic (Reguyal et al., 2017). On the
ble for supporting nZVI and the main mechanisms for Cr6+ other hand, biochars obtained at low temperature consists of
removal by nZVI include adsorption, co-precipitation and reduc- amorphous carbon that can contribute to the fast adsorption of
tion of Cr6+ to Cr3+ via nZVI oxidation (Shang et al., 2017). Chi- pollutants (Chen et al., 2017). However, some studies demon-
tosan, renewable transformed polysaccharide, has been used as strated that pyrolysis at higher temperatures can reduce the
stabilizing to attach ZVI particles on bamboo biochar surface removal ability. Thus, Chen et al. (2015b) studied the adsorption
(Zhou et al., 2014). The use of chitosan on biochar enhanced of N-nitrosodimethylamine, polar and highly water soluble com-
chromium removal compared to unmodified material because pound, onto bamboo biochar obtained at different temperatures
the presence of amine functional groups on chitosan structure 300–700 °C. The best removal efficiency was reached with the bio-
leads to strong chelations to cation metal in aqueous phase. char manufactured at 500 °C. They suggested that at high temper-
Based on previous studies, Zhang et al. (2015a) have developed atures biochar can show structural defects and the micropores on
chitosan modification of magnetic biochar derived from Eichhor- the surface blocked.
nia crassipes, aquatic tropical species characterized as aggressive The biochar surface is heterogeneous as a result of the co-
invasive plant. The results indicated an enhancement on adsorp- existed carbonized and non-carbonized fractions in which different
tion capacity after chitosan modification (120 mg/g) compared to adsorption mechanisms take place. The various mechanisms pro-
raw material (30 mg/g). The effect of background electrolyte on posed (hydrophobic interaction, pore-filling adsorption/partition,
chromium removal may be explained by competitive mechanism electrostatic attraction, and hydrogen bonding or p–p interaction)
owing to monovalent anions slightly compete for positive charge for the interaction of biochar with organic pollutants are summa-
sorption sites on surface whereas multivalent anions restricted rized in Fig. 2.
adsorption process. The sorption of organic pollutants onto biochar can be affected
by their solubility by means of hydrophobic interactions. Thus, sol-
uble pollutants can be attached into hydrophobic biochar if they
4. Organic pollutants have hydrophobic functional group, as for example methyl group,
in their structure (Reguyal et al., 2017).
Several studies have demonstrated that laboratory-produced Diffusion of organic compounds onto biochar can occur by dif-
biochars can adsorb organic pollutants comparably with com- fusion of sorbates into pores (adsorption-pore-filling) or into the
mercial activated carbon and, therefore, provide a cost- organic matter matrix of noncarbonized fractions (partition)
effective alternative to this in water treatment applications (Inyang and Dickenson, 2015). Zhang et al. (2015b) reported that
(Kearns et al., 2014). In addition, the regeneration of biochar trichloroethylene (TCE) removal on soybean stover biochar was
may be carried out easily by thermal treatment (Abdul et al., due to partition to non-carbonized fractions and adsorption to
2017), microwave irradiation (Zhang et al., 2014), using organic the carbonized fraction of the biochar. Some studies have focused
solvents (Reguyal et al., 2017), or milling (Shan et al., 2016). the attention in order to analyze the contribution between adsorp-
Furthermore, organic chemicals can be directly degraded on tion and partition. Thus, Devi and Saroha (2015) studied the frac-
biochar surface by macromolecular free radicals. Thus, Yang tion of adsorption and partition of pentachlorophenol on biochar
et al. (2016) observed obvious degradation of p-nitrophenol in obtained at different temperatures 200–700 °C. They found that
the presence of biochars, through the detection of NO 3 as well the fraction of adsorption increased by increasing pyrolysis tem-
as organic by-products. perature due to the carbonized fraction increases. The sorption of
The recent studies on biochar applications for the remediation organic pollutants onto biochar by pore-filling is function of total
of aquatic environment containing organic compounds are sum- micropore and mesopore volumes. Han et al. (2014) found that
marized in Table 2. The greatest concern of organic pollutants nanoporefilling plays a dominant role in the sorption of hydropho-
has been focused on drugs and pharmaceuticals, plasticizers, bic organic pollutants such as phenanthrene by rice straw biochar
organic solvents, pesticides/herbicides, polycyclic aromatic obtained at 600 °C, which is found to be microporous solids.
hydrocarbons and dyes. Similarly, Sun et al. (2016) stablished a correlation along with
Table 2

184
Adsorption characteristics of various organic contaminants with biochars. Sorption mechanism: p (p-p EDA interaction), HI: (Hydrophobic interaction), Pp (Partition process), H-b (H-bonding), Ea (Electrostatic attraction), Ca (Complexes
adsorption).

Biochar Pyrolysis Pre-treatment (PR) Post- treatment (PO) Pollutant Isotherm N(%) H O (%) C(%) Surface Sorption Reference
Model (%) Area mechanism
(m2/g)
Pine needle 700 °C 2 h at 5 °C/min (PR) dip-coating with graphene oxide Dimethyl phthalate Freundlich 0.39 1.55 9.86 84.98 443.9yy p Abdul et al.
wood oxygen-limited synthesized via thermal reduction Diethyl phthalate (2017)
conditions Dibutyl phthalate
Bamboo 500 °C (PO) HCl + HF (1 M) N-nitrosodimethylamine Slips 0.75 2.44 11.7 85.11 38.47y H-b/HI/Pp Chen et al.
(2015b)
Peanut shell 500 °C 4 h oxygen- n.a. Carbamazepine Dubin- 1.60 2.71 19.1 59.45 64.73 y
HI/p/H-b Chen et al.
limited conditions Ashtakhov (2017)
Peanut shell 700 °C 4h N2 (PO) (1:3, v/v) HNO3/H2SO4 Dimethyl phthalate Freundlich 0.66 2.4 42.46 48.54 281 y
HI/p Ghaffar
atmosphere Diethyl phthalate et al. (2015)
Dibutyl phthalate
Pine wood 600 °C 1 h oxygen- (PO) 0.1 M HCl + NaClO2 + CH3COOH Phenanthrene Freundlich 0 2.2 25.8 65.1 402.1yy Nanoporefilling Han et al.
limited conditions (2014)
y
Sugarcane 600 °C 1 h at 10 °C/min (PR) dip-coating with carboxyl-functionalized Methylene blue Langmuir– 0.7 1.7 11.9 85.7 390 Ea Inyang

E. Rosales et al. / Bioresource Technology 246 (2017) 176–192


bagasse N2 atmosphere carbon nanotubes. Freundlich et al. (2014)
Phragmites 600 °C 2 h oxygen- (PO) 1 M HCl Pentachlorophenol Freundlich 0.61 2.07 8.9 87.8 88.35 y
HI/p Peng et al.
australis limited conditions (2016)
Pinus radiata 650 °C slow pyrolysis (PO) coating with Fe3O4 via oxidative Sulfamethoxazole Redlich- <0.5% 0.82 n.a. 55.77 125.8y HI Reguyal
sawdust hydrolysis in alkaline media Peterson et al. (2017)
Coconut shell 500 °C 1.5h N2 (PO) biochar mixture with Fe3O4 (3:1) was Carbamazepine tetracycline Langmuir n.a. n.a. n.a. 365y Surface Shan et al.
atmosphere milled at 550 rpm for 6 h (40:1 ratio balls complexation with (2016)
mixture) Fe3O4/p
Pelletized pine 850 °C 20 min Top-lit n.a. Sulfamethoxazole n.a. 0.12 0.54 2.54 93.7 405y p Shimabuku
forestry up-draft gassifier et al. (2016)
waste
Rice straw 450 °C 1 h at 10 °C/min (PO) 0.1 M HCl + 1 M HCl and 10% HF Propiconazole Freundlich 1.17 3.71 15.4 72.5 11.5y Pore-filling Sun et al.
oxygen-limited (2016)
Pine wood dust 600 °C 1 h oxygen- (PO) 0.1 M HCl Acetochlor dibutylphthalate Freundlich 0 n.a. 8.13 88.78 544.6yy Pore-filling Wang et al.
limited conditions 17a-ethynylestradiol (2016c)
phenanthrene
Pine wood 700 °C 4 h oxygen- n.a. p-nitrophenol n.a. 0.39 2.94 13.4 82.3 566y Adsorption- Yang et al.
limited conditions degradation by free (2016)
radicals
Rice straw 600 °C 2 h at 7 °C/min (PR) dip-coating in carboxyl functionalized Sulfamethazine Freundlich 0.51 1.83 10.45 87.21 387.2y Pp/p/H-b Zhang et al.
absence of air multi-walled carbon nanotube (2016)
y
Pristine hickory 600 °C 1 h at 10 °C/min (PR) dip-coating with carboxyl-functionalized Sulfapyridine Pb Langmuir- 0.19 2.07 20.05 77.69 359 HI/Ea Inyang
N2 atmosphere carbon nanotubes Freundlich et al. (2015)
Rice 450 °C 2 h + 200 °C 2h (PO) 0.1 M HCl Phenanthrene dibutylphthalate Freundlich 0.83 3.31 11.8 57.9 293.4 yy
Ca/p Jin et al.
N2 atmosphere (2014)
Loblolly pine 300 °C 15min N2 (PO) 4 M NaOH dried 105 + 2 h 800 °C Diclofenac naproxen ibuprofen Freundlich 0.65 0.77 21.3 72.6 1360y p/Ea Jung et al.
chip atmosphere + 0.1 M HCl (2015a)
Bamboo 550 °C (PO) 800 °C 2h N2 flow + 2 h air flow Furfural Langmuir 0.58 1.1 27.6 707.7 492.2y HI/p Li et al.
(2014)
y
Bagasse 600 °C 1h N2 (PRE) suspension with montmorillonite Methylene blue Freundlich 0.75 2.25 18.87 75.31 407 Ion exchange/Ea Yao et al.
atmosphere (2014)
Wheat straw 550 °C 5 min (PO) magnesium hydroxide-coated Direct yellow 12 Langmuir n.a. n.a. n.a. n.a. n.a. H-b Zhang et al.
(2014)
Soybean stover 700 °C 3 h 7 °C/min n.a. Trichloroethylene NR 1.3 1.27 15.45 68.81 420.3y Pp Zhang et al.
(2015b)
Peanut shell 450 °C 2h N2 (PO) impregnated Cu(NO3)2 Doxycycline hydrochloride Langmuir n.a. n.a. n.a. n.a. n.a. Ea/Strong Liu et al.
atmosphere complexation (2017)

n.a. not available.


y
BET.
yy
CO2.
E. Rosales et al. / Bioresource Technology 246 (2017) 176–192 185

the positive relationship between normalized CO2-surface area and nisms. Similar findings were also detected by Zhang et al. (2016).
aromaticity which indicates that pore-filling in nanopores within They reported that the primary mechanisms for the adsorption of
aromatic C dominate nonlinear sorption of propiconazole by ionizable antibiotic, sulfamethazine, included partition caused by
plant-residue and animal waste-derived biochars. Depending on Van der Waals forces and adsorption caused by hydrogen bonding
the biochar type, polar and/or non-polar organic compounds can and p-p EDA interaction.
be sorbed by pore-filling on carbonized fractions of biochar
(Inyang and Dickenson, 2015). In addition the isotherms highly
nonlinear (Freundlich n < 0.71) reflect the predominance of 4.2. Influence of biochar aging and solution chemistry on organic
adsorption/pore-filling mechanisms (Han et al., 2014). pollutant adsorption
Electrostatic interaction between biochar surface and organic
pollutants is another plausible adsorption mechanism. Biochar sur- The solution chemistry (pH, salt content, dissolved organic mat-
faces are normally negatively charged, which could facilitate the ter. . ..) and the aging of biochar can modify the adsorption process.
electrostatic attraction of positively charged organic compounds Therefore, the study of the process under different environments
(Ahmad et al., 2014). Inyang et al. (2014) suggested that electro- and conditions is desirable in order to determine the viability of
static attraction was the dominant mechanisms for the methylene the biochar for practical applications. Accordingly, several authors
blue sorption onto the carbon nanotube–biochar nanocomposites, have evaluated the influence of these parameters in the removal of
though diffusion controlled its rate. Similarly, Yao et al. (2014) organic pollutants. Thus, Peng et al. (2016) determined that the pH
described that the sorption of the same dye on the clay–biochar had significant effects on the adsorption of pentaclorophenol by
composites was mainly controlled by the ion exchange (with clay) reed biochar which could well-explained by hydrophobic, hydro-
and electrostatic attraction (with biochar) mechanisms. gen bonding, electrostatic repulsion and p-p interactions. Equally,
Hydrogen bonding (H-bonding) is another mechanism for the Liu et al. (2015) found strong pH-dependence of the adsorption of
sorption of polar organic pollutants on biochars. The sorption of atrazine on agricultural waste biochars. More atrazine was
dibutyl phthalate by rice straw and swine manure biochars was removed at alkaline than at acidic pH values mainly by the occur-
attributed to H-bonding between H-donor groups or water mole- rence of hydrolysis. Thus, adsorption behavior at lower pH values
cules on biochar and O-atoms on dibutyl phthalate ester group could be explained by a strong affinity between the protonated
(Jin et al., 2014). Chen et al. (2017) demonstrated that the fast biochar surface and atrazine.
adsorption mechanism on carbamazepine on peanut shells biochar The adsorption mechanisms p-p EDA can be greatly affected by
is due to the H-bonding between pollutants and AOH on the sur- pH value. A low pH causes protonation of the carboxyl groups,
face of minerals (CaCO3, KAlO2, and quartz) present in the biochar. which should favor the sorption of strong p -donor compounds
In addition, some researches also detected that the electrostatic affecting to the p-p EDA interactions (Inyang et al., 2014). Xie
repulsion between negatively charged organic compounds and bio- et al. (2014) reported that pH has significant effects on the adsorp-
chars could promote H-bonding and induce adsorption (Ahmad tion of the sulfonamides, mainly by affecting the speciation of the
et al., 2014). adsorbate molecules and thus, their p-electron donating capability
The p-p electron donor acceptor (p-p EDA) interaction has been as well as hydrophobicity. In contrast, Shan et al. (2016) reported
proposed as one of the important mechanisms responsible for the that solution pH values slightly affected the sorption of the phar-
adsorption of chemicals with benzene rings. Thus, benzene rings maceuticals carbamazepine and tetracicline on the different bio-
can donate p electrons to p-electron acceptors as the functional chars studied.
groups. Xie et al. (2014) reported that the adsorption of sulfon- The presence of dissolved organic matter can negatively affect
amides by different pine wood biochars correlate well with the pollutant sorption by biochars as reported by Shimabuku et al.
degree of graphitization, indicating that p-p EDA interaction (2016). Likewise, when electrostatic forces between sorbent sur-
between the sulfonamides molecules and the graphitic surfaces faces and sorbate ions are attractive, an increase in ionic strength
of the biochars was the predominant adsorption mechanism. will decrease the sorption capacity of the sorbate due to competi-
The presence of different mechanisms for pollutant removal tion of cations with positively charged organic pollutants for sorp-
was also reported in different studies. Chen et al. (2015b) sug- tion sites (Inyang et al., 2014). Zhang et al. (2016) also studied the
gested three sorption mechanisms H-bonding between AN@O influence of salt ionic and humic acid strength on the removal of
and the O-containing moieties, hydrophobic force between ACH3 sulfamethazine by biochar carbonaceous nanocomposites. Their
and the ordered graphitic structure, and partition process of N- results showed that an increase in NaCl concentration also
nitrosodimethylamine into the non-carbonization part of biochar. increased slightly pollutant uptake due to the electrostatic screen-
Furthermore, the benzene rings in polychlorinated biphenyls mole- ing of the surface charge by the counterion species added. On the
cules can act as p-acceptor due to the substitution of chlorine contrary, Xie et al. (2014) observed that the presence of Cu2+ and
atoms which would then interact with the electron rich aromatic humic acids inhibited the adsorption of the sulfamethoxazole
moieties of the biochar via p–p EDA interaction. Chen et al. and sulfapyridine through the competition of available O-
(2017) found that hydrophobic and p-p interactions are the dom- functionalities or pores, respectively. Other studies stated that
inated adsorption mechanism of carbamazepine on peanut shell the coexisting cation has different influence on pollutant removal.
biochar. Thus, Liu et al. (2017) observed that doxycycline hydrochloride
It was also reported that the presence of different organic com- adsorption to Cu-biochar was slightly facilitated by Na+, K+ and
pounds could modified the mechanism. Accordingly, Jin et al. inhibited by Ca2+, Mg2+, Mn2+. However, the influence of these
(2014) found that phenanthrene (PHE) and dibutyl phthalate parameters is strongly dependent of the system studied. Thus,
(DBP) adsorption on plant and manure biochars may be attributed solution chemistry such as pH and metal ions did not show obvious
to the sorption of PHE–DBP complexes by biochars because of EDA effect on N-nitrosodimethylamine removal on bamboo biochar
interactions. DBP had more adsorption sites on the biochars in the (Chen et al., 2015b). In the same way, Ferreira et al. (2017)
presence of PHE via the binding form of biochar–PHE–DBP. On the obtained that the variation of operational conditions such as tem-
other hand, Jung et al. (2015a) observed competitive adsorption of perature (8–30 °C) and salinity (0.8–35‰) did not affect the tri-
diclofenac, naproxen, and ibuprofen on NaOH activated biochars. caine methanesulfonate adsorption on pyrolysed paper mill
They established that hydrophobic interactions such as EDA inter- sludge. In addition, they also proved that the presence of organic
actions and polar interactions were the main adsorption mecha- and inorganic matter in aquaculture effluents did not interfere
186 E. Rosales et al. / Bioresource Technology 246 (2017) 176–192

with the adsorption of the studied anesthetic onto the biological The physical pretratement of biochar by using of steam activa-
sludge-based carbon. tion and gas purping may increase the surface area and pore vol-
Incorporation of oxygen during biogeochemical weathering of ume of the biochar by removing the trapped products of
biochar may alter its parent material properties. The elevated oxy- incomplete combustion during thermal treatment. This fact was
gen content due to biochar aging may diversely impact its affinity evidenced by Mondal et al. (2016), who determined the increase
towards variety of organics pollutants (Ghaffar et al., 2015). Devi of surface area of the carbonized char and steam activated char
and Saroha (2014) studied the effect of the magnetic biochar com- increasing the adsorption of ranitidine hydrochloride drug mole-
posites aging on pentachlorophenol removal efficiency. They cule inside the surface of the biochar.
observed that the dechlorination efficiency of the magnetic biochar
composites reduces significantly due to the conversion of zero
5. Continuous flow system for adsorption: Fixed-bed columns
valent iron to iron oxides. But no significant decrease in pen-
tachlorophenol removal efficiency was noticed for the biochar.
Although several types of systems as stirred tank and
fluidized-bed columns can be used to perform the continuous
4.3. Biochar enhacements for organic pollutant removal
flow treatment, fixed-bed columns seem to be the best option
due to several advantages as simplicity and easy scaled up from
4.3.1. Coating biochar
a laboratory-scale (De Gama et al., 2017). Thus, continuous
Recently the use of nanoparticles into biochar production has
fixed-bed columns are selected for large-scale wastewater treat-
been revealed as an interesting approach because the biochar
ment (Zhang et al., 2015b). The knowledge of the breakthrough
matrix and the stabilized carbon nano-materials have good sorp-
curves is fundamental for the design of a continuous flow system
tive properties. Hence, Abdul et al. (2017) obtained large improve-
for adsorption of pollutants, and in general should be determined
ments in the adsorption capacity of phthalic acid esters onto pine
experimentally or modelling based on the kinetic behaviour and
needle wood biochar combined with graphene nanosheets. The
isotherm model (Cobas et al., 2015; Mahmood et al., 2015).
coating treatment transformed the structure and morphology of
Table 3 shows the operational conditions for different fixed-bed
the raw biochar increasing the porosity and surface area as a result
column treatments.
of the modification sp2 and sp3-hybridized carbon atoms that facil-
itate non-covalent linkage with chemicals via hydrophobic interac-
tions, p-p EDA interactions, or non-radiative dipole–dipole 5.1. Operational management
coupling. Zhang et al. (2016) synthetized carbonaceous nanocom-
posites by dip-coating straw biomass in carboxyl functionalized In fixed-bed columns, several general considerations should be
multi-walled carbon nanotubes solution and then pyrolyzed at taken in account in order to avoid operational problems, which are
300 °C and 600 °C in the absence of air. Operating in this way common for the adsorption of different kind of pollutants. Liquid
was possible to improve significantly the surface areas and pore flows at a relatively low velocity through the spaces between the
volume that showed an excellent sulfamethazine sorption. As it particles and, as result of the continuous deposition of these parti-
has been mentioned before, magnetization of biochars has been cles, channelling and/or preferential paths and dead zone appear
disclosed as an interesting alternative to reduce the recovery cost increasing the pressure drop across beds and the resistance to flow
of spent adsorbent. However, several researches detected lower progressively (Sanromán et al., 1994).
sorption of magnetized carbonaceous materials compared to the While one focus of this study is to evaluate the recent studies
non-magnetic carbonaceous material. This fact was due to lower addressed on the suitability of biochar as low-cost adsorbent for
surface area of magnetic composite and the low affinity of the pol- removing organic and inorganic pollutants from water, the main
lutant for the magnetics particles of Fe3O4 (Reguyal et al., 2017). goal is to propose and evaluate strategies to operate at full-scale
in a continuous flow fixed-bed or fluidized-bed column. For this
4.3.2. Physical and chemical pretreatment reason, several operational variables such as particle size, density,
The presence of undesirable residues on the biochar, after the porosity and tortuosity should be considered to guarantee its good
pyrolysis process, can be reduced/eliminated by chemical treat- working order (Russo et al., 2016).
ments. In addition, several researches have evidenced that chemi-
cal treatments can modify the adsorption capacity due to the 5.1.1. Particle size
introduction/removal of functional groups on the carbon surface. One of the main problems to the application of biochar in
Sun et al. (2016) studied the variation in sorption of propiconazole adsorption column is its particle size. Normally, the particle sizes
with biochars after de-ashed treatment with HCl and HF. They of the biochar are highly dependent upon the nature of the raw
reported that the removal of minerals from biochar elevated pollu- material used and sieving before or after thermal treatment.
tant sorption because minerals may exert certain influence on Thus, particle sizes of resultant biochar from organic matter
sorption via affecting spatial arrangement of polar groups and/or feedstock are lower than initial matter due to the shrinkage
organic matter–mineral interactions. In the same way, Peng et al. and attrition during pyrolysis process. In a recent study, physical
(2016) demonstrated that the adsorption capacities for penta- and chemical characterization of six different wood-biochars was
clorophenol on acid-washed reed biochars were higher than that performed. The particle size distribution curves showed that the
of original reed biochars. On the other hand, Chen et al. (2017) percentage of fine particles (<0.075 mm) and grain sizes varies
compared peanut shell-derived biochars treated by different acids considerably among commercially available biochars. However,
founding a decrease in sorption of carbamazepine on acid-washed their average dry bulk density values were less than 1 g/cm3 that
biochars due to the removal of inorganic minerals. Ghaffar et al. reflect the high internal porosity of these materials (Yargicoglu
(2015) studied the structural composition and morphology of the et al., 2015).
peanut shells biochar, before and after oxidation with HNO3/ It is well known that the maximum bed capacities and break-
H2SO4. Despite surface area reduction and pore destruction, they through time is increased when the size of adsorbent particles is
observed an increase on di-alkyl phthalates sorption on oxidized reduced. This fact could be explained due to the finer particles have
biochar surfaces portrayed existence of strong pollutants binding a shorter diffusion path, that facilities the deeper penetration of
sites due to that oxidation introduces polar functional groups (car- adsorbate into the adsorbent particle, increasing the adsorption
boxyl and phenolic groups) on biochar surfaces. rate. In addition, the total external surface area per unit volume
E. Rosales et al. / Bioresource Technology 246 (2017) 176–192 187

for smaller particles inside the column is larger. Thus, the adsorp-

Rojas-Mayorga et al. (2015)


tion breakthrough curves for smaller particles follow a much more

Mayakaduwa et al. (2016)


efficient profile than larger particles, in that the breakthrough time

Shanthakumar (2017)
increased and the curves tended towards the classic shape profile

Mondal et al. (2016)


Zhang et al. (2015a)
Kizito et al. (2016)
Ding et al. (2016)

Jung et al. (2017)


(Malkoc and Nuhoglu, 2006). However, from operational point of

Vilvanathan and

Hu et al. (2015)
view, the use of finer particle sizes reduces the hydraulic conduc-
tivity and increases the high flow resistance of the adsorption col-
Reference

umn that should be avoided (Inglezakis and Grigoropoulou, 2004).


Experiments reported in the literature concluded the depen-
dency of operational conditions on particle size distribution. It is
clear in the results presented in Fig. 3 a in which the change in
the size distributions of sawdust feedstock and produced biochar
Saturation

Others 0.3
Pb,Cu 2.3
time (h)

at two temperatures 450 and 750 °C are showed. This fact could
66.3
10.5
135

149
120

be explained by the decreasing tensile strength of the material


n.a.

n.a.

n.a.

n.a.
4.8

resulting in less resistance to attrition during the biochar


25, 50, 75

processing.
500, 540,

100, 150,
100 each

5, 10, 15

10–100
(mg/L)

18.29
In several cases, the phenomena of the particles aggregation
Conc.

585

100

200

50 take place by changing the size distribution and reducing the


external surface area per unit volume, and as such addressing this
0.89–19.1 53.6 (Pb)

issue is critical to assess the viability of biochar in full-scale sys-


Ni 26.99 Co 23.63
capacity (mg/g)

tems. Several studies reported the advantages of using composite


biochar to avoid biochar’ limitations. Roh et al. (2015) reported
Adsorption

the good adsorption capacity of buffalo weed biochar-alginate


160.77
114.7

35.92

515.1

12.38
104.8

27.2

1.49

18.5

2.36

beads for treatment of cadmium, 2,4,6-trinitrotoluene and 1,3,5-tri


nitro-1,3,5-triazacyclohexane in continuous fixed-bed column
studies. They concluded that the use of biochar-alginate is able
0.7, 1.4, 2.1

to overcome the difficulties like aggregation, swelling, crumbling


Adsorbent

0.5, 1, 1.5
mass (g)

and lack of regeneration efficiency of powder biochar adsorbent


0.529
1.042

10.35
0.407

0.33

materials in large-scale wastewater treatment.


n.a.
1

Furthermore, several studies have been focused on modified


biochar as a consequence of the release of biochar powder into
Flow (mL/

15, 20, 25

2.5, 5, 7.5
1, 2.5, 5

treated effluent. The inclusion of magnetic medium (e.g., mag-


2, 4, 6
min)

netite, c-Fe2O3) by chemical co-precipitation improves the biosor-


3, 6
2

bent affinity and would be a useful method to enable the small


particle powders of biochar to be effectively separated from solu-
Breakthrough

BDST, THM,

BDST, THM,

BDST, THM,

BDST, THM,

BDST, THM,
(Langmuir)

tion with application of external magnetic fields (Chen et al., 2011).


THM, YAN

In several cases, hydrogels are used as dispersing and soldering


model
None

YNM

YNM

YNM

YNM

YNM
THM

agent to attach fine magnetic materials as zero valent iron (ZVI).


Thus, Zhou et al. (2014) synthetized ZVI-biochar composites con-


firmed that chitosan effectively soldered the iron particles onto
porosity

carbonaceous surfaces within the biochar pore networks.


45%

48%
42%

30%
Bed

n.a.

n.a.

n.a.

n.a.

n.a.

n.a.

n.a.

5.1.2. Porosity
3.5, 4.8, 5.8
Bed height

10, 15, 20

The diffusion in porous solids is usually the most important


1, 2, 3

1, 2, 3

controlling factor of mass transfer in adsorption and the tortuosity


(cm)
1.55

n.a.

7.5

1.2

expresses this hindrance. In thermal decomposition process, the


60

pores are formed due to water loss in dehydration process and


the volatiles release to the carbon matrix that permit to obtain dif-
Superheated steam
Alkali modification

ferent size pores encompasses from nanopores (internal diame-


Steam activation

Fe impregnated
Immobilization
alginate beads

ter < 2 nm) to macropores (internal diameter > 50 nm). Thus,


Modification

Dopping Al

biochar porosity varies significantly with pyrolysis temperature


activated

and, usually, larger pore size and surface area are achieved when
None

None

None

the temperature is increased. As example, when the temperature


Fixed-bed biochar column adsorption assays.

of pine needle pyrolysis was increased from 400 to 700 °C, porosity
of biochar increased from 0.044 to 0.19 cm3/g (Li et al., 2017a).
Trichloroethylene
Pb2+, Cu2+ Cd2+,

Furthermore, the biomass feedstock composition is another sig-


hydrochloride
Ammonium

Carbofuran

nificant variable on biochar porosity, obtaining predominantly


Phosphate
Pollutants

Zn2+, Ni2+

Co2+, Ni2+

Ranitide
Fluoride

microporous or macroporous structure biochar when the biomass


is rich in cellulose or lignin, respectively. The results showed in
As

Fig. 3b, clearly demonstrate the effect of these variables and high-
Mung bean husk

light the difference among the different feedstock used (Li et al.,
Soybean stover
Tectona grandis

n.a. not available.


doped bone
Hickory chips

Hickory chips

2017a). Similarly, Zhang et al. (2017b) determined that biochars


Aluminum-
japonica
Hardwood

produced from corn straw, rice straw and wheat straw at 400 °C
Laminaria

Rice husk
BC300
Corncobs

leaves
Sawdust
Biochar

BC700

presented different distribution of surface area and pore size. The


Table 3

abundance of nanopores of the three biochars significantly


increased as the cellulose content of the biomass increased.
188 E. Rosales et al. / Bioresource Technology 246 (2017) 176–192

a) 70

Sawdust feedstock
60 450ºC
750ºC

Particle size distribution (%)


50

40

30

20

10

0
<= 0.6 0.6-1.18 1.18-2.36 2.36-4.75 >4.75

Particle size (mm)

b) 0.20

300
450
600
400
0.15 500
700
900
Porosity (cm /g)
3

0.10

0.05

0.00
Maize straw Pine needle Municipal biosolids Swine manure

Biomass feedstock

Fig. 3. Effect of pyrolysis temperature on: a) particle size distribution of sawdust feedstock (based on data from Downie et al., 2009) and b) biochar’ pore size distribution
(based on data from Li et al., 2017a).

As was mentioned above, efforts have been made to prepare method enhances the porosity of the biochar surface but also pro-
biochar based composite with improved properties to enhance vides the nano-sized MgO particles dispersion, which enabled
adsorption capacity by combining biochar with organic and/or higher phosphate adsorption capability.
inorganic materials through physicochemical methods. In this
framework, the porosity biochar matrix was enhanced by com- 5.2. Regeneration
bined electrochemical modification in which an electrolyte would
generate strong oxidants in solution by electrochemical reaction. The suitability of the designed systems for their regeneration
This method was applied by Jung and Ahn (2016) who reported and reuse in different cycles is also required. Few studies of biochar
the immersion of dried marine macroalgae in MgCl2 solution under regeneration at column scale have been performed. The regenera-
20 V voltage for 10 min prior pyrolysis. The physical and chemical tion of the adsorbents can minimize the operational cost.
characterization of the resultant biochar proved that is a potent The desorption of the pollutants from the columns requires the
additional technique to prepare effectively modified-biochar. This utilization of different eluents related to the compound desired to
E. Rosales et al. / Bioresource Technology 246 (2017) 176–192 189

remove. The use of a strong acid, HCl, as eluting agent has been 5.3.2. Effect of flow rate
reported in the regeneration of biochars containing metals. This Flow rate is another significant parameter with strong influence
eluent was applied to the removal of adsorbed metals Ni and Co on continuous treatment system. An inverse relationship between
on biochar from Tectona grandis leaves (Vilvanathan and flow rate and adsorption capacity has been reported on adsorption
Shanthakumar, 2017), and the removal of Pb from alkali- of different pollutants such as metals, phosphates and carbofuran
modified hickory biochar (Ding et al., 2016). The systems demon- by biochar derived from several biomass feedstocks (Jung et al.,
strated their feasibility during at least four adsorption-desorption 2017; Kizito et al., 2016; Mondal et al., 2016). This fact may be
cycles with a decrease in the efficiency due to the mass loss during explained as a result of scarce contact time on biochar bed which
these cycles. The use of this eluent may suggest an ionic exchange is not enough to attain the equilibrium, leading to a sharper slope
as main mechanism for metal release in the regeneration. How- of breakthrough curve. On the contrary, low flows favour a higher
ever, another eluent NaHCO3 was proposed by Hu et al. (2015) to contact time becoming the diffusion into the pores more effective.
remove As from Fe-impregnated biochar reaching a removal per-
centage around 85%. 5.3.3. Effect of inlet concentration
In the column desorption of an organic compound as TCE, Zhang The design of an efficient operational column requires the study
et al. (2015b) proposed an elution with 25% of methanol for bio- of the inlet concentration. The diffusion usually depends on the
char regeneration. The desorption was performed at a low flow concentration and a high concentration difference provides a high
rate observing initially a sharp increase of the desorption where driving force for the adsorption process. Hence, the available bind-
upon decreased over time. This fact may also suggest multiple ing points are sharply reduced at high pollutant concentration,
adsorption mechanism involved. leading to the decrease of the breakthrough time and biochar sat-
uration. On the contrary, the biochar adsorption capacity achieved
the highest values when the inlet flow was the maximum (Jung
5.3. Modelling fixed bed adsorption
et al., 2017; Kizito et al., 2016; Mondal et al., 2016). This fact
may be explained by the slow transport and lower diffusion value
The design of an appropriate column treatment system requires
or mass transport coefficient.
a good prediction of the effluent behaviour and, therefore, the
breakthrough curves. Breakthrough curves are one of the most
5.3.4. Modelling of the column data
characteristic parameters in a continuous treatment system. The
Several models have been considered to accurately describe the
scale up of the adsorption requires the definition of several param-
experimental data and predict adsorption behaviour on biochar-
eters related to the column design. Among them, bed height, flow
fixed beds. The dynamic behaviour has been studied by empirical
rate and inlet concentration are factors with a significant effect in
or semi-empirical models due to their lower mathematical com-
the adsorption process and consequently the design of new treat-
plexity. Three experimental models (Thomas, Adam’s-Bohart and
ment columns.
Yoon-Nelson) described in Table 4 are commonly used for the anal-
ysis of the experimental data to predict the performance of the
5.3.1. Influence of bed height column.
Bed height represents an important effect on the dynamic per-
formance of the column and, therefore, breakthrough time and 5.3.4.1. Thomas model (THM). This model, widely used for adsorp-
adsorbent bed performance being a determining factor in the bio- tion process without external/internal diffusion limitations,
char capacity for the removal of the interest compounds. assumes that the experimental data follow Langmuir adsorption
Several researches have demonstrated a relationship between isotherm and no axial dispersion is considered based on the
the bed height and the breakthrough time with adsorption ability. assumption that the rate driving force obeys second-order reversi-
This fact was reported by Vilvanathan and Shanthakumar (2017) ble reaction kinetics.
on metal adsorption, Mondal et al. (2016) on carbofuran removal This model has been successfully used to fit the biochar column
and Zhang et al. (2015b) on TCE removal. When the column height adsorption for metals by Tectona grandis biochar (Vilvanathan and
is enlarged, more adsorbent mass is added to the system and Shanthakumar, 2017), ammonium by corncob, hardwood and saw-
results on (i) providing more linking points to the biochar and also dust biochars (Kizito et al., 2016), for carbofuran by rice husk bio-
(ii) enhancing the adsorption capacity. Meanwhile, with the con- char (Mayakaduwa et al., 2016) or TCE by soybean stover biochar
comitant increased adsorption capacity, the residence time also (Zhang et al., 2015b). In all these reported studies, the increase of
increment. Thus, the bed height is directly related to the residence inlet flow originated a rise in the adsorption rate constant, and a
time and consequently, this implies higher diffusion through the decrease in the adsorption capacity of the bed due to the faster
biochar. exhaustion of the biochars. However, this rate constant decreased

Table 4
Empirical and semiempirical breakthrough models.

Model Expression Model parameters


Thomas model (THM) ln(C0/Ct  1) = KThq0W/Q  KThC0t KTh, rate constant (mL/minmg) Ct, outlet concentration at time t (mg/L)
q0, equilibrium uptake (mg/g) Q, flow rate (mL/min)
W, mass of biosorbent (g) t, time (min)
C0, inlet concentration (mg/L)
Adam’s–Bohart model (BDST) ln(Ct/C0) = KABC0t  KABN0(Z/U0) KAB, rate constant (L/mgmin) t, time (min)
C0, inlet concentration (mg/L) N0, sorption capacity of bed (mg/L)
Ct, outlet concentration (mg/L) Z, bed height of the column (cm)
U, linear velocity (cm/min)
Yoon–Nelson model (YNM) ln(Ct/(C0  Ct) = KYNt-sKYN KYN, rate constant (L/min) s, time required
for 50% adsorbate breakthrough (min) t,
breakthrough time (min)
190 E. Rosales et al. / Bioresource Technology 246 (2017) 176–192

and the adsorption capacity of the bed was improved when the ini- Acknowledgements
tial concentration or the bed height was enlarged allowing more
contact time and amount of biochar to remove the pollutants. The authors would like to thank to the Spanish Ministry of
The good fitting to this model also indicates that the main driving Economy and Competitiveness for the financial support of Emilio
force related to the adsorption is the concentration gradient of the Rosales under Juan de la Cierva postdoctoral grant.
pollutant. This model failed to evaluate the breakthrough curves
when the adsorption process implies a significant axial dispersion References
of the pollutant inside the column as reported by Rojas-Mayorga
et al. (2015) in the adsorption of fluoride or drugs as ranitidine Abdelhafez, A.A., Li, J., 2016. Removal of Pb(II) from aqueous solution by using
biochars derived from sugar cane bagasse and orange peel. J. Taiwan Inst. Chem.
hypochloride by mung bean husk biochar (Mondal et al., 2016).
Eng. 61, 367–375.
As it was previously stated, the fitting to this model of the data Abdolali, A., Guo, W.S., Ngo, H.H., Chen, S.S., Nguyen, N.C., Tung, K.L., 2014. Typical
suggested that external and internal diffusions are not the only lignocellulosic wastes and by-products for biosorption process in water and
wastewater treatment: a critical review. Bioresour. Technol. 160, 57–66.
rate limiting steps in the adsorption (Vilvanathan and
Abdul, G., Zhu, X., Chen, B., 2017. Structural characteristics of biochar-graphene
Shanthakumar, 2017). nanosheet composites and their adsorption performance for phthalic acid
esters. Chem. Eng. J. 319, 9–20.
Ahmad, M., Rajapaksha, A.U., Lim, J.E., Zhang, M., Bolan, N., Mohan, D., Vithanage,
5.3.4.2. Adam’s-Bohart model (BDST). This model assumes that the M., Lee, S.S., Ok, Y.S., 2014. Biochar as a sorbent for contaminant management in
equilibrium is not reached instantaneously, and the adsorption soil and water: a review. Chemosphere 99, 19–23.
Ahmed, M.B., Zhou, J.L., Ngo, H.H., Guo, W., Johir, M.A.H., Belhaj, D., 2017.
rate is proportional to both the residual capacity of the biochar
Competitive sorption affinity of sulfonamides and chloramphenicol antibiotics
and the concentration of the adsorbate species. The adsorption rate toward functionalized biochar for water and wastewater treatment. Bioresour.
of the pollutants in the biochar enhances by increasing the flow Technol. 238, 306–312.
rate; however, reduce the bed adsorption capacity, and critical Aller, M.F., 2016. Biochar properties: transport, fate, and impact. Crit. Rev. Environ.
Sci. Technol. 46, 1183–1296.
bed height (Jung et al., 2017; Kizito et al., 2016). Kizito et al. Boutsika, L.G., Karapanagioti, H.K., Manariotis, I.D., 2014. Aqueous mercury sorption
(2016) showed an increment on the bed height and flow increases by biochar from malt spent rootlets. Water Air Soil Pollut. 225 (1), 1–10.
the residence time and the diffusion process in the column. Fur- Chen, B., Chen, Z., Lv, S., 2011. A novel magnetic biochar efficiently sorbs organic
pollutants and phosphate. Bioresour. Technol. 102, 716–723.
thermore, the use of high concentrations and flow rates influence Chen, T., Zhou, Z., Han, R., Meng, R., Wang, H., Lu, W., 2015a. Adsorption of cadmium
the system kinetics and affects to the column behaviour. It has by biochar derived from municipal sewage sludge: impact factors and
been reported that if the adsorption rate increases by increasing adsorption mechanism. Chemosphere 134, 286–293.
Chen, C., Zhou, W., Lin, D., 2015b. Sorption characteristics of N-
the flow rate, the external mass transfer dominates the overall sys- nitrosodimethylamine onto biochar from aqueous solution. Bioresour.
tem kinetics particularly in the first steps of the fixed-bed adsorp- Technol. 179, 359–366.
tion. This fact has been stated for different pollutants such as Chen, J., Zhang, D., Zhang, H., Ghosh, S., Pan, B., 2017. Fast and slow adsorption of
carbamazepine on biochar as affected by carbon structure and mineral
metals (Vilvanathan and Shanthakumar, 2017), pharmaceuticals composition. Sci. Total Environ. 579, 598–605.
(Mondal et al., 2016), phosphates (Jung et al., 2017) or ammonium Cobas, M., Danko, A.S., Pazos, M., Sanromán, M.A., 2015. Removal of metal and
(Kizito et al., 2016). organic pollutants from wastewater by a sequential selective technique.
Bioresour. Technol. 213, 2–10.
Crini, G., 2006. Non-conventional low-cost adsorbents for dye removal: a review.
5.3.4.3. Yoon-Nelson model (YNM). The basis of this method is that Bioresour. Technol. 97, 1061–1085.
de Gama, M.S., Luna, F.M.T., Albarelli, J.Q., Beppu, M.M., Vieira, R.S., 2017.
the rate of decrease in the probability of adsorption for each adsor- Adsorption of copper on glass beads coated with chitosan: Stirred batch and
bate molecule is proportional to the probability of adsorbate fixed bed analysis. Can. J. Chem. Eng. 95, 1164–1170.
adsorption and the probability of adsorbate breakthrough on the Devi, P., Saroha, A.K., 2014. Synthesis of the magnetic biochar composites for use as
an adsorbent for the removal of pentachlorophenol from the effluent. Bioresour.
adsorbent. This model determines the rate constant of the column
Technol. 169, 525–531.
adsorption and the time required for reaching 50% of adsorbate Devi, P., Saroha, A.K., 2015. Effect of pyrolysis temperature on polycyclic aromatic
breakthrough. hydrocarbons toxicity and sorption behaviour of biochars prepared by pyrolysis
The behaviour of the modeled systems in the literature is no of paper mill effluent treatment plant sludge. Bioresour. Technol. 192, 312–320.
Ding, Z., Hu, X., Wan, Y., Wang, S., Gao, B., 2016. Removal of lead, copper, cadmium,
dependent of the considered pollutant. The higher value of the zinc, and nickel from aqueous solutions by alkali-modified biochar: batch and
inlet flow and initial concentration causes a decrease in the resi- column tests. J. Ind. Eng. Chem. 33, 239–245.
dence time due to a faster saturation of the column and an increase Downie, A., Crosky, A., Munroe, P., 2009. Characteristics of biochar-physical
properties of biochar. In: Joseph, S., Lehmann, J. (Eds.), Biochar for
in the adsorption rate (Vilvanathan and Shanthakumar, 2017; Jung Enviromental Management: Science and Technology. Earthscan, London, pp.
et al., 2017; Kizito et al., 2016; Mondal et al., 2016). One of the 13-29.
main drawbacks of this model is that no detailed data concerning EPA, 2017. Drinking water contaminants-Standards and regulations, 2017, https://
www.epa.gov/dwstandardsregulations, accessed 05/28/2017.
to the physical properties of the column bed or about pollutants Ferreira, C.I.A., Calisto, V., Otero, M., Nadais, H., Esteves, V.I., 2017. Removal of
and biochar characteristics are considered. Moreover, the adsorp- tricaine methanesulfonate from aquaculture wastewater by adsorption onto
tion process is considered as a constant and linear pattern. pyrolysed paper mill sludge. Chemosphere 168, 139–146.
Frišták, V., Pipíška, M., Lesný, J., Soja, G., Friesl-Hanl, W., Packová, A., 2015.
Utilization of biochar sorbents for Cd2+, Zn2+, and Cu2+ ions separation from
6. Conclusions aqueous solutions: comparative study. Environ. Monit. Assess. 187, 4093.
Gao, F., Xue, Y., Deng, P., Cheng, X., Yang, K., 2015. Removal of aqueous ammonium
by biochars derived from agricultural residuals at different pyrolysis
Based on the extensive literature reviewed, it is concluded that temperatures. Chem. Spec. Bioavailability 27, 92–97.
biochar can be produced from a huge variety of feedstocks at dif- Ghaffar, A., Ghosh, S., Li, F., Dong, X., Zhang, D., Wu, M., Li, H., Pan, B., 2015. Effect of
biochar aging on surface characteristics and adsorption behavior of dialkyl
ferent thermal conditions and its properties could be improved phthalates. Environ. Pollut. 206, 502–509.
by control of these variables and using strategies that provided Goswami, R., Shim, J., Deka, S., Kumari, D., Kataki, R., Kumar, M., 2016.
functionality and permit the synthesis of biochars ‘‘on demand”. Characterization of cadmium removal from aqueous solution by biochar
produced from Ipomoea fistulosa at different pyrolytic temperatures. Ecol.
Until date, despite an extensive literature related to their physical/
Eng. 97, 444–451.
chemical properties and adsorption kinetics and equilibrium stud- Hadjittofi, L., Prodromou, M., Pashalidis, I., 2014. Activated biochar derived from
ies for the pollutants removal, there is a lack of data concerning the cactus fibres – preparation, characterization and application on Cu(II) removal
operation in continuous flow systems and their regeneration. To from aqueous solutions. Bioresour. Technol. 159, 460–464.
Han, L., Sun, K., Jin, J., Wei, X., Xia, X., Wu, F., Gao, B., Xing, B., 2014. Role of structure
close these knowledge gaps, more studies are needed in further and microporosity in phenanthrene sorption by natural and engineered organic
investigations. matter. Environ. Sci. Technol. 48, 11227–11234.
E. Rosales et al. / Bioresource Technology 246 (2017) 176–192 191

Hou, J., Huang, L., Yang, Z., Zhao, Y., Deng, C., Chen, Y., Li, X., 2016. Adsorption of Meng, J., Feng, X., Dai, Z., Liu, X., Wu, J., Xu, J., 2014. Adsorption characteristics of Cu
ammonium on biochar prepared from giant reed. Environ. Sci. Pollut. Res. 23, (II) from aqueous solution onto biochar derived from swine manure. Environ.
19107–19115. Sci. Pollut. Res. 21, 7035–7046.
Hu, X., Ding, Z., Zimmerman, A.R., Wang, S., Gao, B., 2015. Batch and column Micháleková-Richveisová, B., Frišták, V., Pipíška, M., Duriška, L., Moreno-Jimenez, E.,
sorption of arsenic onto iron-impregnated biochar synthesized through Soja, G., 2017. Iron-impregnated biochars as effective phosphate sorption
hydrolysis. Water Res. 68, 206–216. materials. Environ. Sci. Pollut. Res. 24, 463–475.
Inglezakis, V.J., Grigoropoulou, H., 2004. Effects of operating conditions on the Mohan, D., Sarswat, A., Ok, Y.S., Pittman, C.U., 2014. Organic and inorganic
removal of heavy metals by zeolite in fixed bed reactors. J. Hazard. Mater. 112, contaminants removal from water with biochar, a renewable, low cost and
37–43. sustainable adsorbent - a critical review. Bioresour. Technol. 160, 191–202.
Inyang, M., Dickenson, E., 2015. The potential role of biochar in the removal of Mondal, S., Aikat, K., Halder, G., 2016. Ranitidine hydrochloride sorption onto
organic and microbial contaminants from potable and reuse water: a review. superheated steam activated biochar derived from mung bean husk in fixed bed
Chemosphere 134, 232–240. column. J. Environ. Chem. Eng. 4, 488–497.
Inyang, M., Gao, B., Zimmerman, A., Zhang, M., Chen, H., 2014. Synthesis, Park, S.H., Cho, H.J., Ryu, C., Park, Y., 2016. Removal of copper(II) in aqueous solution
characterization, and dye sorption ability of carbon nanotube-biochar using pyrolytic biochars derived from red macroalga Porphyra tenera. J. Ind.
nanocomposites. Chem. Eng. J. 236, 39–46. Eng. Chem. 36, 314–319.
Inyang, M., Gao, B., Zimmerman, A., Zhou, Y., Cao, X., 2015. Sorption and cosorption Peng, P., Lang, Y., Wang, X., 2016. Adsorption behavior and mechanism of
of lead and sulfapyridine on carbon nanotube-modified biochars. Environ. Sci. pentachlorophenol on reed biochars: PH effect, pyrolysis temperature,
Pollut. Res. 22, 1868–1876. hydrochloric acid treatment and isotherms. Ecol. Eng. 90, 225–233.
Jin, H., Hanif, M.U., Capareda, S., Chang, Z., Huang, H., Ai, Y., 2016. Copper(II) removal Reguyal, F., Sarmah, A.K., Gao, W., 2017. Synthesis of magnetic biochar from pine
potential from aqueous solution by pyrolysis biochar derived from sawdust via oxidative hydrolysis of FeCl2 for the removal sulfamethoxazole
anaerobically digested algae-dairy-manure and effect of KOH activation. J. from aqueous solution. J. Hazard. Mater. 321, 868–878.
Environ. Chem. Eng. 4, 365–372. Roh, H., Yu, M., Yakkala, K., Koduru, J.R., Yang, J., Chang, Y., 2015. Removal studies of
Jin, J., Sun, K., Wu, F., Gao, B., Wang, Z., Kang, M., Bai, Y., Zhao, Y., Liu, X., Xing, B., Cd(II) and explosive compounds using buffalo weed biochar-alginate beads. J.
2014. Single-solute and bi-solute sorption of phenanthrene and dibutyl Ind. Eng. Chem. 26, 226–233.
phthalate by plant- and manure-derived biochars. Sci. Total Environ. 473– Rojas-Mayorga, C.K., Bonilla-Petriciolet, A., Sánchez-Ruiz, F.J., Moreno-Pérez, J.,
474, 308–316. Reynel-Ávila, H.E., Aguayo-Villarreal, I.A., Mendoza-Castillo, D.I., 2015.
Jung, C., Boateng, L.K., Flora, J.R.V., Oh, J., Braswell, M.C., Son, A., Yoon, Y., 2015a. Breakthrough curve modeling of liquid-phase adsorption of fluoride ions on
Competitive adsorption of selected non-steroidal anti-inflammatory drugs on aluminum-doped bone char using micro-columns: effectiveness of data fitting
activated biochars: Experimental and molecular modeling study. Chem. Eng. J. approaches. J. Mol. Liq. 208, 114–121.
264, 1–9. Rosales, E., Ferreira, L., Sanromán, M.Á., Tavares, T., Pazos, M., 2015. Enhanced
Jung, K., Hwang, M., Ahn, K., Ok, Y., 2015b. Kinetic study on phosphate removal from selective metal adsorption on optimised agroforestry waste mixtures.
aqueous solution by biochar derived from peanut shell as renewable adsorptive Bioresour. Technol. 182, 41–49.
media. Int. J. Environ. Sci. Technol. 12, 3363–3372. Russo, V., Masiello, D., Trifuoggi, M., Di Serio, M., Tesser, R., 2016. Design of an
Jung, K., Jeong, T., Choi, B.H., Jeong Kang, H., Ahn, K., 2017. Phosphate adsorption adsorption column for methylene blue abatement over silica: from batch to
from aqueous solution by Laminaria japonica-derived biochar-calcium alginate continuous modeling. Chem. Eng. J. 302, 287–295.
beads in a fixed-bed column: experiments and prediction of breakthrough Sanromán, A., Roca, E., Núñez, M.J., Lema, J.M., 1994. A pulsing device for packed-
curves. Environ. Prog. Sustainable Energy. bed bioreactors: II. Application to alcoholic fermentation. Bioprocess Eng. 10,
Jung, K., Ahn, K., 2016. Fabrication of porosity-enhanced MgO/biochar for removal 75–81.
of phosphate from aqueous solution: application of a novel combined Shan, D., Deng, S., Zhao, T., Wang, B., Wang, Y., Huang, J., Yu, G., Winglee, J., Wiesner,
electrochemical modification method. Bioresour. Technol. 200, 1029–1032. M.R., 2016. Preparation of ultrafine magnetic biochar and activated carbon for
Kearns, J.P., Wellborn, L.S., Summers, R.S., Knappe, D.R.U., 2014. 2,4-D adsorption to pharmaceutical adsorption and subsequent degradation by ball milling. J.
biochars: effect of preparation conditions on equilibrium adsorption capacity Hazard. Mater. 305, 156–163.
and comparison with commercial activated carbon literature data. Water Res. Shang, J., Zong, M., Yu, Y., Kong, X., Du, Q., Liao, Q., 2017. Removal of chromium (VI)
62, 28–29. from water using nanoscale zerovalent iron particles supported on herb-residue
Kim, B., Lee, H.W., Park, S.H., Baek, K., Jeon, J., Cho, H.J., Jung, S., Kim, S.C., Park, Y., biochar. J. Environ. Manage. 197, 331–337.
2016. Removal of Cu2+ by biochars derived from green macroalgae. Environ. Sci. Shimabuku, K.K., Kearns, J.P., Martinez, J.E., Mahoney, R.B., Moreno-Vasquez, L.,
Pollut. Res. 23, 985–994. Summers, R.S., 2016. Biochar sorbents for sulfamethoxazole removal from
Kizito, S., Wu, S., Wandera, S.M., Guo, L., Dong, R., 2016. Evaluation of ammonium surface water, stormwater, and wastewater effluent. Water Res. 96, 236–245.
adsorption in biochar-fixed beds for treatment of anaerobically digested swine Sun, K., Kang, M., Ro, K.S., Libra, J.A., Zhao, Y., Xing, B., 2016. Variation in sorption of
slurry: experimental optimization and modeling. Sci. Total Environ. 563–564, propiconazole with biochars: The effect of temperature, mineral, molecular
1095–1104. structure, and nano-porosity. Chemosphere 142, 56–63.
Lehmann, J., Gaunt, J., Rondon, M., 2006. Bio-char sequestration in terrestrial Tan, G., Sun, W., Xu, Y., Wang, H., Xu, N., 2016. Sorption of mercury (II) and atrazine
ecosystems – a review. Mitig. Adapt. Strategies Global Change 11, 403–427. by biochar, modified biochars and biochar based activated carbon in aqueous
Li, H., Dong, X., da Silva, E.B., de Oliveira, L.M., Chen, Y., Ma, L.Q., 2017a. Mechanisms solution. Bioresour. Technol. 211, 727–735.
of metal sorption by biochars: Biochar characteristics and modifications. Trakal, L., Veselská, V., Šafařík, I., Vítková, M., Číhalová, S., Komárek, M., 2016. Lead
Chemosphere 178, 466–478. and cadmium sorption mechanisms on magnetically modified biochars.
Li, B., Yang, L., Wang, C., Zhang, Q., Liu, Q., Li, Y., Xiao, R., 2017b. Adsorption of Cd(II) Bioresour. Technol. 203, 318–324.
from aqueous solutions by rape straw biochar derived from different Vilvanathan, S., Shanthakumar, S., 2017. Column adsorption studies on nickel and
modification processes. Chemosphere 175, 332–340. cobalt removal from aqueous solution using native and biochar form of Tectona
Li, Y., Shao, J., Wang, X., Deng, Y., Yang, H., Chen, H., 2014. Characterization of grandis. Environ. Prog. Sustainable Energy. http://dx.doi.org/10.1002/ep.12567.
modified biochars derived from bamboo pyrolysis and their utilization for Vu, T.M., Trinh, V.T., Doan, D.P., Van, H.T., Nguyen, T.V., Vigneswaran, S., Ngo, H.H.,
target component (furfural) adsorption. Energy Fuels 28, 5119–5127. 2017. Removing ammonium from water using modified corncob-biochar. Sci.
Liu, N., Charrua, A.B., Weng, C., Yuan, X., Ding, F., 2015. Characterization of biochars Total Environ. 579, 612–619.
derived from agriculture wastes and their adsorptive removal of atrazine from Wang, Z., Guo, H., Shen, F., Yang, G., Zhang, Y., Zeng, Y., Wang, L., Xiao, H., Deng, S.,
aqueous solution: a comparative study. Bioresour. Technol. 198, 55–62. 2015a. Biochar produced from oak sawdust by Lanthanum (La)-involved
Liu, S., Xu, W., Liu, Y., Tan, X., Zeng, G., Li, X., Liang, J., Zhou, Z., Yan, Z., Cai, X., 2017. pyrolysis for adsorption of ammonium (NH+4), nitrate (NO 3 ), and phosphate
Facile synthesis of Cu(II) impregnated biochar with enhanced adsorption (PO34 ). Chemosphere 119, 646–653.
activity for the removal of doxycycline hydrochloride from water. Sci. Total Wang, S., Gao, B., Li, Y., Mosa, A., Zimmerman, A.R., Ma, L.Q., Harris, W.G., Migliaccio,
Environ. 592, 546–553. K.W., 2015b. Manganese oxide-modified biochars: preparation,
Lyu, H., Tang, J., Huang, Y., Gai, L., Zeng, E.Y., Liber, K., Gong, Y., 2017. Removal of characterization, and sorption of arsenate and lead. Bioresour. Technol. 181,
hexavalent chromium from aqueous solutions by a novel biochar supported 13–17.
nanoscale iron sulfide composite. Chem. Eng. J. 322, 516–524. Wang, Z., Liu, G., Zheng, H., Li, F., Ngo, H.H., Guo, W., Liu, C., Chen, L., Xing, B., 2015c.
Mahmood, S., Khalid, A., Mahmood, T., Arshad, M., Loyola-Licea, J.C., Crowley, D.E., Investigating the mechanisms of biochar’s removal of lead from solution.
2015. Biotreatment of simulated tannery wastewater containing reactive Black Bioresour. Technol. 177, 308–317.
5, aniline and CrVI using a biochar packed bioreactor. RSC Adv. 5, 106272– Wang, H., Gao, B., Wang, S., Fang, J., Xue, Y., Yang, K., 2015d. Removal of Pb(II), Cu(II),
106279. and Cd(II) from aqueous solutions by biochar derived from KMnO4 treated
Malkoc, E., Nuhoglu, Y., 2006. Removal of Ni(II) ions from aqueous solutions using hickory wood. Bioresour. Technol. 197, 356–362.
waste of tea factory: Adsorption on a fixed-bed column. J. Hazard. Mater. 135, Wang, Y., Lu, H., Liu, Y., Yang, S., 2016a. Removal of phosphate from aqueous
328–336. solution by SiO2-biochar nanocomposites prepared by pyrolysis of vermiculite
Manariotis, I.D., Fotopoulou, K.N., Karapanagioti, H.K., 2015. Preparation and treated algal biomass. RSC Adv. 6, 83534–83546.
characterization of biochar sorbents produced from malt spent rootlets. Ind. Wang, S., Gao, B., Li, Y., Zimmerman, A.R., Cao, X., 2016b. Sorption of arsenic onto Ni/
Eng. Chem. Res. 54, 9577–9584. Fe layered double hydroxide (LDH)-biochar composites. RSC Adv. 6, 17792–
Mayakaduwa, S.S., Herath, I., Ok, Y.S., Mohan, D., Vithanage, M., 2016. Insights into 17799.
aqueous carbofuran removal by modified and non-modified rice husk biochars. Wang, Z., Han, L., Sun, K., Jin, J., Ro, K.S., Libra, J.A., Liu, X., Xing, B., 2016c. Sorption of
Environ. Sci. Pollut. Res., 1–9 four hydrophobic organic contaminants by biochars derived from maize straw,
192 E. Rosales et al. / Bioresource Technology 246 (2017) 176–192

wood dust and swine manure at different pyrolytic temperatures. Chemosphere residue biochars pyrolyzed at contrasting temperatures: continuous fixed-bed
144, 285–291. experiments. J. Chem. 2015 (647072), 1–6.
Xie, M., Chen, W., Xu, Z., Zheng, S., Zhu, D., 2014. Adsorption of sulfonamides to Zhang, C., Lai, C., Zeng, G., Huang, D., Yang, C., Wang, Y., Zhou, Y., Cheng, M., 2016.
demineralized pine wood biochars prepared under different thermochemical Efficacy of carbonaceous nanocomposites for sorbing ionizable antibiotic
conditions. Environ. Pollut. 186, 187–194. sulfamethazine from aqueous solution. Water Res. 95, 103–112.
Xue, L., Gao, B., Wan, Y., Fang, J., Wang, S., Li, Y., Muñoz-Carpena, R., Yang, L., 2016. Zhang, J., Wang, Q., 2016. Sustainable mechanisms of biochar derived from brewers’
High efficiency and selectivity of MgFe-LDH modified wheat-straw biochar in spent grain and sewage sludge for ammonia–nitrogen capture. J. Cleaner Prod.
the removal of nitrate from aqueous solutions. J. Taiwan Inst. Chem. Eng. 63, 112, 3927–3934. Part 5.
312–317. Zhang, T., Zhu, X., Shi, L., Li, J., Li, S., Lü, J., Li, Y., 2017a. Efficient removal of lead from
Yang, J., Pan, B., Li, H., Liao, S., Zhang, D., Wu, M., Xing, B., 2016. Degradation of p- solution by celery-derived biochars rich in alkaline minerals. Bioresour.
nitrophenol on biochars: role of persistent free radicals. Environ. Sci. Technol. Technol. 235, 185–192.
50, 694–700. Zhang, F., Li, Y., Zhang, G., Li, W., Yang, L., 2017b. The importance of nano-porosity in
Yao, Y., Gao, B., Fang, J., Zhang, M., Chen, H., Zhou, Y., Creamer, A.E., Sun, Y., Yang, L., the stalk-derived biochar to the sorption of 17ß-estradiol and retention of it in
2014. Characterization and environmental applications of clay-biochar the greenhouse soil. Environ. Sci. Pollut. Res. 24, 9575–9584.
composites. Chem. Eng. J. 242, 136–143. Zhang, X.N., Mao, G.Y., Jiao, Y.B., Shang, Y., Han, R.P., 2014. Adsorption of anionic dye
Yargicoglu, E.N., Sadasivam, B.Y., Reddy, K.R., Spokas, K., 2015. Physical and on magnesium hydroxide-coated pyrolytic bio-char and reuse by microwave
chemical characterization of waste wood derived biochars. Waste Manage. irradiation. Int. J. Environ. Sci. Technol. 11, 1439–1448.
36, 256–268. Zhou, Y., Gao, B., Zimmerman, A.R., Chen, H., Zhang, M., Cao, X., 2014. Biochar-
Yoder, J., Galinato, S., Granatstein, D., Garcia-Pérez, M., 2011. Economic tradeoff supported zerovalent iron for removal of various contaminants from aqueous
between biochar and bio-oil production via pyrolysis. Biomass Bioenergy 35, solutions. Bioresour. Technol. 152, 538–542.
1851–1862. Zhou, Z., Liu, Y., Liu, S., Liu, H., Zeng, G., Tan, X., Yang, C., Ding, Y., Yan, Z., Cai, X.,
Zhang, M., Liu, Y., Li, T., Xu, W., Zheng, B., Tan, X., Wang, H., Guo, Y., Guo, F., Wang, S., 2017. Sorption performance and mechanisms of arsenic(V) removal by
2015a. Chitosan modification of magnetic biochar produced from Eichhornia magnetic gelatin-modified biochar. Chem. Eng. J. 314, 223–231.
crassipes for enhanced sorption of Cr(VI) from aqueous solution. RSC Adv. 5, Zhu, N., Yan, T., Qiao, J., Cao, H., 2016. Adsorption of arsenic, phosphorus and
46955–46964. chromium by bismuth impregnated biochar: adsorption mechanism and
Zhang, M., Ahmad, M., Al-Wabel, M.I., Vithanage, M., Rajapaksha, A.U., Kim, H.S., Lee, depleted adsorbent utilization. Chemosphere 164, 32–40.
S.S., Ok, Y.S., 2015b. Adsorptive removal of trichloroethylene in water by crop

You might also like