Вы находитесь на странице: 1из 10

Journal of Environmental Chemical Engineering 4 (2016) 3746

Contents lists available at ScienceDirect

Journal of Environmental Chemical Engineering


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

Adsorption and desorption of phosphate on biochars


P.A. Trazzi, J.J. Leahy, M.H.B. Hayes, W. Kwapinski*
Carbolea Research Group, Department of Chemical and Environmental Science, University of Limerick, Limerick, Ireland

A R T I C L E I N F O

A B S T R A C T

Article history:
Received 5 August 2015
Received in revised form 8 October 2015
Accepted 2 November 2015
Available online 5 November 2015

Biochar (BC) is regarded as a potential carbon sequesterer, a soil fertility enhancer, and a preventer of
nutrient leaching. Phosphorus amended biochar could enhance soil fertility. This work investigated the
adsorption and desorption of phosphate from aqueous solution on two different carbonized materials.
Sugar cane bagasse (SC) and Miscanthus  giganteus(M) samples were carbonized at various temperatures (between 300  C and 700  C) for residence times of 20 or 60 min. The largest surface area and the
highest P adsorption at 20  C and pH 7 was obtained for M BC prepared at 700  C and at the longer
residence time, compared to the SC BC made under the same conditions (approximately 15.5 mg g1 and
12.8 mg g1 for 400 mg dm3 phosphate in solution, respectively). Adsorption of P on BCs was
endothermic and increased with process temperature. The amount of desorbed P was proportional to its
adsorption capacity. Two isotherm models (Freundlich and Langmuir) tted the experimental results of
phosphate adsorption onto the BC, and the Langmuir adsorption model described it better.
Thermodynamic parameters are compared in the text with phosphate adsorption on other BCs reported
in the literature. Our data suggest that adding phosphate to BC could provide a better way to apply P to
soil in order to obetain better agronomic performances.
2015 Elsevier Ltd. All rights reserved.

Keywords:
Miscanthus
Sugar cane bagasse
Carbonization
Pyrolysis
Energy

1. Introduction
Phosphorus deciency is a major factor restricting crop yields.
This deciency especially applies for tropical weathered soils,
where the bioavailability of phosphorus (P) has a major impact on
of crop production [1]. Under natural conditions, the weathering of
rocks and the release of elements essential for crop growth is a
slow process, and much of the P applied as fertilizer becomes xed
in forms such as calcium iron, and aluminium phosphates that are
relatively insoluble and unavailable for plant growth needs. In
these situations non-amended soils are capable of supporting only
slow-growing vegetation and crops adapted to low soil phosphate
availability.
Biochar (BC), produced by pyrolysis under limited air supply
and temperatures greater than 300  C, is applied to soil to facilitate
carbon storage, the ltration of percolating soil water, and the
enhancement of crop productivity. Variabilities in the properties
and performances of BC are attributable to various process
parameters related to its formation, such as: temperature, the
residence time and the heating rate during pyrolysis [2], the
feedstock [3], the particle size [4], and the method of pyrolysis [5].
The degree of alteration of the original structures of the biomass,

* Corresponding author.
E-mail address: witold.kwapinski@ul.ie (W. Kwapinski).
http://dx.doi.org/10.1016/j.jece.2015.11.005
2213-3437/ 2015 Elsevier Ltd. All rights reserved.

through microstructural rearrangement, attrition during processing, and the formation of cracks all depend upon the processing
conditions to which they are exposed [6].
Piterina and Hayes [7] have shown that arbuscular mycorrhizal
fungi in associations with BC can dissolve tricalcium phosphate
and would make the locked P available for the plant having a
symbiotic relationship with the fungus. Recognition of the
properties and mechanisms of BC adsorption and desorption of
phosphate applications to soil is very important for its potential
uses in soil fertilizer applications. Xu et al. [8] concluded that BC
application to soil altered P availability by changing the P sorption
and desorption processes. These effects were dependent on soil
acidity, and have important implications for improving soil
productivity. The phosphorus concentration in carbonized materials vary depending on crops and growing conditions; e.g. the BC
content of rice husk is 4.7 mg P g1 [9], that of Miscanthus 
giganteus (M) is 2.5 mg P g1 [10], and of sugar cane bagasse (SC) is
6.1 mg P g1 [11]. Residue P in BC form is readily available for plants
[12].
The conditions under which BC is produced can inuence its
effects on P availability. As the temperature of pyrolysis increases,
the specic surface area increases, and the volatile matter and
surface functional groups decrease [1315]. All those changes can
inuence P bioavailability. There is evidence to indicate that BC can
decrease P-xation in soil resulting in greater bioavailability of
added inorganic phosphate [5,16,17]. Cui et al. [16] observed that

38

P.A. Trazzi et al. / Journal of Environmental Chemical Engineering 4 (2016) 3746

the presence of BC decreased P adsorption on the Fe-oxides and


thereby enhance P bioavailability. Parvage et al. [17] concluded that
BC from wheat straw can act as a source of soluble P, and low and
high additions of BC can have different effects on soil solution P
concentration. Morales et al. [5] found that BCs, depending on how
produced, can have very different P sorption and desorption
properties: a BC produced by fast pyrolysis decreased the P-xing
capacity of a degraded tropical soil, whereas that formed in slow
pyrolysis had the opposite effect. Xu et al. [8] found that the BC
application altered the availability of P by changing the capacities
of the soils to sorb and desorb P, and these effects were inuenced
by the soil acidity which has important implications for soil
productivity. Some studies presented cation modied BCs, such as
Fe [16,18], Mg [19] or La [20] which inuence sorption as well as
desorption properties. These cation-modied BCs enhance the
sorption of P but might negatively inuence the desorption of P
and its plant availability.
Research on the sorption and desorption properties of BCs is
still in the early stages and number of observations need to be
conrmed. Most studies have worked with the capacities of BCs or
of BC-complexes to adsorb. An understanding of the properties of
BCs and of their inuences on the mechanisms of the sorption and
desorption of P is very important for a better management of soil P
applications. The objective of this study was to investigate the
inuences on the adsorption and desorption of phosphate of the
temperature and of the residence time on the carbonization
process of two feedstock bio-materials.
2. Material and methods
BCs were produced by the pyrolysis of sugar cane bagasse (SC)
and of Miscanthus  giganteus (M). The materials were air-dried
and shredded into pieces ca. 23 cm in length. These pieces were
pyrolyzed at 300, 500 and 700  C for residence times of 20 and
60 min, and are referred to as 300/20; 300/60; 500/20; 500/60;
700/20; and 700/60 BCs, respectively. BCs were produced by slow
pyrolysis in a xed-bed reactor. The pyrolysis apparatus consisted
of a temperature controller cabinet, a quartz tube reactor, and an
electric furnace heater. The BCs obtained were ground, washed
with deionized water, dried in a furnace at 105  C for 24 h, and then
stored in plastic containers.
The surface area was measured by N2 adsorption using the BET
surface area analyser. Ultimate analyses of BC samples were carried
out with a CHNS/O analyzer (Elementar Vario LE Cube). Moisture
contents of the BC samples were determined according to the
standard: ICS 75.160.10, DD CEN/TS 14774-3:2004. Ash contents
and volatile matter contents were determined according to ICS
75.160.10, DD CEN/TS 14775:2004 and ICS 75.160.10, DD CEN/TS
15148:2005, respectively. A Fourier transform infrared spectrometer (FTIR) (Cary 630 FTIR spectrometer, Agilent Technologies Inc.)
with the resolution of 4 cm1 and 64 scans per sample was used to
collect spectra in the range of 6004000 cm1. Samples of 300/60,
500/60 and 700/60 SC BCs were selected before and after
adsorption (BCs residual solids from 100 mg P dm3 at pH 7). A
preliminary study showed that SC and MC BCs were similar in their
peaks and absorbance, and no differences were found between
residence times of 20 and 60 min for the same pyrolysis
temperature.
Determination of zeta potential was carried out for all BCs. To
each sample (0.2 g) 100 ml of de-ionized water at pH 6.5 was
added. The change in the zeta potential from P adsorption was
tested by adding also 0.2 g of 500/60 M BC, with or without the
addition of P solution (100 mg P dm3) to each ask, when the pH
of suspensions had been adjusted to within 3.0 to 8.0 with NaOH or
HCl. Then the analysed solutions were shaken at 250 rpm for

30 min using a mechanical shaker. The suspensions were dispersed


ultrasonically for 1 h at 30  1  C in a bath-type sonicator at a
frequency of 40 kHz and a power of 300 W. After that the solution
was ltered using a lter paper (Whatman 42 lter paper). The zeta
potencial of each supernatant solution obtained was determined
using a Malvern Zetasizer Nano (Malvern Instruments).
To study the effects of pH, 0.2 g of 500/20 SC and of M BC was
mixed with 100 cm3 of solution containing 100 mg KH2PO4. The pH
of the solutions were adjusted to values of 310, before adding the
biochar, using a PHM 84 pH meter (Radiometer, Denmark) with
glass REF 451 and calomel pHG 201-8 electrodes. A separate set of
experiments was set up for each pH measurement. The suspensions were agitated on a shaker at 250 rpm and 25  C for 24 h. Each
sample was ltered through a 0.5-mm syringe lter and the pH of
the ltrate was measured.
For the different phosphate adsorption experiments, to BC
(0.2 g in a conical ask) was added 100 cm3 of phosphate solution
each containing: 25, 50, 100, 200 and 400 mg P dm3 at pH 7 (the
pH values were adjusted by adding HCl or NaOH (0.1 mol dm3)).
An initial study indicated that equilibrium for the adsorption of
inorganic P was attained in a matter of minutes. Nevertheless the
suspensions were agitated on a shaker at 200 rpm and 25  C for
24 h. The suspensions were ltered through a 0.5-mm paper lter.
The phosphate contents of the ltrates were meansured, and the
solids were collected for measurements of phosphate desorption.
The experiments were carried out in triplicates and the BCs from
each sample were combined for the desorption experiments.
For desorption, to BCs residual solids (0.2 g) from each
phosphate concentration applied were added to 100 cm3 of
0.01 mol dm3 of citric acid in a conical ask, and the suspensions
were agitated for 24 h on a shaker at 200 rpm and 25  C. Samples
were ltered as described for the adsorption experiments and the
phosphate concentrations in solution were determined.
The equilibrium adsorption capacity was calculated using
Eq. (1)
qe

VC e  C 0
m

where: qe is the adsorption capacity at equilibrium (mg g1), V the


volume of solution (dm3), C0 and Ce are the initial and equilibrium
concentrations of phosphate (mg dm3), and m is the weight of
adsorbent (g).
Phosphate concentrations were determined using a UVvis4000, Varian Spectrophotometer, using the stannous chloride
method [21]. Each experiment was carried out three times, and
mean values are presented. Differences between sorption for BCs
in each treatment were tested for signicance using a factorial
analysis of variance and Duncans multiple range tests. Differences
are reported as signicant at p < 0.05.
The adsorption data of the phosphate on the BCs were analyzed
using the Langmuir and Freundlich isotherm models. The
Langmuir model, described Eq. (2) [22] is.
qe

kL  C e
1 qm  C e

where: Ce is the equilibrium concentration of phosphate (mg


dm3), the constant qm (mg g1) and KL are the characteristics of
the Langmuir equation (dm3 mg1) and can be determined from
the linearized form (plots of Ce/qe vs. Ce). The Freundlich model is
expressed by Eq. (3) [23]:
qe K F  C e n

3
1

where: KF is the Freundlich adsorption capacity (mg g ), n is the


Freundlich constant. The above equation can be linearized to
calculate the parameters KF and n (plots of log qe vs. log Ce).

P.A. Trazzi et al. / Journal of Environmental Chemical Engineering 4 (2016) 3746

3. Results and discussion

Several methods are used [7] for the determination of the


thermodynamic parameters for adsorption systems (liquidsolid).
The common point in all of these methods is the calculation of the
standard free energy change of the adsorption. The thermodynamic parameters, such as Gibbs free energy (DG0), enthalpy
(DH0) and entropy (DS0), were estimated using Eqs. (4) and (6):

DG0 = R  T ln (Kc)

3.1. Material properties


Hemicelluloses and cellulose will be converted mainly to
gaseous products during the pyrolysis process, whereas lignin is
transformed at higher temperatures, mainly to char [27]. Increasing carbonization temperature trended to raise the xed carbon,
the ash content, and the surface area (BET), and to decrease the
volatile matter in the BCs (Table 1). Increasing the residence time
tended to raise the content of xed carbon and to decrease the
content of volatile matter. Increasing the temperature and/or
residence time during the carbonization process decreased the
atomic ratios of H/C and O/C. These observations and the results
presented in Table 1 are typical for the carbonization of
lignocellulosic biomass. The degree of carbonization of BC can
be promoted by increasing the pyrolysis temperature and the
residence time, indicating the BCs obtained might be benecial for
carbon sequestration. Similar results were found for different
carbonization temperatures in references [20,28].
The adsorption capacity of materials can be identied from
their physical characteristics, including porosity, surface area, and
pore size, and especially in case of BC, to ion-exchange capacity.
The surface area of BC increases with increasing temperature until
it reaches a stage at which deformation occurs, resulting in
subsequent decreases in surface area [6]. The surface area depends
largely upon the C mass removed during processing, creating pores
in the material [29]. This is attributed to a sintering effect, followed
by shrinkage of the BC, and realignment of a structure resulting in
decreased pore sizes [30]. Lua et al. [2] evaluated the importance of
pyrolysis temperature, holding time, nitrogen ow rate, and
heating rate on the properties of the BC produced. They concluded
that temperature has the most signicant effect, followed by the
pyrolysis heating rate. The increase in pyrolysis temperature leads
to the increase of the surface area of BC [13,31], and this facilitates
higher sorption of chemicals [32]. Char made from wheat straw at
500700  C was well carbonized and its surface area was relatively
high (>300 m2 g1), whereas chars formed at 300400  C were
partially carbonized and had a lower surface area (<200 m2 g1)
[32].

(4)
1

1

is the gas constant, and T is


where: R = 8.314 J  K mol
temperature in K. Kc is the equilibrium constant, and can be
calculated from Eq. (5).
Kc

qe
Ce

The equilibrium constant was determined for 500/20 SC and M


BCs, for three temperatures, 20, 35 and 50  C. The equilibrium
constant can be calculated based e.g. on the Langmuir equation
[24] or the Freudlich equation [25], or Eq. (5) [26]. Because the
absolute values of DG0 depend on the method applied, it is
important that the same method be used when comparing values
for different materials.
The DH0 and DS0 values were determined from the slope and
intersection, respectively, from Eq. (6).
lnK c

DH0
DS 0

R  T
R

The energy used for temperature rise (Qd) from ambient (Ta) to
the water boiling temperature can be obtain from Eq. (7).
Qd = m  cp,wet  (100 Ta)

(7)

where: m is the mass of wet biomass (kg), and cp,wet is the specic
heat capacity of a wet material (J kg1 K1).
The energy for water evaporation (Qw) was determined from
Eq. (8)
Qw = mw  lw

(8)

where: mw is the mass of the water in the biomass, and lw is the


water vaporization heat J kg1.
Carbonization energy (Qc) can be calculated from Eq. (9)
Qc = (m mw)  cp,dry  (Tc 100)

39

(9)
3.2. FTIR

where: Tc is nal carbonization temperature ( C), and cp,dry is the


specic heat capacity of the dry material (J kg1 K1).

FTIR spectral analysis is important for identifying characteristic


functional groups which are responsible for adsorbing ions [33].

Table 1
Ultimate and proximate analyses of SC and of M BCs samples formed under different pyrolysis conditions.
Biochar

Ultimate analysis wg%

BET

Proximate analysis wg%

Atomic ratios
C/N

H/C

O/C

m2 g1

Volatile matter

Ash

Fixed carbon

Sugar cane
300/20
300/60
500/20
500/60
700/20

0.3
0.3
0.6
0.6
0.5

51.7
54.0
82.0
83.0
86.5

6.0
6.3
3.7
3.4
1.9

0.0
0.0
0.0
0.0
0.0

42.1
39.5
13.7
13.0
11.1

198
193
142
143
159

0.12
0.12
0.05
0.04
0.02

0.81
0.73
0.17
0.16
0.13

4.93
9.20
10.8
60.5
131

46.9
44.7
21.2
20.1
17.2

3.44
2.53
9.54
9.41
12.5

49.7
52.8
69.3
70.6
70.3

Miscanthus
300/20
300/60
500/20
500/60
700/20
700/60

0.4
0.4
0.5
0.6
0.5
0.5

63.1
63.6
83.0
86.3
89.2
90.4

6.0
5.6
3.6
3.2
1.9
1.7

0.0
0.0
0.0
0.0
0.0
0.0

30.5
30.3
12.9
9.92
8.45
7.46

174
143
163
147
179
192

0.10
0.09
0.04
0.04
0.02
0.02

0.48
0.48
0.16
0.11
0.09
0.08

6.17
6.39
21.8
81.0
228
244

45.0
42.7
23.5
22.5
18.9
13.9

6.54
7.12
10.0
11.3
13.3
14.9

48.5
50.3
66.4
66.3
67.7
71.2

40

P.A. Trazzi et al. / Journal of Environmental Chemical Engineering 4 (2016) 3746

Fig. 1. The FTIR spectra of: (a) 300/60, 500/60 and 700/60 SC BCs, and (b) 300/60, 500/60 and 700/60 SC BCs before and after phosphate adsorption.

Fig. 1a shows a typical FTIR spectrum of 300/60, 500/60 and 700/


60 SC BCs. It was observed that the total functional groups
decreased when the pyrolysis temperature is increased. The peak
assignments in the spectra represented methyl C H stretching
compounds (2916 cm1), aromatic carbonyl/carboxyl CO
(1699 cm1), aromatic CC and CO (1595 cm1), aliphatic
COC and alcohol
OH (1030 cm1), and aromatic C
H
1
(815 cm ) [34,35]. These bands presented different changes
when the pyrolysis temperature was increased, the observation
is consistent with other studies [34,36]. Peaks between the

1000600 cm1 interval are attributed to the aromatic C-H


wagging vibration [33]. The polar groups (OH and C
O) exhibited
the lower magnitude of peaks upon heating at high temperature,
suggesting a decrease in the polar functional groups with an
increase in the pyrolysis temperature [34].
To compare the adsorption by the BCs, the second order
derivatives of the FTIR spectra were obtained in order to give a
better display of the peaks and their intensities (Fig. 1b). It was
observed that, despite of showing the same peaks, the peak
intensities were lower in the BCs after P adsorption, that was

Fig. 2. Effect of the adsorption of phosphate by SC and M BCs on the pH change in solution at various initial pH values.

P.A. Trazzi et al. / Journal of Environmental Chemical Engineering 4 (2016) 3746

41

Fig. 3. Zeta potential of 500/60 M BC before and after phosphate adsorption.

especially evident for SC 700/20 BC, and the highest adsorption


capacity was showed by this BC. Similar results were found by Zeng
et al. [31] and by Halajnia et al. [37] as absence of peaks after
adsorption and FTIR spectra did not show such a strong band for
phosphate.

BCs, and the impact of specic adsorption increased with increased


suspension pH. Tong et al. [44] and Xu et al. [41] found similar
results working on Cu and methyl violet adsorption by crop straws
BCs, respectively.
3.5. Adsorption of phosphate

3.3. Effect of initial pH


The species of P and its stability and dominance depends on pH.
As the pH rises, H2PO4 species begin to form at ca pH 4.0, and at
pH 7 the concentration of H2PO4 and HPO42 are approximately
equal [38,39].
It was also found that initially the pH value difference (pHe
pH0), increases between 3 and 4, and then decreases from pH 4 to
pH 10, for SC as well as for M BC (Fig. 2). It indicates that ion
exchange mechanisms are involved in the adsorption [24].
With an increase in pH, OH
 competes strongly with
phosphate for active sites, which affects the adsorption capacity.
This implies adsorption is favored by low pH values and the
capacity would be higher at low pH [40].
3.4. Zeta potential

The data in Table 2 conrm that adsorption of P on BC was


inuenced by the phosphate concentration in solution, by the
carbonization temperature and by the residence time. The surface
area and the adsorption capacity increased with the carbonization
temperature and the longer residence time (Fig. 4).
There was no statistical difference for the adsorption from
solution of each concentration by SC BC made under the conditions
of 500/60, 700/20 and 700/60. Adsorption of phosphate in the 25,
50, 100 and 200 mg dm3 concentration range was statistically
similar for M BCs. However, adsorption from 400 mg dm3 solution

Table 2
Mean values for the adsorption capacities (qe, mg g1) for phosphate at different
concentrations on BCs from SC and M.
Biochar

There was no signicant difference between zeta potential


values for SC and M BCs (F(11,22) = 1.02; p > 0.05). The zeta potential
at pH 6.5 of the M BCs samples ranged from 26.6 to 36.3 mV,
and of SC BCs samples from 28.4 to 34.9 mV, indicating that the BC
particles carried negative charges on their surfaces. Although, the
zeta potential of the 500/60 M BCs became more negative with
increased pH, suggesting that the amount of negative charge
increased with increased pH (Fig. 3), and adsorption should be
lower as observed at Fig. 2. Similar and consistent results were
found by Xu et al. [41], Inyang et al. [42], Yuan et al. [43]. The
presence of P shifted the zeta potential pH curve of 500/60 M BC
particles to positive values directions. The difference in zeta
potential between the systems (500/60 M BC and 500/60 BC after
sorption of P) raised with increased pH of the suspension. These
results suggested that phosphate can be specically adsorbed by

Initial P concentration, mg dm3


25

50

100

200

400

Sugar cane
300/20 0.63
300/60 2.84
500/20 3.81
500/60 5.32
700/20 5.66

dB
cB
cb B
ab C
ab C

1.27
5.28
5.99
6.98
8.11

b AB
a AB
aA
a BC
a BC

1.63
5.28
6.68
9.34
10.0

c AB
b AB
bA
a AB
a AB

2.28
6.08
7.32
10.3
11.0

cA
bA
bA
aA
aA

2.46
6.52
7.51
11.4
11.5

cA
bA
bA
aA
aA

Miscanthus
300/20 0.69
300/60 3.36
500/20 3.99
500/60 5.89
700/20 6.28
700/60 7.66

dD
cC
cC
bE
ab D
aD

1.40
5.53
6.60
7.72
8.95
9.47

e CD
dB
cB
bD
aC
aC

1.83
5.92
7.64
10.5
11.4
12.1

e BC
d AB
c AB
bC
ab B
aB

2.58
7.08
8.55
12.0
12.8
13.2

c AB
b AB
b AB
aB
aA
aB

3.01
7.75
9.00
13.6
13.7
15.5

dA
cA
cA
bA
bA
aA

Means followed by the same letter, capital letters in rows and lowercase in columns,
do not differ by Duncans test at 5% of probability.

42

P.A. Trazzi et al. / Journal of Environmental Chemical Engineering 4 (2016) 3746

Fig. 4. Equilibrium isotherm plots at 20  C for phosphate sorption on: (a) SC BCs, and (b) M BCs. Solid and dashed lines represent the Langmuir and Freundlich isotherm data
model, respectively.

concentration by the 700/60 BC preparation was statistically


higher than for the others BCs.
For each BC, the maximum adsorption was achieved when
phosphate concentration in the solution was highest. However, for
the 300/20, 300/60 and 500/20 SC BC samples the maximum
concentration was not signicantly different from that for the
50 mg dm3. No statistical differences were observed between the
adsorptions of the 200 and 400 mg dm3 phosphate solution
concentrations by M BCs 300/20, 300/60, 500/20 and 700/20.

The phosphate adsorption on the BCs samples were analyzed


using the Langmuir and Freundlich isotherm models. The Langmuir
isotherm assumes a uniform sorption at all of the binding sites,
whereas the Freundlich isotherm suggests a heterogeneous surface
with a nonuniform distribution of heat of adsorption over the
surface [45]. The Langmuir model gave the higher R2 values,
showing that the adsorption on SC and M BCs could be better
described by that model (Table 3). Similar results were found by
Gao et al. [39] and by Kilpimaa et al. [46] for other chars.

P.A. Trazzi et al. / Journal of Environmental Chemical Engineering 4 (2016) 3746


Table 3
Langmuir and Freundlich isotherm parameters for the adsorption of phosphate on
SC and M BCs.
System

Sugar cane bagasse BC


300/20

300/60

500/20

1.676
4.114
0.784

500/60

700/20

700/60

Freundlich

KF
n
R2

0.177
2.143
0.907

2.499
4.968
0.801

2.971
4.205
0.958

3.615
4.755
0.898

4.754
5.865
0.973

Langmuir

qm
KL
R2

2.953
6.887
7.819
0.014
0.043
0.069
0.994
0.998
0.999
Miscanthus  gigantus BC

12.07
0.042
0.999

11.95
0.065
0.999

13.21
0.057
0.996

Freundlich

KF
n
R2

0.167
1.974
0.907

1.867
3.981
0.873

2.387
4.163
0.832

3.149
3.903
0.974

3.845
4.378
0.936

4.960
5.157
0.987

qm
KL
R2

3.754
0.011
0.995

8.244
0.037
0.998

9.425
0.051
0.999

14.47
0.036
0.998

14.31
0.054
0.999

16.10
0.046
0.995

Langmuir

Based on the results presented in Table 2 we can say that there


is a relationship between the amount of phosphate adsorbed and
the conditions under which a BC was made. The higher the
carbonisation temperature and the longer the residence time the
higher is the adsorption capacity, as the process was endothermic.
However, the difference in the amounts of phosphate adsorbed is
smaller for higher temperatures. This rises a question of optimal
energy usage and its relation to the BC sorption capacity. The
energy required for the pyrolysis process is a sum of the energy
(QS) required to raise the temperature of the wet material to the
boiling water temperature (Qd) plus the energy for water
evaporation (Qw), and the carbonisation energy (Qc). The rst
two energies depend on the moisture content and are independent
from the carbonisation temperature and are the same for each
material. While the energy needed for carbonisation rises with the
nal temperature. We assume no energy loses by conduction or
radiation to the surroundings during these processes, this
simplication does not change the nal conclusions. In terms of
energy expended, the BC formed at 500  C for 60 min performed
best in relation to phosphate sorption, as described in Fig. 5. The
amount of phosphate adsorbed on SC as well as M BCs divided by

43

the sum of energy used for the BCs production was optimum for BC
produced at 500 C for longer residence time.
The adsorption of phosphate by different BCs is listed in Table 4.
Our work shows that the amount of phosphate adsorbed depends
on the carbonization process parameters, as well as on the initial
phosphate concentration in solution. These parameters should be
taken into consideration when comparing the sorption capacities
for various BCs. The data presented in Table 4 indicate that none of
the BCs listed had a signicantly better sorption capacity. The
largest difference presented in Table 4 is between two activated
carbon samples made in relatively similar ways. Our work
indicates that the method of preparation of BC samples has an
important bearing on the sorption of phosphate.
3.6. Adsorption thermodynamics
Table 5 presents the thermodynamic parameters for different
temperatures. The positive values of Gibbs energy suggested that
the adsorption of phosphate onto each BC was non-spontaneous.
Additionally, the positive values of enthalpy changes indicated that
the adsorptions were endothermic. The positive values of the
entropy changes indicate increased randomness at the solid/
solution interface during the adsorption of phosphate onto these
BCs [47]. Results similar to ours were obtained by Foo and Hammed
[49] for activated carbon from palm oil bres. A majority of
researchers have also found phosphate adsorption on BCs to be
endothermic [19,47,49,50] and random [19,47,49,50] However, in
some cases [19,47,50] the Gibbs energy did not always fall with
increasing temperature, indicating that phosphate was adsorbed
efciently at a high solution temperatures. The adsorption of
phosphate increases with temperature, as indicated by Fig. 6. The
data indicate that higher temperatures are favorable for phosphate
adsorption onto these BCs. At higher temperature, the reactivity of
the surface sites and the rate of intraparticle diffusion of sorptive
ions into the pores of adsorbent increased [37,51]. The majority of
researches conrm that phosphate sorption increases with
temperature [19,37,39,47,49,50,52,53]. Opposite results were
obtained for the sorption of phosphate on dolomite mineral
[25,54] and on modied wheat residue [55], suggesting a tendency
for the phosphate ions to escape from the solid phase to the bulk
phase as the temperature of the solution increased.

Fig. 5. The amount of phosphate adsorbed divided by the energy required for the production of SC and M BCs under various conditions.

44

P.A. Trazzi et al. / Journal of Environmental Chemical Engineering 4 (2016) 3746

Table 4
Phosphate adsorption capacities of different BCs formed under different conditions.
Conditions

Material production

Pine sawdust BC
Embauba
Lacre
Inga BCs
Corn BC

Activated sugarcane bagasse


BC

P Adsorption capacity, Refs.


qe
mg g1

126
219

[47]

388
495
382
490

314

10
14
14
20
10
35
40
50
70
11
20
14

501
19.4

0.3
1.2

[18]

285

21

[46]

>350

0.06 to 1.11

[48]

P initial
concentration
mg dm3

550  C
750  C
500  C

for
15 min

300  C
600  C

for 3 h

100

500  C

for
30 min

61.3

30
40

Mg modied corn BC
Oak sawdust BC
La modied oak sawdust BC
Ferrihydrite modied rice
straw BC
Orange peel BC
Fe3+/Fe2+ modied orange peel
BC
Activated carbon residue BC

Surface area
(N2)
m2 g1

600  C

100
(pH 7)
2

700  C with ramp 5  C/min

downdraft gasier 1000  C.


kiln at 400  C for 2 days and activation at 900  C for 100
min

50
(pH 6)
130

[5]

[19]

[20]
[16]

3.7. Desorption of P
Table 5
Thermodynamic parameters at different temperatures for SC and M BCs at
phosphate concentration of 100 mg dm3.
Temperature
( C)

DG0

DS0

DH0

R2

Sugar cane

20
35
50

6.242
5.845
5.386

0.010

9.197

0.990

Miscanthus

20
35
50

5.861
5.289
4.982

0.013

9.509

0.959

Biochar

(kJ mol1)

(kJ mol1 K1)

(kJ mol1)

As for adsorption, the process for desorption of P was inuenced


by the P concentration in solution, by the carbonization temperature, and by the residence time. Results presented in Table 6 show
that SC and M BC samples that adsorbed greater amounts of P,
released more to the equilibrium solution during the desorption
process.
The phosphate desorbability can be dened as the ratio of the
desorbed phosphate to the total adsorbed by the adsorbents, and it
can be used to indicate the degree of phosphate desorption from
the adsorbate [16].
In terms of the phosphate desorbed, there are no statistical
differences for each P concentration between the percentage
desorbed from the SC and M BCs (Fig. 7). For the SC BCs with the

Fig. 6. Effect of temperature on the sorption of phosphate from 50, 100 and 200 mg dm3 initial concentration of phosphate by 500/20 SC and M BCs.

P.A. Trazzi et al. / Journal of Environmental Chemical Engineering 4 (2016) 3746


Table 6
Mean values of phosphate (%) desorbed from different concentrations of phosphate
adsorbed on SC and M BCs.
Biochar

Phosphate concentration, mg dm3


25

50

100

200

400

Sugar cane
300/20
19.6
300/60
20.6
500/20
22.1
500/60
20.9
22.1
700/20
700/60
22.4

a
a
a
a
a
a

E
D
E
E
E
E

25.4
26.4
26.2
27.5
26.3
29.2

a
a
a
a
a
a

D
C
D
D
D
D

29.0
28.9
30.7
32.6
31.1
32.5

a
a
a
a
a
a

C
C
C
C
C
C

36.3
34.2
36.6
38.0
38.1
38.3

a
a
a
a
a
a

B
B
B
B
B
B

44.4
44.8
43.5
45.3
46.3
45.6

a
a
a
a
a
a

A
A
A
A
A
A

Miscanthus
19.1
300/20
300/60
21.0
500/20
20.5
500/60
19.7
700/20
23.7
700/60
21.0

a
a
a
a
a
a

D
C
C
E
E
D

24.7
27.7
25.5
29.3
25.2
30.7

a
a
a
a
a
a

D
C
C
D
D
C

28.2
30.2
29.7
32.2
32.5
31.6

a
a
a
a
a
a

C
C
B
C
C
BC

36.9
32.7
37.8
38.3
40.5
36.3

a
a
a
a
a
a

B
B
A
B
B
B

44.5
48.1
40.7
44.5
48.3
47.2

a
a
a
a
a
a

A
A
A
A
A
A

Means followed by the same letter, capital letters in rows and lowercase in columns,
do not differ by Duncans test at 5% of probability.

45

lowest and the highest initial P concentrations, the desorption was


19.6 to 22.4% (F(5,10) = 0.39; p > 0.05) and 43.5 to 46.3% (F(5,10) = 0.38;
p > 0.05), respectively. Very similar results were observed for the M
BC, with corresponding gures of 19.1 to 23.7% (F(5,10) = 0.92;
p > 0.05), for the lowest initial phosphate concentration, and 40.7
to 48.3% (F(5,10) = 1.38; p > 0.05), for the highest.
Results are consistent with those obtained by Cui et al. [16]; the
rate of phosphate desorption from the BC complex is clearly
dependent on the desorption conditions employed, the gradient of
phosphate concentration, and the concentrations of various ions in
solution. According to Morales et al. [5], the feedstock makes a
signicant difference in phosphate desorption properties and
opens the possibility for designing BC preparations for specic soil
management objectives.
4. Conclusions
Increasing the carbonization temperature and residence time
for biomass gave rise to incresed xed carbon and surface area.

Fig. 7. Effect of the initial concentration on phosphate desorption from SC and M BCs.

46

P.A. Trazzi et al. / Journal of Environmental Chemical Engineering 4 (2016) 3746

Miscanthus biochar has a higher surface area and higher phosphate


adsorption capacity than sugar cane bagasse biochar. The greatest
changing pH adsorption corresponded to the highest adsorption.
Desorption was greatest in cases where most phosphate was
adsorbed. Biochars formed at 500  C for 60 min performed best in
relation to phosphate sorption in terms of energy expended.
The experimental data tted with the Langmuir isotherm
marginally better than with the Freundlich isotherm, indicating a
uniform phosphate sorption at all of the binding sites. Adsorption
of phosphate on the biochars was an endothermic non-spontaneous process and the adsorption incresed with the temperature.
Acknowledgements
Paulo Trazzi thanks to Conselho Nacional de Desenvolvimento
Cientco (CNPq) for receiving his Post-Doctoral grant.
References
[1] L.H. Bohn, B.L. McNeal, G.A. OConnor, Soil Chemistry, John Wiley and Sons Inc.,
New York, 2001, pp. 322.
[2] A.C. Lua, T. Yang, J. Guo, J. Anal. Appl. Pyrolysis 72 (2004) 279.
[3] K.A. Spokas, K.B. Cantrell, J.M. Novak, D.W. Archer, J.A. Ippolito, H.P. Collins, A.A.
Boateng, I.M. Lima, M.C. Lamb, A.J. McAloon, R.D. Lentz, K.A. Nichols, J. Environ.
Qual. 41 (2012) 973.
[4] S.D. Joseph, M. Camps-Arbestain, Y. Lin, P. Munroe, C.H. Chia, J. Hook, L. van
Zwieten, S. Kimber, A. Cowie, B.P. Singh, J. Lehmann, N. Foidl, R.J. Smernik, J.E.
Amonette, Aust. J. Soil Res. 48 (2010) 501.
[5] M.M. Morales, N. Comerford, I.A. Guerrini, N.P.S. Falcao, J.B. Reeves, Soil Use
Manag. 29 (2013) 306.
[6] A. Downie, A. Crosky, P. Munroe, Biochar for Environmental Management:
Science and Technology, in: J. Lehmann, S. Joseph (Eds.), Earthscan, London,
2009, pp. 13.
[7] A.V. Piterina, M.H.B. Hayes, Inuence of the addition of herbaceous biochar on
the metabolic proles of the maize rhizosphere microbial community, 4th
International Biochar Congress, Beijing, 2012, pp. 74.
[8] G. Xu, J.N. Sun, H.B. Shao, S.X. Chang, Ecol. Eng. 62 (2014) 54.
[9] T.T. Qian, X.S. Zhang, J.Y. Hu, H. Jiang, Chemosphere 93 (2013) 2069.
[10] C.M. Monterumici, D. Rosso, E. Montoneri, M. Ginepro, A. Baglieri, E.H.
Novotny, W. Kwapinski, M. Negre, Int. J. Mol. Sci. 16 (2015) 8826.
[11] R.Z. Yin, R.H. Liu, Y.F. Mei, W.T. Fei, X.Q. Sun, Fuel 112 (2013) 96.
[12] R. Chintala, T.E. Schumacher, L.M. McDonald, D.E. Clay, D.D. Malo, S.K.
Papiernik, S.A. Clay, J.L. Julson, Clean-Soil Air Water 42 (2014) 626.
[13] R.A. Brown, A.K. Kercher, T.H. Nguyen, D.C. Nagle, W.P. Ball, Org. Geochem. 37
(2006) 321.
[14] J. Lehmann, Front. Ecol. Environ. 5 (2007) 381.
[15] S. Joseph, C. Peacocke, J. Lehmann, P. Munroe, Biochar for Environmental
Management: Science and Technology, in: J. Lehmann, S. Joseph (Eds.),
Earthscan, London, 2009, pp. 107.
[16] H.J. Cui, M.K. Wang, M.L. Fu, E. Ci, J. Soils Sediments 11 (2011) 1135.
[17] M.M. Parvage, B. Ulen, J. Eriksson, J. Strock, H. Kirchmann, Biol. Fertil. Soils 49
(2013) 245.
[18] B.L. Chen, Z.M. Chen, S.F. Lv, Bioresour. Technol. 102 (2011) 716.

[19] C. Fang, T. Zhang, P. Li, R.F. Jiang, Y.C. Wang, Int. J. Environ. Res. Public Health 11
(2014) 9217.
[20] Z.H. Wang, H.Y. Guo, F. Shen, G. Yang, Y.Z. Zhang, Y.M. Zeng, L.L. Wang, H. Xiao,
S.H. Deng, Chemosphere 119 (2015) 646.
[21] American Public Health Association (APHA) Standard Methods, Method 4500P D, 21st Ed., 2005, Washington, D.C.
[22] I. Langmuir, J. Am. Chem. Soc. 38 (1916) 2221.
[23] H.F.M. Freundlich, Z. Phys. Chem. 57 (1906) 385.
[24] D. Kolodynska, R. Wnetrzak, J.J. Leahy, M.H.B. Hayes, W. Kwapinski, Z. Hubicki,
Chem. Eng. J. 197 (2012) 295.
[25] W.X. Xiaoli Yuan, J. An, Jianguo Yin, Xuejiao Zhou, Wenqiang Yang, J. Chem.
853105 (2015) 1 ID.
[26] P.N. Diagboya, B.I. Olu-Owolabi, D. Zhou, B.H. Han, Carbon 79 (2014) 174.
[27] R. Brown, Biochar for Environmental Management: Science and Technology,
in: J. Lehmann, S. Joseph (Eds.), Earthscan, London, 2009, pp. 127.
[28] W.P. Song, M.X. Guo, J. Anal. Appl. Pyrolysis 9 (2012) 138.
[29] A. Zabaniotou, G. Stavropoulos, V. Skoulou, Bioresour. Technol. 99 (2008) 320.
[30] J. Guo, A.C. Lua, J. Anal. Appl. Pyrolysis 46 (1998) 113.
[31] Z. Zeng, S.D. Zhang, T.Q. Li, F.L. Zhao, Z.L. He, H.P. Zhao, X.E. Yang, H.L. Wang, J.
Zhao, M.T. Raq, J. Zhejiang Univ.-Sci. B 14 (2013) 1152.
[32] J.C. Tang, W.Y. Zhu, R. Kookana, A. Katayama, J. Biosci. Bioeng. 116 (2013) 653.
[33] Z.H. Zhou, J.H. Yuan, M.H. Hu, Environ. Progress Sustainable Energy 34 (2015)
655.
[34] X.P. Gai, H.Y. Wang, J. Liu, L.M. Zhai, S. Liu, T.Z. Ren, H.B. Liu, PLoS One 9 (2014) .
[35] K. Sun, K. Ro, M.X. Guo, J. Novak, H. Mashayekhi, B.S. Xing, Bioresour. Technol.
102 (2011) 5757.
[36] B.L. Chen, D.D. Zhou, L.Z. Zhu, Environ. Sci. Technol. 42 (2008) 5137.
[37] A. Halajnia, S. Oustan, N. Naja, A.R. Khataee, A. Lakzian, Appl. Clay Sci. 8081
(2013) 305.
[38] K. Kaikake, T. Sekito, Y. Dote, Waste Manag. 29 (2009) 1084.
[39] Y. Gao, N. Chen, W.W. Hu, C.P. Feng, B.G. Zhang, Q. Ning, B. Xu, J. Solution Chem.
42 (2013) 691.
[40] Y.J. Xue, H.B. Hou, S.J. Zhu, J. Hazard. Mater. 162 (2009) 973.
[41] R.K. Xu, S.C. Xiao, J.H. Yuan, A.Z. Zhao, Bioresour. Technol. 102 (2011) 10293.
[42] M. Inyang, B. Gao, P. Pullammanappallil, W.C. Ding, A.R. Zimmerman,
Bioresour. Technol. 101 (2010) 8868.
[43] J.H. Yuan, R.K. Xu, H. Zhang, Bioresour. Technol. 102 (2011) 3488.
[44] X.J. Tong, J.Y. Li, J.H. Yuan, R.K. Xu, Chem. Eng. J. 172 (2011) 828.
[45] K.L. Wasewar, M. Atif, B. Prasad, I.M. Mishra, Clean-Soil Air Water 36 (2008)
320.
[46] S. Kilpimaa, H. Runtti, T. Kangas, U. Lassi, T. Kuokkanen, Chem. Eng. Res. Des. 92
(2014) 1923.
[47] F. Peng, P.W. He, Y. Luo, X. Lu, Y. Liang, J. Fu, Clean-Soil Air Water 40 (2012) 493.
[48] M.N. Liang, H.H. Zeng, Y.N. Zhu, Z.L. Xu, H.L. Liu, Environmental Biotechnology
and Materials Engineering, Pts 13, in: Y.G. Shi, J.L. Zuo (Eds.), Trans Tech
Publications Ltd., Stafa-Zurich, 2011, pp. 1046.
[49] K.Y. Foo, B.H. Hameed, Fuel Process. Technol. 99 (2012) 103.
[50] C. Namasivayam, D. Sangeetha, J. Colloid Interface Sci. 280 (2004) 359.
[51] R. Chitrakar, S. Tezuka, A. Sonoda, K. Sakane, K. Ooi, T. Hirotsu, J. Colloid
Interface Sci. 290 (2005) 45.
[52] W. Jutidamrongphan, K.Y. Park, S. Dockko, J.W. Choi, S.H. Lee, Environ. Chem.
Lett. 10 (2012) 21.
[53] L. Zeng, X.M. Li, J.D. Liu, Water Res. 38 (2004) 1318.
[54] S. Karaca, A. Gurses, M. Ejder, M. Acikyildiz, J. Colloid Interface Sci. 277 (2004)
257.
[55] X. Xu, B.Y. Gao, W.Y. Wang, Q.Y. Yue, Y. Wang, S.Q. Ni, Colloids Surfaces BBiointerfaces 70 (2009) 46.

Вам также может понравиться