You are on page 1of 9

Journal of Analytical and Applied Pyrolysis 104 (2013) 1018

Contents lists available at ScienceDirect

Journal of Analytical and Applied Pyrolysis


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

Optimization of pyrolysis conditions and adsorption properties of


bone char for uoride removal from water
C.K. Rojas-Mayorga a , A. Bonilla-Petriciolet a, , I.A. Aguayo-Villarreal a ,
V. Hernndez-Montoya a , M.R. Moreno-Virgen a , R. Tovar-Gmez a , M.A. Montes-Morn b
a
b

Instituto Tecnolgico de Aguascalientes, Aguascalientes, Av. Lpez Mateos No. 1801, C.P. 20256, Mxico
Instituto Nacional del Carbn, INCAR-CSIC, Apartado 73 E-33080, Oviedo, Spain

a r t i c l e

i n f o

Article history:
Received 17 June 2013
Accepted 29 September 2013
Available online 9 October 2013
Keywords:
Fluoride adsorption
Pyrolysis
Bone char
Water treatment

a b s t r a c t
This study reports the optimization of a pyrolysis process for the synthesis of bone char for uoride
removal from water. Specically, we have performed a detailed analysis of the adsorption properties of
bone char samples obtained at different operating conditions of pyrolysis. Results show that the pyrolysis temperature plays a major role to synthetize an effective bone char for water deuoridation. In
particular, the best adsorption properties of bone char for uoride removal are obtained with those samples synthetized at 700 C. Pyrolysis temperatures higher than 700 C cause the dehydroxylation of the
hydroxyapatite of bone char reducing its uoride adsorption capacity. The maximum uoride adsorption capacity of the bone char obtained in this study (i.e., 7.32 mg/g) is higher than those reported for
commercial bone chars. Finally, adsorption experiments were performed using the optimized bone char
for determining kinetic, equilibrium and thermodynamic parameters of the uoride removal from water
using this adsorbent. In summary, this study shows that the optimization of pyrolysis conditions for the
synthesis of bone char is useful to obtain an effective adsorbent for uoride removal from water.
2013 Elsevier B.V. All rights reserved.

1. Introduction
Pyrolysis has been traditionally employed for the synthesis and
production of adsorbents for air pollution control and wastewater
treatment [1,2]. In particular, pyrolysis is useful for the synthesis
of bone char, which has been recognized as an effective adsorbent
for uoride removal from water [36]. Deuoridation of drinking
water for human consumption is a relevant concern in the context
of health protection and water pollution control [7]. It has been
estimated that 260 million human beings worldwide are exposed
to the consumption of drinking water with uoride concentrations
higher than 1 mg/L, which has been established as the uoride concentration limit to prevent the development of dental and skeletal
uorosis in exposed population [8]. Therefore, the production of
low-cost bone char for the treatment of drinking water polluted by
uoride ions has increased its importance and relevance especially
in developing countries.
Bone char is considered as a mixed adsorbent constituted by
carbon and calcium phosphate, which is in the hydroxyapatite form
[9]. This adsorbent has shown uoride adsorption capacities higher
than those obtained for other carbonaceous materials [4,10] and

Corresponding author. Tel.: +52 449 9105002.


E-mail address: petriciolet@hotmail.com (A. Bonilla-Petriciolet).
0165-2370/$ see front matter 2013 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.jaap.2013.09.018

commercial adsorbents such as activated alumina [11,12]. Several


studies have concluded that the uoride adsorption properties of
bone char are attributed to their content of mineral components,
especially, the hydroxyapatite content [6,11,12].
Bone char can be synthetized by the calcination of bones under
a low-oxygen environment (i.e., partial calcination) or by pyrolysis, where no oxygen is present during the thermal treatment.
In particular, thermal processes with restricted presence of air at
500600 C are commonly used for the synthesis of bone char for
water deuoridation [9]. In fact, some authors have suggested that
bone char produced by partial calcination are suitable for uoride
removal if the synthesis temperatures do not exceed 500 C [3].
However, several studies available in the literature have reported
discrepancies and different results with respect to the effect of the
synthesis conditions on the adsorption properties of bone char for
uoride removal from water. For example, Mwaniki [13] determined the uoride removal performance of white, gray and black
bovine bone chars, which were obtained from thermal treatments
at 350 C (black), 450 C (gray) and 600 C (white). This study concluded that black bone char showed the best adsorption properties.
On the other hand, Phantumvanit and LeGeros [14] reported a comparison of uoride removal performances of bone chars obtained by
calcination of bone meal (i.e., cow and pig bones) for 30 min at different temperatures: 400, 600 and 800 C. This study concluded that
bone char obtained at 400 C showed the highest uoride uptakes.

C.K. Rojas-Mayorga et al. / Journal of Analytical and Applied Pyrolysis 104 (2013) 1018

In other study, Kaseva [15] reported that thermal treatment of cattle


bones at temperatures higher than 600 C may damage the apatite
structure affecting the uoride adsorption properties of bone char,
while thermal treatments of bones at temperatures below 400 C
may produce a bone char that is not hygienically acceptable for
water deuoridation since the water treated with this adsorbent
may have bad taste and odor. Recently, Kawasaki et al. [5] studied
the synthesis of bone char using cow, pig, chicken and sh bones
and a pyrolysis process at 800 and 1000 C. This study concluded
that the uoride adsorption properties of bone char obtained at low
pyrolysis temperature are higher than those obtained at 1000 C.
The uoride uptakes of synthetized bone chars ranged from 0.5 to
2.5 mg/g. Finally, Leyva-Ramos et al. [6] reported that the bone char
produced by partial calcination at 550 C and a charring time of 4 h
is suitable for the uoride removal from water. Even though these
studies have reported inconsistent results, it is clear that the operating conditions for the synthesis of bone char play an important role
to determine the nal adsorption properties of the adsorbent and,
consequently, its performance in water deuoridation. In general,
the uoride adsorption capacities of commercial bone chars may
range from 1.0 to 3.0 mg/g [11,12,16], where bone chars obtained
from pyrolysis process may show the best adsorption properties.
Herein, it is convenient to remark that, in a controlled synthesis
process, the precursor is pyrolyzed under a suitable temperature and inert environment; and the adsorption properties of the
obtained adsorbent depend on the operating conditions used. The
experimental parameters of pyrolysis that have the largest inuence on the physicochemical properties of the nal adsorbent
are the precursor particle size, temperature of thermal treatment,
nitrogen ow rate, residence time and heating rate [17,18]. These
process conditions can be optimized to improve the adsorption
properties of the nal product for a desired application.
Based on these facts, this study reports a detailed analysis of the
effect of pyrolysis conditions on the physicochemical properties of
the bone chars and their ability for uoride removal from water.
Specically, the optimal conditions of the pyrolysis process that
improve the uoride uptakes of bone char have been determined.
Results show that the bone char prepared using the optimum pyrolysis conditions has uoride uptakes higher than those reported for
commercial bone chars available in the market.

2. Methodology
2.1. Pyrolysis conditions for the synthesis of bone char
Cow femur residues were used as precursor for the synthesis
of bone char. These bone residues were washed with boiling and
deionized water to eliminate fat and meat residues. Bone samples
were then dried for 24 h, crushed and sieved to obtain a particle
size of 1 mm. These bone samples were used for the preparation
of bone char using a pyrolysis process.
A tubular furnace Carbolite Eurotherm CTF 12165/550 with a
quartz sample holder was used for the synthesis of bone char samples. Samples of bone char were synthetized at specic conditions
of heating rate, pyrolysis temperature and thermal treatment time,
which were dened using a full factorial design Ak , see Table 1. For
pyrolysis process, nitrogen gas (400 mL/min) was used to provide
an inert atmosphere during the synthesis of bone char. All synthetized bone char samples were washed with deionized water
until obtaining a constant pH and dried for their use in uoride
adsorption experiments.
For all experiments established in the experimental design
(Table 1), the product yield was quantied and the uoride adsorption properties of bone char samples were determined. Specically,
the adsorbed amount of uoride on bone chars was used as

11

the response variable of this experimental design. This uoride


uptake was measured in batch adsorption experiments, which were
performed by triplicate using a uoride solution with an initial concentration of 60 mg/L and an adsorbent dosage of 2 g/L at 30 C and
pH 7. Fluoride uptakes of bone chars (mg/g) were calculated using
a mass balance


qF =

[F ]0 [F ]f
m


V

(1)

where [F ]0 and [F ]f are the initial and nal uoride concentration in adsorption experiment given in mg/L, m is the mass of bone
char samples in g and V is the volume of uoride solution given in
L, respectively. Fluoride concentration in solution was quantied
using a selective electrode and TISAB chemical reagent according
to the procedure described in the Standard Methods of Examination
of Water and Wastewater [19].
A statistical analysis of the results obtained from this experimental design was performed to determine the effect of pyrolysis
conditions on the uoride adsorption properties of bone char. In
addition, this analysis was used to identify those operating parameters of pyrolysis process that improve the performance of this
adsorbent for uoride removal from water. Statistica software
was used for data analysis.
The best pyrolysis conditions were identied and the synthetized adsorbent was used in additional removal experiments for
determining kinetics, isotherms and thermodynamic parameters of
the uoride adsorption on bone char at different conditions of temperature and pH. Specically, the optimized bone char sample was
used for determining the uoride adsorption rates at initial uoride
concentrations of 10, 30 and 50 mg/L, respectively. On the other
hand, the uoride adsorption isotherms using this bone char were
obtained at pH 6, 7 and 8 and 20, 30 and 40 C. Residence time used
for batch adsorption experiments ranged from 0.5 to 24 h. These
results were used for the calculation of adsorption thermodynamic
parameters.

2.2. Physicochemical characterization of raw precursor and bone


chars
Several characterization techniques were used for determining
the most relevant physicochemical properties of both raw precursor (i.e., bovine femur) and bone char samples obtained in this
study. Specically, the thermogravimetric analysis of the raw bone
was performed from room temperature to 1000 C under an Ar
atmosphere employing an SDT Q600 automatic analyzer from TA
Instruments. The content of C, H, N and S in the raw bone and bone
char samples was determined using appropriate elemental micro
analyzers from LECO. The functional groups were determined using
a Nicolet-8700 FT-IR spectrometer (Thermo Electron Co.) equipped
with a deuterated triglycine sulfate (DTGS) detector. FT-IR spectra were collected in the spectral range of 40004500 cm1 . On
the other hand, SEM/EDX analyses were performed in a FE-SEM
system (Quanta FEG 650, FEI) to determine possible changes in
the morphology and composition of both raw bone wastes and
bone char samples. The crystalline structure of bone char samples was analyzed using a Bruker D8 Advance X-ray diffractometer.
Finally, the textural parameters of adsorbent samples were determined using nitrogen adsorptiondesorption isotherms at 77 K
with a Micromeritics ASAP 2420. Results of physicochemical characterization of the bone char samples were related to the uoride
adsorption and used for analyzing the uoride removal mechanism.

12

C.K. Rojas-Mayorga et al. / Journal of Analytical and Applied Pyrolysis 104 (2013) 1018

Table 1
Experimental design used for the synthesis of bone char via pyrolysis. N2 ow: 400 mL/min.
Pyrolysis conditions

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20

650

700

800

900

1000

Bone char performance


Heating rate, C/min

Residence time, h

Yield, %

Fluoride uptake, mg/g

5
5
10
10
5
5
10
10
5
5
10
10
5
5
10
10
5
5
10
10

2
4
2
4
2
4
2
4
2
4
2
4
2
4
2
4
2
4
2
4

75.33
73.45
74.13
73.18
75.38
73.29
74.25
73.08
75.41
73.24
74.28
73.12
73.59
71.79
72.91
71.23
72.73
71.70
71.91
70.53

6.70
6.62
6.51
6.63
6.96
7.06
7.32
7.16
6.67
6.64
6.71
6.57
2.99
3.08
3.03
3.01
1.34
1.28
1.24
1.25

3. Results and discussion


3.1. Optimization of pyrolysis conditions for the synthesis of bone
char for uoride removal
Product yields and uoride adsorption capacities of the bone
char samples obtained at different pyrolysis conditions of the
experimental design are reported in Table 1. Specically, the yields
of bone char ranged from 70.53 to 75.41% at tested experimental
conditions and these results indicate that the tested pyrolysis conditions do not have a signicant impact on the product yield of
bone chars, see Table 1. However, an increment of the residence
time seems to systematically cause a slight decrement (3%) of the
bone char yield. This trend agrees to results reported in other studies, e.g. [5]. Overall, the product yields of bone char obtained in this
study are higher (i.e., >70%) than those obtained by Kawasaki et al.
[5] using bones from cow, pig, chicken and sh as precursors.
On the other hand, the results of uoride adsorption experiments for the full factorial experimental design are also reported
in Table 1. Statistical analysis (i.e., variance analysis) of the experimental design indicated that the pyrolysis temperature has the
most signicant effect on the uoride adsorption properties of
bone char (p-level < 0.01), whereas both the heating rate and thermal treatment time do not affect signicantly the performance
of the bone chars (p-level > 0.05). It is interesting to remark that,
when the pyrolysis temperature is higher than 800 C, the uoride adsorption capacity of bone char decreased signicantly from
6.0 mg/g to 3.0 mg/g. In fact, the removal performance of bone
char synthetized at 1000 C decreased signicantly showing uoride adsorption capacities of 1.0 mg/g. This trend of adsorbent
performance is in agreement with the study of Kawasaki et al. [5],
which indicates that bone chars obtained at higher temperatures
may show low uoride uptakes due to the thermal degradation
of functional groups that may be involved in uoride removal.
With illustrative purposes, Fig. 1a shows the uoride adsorption
isotherms of bone chars obtained at different pyrolysis temperatures from 650 to 1000 C, using a residence time and heating
rate of 2 h and 10 C/min, respectively. Maximum uoride adsorption capacities ranged from 1.26 to 7.32 mg/g, respectively. On the
other hand, adsorption isotherms reported in Fig. 1b conrmed that
dwell time and heating rate of the thermal treatment do not have
a signicant impact on the uoride adsorption properties of bone
char.

A detailed analysis of the effect of pyrolysis temperature was


performed and additional bone char samples were prepared at 690,
710, 720 and 730 C, keeping all other variables of the pyrolysis process constant (10 C/min and 2 h). Maximum uoride adsorption
capacities of the synthetized bone chars are reported in Fig. 2. These
results show that the best uoride removal performance of bone
char corresponds to a pyrolysis temperature of 700 C. In fact, this
pyrolysis temperature is the optimum condition for the synthesis
of bone char where

qF
Tp

=0

(2)

tp ,rp

a)

8
7
6
qe, mg/g

Temperature, C

5
Pyrolysis temperature

650C
700C
800C
900C
1000C

3
2
1
0
0

10

20

30

40
50
[F-]e, mg/L

60

70

80

b)

8
7
6
qe, mg/g

Sample

5
Pyrolysis conditions

5C/min, 2 h

5C/min, 4 h
2

10C/min, 2 h

10C/min, 4 h

0
0

10

20

30

40
50
[F-]e, mg/L

60

70

80

Fig. 1. Fluoride adsorption isotherms at pH 7 and 30 C on bone chars obtained at


different pyrolysis conditions.

8
7
6
5
4
3
2
1
0

13

1
0.9

R = 0.9972

0.8
0.7
0.6
[F-]t/[F-]0

Maximum fluoride uptake, mg/g

C.K. Rojas-Mayorga et al. / Journal of Analytical and Applied Pyrolysis 104 (2013) 1018

0.5
0.4
[F-]0, mg/L

0.3

650

700

750
800
850
900
Pyrolysis temperature, C

950

10
30
50

0.2

1000

0.1
0

Fig. 2. Effect of pyrolysis temperature on the maximum uoride uptake of bone


chars at pH 7 and 30 C.

where qF is the uoride adsorption capacity of bone char, rp is the


residence time and Tp is the pyrolysis temperature, respectively.
For this pyrolysis temperature, the bone char showed the maximum uoride adsorption capacity, i.e., 7.32 mg/g at pH 7 and 30 C.
Based on the statistical analysis of the experimental design, the
best conditions for the synthesis of bone char via pyrolysis are: a
pyrolysis temperature of 700 C, a heating rate of 10 C/min and a
residence time of 2 h. These pyrolysis conditions fulll the lowest
energy requirements for the synthesis of bone char without compromising the product yield and its uoride adsorption properties.
It is convenient to highlight that the maximum uoride adsorption
capacity of bone char prepared in this study is higher than those
reported for commercial bone chars, e.g.: Fija uor 5.0 mg/g [12],
Brimac 3.5 mg/g and Carbones Mexicanos 4.5 mg/g [16]. Also,
the removal performance of this bone char outperforms the results
reported by Abe et al. [4] and Kawasaki et al. [5], where the uoride
adsorption capacities ranged from 1.0 to 3.0 mg/g. In summary,
the optimization of pyrolysis conditions for the synthesis of bone
char is useful to obtain an effective adsorbent for uoride removal
from water, which shows better uoride adsorption properties than
those reported for other commercial bone chars.

3.2. Kinetic, equilibrium and thermodynamic parameters of


uoride adsorption on bone char prepared using the optimal
pyrolysis conditions
Fluoride adsorption kinetics at pH 7 and 30 C for the bone char
obtained at the best pyrolysis conditions are given in Fig. 3 and the
adsorption rates were calculated using the pseudo-second order
kinetic model, see Table 2. This kinetic model was used because preliminary calculations indicated that it may offer a better correlation
performance than those obtained for other traditional adsorption
kinetic models. Adsorption rates of uoride on bone char ranged
from 8.02E04 to 2.94E03 g/mg min at tested experimental conditions. In particular, the initial adsorption rate (h) increased with

10

12
t, h

14

16

18

20

22

24

Fig. 3. Fluoride adsorption kinetics at pH 7 and 30 C on the bone char synthetized


at the optimal pyrolysis conditions.

the initial concentration of uoride in the solution and the adsorption equilibrium was obtained at 24 h.
Kinetic data of uoride adsorption were also analyzed using the
WeberMorris intraparticle diffusion model. Fig. 4 shows the plots
of qt versus t0.5 for uoride adsorption kinetic data at pH 7 and 30 C.
It is clear that several steps are involved in the adsorption process
of uoride on bone char because the plot qt t0.5 is multi-linear.
First stage of uoride adsorption is related to surface diffusion, second stage comprises the intraparticle or pore diffusion, and the
nal stage is given by the adsorption equilibrium. So, the uoride
adsorption rate is controlled by these stages. Note that the value of
intercept C of diffusion model increased with the uoride concentration (see Table 2), thus indicating that the surface diffusion has
a larger role as the rate-limiting step in uoride removal using this
adsorbent.
Fluoride adsorption isotherms on the optimized bone char at
different conditions of pH and temperature are reported in Fig. 5.
Overall, the uoride uptakes of bone char decreased with increments of solution pH, while the uoride adsorption capacities
increased with the temperature. These results are consistent with
studies performed by Abe et al. [4], Kawasaki et al. [5] and MedellinCastillo et al. [12], which have reported the performance of different
bone chars for water deuoridation. In particular, solution pH has a
higher impact on the uoride uptakes of bone char than the adsorption temperature. Results of removal experiments indicated that
the uoride adsorption process using bone char is endothermic.
Thermodynamic parameters of uoride adsorption on bone char
were calculated using the results of adsorption isotherms at 20,
30 and 40 C. Specically, the Gibbs free energy (G0 , kJ/mol) was
determined using
G0 = RT ln Ka

(3)

Table 2
Kinetic data of uoride adsorption from water on the bone char synthetized at optimal pyrolysis conditions.
Model

Parameter

10

30

50

R
k, g/mg min
qte , mg/g
h = q2te k

0.99
2.94E03
3.38
3.36E02

0.98
8.02E04
7.12
4.07E02

0.99
1.19E03
7.78
7.20E02

R2
kd , mg/g min0.5
C

0.90
0.06
1.28

0.95
0.16
1.38

0.90
0.14
2.87

Pseudo-second order
kinetic: qt =

q2 kt

te
1+qte kt

Intraparticle diffusion:
qt = kd t0.5 + C

[F ]0 , mg/L

14

C.K. Rojas-Mayorga et al. / Journal of Analytical and Applied Pyrolysis 104 (2013) 1018

a)

3.5

7
6
qe, mg/g

2.5

qt, mg/g

a)

5
4
3

1.5

pH
6
7
8

0.5

0
0

10 mg/L

10

20

30

0
0

10

15 20 25 30
t 0.5, min 0.5

35

40
50
[F-]e, mg/L

60

70

80

40
b)

10
9
8

b)

7
qe, mg/g

6
qt, mg/g

6
5
4

30
40

0
0

10

20

30 mg/L

0
5

10

15

20

25

30

35

40

t 0.5, min 0.5

qt, mg/g

20

50 mg/L
5

R
S0

H 0
RT

40
50
[F-]e, mg/L

60

70

80

Fig. 5. Fluoride adsorption isotherms at different conditions of pH and temperature


using bone char synthetized at the optimal pyrolysis conditions.

3.3. Physicochemical characterization of bone char obtained at


optimal pyrolysis conditions and its relationship with uoride
removal mechanism

10 15 20 25 30 35 40
t 0.5, min 0.5

Fig. 4. Linear analysis of uoride adsorption kinetic data at pH 7 and 30 C using


bone char synthetized at the optimal pyrolysis conditions.

S 0

30

the adsorption equilibrium constant, respectively. The adsorption


equilibrium constant Ka was obtained from data tting of uoride adsorption isotherms using Langmuir model [20]. Results of
these thermodynamic calculations are reported in Table 3. Thermodynamic calculations conrmed that the uoride adsorption on
bone char is endothermic and spontaneous. The positive values
of standard entropy and enthalpy changes indicate a strong afnity of the bone char to uoride ions. These results are consistent
to those ndings of Abe et al. [4] and Kawasaki et al. [5]. Finally,
Langmuir and Freundlich models were used for tting the uoride
adsorption isotherms obtained at tested experimental conditions,
see Table 4. Langmuir model provides the best correlation results
for uoride adsorption isotherms using bone char and the monolayer adsorption capacities ranged from 1.39 to 8.96 mg/g at tested
experimental conditions.

c)

8
7
6
5
4
3
2
1
0
0

ln Ka =

T, C

(4)
H0

is the standard entropy change in kJ/mol K,


is
where
the standard enthalpy change in kJ/mol, R is the gas universal
constant (8.3145E03 kJ/mol K), T is temperature in K and Ka is

Fig. 6 shows the TG and DTG curves of bovine bone used as precursor for the synthesis of bone chars. There is a weight loss (10%)
Table 3
Thermodynamic parameters of the uoride adsorption from water on the bone char
synthetized at optimal pyrolysis conditions.
T, C

H , kJ/mol

S , kJ/mol K

20
30
40

42.80

146.76

G , kJ/mol
0.02
2.11
2.93

C.K. Rojas-Mayorga et al. / Journal of Analytical and Applied Pyrolysis 104 (2013) 1018

15

Table 4
Results of data modeling of uoride adsorption isotherms on the bone char samples obtained from pyrolysis process.
Conditions

Adsorbent

Results of data tting for isotherm model

T, C

pH

Langmuir

Freundlich

7
7
7
7
7
7
7
6
8

30
30
30
30
30
20
40
30
30

Bone char at 650 C


Bone char at 700 C
Bone char at 800 C
Bone char at 900 C
Bone char at 1000 C
Bone char at 700 C
Bone char at 700 C
Bone char at 700 C
Bone char at 700 C

Ka , L/mg

qml , mg/g

R2

KF , mg/g

0.97
0.98
0.99
0.99
0.98
0.98
0.92
0.98
0.92

0.23
0.31
0.13
0.11
0.15
0.13
0.52
0.25
0.47

6.88
7.61
7.69
3.53
1.39
7.12
8.43
8.96
4.70

0.89
0.87
0.88
0.85
0.87
0.87
0.95
0.94
0.90

1.92
2.27
1.42
0.64
0.39
1.31
3.22
2.29
1.97

0.31
0.30
0.40
0.39
0.29
0.42
0.25
0.36
0.22

100
TG
TGA
DTGA
dTG

0.14
0.13

90

0.11

85

0.10

80

0.08
0.07

75

DTG

Weight, %

95

0.05
70

0.04

65

0.02

60
0

100

200

300

400 500 600


T, C

700

800

0.01
900 1000

Fig. 6. Results of TGA in Ar of raw bone sample.

from room temperature to 200 C, which is related to the loss of


moisture. From 200 to 600 C, there is a continuous weight loss (i.e.,
18%) caused by the thermal degradation of organic substances such
as proteins, collagen and fat tissues [21,22]. Finally, a weight loss of
6% was observed between 600 and 1000 C, which is mainly caused
by the partial dehydroxylation process of hydroxyapatite and the
decomposition of carbonates [2123]. In the dehydroxylation process, the hydroxyapatite loses OH during thermal treatment and
this process is given by [24]
Ca10 (PO4 )6 (OH)2 Ca10 (PO4 )6 (OH)(22x) Ox x + xH2 O

(5)

where  is a vacancy. This dehydroxylation process may begin at


600800 C depending on the operating conditions of the thermal
treatment [9,24].
Fig. 7 shows the morphology of bone char samples obtained at
650, 700 and 800 C. Overall, these samples have a similar morphology, which is rough without a dened geometry. Table 5 shows the
results of elemental composition and EDX analysis for both raw
bone and bone char samples obtained at different pyrolysis temperatures. Overall, the EDX analysis indicates that the raw bone
is mainly composed of calcium (Ca) and phosphorus (P) and there
are also some minor components such as oxygen (O) and carbon

(C). On the other hand, the calculated molar ratio Ca/P of bone
char samples, which can be associated to the crystallinity index,
decreased with the pyrolysis temperature used for the synthesis of
these adsorbents. In general, the Ca/P ratios ranged from 2.46 to
2.06 for bone char samples obtained at 650 and 1000 C, respectively. These Ca/P ratios are higher than the stoichiometric value of
1.67 and are in agreement with the results reported in other studies
[21,23]. Particularly, the presence of carbonate in the hydroxyapatite tends to decrease the crystallinity index [25].
FT-IR spectra of raw bone and bone char samples obtained at
different pyrolysis conditions are reported in Fig. 8. Specically,
Fig. 8a shows the FT-IR spectrum of the raw precursor where
the band at 3400 cm1 corresponds to the stretching vibration
of O H [23] and the band at 2930 cm1 is assigned to the CH
stretching vibration [25]. Additionally, the bands of groups C O,
C C and C N corresponding to the organic phase of the bone
matrix are identied at 14651744 cm1 [25]. The band of the phosphate group PO4 3 was observed at 1030 cm1 , the peaks of the
carbonate group CO3 2 are evident at 870, 1410 and 1445 cm1
[25,26]. On the other hand, Fig. 8bf shows the FT-IR spectra of
the bone char samples obtained at different temperature of pyrolysis. Overall, FT-IR spectra of bone char has 9 bands at 3420,
1620, 1450, 1445, 1410, 1030, 870, 600 and 565 cm1 , respectively.
These bands correspond to the structure of natural hydroxyapatite (i.e., non-stoichiometric), which contains carbonate groups
[23,25]. The main differences between FT-IR spectra of bone char
samples obtained at different pyrolysis temperatures are related
to the bands of OH and CO3 2 groups; see Fig. 8bf. Therefore,
the dehydroxylation process of hydroxyapatite and the loss of carbonate groups of bone chars caused that the absorption bands
of these groups disappear with the increment of the pyrolysis
temperature used for the synthesis of adsorbents especially at
>800 C. Similar ndings have been also reported by other studies using bone calcination process [21,23,25]. Note that the bands
corresponding to the phosphate group remain constant regardless
of the pyrolysis temperature used for the bone char synthesis.
Fig. 9 reports the XRD patterns of raw bone and bone char samples. XRD patterns conrmed that the diffraction peaks correspond

Table 5
Results of elemental analysis and EDX of the raw precursor and bone char samples.
Sample

Elemental analysis, wt%

EDX, wt%

Ca

Raw bone
Bone char at 650 C
Bone char at 700 C
Bone char at 800 C
Bone char at 900 C
Bone char at 1000 C

11.96
6.28
6.14
5.62
5.04
4.7

2.04
0.67
0.62
0.53
0.13
0.03

3.66
0.87
0.94
0.91
1.02
1.05

0.06
0.02
0.04
0.03
0.04
0.03

31.72
26.16
31.12
27.83
31.04
33.20

16.65
18.97
18.16
18.79
19.72
18.24

40.31
46.69
40.57
44.66
39.81
37.72

16

C.K. Rojas-Mayorga et al. / Journal of Analytical and Applied Pyrolysis 104 (2013) 1018

Fig. 7. SEM images of bone chars obtained at different pyrolysis temperatures: (a, b) 650 C, (c, d) 700 C and (e, f) 800 C. Heating rate: 10 C/min and residence time of 2 h.

to the crystal structure of hydroxyapatite in all samples. Note that


XRD patterns showed a gradual increase in the intensity of peaks
for those bone char samples obtained at high pyrolysis temperature. The differences of XRD patterns (i.e., their narrowness and
shape) suggest that the analyzed samples show different degrees
of crystallinity. These ndings are in agreement to results reported
by Younesi et al. [23]. In addition, the dehydroxylation of the
hydroxyapatite may also cause a shifting in the peaks of XRD
patterns [24].
Finally, the textural parameters of selected bone char samples obtained at 650, 700 and 800 C are reported in Table 6.
The specic surface areas of selected bone chars is relatively
low and they do not show signicant differences. The bone char
obtained from the optimized pyrolysis process showed the following main textural parameters: specic surface area = 110 m2 /g, total

pore volume = 0.233 cm3 /g, and pore size = 8.47 nm. The volume of
mesopores represents approximately 81.5% of the total pore volume of this adsorbent. The textural properties of commercial bone
char have been reported in previous studies and the values obtained
in this study are consistent to the results reported in the literature
[12,16].
In summary, the characterization results suggest that the dehydroxylation process of the hydroxyapatite with temperature should
play an important role for determining the uoride adsorption
properties of bone char. The hydroxyl groups present on bone chars
are thought to be in the uoride removal using bone char according
to the reaction [27,28]
Ca10 (PO4 )6 (OH)2 + 2F Ca10 (PO4 )6 F2 + 2OH

(6)

Table 6
Textural parameters of selected bone chars prepared at different temperatures of pyrolysis.
Pyrolysis temperature, C

Textural parameters of bone char


SBET , m2 /g

650
700
800

118
110
96

VTotal , cm3 /g

VDR , cm3 /g

VMesopore , cm3 /g

Pore size, nm

0.240
0.233
0.224

0.046
0.043
0.036

0.194
0.190
0.188

8.13
8.47
9.33

C.K. Rojas-Mayorga et al. / Journal of Analytical and Applied Pyrolysis 104 (2013) 1018

Then, by reducing the OH groups decreases the exchange


between ions OH and F , directly affecting the adsorption capacity of the bone char obtained from a pyrolysis process. Therefore,
the synthesis of bone char at 700 C helps to reduce the effect of the
hydroxyapatite dehydroxylation on its uoride adsorption properties.

f)

e)

4. Conclusions

d)

Transmittance, (%)

17

c)

b)

a)

4000 3600 3200 2800 2400 2000 1600 1200

800

400

This study has performed a systematic analysis and the optimization of the pyrolysis conditions for the synthesis of bone char
for uoride removal from water. Pyrolysis temperature is a critical operating parameter for the synthesis of bone char for uoride
removal from water. Specically, this temperature has a major
effect on the uoride adsorption properties of bone char due to
the dehydroxylation process of the hydroxyapatite contained in
this adsorbent. An optimum uoride uptake of 7.32 mg/g can be
obtained if a pyrolysis temperature of 700 C is used for bone
char synthesis. The uoride removal performance of the optimized
bone char is better than those reported for several commercial
bone chars up to 63%. Fluoride adsorption isotherms on bone char
were well tted using the Langmuir model and the thermodynamic parameters of adsorption process indicated that the uoride
removal from water using this adsorbent is a spontaneous and
endothermic process. Finally, this study highlights the relevance
of optimizing the pyrolysis conditions for the synthesis of effective
adsorbents for water and wastewater treatment.

Wavenumber, cm-1
References
Fig. 8. FT-IR spectra of raw precursor and bone chars obtained at different pyrolysis
temperatures. Sample: (a) raw bone, (b) 650 C, (c) 700 C, (d) 800 C, (e) 900 C and
(f) 1000 C.

Ca10 (PO4)6(OH)2
f)

e)

Intensity, a.u.

d)

c)

b)

a)

10

20

30

40

50

60

70

80

,
Fig. 9. XRD patterns of raw precursor and bone chars obtained at different pyrolysis
temperatures. Sample: (a) raw bone, (b) 650 C, (c) 700 C, (d) 800 C, (e) 900 C and
(f) 1000 C.

[1] K. Mahapatra, D.S. Ramteke, L.J. Paliwal, Production of activated carbon from
sludge of food processing industry under controlled pyrolysis and its application for methylene blue removal, J. Anal. Appl. Pyrol. 95 (2012) 7986.
[2] C. Wu, M. Song, B. Jin, Y. Wu, Z. Zhong, Y. Huang, Adsorption of sulfur dioxide
using nickel oxide/carbon adsorbents produced by one-step pyrolysis method,
J. Anal. Appl. Pyrol. 99 (2013) 137142.
[3] J. Albertus, H. Bregnhoj, M. Kongpun, Bone char quality and deuoridation
capacity in contact precipitation, in: Proceedings of 3rd International Workshop on Fluorosis Prevention and Deuoridation of Water, 2010, pp. 6172.
[4] I. Abe, S. Iwasaki, T. Tokimoto, N. Kawasaki, T. Nakamura, S. Tanada, Adsorption
of uoride ions onto carbonaceous materials, J. Colloid Interface Sci. 275 (2004)
3539.
[5] N. Kawasaki, F. Ogata, H. Tominaga, I. Yamaguchi, Removal of uoride ion by
bone char produced from animal biomass, J. Oleo Sci. 58 (2009) 529535.
[6] R. Leyva-Ramos, J. Rivera-Utrilla, N.A. Medellin-Castillo, M. Sanchez-Polo,
Kinetic modeling of uoride adsorption from aqueous solution onto bone char,
Chem. Eng. J. 158 (2010) 458467.
[7] M. Mohapatra, S. Anand, B.K. Mishra, D.E. Giles, P. Singh, Review of uoride
removal from drinking water, J. Environ. Manage. 91 (2009) 6777.
[8] A. Bhatnagar, E. Kumar, M. Sillanpaa, Fluoride removal from water adsorption
a review, Chem. Eng. J. 171 (2011) 811840.
[9] J.C. Moreno-Pirajn, R. Gmez-Cruz, V.S. Garca-Cuello, L. Giraldo, Binary system Cu(II)/Pb(II) adsorption on activated carbon obtained by pyrolysis of cow
bone study, J. Anal. Appl. Pyrol. 89 (2010) 122128.
[10] V. Hrnandez-Montoya, L.A. Ramrez-Montoya, A. Bonilla-Petriciolet, M.A.
Montes-Morn, Optimizing the removal of uoride from water using new carbons obtained by modication of nut shell with a calcium solution from egg
shell, Biochem. Eng. J. 62 (2012) 17.
[11] V. Hrnandez-Montoya, M.P. Elizalde-Gonzlez, R. Trejo-Vzquez, Screening
of commercial sorbents for removal of uoride in synthetic and groundwater,
Environ. Technol. 28 (2007) 595607.
[12] N.A. Medellin-Castillo, R. Leyva-Ramos, R. Ocampo-Perez, R.F. Garcia de la Cruz,
J.M. Martinez-Rosales, R.M. Guerrero-Coronado, L. FuentesA. Aragon-Pina,
Rubio, Adsorption of uoride from water solution on bone char, Ind. Eng. Chem.
Res. 46 (2007) 92059212.
[13] D.L. Mwaniki, Fluoride sorption characteristics of different grades of bone charcoal based on batch tests, J. Dent. Res. 71 (1992) 13101315.
[14] P. Phantumvanit, R.Z. LeGeros, Characteristics of bone char related to efcacy of
uoride removal from highly-uoridated water, Fluoride 30 (1997) 207218.
[15] M.E. Kaseva, Optimization of regenerated bone char for uoride removal in
drinking water: a case study in Tanzania, J. Water Health 4 (2006) 139147.
[16] R. Tovar-Gmez, M.R. Moreno-Virgen, J.A. Dena-Aguilar, V. HernndezMontoya, A. Bonilla-Petriciolet, M.A. Montes-Morn, Modeling of xed-bed
adsorption of uoride on bone char using a hybrid neural network approach,
Chem. Eng. J. 228 (2013) 10981109.

18

C.K. Rojas-Mayorga et al. / Journal of Analytical and Applied Pyrolysis 104 (2013) 1018

[17] O. Ioannidou, A. Zabaniotou, Agricultural residues as precursors for activated carbon production a review, Renew. Sustain. Energy Rev. 11 (2007)
19662005.
[18] A.R. Mohamed, M. Mohammadi, G.N. Darzi, Preparation of carbon molecular
sieve from lignocellulosic biomass: a review, Renew. Sustain. Energy Rev. 14
(2010) 15911599.
[19] Standard Methods for Examination of Water and Wastewater, 20th edition,
American Public Health Association, 1998.
[20] Y.S. Ho, A.E. Ofomaja, Kinetics and thermodynamics of lead ion sorption on palm
kernel bre from aqueous solution, Process Biochem. 40 (2005) 34553461.
[21] C.Y. Ooi, M. Hamdi, S. Ramesh, Properties of hydroxyapatite produced by
annealing of bovine bone, Ceram. Int. 33 (2007) 11711177.
[22] A. Sobczak-Kupiec, Z. Wzorek, The inuence of calcination parameters on
free calcium oxide content in natural hydroxyapatite, Ceram. Int. 38 (2012)
641647.

[23] M. Younesi, S. Javadpour, M.E. Bahrololoom, Effect of heat treatment temperature on chemical compositions of extracted hydroxyapatite from bovine bone
ash, J. Mater. Eng. Perform. 20 (2011) 14841490.
[24] P.E. Wang, T.K. Chaki, Sintering behavior and mechanical properties of hydroxyapatite and dicalcium phosphate, J. Mater. Sci. Mater. Med. 4 (1993) 150158.
[25] M. Figueiredo, A. Fernando, G. Martins, J. Freitas, F. Judas, H. Figueiredo, Effect
of the calcination temperature on the composition and microstructure of
hydroxyapatite derived from human and animal bone, Ceram. Int. 36 (2010)
23832393.
[26] S. Lurtwitayapont, T. Srisatit, Comparison of lead removal by various types of
swine bone adsorbents, EnvironmentAsia 3 (2010) 3238.
[27] X. Fan, D.J. Parker, M.D. Smith, Adsorption kinetics of uoride on low cost
materials, Water Res. 37 (2003) 49294937.
[28] S.M. Rao, B.V.V. Reddy, S. Lakshmikanth, N.S. Ambika, Re-use of uoride contaminated bone char sludge in concrete, J. Hazard. Mater. 166 (2009) 751756.

You might also like