Você está na página 1de 7

Microporous and Mesoporous Materials 209 (2015) 3844

Contents lists available at ScienceDirect

Microporous and Mesoporous Materials


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

A new synthesis route for bone chars using CO2 atmosphere


and their application as fluoride adsorbents
C.K. Rojas-Mayorga a, J. Silvestre-Albero b, I.A. Aguayo-Villarreal a, D.I. Mendoza-Castillo a,
A. Bonilla-Petriciolet a,
a
b

Laboratorio de Ingeniera y Tecnologa del Agua, Instituto Tecnolgico de Aguascalientes, C.P. 20256 Aguascalientes, Mexico
Laboratorio de Materiales Avanzados, Departamento de Qumica Inorgnica-Instituto Universitario de Materiales, Universidad de Alicante, Apartado 99, E-3080 Alicante, Spain

a r t i c l e

i n f o

Article history:
Received 7 June 2014
Accepted 1 September 2014
Available online 16 September 2014
Keywords:
CO2
Bone char
Carbonization
Fluoride adsorption

a b s t r a c t
This study describes a new synthesis route for bone chars using a CO2 atmosphere and their behavior as
adsorbent for fluoride removal from water. Specifically, we have performed a detailed analysis of the
adsorption properties of bone char samples obtained at different carbonization conditions and a comparative study with samples of bone char obtained via pyrolysis under nitrogen. Experimental results show
that the nature of the gas atmosphere (CO2 versus N2) and the carbonization temperature play a major
role to achieve an effective bone char for water defluoridation. In particular, the best adsorption properties of bone char for fluoride removal are obtained with those samples synthesized at 700 !C. Carbonization temperatures above 700 !C under CO2 atmosphere cause the dehydroxylation of the hydroxyapatite
in the bone char, thus reducing its fluoride adsorption capacity. The maximum fluoride adsorption capacity for the bone char obtained in this study under CO2 atmosphere (i.e., 5.92 mg/g) is higher than those
reported for commercial bone chars.
" 2014 Elsevier Inc. All rights reserved.

1. Introduction
Nowadays, the research and development of low cost materials
for the mitigation of environmental pollution have gained interest.
Recently, the attention has been paid to the application of materials
such as agricultural and industrial wastes, since the cost of these
materials is lower than the cost of commercial adsorbents, such
as ion-exchange resins or activated carbon [1]. Extensive research
has been carried out during the last decade to find low-cost materials and high capacity adsorbents for the removal of water pollutants. In this context, a wide variety of adsorbents has been
developed and tested and they include natural and synthetic polymers, zeolites, clays, activated aluminas, ashes, biomasses, industrial and agro-industrial wastes and several activated carbons,
among others [211]. Other material with potential for adsorption
is the bone char [1]. This material has acquired relevance in the
treatment of wastewater due to its versatility and economic advantages over other adsorbents. The bone char is a relatively inexpensive adsorbent since it can be obtained as waste from the food
industry. Globally, more than 60 million ton of beef are produced
annually and 58 million of them are consumed [12]. In particular,
Corresponding author.

E-mail address: petriciolet@hotmail.com (A. Bonilla-Petriciolet).

http://dx.doi.org/10.1016/j.micromeso.2014.09.002
1387-1811/" 2014 Elsevier Inc. All rights reserved.

Mexico contributes 3 million to this total, being one of the 7 countries with greater production and consumption of cattle [12]. Animal bones are composed about 70% of inorganic matter, mainly
hydroxyapatite. The chemical composition of hydroxyapatite is
Ca10(PO4)6(OH)2. The remaining part of bones is composed of
organic matter, mainly fibrous protein such as collagen and osteocalcin. Also, the bone char is an adsorbent, which is mainly composed of calcium phosphate as hydroxyapatite (7076%), carbon
(911%) and calcite (79%) [1315]. It is important to highlight that
the organic fraction of bone char is primarily linked to the property
of adsorbing nonpolar organic species, while the inorganic component provides the ability to adsorb ions [1618]. Bone char has been
applied as a versatile adsorbent for a wide variety of pollutants,
including dyes, heavy metals, arsenate and fluoride [19,20]. Therefore, the production of bone char from animal bones sub-products
serves a double purpose. First, it converts unwanted surplus meat
industry waste, of which billions of kilograms are produced annually, to useful value-added adsorbents. Second, the bone chars are
increasingly used in the removal of water pollutant. However, the
removal efficiency and the quality of treated water depend largely
on the carbonization conditions of bone char. Under this context,
this paper reports the evaluation of the synthesis of bone chars
under partially oxidative conditions using a CO2 atmosphere
compared to the traditional pyrolysis process. This study aims to

39

C.K. Rojas-Mayorga et al. / Microporous and Mesoporous Materials 209 (2015) 3844

evaluate the nature of the gas atmosphere in the synthesis process


of the bone char and its adsorption behavior in the removal of fluoride from water.
2. Experimental
2.1. Synthesis of bone chars via CO2
Cow femur bone was used as precursor for the synthesis of bone
chars. The preparation and conditioning of the raw material was
performed according to the methodology proposed in the literature
by Rojas-Mayorga et al. [20]. Bone char synthesis was performed
by a carbonization process using a partially oxidative atmosphere
of CO2 (400 mL/min) in a tubular furnace Carbolite Eurotherm
CTF 12165/550 with a quartz sample holder. Samples of bone char
were synthesized at specific conditions of heating rate, carbonization temperature and duration of thermal treatment, which were
defined using a full factorial design NK (see Table 1). All synthesized bone char samples were washed with deionized water until
constant pH and dried overnight before their use in fluoride
adsorption experiments.
2.2. Fluoride adsorption experiments
For all samples established in the experimental design (Table 1),
the product yield and the fluoride adsorption properties of bone
char samples were determined. Particularly, the adsorbed amount
of fluoride on bone chars was used as the response variable of this
experimental design. The amount of fluoride adsorbed on the different samples was calculated using the data obtained from batch
adsorptions experiments, which were performed by triplicated and
using an initial fluoride concentration of 60 mg/L. The experimental conditions were 30 !C, pH 7, equilibrium time of 24 h and an
adsorbent dosage of 2 g/L. Fluoride uptake of bone chars (mg/g)
were calculated using a mass balance:

qF !

! !
"
F %0 ! F ! %f
V
m

where [F ]0 and [F ]f are the initial and final fluoride concentration


in the adsorption experiments given in mg/L, m is the mass in g of
bone char and V is the volume of fluoride solution given in L, respectively. Fluoride concentration in solution was quantified using a
!

selective electrode and TISAB chemical reagent according to the


procedure described in the Standard Methods of Examination of
Water and Wastewater [21]. Statistica# software was used for data
analysis of the results obtained from this experimental design. This
statistical analysis was performed to determine the effect of carbonization conditions in the adsorption behavior of bone char for fluoride removal. This analysis was also used to identify those operating
parameters of the carbonization process that improve the adsorption behavior of this adsorbent for water defluoridation. Fluoride
adsorptions isotherms were obtained at a concentration range of
580 mg/L, 30 !C, pH 7 and equilibrium time of 24 h. These conditions were used to achieve the saturation of the adsorbent and to
determine the maximum adsorption capacity of selected samples
of bone chars.
2.3. Characterization of raw bone and bone chars
Several characterization techniques were used for determining
the most relevant physicochemical properties of bone char samples obtained in this study. Specifically, the textural parameters
of synthesized samples were determined using N2 adsorption
desorption isotherms at !196 !C using a home-made fully automated equipment designed and constructed by the Advanced
Materials group (LMA), now commercialized as N2 Gsorb-6 (Gas
to Materials Technologies; www.g2mtech.com). The functional
groups were determined using a Transmission FT-IR spectra (KBr)
recorded on a Bruker IFS 66/S spectrophotometer and the sample
was analyzed together with spectroscopic grade KBr, where 200
scans and resolutions of 4 cm!1 were used. The crystalline structure of bone char sample was analyzed using an X-ray diffractometer Bruker D8-Advance with mirror Goebel having a tube with
copper anode RX and radiation Cu Ka (k = 1.5406 ). The diffractograms were obtained on a range of 10 6 2h 6 80. The database
available is from ICDD (International Center for Diffraction Data).
The average crystal size can be estimated semiquantitatively by
the Scherrer equation:

Kk
Dcosh

where b is the average crystal size, K is the crystal form factor (0.9),
k is the wavelength of the X radiation used given in , D is the width
at half height of the diffraction peak at 2h! and h is the angle of the

Table 1
Experimental design used for the synthesis of bone char via CO2. Gas flow: 400 mL/min.
Sample name

Carbonization conditions
Temperature (!C)

C-BC1
C-BC2
C-BC3
C-BC4
C-BC5
C-BC6
C-BC7
C-BC8
C-BC9
C-BC10
C-BC11
C-BC12
C-BC13
C-BC14
C-BC15
C-BC16
C-BC17
C-BC18
C-BC19
C-BC20
Calcined (air)

650

700

800

900

1000

700

Bone char performance

Heating rate (!C/min)

Residence time (h)

Color samples

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
10

2
4
2
4
2
4
2
4
2
4
2
4
2
4
2
4
2
4
2
4
6

Black
Black
Black
Black
Black
Black
Black
Black
Grey
Grey
Grey
Grey
White
White
White
White
White
White
White
White
White/grey

76.41
73.14
75.47
75.57
76.49
75.87
73.97
73.45
70.61
68.28
68.02
68.56
69.51
67.19
66.24
67.61
69.08
68.68
66.60
66.07
68.04

5.52 0.08
5.44 0.06
5.33 0.14
5.45 0.06
5.78 0.05
5.88 0.17
5.92 0.03
5.72 0.09
0.74 0.13
0.71 0.02
0.81 0.02
0.67 0.08
0
0
0
0
0
0
0
0
0.49 0.01

C.K. Rojas-Mayorga et al. / Microporous and Mesoporous Materials 209 (2015) 3844

diffraction peak. Finally, X-ray photoelectron spectroscopy (XPS)


measurements were used to estimate the fluorine content at the
surface of the bone char samples. This analysis was performed with
a Prevac photoelectron spectrometer equipped with a hemispherical analyzer (VG SCIENTA R3000). Spectra were taken using a monochromatized aluminum source Al Ka (E = 1486.6 eV). The base
pressure in the analytical chamber was 5 ' 10!9 mbar. The binding
energy signal was calibrated using the Au 4f7/2 line of a cleaned
gold sample at 84.0 eV. Finally, the surface composition was analyzed taking into account the areas and binding energies of F 1s core
level.

8
7
6
5

qe (mg/g)

40

Bone char samples:


N-BC
C-BC3
C-BC7
C-BC11
C-BC15
C-BC19

4
3

3. Results and discussion

3.1. Adsorption studies


1

Table 1 describes the color of the synthesized samples, the


product yields and the fluoride adsorption capacities for the bone
chars obtained at different carbonization conditions. Experimental
results show that, as the carbonization temperature is increased, a
transition in the color of the bone char samples synthesized is
observed (see Table 1). The color of the raw bone was light yellow,
but during the heat treatment at temperatures from 650 to 800 !C,
the bone char color changes from black to gray. For temperatures
above 900 !C, samples become white thus suggesting complete
elimination of the organic matter of the material. According to literature, these color changes are associated with the thermal degradation of the organic matrix (i.e., collagen) of bone char with
temperature. On the other hand, darker colors indicate the incomplete removal of organic compounds [22], the color arising from
elemental carbon present in the inorganic structure. Based on
these observations, our results indicate that it is possible to synthesize bone chars under a CO2 atmosphere in the temperature range
650700 !C. The yields of bone char range from 66.07% to 76.49% at
evaluated experimental conditions, thus suggesting that the tested
carbonization conditions do not have a significant impact on the
product yield of bone chars. Concerning the adsorption performance, the best fluoride uptake was 5.92 mg/g for the sample synthesized at 700 !C, using a residence time and heating rate of 2 h
and 10 !C/min, respectively. It is important to highlight that, when
the carbonization temperature is higher than 700 !C, the fluoride
adsorption capacity of the bone chars decreases significantly from
5.92 mg/g to 0 mg/g. In fact, the removal performance for bone
chars synthesized at 800, 900 and 1000 !C is unsatisfactory for
water defluoridation (i.e., adsorption capacity < 0.8 mg/g). This
trend in the adsorption performance is in agreement with the
study of Kawasaki et al. [18], which indicates that bone chars
obtained at higher temperatures may show low fluoride uptakes
due to the thermal degradation of surface functional groups that
must be involved in the fluoride removal process. The results of
fluoride adsorption reported in Table 1 and the statistical analysis
(i.e., variance analysis) of the full factorial experimental design
indicate that the carbonization temperature has the most significant effect on the fluoride adsorption properties of bone chars.
To provide more insight into the role of remaining carbon in the
adsorption behavior of bone chars, the fluoride adsorption capacities of bone char samples obtained via a calcination process using
air are reported in Table 1. In this case, the fluoride uptake was
<0.5 mg/g. At this point it is important to highlight that this low
adsorption capacity may be associated to the change in crystallinity of the adsorbent due to the loss of elemental carbon and to the
dehydroxylation of hydroxyapatite, both processes being highly
sensitive to the atmosphere used in the bone char synthesis.
Fig. 1 shows the fluoride adsorption isotherms for the bone char
samples obtained via CO2 at different carbonization temperatures

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 synthesis conditions.

(i.e., 6501000 !C). For the sake of comparison, the bone char
sample obtained via pyrolysis under N2 (sample N-BC) is included.
The best values of fluoride adsorption capacity are 7.32 and
5.92 mg/g, which correspond to the samples N-BC (using N2) and
C-BC7 (using CO2), respectively. Based on the statistical analysis
of the experimental design, the best conditions for the synthesis
of bone chars via CO2 are: a carbonization temperature of 700 !C,
a heating rate of 10 !C/min and a residence time of 2 h. Bone char
samples obtained via CO2 showed competitive fluoride adsorption
capacities compared with the results of bone char samples via pyrolysis under N2 reported in a previous study by Rojas-Mayorga et al.
[20]. It is convenient to highlight that the best fluoride adsorption
capacity of bone char prepared in this study using a CO2 atmosphere is higher than those reported in the literature for commercial bone chars [15,16]. Furthermore, this value outperforms the
results reported in the literature where the fluoride adsorption
capacities ranged from (1.0 to 3.0 mg/g [17,18]. In summary, the
synthesis of bone chars under a CO2 atmosphere gives rise to competitive adsorbents for fluoride removal from water, where the
obtained materials show an improved adsorption performance for
fluoride than those reported for other commercial bone chars.
3.2. Textural properties
Fig. 2a shows the nitrogen adsorption isotherms for the different bone chars. According to the IUPAC classification, samples
N-BC, C-BC3 and C-BC7 exhibit a type IV isotherm, which is
characteristic of mesoporous materials with 250 nm pore size
[2325]. On the other hand, samples C-BC11, C-BC15 and C-BC19
exhibit a type III isotherm, which corresponds to systems with
no porosity and characterized by a weak adsorbate-adsorbent
interaction [2325]. Pore size distributions (PSDs) calculated with
the QSDFT model are reported in Fig. 2b. In general, PSDs profiles
show broad peaks with pore size >2 nm, which correspond to mesoporous materials. However, N-BC and C-BC7 samples have a better development of mesoporosity with three peaks at 2.2, 3.5 and
4.4 nm, respectively. The textural parameters of selected bone char
samples are reported in Table 2. Results show that the temperature
and synthesis atmosphere of bone char affect substantially the textural properties of samples obtained. Firstly, the N-BC sample
obtained at 700 !C via N2 presents the largest specific surface area

C.K. Rojas-Mayorga et al. / Microporous and Mesoporous Materials 209 (2015) 3844

(a) 180
Bone char samples:
N-BC
C-BC3
C-BC7
C-BC11
C-BC15
C-BC19

Adsorbed amount of N2 (cm /g STP)

160
140
120
100
80
60
40
20
0

0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.0

Relative Pressure (P/P0)

(b) 0.00014

C-BC7

0.00007

dV(w) (cc/A/g)

the N-BC, C-BC3 and C-BC7 samples have some microporosity in


their inorganic structure with micropore volumes (VDR) of 0.04,
0.03 and 0.02 cm3/g, respectively. In contrast, samples C-BC11,
C-BC15 and C-BC19 have values of VDR close to 0, which indicates
the absence of microporosity. This microporosity is located within
the bone char and corresponds to pores formed between hydroxyapatite crystals. Although the micropore volume of the samples is
small, this parameter may be important to improve the interaction
between F- ions and the active sites of the inorganic bone char
structure, thus favoring the removal of this water pollutant. With
regard to the volume of mesopores, N-BC and C-BC7 samples
showed the same value of 0.20 cm3/g. For all other samples, this
value decreases significantly. Interestingly, the volume of mesopores represents approximately 80% of the total pore volume on
these adsorbents. It is noteworthy to mention that the C-BC7 bone
char obtained via CO2 has the best fluoride adsorption properties,
and its textural parameters are similar to the N-BC bone char
obtained via N2. Finally, 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 [15,26].
3.3. Crystalline structure

0.00000
0.0032

C-BC3

0.0016
0.0000
0.0042

N-BC

0.0021
0.0000

41

10

20

30

40

50

Pore Width (A)

Fig. 3 shows the X-ray diffractograms for the raw bone (Fig. 3a)
and the selected samples of bone chars (Fig. 3bg). All samples
have the same crystallographic planes, the analysis confirming that
the diffraction peaks correspond to the crystalline structure of the
hydroxyapatite [Ca10(PO4)6(OH)2] [2732]. However, there are different degrees of crystallinity of the samples with respect to the
carbonization temperature. Samples C-BC3 and C-BC7 obtained
via CO2 at 650 and 700 !C (Fig. 3c and d) have a similar crystalline
structure to the N-BC bone char obtained via N2 (Fig. 3b). These
samples show broader diffraction peaks mainly due to the small
crystal size of the hydroxyapatite as a consequence of the presence
of remaining elemental carbon in the inorganic structure. XRD pro-

Fig. 2. (a) N2 adsorption isotherms at 77 K for bone char samples (filled symbols:
adsorption branch; empty symbols: desorption branch) and (b) pore size distribution obtained with the QSDFT model for selected samples.

Table 2
Crystal size and textural parameters of selected bone chars prepared at different
temperatures of carbonization.
Sample
name

N-BCa
C-BC3
C-BC7
C-BC11
C-BC15
C-BC19
a

Crystal size
(nm)

11.02
11.60
11.62
29.18
38.05
45.37

Textural parameters of bone char samples


SBET
(m2/g)

VTotal
(cm3/g)

VDR
(cm3/g)

VMesopore
(cm3/g)

85
62
69
9
4
2

0.24
0.20
0.23
0.16
0.04
0.02

0.04
0.02
0.03
0.00
0.00
0.00

0.20
0.18
0.20
0.16
0.04
0.02

Optimum bone char synthesized via pyrolysis (nitrogen flow: 400 ml/min).

(SBET) with a value of 85 m2/g. Thereafter, samples C-BC3 and


C-BC7 (which were obtained at 650 and 700 !C via CO2) showed
lower SBET values of 62 and 69 m2/g, respectively. However, the
samples C-BC11, C-B15 and C-BC19 show a significant decline in
this property with values of SBET < 10 m2/g. According to the literature, when the SBET is increased, the adsorbent-adsorbate contact
area increases, thus favoring the adsorption process [2425]. In
this context, the results of fluoride adsorption from water are consistent, since the adsorption capacity of the bone chars is favored
with increasing SBET parameter. Moreover, it was found that only

(g)
(f)
(e)
(d)
(c)
(b)
(a)

Fig. 3. XRD patterns of raw precursor and bone chars obtained at different
synthesis conditions. Sample: (a) raw bone, (b) N-BC, (c) C-BC3, (d) C-BC7,
(e) C-BC11, (f) C-BC15 and (g) C-BC19.

C.K. Rojas-Mayorga et al. / Microporous and Mesoporous Materials 209 (2015) 3844

files for samples obtained via CO2 at higher temperature (Fig. 3eg)
show narrower and more intense diffraction peaks, which indicates that the degree of crystallinity of the sample is favored when
the carbonization temperature is increased. Based on these results,
it can be summarized that changes in the crystallinity of the samples may be associated with the loss of elemental carbon present in
the inorganic structure, i.e. elemental carbon (C) may react with
CO2 used as synthesis atmosphere, with the release of CO.
Consequently, the structure of bone char is free of C atoms after
the thermal treatment, thus favoring the rearrangement of the
hydroxyapatite atoms. The molecular arrangement becomes more
orderly, crystallinity index is favored, and a more crystalline structure is formed with higher density and lesser gaps between the
hydroxyapatite crystals. According to Markovic et al. [33], the
width of the diffraction peaks is inversely proportional to the size
and perfection of the crystal lattice, so that diffraction profiles will
be sharper for crystalline materials constituted by larger crystallites and free of deformations. Herein, it is convenient to highlight
that the diffractograms of Fig. 3bd are very similar among them,
which indicates that the synthesis atmosphere does not affect significantly the crystallinity of the bone chars, but only under these
experimental conditions. However, the synthesis of bone chars via
CO2 accelerates the crystallinity of the samples as compared with
the synthesis of bone char via pyrolysis (under nitrogen atmosphere). This effect can be due to the faster removal of elemental
carbon and dehydroxylation of the hydroxyapatite under CO2,
which may occur already at 700 !C [20,34]. In addition, the average
crystal size increases with a further increase in the carbonization
temperature up to 1000 !C (see Table 2). This effect must be attributed to the aggregation of the hydroxyapatite crystals at high temperature and in the absence of elemental carbon [32].
3.4. Functional groups
Fig. 4 shows the FT-IR spectra of raw bone, bone char sample
obtained via pyrolysis and bone char samples obtained using CO2
at different carbonization temperatures. In general, all spectra
show the 6 characteristic bands of hydroxyapatite centered at
3420, 1620, 1450, 1030, 600 and 565 cm!1. The FT-IR spectrum
of raw bone (Fig. 4a) indicates the presence of OH groups
(30003600 cm!1), CH stretching of hydrocarbon (2853
2923 cm!1), C@O, C@C, C@N vibrations of protein and collagen
(14651744 cm!1), a broad and strong band of phosphate PO3!
4
!1
group (1100900 cm!1), a carbonate CO2!
), and
3 peak (720 cm
an additional peak due to calcium Ca2+ (550610 cm!1). After the
thermal treatment, the protein and collagen bands, the OH group
(30003600 cm!1) and the hydrocarbon CH stretching vibration
(28532923 cm!1) lose their intensity due, mainly, to the degradation of the organic matrix as well as the removal of moisture present in the raw material. For bone char N-BC obtained using
nitrogen (Fig. 4b) and the samples obtained using CO2 (Fig. 4cg),
there is a band at 3420 cm!1 corresponding to a high energy
elongation peak coming from !OH groups on the hydroxyapatite
structure [35]. Note that the absorption band of the !OH group
loses intensity with an increase in the carbonization temperature,
and a drastic change takes place above 700 !C. This effect may be
due to dehydroxylation of hydroxyapatite and its structural change
[22,36]. It is important to highlight that this change in the surface
chemistry of bone chars above 700 !C is associated with the
decrease in the fluoride adsorption capacity. FTIR signal at
1620 cm!1 is assigned to the vibration of quinone C@O, while the
band around 1450 cm!1 is assigned to the stretching vibration of
the carboxyl group CO [37,38]. Moreover, the characteristic bands
of the phosphate group around 600 and 1040 cm!1 are observed.
These bands are assigned to PO symmetric stretching mode
(570600 cm!1) and as a major peak is the PO asymmetric

(g)
(f)
(e)
Transmittance (%)

42

(d)
(c)

(b)
(a)

4000

3600

3200

2000 1600 1200 800

400

-1

Wavenumber (cm )
Fig. 4. FT-IR spectra patterns of raw precursor and bone chars obtained at different
synthesis conditions. Sample: (a) raw bone, (b) N-BC, (c) C-BC3, (d) C-BC7,
(e) C-BC11, (f) C-BC15 and (g) C-BC19.

stretching band (1100960 cm!1) [39,40]. Note that the bands corresponding to the phosphate group are kept regardless of the carbonization temperature. Finally, the absorption band at 565 cm!1
corresponds to the calcium present in the inorganic structure; specifically, this band is assigned to the bond between calcium and the
phosphate group [4042].
The bone char sample obtained via CO2 with the best fluoride
adsorption properties was C-BC7. This sample recovered after the
fluoride adsorptions experiments was labeled as C-BC7-F and
was selected for further characterization.
3.5. Crystalline structure after fluoride adsorption studies
Fig. 5 shows the X-ray diffractograms for samples C-BC7 and
C-BC7-F, before and after fluoride adsorptions experiments, respectively. XRD analysis shows that the crystallinity of the sample
increases with the incorporation of F! on the structure. However,
the crystallinity is affected by the preferential orientation of the
crystallites masking the amorphous phase. In fact, fluorapatite
has a lower solubility and higher thermal stability compared to
hydroxyapatite, and produces highly crystalline samples [10]. A
pure hydroxyapatite has a stoichiometric formula in the form of
Ca10(PO4)6(OH)2. In the fluoride removal experiments, some F! ions
replace part of the OH! ions and forms Ca10(PO4)6Fx(OH)2!x, that is,
the so-called fluoridated hydroxyapatite. In the ideal case, fluorine
replaces all !OH group and thus forms fluorapatite Ca10(PO4)6F2
[43,44]. In a hydroxyapatite, the F/Ca ratio is equal to 0, since there
is no fluorine. In a fluoridated hydroxyapatite, the F/Ca ratio varies
from 0 to <0.2. In a pure fluorapatite, the F/Ca ratio is 0.2, which is
the stoichiometric limit. In our case, the ratio F/Ca = 0.1082 indicating that the crystalline structure formed is the fluoridated hydroxyapatite. It is important to note that the F/Ca ratio of the C-BC7-F
bone char was calculated from the atomic percentage that was
obtained through X-ray photoelectron spectroscopy analysis
(XPS), (see Table 3).

43

C.K. Rojas-Mayorga et al. / Microporous and Mesoporous Materials 209 (2015) 3844
Table 3
XPS surface elemental analysis.
Sample
name

at.%

C-BC7
C-BC7-Fa
C-BC19

Ca

Ca

16.03
15.65
17.55

23.66
23.75
24.94

60.31
58.01
57.51

0.00
2.59
0.00

20.60
20.08
22.07

39.36
39.44
40.58

40.04
38.45
37.35

0.00
2.04
0.00

Bone char C-BC7 after of fluoride removal experiments.

* Fluoridated hydroxyapatite
+ Hydroxyapatite

(b)

Transmittance (%)

wt.%

(a)

Intensity (a.u.)

(b)

*
*

**
+

*
*

+ *
+ ***
++
+ **
** + *

* * ++

4000 3600

3200 2800 2000 1600 1200

800

400

-1

Wavenumber (cm )
Fig. 6. (a) FT-IR spectra of C-BC7 and (b) C-BC7-F after of fluoride removal.

(a)
+ +

+
++

+ ++
+

+
+ +

CaF2

++

684.67 eV

20

30

40

50

60

70

80

Fig. 5. (a) XRD patterns of C-BC7 and (b) C-BC7-F after of fluoride removal.

Intensity (a.u.)

2 ()

3.6. Functional groups after fluoride adsorption studies


Additionally, Fig. 6 shows the FT-IR spectra of samples C-BC7
and C-BC7-F. In these spectra we can identify the change in
the absorption band corresponding to OH groups (3000
3600 cm!1). The band intensity decreases after fluoride adsorption
experiments, and this effect can be occurring due to the ionic
exchange between F! and OH! [44].

3.7. XPS analysis


Finally, F1s XPS spectra of the bone char samples before and
after adsorption C-BC7 and C-BC7-F, respectively, are reported in
Fig. 7. In the bone char C-BC7 before fluoride adsorption experiments the F1s signal does not exist; however, after fluoride adsorption experiments, a new and unique peak is found. This energy
band corresponds to F1s with a binding energy at 684.67 eV.
According to Hwang et al. [45], this F1s peak corresponds to the
fluoride bonded to calcium (CaF2), which is formed between the
calcium of hydroxyapatite and the fluoride present in solution.
Therefore, with this study we confirm that surface reactions exist
between fluoride ions and calcium from hydroxyapatite [4548].
Also, this analysis is in agreement with XRD and FT-IR results.

C-BC7-F
C-BC7
690

689

688

687

686 685 684 683


Binding energy (eV)

682

681

680

Fig. 7. XPS F1s spectrum of C-BC7-F and C-BC7.

4. Conclusions
It is possible to obtain bone char with outstanding fluoride
adsorption properties using CO2 as synthesis atmosphere. Carbonization temperature is a critical operating parameter for the synthesis of bone char via CO2 for fluoride removal from water.
Specifically, this temperature has a major effect on the fluoride
adsorption properties of bone char due to the change in crystallinity of the sample and to the dehydroxylation process of the
hydroxyapatite contained in this adsorbent. Also, the surface area
decreased with increments in the carbonization temperature. The

44

C.K. Rojas-Mayorga et al. / Microporous and Mesoporous Materials 209 (2015) 3844

low crystallinity, higher specific surface area and !OH functional


groups of bone char improve the fluoride adsorption capacity.
The best fluoride uptake of 5.92 mg/g can be obtained if a carbonization temperature of 700 !C is used for bone char synthesis via
CO2. The fluoride removal performance of the bone char is better
than those reported for several commercial bone chars up to 31%.
The bone char C-BC7 obtained via CO2 showed a competitive
performance in the fluoride removal from water (i.e., 5.92 mg/g)
compared with the optimum bone char obtained via pyrolysis
N-BC (i.e., 7.32 mg/g). FT-IR spectroscopy and XPS measurements
proved that fluoride removal by this adsorbent occurred via the
ionic exchange of surface hydroxyl groups with fluoride. F1s XPS
spectra confirmed this theory and the reaction between fluorine
and calcium, while the DRX study indicates the fluoridated
hydroxyapatite formation.
Acknowledgements
Authors acknowledge the financial support provided by
CONACYT, DGEST, Instituto Tecnolgico de Aguascalientes and
Universidad de Alicante.
References
[1] L.R. Brunson, D.A. Sabatini, Sci. Total Environ. 488 (2014) 580587.
[2] N.I. Chubar, V.F. Samanidou, V.S. Kouts, G.G. Gallios, V.A. Kanibolotsky, V.V.
Strelko, Zhuravlev, J. Colloid. Interface Sci. 291 (2005) 6774.
[3] M. Islam, R.K. Patel, J. Hazard. Mater. 143 (2007) 303310.
[4] S.P. Kamble, S. Jagtap, N.K. Labhsetwar, D. Thakare, S. Godfrey, S. Devotta, S.S.
Rayalu, Chem. Eng. J. 129 (2007) 173180.
[5] B. Nagappa, G.T. Chandrappa, Microporous Mesoporous Mater. 106 (2007)
212218.
[6] W. Nigussie, F. Zewge, B.S. Chandravanshi, J. Hazard. Mater. 147 (2007) 954
963.
[7] S.V. Ramanaiah, S.V. Mohan, P.N. Sarma, Ecol. Eng. 31 (2007) 4756.
[8] S. Venkata Mohan, S.V. Ramanaiah, B. Rajkumar, P.N. Sarma, Bioresource.
Technol. 98 (2007) 10061011.
[9] A. Eskandarpour, M.S. Onyango, A. Ochieng, S. Asai, J. Hazard. Mater. 152
(2008) 571579.
[10] N. Hamdi, E. Srasra, Desalination 206 (2007) 238244.
[11] S. Meenakshi, C.S. Sundaram, R. Sukumar, J. Hazard. Mater. 153 (2008) 164
172.
[12] J.A.P. Pea, D.B. Varela, H.J. Severiano, F.G.R. Rincn, G.B. Rodrguez, SAGARPA
5 (2013).
[13] J.A. Wilson, I.D. Pulford, S. Thomas, Environ. Geochem. Health 25 (2003) 51
56.
[14] K.K.H. Choy, G. McKay, Chemosphere 60 (2005) 1141150.

[15] N.A. Medelln-Castillo, R. Leyva-Ramos, R. Ocampo-Prez, R.F. Garca de la


Cruz, A. Aragn-Pia, J.M. Martnez-Rosales, R.M. Guerrero-Coronado, L.
Fuentes-Rubio, Ind. Eng. Chem. Res. 46 (2007) 92059212.
[16] P. Dasgupta, A. Singh, S. Adak, K.M. Purohit, Int. Symp. Res. Stud. Mater. Sci.
Eng. (2004) 16.
[17] I. Abe, S. Iwasaki, T. Tokimoto, N. Kawasaki, T. Nakamura, S. Tanada, J. Colloid
Interface Sci. 275 (2004) 3539.
[18] N. Kawasaki, F. Ogata, H. Tominaga, I. Yamaguchi, J. Oleo Sci. 58 (2009) 529
535.
[19] J.C. Moreno-Pirajn, R. Gmez-Cruz, V.S. Garca-Cuello, L. Giraldo, J. Anal. Appl.
Pyrol. 89 (2010) 122128.
[20] C.K. Rojas-Mayorga, A. Bonilla-Petriciolet, I.A. Aguayo-Villarreal, V. HernndezMontoya, M.R. Moreno-Virgen, R. Tovar-Gmez, M.A. Montes-Morn, J. Anal.
Appl. Pyrol. 104 (2013) 1018.
[21] Standard Methods for Examination of Water and Wastewater, 20th edition,
American Public Health Association, 1998.
[22] C.Y. Ooi, M. Hamdi, S. Ramesh, Ceram. Int. 33 (2007) 11711177.
[23] S.J. Gregg, K.S.W. Sing, Adsorption Surface Area and Porosity, Ed. Academic
Press, second ed., London, 1982.
[24] B.C. Pergher, A. Corma, V. Forns, Quim. Nova 26 (2003) 795802.
[25] F. Rouquerol, J. Rouquerol, K. Sing, Adsorption by Powders and Porous Solids:
Principles Methodology and Applications, Academic Press, San Diego, 1999.
[26] R. Tovar-Gmez, M.R. Moreno-Virgen, J.A. Dena-Aguilar, V. HernndezMontoya, A. Bonilla-Petriciolet, M.A. Montes-Morn, Chem. Eng. J. 228
(2013) 10981109.
[27] L.R. Brunson, D.A. Sabatini, Environ. Eng. Sci. 26 (2009) 17771784.
[28] F.C.M. Driessens, R.M.H. Verbeeck, Biominerals, CRC Press, Florida, 2000.
[29] L. Caldern, M.J. Stott, A. Rubio, Phys. Rev. A 67 (2003) 17.
[30] P. Fernigrini, O.R. Cmara, F.Y. Oliva, Asociacin Argentina de materiales, 2do.
Encuentro de Jvenes Investigadores en Ciencia y Tecnologa de Materiales,
2008.
[31] A.B. Martnez-Valencia, H.E. Esparza-Ponce, G. Carbajal-De la Torre, J. OrtizLanderos, Soc. Mex. Ciencia Tecnol. Superficies Mater. 21 (2008) 1821.
[32] L.G. Sequeda, J.M. Daz, S.G. Gutirrez, S.G. Perdomo, O.L. Gmez, Rev. Colomb.
Ciencia Qumica Farm. 41 (2012) 5066.
[33] M. Markovic, B. Fowler, M. Tung, J. Res. Natl. Inst. Stan. 109 (2004) 553568.
[34] P.E. Wang, T.K. Chaki, J. Mater. Sci. - Mater. Med. 4 (1993) 150158.
[35] F. Miyaji, Y. Kono, Y. Suyama, Mater. Res. Bull. 40 (2005) 209220.
[36] M. Younesi, S. Javadpour, M.E. Bahrololoom, J. Mater. Eng. Perform. 20 (2011)
14841490.
[37] A.A.M. Daifullah, B.S. Girgis, H.M.H. Gad, Mater. Lett. 57 (2013) 17231731.
[38] C.J. Durn-Valle, M. Gmez-Corzo, J. Pastor-Villegas, V. Gmez-Serrano, J. Anal.
Appl. Pyrol. 73 (2005) 5967.
[39] C. Akemi, A.M. De Guzzi, Rev. Bras. Med. 17 (2001) 123130.
[40] S. Lurtwitayapont, T. Srisatit, Environ. Asia 3 (2010) 3238.
[41] J.H.G. Rocha, A.F. Lemos, S. Kannan, S. Agathopoulos, J.M.F. Ferreira, J. Mater.
Chem. 15 (2005) 50075011.
[42] H.K. Varma, S. Babu, Ceram. Int. 31 (2005) 109114.
[43] Y. Chen, L. Chai, Y. Shu, J. Hazard. Mater. 160 (2008) 168172.
[44] S.V. Dorozhkin, Mater. Sci. 2 (2009) 399498.
[45] S.M. Hwang, A. Izumi, K. Tsutsui, S. Furukawa, Appl. Surf. Sci. (1994) 523527.
[46] P. Xiangliang, J. Wang, D. Zhang, Desalination 249 (2009) 609614.
[47] C. Kui, Z. Sam, W. Wenjian, Surf. Coat. Technol. 198 (2005) 237240.
[48] D. Xiaomin, M. Dinesh, U. Charles, J. Pittman, Y. Shuo, Chem. Eng. J. 198199
(2012) 236245.

Você também pode gostar