Вы находитесь на странице: 1из 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 uoride 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
Laboratorio de Ingeniera y Tecnologa del Agua, Instituto Tecnolgico de Aguascalientes, C.P. 20256 Aguascalientes, Mexico
b
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 a b s t r a c t

Article history: This study describes a new synthesis route for bone chars using a CO2 atmosphere and their behavior as
Received 7 June 2014 adsorbent for uoride removal from water. Specically, we have performed a detailed analysis of the
Accepted 1 September 2014 adsorption properties of bone char samples obtained at different carbonization conditions and a compar-
Available online 16 September 2014
ative 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
Keywords: role to achieve an effective bone char for water deuoridation. In particular, the best adsorption proper-
CO2
ties of bone char for uoride removal are obtained with those samples synthesized at 700 C. Carboniza-
Bone char
Carbonization
tion temperatures above 700 C under CO2 atmosphere cause the dehydroxylation of the hydroxyapatite
Fluoride adsorption in the bone char, thus reducing its uoride adsorption capacity. The maximum uoride adsorption capac-
ity 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 Mexico contributes 3 million to this total, being one of the 7 coun-
tries with greater production and consumption of cattle [12]. Ani-
Nowadays, the research and development of low cost materials mal bones are composed about 70% of inorganic matter, mainly
for the mitigation of environmental pollution have gained interest. hydroxyapatite. The chemical composition of hydroxyapatite is
Recently, the attention has been paid to the application of materials Ca10(PO4)6(OH)2. The remaining part of bones is composed of
such as agricultural and industrial wastes, since the cost of these organic matter, mainly brous protein such as collagen and osteo-
materials is lower than the cost of commercial adsorbents, such calcin. Also, the bone char is an adsorbent, which is mainly com-
as ion-exchange resins or activated carbon [1]. Extensive research posed of calcium phosphate as hydroxyapatite (7076%), carbon
has been carried out during the last decade to nd low-cost mate- (911%) and calcite (79%) [1315]. It is important to highlight that
rials and high capacity adsorbents for the removal of water pollu- the organic fraction of bone char is primarily linked to the property
tants. In this context, a wide variety of adsorbents has been of adsorbing nonpolar organic species, while the inorganic compo-
developed and tested and they include natural and synthetic poly- nent provides the ability to adsorb ions [1618]. Bone char has been
mers, zeolites, clays, activated aluminas, ashes, biomasses, indus- applied as a versatile adsorbent for a wide variety of pollutants,
trial and agro-industrial wastes and several activated carbons, including dyes, heavy metals, arsenate and uoride [19,20]. There-
among others [211]. Other material with potential for adsorption fore, the production of bone char from animal bones sub-products
is the bone char [1]. This material has acquired relevance in the serves a double purpose. First, it converts unwanted surplus meat
treatment of wastewater due to its versatility and economic advan- industry waste, of which billions of kilograms are produced annu-
tages over other adsorbents. The bone char is a relatively inexpen- ally, to useful value-added adsorbents. Second, the bone chars are
sive adsorbent since it can be obtained as waste from the food increasingly used in the removal of water pollutant. However, the
industry. Globally, more than 60 million ton of beef are produced removal efciency and the quality of treated water depend largely
annually and 58 million of them are consumed [12]. In particular, on the carbonization conditions of bone char. Under this context,
this paper reports the evaluation of the synthesis of bone chars
Corresponding author. under partially oxidative conditions using a CO2 atmosphere
E-mail address: petriciolet@hotmail.com (A. Bonilla-Petriciolet). compared to the traditional pyrolysis process. This study aims to

http://dx.doi.org/10.1016/j.micromeso.2014.09.002
1387-1811/ 2014 Elsevier Inc. All rights reserved.
C.K. Rojas-Mayorga et al. / Microporous and Mesoporous Materials 209 (2015) 3844 39

evaluate the nature of the gas atmosphere in the synthesis process selective electrode and TISAB chemical reagent according to the
of the bone char and its adsorption behavior in the removal of uo- procedure described in the Standard Methods of Examination of
ride from water. Water and Wastewater [21]. Statistica software was used for data
analysis of the results obtained from this experimental design. This
2. Experimental statistical analysis was performed to determine the effect of carbon-
ization conditions in the adsorption behavior of bone char for uo-
2.1. Synthesis of bone chars via CO2 ride removal. This analysis was also used to identify those operating
parameters of the carbonization process that improve the adsorp-
Cow femur bone was used as precursor for the synthesis of bone tion behavior of this adsorbent for water deuoridation. Fluoride
chars. The preparation and conditioning of the raw material was adsorptions isotherms were obtained at a concentration range of
performed according to the methodology proposed in the literature 580 mg/L, 30 C, pH 7 and equilibrium time of 24 h. These condi-
by Rojas-Mayorga et al. [20]. Bone char synthesis was performed tions were used to achieve the saturation of the adsorbent and to
by a carbonization process using a partially oxidative atmosphere determine the maximum adsorption capacity of selected samples
of CO2 (400 mL/min) in a tubular furnace Carbolite Eurotherm of bone chars.
CTF 12165/550 with a quartz sample holder. Samples of bone char
were synthesized at specic conditions of heating rate, carboniza- 2.3. Characterization of raw bone and bone chars
tion temperature and duration of thermal treatment, which were
dened using a full factorial design NK (see Table 1). All synthe- Several characterization techniques were used for determining
sized bone char samples were washed with deionized water until the most relevant physicochemical properties of bone char sam-
constant pH and dried overnight before their use in uoride ples obtained in this study. Specically, the textural parameters
adsorption experiments. of synthesized samples were determined using N2 adsorption
desorption isotherms at 196 C using a home-made fully auto-
2.2. Fluoride adsorption experiments mated equipment designed and constructed by the Advanced
Materials group (LMA), now commercialized as N2 Gsorb-6 (Gas
For all samples established in the experimental design (Table 1), to Materials Technologies; www.g2mtech.com). The functional
the product yield and the uoride adsorption properties of bone groups were determined using a Transmission FT-IR spectra (KBr)
char samples were determined. Particularly, the adsorbed amount recorded on a Bruker IFS 66/S spectrophotometer and the sample
of uoride on bone chars was used as the response variable of this was analyzed together with spectroscopic grade KBr, where 200
experimental design. The amount of uoride adsorbed on the dif- scans and resolutions of 4 cm1 were used. The crystalline struc-
ferent samples was calculated using the data obtained from batch ture of bone char sample was analyzed using an X-ray diffractom-
adsorptions experiments, which were performed by triplicated and eter Bruker D8-Advance with mirror Goebel having a tube with
using an initial uoride concentration of 60 mg/L. The experimen- copper anode RX and radiation Cu Ka (k = 1.5406 ). The diffracto-
tal conditions were 30 C, pH 7, equilibrium time of 24 h and an grams were obtained on a range of 10 6 2h 6 80. The database
adsorbent dosage of 2 g/L. Fluoride uptake of bone chars (mg/g) available is from ICDD (International Center for Diffraction Data).
were calculated using a mass balance: The average crystal size can be estimated semiquantitatively by
   the Scherrer equation:
F 0  F  f
qF  V 1 Kk
m b 2
Dcosh
 
where [F ]0 and [F ]f are the initial and nal uoride concentration
in the adsorption experiments given in mg/L, m is the mass in g of where b is the average crystal size, K is the crystal form factor (0.9),
bone char and V is the volume of uoride solution given in L, respec- k is the wavelength of the X radiation used given in , D is the width
tively. Fluoride concentration in solution was quantied using a 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 ow: 400 mL/min.

Sample name Carbonization conditions Bone char performance


Temperature (C) Heating rate (C/min) Residence time (h) Color samples Yield (%) Fluoride uptake (mg/g)
C-BC1 650 5 2 Black 76.41 5.52 0.08
C-BC2 5 4 Black 73.14 5.44 0.06
C-BC3 10 2 Black 75.47 5.33 0.14
C-BC4 10 4 Black 75.57 5.45 0.06
C-BC5 700 5 2 Black 76.49 5.78 0.05
C-BC6 5 4 Black 75.87 5.88 0.17
C-BC7 10 2 Black 73.97 5.92 0.03
C-BC8 10 4 Black 73.45 5.72 0.09
C-BC9 800 5 2 Grey 70.61 0.74 0.13
C-BC10 5 4 Grey 68.28 0.71 0.02
C-BC11 10 2 Grey 68.02 0.81 0.02
C-BC12 10 4 Grey 68.56 0.67 0.08
C-BC13 900 5 2 White 69.51 0
C-BC14 5 4 White 67.19 0
C-BC15 10 2 White 66.24 0
C-BC16 10 4 White 67.61 0
C-BC17 1000 5 2 White 69.08 0
C-BC18 5 4 White 68.68 0
C-BC19 10 2 White 66.60 0
C-BC20 10 4 White 66.07 0
Calcined (air) 700 10 6 White/grey 68.04 0.49 0.01
40 C.K. Rojas-Mayorga et al. / Microporous and Mesoporous Materials 209 (2015) 3844

diffraction peak. Finally, X-ray photoelectron spectroscopy (XPS) 8


measurements were used to estimate the uorine content at the
surface of the bone char samples. This analysis was performed with 7
a Prevac photoelectron spectrometer equipped with a hemispheri-
cal analyzer (VG SCIENTA R3000). Spectra were taken using a mono-
6
chromatized aluminum source Al Ka (E = 1486.6 eV). The base
pressure in the analytical chamber was 5  109 mbar. The binding
energy signal was calibrated using the Au 4f7/2 line of a cleaned 5
Bone char samples:
gold sample at 84.0 eV. Finally, the surface composition was ana-

qe (mg/g)
N-BC
lyzed taking into account the areas and binding energies of F 1s core 4 C-BC3
level. C-BC7
C-BC11
3
C-BC15
C-BC19
3. Results and discussion
2

3.1. Adsorption studies


1
Table 1 describes the color of the synthesized samples, the
product yields and the uoride adsorption capacities for the bone 0
chars obtained at different carbonization conditions. Experimental 0 10 20 30 40 50 60 70 80
results show that, as the carbonization temperature is increased, a [F-]e (mg/L)
transition in the color of the bone char samples synthesized is
observed (see Table 1). The color of the raw bone was light yellow, Fig. 1. Fluoride adsorption isotherms at pH 7 and 30 C on bone chars obtained at
but during the heat treatment at temperatures from 650 to 800 C, different synthesis conditions.
the bone char color changes from black to gray. For temperatures
above 900 C, samples become white thus suggesting complete (i.e., 6501000 C). For the sake of comparison, the bone char
elimination of the organic matter of the material. According to lit- sample obtained via pyrolysis under N2 (sample N-BC) is included.
erature, these color changes are associated with the thermal degra- The best values of uoride adsorption capacity are 7.32 and
dation of the organic matrix (i.e., collagen) of bone char with 5.92 mg/g, which correspond to the samples N-BC (using N2) and
temperature. On the other hand, darker colors indicate the incom- C-BC7 (using CO2), respectively. Based on the statistical analysis
plete removal of organic compounds [22], the color arising from of the experimental design, the best conditions for the synthesis
elemental carbon present in the inorganic structure. Based on of bone chars via CO2 are: a carbonization temperature of 700 C,
these observations, our results indicate that it is possible to synthe- a heating rate of 10 C/min and a residence time of 2 h. Bone char
size bone chars under a CO2 atmosphere in the temperature range samples obtained via CO2 showed competitive uoride adsorption
650700 C. The yields of bone char range from 66.07% to 76.49% at capacities compared with the results of bone char samples via pyro-
evaluated experimental conditions, thus suggesting that the tested lysis under N2 reported in a previous study by Rojas-Mayorga et al.
carbonization conditions do not have a signicant impact on the [20]. It is convenient to highlight that the best uoride adsorption
product yield of bone chars. Concerning the adsorption perfor- capacity of bone char prepared in this study using a CO2 atmo-
mance, the best uoride uptake was 5.92 mg/g for the sample syn- sphere is higher than those reported in the literature for commer-
thesized at 700 C, using a residence time and heating rate of 2 h cial bone chars [15,16]. Furthermore, this value outperforms the
and 10 C/min, respectively. It is important to highlight that, when results reported in the literature where the uoride adsorption
the carbonization temperature is higher than 700 C, the uoride capacities ranged from 1.0 to 3.0 mg/g [17,18]. In summary, the
adsorption capacity of the bone chars decreases signicantly from synthesis of bone chars under a CO2 atmosphere gives rise to com-
5.92 mg/g to 0 mg/g. In fact, the removal performance for bone petitive adsorbents for uoride removal from water, where the
chars synthesized at 800, 900 and 1000 C is unsatisfactory for obtained materials show an improved adsorption performance for
water deuoridation (i.e., adsorption capacity < 0.8 mg/g). This uoride than those reported for other commercial bone chars.
trend in the adsorption performance is in agreement with the
study of Kawasaki et al. [18], which indicates that bone chars 3.2. Textural properties
obtained at higher temperatures may show low uoride uptakes
due to the thermal degradation of surface functional groups that Fig. 2a shows the nitrogen adsorption isotherms for the differ-
must be involved in the uoride removal process. The results of ent bone chars. According to the IUPAC classication, samples
uoride adsorption reported in Table 1 and the statistical analysis N-BC, C-BC3 and C-BC7 exhibit a type IV isotherm, which is
(i.e., variance analysis) of the full factorial experimental design characteristic of mesoporous materials with 250 nm pore size
indicate that the carbonization temperature has the most signi- [2325]. On the other hand, samples C-BC11, C-BC15 and C-BC19
cant effect on the uoride adsorption properties of bone chars. exhibit a type III isotherm, which corresponds to systems with
To provide more insight into the role of remaining carbon in the no porosity and characterized by a weak adsorbate-adsorbent
adsorption behavior of bone chars, the uoride adsorption capaci- interaction [2325]. Pore size distributions (PSDs) calculated with
ties of bone char samples obtained via a calcination process using the QSDFT model are reported in Fig. 2b. In general, PSDs proles
air are reported in Table 1. In this case, the uoride uptake was show broad peaks with pore size >2 nm, which correspond to mes-
<0.5 mg/g. At this point it is important to highlight that this low oporous materials. However, N-BC and C-BC7 samples have a bet-
adsorption capacity may be associated to the change in crystallin- ter development of mesoporosity with three peaks at 2.2, 3.5 and
ity of the adsorbent due to the loss of elemental carbon and to the 4.4 nm, respectively. The textural parameters of selected bone char
dehydroxylation of hydroxyapatite, both processes being highly samples are reported in Table 2. Results show that the temperature
sensitive to the atmosphere used in the bone char synthesis. and synthesis atmosphere of bone char affect substantially the tex-
Fig. 1 shows the uoride adsorption isotherms for the bone char tural properties of samples obtained. Firstly, the N-BC sample
samples obtained via CO2 at different carbonization temperatures obtained at 700 C via N2 presents the largest specic surface area
C.K. Rojas-Mayorga et al. / Microporous and Mesoporous Materials 209 (2015) 3844 41

(a) 180 the N-BC, C-BC3 and C-BC7 samples have some microporosity in
their inorganic structure with micropore volumes (VDR) of 0.04,
160 Bone char samples: 0.03 and 0.02 cm3/g, respectively. In contrast, samples C-BC11,
N-BC
C-BC15 and C-BC19 have values of VDR close to 0, which indicates
140 C-BC3
the absence of microporosity. This microporosity is located within
Adsorbed amount of N2 (cm /g STP)

C-BC7
C-BC11 the bone char and corresponds to pores formed between hydroxy-
120
C-BC15 apatite crystals. Although the micropore volume of the samples is
3

C-BC19 small, this parameter may be important to improve the interaction


100
between F- ions and the active sites of the inorganic bone char
80 structure, thus favoring the removal of this water pollutant. With
regard to the volume of mesopores, N-BC and C-BC7 samples
60 showed the same value of 0.20 cm3/g. For all other samples, this
value decreases signicantly. Interestingly, the volume of mesop-
40
ores represents approximately 80% of the total pore volume on
20
these adsorbents. It is noteworthy to mention that the C-BC7 bone
char obtained via CO2 has the best uoride adsorption properties,
0 and its textural parameters are similar to the N-BC bone char
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 obtained via N2. Finally, the textural properties of commercial bone
Relative Pressure (P/P0) char have been reported in previous studies and the values
obtained in this study are consistent to the results reported in
(b) 0.00014 C-BC7 the literature [15,26].
0.00007
3.3. Crystalline structure
0.00000
Fig. 3 shows the X-ray diffractograms for the raw bone (Fig. 3a)
dV(w) (cc/A/g)

0.0032
C-BC3
and the selected samples of bone chars (Fig. 3bg). All samples
0.0016 have the same crystallographic planes, the analysis conrming that
the diffraction peaks correspond to the crystalline structure of the
0.0000 hydroxyapatite [Ca10(PO4)6(OH)2] [2732]. However, there are dif-
0.0042
N-BC
ferent degrees of crystallinity of the samples with respect to the
carbonization temperature. Samples C-BC3 and C-BC7 obtained
0.0021 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
0.0000 samples show broader diffraction peaks mainly due to the small
0 10 20 30 40 50
crystal size of the hydroxyapatite as a consequence of the presence
Pore Width (A)
of remaining elemental carbon in the inorganic structure. XRD pro-
Fig. 2. (a) N2 adsorption isotherms at 77 K for bone char samples (lled symbols:
adsorption branch; empty symbols: desorption branch) and (b) pore size distribu-
tion 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 Crystal size Textural parameters of bone char samples (g)


name (nm)
SBET VTotal VDR VMesopore
(m2/g) (cm3/g) (cm3/g) (cm3/g) (f)
N-BCa 11.02 85 0.24 0.04 0.20
C-BC3 11.60 62 0.20 0.02 0.18 (e)
C-BC7 11.62 69 0.23 0.03 0.20
C-BC11 29.18 9 0.16 0.00 0.16
C-BC15 38.05 4 0.04 0.00 0.04 (d)
C-BC19 45.37 2 0.02 0.00 0.02
a
Optimum bone char synthesized via pyrolysis (nitrogen ow: 400 ml/min). (c)

(b)
(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
(a)
lower SBET values of 62 and 69 m2/g, respectively. However, the
samples C-BC11, C-B15 and C-BC19 show a signicant decline in
this property with values of SBET < 10 m2/g. According to the liter-
ature, when the SBET is increased, the adsorbent-adsorbate contact
area increases, thus favoring the adsorption process [2425]. In
this context, the results of uoride adsorption from water are con- Fig. 3. XRD patterns of raw precursor and bone chars obtained at different
sistent, since the adsorption capacity of the bone chars is favored synthesis conditions. Sample: (a) raw bone, (b) N-BC, (c) C-BC3, (d) C-BC7,
with increasing SBET parameter. Moreover, it was found that only (e) C-BC11, (f) C-BC15 and (g) C-BC19.
42 C.K. Rojas-Mayorga et al. / Microporous and Mesoporous Materials 209 (2015) 3844

les for samples obtained via CO2 at higher temperature (Fig. 3eg)
show narrower and more intense diffraction peaks, which indi- (g)
cates 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 sam-
(f)
ples may be associated with the loss of elemental carbon present in
the inorganic structure, i.e. elemental carbon (C) may react with
(e)
CO2 used as synthesis atmosphere, with the release of CO.
Consequently, the structure of bone char is free of C atoms after

Transmittance (%)
the thermal treatment, thus favoring the rearrangement of the (d)
hydroxyapatite atoms. The molecular arrangement becomes more
orderly, crystallinity index is favored, and a more crystalline struc-
ture is formed with higher density and lesser gaps between the (c)
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 proles will
(b)
be sharper for crystalline materials constituted by larger crystal-
lites and free of deformations. Herein, it is convenient to highlight
that the diffractograms of Fig. 3bd are very similar among them, (a)
which indicates that the synthesis atmosphere does not affect sig-
nicantly 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 atmo- 4000 3600 3200 2000 1600 1200 800 400
sphere). This effect can be due to the faster removal of elemental -1
carbon and dehydroxylation of the hydroxyapatite under CO2, Wavenumber (cm )
which may occur already at 700 C [20,34]. In addition, the average
Fig. 4. FT-IR spectra patterns of raw precursor and bone chars obtained at different
crystal size increases with a further increase in the carbonization synthesis conditions. Sample: (a) raw bone, (b) N-BC, (c) C-BC3, (d) C-BC7,
temperature up to 1000 C (see Table 2). This effect must be attrib- (e) C-BC11, (f) C-BC15 and (g) C-BC19.
uted to the aggregation of the hydroxyapatite crystals at high tem-
perature and in the absence of elemental carbon [32].
stretching band (1100960 cm1) [39,40]. Note that the bands cor-
3.4. Functional groups responding to the phosphate group are kept regardless of the car-
bonization temperature. Finally, the absorption band at 565 cm1
Fig. 4 shows the FT-IR spectra of raw bone, bone char sample corresponds to the calcium present in the inorganic structure; spe-
obtained via pyrolysis and bone char samples obtained using CO2 cically, this band is assigned to the bond between calcium and the
at different carbonization temperatures. In general, all spectra phosphate group [4042].
show the 6 characteristic bands of hydroxyapatite centered at The bone char sample obtained via CO2 with the best uoride
3420, 1620, 1450, 1030, 600 and 565 cm1. The FT-IR spectrum adsorption properties was C-BC7. This sample recovered after the
of raw bone (Fig. 4a) indicates the presence of OH groups uoride adsorptions experiments was labeled as C-BC7-F and
(30003600 cm1), CH stretching of hydrocarbon (2853 was selected for further characterization.
2923 cm1), C@O, C@C, C@N vibrations of protein and collagen
(14651744 cm1), a broad and strong band of phosphate PO3 4 3.5. Crystalline structure after uoride adsorption studies
group (1100900 cm1), a carbonate CO2 3 peak (720 cm
1
), and
an additional peak due to calcium Ca2+ (550610 cm1). After the Fig. 5 shows the X-ray diffractograms for samples C-BC7 and
thermal treatment, the protein and collagen bands, the OH group C-BC7-F, before and after uoride adsorptions experiments, respec-
(30003600 cm1) and the hydrocarbon CH stretching vibration tively. XRD analysis shows that the crystallinity of the sample
(28532923 cm1) lose their intensity due, mainly, to the degrada- increases with the incorporation of F on the structure. However,
tion of the organic matrix as well as the removal of moisture pres- the crystallinity is affected by the preferential orientation of the
ent in the raw material. For bone char N-BC obtained using crystallites masking the amorphous phase. In fact, uorapatite
nitrogen (Fig. 4b) and the samples obtained using CO2 (Fig. 4cg), has a lower solubility and higher thermal stability compared to
there is a band at 3420 cm1 corresponding to a high energy hydroxyapatite, and produces highly crystalline samples [10]. A
elongation peak coming from OH groups on the hydroxyapatite pure hydroxyapatite has a stoichiometric formula in the form of
structure [35]. Note that the absorption band of the OH group Ca10(PO4)6(OH)2. In the uoride removal experiments, some F ions
loses intensity with an increase in the carbonization temperature, replace part of the OH ions and forms Ca10(PO4)6Fx(OH)2x, that is,
and a drastic change takes place above 700 C. This effect may be the so-called uoridated hydroxyapatite. In the ideal case, uorine
due to dehydroxylation of hydroxyapatite and its structural change replaces all OH group and thus forms uorapatite Ca10(PO4)6F2
[22,36]. It is important to highlight that this change in the surface [43,44]. In a hydroxyapatite, the F/Ca ratio is equal to 0, since there
chemistry of bone chars above 700 C is associated with the is no uorine. In a uoridated hydroxyapatite, the F/Ca ratio varies
decrease in the uoride adsorption capacity. FTIR signal at from 0 to <0.2. In a pure uorapatite, the F/Ca ratio is 0.2, which is
1620 cm1 is assigned to the vibration of quinone C@O, while the the stoichiometric limit. In our case, the ratio F/Ca = 0.1082 indicat-
band around 1450 cm1 is assigned to the stretching vibration of ing that the crystalline structure formed is the uoridated hydroxy-
the carboxyl group CO [37,38]. Moreover, the characteristic bands apatite. It is important to note that the F/Ca ratio of the C-BC7-F
of the phosphate group around 600 and 1040 cm1 are observed. bone char was calculated from the atomic percentage that was
These bands are assigned to PO symmetric stretching mode obtained through X-ray photoelectron spectroscopy analysis
(570600 cm1) and as a major peak is the PO asymmetric (XPS), (see Table 3).
C.K. Rojas-Mayorga et al. / Microporous and Mesoporous Materials 209 (2015) 3844 43

Table 3
XPS surface elemental analysis.

Sample at.% wt.%


name (b)
P Ca O F P Ca O F
C-BC7 16.03 23.66 60.31 0.00 20.60 39.36 40.04 0.00
C-BC7-Fa 15.65 23.75 58.01 2.59 20.08 39.44 38.45 2.04
C-BC19 17.55 24.94 57.51 0.00 22.07 40.58 37.35 0.00
a

Transmittance (%)
Bone char C-BC7 after of uoride removal experiments.

(a)

* Fluoridated hydroxyapatite
* + Hydroxyapatite

*
(b)
*
+ *
Intensity (a.u.)

* * + *** ++
** * + ** ** + * * * ++ 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 uoride removal.

(a)
+ + + + ++
+
+
+ + +
CaF2
++ + ++
684.67 eV

20 30 40 50 60 70 80
2 ()
Intensity (a.u.)

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

3.6. Functional groups after uoride adsorption studies

Additionally, Fig. 6 shows the FT-IR spectra of samples C-BC7


C-BC7-F
and C-BC7-F. In these spectra we can identify the change in
the absorption band corresponding to OH groups (3000
3600 cm1). The band intensity decreases after uoride adsorption C-BC7
experiments, and this effect can be occurring due to the ionic
exchange between F and OH [44].
690 689 688 687 686 685 684 683 682 681 680
Binding energy (eV)

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

Finally, F1s XPS spectra of the bone char samples before and
after adsorption C-BC7 and C-BC7-F, respectively, are reported in 4. Conclusions
Fig. 7. In the bone char C-BC7 before uoride adsorption experi-
ments the F1s signal does not exist; however, after uoride adsorp- It is possible to obtain bone char with outstanding uoride
tion experiments, a new and unique peak is found. This energy adsorption properties using CO2 as synthesis atmosphere. Carbon-
band corresponds to F1s with a binding energy at 684.67 eV. ization temperature is a critical operating parameter for the syn-
According to Hwang et al. [45], this F1s peak corresponds to the thesis of bone char via CO2 for uoride removal from water.
uoride bonded to calcium (CaF2), which is formed between the Specically, this temperature has a major effect on the uoride
calcium of hydroxyapatite and the uoride present in solution. adsorption properties of bone char due to the change in crystallin-
Therefore, with this study we conrm that surface reactions exist ity of the sample and to the dehydroxylation process of the
between uoride ions and calcium from hydroxyapatite [4548]. hydroxyapatite contained in this adsorbent. Also, the surface area
Also, this analysis is in agreement with XRD and FT-IR results. 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 specic surface area and OH functional [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.
groups of bone char improve the uoride adsorption capacity.
Fuentes-Rubio, Ind. Eng. Chem. Res. 46 (2007) 92059212.
The best uoride uptake of 5.92 mg/g can be obtained if a carbon- [16] P. Dasgupta, A. Singh, S. Adak, K.M. Purohit, Int. Symp. Res. Stud. Mater. Sci.
ization temperature of 700 C is used for bone char synthesis via Eng. (2004) 16.
CO2. The uoride removal performance of the bone char is better [17] I. Abe, S. Iwasaki, T. Tokimoto, N. Kawasaki, T. Nakamura, S. Tanada, J. Colloid
Interface Sci. 275 (2004) 3539.
than those reported for several commercial bone chars up to 31%. [18] N. Kawasaki, F. Ogata, H. Tominaga, I. Yamaguchi, J. Oleo Sci. 58 (2009) 529
The bone char C-BC7 obtained via CO2 showed a competitive 535.
performance in the uoride removal from water (i.e., 5.92 mg/g) [19] J.C. Moreno-Pirajn, R. Gmez-Cruz, V.S. Garca-Cuello, L. Giraldo, J. Anal. Appl.
Pyrol. 89 (2010) 122128.
compared with the optimum bone char obtained via pyrolysis [20] C.K. Rojas-Mayorga, A. Bonilla-Petriciolet, I.A. Aguayo-Villarreal, V. Hernndez-
N-BC (i.e., 7.32 mg/g). FT-IR spectroscopy and XPS measurements Montoya, M.R. Moreno-Virgen, R. Tovar-Gmez, M.A. Montes-Morn, J. Anal.
proved that uoride removal by this adsorbent occurred via the Appl. Pyrol. 104 (2013) 1018.
[21] Standard Methods for Examination of Water and Wastewater, 20th edition,
ionic exchange of surface hydroxyl groups with uoride. F1s XPS American Public Health Association, 1998.
spectra conrmed this theory and the reaction between uorine [22] C.Y. Ooi, M. Hamdi, S. Ramesh, Ceram. Int. 33 (2007) 11711177.
and calcium, while the DRX study indicates the uoridated [23] S.J. Gregg, K.S.W. Sing, Adsorption Surface Area and Porosity, Ed. Academic
Press, second ed., London, 1982.
hydroxyapatite formation. [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.
Acknowledgements [26] R. Tovar-Gmez, M.R. Moreno-Virgen, J.A. Dena-Aguilar, V. Hernndez-
Montoya, A. Bonilla-Petriciolet, M.A. Montes-Morn, Chem. Eng. J. 228
Authors acknowledge the nancial support provided by (2013) 10981109.
[27] L.R. Brunson, D.A. Sabatini, Environ. Eng. Sci. 26 (2009) 17771784.
CONACYT, DGEST, Instituto Tecnolgico de Aguascalientes and [28] F.C.M. Driessens, R.M.H. Verbeeck, Biominerals, CRC Press, Florida, 2000.
Universidad de Alicante. [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,
References 2008.
[31] A.B. Martnez-Valencia, H.E. Esparza-Ponce, G. Carbajal-De la Torre, J. Ortiz-
[1] L.R. Brunson, D.A. Sabatini, Sci. Total Environ. 488 (2014) 580587. Landeros, Soc. Mex. Ciencia Tecnol. Supercies Mater. 21 (2008) 1821.
[2] N.I. Chubar, V.F. Samanidou, V.S. Kouts, G.G. Gallios, V.A. Kanibolotsky, V.V. [32] L.G. Sequeda, J.M. Daz, S.G. Gutirrez, S.G. Perdomo, O.L. Gmez, Rev. Colomb.
Strelko, Zhuravlev, J. Colloid. Interface Sci. 291 (2005) 6774. Ciencia Qumica Farm. 41 (2012) 5066.
[3] M. Islam, R.K. Patel, J. Hazard. Mater. 143 (2007) 303310. [33] M. Markovic, B. Fowler, M. Tung, J. Res. Natl. Inst. Stan. 109 (2004) 553568.
[4] S.P. Kamble, S. Jagtap, N.K. Labhsetwar, D. Thakare, S. Godfrey, S. Devotta, S.S. [34] P.E. Wang, T.K. Chaki, J. Mater. Sci. - Mater. Med. 4 (1993) 150158.
Rayalu, Chem. Eng. J. 129 (2007) 173180. [35] F. Miyaji, Y. Kono, Y. Suyama, Mater. Res. Bull. 40 (2005) 209220.
[5] B. Nagappa, G.T. Chandrappa, Microporous Mesoporous Mater. 106 (2007) [36] M. Younesi, S. Javadpour, M.E. Bahrololoom, J. Mater. Eng. Perform. 20 (2011)
212218. 14841490.
[6] W. Nigussie, F. Zewge, B.S. Chandravanshi, J. Hazard. Mater. 147 (2007) 954 [37] A.A.M. Daifullah, B.S. Girgis, H.M.H. Gad, Mater. Lett. 57 (2013) 17231731.
963. [38] C.J. Durn-Valle, M. Gmez-Corzo, J. Pastor-Villegas, V. Gmez-Serrano, J. Anal.
[7] S.V. Ramanaiah, S.V. Mohan, P.N. Sarma, Ecol. Eng. 31 (2007) 4756. Appl. Pyrol. 73 (2005) 5967.
[8] S. Venkata Mohan, S.V. Ramanaiah, B. Rajkumar, P.N. Sarma, Bioresource. [39] C. Akemi, A.M. De Guzzi, Rev. Bras. Med. 17 (2001) 123130.
Technol. 98 (2007) 10061011. [40] S. Lurtwitayapont, T. Srisatit, Environ. Asia 3 (2010) 3238.
[9] A. Eskandarpour, M.S. Onyango, A. Ochieng, S. Asai, J. Hazard. Mater. 152 [41] J.H.G. Rocha, A.F. Lemos, S. Kannan, S. Agathopoulos, J.M.F. Ferreira, J. Mater.
(2008) 571579. Chem. 15 (2005) 50075011.
[10] N. Hamdi, E. Srasra, Desalination 206 (2007) 238244. [42] H.K. Varma, S. Babu, Ceram. Int. 31 (2005) 109114.
[11] S. Meenakshi, C.S. Sundaram, R. Sukumar, J. Hazard. Mater. 153 (2008) 164 [43] Y. Chen, L. Chai, Y. Shu, J. Hazard. Mater. 160 (2008) 168172.
172. [44] S.V. Dorozhkin, Mater. Sci. 2 (2009) 399498.
[12] J.A.P. Pea, D.B. Varela, H.J. Severiano, F.G.R. Rincn, G.B. Rodrguez, SAGARPA [45] S.M. Hwang, A. Izumi, K. Tsutsui, S. Furukawa, Appl. Surf. Sci. (1994) 523527.
5 (2013). [46] P. Xiangliang, J. Wang, D. Zhang, Desalination 249 (2009) 609614.
[13] J.A. Wilson, I.D. Pulford, S. Thomas, Environ. Geochem. Health 25 (2003) 51 [47] C. Kui, Z. Sam, W. Wenjian, Surf. Coat. Technol. 198 (2005) 237240.
56. [48] D. Xiaomin, M. Dinesh, U. Charles, J. Pittman, Y. Shuo, Chem. Eng. J. 198199
[14] K.K.H. Choy, G. McKay, Chemosphere 60 (2005) 1141150. (2012) 236245.

Вам также может понравиться