Sie sind auf Seite 1von 41

Environmental Technology

ISSN: 0959-3330 (Print) 1479-487X (Online) Journal homepage: http://www.tandfonline.com/loi/tent20

Iron-oxide Nanoparticles by Green Synthesis


Method Using Moringa oleifera Leaf Extract for
Fluoride Removal

Carole Silveira, Quelen Letícia Shimabuku, Marcela Fernandes Silva &


Rosângela Bergamasco

To cite this article: Carole Silveira, Quelen Letícia Shimabuku, Marcela Fernandes Silva
& Rosângela Bergamasco (2017): Iron-oxide Nanoparticles by Green Synthesis Method
Using Moringa oleifera Leaf Extract for Fluoride Removal, Environmental Technology, DOI:
10.1080/09593330.2017.1369582

To link to this article: http://dx.doi.org/10.1080/09593330.2017.1369582

Accepted author version posted online: 21


Aug 2017.

Submit your article to this journal

View related articles

View Crossmark data

Full Terms & Conditions of access and use can be found at


http://www.tandfonline.com/action/journalInformation?journalCode=tent20

Download by: [Australian Catholic University] Date: 21 August 2017, At: 02:06
Publisher: Taylor & Francis & Informa UK Limited, trading as Taylor & Francis Group

Journal: Environmental Technology

DOI: 10.1080/09593330.2017.1369582

Iron-oxide Nanoparticles by Green Synthesis Method Using Moringa oleifera Leaf

Extract for Fluoride Removal

Carole Silveira1, Quelen Letícia Shimabuku2, Marcela Fernandes Silva3, Rosângela

Bergamasco4
Downloaded by [Australian Catholic University] at 02:06 21 August 2017

1
Department of Chemical Engineering, State University of Maringá, Avenida Colombo,

5790. Bloco D90, 87020–900, Maringá, Paraná, Brazil. e-mail: carole_silveira@hotmail.com

2
Department of Chemical Engineering, University of West Parana, street of Faculdade,

645, 85903–000, Toledo, Paraná, Brazil. e-mail: le.shimabuku@gmail.com

3
Department of Chemical Engineering, State University of Maringá, Avenida Colombo,

5790. Bloco D90, 87020–900, Maringá, Paraná, Brazil. e-mail: celafs@gmail.com

4
Department of Chemical Engineering, State University of Maringá, Avenida Colombo,

5790. Bloco D90, 87020–900, Maringá, Paraná, Brazil. e-mail: rbergamasco@uem.br

Abstract

The aim of this work was to synthesize iron-oxide nanoparticles (NPsFeO) via a non-polluting method

(green synthesis), using Moringa oleífera leaf extract, and evaluate its fluoride ion adsorption

potential, comparing its efficiency with a commercially available adsorbent (activated carbon of bone

[BGAC]). The adsorbent materials were characterized using X-ray diffraction, transmission, and

scanning electronic microscopy, X-ray dispersive energy spectrometry, and N2 adsorption/desorption.

The results showed that the fluoride ion adsorption is favorable in neutral pH values, with maximum

adsorption at pH 7 for NPsFeO, and pH 5 for the BGAC. Adsorption kinetics tests showed that the
equilibrium was reached in 40 min for the NPsFeO, and 90 min for BGAC, with adsorption potential

of 1.40 mg g-1 and 1.20 mg g-1, respectively. The model that best described the kinetic data was

pseudo-first-order for NPsFeO and pseudo-second-order for BGAC. The Langmuir isotherm had a

better fit for both adsorbents. The thermodynamic parameters indicated spontaneous and endothermic

adsorption at 30, 40, and 50° C for BGAC, and at 30°C for NPsFeO. The regeneration process showed

that is possible to reuse NPsFeO three times in the fluoride ion adsorption process. As a result of its

adsorption capabilities and the shortest contact time to achieve equilibrium, the synthesized

nanoparticles in this work are a highly promising material for fluoride ion removal.

Keywords: Iron oxide nanoparticles; Green synthesis; Moringa oleifera; Water treatment;
Downloaded by [Australian Catholic University] at 02:06 21 August 2017

Fluoride

Funding

The authors gratefully acknowledge CAPES (Coordenação de Aperfeiçoamento de Pessoal de

Nível Superior) for its financial support.

1. Introduction

Fluoride is an element found in water, in the concentration range of 0.5–1.5 mg L-1,

and has benefits for human health. However, water consumption at concentrations above

that value for extended periods of time can cause skeletal fluorosis, dental, neurological

diseases, and affect fetal brain function, among others. [1, 2, 3]. Among the available

methods for fluoride ion removal from water, we can mention coagulation-precipitation [4],

electrochemical processes [5], Ion Exchange [6], membrane separation [7] and adsorption [8].

The adsorption process stands out for its simple application and provides satisfactory results

for the removal of fluoride ions [7].


Several adsorbents have already been evaluated for water-fluoride removal, among

them: activated alumina, biosorbents, resins, calcium-based adsorbents, and activated

charcoal. [9, 10, 11]. Nanotechnology has been widely addressed in relation to adsorption

processes due to its differentiated characteristics, such as the large surface area available

compared to the small volume of the particles. Nanoparticles of iron oxides can be easily

synthesized, and are used in the preparation of different adsorbents for the adsorption of

cations, anions, organic components, among other substances present in contaminated water

[12, 13].
Downloaded by [Australian Catholic University] at 02:06 21 August 2017

Among the most used methods for the metal nanoparticles synthesis hydrothermal

synthesis can be cited [14]. It´s a chemical approach in which metal ions in the solution are

reduced in conditions that favor the subsequent formation of clusters or metal aggregates on

the nanometer scale [15, 16]. Plant extracts can be used in green synthesis processes, and can

act as reducing agents for the development of metallic nanoparticles [17, 18]. It is known that

the source of the plant extract can influence the nanoparticles characteristics. This occurs

because different extracts contain different concentration combinations of organic reducing

agents [19]. It is believed that the principal active agents in some of these syntheses are the

polyphenols present, for example, in tea, wine, red=grape bagasse, pomegranate leaf, and

Moringa oleífera. The green synthesis of nanoparticles provides an advance over other

methods because it is simple and inexpensive, relatively reproducible, and often results in

more stable materials [20, 21, 22].

The Moringa oleífera is a plant of the Moringaceae family, and is native to India,

Pakistan, Bangladesh, and Afghanistan. [23] used the Moringa oleífera flower extract to

synthesize hydroxyapatite nanoparticles, crystalline calcium phosphate. [24] synthesized

nanoparticles of zinc oxide using Moringa oleífera leaf extract as a reducing agent. [21]

studied the green synthesis of gold nanoparticles using Moringa oleífera petal extract as a
reducing agent. This study presents the synthesis of iron-oxide nanoparticles (NPsFeO) using

Moringa oleífera extract, and its use in fluoride ion adsorption, comparing its adsorption

efficiency with a commercial adsorbent of granulated charcoal of bone (BGAC).

2. Materials and methods

2.1. Adsorbents
Downloaded by [Australian Catholic University] at 02:06 21 August 2017

NPsFeO adsorbent was prepared from iron nitrate III (Fe(NO3)3.9H2O) synthesized

by green synthesis using Moringa oleífera leaf. The commercially available adsorbent

(BGAC), from the Bonecher company, was used for comparison.

2.1.1. Synthesis of NPsFeO using Moringa oleifera leaf extract

The green synthesis of NPsFeO has been described by Shahwan, Abu Sirriah [25] and

by Huang, Weng [26]. In this study, the Moringa oleífera leaf extract was prepared by heating

an aqueous mixture of leaves to 80° C, in a 60 g L-1 concentration, for 1 hour. Then the

extract was vacuum-filtered. Then, a solution of 0.1 M of iron nitrate III was added to the

Moringa oleífera extract, with 99.10% purity in a 1:2 (v/v) ratio under stirring. It was possible

to visualize the NPsFeO formation immediately by the color change of the mixture, from

green to black. NPsFeO separation was first made by water evaporation at room temperature,

then by drying at 50° C in an oven.

2.2. Fluorinated Solution


A fluoride standard solution of 1.000 mg L-1 was prepared by dissolving 2.21 g of NaF

(P.A.) in distilled water and, from this, dilutions were made of the initial concentration of 6

mg L-1 [27, 28].

2.3. Adsorbents characterization

The NPsFeO and BGAC were characterized by X-ray diffraction (XRD),

Transmission Electron Microscopy (TEM), Scanning Electron Microscopy (SEM),


Downloaded by [Australian Catholic University] at 02:06 21 August 2017

energy dispersive X-ray Spectrometry (EDX), and N2 adsorption/desorption by BET method.

Diffractograms of the adsorbents were obtained for the structural characterization of

the samples using a diffractometer (Bruker-AXS D8, Advance). A copper emission radiation

source (CuKα, λ = 0.154 nm), 40 kV voltage, 30 mA current, and 2θ scan ranging from 15 °

to 75 °, with a 2 ° min-1 step, were used. The diffractograms were interpreted using a JCPDS

87-1166 diffraction pattern. The NPsFeO size was estimated using the Scherrer equation

(Equation 1).

⁄ (1)

In this equation is the radiation wavelenght, 0.89 is the constant related to the

spherical shape approximation, B is the width of the peak at half height, and B is the Bragg

angle [29].

Morphological characteristics were analyzed by TEM. The adsorbent samples were

deposited in a sample port (grid) of Cu (200 mesh), and covered with a thin film of pure

carbon (CF200-Cu, EMS) in JEM-1400, with JEOL equipment 120 kV. Using SEM the
adsorbent samples were fixed on metallic disks, and then covered with a thin film of gold

using Quanta 250 - Fei, coupled with an EDX system (Oxford Instruments x-act).

Adsorption/Desorption of N2 (77 K) were estimated in a gas-sorption analyzer

(Quantachrome Autosorb 1® volumetric analyzer) to verify the textural characteristics of the

samples, such as the specific surface area (SBET), average pore diameter (Dp), and pore

volume (VP).

2.5. Adsorption of fluoride ions in NPsFeO and BGAC


Downloaded by [Australian Catholic University] at 02:06 21 August 2017

All tests explained as follows were performed for the NPsFeO and BGAC adsorbents.

The effect of pH on the adsorption was analyzed using 0.5 g adsorbent mass and a volume of

50 mL of solution with 6 mg L-1 concentration. The pH values were varied at 3, 4, 5, 7, and

10, using HNO3 (1 mol/L) and/or NaOH (0.1 mol/L). During the batch adsorption process, the

solutions with the adsorbents were placed in sealed containers in a Dubnoff bath (304-TPA),

at 130 rpm agitation for 6 hours, then they were vacuum-filtered and read in a UV-Vis

spectrophotometer using the SPADNS method—4500–F- D.—Standard Methods [30].

Adsorption kinetics tests were conducted using 0.5 g adsorbent mass and 50 mL of

solution, with a fluoride concentration of 6 mg L-1, pH 5.0 for BGAC, and pH 7.0 for

NPsFeO. Samples were placed in a bath, with agitation of 130 rpm and 25° C, and drawn at

different time intervals, filtered, and the final concentration of fluoride ions determined.

The adsorption isotherm tests were performed at pH 5.0 for BGAC, and pH 7.0 for

NPsFeO, with fluoride concentration of 6 mg L-1, and the adsorbent mass was varied between

0.05 and 0.5 g. The solutions with the adsorbents were stirred at 130 rpm in the bath until

equilibrium time was reached, and the final concentration of fluoride ions was determined.
2.6. Regenaration of NPsFeO

The regeneration of NPsFeO was investigated in order to evaluate the possibility of

reuse of the adsorbent in the fluoride ion adsorption process. The methodology used was

described by [31]. After impurity removal the adsorbent was tested in three successive reuses.

NPsFeO was cleaned by washing with a 20% ethanol solution. The mixture was stirred for 10

min at room temperature. After washing, the supernatant was removed by magnetic

separation, and the wash was repeated three times.


Downloaded by [Australian Catholic University] at 02:06 21 August 2017

3. Results and discussions

The results are presented that were obtained for the characterizations of the

adsorbent materials NPsFeO and BGAC, and the fluoride ions adsorption results in these

adsorbents.

3.1. Adsorbents Characterization

Diffractograms obtained by XRD of NPsFeO and BGAC are shown in Figure 1. The

diffractograms were used to the structural/compositional characteristic determinations of the

samples. For the sample BGAC (Figure 1a) shows that the main phosphorus phase is the

hydroxyapatite (bone charcoal consists of 76% CaHAP (Ca10(PO4)6(OH)2) [32]. The NPsFeO

sample (Figure 1b), despite the high noise level, caused by scattering due to the use of a

copper source, it is possible to observe characteristic peaks at 30.3º, 33º, 35.6º, 43.4º , 49.5°,

54.5°, 57.5°, and 62.8°, for diffraction planes of iron oxide in the hematite (α-Fe2O3)
structural form, according to the diffraction pattern JCPDS 87-1166 [33]. The NPsFeO

average diameter was estimated at 7nm by the Scherrer equation (Equation 1).

Figure 1 near here

The morphologies of the adsorbents were analyzed by means of SEM, the

magnitude being 20000x in relation to the real size of NPsFeO and BGAC. The micrographs

and EDX spectra are presented in Figure 2.


Downloaded by [Australian Catholic University] at 02:06 21 August 2017

Figure 2 near here

As can be seen in Figure 2a, the NPsFeO synthesized using Moringa oleífera leaf

extracts presented spherical agglomerated. Some particle agglomerates can also be observed

due to magnetic properties [34]. Similar results were obtained by other authors [35]. The

BGAC presents a porous and irregular structure (Figure 2c), which is a typical characteristic

of activated materials [36]. The NPsFeO EDX pattern (Figure 2.b) detected the predominant

presence of Fe and O, and a small percentage of K and Ca. The presence of oxygen indicates

the formation of Fe oxide nanoparticles, as previously demonstrated by XRD. The presence of

K and Ca originated from the compounds present in the Moringa oleífera leaf [37]. The EDX

spectrum of BGAC (Figure 2d) showed high values of calcium and phosphorus due to the

main phase being hydroxyapatite [32].

The TEM micrographs of NPsFeO and BGAC are shown in Figure 3.

Figure 3 near here


Micrographs of NPsFeO (Figure 3. a) show that the particles have an average

diameter of less than 100nm. As can be seen, the obtained particles have nanometric

structures, in which small particles associate themselves, forming darker coloration

agglomerates, which are characteristic of metallic particles [38, 39]. These images confirmed

that the green synthesis using Moringa oleífera leaf extract is efficient in obtaining iron

compound nanoparticles. In the activated charcoal sample (Figure 3. b) only grayish

formations of hydroxyapatite crystals are noticed [40]

The textural characterization parameters were estimated by the N2


Downloaded by [Australian Catholic University] at 02:06 21 August 2017

adsorption/desorption method (77 K), and showed that the specific surface areas (SBET) were

133.9 m2 g-1 and 99.79 m2 g-1, with a pore volume of 0.3283 and 0.0870 cm3 g-1 for the BGAC

adsorbent and the NPsFeO, respectively. The mean pore diameters for the BGAC were 9.80

nm and 4.14 nm for the NPsFeO, which characterizes the predominance of mesopores.

Figure 4 shows the adsorption/desorption isotherms of N2 corresponding to adsorbent

materials. According to the IUPAC classification, the observed isotherms of type III and V for

BGAC (Figure 4a) are related to very weak interactions in systems containing macro and

mesopores, whereas those of type IV for NPsFeO (Figure 3.b), are related to the presence of

hysteresis and are characteristic of mesopores, which confirms the results previously found

[41, 42].

Figure 4 near here

3.2. Adsorption tests

The pH of the solution is one of the main factors that can influence the adsorption

capacity by affecting the adsorbents’ surface charge [43]. To determine the most favorable
chemical condition for the fluoride ion adsorption on the activated carbon and NPsFeO,

adsorption studies were performed at different pH values (3, 5, 7, and 10). In Figure 5 it is

possible to observe that the adsorption was influenced by the pH of the environment. So, for

all subsequent adsorption tests, pH 7 was adopted for NPsFeO, and pH 5 for BGAC, with

maximum adsorption capacities of 1.40 and 1.47 mg g-1, respectively. Other authors have

also used this pH range for fluoride ion removal [44, 45, 46]. According to studies reported in

the literature [47, 48], in an alkaline environment the hydroxyl groups will compete with

fluorine ions at the active site of the adsorbent, reducing the adsorption potential to basic pH.
Downloaded by [Australian Catholic University] at 02:06 21 August 2017

Figure 5 near here

Defluoridation kinetic tests were performed to determine the adsorbents’ equilibrium

point. Figure 4 shows the fluoride ion kinetic adsorption behavior in NPsFeO and BGAC,

using an aqueous solution of 6 mg L-1 fluoride.

Figure 6 near here

The adsorption process was fast at the beginning of the process, due to a large number

of active sites available for adsorption. Over time the adsorbate removal decreased, reaching

equilibrium point at 40 min when NPsFeO was used, and 180 min for the BGAC, with an

adsorption capacity of 1.40 mg g-1 for NPsFeO and 1.20 mg g-1 for the BGAC. The adsorption

capacity of both adsorbents was similar, but the adsorbent synthesized in this work was

equilibrated in a significantly shorter time than BGAC.

To analyze the adsorption process mechanism, the data obtained in the kinetic tests

were adjusted for the following nonlinear models: pseudo-first-order—this model assumes
that the adsorbate removal rate of in relation to time is directly proportional to the difference

in saturation concentration, and the number of active sites present in the adsorbent [49, 50].

Pseudo-second order it is assumed that the adsorption capacity of the adsorbent depends on

the number of active sites occupied in the same [51], represented by Equations 2 and 3, using

OringinPro 9 64-bit software.

(2)
Downloaded by [Australian Catholic University] at 02:06 21 August 2017

⁄ (3)

where and are pseudo-first-order and pseudo-second-order rate constants,

respectively, is the amount of ions adsorbed at time t, and is the adsorption capacity at

equilibrium. The parameters of the abovementioned models and the ratio coefficients for

adsorption of fluoride ions are shown in Table 1.

Table 1 near here

The results suggest that the fluoride adsorption processes were well represented by

both the pseudo-first-order model and the pseudo-second-order model, with r2 values varying

from 0.996–0.999. However, for the process using BGAC, the experimental qeq value and the

derivative of the pseudo-second order model are different, according to García-Sánchez,

Solache-Ríos [52]. This means that fluoride adsorption is a first-order reaction.

The adsorption isotherm is important for estimating the affinity between adsorbent and

adsorbate, and the maximum adsorption capacity. [36, 53]. The fluoride ion adsorption
experimental data using NPsFeO and BGAC were adjusted for the Langmuir, Freundlic,

Temkin, Toth, Sips, and Redlich-Paterson models (Table 2), in order to better understand the

adsorption behavior. The experimental data with model adjustments are shown in Figure 7

and the adjusted parameter values for each isotherm, as well as the correlation coefficient r2,

are shown in Table 3.

Table 2 near here


Downloaded by [Australian Catholic University] at 02:06 21 August 2017

Figure 7 near here

Table 3 near here

The results indicate that the experimental data of adsorption isotherms for BGAC and

NPsFeO have been described for all models (Langmuir, Freundlic, Temkin Toth, Sips and

Redlich-Paterson), since the correlation coefficients vary from 0.9482 to 0.9868 (Table 3).

However, the Langmuir isotherm defines a better fit for both BGAC and NPsFeO, with a

correlation coefficient of 0.9824 and 0.9868, and maximum adsorption capacity for the lowest

adsorbent mass used of 4.13 and 4.44 mg g-1, respectively, which implies the adsorption of

fluoride ions is homogeneous and monolayer, and that adsorption only occurs in a fixed

number of equivalent and well-defined sites, accommodating only one adsorbed atom.

The parameter is an important datum provided by the Langmuir isotherm, which

relates the Langmuir affinity constant (b), through Equation 4, and provides the solute affinity

degree by the sites of the adsorbent material.

⁄ (4)
where b is the Langmuir isothermal constant, and C0 is the concentration of fluoride

ions (mg g-1). For both adsorbents studied the values of RL were calculated, and the values are

shown in Figure 8. The equilibrium parameters presented values greater than zero and less

than one, indicating that the fluoride ion adsorption process is favorable, since for RL> 1, the

process is unfavorable, for = 1 it is linear, for 0 < <1 it is favorable, and for =0 it is

irreversible [54].
Downloaded by [Australian Catholic University] at 02:06 21 August 2017

Figure 8 near here

Analyzing the adsorption capacity of the nanoparticles synthesized in this work

(NPsFeO), and comparing with other adsorbents already used for the removal of fluoride ions

(Table 4), it can be observed that the NPsFeO presented a higher pollutant adsorption capacity

in some cases, proving to be a good alternative for the removal of fluoride ions.

Table 4 near here

3.3. Thermodynamic Study in Adsorption Process

The main purpose of the thermodynamic study is to determine whether the process is

spontaneous, endothermic, or exothermic. To calculate the free energy of Gibbs for the

fluoride ion adsorption process at temperatures of 30, 40, and 50° C, the enthalpy (ΔH°) and

entropy (ΔS°) parameters in Equations 5 and 7 were used [55].

(5)
(6)

⁄ ⁄ ⁄ (7)

in which ΔG° is the Gibbs free energy variation, ΔH° is the enthalpy variation, ΔS° is

the entropy variation, R is the ideal gas constant, T is the absolute temperature in kelvin, and

K is the constant of balance. The thermodynamic adsorption parameters are shown in Table 5.
Downloaded by [Australian Catholic University] at 02:06 21 August 2017

Table 5 near here

The ΔG° indicates the degree of spontaneity of the adsorptive process.

The negative values found for the fluoride ions adsorption using BGAC confirm the

spontaneity of the process, and the thermodynamic viability for all the analyzed temperatures

[56]. For the NPsFeO it can be observed that the process is only spontaneous at the lowest

temperature. The endothermic nature of the process can be verified by the positive values of

ΔH° for the two adsorbents. However, the positive value found for ΔS° in the BGAC process

suggests an increase in the degree of randomness at the solid-liquid interface, in the

adsorption of fluoride ions. For the NPsFeO, the negative ΔS° implies the fact that the

fluoride ion molecules decrease their randomness, increasing the interaction of the adsorbate

with the adsorbent.

3.4. NPsFeO regeneration


Based on the results of the adsorption tests (better adsorption capacity and lower

equilibration time), the regeneration was only done for NPsFeO as the adsorbent of interest in

the present study.

After using NPsFeO in the fluoride ion adsorption process, these were washed with

20% ethanol solution to determine the reuse potential of the adsorbent. Adsorption tests show

that it is possible to reuse the NPsFeO three times (Figure 9), compared to the control sample

(qeq of fluoride ion removal before washing). A reduction of 80.94% to 80.44% can be

observed in the adsorption potential after the first wash, and 75% in the following washes.
Downloaded by [Australian Catholic University] at 02:06 21 August 2017

According toOkoli, Boutonnet [31] the decrease in adsorption efficiency may have occurred

due to the incomplete removal of the fluoride ion from the surface of NPsFeO. The major

advantage of regenerating NPsFeO is the reduction in the adsorption process cost.

Figure 9 near here

4. Conclusion

In the present study, iron-oxide nanoparticles were successfully synthesized using

Moringa oleífera leaf extract by green synthesis. Its efficiency was tested in the adsorption

process of fluoride ions in a batch system, comparing it with a bone activated charcoal

commercial adsorbent

The characterization techniques showed that the adsorbent materials are porous with a

predominance of mesopores, and with an irregular surface for the BGAC and spherical forms

for NPsFeO. In addition, the EDX and XRD spectra confirmed the formation of iron-oxide

nanoparticles.
In the evaluation of batch process adsorbents, NPsFeO obtained a higher efficiency,

with equilibrium time reached in 40 min, and an adsorption capacity of 1.40 mg g-1 at pH 7,

compared to BGAC that reached equilibrium time in 180 min, and whose maximum

adsorption capacity was 1.20 mg g-1 at pH 5. The adsorption experimental data of both

adsorbents showed a good fit for the Langmuir isotherm, with correlation coefficients equal to

0.9868 and 0.9824 for NPsFeO and BGAC, respectively. The parameter indicated that the

fluoride ion adsorption process is favorable for the two adsorbents studied

The regeneration study of NPsFeO shows that even though a reduction of the
Downloaded by [Australian Catholic University] at 02:06 21 August 2017

adsorption capacity occurs after the first washing, the reuse of the same in the process is

possible, which can lead to a cost reduction.

Based on the results obtained in this work, it can be concluded that the green synthesis

of iron-oxide nanoparticles was successful and could be a favorable alternative for the

environment. Besides its easy application, there is no need to use chemicals, high

temperatures, and pressure. In addition, it presented an excellent adsorption capacity for

fluoride ions, compared to the commercial adsorbent.

Acknowledgements

The authors are grateful to COMCAP/UEM (Complexo de Centraisde Apoio à

Pesquisa) for DRX, TEM and SEM analyses.

References

1. Viswanathan G, Jaswanth A, Gopalakrishnan S, et al. Determining the optimal fluoride


concentration in drinking water for fluoride endemic regions in South India. Science of the
Total Environment. 2009;407(20):5298-5307. doi:
http://dx.doi.org/10.1016/j.scitotenv.2009.06.028.
2. Ganvir V, Das K. Removal of fluoride from drinking water using aluminum hydroxide coated
rice husk ash. Journal of Hazardous Materials. 2011;185(2–3):1287-1294. doi:
http://dx.doi.org/10.1016/j.jhazmat.2010.10.044.
3. Chai L, Wang Y, Zhao N, et al. Sulfate-doped Fe3O4/Al2O3 nanoparticles as a novel adsorbent
for fluoride removal from drinking water. Water Research. 2013;47(12):4040-4049. doi:
http://dx.doi.org/10.1016/j.watres.2013.02.057.
4. Meenakshi, Maheshwari RC. Fluoride in drinking water and its removal. Journal of Hazardous
Materials. 2006;137(1):456-463. doi: http://dx.doi.org/10.1016/j.jhazmat.2006.02.024.
5. Pulkka S, Martikainen M, Bhatnagar A, et al. Electrochemical methods for the removal of
anionic contaminants from water – A review. Separation and Purification Technology.
2014;132:252-271. doi: http://dx.doi.org/10.1016/j.seppur.2014.05.021.
6. Meenakshi S, Viswanathan N. Identification of selective ion-exchange resin for fluoride
sorption. Journal of Colloid and Interface Science. 2007;308(2):438-450. doi:
http://dx.doi.org/10.1016/j.jcis.2006.12.032.
7. Mohapatra M, Anand S, Mishra BK, et al. Review of fluoride removal from drinking water.
Downloaded by [Australian Catholic University] at 02:06 21 August 2017

Journal of environmental management. 2009;91(1):67-77. doi:


http://dx.doi.org/10.1016/j.jenvman.2009.08.015.
8. Yu Y, Wang C, Guo X, et al. Modification of carbon derived from Sargassum sp. by lanthanum
for enhanced adsorption of fluoride. Journal of Colloid and Interface Science. 2015;441:113-
120. doi: http://dx.doi.org/10.1016/j.jcis.2014.10.039.
9. Ghorai S, Pant KK. Equilibrium, kinetics and breakthrough studies for adsorption of fluoride
on activated alumina. Separation and Purification Technology. 2005;42(3):265-271. doi:
http://dx.doi.org/10.1016/j.seppur.2004.09.001.
10. Leyva-Ramos R, Rivera-Utrilla J, Medellin-Castillo NA, et al. Kinetic modeling of fluoride
adsorption from aqueous solution onto bone char. Chemical Engineering Journal.
2010;158(3):458-467. doi: http://dx.doi.org/10.1016/j.cej.2010.01.019.
11. Bhatnagar A, Kumar E, Sillanpää M. Fluoride removal from water by adsorption—A review.
Chemical Engineering Journal. 2011;171(3):811-840. doi:
http://dx.doi.org/10.1016/j.cej.2011.05.028.
12. Kobayashi A, Gulshan F, Kameshima Y, et al. Preparation and simultaneous ion uptake
properties of CaO-Fe2O3-SiO2 compounds. Journal of the Ceramic Society of Japan.
2008;116(1350):187-191.
13. Gulshan F, Kameshima Y, Nakajima A, et al. Preparation of alumina-iron oxide compounds by
gel evaporation method and its simultaneous uptake properties for Ni2+, NH4+ and H2PO4-.
Journal of Hazardous Materials. 2009 Sep 30;169(1-3):697-702. doi: DOI
10.1016/j.jhazmat.2009.04.009. PubMed PMID: ISI:000268967200096; English.
14. Merkache R, Fechete I, Maamache M, et al. 3D ordered mesoporous Fe-KIT-6 catalysts for
methylcyclopentane (MCP) conversion and carbon dioxide (CO2) hydrogenation for energy
and environmental applications. Applied Catalysis A: General. 2015. doi:
http://dx.doi.org/10.1016/j.apcata.2015.03.032.
15. Khomutov GB, Gubin SP. Interfacial synthesis of noble metal nanoparticles. Materials Science
and Engineering: C. 2002;22(2):141-146. doi: http://dx.doi.org/10.1016/S0928-
4931(02)00162-5.
16. Oliveira MM, Ugarte D, Zanchet D, et al. Influence of synthetic parameters on the size,
structure, and stability of dodecanethiol-stabilized silver nanoparticles. Journal of Colloid and
Interface Science. 2005;292(2):429-435. doi: http://dx.doi.org/10.1016/j.jcis.2005.05.068.
17. Philip D. Biosynthesis of Au, Ag and Au–Ag nanoparticles using edible mushroom extract.
Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy. 2009;73(2):374-381.
doi: http://dx.doi.org/10.1016/j.saa.2009.02.037.
18. Thakkar KN, Mhatre SS, Parikh RY. Biological synthesis of metallic nanoparticles.
Nanomedicine: Nanotechnology, Biology and Medicine. 2010;6(2):257-262. doi:
http://dx.doi.org/10.1016/j.nano.2009.07.002.
19. Mukunthan K, Balaji S. Cashew apple juice (Anacardium occidentale L.) speeds up the
synthesis of silver nanoparticles. International Journal of Green Nanotechnology.
2012;4(2):71-79.
20. Abou El-Nour KMM, Eftaiha Aa, Al-Warthan A, et al. Synthesis and applications of silver
nanoparticles. Arabian Journal of Chemistry. 2010;3(3):135-140. doi:
http://dx.doi.org/10.1016/j.arabjc.2010.04.008.
21. Anand K, Gengan RM, Phulukdaree A, et al. Agroforestry waste Moringa oleifera petals
mediated green synthesis of gold nanoparticles and their anti-cancer and catalytic activity.
Journal of Industrial and Engineering Chemistry. 2015;21:1105-1111. doi:
http://dx.doi.org/10.1016/j.jiec.2014.05.021.
22. Byranvand MM, Kharat AN. One pot green synthesis of gold nanowires using pomegranate
juice. Materials Letters. 2014;134:64-66. doi:
http://dx.doi.org/10.1016/j.matlet.2014.07.046.
23. Sundrarajan M, Jegatheeswaran S, Selvam S, et al. The ionic liquid assisted green synthesis of
hydroxyapatite nanoplates by Moringa oleifera flower extract: A biomimetic approach.
Downloaded by [Australian Catholic University] at 02:06 21 August 2017

Materials & Design. 2015;88:1183-1190. doi:


http://dx.doi.org/10.1016/j.matdes.2015.09.051.
24. Elumalai K, Velmurugan S, Ravi S, et al. Green synthesis of zinc oxide nanoparticles using
Moringa oleifera leaf extract and evaluation of its antimicrobial activity. Spectrochimica Acta
Part A: Molecular and Biomolecular Spectroscopy. 2015;143:158-164. doi:
http://dx.doi.org/10.1016/j.saa.2015.02.011.
25. Shahwan T, Abu Sirriah S, Nairat M, et al. Green synthesis of iron nanoparticles and their
application as a Fenton-like catalyst for the degradation of aqueous cationic and anionic
dyes. Chemical Engineering Journal. 2011;172(1):258-266. doi:
http://dx.doi.org/10.1016/j.cej.2011.05.103.
26. Huang L, Weng X, Chen Z, et al. Green synthesis of iron nanoparticles by various tea extracts:
Comparative study of the reactivity. Spectrochimica Acta Part A: Molecular and Biomolecular
Spectroscopy. 2014;130(0):295-301. doi: http://dx.doi.org/10.1016/j.saa.2014.04.037.
27. Karkar S, Debnath S, De P, et al. Preparation, characterization and evaluation of fluoride
adsorption efficiency from water of iron-aluminum oxide-graphene oxide composite
material. Chemical Engineering Journal. 2016. doi:
http://dx.doi.org/10.1016/j.cej.2016.07.037.
28. Massoudinejad M, Ghaderpoori M, Shahsavani A, et al. Adsorption of fluoride over a metal
organic framework Uio-66 functionalized with amine groups and optimization with response
surface methodology. Journal of Molecular Liquids. 2016;221:279-286. doi:
http://dx.doi.org/10.1016/j.molliq.2016.05.087.
29. Koch C. Structural Nanocrystalline Materials: Fundamentals and Applications. Cambridge
University Press; 2007.
30. Federation WE, Association APH. Standard methods for the examination of water and
wastewater. American Public Health Association (APHA): Washington, DC, USA. 2005.
31. Okoli C, Boutonnet M, Järås S, et al. Protein-functionalized magnetic iron oxide
nanoparticles: time efficient potential-water treatment [journal article]. Journal of
Nanoparticle Research. 2012;14(10):1194. doi: 10.1007/s11051-012-1194-9.
32. Ghaneian MT, Ghanizadeh G, Alizadeh MTH, et al. Equilibrium and kinetics of phosphorous
adsorption onto bone charcoal from aqueous solution. Environmental technology.
2014;35(7):882-890.
33. Silva MF, de Oliveira LAS, Ciciliati MA, et al. Nanometric particle size and phase controlled
synthesis and characterization of γ-Fe2O3 or (α + γ)-Fe2O3 by a modified sol-gel method.
Journal of Applied Physics. 2013;114(10):104311. doi:
doi:http://dx.doi.org/10.1063/1.4821253.
34. Pattanayak M, Nayak P. Ecofriendly green synthesis of iron nanoparticles from various plants
and spices extract. International Journal of Plant, Animal and Environmental Sciences.
2013;3(1):68-78.
35. Devatha CP, Thalla AK, Katte SY. Green synthesis of iron nanoparticles using different leaf
extracts for treatment of domestic waste water. Journal of Cleaner Production.
2016;139:1425-1435. doi: http://dx.doi.org/10.1016/j.jclepro.2016.09.019.
36. Marin P, Módenes AN, Bergamasco R, et al. Synthesis, Characterization and Application of
ZrCl4-Graphene Composite Supported on Activated Carbon for Efficient Removal of Fluoride
to Obtain Drinking Water [journal article]. Water, Air, & Soil Pollution. 2016;227(12):479. doi:
10.1007/s11270-016-3188-1.
37. Gopalakrishnan L, Doriya K, Kumar DS. Moringa oleifera: A review on nutritive importance
and its medicinal application. Food Science and Human Wellness. 2016;5(2):49-56. doi:
http://dx.doi.org/10.1016/j.fshw.2016.04.001.
38. Das MR, Sarma RK, Saikia R, et al. Synthesis of silver nanoparticles in an aqueous suspension
of graphene oxide sheets and its antimicrobial activity. Colloids and Surfaces B: Biointerfaces.
Downloaded by [Australian Catholic University] at 02:06 21 August 2017

2011;83(1):16-22. doi: http://dx.doi.org/10.1016/j.colsurfb.2010.10.033.


39. Cheng K-m, Hung Y-w, Chen C-c, et al. Green synthesis of chondroitin sulfate-capped silver
nanoparticles: Characterization and surface modification. Carbohydrate Polymers.
2014;110:195-202. doi: http://dx.doi.org/10.1016/j.carbpol.2014.03.053.
40. Mateus AP, Monteiro FJ, Ferraz MP. Nanoparticles of hydroxyapatite: preparation,
characterization and cellular approach-An Overview. Mutis. 2013;3(2):43-56.
41. Nomenclature and terminology of graphite intercalation compounds (IUPAC
Recommendations 1994), 66 (1994).
42. Lucena SMP, Oliveira JCA, Gonçalves DV, et al. Second-generation kernel for characterization
of carbonaceous material by adsorption. Carbon. 2017 2017/08/01/;119:378-385. doi:
http://dx.doi.org/10.1016/j.carbon.2017.04.061.
43. Qin Q, Wang Q, Fu D, et al. An efficient approach for Pb(II) and Cd(II) removal using
manganese dioxide formed in situ. Chemical Engineering Journal. 2011;172(1):68-74. doi:
http://dx.doi.org/10.1016/j.cej.2011.05.066.
44. Poursaberi T, Hassanisadi M, Torkestani K, et al. Development of zirconium (IV)-
metalloporphyrin grafted Fe3O4 nanoparticles for efficient fluoride removal. Chemical
Engineering Journal. 2012;189–190:117-125. doi:
http://dx.doi.org/10.1016/j.cej.2012.02.039.
45. Rathore VK, Dohare DK, Mondal P. Competitive adsorption between arsenic and fluoride
from binary mixture on chemically treated laterite. Journal of Environmental Chemical
Engineering. 2016;4(2):2417-2430. doi: http://dx.doi.org/10.1016/j.jece.2016.04.017.
46. Zazouli MA, Mahvi AH, Dobaradaran S, et al. Adsorption of fluoride from aqueous solution by
modified Azolla filiculoides. Fluoride 2014; 47 (4): 349. 2014;58.
47. Mitchell-Koch JT, Pietrzak M, Malinowska E, et al. Aluminum(III) porphyrins as ionophores for
fluoride selective polymeric membrane electrodes [Article]. Electroanalysis. 2006;18(6):551-
557. doi: 10.1002/elan.200503450.
48. Malinowska E, Górski Ł, Meyerhoff ME. Zirconium(IV)-porphyrins as novel ionophores for
fluoride-selective polymeric membrane electrodes. Analytica Chimica Acta. 2002;468(1):133-
141. doi: http://dx.doi.org/10.1016/S0003-2670(02)00635-9.
49. Largergren S. Zur theorie der sogenannten adsorption geloster stoffe. Kungliga Svenska
Vetenskapsakademiens. Handlingar. 1898;24:1-39.
50. Fiorentin LD, Trigueros DEG, Módenes AN, et al. Biosorption of reactive blue 5G dye onto
drying orange bagasse in batch system: Kinetic and equilibrium modeling. Chemical
Engineering Journal. 2010 2010/09/15/;163(1):68-77. doi:
http://dx.doi.org/10.1016/j.cej.2010.07.043.
51. Ho Y, McKay G, Wase D, et al. Study of the sorption of divalent metal ions on to peat.
Adsorption Science & Technology. 2000;18(7):639-650.
52. García-Sánchez JJ, Solache-Ríos M, Martínez-Gutiérrez JM, et al. Modified natural magnetite
with Al and La ions for the adsorption of fluoride ions from aqueous solutions. Journal of
Fluorine Chemistry. 2016;186:115-124. doi:
http://dx.doi.org/10.1016/j.jfluchem.2016.05.004.
53. Weber WJ, Morris JC. Kinetics of adsorption on carbon from solution. Journal of the Sanitary
Engineering Division. 1963;89(2):31-60.
54. Foo KY, Hameed BH. Insights into the modeling of adsorption isotherm systems. Chemical
Engineering Journal. 2010;156(1):2-10. doi: http://dx.doi.org/10.1016/j.cej.2009.09.013.
55. Saucier C, Adebayo MA, Lima EC, et al. Microwave-assisted activated carbon from cocoa shell
as adsorbent for removal of sodium diclofenac and nimesulide from aqueous effluents.
Journal of Hazardous Materials. 2015;289:18-27. doi:
http://dx.doi.org/10.1016/j.jhazmat.2015.02.026.
56. Isah A U, Abdulraheem G, Bala S, et al. Kinetics, equilibrium and thermodynamics studies of
C.I. Reactive Blue 19 dye adsorption on coconut shell based activated carbon. International
Biodeterioration & Biodegradation. 2015;102:265-273. doi:
Downloaded by [Australian Catholic University] at 02:06 21 August 2017

http://dx.doi.org/10.1016/j.ibiod.2015.04.006.
Figure 1. Diffractograms obtained by XRD of adsorbents: a) BGAC; and b) NPsFeO

Figure 2 a) NPsFeO micrographs obtained by SEM; b) NPsFeO spectrum obtained by


EDX; c) BGAC micrographs obtained by SEM; d) BGAC spectrum obtained by EDX.

Figure 3. Micrographs obtained by TEM: (a) NPsFeO; (b) BGAC

Figure 4. N2 adsorption-desorption isotherm: (a) BGAC; (b) NPsFeO

Figure 5. Effect of pH on removal of fluoride ions by NPsFeO and BGAC (pH range of
Downloaded by [Australian Catholic University] at 02:06 21 August 2017

3–10; concentration of fluoride ions solution 6 mg L-1).

Figure 6. Adsorption of fluoride ions by NPsFeO and BGAC as a function of time:


experimental and simulated data by pseudo-first and pseudo-second order kinetic
models (C0 = 6 mg L−1, m = 0,5 g, Agitation = 130 rpm, T = 25 °C).

Figure 7. Batch process adsorption isotherm, experimental data fitted to Langmuir,


Freundlic, Temkin Toth, Sips and Redlich-Paterson models: a) BGAC; b) NPsFeO.

Figure 8. Effect of initial concentration of adsorbents BGAC and NPsFeO on adsorption


of fluoride ions on RL values

Figure 9. Regeneration of NPsFeO after fluoride ion adsorption test


Table 1. Kinetic parameters of fluoride ion adsorption by NPsFeO and BGAC
Kinetic model Adsorbent
NPsFeO BGAC
qexperimental (mg g-1) 1.40 1.20
Pseudo-first order model
K1 (1 min-1) 0.175 0.016
qeq (mg g-1) 1.421 1.280
r2 0.996 0.998
Pseudo-second order model
-1 -1
K2 (g min mg ) 0.016 0.011
Downloaded by [Australian Catholic University] at 02:06 21 August 2017

qeq (mg g-1) 1.452 1.538


r2 0.999 0.990
Table 2. Langmuir, Freundlich, Temkin Toth, Sips, and Redlich-Paterson isothermal
models.
Application and
Isotherm Expression
Characteristic
Homogeneous
surfaces
Langmuir ⁄ Mono Layer
Two theoretical
parameters
Downloaded by [Australian Catholic University] at 02:06 21 August 2017

Heterogeneous
surfaces
Freundlich
Multi-layer
Two empirical
parameters

Mono layer
Temkin ⁄ chemisorption
Two theoretical
parameters

Heterogeneous
surfaces
Toth [ ⁄ ⁄ ]
Multi-layer
Three empirical
parameters

Combined
Langmuir – Freundlich
Sips [ ⁄ ]
Three semi-
empirical parameters
Downloaded by [Australian Catholic University] at 02:06 21 August 2017

Paterson
Redlich-
[

]
surfaces

parameters
Heterogeneous

Three empirical
Table 3. Parameters of adsorption isotherm of fluoride ions by NPsFeO and BGAC

Langmuir Freundlich Temkin Toth Sips Redlich-

qmax=
qmax=
9.629
qmax= 6.824 KF= 2.162 B= 1.256 8.667 kRP= 4.42
bs= 0.8321
b=0.4960 nF= 0.607 kT=7.471 bT=0.5803 n= 0.749
Ks=
R2=0.9824 R2= 0.9727 R2=0.9482 nT=0.5649 R2= 0.97
0.3010
R2=0.9740
R2=0.9747
qmax=8.66 qmax=9.62
qmax=6.036 KF=2.162 B=1.256 7 9 kRP=4.42
Downloaded by [Australian Catholic University] at 02:06 21 August 2017

b=0.680 nF=0.607 kT=7.4715 bT=0.580 bs=0.832 n=0.749


2 2 2
R =0.9868 R =0.9804 R =0.9628 nT=0.565 Ks=0.301 R2=0.981
R2=0.9816 R2=0.9822
Table 4. Adsorption capacity of different adsorbents
Adsorbent qmax (mg g-1) Reference
Nanoparticle of iron
compounds modified with Moringa
4.44 In this work
oleifera leaf

Bone Charcoal 4.13 In this work


Magnetite modified with Al 0.43
[46]
Magnetite modified with La 0.46
Activated Alumina 0.76 [49]
Downloaded by [Australian Catholic University] at 02:06 21 August 2017

Granular Ceramics 0.99 [50]


Granular ceramics modified
with FeSO4 – 7H2O 0.47
[51]
Granular ceramics modified 0.26
with Fe2O3
Hydrous ferric oxide 6.71 [52]
Table 5. Thermodynamic parameters of fluoride ion adsorption using NPsFeO and
BGAC
ΔG° (kJ / mol) ΔH° (kJ / mol) ΔS° (kJ / mol K)
T (K)=303 T=313 T=323
BGAC -3.50337 -4.30964 -6.06234 35.10753 0.126938
NPsFeO -4.30345 3.92034 3.49347 123.8703 -0.39906
Downloaded by [Australian Catholic University] at 02:06 21 August 2017
Downloaded by [Australian Catholic University] at 02:06 21 August 2017
Downloaded by [Australian Catholic University] at 02:06 21 August 2017
Downloaded by [Australian Catholic University] at 02:06 21 August 2017
Downloaded by [Australian Catholic University] at 02:06 21 August 2017
Downloaded by [Australian Catholic University] at 02:06 21 August 2017
Downloaded by [Australian Catholic University] at 02:06 21 August 2017
Downloaded by [Australian Catholic University] at 02:06 21 August 2017
Downloaded by [Australian Catholic University] at 02:06 21 August 2017
Downloaded by [Australian Catholic University] at 02:06 21 August 2017
Downloaded by [Australian Catholic University] at 02:06 21 August 2017
Downloaded by [Australian Catholic University] at 02:06 21 August 2017
Downloaded by [Australian Catholic University] at 02:06 21 August 2017
Downloaded by [Australian Catholic University] at 02:06 21 August 2017

Das könnte Ihnen auch gefallen