You are on page 1of 44

Accepted Manuscript

Magnetic MnFe2O4-Graphene Hybrid Composite for Efficient Removal of


Glyphosate from Water

Natália Ueda Yamaguchi, Rosângela Bergamasco, Safia Hamoudi

PII: S1385-8947(16)30303-5
DOI: http://dx.doi.org/10.1016/j.cej.2016.03.051
Reference: CEJ 14902

To appear in: Chemical Engineering Journal

Received Date: 28 December 2015


Revised Date: 9 March 2016
Accepted Date: 10 March 2016

Please cite this article as: N.U. Yamaguchi, R. Bergamasco, S. Hamoudi, Magnetic MnFe2O4-Graphene Hybrid
Composite for Efficient Removal of Glyphosate from Water, Chemical Engineering Journal (2016), doi: http://
dx.doi.org/10.1016/j.cej.2016.03.051

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers
we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting proof before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
Magnetic MnFe2O4-Graphene Hybrid Composite for Efficient Removal of Glyphosate

from Water

Natália Ueda Yamaguchia,b; Rosângela Bergamascob and Safia Hamoudia*

aDé partement de Sols et Gé nie Agroalimentaire, Centre en Chimie Verte et Catalyse, Université

Laval, Québec (QC), Canada, G1V 0A6.

b Departamento de Engenharia Química, Universidade Estadual de Maringa, Av. Colombo,

5790, CEP 87020-900, Bloco D90, Maringá , PR, Brazil.

*
Corresponding author: Phone (418) 656 2131 ext: 8460; Fax (418) 656 3723; email: safia.hamoudi@fsaa.ulaval.ca

1
Abstract

Reduced graphene oxide decorated with MnFe2O4 microspheres composed a hybrid

composite (MnFe2O4-G) that was synthesized by a facile and green strategy involving

immobilizing the MnFe2O4 microspheres on the graphene nanosheets via a simple one-pot

solvothermal route. The composites were characterized using transmission and scanning

electron microscopy, Fourier transformed infrared spectroscopy, X-ray diffraction, X-ray

photoelectron spectroscopy and surface area measurements. It was demonstrated that the

composite of single-layer graphene oxide with manganese ferrite magnetic microspheres

exhibited adsorption properties for efficient removal of Glyphosate from contaminated water.

The as-prepared graphene manganese ferrite composite was tested for the adsorption of

glyphosate by analytical methods under diverse experimental conditions. With respect to

contact time measurements, the adsorption of glyphosate increased and reached equilibrium

within 8 h at 25°C with a maximum adsorption capacity of 39 mg g-1 at 5°C. The Freundlich

model and the pseudo-second-order kinetic model correlated satisfactorily to the

experimental data. Thermodynamic studies revealed that the adsorption of glyphosate onto

MnFe2O4 -G was spontaneous, exothermic and feasible in the range of 5–45°C. Therefore, the

MnFe2O4-G hybrids synthesized in this research may offer an attractive adsorbent candidate

for highly efficient removal of glyphosate from contaminated water for water treatment and

purification processes.

Keywords: MnFe2O4, Graphene, Magnetic nanoparticles, Hybrids, Glyphosate, Adsorption.

2
1. Introduction

Widespread use of pesticides and herbicides for agriculture production has led to

ubiquitous contamination of soil surface and water resources, resulting in serious

environmental problems [1]. Glyphosate (N-(phosphonomethyl) glycine), an

organophosphorus pesticide, is a broad-spectrum, post-emergence, non-selective, systemic

herbicide and one of the most widely applied herbicides in both agricultural and non-

agricultural areas [2]. Due to the enormous quantities used worldwide, the concerns of its

impacts on the environment and human safety are increased, since it is introduced via various

routes to the environment, during its manufacture, use and runoff after use [3]. While

glyphosate is generally believed to be an environmentally safe herbicide, recently its

toxicological effect is questioned [4]. Many studies suggest that glyphosate is responsible for

acute toxic effects, such as endocrinal effects [5], mutagenic and carcinogenic effects [6],

genotoxic effects [7] and also may cause neurologic problems [8]. Therefore, water glyphosate

pollution is a major public concern and its removal from water is a subject of public

requirement.

Within the several processes to remove such contaminants from water, adsorption is a

top remediation technology in terms of low costs adsorbents, flexibility and simplicity of

design, easy operation and high efficiency [9, 10]. Glyphosate adsorptive removal studies are

reported using several different materials, such as biochar [11], water industrial residues

[12], zeolite combined with polyaniline [13] and layered double hydroxides [1, 14].

In view of the wide variety of nanomaterials for adsorption, graphene (G), which is

emerging as a new two-dimensional one-atom thick carbon material has attracted extensive

attention from the scientific communities due its outstanding mechanical, electrical, thermal

and optical properties as well as very high specific surface area [15]. Even though graphene

and graphene oxide (GO) are good adsorbents for many pollutants, their efficient removal
3
from water after the treatment process is still a challenge [16]. To overcome this issue, an

innovative technology that has gained much attention is the use of magnetic materials for an

easy separation process. Magnetic separation can be done using magnetic fields for its phase

separation from aqueous solutions, providing an attractive and cost-effective method for

practical operation [17]. Several efforts to integrate graphene and magnetic nanoparticles

have been pursued, since the new hybrid is also likely to possess enhanced functionalities

with respect to adsorption [18].

In the current work a facile approach for preparing magnetic MnFe2O4 microspheres

grown on graphene layers using a straightforward one-pot solvothermal process was

reported. This work was originally motivated to investigate the performance of glyphosate

adsorption using MnFe2O4 -G composite magnetically separable from water. The properties of

MnFe2O4 and MnFe2O4-G hybrids were characterized and a series of experimental parameters

were systematically analyzed and optimized. Isotherm models, kinetics of adsorption and

thermodynamic properties were also studied to describe the adsorption process.

2. Experimental

2.1 Reagents and materials

Graphite flakes 325 mesh were purchased from Alfa Aesar (99.8%). Hydrochoric acid (HCl,

36% Fisher), sulfuric acid (H2SO4, 95-98% Fisher), Potassium persulfate (K2S2O8, 99-100%

Anachemia), Phosphorus pentoxide (P2O5, 99% JT Baker), Potassium Permanganate (KMnO4,

97-100% Anachemia), Hydrogen peroxide solution 30% (30% EMD), Glyphosate (N-

(Phosphonomethyl)glycine, 96% Sigma Aldrich), Ethylene glycol (BDH), Ferric chloride

hexahydrate (FeCl3·6H2O Sigma Aldrich), Manganese dichloride tetrahydrate (MnCl2·4H2O

Sigma Aldrich), Sodium acetate anhydrous (NaAc, >99% Sigma Aldrich) were used without

4
further purification. Deionized water was used in all process of aqueous solutions,

suspensions and washing.

2.2 Preparation of graphene oxide

Graphene oxide was synthesized following the modified Hummers method [19]. Graphite

preoxidation was carried out in a 250 ml round-bottom flask. Firstly, 5 g of graphite, 18 ml of

concentrated H2SO4, 2.5 g of K2S2O8 and 2.5 g of P2O5 were added to a reflux system at 80°C

under a constant magnetic stirring for 5 h. Subsequently, heating was stopped and the

mixture was diluted with 1 L of deionized water. Later, the resultant solution was then

filtered and washed to remove excess acid. The solid product was dried in an oven at 60°C

overnight resulting in graphite pre-oxidized. Further, 1 g of pre-oxidized graphite was added

to 23 ml of concentrated H2SO4 under ice bath conditions and stirred. 3 g of KMnO4 was

slowly added to the mixture. Then, temperature was raised to 35 C and the reaction was

continued for 2 h. Subsequently, 46 ml of deionized water were added carefully keeping the

temperature below 50°C. The reaction was stirred for 2 h and then 140 ml of deionized water

and 2,5 ml of 30% H2O2 were added. The mixture was kept at room temperature for 24 h and

the supernatant was carefully decanted. The settled dispersion was centrifuged and washed

with 10% HCl followed by distilled water. This was repeated several times, then the final

product was dried at 60°C overnight. Graphene oxide (GO) was prepared after ultrasonication

[20].

2.3 Preparation of MnFe2O4-G hybrid nanocomposite

Synthesis of MnFe2O4-G hybrids was based on a facile one-pot solvothermal method using

FeCl3·6H2O and MnCl2·4H2O as starting materials. In a typical synthesis, 0.1 g of GO, 1 g of

FeCl36H2O and 0.376 g of MnCl24H2O were dispersed in 30 mL of ethylene glycol with

5
ultrasonication for 3 h. Later, 3 g of NaAc were added, followed by stirring for 30 min. The

mixture was then transferred into a 40 mL Teflon-lined stainless steel autoclave and heated at

200°C for 10 h. Solid black product was obtained and washed several times by deionized

water and ethanol and dried in an oven at 60°C overnight. Bare MnFe2O4 nanoparticles were

also synthesized by a similar approach but in the absence of GO [21-23]. Also, bare reduced

graphene oxide denoted G was prepared under the same hydrothermal conditions but

without MnFe2O4 nanoparticles.

2.4 Materials Characterization

The crystalline structure of the as-synthesized nanocomposite was identified by X-ray

diffraction (XRD) data using a Rigaku monochromatic diffractometer (Model D Ultima III),

with Cu Kα radiation source (wavelength λ = 1.5406 AU ). Fourier transform infrared (FTIR)

spectra were collected between 350 and 4000 cm−1 on a Varian 1000 (Scimitar series)

spectrometer with 2 cm−1 resolution using KBr pellets at room temperature. The morphology

and elemental composition studies were examined using a Scanning Electron Microscope

JEOL 840-A with an energy dispersive X-ray spectrometry (EDX) and a Transmission Electron

Microscope JEOL, model JEM-1230. X-ray photoelectron spectroscopy (XPS) analysis were

carried out by using AXIS ULTRA from Kratos Analytical (Manchester, UK) equipped with a

double X-ray source for non-monochromatic Al-Mg (Kα) X-ray irradiation, a monochromatic

Al (Kα) X-ray irradiation source (1486.6 eV), and an electrostatic analyzer of large radius.

Casa XPS version 2.3.13 software was used for background subtraction and fitting of the

curve. The textural properties and surface analysis of the samples were investigated using

nitrogen adsorption/desorption analyses at 77 K using a volumetric adsorption analyzer

(Model Autosorb-1, Quantachrome Instruments, Boyton Beach, FL). The specific surface area

was determined from the linear part of the Brunauer-Emmett-Teller (BET) plot (P/P0 = 0.05-

6
0.30). The pore size distributions curves were calculated from the desorption branch, using

the Horvath-Kawazoe (HK) and Barrett-Joyner-Hallenda (BJH) methods for the microporous

(pore size <2.0 nm) and mesoporous (pore size >2.0 nm) materials, respectively. The total

pore volume was evaluated from the adsorbed amount at a relative pressure P/P0 = 0.990

single point. The zeta potential analyses were realized by electrophoretic laser Doppler

anemometry at different pH values for MnFe2O4-G, MnFe2O4 and GO using a Delsa NanoTM C

Beckman Coulter.

2.5 Adsorption Experiments and Glyphosate Analysis

The adsorption capability of the synthesized adsorbent material toward glyphosate was

investigated. Adsorption kinetics experiments were conducted to find out the equilibrium

time and the kinetics models of glyphosate adsorption by MnFe2O4-G. The experiments were

carried out using Erlenmeyer flasks with 80 mg of MnFe2O4-G in 80 ml of 20 ppm glyphosate

solution in a shaker at 25°C. At preset time intervals, 15 min – 24 h, the dispersion was drawn

and separated immediately using a magnet. The supernatant was filtered with a microsyringe

(0.2 μm) and the residual glyphosate concentration was measured by ion chromatography

using an ICS 2500 from Dionex, equipped with an IonPac AS18 (4 × 250 mm) column, an

electrochemical detector and an online eluent generator. The unit was operated isocratically

at 1 mL/min using a 20 mM KOH solution as a mobile phase, with a current of 50 mA.

Adsorption isotherms were obtained performing batchwise experiments in 50 ml Erlenmeyer

flasks using an orbital shaker at 150 rpm under the following conditions: temperature, 5-

45°C; adsorbent loading, 1 g/L and contaminant initial concentration, 5-80 mg/L. The stirring

was continued for 8 h before aliquots of the solutions were withdrawn with the aid of a

magnet, and the concentration of the remaining glyphosate was analyzed. All the adsorption

experiments were performed in duplicate and analyzed twice. The data processing was

performed using Chromeleon software from Dionex. The parameters of kinetic and isotherm

7
models with statistical evaluation data were defined by nonlinear regressions using the

OriginPro software (ver. 8.0, OriginLab Co., USA).

2.6 Effect of competing anions

The effect of competing anions on the adsorptive removal of glyphosate using MnFe2O4-G was

investigated for fluoride, chloride, nitrate, sulfate and phosphate anions. Competing anions

experiments were performed separately for each anion with two different concentrations of

the competing anions (0.1 and 1.0 mM). The experiments were carried out batchwise in an

orbital shaker at 150 rpm, 35°C, 20 ppm of glyphosate initial concentration and 1 g/L of

adsorbent loading. After 24 h, the dispersion was separated using a magnet and the

concentration of the residual glyphosate on the supernatant was measured by ion

chromatography.

3. Results and Discussion

3.1 Adsorbent Characterization

3.1.1 Morphological analysis

Morphological structure of bare MnFe2O4 and MnFe2O4-G hybrid materials has been verified

by SEM and TEM techniques as shown in Fig. 1A,B. The bare MnFe2O4 prepared in the absence

of GO showed microspherical particles with severe aggregation, presenting an average

particle size of 350 nm (Fig. 1A). In comparison, the microspheres for MnFe2O4-G hybrids

were fully uniformly anchored on transparent crumpled graphene sheets indicated by the

arrows in Fig. 1C. Similar structures were found in the literature [21, 24]. As seen from the

image inset (Fig. 1D), the MnFe2O4 microspheres are actually the aggregation of a great

number of smaller MnFe2O4 nanoparticles with a size around 15 nm with an estimated cluster
8
size ranging between 100 and 400 nm [22]. The graphene structure is expected to prevent the

agglomeration of the microspheres to some extent, and ensure a large specific surface area

due to the intimate interaction between bare MnFe2O4 microspheres and graphene sheets

[24]. It is noteworthy that MnFe2O4 particles were still tightly anchored on the surface of

graphene even after the preparation of the TEM sample (mechanical stirring and sonication),

suggesting a strong interaction between MnFe2O4 and graphene and enhanced mechanical

stability [25]. Furthermore, the EDX measurement confirmed the composition of the hybrids

(Fig. 1E), which demonstrated the presence of Mn, Fe, C, and O. The C peak is mainly due the

basal plane of graphene nanosheets, while O is attributed to both MnFe2O4 and the residual

oxygen- containing functional groups of graphene [26]. These results are in agreement with

those reported in the literature [21, 27, 28].

9
Fig. 1 – SEM images of (A) MnFe2O4 microspheres, (B) MnFe2O4-G composite; TEM images of

(C) MnFe2O4-G composite and (D) inset of MnFe2O4-G composite image of selected region; (E)

EDX analysis of MnFe2O4-G composite.

10
3.1.2 X-ray diffraction

The crystallinity and phase purity of the natural graphite, the as-prepared samples of GO, G,

MnFe2O4 and MnFe2O4-G were characterized using XRD analysis (Fig. 2). The sharp and

intense well defined peak at 2θ = 26.5° observed in Fig. 2A corresponds to the (002) plane of

natural graphite. After the oxidative modification of the graphite using Hummers method, the

obtained GO exhibited a specific XRD pattern (see Fig. 2B) characterized by the peak at

scattering angle 2θ = 9.4° corresponding to (001) plane of GO [27]. Using Bragg’s law λ =

2dsinθ (where λ is the wavelength of the Cu X-ray beam, d is the spacing between the

adjacent layers, θ is the diffraction angle) [30], it was possible to evaluate the spacing

between layers in GO and natural graphite samples. The obtained values for the d-spacing

were 0.34 and 0.93 nm for natural graphite and GO, respectively. This great increase in the

distance between the interlayers of graphene oxide nanosheets can be attributed to the

functionalization of graphite sheets with oxygen-containing groups during the oxidation

process [27, 31, 32].

As for the graphene synthesized via the solvothermal method, it can be seen that the XRD

diffractogram depicted in Fig. 2C exhibited a main peak of graphene around 2θ = 24.5°. The

corresponding d-spacing was evaluated to be 0.36 nm very slightly greater than the distance

between the carbon layers of natural graphite. The great decrease in the d-spacing induced

by the solvothermal treatment of graphene oxide can be attributed to the reductive removal

of the oxygen-containing functional groups entrapped between the GO sheets thus leading to

interlayer spacing close to graphene nanosheets [31].

11
Fig. 2 – XRD patterns of natural graphite (A), the as-synthesized GO (B), reduced graphene

oxide (C), MnFe2O4 (D) and MnFe2O4-G (E).

Regarding the MnFe2O4-G sample (Fig. 2E) it is possible to notice the disappearance of the

(001) diffraction peak for GO and (002) peak of G. This can be attributed to the destruction of

the regular layer stacking of GO by the crystal growth of MnFe2O4 between the interlayers

during the reduction reaction in accordance with the earlier reports [27].

Furthermore, both MnFe2O4 and MnFe2O4-G samples exhibited similar XRD patterns (Fig

2D,E). The observed sharp and intense well-defined peaks reveal the good crystallinity of the

MnFe2O4 sample (Fig 2D). The peaks at the 2θ values of 17.78°, 30.04°, 35.50°, 42.98°, 53.32°,

12
56.74°, 62.56° and 73.46° can be indexed as the (111), (220), (311), (400), (422), (511), (440)

and (533) crystal planes of MnFe2O4, respectively [24, 33].

The average crystallite size of MnFe2O4 was estimated from the Debye-Scherrer equation D =

Kλ/βcosθ (K is a constant having value 0.9 and β is the full-width at half-maximum of the

diffraction peak) and it was found to be 22.3 nm for bare MnFe2O4 and 16.7 nm for MnFe2O4

in MnFe2O4-G. The decrease in the size of the nanoparticles in the hybrid composite may be

attributed to the fact that one side of the nanoparticles growth was blocked when grown in

situ onto the surface of the graphene [16].

3.1.3 BET analysis

Textural properties of the MnFe2O4-G and MnFe2O4 materials were investigated using N2

adsorption–desorption volumetric analysis. BET surface area of MnFe2O4-G was found to be

305.7 m2/g, which value is much larger when compared to BET surface area of 65.7 m2/g

obtained for bare MnFe2O4. Both specific areas are larger than the values reported elsewhere

[16, 22, 26]. The high surface area of the hybrid provides more adsorption and reaction sites,

consequently leading to high adsorption capacity of these adsorbents [26]. According to the

IUPAC (International Union of Pure and Applied Chemistry) classification, the nitrogen

adsorption–desorption isotherm for the MnFe2O4 -G composite (Fig. 3A) shows type IV curve

and H3 hysteresis loop. This behavior shows the predominance of mesopores [34, 35] and

also indicates the non-ordered mesoporosity with interconnection of pores [26, 36]. The

Barrett–Joyner–Halenda (BJH) method was used to calculate the average pore diameter and it

was found to be centered at 3.4 nm for both samples, and the corresponding total pore

volume was 0.96 cm3 g−1 and 0.17 cm3 g−1 for MnFe2O4-G (Fig. 3A) and MnFe2O4 (Fig. 3B),

13
respectively, showing the increase of the volume of mesopores in the hybrid nanocomposite

[22].

Fig. 3 - N2 adsorption-desorption isotherm for MnFe2O4-G nanocomposite (A) and bare

MnFe2O4 (B).

3.1.4 FTIR analysis

Fig. 4 exhibits the FTIR spectra of GO, MnFe2O4 and MnFe2O4-G recorded in the range of 4000-

350cm-1. All samples FTIR spectra presented a broad band at 3400 cm-1 that corresponds to

the stretching and bending vibration of OH groups of water molecules adsorbed on the

samples. The characteristic absorption peaks of graphite oxide appeared at 1706 cm-1

showing the possible occurrence of carbonyl and carboxyl groups (C = O of stretching COOH

groups), at 1566 cm-1 which can be attributed to aromatic C=C bond, at 1194 cm-1 and at 842

cm-1 that can be related to the epoxy group (C – O – C groups vibrations) and at 1041 cm-1

14
assigned to the alkoxy group (C – OH) [37]. These evidences indicate that during the oxidation

process of graphite with KMnO4 and sulfuric acid, the graphite conjugated π-orbital structure

were destroyed and oxygen functionalities were inserted into its composition [38].

Comparing the FTIR patterns of MnFe2O4-G and GO it is possible to notice the disappearance

of most of oxygen-containing groups peaks, revealing that the material has been reduced to

graphene sheets during the solvothermal process. The MnFe2O4-G sample presents only the

peaks of the stretching vibration of the aromatic C = C at 1535 cm-1, the OH groups at 1373

cm-1 and the C – O – C groups vibrations at 1110 cm-1 besides the MnFe2O4 peak [39]. The

spectrum of MnFe2O4 shows the presence of MnFe2O4 at 567 cm-1 and 389 cm-1 that

correspond to the formation of metal-oxygen bonds at octahedral and tetrahedral sites,

respectively [40, 41]. Moreover, the peak at 1576 cm-1 can also be attributed to the

asymmetric vibration of –COO- from the NaAc coordinated with Fe3+ [42], and the peak at

1022 cm-1 that represents the bonded hydroxyl groups on the metal of the oxide surface [43].

15
Fig. 4 – FTIR spectra of GO, MnFe2O4-G and MnFe2O4.

3.1.5 XPS analysis

To further verify the chemical composition and surface characteristics of the as-synthesized

MnFe2O4-G hybrids, XPS measurements were recorded. Distinct peaks due the elements Mn,

Fe, C, and O are evident in the wide scan XPS survey of MnFe2O4-G (Fig. 5). The peaks obtained

at 285, 530, 642 and 711 eV correspond to the C 1s in sp2 carbon, O1s of oxygen, Mn 2p, and

Fe 2p species respectively, which is in conformity with previously reported results [22]. Fig. 6

shows the deconvoluted spectrum of element peaks in the composite. The C 1s deconvoluted

spectrum of MnFe2O4-G composite (Fig. 6A) presents four Gaussian peaks centered at 284.4,

286.1, 287.6 and 288.8 eV associated with the binding energy of sp2 C–C, C–O, C=O and O=C-O
16
bonds, respectively [44]. The O 1s XP spectrum shows two peaks (Fig. 6B). The peak at 530.2

eV that is characteristic of the lattice oxide oxygen of the metal oxides as Fe–O and Mn–O of

MnFe2O4, and the other peak at 531.6 eV that belongs to the surface adsorbed oxygen groups

(such as OH, water molecules or COOH) bonded with carbon atoms in the graphene sheets

and in the oxygen MnFe2O4 interfacial bonding structure due to contact with air or organic

compounds such as ethylene glycol [26, 45]. In Fig. 6D, the Fe 2p spectra shows two main

peaks at binding energies of 711.7 and 725.3 eV, which are assigned to the Fe 2p3/2, and Fe

2p1/2 of Fe3+, respectively [46]. Moreover, from the Mn2p XPS spectra (Fig. 2C), the two

peaks centered at 641.4 and 653.9 correspond to the binding energy for Mn 2p3/2 and Mn

2p1/2, indicating the oxidation state of Mn2+ in MnFe2O4 [46]. The obtained results confirm

again that both iron and manganese elements are present in MnFe2O4-G hybrids.

Fig. 5 - Wide scan XPS spectra and element contents of MnFe2O4-G composite.

17
Fig. 6 -Deconvolution XPS Spectra of MnFe2O4-G (A) C1s (B) O1s (C) Mn 2p (D) Fe 2p

3.1.6 Zeta potential analysis

It is known that measuring the zeta potential of MnFe2O4-G hybrids as a function of pH, the

acidity or basicity of surfaces and isoelectric point (IEP) can be determined. The zeta potential

measurements are shown in Fig. 7. It was observed that IEP of MnFe2O4-G was found to be

approximately at pH 4. This indicates that at pH > IEP, the MnFe2O4-G hybrids exhibits

negative surface charge and can act as anion exchanger, while at pH < IEP, the surface charge

is positive, which is beneficial for adsorbing anions by electrostatic attraction [47].

18
Fig. 7 - pH dependent zeta-potential plots of GO, MnFe2O4 microspheres, MnFe2O4-G hybrids.

Glyphosate belongs to the aminophosphonic group, its molecule is amphoteric, ranging from a

univalent positive charge due the amino group to a trivalent negative charge from the

phosphonic group and it presents a sequence of equilibrium constants of acid dissociation

shown in the Fig. 8 [1, 13]. Glyphosate has a positive charge in very acidic pH. At pH values

between 0.78 and 2.29, glyphosate presents a neutral charge and at pH values above 2.29,

glyphosate presents a global negative charge. The number of negative charges increases with

the pH. Therefore, at a pH of 4.7 reached under the range of glyphosate concentrations

investigated in the present work, the global charge of glyphosate is negative.

19
Fig. 8 – The acid-base forms of glyphosate and equilibrium constants of acid dissociation.

The zeta potential values are more positive for MnFe2O4 compared to MnFe2O4-G. This is due

the negativity of GO, as shown in Fig. 7. The positive charge of nanoparticles increases the

adsorption capacity of anions, as glyphosate at pH 4.7. The incorporation of MnFe2O4

nanoparticles on graphene sheets avoids the agglomerations of graphene sheets (as shown in

Fig. 1B, SEM images) and hence enhances the adsorption capacity of the composite [22].

At pH below IEP (pH<4) the surface of the hybrid composite is protonated and has a positive

charge because of the following reaction [48]:



 
  →   


 
  →    (1)

20
Where, Me denoted the surface of MnFe2O4.

The positively charged surface attracts the negatively charged glyphosate through Coulombic

interaction as illustrated in Fig. 9.

Fig. 9 – Proposed adsorption mechanism by MnFe2O4-G composite at pH < 4.

When pH is above IEP (pH>4), the hybrid composite surface is hydroxylated and develops a

negative charged surface of the MnFe2O4-G and can exchange glyphosate anions as shown in

eqs. 2 and 3 [49].


≡ 
 →≡    (2)

≡      →≡      (3)

Glyphosate phosphonates have a high affinity with metal oxide via free radical pathway [50],

contributing to glyphosate replacement of the hydroxide from the surface of the hydrolyzed

metal oxides. It is proposed herein, that ion exchange is involved along with electrostatic

21
repulsion at the work pH range. The adsorption of Glyphosate on MnFe2O4-G is through a

combination of an electrostatic interaction with an ion exchange as previously proposed by

Hu, Lo and Chen [48].

3.2 Adsorption Studies

To explore the adsorption behavior of the as-synthesized MnFe2O4-G, a batch technique was

carried out for adsorption experiments as explained in the previous section. The adsorption

percentage and the amount of glyphosate retained in the adsorbent phase (qe in mg/g) at

equilibrium were calculated as follows:

"# "$
 %  × 100 (4)
"#

"# "$ *
()  +
(5)

Where C0 and Ce are the initial concentration and the concentration of glyphosate at

equilibrium (mg L-1), qe is the equilibrium adsorption capacity (mg g-1), V is the volume of

glyphosate solution (L), and m is the mass of adsorbent (g).

Fig. 10 – Glyphosate removal percentage of different materials synthesized in this work.

22
The adsorption removal of different materials synthesized is shown in Fig. 10. It was observed

that graphene oxide presented very low capacity for glyphosate removal. This can be related

to graphene oxide negative charge at pH 4.7, since glyphosate also presents a negative charge

at this pH conditions due to the carboxyl and phosphonic groups as shown in Fig 8. The

glyphosate removal of 33% obtained using GO as adsorbent can be related to the interaction

between the positively charged amino groups of glyphosate and the negative charge of GO as

shown in Fig 7.

The nanoparticles alone have shown a high removal of glyphosate (92%) but the MnFe2O4-G

composite demonstrated the best performances (97 %).

The prepared MnFe2O4 microspheres are porous, and as observed from the TEM images (Fig.

1D), small sized MnFe2O4 nanoparticles aggregate as larger microspheres and exhibiting

porous nature thus contributing to the glyphosate adsorption [22]. In the case of MnFe2O4-G,

it is worth mentioning that MnFe2O4 presents smaller nanoparticles, 22.3 nm for bare

MnFe2O4 and 16.7 nm in the MnFe2O4-G composite (calculated previously by Scherrer

equation), there are better dispersed on the graphene nanosheets surface combined with the

intrinsic adsorption capacity of the graphene itself, explaining the contribution of graphene

leading to increased number of surface active sites in turn improving the adsorption effect.

2.2.1 Adsorption Kinetics

Kinetic study provides important information about the mechanism of glyphosate adsorption

onto MnFe2O4-G. Graphical representation of the adsorption of glyphosate on MnFe2O4-G

versus equilibrium time is presented in Fig. 11. The result clearly shows that the adsorption

equilibrium for glyphosate is reached after 8h of adsorption. Therefore, all the further

23
experiments were continued for 8h before aliquots of the solutions were withdrawn for

analysis.

Fig. 11 – Glyphosate adsorption kinetics by MnFe2O4-G nanocomposite.

In order to elucidate the adsorption kinetics process of glyphosate onto MnFe2O4-G hybrids,

the experimental results of Fig. 11 were adjusted using three kinetics models, the Lagergren

pseudo-first-order, pseudo-second-order and the intra-particle model. The linear form of

pseudo-first-order kinetic model is generally expressed by the following equation:

, () − (.  , () − /0 (6)

where K1 is the rate constant of adsorption (min-1), qe and qt are the amount of glyphosate

adsorbed in mg g-1 at equilibrium and at time t (min), respectively. The first-order rate

constant K1 and qe can be determined from the intercept and slope of the plot ln(qe – qt)

versus t, respectively. The linear form of pseudo-second-order model can be expressed by:
. 0 .
1
2 3  (7)
3$ $

24
where K2 (g mg-1 min-1) is the kinetic rate constant for the pseudo-second-order adsorption

process. The kinetic rate constant K2 and qe can be calculated by the linear plot of t/qt versus

t. The intraparticle diffusion model is described by the equation:

(.  /4  5.7 8 (8)

where Kp is the intraparticle diffusion constant (mg g-1 min-0.5) which can be obtained from

the plot of Qt versus t0.5.

The adsorption kinetics data fitted into the kinetics models are shown in Fig. 12 and the

obtained parameters are given in Table 1.

Table 1 – Adsorption kinetics models parameters

Experimental Pseudo-first-order Pseudo-second-order Intraparticle diffusion

qe K1 qe, cal R2 K2 qe, cal R2 Kp C R2

(mg g-1) (min-1) (mg g-1) (g mg-1 min-1) (mg g-1) (mg g-1 min-0.5)

30.00 0.006 9.126 0.988 0.003 29.586 0.999 0.535 19.52 0.961

25
Fig. 12 – Pseudo-first-order kinetics model, Pseudo-first-order kinetics model, Intra-particle

kinetics model for glyphosate adsorption by MnFe2O4-G nanocomposite.

The experimental value of qe calculated by the first-order model does not agree with

the calculated one, obtained from the linear plots (Fig. 12A), even though, the values of

correlation coefficient (R2) are relatively good for the experimental data (Table 1). This shows

that the adsorption process may not be reasonably fitted to the first-order equation.

The linear plot of t/qt versus t (Fig. 12B) for the pseudo-second-order model shows a

good agreement between experimental and calculated qe value presented in Table 1 and also

presented a higher correlation coefficient (R2 = 0.99). Thus, the above result indicated the

applicability of the pseudo-second-order kinetic model to describe the adsorption of

glyphosate onto MnFe2O4-G for the experimental data providing a better fit. Based on the

assumption of the successful fitting with the pseudo-second-order kinetic model, it can be

suggested that chemisorption was the rate-controlling step [17, 51].

26
The constant C obtained from the intra-particle model represented in Fig. 12C

indicates if the intraparticle diffusion is a controlling step or not. When C is different from

zero, the adsorption mechanism is complex, and if C = 0, adsorption kinetics is controlled by

intraparticle diffusion only. Most of studies presents a non-zero value, indicating that the

adsorption process is a complex pathway and it is not only controlled by the intraparticle

diffusion [52].

3.2.2 Adsorption Isotherms Modeling

The adsorption isotherm model is fundamental in describing the interactive behavior

between adsorbate and adsorbent. The analysis of adsorption isotherm data was conducted to

investigate the effect of temperature on the adsorption of glyphosate on MnFe2O4-G and to

predict the maximum adsorption capacities of the adsorbent, which is one of the main

parameters required for designing an adsorption system [53]. The adsorption isotherms of

glyphosate on MnFe2O4-G for five different temperatures are depicted in Fig. 13. As shown, by

increasing the temperature, the equilibrium uptake decreased, indicating the exothermic

character of the adsorption process. Maximum adsorption capacity of 39 mg/g was obtained

at the lowest temperature investigated (5°C).

27
Fig. 13 – Glyphosate adsorption isotherms on MnFe2O4-G at different temperatures.

The adsorption equilibrium isotherms data studies were restricted to the two commonly used

adsorption isotherms models: Langmuir and Freundlich isotherm models which are

represented by the following equations:

2 9: ;9< "$
Langmuir equation: ()  029: "$
(9)

0/?
Freundlich equation: ()  /= 8) (10)

Where KL (mg g-1) is the Langmuir adsorption equilibrium constant; qm is the maximum

adsorption capacity at monolayer coverage (mg g-1), and qe is the amount of glyphosate

adsorbed at the equilibrium (mg g-1), KF ((mg g-1)(L mg-1)1/n) and n are Freundlich
28
characteristic constants related to the adsorption capacity and adsorption intensity,

respectively. The regression coefficient (R2) was utilized as the factor that validates the

isotherm model.

The fitting curves for the predicted isotherms models for the temperature of 5°C are

presented in Fig. 14, and the calculated isotherm parameters are listed in Table 2.

Table 2 – Langmuir and Freundlich models parameters

T (C) Langmuir model Freundlich model

qm KL R2 KF 1/n R2

5 40.78954 0.28634 0.73853 18.07902 0.19203 0.999043

15 36.54044 0.27288 0.68077 15.86366 0.19618 0.9856

25 33.44692 0.26008 0.83176 14.69927 0.191 0.98635

35 30.26586 0.25876 0.86842 13.33627 0.19618 0.9967

45 27.77169 0.23498 0.98705 11.8094 0.2027 0.97841

Reasonable agreement was obtained between the experimental data and the different models

predicted isotherms for glyphosate adsorption. The adsorption data fitted better by the

Freundlich model than by the Langmuir model, suggesting that glyphosate adsorption on

MnFe2O4-G is multilayer with heterogeneous surface. The Freundlich constant (1/n) is related

to the adsorption intensity, being easy to adsorb when 0.1<1/n≤ 0.5, having some difficulties

with the absorption when 0.5≤1/n≤1 and being difficult to adsorb when 1/n>1 [54]. The

obtained results indicate that the glyphosate is easily adsorbed by the MnFe2O4-G composite.

This result can be confirmed by the morphology characterization, that is evidently

heterogeneous surface. The Langmuir isotherm model assumes that a monolayer adsorption
29
exists on the adsorption surface with a finite number of identical sites, that are energetically

equivalent and there is no interaction between the adsorbed molecules [53].

Fig. 14 – Isotherm models fitting for Glyphosate adsorption on MnFe2O4-G hybrids at 5°C.

The favorability of the adsorption process between glyphosate and the adsorbent can be

assumed with the Langmuir dimensionless constant separation factor (RL) according with the

following equation:
0
@A  02 (11)
B "$

Where Ce is the equilibrium concentration of adsorbate (mg L-1), KL is the Langmuir constant.

The value of RL indicates if the type of isotherm is unfavorable (RL>1), linear (RL=1) or

favorable (0<RL<1). The calculated values are shown in Table 3.

30
Table 3 – RL values calculated for all temperatures and concentrations.

Temperature 5°C 15°C 25°C 35°C 45°C

C0 (mg L-1) RL

5 0.998 0.996 0.993 0.989 0.992

10 0.983 0.978 0.961 0.935 0.908

20 0.479 0.455 0.433 0.424 0.379

30 0.234 0.226 0.229 0.225 0.226

40 0.169 0.162 0.165 0.157 0.161

60 0.087 0.083 0.083 0.084 0.087

80 0.054 0.056 0.058 0.058 0.060

The RL values were found to be favorable adsorption of glyphosate onto MnFe2O4-G composite

at all different initial concentrations and temperatures, varying in the range of 0.054 to 0.998

(0<RL<1).

3.3 Thermodynamic Parameters

It is proposed in this section to estimate thermodynamic parameters. To determine the

thermodynamic parameters and the temperature effect on glyphosate adsorption onto

MnFe2O4-G composite, adsorption experiments were conducted at five different

temperatures, 5-45°C (Fig. 13). It is observed that the adsorption process is temperature

dependent. The standard Gibb’s free energy change, ΔG°, can be calculated by the following

equation:

∆ D  −@E, /" (12)

31
Where R is the universal gas constant (8.314 J mol-1 K-1), and T is the absolute temperature

(K). Thermodynamic equilibrium constant KC is defined by:



/F  "$ (13)
$

That is also known as Henry adsorption equation at its linear isotherm equation qe = KCCe. KC

can be calculated by plotting ln (q e/Ce) as a function of qe. This method presupposes that at

low concentration, the sorption isotherm is a straight line and it is known as equilibrium

constant method [55].

The average standard enthalpy change (ΔH°) and entropy change (ΔS°) can be calculated

according to Van’t Hoff equation:

∆G H ∆J H
, /"  I
− IK
(14)

Plotting ln(KC) against 1/T gives a straight line with slope and intercept equal to –ΔH/R and

ΔS/R, respectively. The thermodynamic parameters are reported in Table 4.

Table 4 - Thermodynamic parameters

Temperature (°C) ΔG° (kJ mol-1) ΔH° (kJ mol-1) ΔS° (J mol-1 K-1)

5 -4.737

15 -4.733

25 -4.789 -5.764 -3.593

35 -4.555

45 -4.642

32
As observed, the negative values of ΔG° imply that the adsorption of glyphosate by the hybrid

composite is a spontaneous process. The negative standard ΔH° suggests that the interaction

between glyphosate and MnFe2O4-G is an exothermic process, which is supported by the

increase of adsorption with the decrease of the temperatures [52].

3.4 Effect of competing anions

Water and wastewater treatment requires a competing anion studies. There exist many ions
in contaminated water that can affect the adsorption of the target ion contaminant. These ions
compete to bind to the adsorbent surface, affecting the efficiency of the adsorption of the
target ion [16]. The effects of the investigated competing anions are summarized in Table 5.
The removal of glyphosate was affected by the concentration of different anions. Interestingly,
the adsorption of glyphosate was positively influenced in the presence of fluoride, chloride,
nitrate and sulfate even at the highest concentration investigated, i.e. 1 mM anion. This
behavior can be attributed to the adsorption of glyphosate as strong complexes on the surface
of MnFe2O4-G [56]. The increase in anions concentration reduced the electrostatic repulsion
between the negative charge of glyphosate ions and the MnFe2O4-G surfaces leading to an
increase in the number of collisions between glyphosate ions and MnFe2O4-G [57], thus
increasing the glyphosate adsorptive removal.
However, phosphate, known as a strong competing ion, showed some competitive effect on
the adsorption of glyphosate. Using 20 ppm (≈0.1 mM) of glyphosate there was a very slight
decrease of adsorption removal when using the same amount of phosphate anions (0.1 mM).
Increasing phosphate concentration to 10 folds that of glyphosate (1 mM) resulted in a
decrease in glyphosate removal, reaching 51%. This behavior can be related to the strong
phosphate adsorption on iron oxides via ligand exchange with hydroxyl groups [58].
In summary, the obtained results showed that most anions did not compete with glyphosate
adsorption, and instead they improved its removal using MnFe2O4-G nanohybrid as
adsorbent. Therefore, these results are very positive because in polluted waters, there are
many different co-existing anions which can compensate for the slight decrease registered in
the presence of phosphate, thus contributing to maintain the promising performances of
MnFe2O4-G nanohybrid adsorbent.

33
Table 5 – Effect of competing anions on the adsorption of glyphosate on MnFe2O4-G
Competing % of glyphosate adsorbed on MnFe2O4-G (1 g L-1)
ion
concentration F- Cl- NO3- SO42- H2PO4-
(mM)
0 80 80 80 80 80
0.1 86 86 81 90 72
1 92 88 91 91 51

4. Conclusions

Magnetic MnFe2O4-G hybrid composite was successfully synthesized via a facile solvothermal

method. The material was characterized using various techniques for their crystal structures

and surface properties (XRD, FTIR, XPS, SEM, TEM, EDX), and then applied to remove

glyphosate from water. The hybrid composite presented a large specific area observed from

BET analysis, 305.5 m2/g, due to the well dispersed microspheres on the graphene

nanosheets observed in the SEM and TEM analysis. The proposed adsorption mechanism of

glyphosate on MnFe2O4-G is through a combination of an electrostatic interaction with an ion

exchange. The adsorption capacity of MnFe2O4-G hybrids to remove glyphosate from aqueous

solutions was investigated. The results indicated that the MnFe2O4-G hybrid showed an

excellent adsorption capacity and a high adsorption rate for glyphosate. The maximum

adsorption capacity of MnFe2O4-G for glyphosate was 39 mg/g at 5°C. Low temperature was

determined to be favorable for the adsorption. The kinetics and isotherm data fitted well with

pseudo-second-order kinetics and the Freundlich isotherm model, respectively. The

adsorption of glyphosate on the MnFe2O4-G hybrid was determined to be a spontaneous

exothermic process. This study demonstrated that the new synthesized magnetic material

MnFe2O4–G hybrids proved to be effective as a potential adsorbent for the removal of

34
glyphosate from aqueous solution and a promising adsorbent for further utilization in water

purification.

Acknowledgements

The authors would like to thank NSERC (Natural Sciences and Engineering Research Council

of Canada), CFI (Canada Foundation for Innovation) and CAPES (Coordenação de

Aperfeiçoamento de Pessoal de Nível Superior) for supporting this project. Alain Adnot

(CERPIC) and Richard Janvier are acknowledged for XPS and TEM characterizations. NUY is

grateful for Emerging Leaders in the Americas Program (ELAP) of Canada Government for the

generous scholarship.

References

[1] F. Li, Y. Wang, Q. Yang, D.G. Evans, C. Forano, X. Duan, Study on adsorption of glyphosate

(N-phosphonomethyl glycine) pesticide on MgAl-layered double hydroxides in aqueous

solution, Journal of Hazardous Materials 125 (2005) 89-95.

[2] G.A. Khoury, T.C. Gehris, L. Tribe, R.M. Torres Sánchez, M. dos Santos Afonso, Glyphosate

adsorption on montmorillonite: An experimental and theoretical study of surface complexes,

Applied Clay Science 50 (2010) 167-175.

[3] K. Solomon, D. Thompson, Ecological Risk Assessment for Aquatic Organisms from Over-

Water Uses of Glyphosate, Journal of Toxicology and Environmental Health, Part B 6 (2003)

289-324.

[4] M. Hagner, O.-P. Penttinen, K. Tiilikkala, H. Setälä, The effects of biochar, wood vinegar and

plants on glyphosate leaching and degradation, European Journal of Soil Biology 58 (2013) 1-

7.

35
[5] J. Sanchis, L. Kantiani, M. Llorca, F. Rubio, A. Ginebreda, J. Fraile, T. Garrido, M. Farre,

Determination of glyphosate in groundwater samples using an ultrasensitive immunoassay

and confirmation by on-line solid-phase extraction followed by liquid chromatography

coupled to tandem mass spectrometry, Analytical and bioanalytical chemistry 402 (2012)

2335-2345.

[6] N. Lin, V.F. Garry, In vitro studies of cellular and molecular developmental toxicity of

adjuvants, herbicides, and fungicides commonly used in Red River Valley, Minnesota, Journal

of toxicology and environmental health. Part A 60 (2000) 423-439.

[7] F. Manas, L. Peralta, J. Raviolo, H.G. Ovando, A. Weyers, L. Ugnia, M.G. Cid, I. Larripa, N.

Gorla, Genotoxicity of glyphosate assessed by the comet assay and cytogenetic tests,

Environmental toxicology and pharmacology 28 (2009) 37-41.

[8] A. Paganelli, V. Gnazzo, H. Acosta, S.L. Ló pez, A.s.E. Carrasco, Glyphosate-Based Herbicides

Produce Teratogenic Effects on Vertebrates by Impairing Retinoic Acid Signaling, Chemical

Research in Toxicology 23 (2010) 1586-1595.

[9] J.M. Salman, F.M. Abid, Preparation of mesoporous activated carbon from palm-date pits:

optimization study on removal of bentazon, carbofuran, and 2,4-D using response surface

methodology, Water science and technology : a journal of the International Association on

Water Pollution Research 68 (2013) 1503-1511.

[10] M. Ghaedi, A. Ansari, M.H. Habibi, A.R. Asghari, Removal of malachite green from aqueous

solution by zinc oxide nanoparticle loaded on activated carbon: Kinetics and isotherm study,

Journal of Industrial and Engineering Chemistry 20 (2014) 17-28.

[11] S.S. Mayakaduwa, P. Kumarathilaka, I. Herath, M. Ahmad, M. Al-Wabel, Y.S. Ok, A. Usman,

A. Abduljabbar, M. Vithanage, Equilibrium and kinetic mechanisms of woody biochar on

aqueous glyphosate removal, Chemosphere (2015).

[12] Y.S. Hu, Y.Q. Zhao, B. Sorohan, Removal of glyphosate from aqueous environment by

adsorption using water industrial residual, Desalination 271 (2011) 150-156.


36
[13] M. Milojević-Rakić, A. Janošević, J. Krstić, B. Nedić Vasiljević, V. Dondur, G. Ćirić-

Marjanović, Polyaniline and its composites with zeolite ZSM-5 for efficient removal of

glyphosate from aqueous solution, Microporous and Mesoporous Materials 180 (2013) 141-

155.

[14] A. Khenifi, Z. Derriche, C. Mousty, V. Prévot, C. Forano, Adsorption of Glyphosate and

Glufosinate by Ni2AlNO3 layered double hydroxide, Applied Clay Science 47 (2010) 362-371.

[15] A. Martín, A. Escarpa, Graphene: The cutting–edge interaction between chemistry and

electrochemistry, TrAC Trends in Analytical Chemistry 56 (2014) 13-26.

[16] S. Kumar, R.R. Nair, P.B. Pillai, S.N. Gupta, M.A.R. Iyengar, A.K. Sood, Graphene Oxide–

MnFe2O4 Magnetic Nanohybrids for Efficient Removal of Lead and Arsenic from Water, ACS

Applied Materials & Interfaces 6 (2014) 17426-17436.

[17] C. Wang, C. Feng, Y. Gao, X. Ma, Q. Wu, Z. Wang, Preparation of a graphene-based magnetic

nanocomposite for the removal of an organic dye from aqueous solution, Chemical

Engineering Journal 173 (2011) 92-97.

[18] Y. Zhang, B. Chen, L. Zhang, J. Huang, F. Chen, Z. Yang, J. Yao, Z. Zhang, Controlled

assembly of Fe3O4 magnetic nanoparticles on graphene oxide, Nanoscale 3 (2011) 1446-

1450.

[19] N.I. Kovtyukhova, P.J. Ollivier, B.R. Martin, T.E. Mallouk, S.A. Chizhik, E.V. Buzaneva, A.D.

Gorchinskiy, Layer-by-Layer Assembly of Ultrathin Composite Films from Micron-Sized

Graphite Oxide Sheets and Polycations, Chemistry of Materials 11 (1999) 771-778.

[20] S.M. Maliyekkal, T.S. Sreeprasad, D. Krishnan, S. Kouser, A.K. Mishra, U.V. Waghmare, T.

Pradeep, Graphene: A Reusable Substrate for Unprecedented Adsorption of Pesticides, Small 9

(2013) 273-283.

[21] W. Cai, T. Lai, W. Dai, J. Ye, A facile approach to fabricate flexible all-solid-state

supercapacitors based on MnFe2O4/graphene hybrids, Journal of Power Sources 255 (2014)

170-178.
37
[22] S. Chella, P. Kollu, E.V.P.R. Komarala, S. Doshi, M. Saranya, S. Felix, R. Ramachandran, P.

Saravanan, V.L. Koneru, V. Venugopal, S.K. Jeong, A. Nirmala Grace, Solvothermal synthesis of

MnFe2O4-graphene composite—Investigation of its adsorption and antimicrobial properties,

Applied Surface Science 327 (2015) 27-36.

[23] Z. Li, K. Gao, G. Han, R. Wang, H. Li, X.S. Zhao, P. Guo, Solvothermal synthesis of MnFe2O4

colloidal nanocrystal assemblies and their magnetic and electrocatalytic properties, New

Journal of Chemistry 39 (2015) 361-368.

[24] H. Liu, S. Ji, Y. Zheng, M. Li, H. Yang, Modified solvothermal synthesis of magnetic

microspheres with multifunctional surfactant cetyltrimethyl ammonium bromide and directly

coated mesoporous shell, Powder Technology 246 (2013) 520-529.

[25] Y. Yao, S. Miao, S. Liu, L.P. Ma, H. Sun, S. Wang, Synthesis, characterization, and adsorption

properties of magnetic Fe3O4@graphene nanocomposite, Chemical Engineering Journal 184

(2012) 326-332.

[26] Y. Yao, Y. Cai, F. Lu, F. Wei, X. Wang, S. Wang, Magnetic recoverable MnFe2O4 and

MnFe2O4-graphene hybrid as heterogeneous catalysts of peroxymonosulfate activation for

efficient degradation of aqueous organic pollutants, Journal of Hazardous Materials 270

(2014) 61-70.

[27] P. Xiong, C. Hu, Y. Fan, W. Zhang, J. Zhu, X. Wang, Ternary manganese

ferrite/graphene/polyaniline nanostructure with enhanced electrochemical capacitance

performance, Journal of Power Sources 266 (2014) 384-392.

[28] K. Wu, G. Hu, Y. Cao, Z. Peng, K. Du, Facile and green synthesis of MnFe2O4/reduced

graphene oxide nanocomposite as anode materials for Li-ion batteries, Materials Letters 161

(2015) 178-180.

[29] X. Bao, Z. Qiang, W. Ling, J.-H. Chang, Sonohydrothermal synthesis of MFe2O4 magnetic

nanoparticles for adsorptive removal of tetracyclines from water, Separation and Purification

Technology 117 (2013) 104-110.


38
[30] K.V. Sankar, R.K. Selvan, The ternary MnFe2O4/graphene/polyaniline hybrid composite

as negative electrode for supercapacitors, Journal of Power Sources 275 (2015) 399-407.

[31] Y. Sheng, X. Tang, E. Peng, J. Xue, Graphene oxide based fluorescent nanocomposites for

cellular imaging, Journal of Materials Chemistry B 1 (2013) 512-521.

[32] G. Srinivas, Y. Zhu, R. Piner, N. Skipper, M. Ellerby, R. Ruoff, Synthesis of graphene-like

nanosheets and their hydrogen adsorption capacity, Carbon 48 (2010) 630-635.

[33] B. Saner, F. Okyay, Y. Yürüm, Utilization of multiple graphene layers in fuel cells. 1. An

improved technique for the exfoliation of graphene-based nanosheets from graphite, Fuel 89

(2010) 1903-1910.

[34] L. Chen, S.-F. Yin, S.-L. Luo, R. Huang, Q. Zhang, T. Hong, P.C.T. Au, Bi2O2CO3/BiOI

Photocatalysts with Heterojunctions Highly Efficient for Visible-Light Treatment of Dye-

Containing Wastewater, Industrial & Engineering Chemistry Research 51 (2012) 6760-6768.

[35] R. Yuan, S.N. Ramjaun, Z. Wang, J. Liu, Effects of chloride ion on degradation of Acid

Orange 7 by sulfate radical-based advanced oxidation process: Implications for formation of

chlorinated aromatic compounds, Journal of Hazardous Materials 196 (2011) 173-179.

[36] M.-Q. Yang, B. Weng, Y.-J. Xu, Improving the Visible Light Photoactivity of In2S3–

Graphene Nanocomposite via a Simple Surface Charge Modification Approach, Langmuir 29

(2013) 10549-10558.

[37] H. Beitollahi, M. Hamzavi, M. Torkzadeh-Mahani, Electrochemical determination of

hydrochlorothiazide and folic acid in real samples using a modified graphene oxide sheet

paste electrode, Materials Science and Engineering: C 52 (2015) 297-305.

[38] P. Lian, X. Zhu, S. Liang, Z. Li, W. Yang, H. Wang, Large reversible capacity of high quality

graphene sheets as an anode material for lithium-ion batteries, Electrochimica Acta 55 (2010)

3909-3914.

39
[39] N. Pan, D. Guan, Y. Yang, Z. Huang, R. Wang, Y. Jin, C. Xia, A rapid low-temperature

synthetic method leading to large-scale carboxyl graphene, Chemical Engineering Journal 236

(2014) 471-479.

[40] R.H. Vignesh, K.V. Sankar, S. Amaresh, Y.S. Lee, R.K. Selvan, Synthesis and characterization

of MnFe2O4 nanoparticles for impedometric ammonia gas sensor, Sensors and Actuators B:

Chemical 220 (2015) 50-58.

[41] Y. Zhou, B. Xiao, S.-Q. Liu, Z. Meng, Z.-G. Chen, C.-Y. Zou, C.-B. Liu, F. Chen, X. Zhou, Photo-

Fenton degradation of ammonia via a manganese–iron double-active component catalyst of

graphene–manganese ferrite under visible light, Chemical Engineering Journal 283 (2016)

266-275.

[42] Y.-P. Chang, C.-L. Ren, J.-C. Qu, X.-G. Chen, Preparation and characterization of

Fe3O4/graphene nanocomposite and investigation of its adsorption performance for aniline

and p-chloroaniline, Applied Surface Science 261 (2012) 504-509.

[43] Y. Ren, N. Li, J. Feng, T. Luan, Q. Wen, Z. Li, M. Zhang, Adsorption of Pb(II) and Cu(II) from

aqueous solution on magnetic porous ferrospinel MnFe2O4, Journal of Colloid and Interface

Science 367 (2012) 415-421.

[44] V. Chandra, J. Park, Y. Chun, J.W. Lee, I.-C. Hwang, K.S. Kim, Water-Dispersible Magnetite-

Reduced Graphene Oxide Composites for Arsenic Removal, ACS Nano 4 (2010) 3979-3986.

[45] J. Xiao, G. Xu, S.-G. Sun, S. Yang, MFe2O4 and MFe@Oxide Core–Shell Nanoparticles

Anchored on N-Doped Graphene Sheets for Synergistically Enhancing Lithium Storage

Performance and Electrocatalytic Activity for Oxygen Reduction Reactions, Particle & Particle

Systems Characterization 30 (2013) 893-904.

[46] H. Kim, D.-H. Seo, H. Kim, I. Park, J. Hong, K.-Y. Park, K. Kang, Multicomponent Effects on

the Crystal Structures and Electrochemical Properties of Spinel-Structured M3O4 (M = Fe, Mn,

Co) Anodes in Lithium Rechargeable Batteries, Chemistry of Materials 24 (2012) 720-725.

40
[47] A. Prakash, S. Chandra, D. Bahadur, Structural, magnetic, and textural properties of iron

oxide-reduced graphene oxide hybrids and their use for the electrochemical detection of

chromium, Carbon 50 (2012) 4209-4219.

[48] J. Hu, Lo, G. Chen, Fast Removal and Recovery of Cr(VI) Using Surface-Modified Jacobsite

(MnFe2O4) Nanoparticles, Langmuir 21 (2005) 11173-11179.

[49] M.S. Podder, C.B. Majumder, SD/MnFe2O4 composite, a biosorbent for As(III) and As(V)

removal from wastewater: Optimization and isotherm study, Journal of Molecular Liquids 212

(2015) 382-404.

[50] K.A. Barrett, M.B. McBride, Oxidative Degradation of Glyphosate and

Aminomethylphosphonate by Manganese Oxide, Environmental Science & Technology 39

(2005) 9223-9228.

[51] S. Yang, L. Li, Z. Pei, C. Li, J. Lv, J. Xie, B. Wen, S. Zhang, Adsorption kinetics, isotherms and

thermodynamics of Cr(III) on graphene oxide, Colloids and Surfaces A: Physicochemical and

Engineering Aspects 457 (2014) 100-106.

[52] L. Fan, C. Luo, X. Li, F. Lu, H. Qiu, M. Sun, Fabrication of novel magnetic chitosan grafted

with graphene oxide to enhance adsorption properties for methyl blue, Journal of Hazardous

Materials 215–216 (2012) 272-279.

[53] Q. Wu, C. Feng, C. Wang, Z. Wang, A facile one-pot solvothermal method to produce

superparamagnetic graphene–Fe3O4 nanocomposite and its application in the removal of dye

from aqueous solution, Colloids and Surfaces B: Biointerfaces 101 (2013) 210-214.

[54] T. Liu, Y. Li, Q. Du, J. Sun, Y. Jiao, G. Yang, Z. Wang, Y. Xia, W. Zhang, K. Wang, H. Zhu, D. Wu,

Adsorption of methylene blue from aqueous solution by graphene, Colloids and Surfaces B:

Biointerfaces 90 (2012) 197-203.

[55] S. Salvestrini, V. Leone, P. Iovino, S. Canzano, S. Capasso, Considerations about the correct

evaluation of sorption thermodynamic parameters from equilibrium isotherms, The Journal of

Chemical Thermodynamics 68 (2014) 310-316.


41
[56] S. Goldberg, Inconsistency in the triple layer model description of ionic strength

dependent boron adsorption, Journal of Colloid and Interface Science 285 (2005) 509-517.

[57] X.-j. Hu, Y.-g. Liu, H. Wang, G.-m. Zeng, X. Hu, Y.-m. Guo, T.-t. Li, A.-w. Chen, L.-h. Jiang, F.-y.

Guo, Adsorption of copper by magnetic graphene oxide-supported β-cyclodextrin: Effects of

pH, ionic strength, background electrolytes, and citric acid, Chemical Engineering Research

and Design 93 (2015) 675-683.

[58] D. Jia, C. Zhou, C. Li, Adsorption of Glyphosate on Resin Supported by Hydrated iron

Oxide: Equilibrium and Kinetic Studies, Water Environment Research 83 (2011) 783-790.

42
• This is the first application of graphene-based adsorbent for glyphosate removal from water
• The magnetic character of the adsorbent makes it easily separable
• A mechanism of glyphosate adsorption was proposed

43

You might also like