Functionalization of Magnetite Nanoparticles as Oil Spill ... - MDPI

1 downloads 0 Views 1MB Size Report
Mar 26, 2015 - Iron ions and iodine were prepared by the reaction between ferric chloride and potassium iodide. The structure and morphology of magnetite ...
Int. J. Mol. Sci. 2015, 16, 6911-6931; doi:10.3390/ijms16046911 OPEN ACCESS

International Journal of

Molecular Sciences ISSN 1422-0067 www.mdpi.com/journal/ijms Article

Functionalization of Magnetite Nanoparticles as Oil Spill Collector Ayman M. Atta 1,2,*, Hamad A. Al-Lohedan 1 and Sami A. Al-Hussain 1,3 1

2 3

Surfactants Research Chair, Chemistry Department, College of Science, King Saud University, Riyadh 11541, Saudi Arabia; E-Mails: [email protected] (H.A.A.-L.); [email protected] (S.A.A.-H.) Egyptian Petroleum Research Institute, 1 Ahmad Elzomor St., Nasr City, Cairo 11727, Egypt Chemistry Department, Faculty of Science, Al-Imam Muhammad Bin Saud Islamic University, Riyadh 11632, Saudi Arabia

* Author to whom correspondence should be addressed; E-Mail: [email protected] or [email protected]; Tel.: +966-11-467-5998; Fax: +966-11-467-5992. Academic Editor: O. Thompson Mefford Received: 17 February 2015 / Accepted: 20 March 2015 / Published: 26 March 2015

Abstract: In the present study, a new magnetic powder based on magnetite can be used as a petroleum crude oil collector. Amidoximes based on rosin as a natural product can be prepared from a reaction between hydroxylamine and rosin/acrylonitrile adducts. The produced rosin amidoximes were used as capping agents for magnetite nanoparticles to prepare hydrophobic coated magnetic powders. A new class of monodisperse hydrophobic magnetite nanoparticles was prepared by a simple and inexpensive co-precipitation method. Iron ions and iodine were prepared by the reaction between ferric chloride and potassium iodide. The structure and morphology of magnetite capped with rosin amidoxime were characterized by Fourier transform infrared spectroscopy (FTIR), X-ray diffraction (XRD), transmission electron microscopy (TEM), zeta potential, thermogravimetric analysis (TGA) and dynamic light scattering (DLS). The magnetic properties were determined from vibrating sample magnetometer (VSM) analyses. These prepared magnetite nanoparticles were tested as bioactive nanosystems and their antimicrobial effects were investigated. The prepared nanomaterials were examined as a crude oil collector using magnetic fields. The results show promising data for the separation of the petroleum crude oil from aqueous solution in environmental pollution cleanup.

Int. J. Mol. Sci. 2015, 16

6912

Keywords: amidoxime; rosin; coated magnetite; magnetic properties; oil collectors; magnetic fluids

1. Introduction Magnetite nanoparticles and their combinations (core—shell NPs and other composite nanostructures) has attracted great attention in environmental and medical applications [1–5]. In order to enhance the performance of magnetite nanoparticles, they must satisfy certain criteria such as minimal cytotoxicity, excellent colloidal stability under physiological conditions, and particle shape and size distribution [6–8]. The chemical modification of the magnetite NP surface is very important to achieve the necessary domains of applications. However, control over particle size distribution, colloidal stability and low yielding multistep cost effective preparation methods are still limited for many magnetic NP systems. It is well known that the crystal structure and the cessation of the growth of NPs are fundamentally regulated by the ligands such as surfactants or modified polymers [9–12]. It can be reported that the magnetite nanoparticles should form physically and chemically stable colloidal suspensions in which the NPs do not aggregate, dissociate, or chemically react with the surround or any dissolved gas with time. The advantage of small molecules as ligands is that they offer a certain degree of physical barrier, similar to polymeric ligands, but give a smaller hydrodynamic radius [1]. There are several methods used to synthesize magnetite nanoparticles (MNPs) in solvent, gel template, and solvent-free methods, such as chemical vapor deposition, electrical explosion and mechanical milling [13–19]. Water soluble NPs may be prepared in a one-step synthesis in organic solvents by judicious choice of stabilization agents. However, raw MNPs are easily aggregated in aqueous solutions due to anisotropic dipolar attraction between particles, which restricts the application range [20]. The surface modifications of MNPs by covalent binding or physical coating have been widely explored to prevent these shortages; it was reported that they can serve as a more ideal shell component for magnetite due to stability under acidic conditions, inertia to redox reactions, and introduced surface hydroxyl groups which have a strong ability to link special functional groups [21–25]. It is well known that amidoxime, an excellent amphoteric functional group due to the presence of both acidic and basic sites, has been grafted onto the surfaces of magnetite-coated silica for recovery and removal of uranium U (VI) with fast sorption rate, high U (VI) loading capacity, and safety of the environment [4]. In our previous work, we succeeded to prepare capped magnetite using an inexpensive one-step method [26]. The objective of the present work is to synthesis amidoxime from natural products derived from rosin gum to prepare highly dispersed magnetite nanoparticles using an inexpensive one-step method. The effect of amidoxime organic coatings on magnetic properties of magnetite is another objective of the present work. 2. Results and Discussion It is well known that magnetite (Fe3O4) nanoparticles can be oxidized by air oxidation and are easily aggregated in aqueous solutions due to anisotropic dipolar attraction, which restricts its applications [27]. The surface modifications of magnetite using covalent bonding or physical coating can be used to

Int. J. Mol. Sci. 2015, 16

6913

compensate for these shortages without reduction of the magnetic properties of the nanoparticles. In this work, we use amidoxime based on natural gum, such as rosin, to bind with magnetite due to the excellent amphoteric functional group, the presence of both acidic and basic sites and safety of the environment. The reaction scheme to prepare magnetite coated rosin amidoximes is illustrated in Scheme 1.

Scheme 1. Synthesis of magnetite coated rosin amidoxime. The rosin amidoximes can be prepared by formation of Diels Alder adducts between rosin and acrylonitrile (AN) followed by reaction of adducts with hydroxylamine in basic medium as described in the Experimental Section and represented in Scheme 2.

Scheme 2. Synthesis of R-AN and RK-AN amidoximes.

Int. J. Mol. Sci. 2015, 16

6914

The cycloaddition reaction between RA and AN was completed by isomerization of abietic acid (the major constituent of RA) into leveopimaric acid at a high reaction temperature (185–190 °C) and with a prolonged reaction duration (10 h) developed with high yields (ca. 95%) [28]. The isomerization occurred and developed together with the cycloaddition reaction. The RK-AN was produced by dehydrodecarboxylation R-AN in the presence of p-toluene sulfonic acid (PTSA as catalyst at temperature of 240 °C as illustrated in the Experimental Section and represented in the Scheme 2. The cyanide functional groups of R-AN and RK-AN were reacted with hydroxylamine in an alkaline medium to yield amidoxime groups. The mole ratio of hydroxylamine was in excess to R-AN and RK-AN, the reaction was done for 2 h at 70 °C. A mixture of methanol-water ratio 5:1 was used as reaction medium. Upon amidoximation, the nitrogen content of R-AN and RK-AN changed from 3.93% and 4.33% to 7.31% and 7.91%, respectively. The data of the experimental nitrogen contents of R-AN and RK-AN and their amidoxime derivatives agree with the theoretical calculation which indicated the purity of the prepared rosin amidoxime derivatives and supported the rosin functionalization by post reaction. The FTIR spectra of R-AN was selected as representative for formation of Diels Alder adduct between rosin and AN represented in Figure 1. Moreover, 1H and 13C NMR spectra of RK-AN amidoximes are selected and represented in Figures 2 and 3. FT-IR spectra of R-AN and RK-AN (not represented for brevity; Figure 1), show the characteristic absorption band at 2245 cm−1 due to CN stretching modes in addition of absorption bands of rosin which included C–H stretching at 2936–2870 cm−1, C=O stretching at 1694 cm−1 and C=C stretching in hydrophenanetherene moiety at 1462 cm−1. After amidoxime group preparation (Figure 1B), the CN band at 2245 cm−1 disappeared, and a new band of amidoxime at 1652 cm−1, and amide II band of N–H at 1568 cm−1, respectively, appeared. FTIR spectra confirm that the nitrile groups are converted to amidoxime groups under the previously described experimental conditions. The band at 938 cm−1 may be assigned to the N–O bond of the amidoxime group [29]. The chemical structures of the produced R-AN, RK-AN amidoxime were confirmed by 1H NMR and 13C NMR as represented in Figures 2 and 3 respectively. The 1H NMR spectra of R-AN, RK-AN amidoxime (Figure 2) show new R-AN, RK-AN amidoxime peaks at 2.2, 3.5 and 5.7 ppm, which are due to hydrogen atoms of OH, CH (CH–C=N(NH2)–OH) and NH2 amidoxime groups, respectively. The presence of some residual cyanoethyl groups was confirmed by the band at 2.8 ppm, which corresponds to the –CH2–CN methylene protons. Figure 3 shows the 13C NMR spectra of the R-AN and RK-AN amidoximes. The appearance of the signals at 185.30 ppm, characteristic of carboxyl carbon, and at 122.73 ppm, characteristic of nitrile carbon, demonstrate the performance of the addition reaction between rosin and AN. The signals at 122.22 and 150.71 ppm show the unsaturated carbons from the hydrophenanetherene unit, and the signals in the range of 16.14–60.33 ppm evidence the methyl, methylene, methine, and quaternary carbons from the same hydrophenanetherene unit and the remaining hydrocarbons belonging to AN. The disappearance of signals at 19.7, 64.5 and 121.5 ppm and appearance of new signals at 32.4, 66.9 and 156.3 ppm correspond to the amidoxime groups indicated the conversion of all cyano group to amidoxime.

Int. J. Mol. Sci. 2015, 16

Figure 1. FTIR spectra of (A) R-AN and (B) RK-AN amidoxime derivatives.

6915

Int. J. Mol. Sci. 2015, 16

6916

Figure 2. 1H NMR spectrum of R-AN amidoxime.

Figure 3. 13C NMR spectrum of RK-AN amidoxime. 2.1. Chemical Structure of Amidoxime-Coated Magnetite Nanoparticles Herein we extend the previous works [26,30–32] to modify the surface of magnetite nanoparticles using a new method for synthesizing stabilized magnetic nanoparticle colloids. A new class of monodisperse water-soluble magnetite nanoparticles was prepared by a simple and cheap co-precipitation method. Iron ions and iodine were prepared by the reaction between ferric chloride and potassium iodide as represented

Int. J. Mol. Sci. 2015, 16

6917

in the Experimental Section. The ferrous and ferric ions were hydrolyzed at low temperature at pH 9 in the presence of iodine to produce iron oxide nanoparticles. Moreover, amidoxime derivatives of R-AN and RK-AN were used as capping agent to produce highly dispersed coated magnetite nanoparticles. The condensation of R-on surfaces of Fe3O4 cores resulted in core-shell microspheres as shown in Scheme 1. In our previous work [26], we found that the presence of iodine during formation of magnetite nanoparticles modifies the surface of magnetite with hydroxyl groups. In the present work, it was expected that amidoxime derivatives of R-AN and RK-AN can be bonded either with covalent bonding or with physical adsorption of amidoxime derivatives with magnetite surfaces as represented in Scheme 1. It was postulated that the amidoxime or carboxylic groups in amidoxime derivatives of R-AN and RK-AN are responsible for its chemical adsorption on the surface of bare magnetite. The carboxylic functional group of R-AN amidoxime was used to form covalent bonds to the MNPs using an esterification reaction between OH of magnetite and COOH of R-AN amidoxime. Moreover, the adsorption on magnetite nanoparticles (MNPs) co-precipitated with amidoxime derivatives of R-AN and RK-AN was explained by more complex mechanisms involving intermolecular hydrogen bonding. The structure and morphology of the magnetite capped with amidoxime derivatives of R-AN and RK-AN were characterized by FTIR and XRD analyses. In this respect FTIR spectra of Fe3O4 capped with amidoxime derivatives of R-AN and RK-AN were represented in Figure 4. In the previous work [26] with FTIR spectrum of uncapped iron oxides, not represented here for brevity, indicated the presence of α-Fe2O3 (hematite) beside magnetite nanoparticles. Figure 4 indicated that the presence of amidoxime derivatives of R-AN and RK-AN during the formation of magnetite prevent its air oxidation and produces Fe3O4 nanoparticles without any contamination of other iron oxides. This observation was confirmed by the appearance of a strong absorption band at 576 cm−1 which was assigned to the Fe–O vibration frequency of magnetite. Moreover, new absorption band at 1745 cm−1 for R-AN amidoxime (Figure 4A) indicated the formation of an ester link between COOH of R-AN amidoxime and OH of magnetite. The appearance of the band arises at 1652 cm−1 corresponding to the C=N stretching vibrations of amidoxime groups. The presence of absorption bands around 2865 and 2940 cm−1 corresponded to symmetrical and asymmetrical of CH2 stretching vibrations of the rosin hydrophenancerene ring. A broad absorption bands was found around 3299 and 3450 cm−1 that attributed to stretching vibration of hydroxyl groups and NH2 in the amidoxime derivatives of R-AN and RK-AN molecules. Consequently, these FT-IR spectra provided supportive evidence that amidoxime derivatives of R-AN and RK-AN attached to Fe3O4 nanoparticles via the interaction between magnetite and C=N, OH and NH2 groups of amidoxime derivatives of R-AN and RK-AN.

Figure 4. FTIR spectra of magnetite coated with (A) R-AN and (B) RK-AN amidoximes.

Int. J. Mol. Sci. 2015, 16

6918

XRD analysis is used to detect the phase and crystal shape of the capped and uncapped magnetite nanoparticles. The effect of capping agents on the magnetite crystal shapes was previously discussed [33]. It was found that the amount of capping agent such as sodium dodecyl sulfate (SDS) changed the morphology of magnetite from nanospheres to nanoneedles and to nanocubes by varying the amount of SDS. In previous work [26] we observed that α-phase of iron oxide nanoparticles (α-Fe2O3, hematite) can be formed during the preparation process of magnetite in the presence of iodine. In the present work, the XRD diffractograms of magnetite capped with amidoxime derivatives of R-AN and RK-AN were represented in Figure 5. XRD patterns of Figure 5 displayed eight characteristic peaks at 2θ = 20.12°, 30.17°, 35.53°, 43.16°, 53.51°, 57.27°, 62.77°, and 73.79° revealing a characteristic peaks for Fe3O4 marked by their indices ((111), (220), (311), (400), (422), (511), (440), and (622)). These peaks are consistent with the database in the Joint Committee on Powder Diffraction Standards (JCPDS) file (PDF No. 65-3107). Moreover, that both the two samples show weak diffraction peaks at about 2θ = 20° was attributed to amorphous capping agents [34,35], which was also indicated since Fe3O4 was coated by amidoxime derivatives of R-AN and RK-AN, and its structure was not destroyed. This result indicates that the core-shell microspheres have been successfully synthesized without damaging the crystal structure of the Fe3O4 core during surface functionalization. The position and relative intensity of all diffraction peaks matched well with the standard of the Fe3O4 reflections. The data indicated that no other impurity phase was observed. The formation of a broad peak at 35.53° indicates that the synthesized magnetite nanoparticles have nanospheres [33]. The strong and sharp peaks of R-AN amidoxime revealed that Fe3O4 particles were highly crystallized. The diffraction peak broadening suggested the small size of magnetite nanoparticles. The mean size of Fe3O4 nanocrystallites was estimated using Scherrer’s formula [35]: DS = K × λ/β × cos θ, where DS represents the crystallite size, θ represents the position of the Bragg peak and β is the half-width of the diffraction peak. The constant K in the Scherrer’s equation [35] depends on the morphology of the crystallites and K = 1 is usually assumed. The calculated crystal grain size of the magnetite nanospheres capped with R-AN and RK-AN amidoximes were 22.3, and 33.8 nm, respectively. In the previous work [26], the size of the uncapped iron oxides prepared in the presence of iodine was 6.5 nm. These data confirmed that magnetite was capped with R-AN and RK-AN amidoxime and forms a core-shell structure. The size of magnetite was reduced by using R-AN amidoxime due to the high possibility to form covalent bond with the magnetite surfaces as described in Scheme 1.

Figure 5. XRD diffractograms of magnetite coated with (A) R-AN and (B) RK-AN amidoximes.

Int. J. Mol. Sci. 2015, 16

6919

The amounts of adsorbed layers of R/AN and RK/AN amidoximes or bonded with magnetite can be determined from TGA as presented in Figure 6, an 10.25% weight loss (%) of the magnetite was observed in the temperature range from room temperature to 120 °C. This indicated that water (including bound water) was contained in the magnetite nanoparticles. The R/AN and RK/AN amidoxime shell was decomposed at about 180 °C and finished at about 460 °C. The R/AN and RK/AN extent was determined to be 26.7% and 16.8%, respectively. The magnetite content of Fe3O4 was calculated for coated R/AN and the RK/AN amidoxime to be 73.3% and 83.2% in the dried state, respectively. These data indicated that RK/AN chelated magnetite in their structure more than the R/AN amidoxime shell.

Figure 6. Thermogravimetric thermograms of (A) magnetite; (B) magnetite coated with RK/AN and (C) magnetite coated with R/AN amidoximes. 2.2. Morphology and Particle Size of Magnetite-Coated Nanoparticles The morphology and microstructure of Fe3O4 coated with R-AN and RK-AN amidoxime were characterized using TEM. Figure 7A,B illustrated that the capped Fe3O4 with R-AN and RK-AN amidoxime were spherical-shaped magnetic composites. The figures indicated that Fe3O4 coated with RK-AN amidoxime was essentially through mono-dispersion. Moreover, it was clear that the prepared magnetite nanoparticles could be divided into the core and the shell coating layer, which were assigned to be Fe3O4 and amidoxime, respectively. From previous work [26], the diameters of Fe3O4 was 6.2 nm while the diameter of capped Fe3O4 with R-AN and RK-AN amidoxime were about 15–23 and 32–48 nm, respectively. Figure 6 indicated that the appearance of dark colored Fe3O4 cores and gray colored R-AN and RK-AN amidoxime shells. The clear observation of the core-shell structure was referred to the distinct density contrast between these two components. In addition, surface modification with amidoxime can be hardly identified from the TEM image, which could be attributed to the fact that the modification reaction occurs only on the particle surface and the molecules grafted are not thick enough to be identified by TEM. The schematic presentation of the synthesis process is shown in Scheme 1. Figure 7 shows the TEM images of magnetic rosin amidoximes, in which Fe3O4 particles are closely anchored on the surface of rosin amidoxime rods, suggesting a strong interaction between Fe3O4 particles and R-AN and RK-AN amidoxime rods. The sizes of capped Fe3O4 with R-AN and RK-AN amidoxime nanoparticles calculated from the XRD patterns were in good agreement with the

Int. J. Mol. Sci. 2015, 16

6920

average sizes determined by TEM images (Figure 7). The accordant results about the size indicated the single crystalline nature of Fe3O4 nanoparticles. It is important to measure the particle size diameter and distribution. DLS is one of the most popular methods used to determine the size of particles, particle distribution and for monitoring the colloidal stability of the MNP suspension [36]. As represented in Scheme 1, it was expected that the hydrophilicity of magnetite will be affected by capping with R-AN or RK-AN amidoximes based on core-shell morphology due to chelation of magnetite with amidoxime groups which leads to increase in the hydrophobicity of magnetite by hydrophobic hydrophenancerene of rosin. Surface charges and size distributions of the magnetic nanoparticles are key factors for their dispersability and stability, as magnetic particles with higher charge and smaller size distribution are less likely to combine with each other. This speculation confirmed the formation of amphiphile magnetite having a hydrophilic core and hydrophobic shell. It was found that the prepared magnetite nanoparticles capped with amidoximes cannot disperse easily in water due to hydrophobicity of nanoparticles. It was also expected that the amidoximes are used for both single-layer and bi-layer coating of the iron oxide nanoparticles. The ethanol/water mixture (4/1 volume %) was found to be the suitable solvent to disperse magnetite nanoparticles capped with amidoximes. DLS data of magnetite nanoparticles capped with amidoximes are represented in Figure 8. DLS measurement of magnetite coated with R-AN amidoxime (Figure 8A) yields a hydrodynamic average size of 18 nm with a polydispersity index (PDI) of 0.731. The DLS data of magnetite-coated with RK-AN amidoxime (Figure 8B) indicate that the size and the polydispersity index (PDI) are 48 nm and 0.038, respectively. The data indicate that particle sizes increase with coating with RK-AN amidoxime and PDIs decreased and became monodisperse. This can be attributed to the generation of a high amount of magnetite NPs which are encapsulated in RK-AN amidoxime by the homogeneous nucleation mechanism. The distribution of magnetite NPs inside RK-AN amidoxime is an important factor, which can influence the chemical stability of particles. DLS data showed larger particle size than TEM, which may be caused by the mutual magnetic interaction of the particles to form magnetic aggregates, thus disturbing DLS. It is well known that the size distributions and surface charges of the magnetite nanoparticles are responsible for stability and dispersability of particles in solutions [37]. The higher charge via smaller size particles repulse each other and cannot combine to form aggregates. In this respect, the zeta potential of magnetite and its composites with R/AN and RK/AN amidoximes were determined as represented in Figure 9.

(A)

(B)

Figure 7. TEM micrographs of magnetite coated with (A) R-AN and (B) RK-AN amidoximes.

Int. J. Mol. Sci. 2015, 16

6921

Figure 8. DLSA data of magnetite coated with (A) R-AN and (B) RK-AN amidoximes in (ethanol/water; 60/40 volume %) 1 mM KCl at pH = 7. It was found that the zeta potential of the magnetite increased from +3.0 mV to +20.0 and +45.0 mV when magnetite coated with R/AN and RK/AN amidoximes, respectively. The positive charges of zeta potential were probably attributed to amine groups [38] of rosin amidoxime which increased for the RK/AN amidoxime. It was previously reported that a nanometer system is stable enough when the absolute value of the zeta potential is higher than 30 mV [37]. So, the magnetite coated with RK/amidoxime particles with a zeta potential of +45 mV had a better stability and good dispersability in ethanol/water (60/40) than that coated with R/AN amidoxime. Moreover, the R/AN and RK/AN amidoximes layer could provide stability for aggregation, which increased the possibility to use magnetite nanoparticles for medical and environmental applications.

Figure 9. Zeta potential of (A) magnetite; (B) magnetite coated with R/AN and (C) magnetite RK/AN coated with amidoximes in (ethanol/water; 60/40 volume %) 1 mM KCl at pH = 7. 2.3. Antibacterial Activity of Magnetite-Coated Nanoparticles Toxicity of chemicals used to control marine pollution by oil is one of the difficult challenges that face researchers for using the dispersant to control the oil spill that is toxic to marine life. Moreover, chemical dispersants have been shown to exert a synergistic effect when mixed with oil, increasing its toxicity. The nanoparticles have dimensions typically below 100 nm and have many properties that are different from those of traditionally used materials [39]. However, only the iron oxide MNPs have been approved for clinical use by the Food and Drug Administration due to their inherent properties [39]. Even though magnetite nanoparticles display great antimicrobial effect, many studies reported that these nanostructures may be highly toxic for hosts in higher concentrations or even active doses [40]. The present work will investigate the antibacterial properties of magnetite nanoparticles coated with R/AN

Int. J. Mol. Sci. 2015, 16

6922

and RK/AN amidoxime against Escherichia coli (E. coli) and Staphylococcus aureous (S. aureous) to determine the minimum inhibition concentration and percent of inhibition of colony growth. The data of antimicrobial inhibition of magnetite-coated nanoparticles were determined and listed in Table 1. The data indicated that the magnetite coated with RK/AN amidoxime is highly effective to inhibit E. coli bacteria more than that coated with R/AN amidoxime at lower concentration 2500 µg·mL−1 as listed in Table 1. Table 1. MIC (minimum inhibition concentration) values (mg·mL−1) of sample against Escherichia coli ATCC 8739, Staphylococcus aureus ATCC 6538 strains. Antimicrobial Materials Magnetite/R-AN amidoxime Magnetite/RK-AN amidoxime

E. coli S. aureus E. coli S. aureus

MIC (µg·mL−1) 10,000 5000 2500 -

The Inhibition of Colony Growth Achieved Standard Percent (%) Deviation 30 ± 4 1.85 39 ± 5 1.58 47 ± 7 3.59 -

The reason of this exceptional antimicrobial activity was the higher surface area to volume ratio of the MNPs coated with RK/AN amidoxime as determined from DLS and TEM analyses and hence the greater available surface of the antibiotic that was in contact with the microorganisms but also the control release ratio. These data indicated that the RK/AN amidoxime had a positive effect on the antimicrobial activity through encapsulation of bacteria with magnetite nanomaterials which making them available for interacting with the bacterial microorganisms. The high contents of the hydroxyl groups of magnetite-coated with RK/AN amidoxime play significant roles in the bactericidal efficacy and mechanism [40]. The nanoparticles have low size and are also able to penetrate and damage the mitochondria [41]. Moreover, it was established that there are possible mechanisms used to explain the antimicrobial activity of metal and metal oxide nanoparticles, based on their ability to attach or penetrate to the cell membrane of bacteria and disturb its function [42,43]. The transition metals nanoparticles (iron, copper etc.) generate reactive oxygen and form hydroxyl radical that attack biological molecules [44]. It was also observed that magnetite coated with R/AN inhibited the growth of S. aureus, but that magnetite-coated with RK/AN amidoxime was not able to inhibit the growth of the S. aureus strain. The bond between superoxide radical and hydroxide of Fe3O4 and NH2 or OH groups of rosin amidoxime (Scheme 1) enhanced antimicrobial activity of the resultant nanocomposite. These data demonstrated that magnetite-coated with R/AN amidoxime had very good antimicrobial activity against S. aureus. It was previously reported that, mixing of silver nitrate and titanium dioxide nanoparticles induced a 100% reduction in the viable Escherichia coli and S. aureus observed in the coated mask materials after 48 h of incubation [45]. The magnetite nanomaterials coated with oleic acid to improve their antibiofilm properties of textile dressing was tested in vitro against monospecific Candida albicans biofilms. The antimicrobial data proved that the coated materials were more resistant to C. albicans colonization as compared to the uncoated ones. The literature [46] indicated that, at the present time, magnetic iron oxide nanoparticles are applied as contrast agents for targeting organs (liver and spleen) or lymph nodes due to their antimicrobial activity. Moreover, magnetite developed to use through molecular imaging and cell tracking face a challenge with functionalization of iron oxide

Int. J. Mol. Sci. 2015, 16

6923

nanoparticle surfaces. Thus, such magnetite-coated nanoparticles with R/AN and RK/AN can be used for oil spill collector without being cytotoxic or influencing the HCT8 eukaryotic cell cycle. Our results demonstrate that the novel synthesized Fe3O4 coated with R/AN and RK/AN amidoxime nanosystems resulted in a novel safe nano-bio composite. Due to these characteristics synthesized Fe3O4 coated with R/AN and RK/AN amidoxime thin films represent a competitive candidate for the development of novel biomedical surfaces or devices with low costs and a high efficiency. 2.4. Application of Magnetite-Coated Nanoparticles as Oil Spill Collector It is very important to measure the magnetic properties of the core/shell magnetic and superparamagnetic properties for practical applications. The magnetic properties of the magnetite nanoparticles capped with R-AN and RK-AN amidoximes were also studied by a VSM. Figure 10 shows the hysteresis loops of the Fe3O4 coated with R-AN and RK-AN amidoximes at room temperature. The saturation magnetization (Ms), of bulk magnetite (92 emu·g−1) [47] was reduced to 80 emu·g−1 for magnetite nanoparticles which is more than sufficient for magnetic separation with a conventional magnet for use in the environmental applications in water purification from oil or heavy metals. The saturation magnetization of the prepared magnetite particles, which was found to be equal to 78.6 emu/g, can be compared to the theoretical value of 92 emu/g. It was found that the value 84 emu/g was the highest value found in the literature for synthetic magnetite particles [48]. The magnetization of magnetite nanoparticles were reduced from 108.1 to 75.9 emu/g by changing the crystal shapes from needle to sphere nanoparticles. In our previous work [26], the magnetization of iron oxide cubic nanoparticles prepared in the presence of iodine was 78.6 emu/g. This result indicates the formation of the well-defined crystalline structure for the as-synthesized magnetite nanoparticles. In the present work the relation between particle size and the magnetic behavior of capped Fe3O4 nanoparticles is discussed. It is known that the magnetization of a magnetic particle in an external field is proportional to its size value. It was reported that the decrease in particle size means an increasing surface-to-volume ratio, which causes more surface spin disorder, and consequently the saturation magnetization will reduce [49,50]. Therefore, the increase in the crystallites size will increase the value of magnetization. These facts are in agreement with the data of magnetic saturation magnetite nanoparticles capped with R-AN and RK-AN amidoximes which are 60.9 and 70.8 emu/g, respectively. The data indicated that the existence of the nonmagnetic shell would also lead to a decreased saturation magnetization. Therefore, the reduced magnetization may be related to the nature of the core or core surface [51]. Although the saturation magnetization decreased due to the decrease of the magnetite fraction after R-AN and RK-AN amidoximes coating, complete magnetic separation can be achieved in 15 s by placing a magnet near the vessels containing the aqueous dispersion of Fe3O4 particles. This advantage further facilitated the practical application of these magnetic nanoparticles in the petroleum crude oil collector from sea water. Different methods were used to control oil spill pollution using mechanical tools, chemical dispersants, bioremediation, physical methods and oil sorbents [12]. The present work aims to use a new method to convert petroleum crude oil to magnetic fluids followed by collecting from the water surface using a magnetic field. In this respect, the magnetite materials should have: been selectively dispersed in crude oil more than water, strong magnetic properties, low density, more hydrophobic and

Int. J. Mol. Sci. 2015, 16

6924

easily handled. The present work prepared hydrophobic magnetite with strong magnetic properties as illustrated in the previous section. The efficiency of the magnetite as oil collector will be illustrated in this section. The data of oil recovery efficiency correlated to magnetite oil ratio (MOR) and is listed in Table 2.

Figure 10. VSM magnetization curves of magnetite coated with (A) R-AN and (B) RK-AN amidoximes. Table 2. Efficiency of magnetite nanoparticles at different MOR with Arabian heavy crude oil. Magnetite Uncoated magnetite

Magnetite/R-AN amidoxime

Magnetite/RK-AN amidoxime

MOR 1:1 1:5 1:10 1:1 1:5 1:10 1:1 1:5 1:10

Efficiency Index (%EI) 45 20 10 80 75 70 95 90 80

The data indicated that magnetite coated with RK-AN amidoxime shows the good oil collection of oil without magnetite precipitation in water at MOR 1:5. This data can be correlated to the data of magnetic saturation which indicated that the magnetite coated with RK-AN amidoximes stronger than that coated with R-AN amidoxime. Moreover, although uncoated magnetite has strong magnetic saturation [26], it achieved low oil spill collector (Table 2) due to low hydrophobicity of uncoated magnetite. 3. Experimental Section 3.1. Materials Rosin acid (acid number = 184 mg of KOH·g−1, 99%) was separated and purified by crystallization from cooled and concentrated acetone solutions of commercial rosin (Khangzhou Forest Chemical Plant, Deqing, China). Acrylonitrile AN (Fluka; Buchs, Switzerland), was purified by distillation under reduced pressure. Distilled AN was stabilized with 0.1% hydroquinone. The Diels-Alder adduct of rosin acid with acrylonitrile (R-AN) was efficiently prepared and purified as described previously [29]. Hydroxylamine hydrochloride (Fluka), and p-toluene-4-sulfonic acid monohydrate (PTSA; Aldrich,

Int. J. Mol. Sci. 2015, 16

6925

St Louis, MO, USA) were used as received. Hydroxylamine hydrochloride (40.1 g) was dissolved in 290 mL of methanol solution (methanol: water 5:1). The hydroxyl amine was neutralized by sodium hydroxide solution till pH 10 and the precipitated NaCl was removed by filtration. The organic solvents were analytical-grade. The Arabian heavy crude oil produced by ARAMCO, Ras Tanora oil field has API degree 25 and asphaltene content 15.3 (wt %) was used. 3.2. Preparation Technique 3.2.1. Preparation of Diels Alder of Rosin Adduct Dehydrocaboxylation of R-AN was carried out to prepare di-rosin ketone-acrylonitrile adduct (RK-AN). In this respect, R-AN (144 g, 0.4 mol) was mixed with toluene (40 g) and heated to reflux at 120 °C for 10 min under nitrogen atmosphere. PTSA (1.44 g, 1.0 wt %, referred to the R-AN weight) was added to the reaction mixture. The ascendant condenser was replaced with a Dean-Stark trap to remove water from the reaction mixture and toluene. The reaction temperature was increased to 235 °C to release carbon dioxide and water from the reaction for 4 h until the water was completely removed. The reaction temperature was kept at 240 °C for 1 h. The unreacted R-AN and PTSA were removed by the dissolution of the crude product (131 g) into acetone (200 g), neutralization with 0.2 N alcoholic NaOH and heating at 35 °C for 30 min. The RK-AN was separated by precipitation of the solution in ice-cooled water under vigorous stirring. The precipitate was washed with hot distilled water until the pH of the water became neutral. The dry precipitate was washed with petroleum ether, drying, dissolution into diethyl ether, filtering, and finally drying. Finally, 116 g (80.5%) of pure RK-AN (a.n. = 1.2 mg of KOH·g−1) was obtained. 3.2.2. Preparation of Rosin Amidoxime The resins R-AN or RK-AN (30 g) having different mole of ratios were reacted with the prepared hydroxylamine solution at 70 °C for four hours. The prepared rosin amidoxime resin was separated by filtration and washed several time by methanol solution then treated with 0.1 M HCl solution for at least 5 min. Finally, resin was filtered and washed several times by methanol solution and dried at 50 °C to constant weight. 3.2.3. Preparation of Magnetite Coated Amidoxime Nanoparticles The procedure for preparation of magnetite nanoparticles capped with rosin amidoxime is as follows: anhydrous FeCl3 (40 g) dissolved in distilled water (300 mL) and potassium iodide (13.2 g, 0.08 mol) dissolved in distilled water (50 mL) were mixed together at room temperature, stirred and allowed to reach equilibrium for one hour while bubbling with pure N2. R-An or RK-An amidoximes (10 g) was solubilized in ethanol-water solvent (100 mL, 4:1 volume %) and added dropwise to the reaction mixture (included iodine as precipitate) at the same time as 25% ammonia solution (200 mL). The reaction mixture was then heated to 45 °C under stirring and N2 bubbling for 4 h until complete precipitation of black magnetite was achieved. The precipitate was then left to settle, filtered, washed

Int. J. Mol. Sci. 2015, 16

6926

with distilled water, ethanol, dried under vacuum at 30 °C (the precipitate was dried without heating) and weighed. The percentage yield was 95.9% and 99.5% for R-AN and RK-AN amidoxime, respectively. 3.3. Characterization FTIR spectra were analyzed with a Nicolet FTIR spectrophotometer (Madison, WI, USA) using KBr in the wavenumber range of 4000–500 cm−1 with a resolution accuracy of 4 cm−1. All samples were ground and mixed with KBr and then pressed to form pellets. 1 H NMR and 13C NMR spectra of the prepared resins were recorded on a 400 MHz Bruker Avance DRX-400 spectrometer (Santa Clara, CA, USA). The obtained samples were characterized on a BDX-3300 diffractometer (Eindhoven, The Netherlands) using CuKα radiation (wavelength, λ = 1.5406 Å) with variable slits at 45 kV/40 mA to obtain X-ray powder diffraction (XRD) patterns. Transmission electron microscopy (TEM) micrographs were taken with a JEOL JEM-2100F (JEOL, Tokyo, Japan). A few drops of magnetite nanoparticle solution were diluted into 1 mL of ethanol, and the resulting ethanol solution was placed onto a carbon coated copper grid and allowed to evaporate. Dynamic light scattering (DLS) measurements were performed on a Zetasizer Nano ZS (Malvern Instruments, Malvern, Montereal, QC, Canada) with a 633 nm He-Ne laser using ethanol/water (4/1) as solvent. The nanoparticles size was further measured by DLS at 25 °C. The DLS measurements were performed on a Brookhaven Instruments system (Santa Barbara, CA, USA) with a 514.5 nm argon ion laser (model 85 Lexel Laser) as the light source. Thermogravimetric analysis (TGA) was determined by a simultaneous DTA-TG (Shimadzu, Tokyo, Japan) DTG-60 M) and DSC apparatus (Shimadzu, DSC-60) by heating the samples from room temperature to 700 °C under N2 atmosphere at a heating rate of 10 °C·min−1. The magnetic measurement of the products was carried out in a USALDJ9600-1 vibrating sample magnetometer (VSM; LDJ Electronics Co., Troy, MI, USA). Magnetic hysteresis loops were recorded at room temperature in a field of 10,000 Oe to determine the saturation magnetization (Ms) for the samples. 3.4. Antimicrobial Properties of Samples Antimicrobial effects and minimum inhibitory concentration (MIC) and minimum bactericidal concentration (MBC) of the samples were determined using the broth-micro dilution test against three common strains of bacteria; Escherichia coli ATCC 8739, Staphylococcus aureus ATCC 6538, Bacillus subtilis ATCC 6633 and Pseudomonas aeruginosa ATCC 10145. Stock culture (0.1 mL) containing approximately 10 × 106 CFU·mL−1 of each bacterial suspension and ~2.5–100 mg of sterile samples were inoculated in 9.9 mL nutrient broth. Incubation was carried out at 35 °C for 18–24 h. For enumeration, the cultures in each media were serially diluted 10 fold using sterile FTS solution, and 100 µL of each diluted sample was placed on nutrient agar and incubated for 18–24 h at 35 °C, and finally, the colonies were counted.

Int. J. Mol. Sci. 2015, 16

6927

3.5. Evaluation of the Coated Magnetite/Rosin Amidoxime Nanoparticles as Oil Collector The water bath experiment depicted by Chun et al. [52] was modified to determine the efficiency of the prepared magnetite/rosin amidoxime nanoparticles as oil collector as follows: The crude oil was (1 mL) spread over 250 mL of sea water. The magnetite powder was carefully added to the oil surface and then a 1 cm deep vortex was created by slow stirring. The magnetite powder was added to the center of the vortex and the stirring rate was immediately increased and maintained at a maximum rate of 200 rpm over a 60 s period and then stopped. After 5 min, the magnetic particles bound with the surfactant were recovered with a permanent magnet of 25 mm diameter and 10 mm thickness, made with Nd-Fe-B (4300 Gauss). The ratios between magnetite powders and crude oil were changed from 1:1, 1:5, 1:10 to study the effect of magnetite contents on the oil recovery. The remained oil was extracted by chloroform. The chloroform extraction was repeated twice more. The absorbance of chloroform extract was made against a chloroform blank at 580 nm in glass of 5 mm path length. The calibration graph was then used to calculate the weight of oil contained in the 50 cm3 oily water sample. The Efficiency Index, EI, is calculated from the following equation: %EI = Weight of oil in 50 cm3 sample of oily water × 500

(1)

4. Conclusions Fe3O4 nanoparticles coated with rosin amidoxime were synthesized via a simple and inexpensive co-precipitation reaction at a temperature of 45 °C. This methodology should lead to a variety of magnetite capped nanoparticles that have different dispersabilities in organic solvents and stabilities which can be suitable for further investigation in environmental applications. Moreover, properties of magnetite were accomplished using Fourier transformed infrared (FTIR), X-ray powder diffraction (XRD), transmission electron microscopy (TEM) and vibrating sample magnetometer (VSM) analyses. The shape-controlled magnetite capped rosin amidoxime nanoparticles have been successfully prepared and the morphology changed from nanocubes to nanosphere by capping magnetite with rosin amidoximes. The prepared magnetite capped nanoparticles displayed the perfect regular nanostructure with monodispersion and a narrow distribution in particle size. This work reports the successful preparation of bioactive thin films based on Fe3O4 coated with R/AN and RK/AN amidoxime magnetic nanospheres with diameters between 20 and 40 nm. These nanocoatings displayed antimicrobial activity properties, inhibiting the growth of S. aureus and P. aeruginosa. The collection of petroleum crude oil using a magnetic field was successfully accomplished using magnetic rosin amidoxime. In addition, the magnetic amidoxime can be easily separated by the external magnetic field. Accordingly, the problem of phase separation using the traditional oil sorbents can be resolved. Acknowledgments This project was supported by King Saud University, Deanship of Scientific Research, Research group RGP-235.

Int. J. Mol. Sci. 2015, 16

6928

Author Contributions Ayman M. Atta suggested the idea and contributed the discussion, Hamad A. Al-Lohedan contributed in discussion and Sami A. Al-Hussain performed the experimental work. Conflicts of Interest The authors declare no conflict of interest. References 1.

Thanh, N.T.K.; Greena, L.A.W. Functionalisation of nanoparticles for biomedical applications. Nano Today 2010, 5, 213–230. 2. Madrakian, T.; Afkhami, A.; Zolfigol, M.A.; Ahmadi, M.; Koukabi, N. Application of modified silica coated magnetite nanoparticles for removal of iodine from water samples. Nano Micro Lett. 2012, 4, 57–63. 3. Zhao, X.; Shi, Y.; Wang, T.; Cai, Y.; Jiang, G. Preparation of silica-magnetite nanoparticle mixed hemimicelles sorbents for extraction of several typical phenolic compounds from environmental water samples. J. Chrom. A 2008, 1188, 140–147. 4. Zhao, Y.; Li, J.; Zhao, L.; Zhang, S.; Huang, Y.; Wu, X.; Wang, X. Synthesis of amidoxime-functionalized Fe3O4@SiO2 core-shell magnetic microspheres for highly efficient sorption of U(VI). Chem. Eng. J. 2014, 235, 275–283. 5. Akin, I.; Arslan, G.; Tor, A.; Ersoz, M.; Cengeloglu, Y. Arsenic(V) removal from underground water by magnetic nanoparticles synthesized from waste red Mud. J. Hazard. Mater. 2012, 235, 62–68. 6. Chandra, V.; Park, J.; Chun, Y.; Lee, J.W.; Hwang, I.C.; Kim, K.S. Water-dispersible magnetite-reduced graphene oxide composites for arsenic removal. ACS Nano 2010, 4, 3979–3986. 7. Koo, H.Y.; Lee, H.J.; Go, H.A.; Lee, Y.B.; Bae, T.S.; Kim, J.K.; Choi, W.S. Graphene-based multifunctional iron oxide nanosheets with tunable properties. Chem. Eur. J. 2011, 17, 1214–1219. 8. Potapova, E.; Grahn, M.; Holmgren, A.; Hedlund, J. The effect of polymer adsorption on the wetting properties of partially hydrophobized magnetite. J. Colloid Interface Sci. 2012, 367, 478–484. 9. Akl, M.A.; Atta, A.M.; Yousef, A.-F.M.; Alaa, M.I. Characterization of stabilized porous magnetite core-shell nanogel composites based on crosslinked acrylamide/sodium acrylate copolymers. Polym. Int. 2013, 62, 1667–1677. 10. Bronstein, L.M.; Shtykova, E.V.; Malyutin, A.; Dyke, J.C.; Gunn, E.; Gao, X.; Stein, B.; Konarev, P.V.; Dragnea, B.; Svergun, D.I. Hydrophilization of magnetic nanoparticles with modified alternating copolymers. Part 1: The influence of the grafting. J. Phys. Chem. B 2009, 113, 647–655. 11. Tan, W.L.; Abu Bakar, M. The Effect of additives on the size of Fe3O4 particles. J. Phys. Sci. 2006, 17, 37–50. 12. Burke, N.A.D.; H. Stover, H.D.; Dawson, F.P. Magnetic nanocomposites: Preparation and characterization of polymer-coated iron nanoparticles. Chem. Mater. 2002, 14, 4752–4761.

Int. J. Mol. Sci. 2015, 16

6929

13. Shang, H.; Chang, W.-S.; Kan, S.; Majetich, S.A.; Lee, G.U. Synthesis and characterization of paramagnetic microparticles through emulsion-templated free radical polymerization. Langmuir 2006, 22, 2516–2522. 14. Daou, T.J.; Grenèche, J.M.; Pourroy, G.; Buathong, S.; Derory, A.; Ulhaq-Bouillet, C.; Donnio, B.; Guillon, D.; Begin-Colin, S. Coupling agent effect on magnetic properties of functionalized magnetite-based nanoparticles. Chem. Mater. 2008, 20, 5869–5875. 15. Barth, S.; Estrade, S.; Hernandez-Ramirez, F.; Peiro, F.; Arbiol, J.; Romano-Rodriguez, A.; Morante, J.R.; Mathur, S. Studies on surface facets and chemical composition of vapor grown one-dimensional magnetite nanostructures. Cryst. Growth Des. 2009, 9, 1077–1081. 16. Miles, W.C.; Huffstetler, P.P.; Goff, J.D.; Chen, A.Y.; Riffle, J.S.; Davis, R.M. Design of stable polyether_magnetite complexes in aqueous media: Effects of the anchor group, molecular weight, and chain density. Langmuir 2011, 27, 5456–5463. 17. Astete, C.E.; Kumar, C.S.S.R.; Sabliov, C.M. Size control of poly(D,L-lactide-co-glycolide) and poly(D,L-lactide-co-glycolide)-magnetite nanoparticles synthesized by emulsion evaporation technique. Colloid Surf. A 2007, 299, 209–216. 18. Yathindranath, V.; Rebbouh, L.; Moore, D.F.; Miller, D.W.; Lierop, J.; Hegmann, T. A versatile method for the reductive, one-pot synthesis of bare, hydrophilic and hydrophobic magnetite nanoparticles. Adv. Funct. Mater. 2011, 21, 1457–1464. 19. Yang, D.; Hu, J.; Fu, S. Controlled synthesis of magnetite-silica nanocomposites via a seeded sol-gel approach. J. Phys. Chem. C 2009, 113, 7646–7651. 20. Zhao, G.X.; Wen, T.; Yang, X.; Yang, S.B.; Liao, J.L.; Hu, J.; Shao, D.D.; Wang, X.K. Preconcentration of U(VI) ions on few-layered graphene oxide nanosheets from aqueous solutions. Dalton Trans. 2012, 41, 6182–6188. 21. Chen, H.M.; Deng, C.H.; Zhang, X.M. Synthesis of Fe3O4@SiO2@PMMA core-shell-shell magnetic microspheres for highly efficient enrichment of peptides and proteins for MALDI-ToF MS analysis. Angew. Chem. Int. Ed. 2010, 49, 607–611. 22. Park, M.; Seo, S.; Lee, I.S.; Jung, J.H. Ultraefficient separation and sensing of mercury and methylmercury ions in drinking water by using aminonaphthalimide-functionalized Fe3O4@SiO2 core/shell magnetic nanoparticles. Chem. Commun. 2010, 46, 4478–4480. 23. Sun, L.; Li, Y.X.; Sun, M.D.; Wang, H.G.; Xu, S.F.; Zhang, C.Q.; Yang, Q.B. Porphyrin functionalized Fe3O4@SiO2 core/shell magnetic colorimetric material for detection, adsorption and removal of Hg2+ in aqueous solution. New J. Chem. 2011, 35, 2697–2704. 24. Rezaei, A.; Khani, H.; Masteri-Farahani, M.; Rofouei, M.K. A novel extraction and preconcentration of ultra-trace levels of uranium ions in natural water samples using functionalized magnetic-nanoparticles prior to their determination by inductively coupled plasma-optical emission spectrometry. Anal. Methods 2012, 4, 4107–4114. 25. Sadeghi, S.; Azhdari, H.; Arabi, H.; Moghaddam, A.Z. Surface modified magnetic Fe3O4 nanoparticles as a selective sorbent for solid phase extraction of uranyl ions from water samples. J. Hazard. Mater. 2012, 215, 208–216. 26. Atta, A.M.; al-Lohedan, H.A.; al-Hussain, S.A. Synthesis of stabilized myrrh-capped hydrocolloidal magnetite nanoparticles. Molecules 2014, 19, 11263–11278.

Int. J. Mol. Sci. 2015, 16

6930

27. Zong, P.F.; Wang, S.F.; Zhao, Y.L.; Wang, H.; Pan, H.; He, C.H. Synthesis and application of magnetic graphene/iron oxides composite for the removal of U(VI) from aqueous solutions. Chem. Eng. J. 2013, 220, 45–52. 28. Arranz, F.; Sanchez-Chaves, M.; Gallego, M.M. Synthesis of poly(viny1 alcohol) derivatives containing amidoxime groups. Die Angew. Makromol. Chem. 1994, 218, 183–196. 29. Bicu, I.; Mustata, F. Synthesis and characterization of highly dispersed acid–acrylonitrile Diels–Alder adduct. J. Polym. Sci. 2005, 43, 6308–6322. 30. Atta, A.M.; Dyab, A.K.F. Coated Magnetite Nanoparticles, Method for the Preparation Thereof and Their Use. Eur. Pat. 13167616.5, 18 March 2013. 31. Khalil, M.I. Process for Preparing Magnetic (Fe3O4) and Derivatives Thereof. Eur. Pat. 2505558 B1, 16 January 2013. 32. El Mahdy, G.A.; Atta, A.M.; Dyab, A.K.F.; Al-Lohedan, H.A. Protection of petroleum pipeline carbon steel alloys with new modified core-shell magnetite nanogel against corrosion in acidic medium. J. Chem. 2013, 2013, doi:10.1155/2013/125731. 33. Shen, L.; Qiao, Y.; Guon, Y.; Meng, S.; Yang, G.; Wu, M.; Zhao, J. Facile co-precipitation synthesis of shape-controlled magnetite nanoparticles. Ceram. Int. 2014, 40, 1519–1524. 34. Ni, Y.H.; Ge, X.W.; Zhang, Z.C.; Ye, Q. In situ single-step synthesis of gold/polyacrylamide nanocomposites in an ethanol system. Mater. Lett. 2002, 55, 171–174. 35. Scherrer, P. Bestimmung der Grösse und Derinneren Struktur von Kolloidteilchen Mittels Röntgensrahlen. Available online: http://www.sajs.co.za/sites/default/files/publications/pdf/ Kroon_Sci%20Corr.pdf (accessed on 21 July 2014). 36. Lim, J.; Yeap, S.P.; Che, H.X.; Low, S.C. Characterization of magnetic nanoparticle by dynamic light scattering. Nanoscale Res. Lett. 2013, 8, 381–388. 37. Sun, H.; Yu, J.; Gong, P.; Xu, D.; Zhang, C. Novel core–shell magnetic nanogels synthesized in an emulsion-free aqueous system under UV irradiation for targeted radiopharmaceutical applications. J. Magn. Magn. Mater. 2005, 294, 273–280. 38. Ho, K.M.; Li, P. Design and synthesis of novel magnetic core-shell polymeric particles. Langmuir 2008, 24, 1801–1807. 39. Wilczewska, A.Z.; Niemirowicz, K.; Markiewicz, K.H.; Car, H. Nanoparticles as drug delivery systems. Pharmacol. Rep. 2012, 64, 1020–1037. 40. Iconaru, S.L.; Prodan, A.M.; Motelica-Heino, M.; Sizaret, S.; Predoi, D. Synthesis and characterization of polysaccharide-maghemite composite nanoparticles and their antibacterial properties. Nanoscale Res. Lett. 2012, 7, 576–582. 41. Li, N.; Sioutas, C.; Cho, A.; Schmitz, D.; Misra, C.; Sempf, J.; Wang, M.; Oberley, T.; Froines, J.; Nel, A. Ultrafine particulate pollutants induce oxidative stress and mitochondrial damage. Environ. Health Perspect. 2003, 111, 455–460. 42. Morones, J.R.; Elechiguerra, J.L.; Camacho, A.; Holt, K.; Kouri, J.B.; Tapia, J.; Yacaman, M.J. The bactericidal effect of silver nanoparticles. Nanotechnology 2005, 16, 2346–2353. 43. Buzea, C.; Pacheco Blandino, I.; Robbie, K. Nanomaterials and nanoparticles: Sources and toxicity. Biointerphases 2007, 2, MR17–MR71. 44. Risom, L.; Moller, P.; Loft, S. Oxidative stress-induced DNA damage by particulate air pollution. Mutat. Res. 2005, 592, 119–137.

Int. J. Mol. Sci. 2015, 16

6931

45. Li, Y.; Leung, P.; Yao, L.; Song, Q.W.; Newton, E. Antimicrobial effect of surgical masks coated with nanoparticles. J. Hosp. Infect. 2006, 62, 58–63. 46. Laurent, S.; Forge, D.; Port, M.; Roch, A.; Robic, C.; Elst, L.V.; Muller, R.N. Magnetic iron oxide nanoparticles: Synthesis, stabilization, vectorization, physicochemical characterizations, and biological applications. Chem. Rev. 2008, 108, 2064–2110. 47. Zaitsev, V.S.; Filimonov, D.S.; Presnyakov, I.A.; Gambino, R.J.; Chu, B. Physical and chemical properties of magnetite and magnetite-polymer nanoparticles and their colloidal dispersions. J. Colloid Interface Sci. 1999, 212, 49–57. 48. Dresco, P.A.; Zaitsev, V.S.; Gambino, R.J.; Chu, B. Preparation and properties of magnetite and polymer magnetite nanoparticles. Langmuir 1999, 15, 1945–1951. 49. Rajendran, M.; Pullar, R.C.; Bhattacharya, A.K.; Das, D.; Chintalapudi, S.N.; Majumdar, C.K. Magnetic properties of nanocrystalline CoFe2O4 powders prepared at room temperature: Variation with crystalline size. J. Magn. Magn. Mater. 2001, 232, 71–83. 50. Roy, S.; Dubenko, I.; Edorth, D.D.; Ali, N. Size induced variations in structural and magnetic properties of double exchange Ls0.8Sr0.2MnO3_δ nano-ferromagnet. J. Appl. Phys. 2004, 96, 1202–1208. 51. Fingas, M. Oil Spill Science and Technology Prevention, Response, and Cleanup; Gulf Professional Publishing; Elsevier 30 Corporate Drive: Burlington, MA, USA, 2011; pp. 26–88. 52. Chun, C.-L.; Park, J.-W. Oil spill remediation using magnetic separation. J. Environ. Eng. 2001, 127, 443–449. © 2015 by the authors; licensee MDPI, Basel, Switzerland. This article is an open access article distributed under the terms and conditions of the Creative Commons Attribution license (http://creativecommons.org/licenses/by/4.0/).