Photocatalytic degradation of methylene blue with hematite nanoparticles synthesized by thermal decomposition of fluoroquinolones oxalato–iron(III) complexes

Ahmed M. Mansour*
Chemistry Department, Faculty of Science, Cairo University, Gamaa Street, Giza 12613, Egypt. E-mail: inorganic_am@yahoo.com; mansour@sci.cu.edu.eg; Fax: +20 2 35728843; Tel: +20 2 01222211253

Received 24th June 2015 , Accepted 10th July 2015

First published on 10th July 2015


Abstract

[Fe(C2O4)(FQ)(H2O)2] complexes (H-FQ = ciprofloxacin (1), lomefloxacin (2) and norfloxacin (3)) were synthesized and characterized using a variety of analytical and spectral techniques such as elemental analysis, infrared spectroscopy, thermogravimetric analysis, ultraviolet-visible spectroscopy, magnetic and conductance measurements. The experimental studies were complemented by quantum chemical calculations in terms of geometry optimization, natural bond orbital analysis and molecular electrostatic potential maps. Electronic structures were discussed by TD-DFT. Hematite (α-Fe2O3) nanoparticles, as promising materials for different catalytic applications, were prepared in air via the controlled thermal decomposition of 1–3. Powder X-ray diffraction was used to identify the polymorph of iron oxide. The morphology of nano-hematite was investigated by a field emission scanning electron microscope coupled to energy-dispersive X-ray spectroscopy for surface analysis. The catalytic degradation of methylene-blue (MB), as an industrial pollutant, exposed to UV radiation in the presence of nano-α-Fe2O3 as a catalyst and hydrogen peroxide as the oxidant was studied at room temperature in water.


Introduction

So far, the oxalato ligand has appeared as a fertile key for design complexes of intriguing structural research applications.1 Normally, the oxalato can act as a chelating or bridging ligand.2,3 The [MIII(C2O4)3]3− complexes of Al and transition metal ions have been extensively studied.4–16 In 1939,5 Blair and Jones synthesized potassium tris(oxalato)ferrate(III) trihydrate (KOF) from FeC2O4. Recently, a modified synthesis method has been introduced by Hussain et al.6 through the reaction of a mixture of K2C2O4·H2O and H2C2O4·2H2O with Fe(OH)3. Crystal structures of anhydrous6 and hydrated forms7 of KOF showed the existence of infinite anionic distorted octahedral [Fe(C2O4)3]3− units and the three K+ ions are coordinated by water or oxalato ligands. The importance of KOF is ascribed to being a chemical actinometer. This complex undergoes a color change when it is heated8 or if the aqueous solution is exposed to UV radiation9 suggesting the reduction of Fe3+ and formation of an ion-radical intermediate, C2O4˙.10,11 The photo-chemical significance of KOF has been extended to the field of polymer chemistry e.g. the photo-polymerization of acrylamide12 and methyl methacrylate13 in the aqueous acid solution of KOF has been achieved. Another application of KOF has been gained from the preparation of nano-Fe2O3 via the solid state thermolysis process.14 Different polymorphs nanoparticles were obtained at the different temperatures.8,15,16

Fluoroquinolones (FQ's) represent a group of synthetic antibiotics distinguished by a good bioavailability, safety profile, long half-life (with once-daily dosing) and a broad antibacterial spectrum against Gram(−) bacteria.17,18 Ciprofloxacin (H-Cf) (Scheme 1) is a second-generation quinolone drug used for the treatment of urinary zone infections, bone and joint infections, infectious diarrhea, typhoid fever and acute sinusitis.19 Norfloxacin (H-Nf) is the first FQ drug introduced to the market. Its antibacterial activity involves multi-antibiotic resistant, Gram(−) rods and beta-lactamase producing organisms. Lomefloxacin (H-Lf) is a third-generation quinolone used in cure of respiratory and urinary tracts, gynecological, ophthalmological and soft tissue infections. It has the advantage of being effective against some anaerobic bacteria.20


image file: c5ra12157d-s1.tif
Scheme 1 (a) Structures of the investigated fluoroquinolones (FQ's) and (b) synthesis of [Fe(C2O4)(FQ)(H2O)2] complexes.

The formation of FQ's complexes with transition and non-transition metal ions was the question of a number of literature reports.21 Several studies22–31 concerning the interactions between H-Cf, H-Lf and H-Nf and Fe3+ ions were reported. A simple rapid spectral method was developed for determination of H-Cf in its pharmaceutical formulations using Fe(NO3)3.22 Formation constants of Fe3+ complexes of H-Cf and H-Nf were determined.23 The water-soluble Fe(Nf)2+ (λmax = 377 nm) formed in the acidic medium was used for determination of Fe3+.24 The complexation of H-Lf with Fe3+ ions in terms of stability constants and stoichiometries was studied by Riley et al.25 Complexes of the type [Fe(Nf)2(H2O)2]Cl3·6H2O,26 [Fe(Nf)3]Cl3·12H2O,27 [Fe(Cf)(nitrilotriacetate)]·3.5H2O,28 [Fe(Cf)3]29 and [Fe(Lf)(H2O)4]Cl3·H2O30 were proposed, where FQ's interact with Fe3+ via the carboxylate and pyridone oxygen atoms. Hydrothermal reaction of H-Cf with Fe(OH)3 in presence of oxalate yielded [Fe(H-Cf)(C2O4)2]·H2Cf·5H2O.31

In the present study, synthesis, spectroscopic and thermal characterization of three mononuclear iron(III)–oxalato complexes containing fluoroquinolones (ciprofloxacin, lomefloxacin and norfloxacin) as a primary ligand (Scheme 1) are reported both experimentally and theoretically. Due to the wide variety of applications for iron oxide nanoparticle polymorphs in the fields of catalysis,32 sensors,33 clinical diagnosis,34 and magnetic storage,35 a simple and nontoxic method for preparation of α-Fe2O3 nanoparticles based upon the thermal decomposition of [Fe(C2O4)(FQ)(H2O)2] complexes has been explored. The morphology has been characterized by different physical methods such as XRD, and FE-SEM. The surface analysis has been carried out by EDX. Finally, the degradation of MB by Fenton-like catalytic performance of α-Fe2O3 nanoparticles has been investigated with the aid of the ultraviolet radiation.

Results and discussion

Spectral and magnetic characterization

The reaction of KOF with one equivalent of FQ's (Scheme 1) affords complexes of the type [Fe(C2O4)(FQ)(H2O)2] (H-FQ = H-Cf (1), H-Lf (2) and H-Nf (3)). These complexes were characterized by elemental analysis, TGA, IR, UV-Vis, magnetic and molar conductance measurements. The active IR bands at 1707, 1624 cm−1 (H-Cf·HCl), 1723, 1620 cm−1 (H-Lf·HCl) and 1708, 1620 cm−1 (H-Nf) (Fig. S1) are assigned to ν(COOH)carb and ν(C[double bond, length as m-dash]O)py (carb: carboxyl, py: pyridone). In complexes, the vanishing of ν(COOH)carb mode might be a sign for the ionization of –COOH and consequently its involvement in chelation.36 In general, the carboxylate ion gives rise to two bands: a strong asymmetric stretching band near 1650–1550 cm−1 and a weaker symmetrical stretching one near 1400 cm−1.37 For example, complex 2 shows the νass(C[double bond, length as m-dash]O) and νss(C[double bond, length as m-dash]O) modes at 1554 and 1360 cm−138 with Δν ≈ 200 cm−1 characteristic of unidentate COO group,39 while complex 3 displays only the asymmetric vibration at 1548 cm−1. These modes are overlapped in 1. The pyridone C[double bond, length as m-dash]O group is liberated from the intra-molecular H-bond interaction via the complex formation and interacts with metal ion giving a band at the higher wave numbers; 1628 (1), 1630 (2) and 1627 (3).40 In contrast, the piperazine NH is not involved in chelation, since its stretching mode is located at higher wave numbers,41 3436–3442 cm−1 (1–3) compared with 3429–3436 cm−1 observed in the free ligands. The weak/medium intensity bands at 1714, 1668 and 1476 cm−1 (1), 1714, 1670 and 1459 cm−1 (2), 1713, 1665 and 1477 cm−1 (3) (Fig. S2) allocated for ν(C[double bond, length as m-dash]O)ox, νass(C[double bond, length as m-dash]O)ox and νss(C[double bond, length as m-dash]O)ox (ox: oxalato) are taken as evidence for the bidentate nature of the oxalato ligand in the mononuclear complexes.3,42

The electronic absorption spectra (Fig. S3) of 1–3 in DMSO exhibit the intra-ligand transitions at 335, 320 and 285 nm (1), 340, 325 and 292 nm (2), 335, 320 and 285 nm (3).38 An additional weak band at about 625 nm assigned to 6A1g5T1g transition in a distorted octahedral stereochemistry is observed in the studied complexes.29 The effective magnetic moment values at 298 K, calculated by the spin-only terms, are 4.86, 6.19 and 6.14μB for complexes 1–3 that are in agreement with those expected for high spin arrangement Fe(III) complexes (S = 5/2; μeff = 5.0–6.1μB).43

DFT/TD-DFT

To obtain an insight into the geometrical and electronic structures of the investigated complexes, cis-[Fe(C2O4)(FQ)(H2O)2] were optimized (Fig. 1) by DFT/B3LYP method combined with 6-31G(d). It was found that the cis-isomer is more stable than the trans-isomer by 18.543 kJ mol−1. The complexes were characterized as local minima through harmonic frequency analysis. The Fe3+ ion displays a distorted FeO6 octahedral geometry. Four oxygen atoms coming from the chelating oxalato [FeO2 = 1.840 Å and FeO7 ≈ 1.882 Å] and bidentate FQ drug [FeO3 ≈ 1.893 Å and FeO4 ≈ 1.877 Å] are coordinated to the metal ion. The octahedral geometry is completed by two water molecules [FeO5 = 2.033 Å and FeO6 = 2.026 Å] lying in a cis-position. The Fe3+–oxalato bonds are not equal, since the oxalato oxygen atoms are participated in two H-bonds of different strengths [O2⋯H54–O6 (2.403 Å, <95.5°) and O7⋯H51–O5 (2.094 Å, <107.3°)]. As shown in Fig. 1, another two H-bonds affecting the regularity of the geometry [O4⋯H53–O6 (1.989 Å, <110.0°) and O6⋯H52–O5 (2.543 Å, <95.5°)] are also observed. The strongest H-bond in the complexes is O4⋯H53–O6, since the angle of this interaction is 110.0°.37 As shown in Table 1, the optimized bond lengths and angles around the iron(III) ion are unaffected by the change in the type of substituent attached to the studied FQ's drugs.
image file: c5ra12157d-f1.tif
Fig. 1 Local minimum structures of complexes (a) 1, (b) 2 and (c) 3 obtained at DFT/B3LYP/6-31G(d) level of theory.
Table 1 Selected bond lengths (Å) and angles (°) for complexes 1–3 calculated by DFT/B3LYP/6-31G(d) method
Bond lengths (Å) Angles (°)
  1 2 3   1 2 3
FeO2 1.840 1.840 1.840 O2FeO3 93.6 93.4 93.6
FeO3 1.893 1.895 1.893 O2FeO4 97.1 97.1 97.1
FeO4 1.877 1.876 1.877 O2FeO5 165.9 166.1 165.9
FeO5 2.033 2.033 2.033 O2FeO6 87.5 87.7 87.5
FeO6 2.026 2.024 2.026 O2FeO7 85.8 85.8 85.8
FeO7 1.882 1.881 1.882 O3FeO4 96.1 96.1 96.1
        O3FeO6 175.7 175.8 175.7
        O4FeO7 174.5 174.5 174.5
        O4FeO6 79.6 79.7 79.6
        O6FeO7 95.9 95.8 95.9


The nature of the electronic transitions observed in the UV-Vis spectra of the complexes has been studied by time-dependent DFT.44–46 The lowest 30 singlet-to-singlet spin-allowed excitation states were calculated using the same functional and basis set for the optimization. The effect of solvent (DMSO) was performed using the default polarizable continuum model. The calculated d–d excitation wave lengths, energies of other excitation transitions (f > 0.002) and their assignments are taken into consideration. The stimulated spectrum of complex 1 (Fig. S4) shows three transitions at 445, 404, and 374 nm assigned to H(β) → L/L+1(β), H(β) → L(β) and H-4/H-5(β) → L(β) (H: HOMO; L: LUMO). The transitions at 445 nm have a ground state composed of Fe d character, whereas the excitation states are of Fe-dz2 (LUMO) and Fe-dx2y2/(FQ and oxalate)-π*orbitals (LUMO+1) forming d → dz2 and d → dx2y2/LMCT transitions (Fig. 2) characterized to the octahedral Fe3+ complexes. The band at 374 nm has a LMCT character, originating from the piperazine moiety (HOMO-4) and (FQ and oxalate)-π orbitals (HOMO-5) and going to Fe dz2 orbital. The TD-DFT spectrum of 2 is characterized by five transitions at 465, 420, 375, 360 and 325 nm with oscillator strengths of 0.0003, 0.0153, 0.0848, 0.0437 and 0.0101 corresponding to H(β) → L/L+1(β), H(β) → L(β), H(β) → L+1(β), H-4/H-2(β) → L(β) and H-1(β) → L+1(β). The descriptions of frontier MO's and the relocation of the electron density of complexes are nearly the same as those described in complex 1. This is in agreement with the experimental findings, since the three complexes show similar electronic spectrum (Fig. S4). The calculated spectrum of 3 exhibits the same transitions as observed in 1. Comparison between the calculated and experimental spectra indicated the suitability of the applied calculation method for this size of compounds, where the experimental bands are deviated from the calculated ones by about 10 nm.


image file: c5ra12157d-f2.tif
Fig. 2 Theoretical electronic absorption transitions of complex 1 in DMSO.

Natural bond orbital analysis and molecular electrostatic potential

Natural bond orbital (NBO) analysis47,48 and second order perturbation theory analysis of Fock matrix provide details about the type of hybridization, nature of bonding and strength of the interactions between Fe3+ ion and donor sites.49 The electronic arrangement of Fe in complexes 1–3 is [Ar]4s0.283d6.074p0.435s0.015p0.01 with 6.780 valence electrons. The natural charge is 1.200 and the occupancies of Fe 3d are: dxy1.900 dxz1.608 dyz1.352 dx2y20.563 dz20.644 (1), dxy1.925 dxz1.539 dyz1.338 dx2y20.558 dz20.707 (2) and dxy1.919 dxz1.609 dyz1.356 dx2y20.539 dz20.644 (3). The 3d-electronic population is corresponding to the oxidation state Fe(I), not Fe(III) that is in agreement with ligand to dFe electron transfer. The strength of interactions between Fe3+ ion and donor sites can be estimated by the second order perturbation theory. The larger the E2 value, the more intensive is the interaction between the electron donors and electron acceptors. For complex 1, the E2 values are 1.07, 0.72, 1.27, 1.33, 0.89 and 1.32 kcal mol−1 for LP(3)O2 → RY*(3)Fe, LP(2)O3 → RY*(4)Fe, LP(3)O4 → RY*(2)Fe, LP(2)O5 → RY*(3)Fe, LP(2)O6 → RY*(4)Fe and LP(3)O7 → RY*(2)Fe respectively. Similar values are observed for the other investigated complexes as tabulated in Table S1.

Molecular electrostatic potential (MEP) map is a very useful descriptor in understanding the sites for electrophilic and nucleophilic reactions as well as H-bonding interactions.44,46 The MEP plots (Fig. 3) of 1–3 are characterized by a positive region (blue) coming from H-atoms of FQ's and the coordinated water molecules. The negative charge region is mainly located on the oxalato O atoms and the minor contribution is coming from the carboxylate group. A zero potential area (green) is covered the remaining parts of the investigated complexes.


image file: c5ra12157d-f3.tif
Fig. 3 Molecular electrostatic potential map for complexes (a) 1, (b) 2 and (c) 3. The electron density isosurface is 0.004 a.u. Different values of the electrostatic potential are represented by different colors: red represents the regions of the most electro negative potential, blue represents regions of most positive electrostatic potential and green represents regions of zero potential. Potential increases in the following order: red < orange < yellow < green < blue.

Thermogravimetric analysis

The thermogram of 1 exhibits four decomposition steps at 90, 233, 325 and 477 °C. In the 1st step, only one coordinated water molecule is eliminated (mass loss: found, 3.73%; calcd, 3.53%). The 2nd stage is accompanied by a mass loss amounts to 20.74% (up to 310 °C), which finds a parallelism with the calculated value (20.78%) responsible for loss of another water molecule and 2CO2. The 3rd and 4th steps bring the total mass loss up to 90.57%. Such mass loss is close to the calculated value (89.02%) expected for the formation of metallic iron as a residue. Similar, four degradation stages are observed in the TG curve of 2 at 143, 214, 331 and 405 °C. The 1st step reveals loss of 1.5H2O molecules (mass loss, found, 4.73%; calcd, 5.09%). The 2nd step is assigned to removal of 0.5H2O + 2CO2 with a mass loss amounts to 18.25% (calcd 18.30%). The 3rd and 4th stages are attributed to degradation of one Lf molecule giving a metallic iron as a final residue (found, 10.33%; calcd 10.56%). The thermogram of 3 displays five decomposition steps at 102, 235, 315, 426 and 454 °C. The 1st stage (up to 200 °C) is responsible for loss of 2H2O (mass loss, found, 7.18%; calcd 7.22%). The 2nd thermal event is accompanied by a mass loss amounts to 21.54%. This value goes parallel to the calculated one (21.48%) that is attributed to the elimination of 2CO2 + 1/2F2 molecules. The 3rd and 4th stages are corresponding to the pyrolysis of the rest of the organic part with overall mass loss amounts to 88.83% (calcd 88.75%) leaving metallic iron as a residue.

α-Fe2O3 nanoparticles were prepared by heating complexes 1–3 in an electric furnace (800 °C) under the ambient air condition before cooling to room temperature. Their XRD patterns (Fig. S5) indicated that samples 1 and 2 are composed of single rhombohedral phase characteristic of α-Fe2O3 nanoparticles. All the diffraction peaks are in agreement with the reported JCPDS card no. 04-011-9585. Sample 3 is mainly α-Fe2O3 phase contaminated with α-FeOOH (goethite) (JCPDS card no. 00-034-1266). Fig. 4 shows FE-SEM images of the investigated α-Fe2O3 nanoparticles. Other FE-SEM images with a diameter of 2 μm are given in Fig. S6. As expected, the pristine nanoparticles of iron oxides bend to aggregate into large clusters because of anisotropic dipolar attraction in nanosized form of Fe2O3. In other words, the starting of sintering led to many grown agglomerates. The influence of the type of the coordinated FQ's has been reflected in size of the nanoparticles. The diameter of the nanoparticles is found to be 50–120 nm (1), 28–54 nm (2) and 47–82 nm (3). Therefore, the thermal decomposition of lomefloxacin complex 3 gave rises to smallest iron oxide nanoparticles. EDX analysis of these particles indicates the presence of Fe and O composition in the pure as-grown iron oxide as no other peak related with any impurity has been detected (Fig. 5). The sizes of the current nanoparticles are comparable to the results of several previous reports prepared iron oxide nanoparticles from the thermal decomposition conditions (13–180 nm).8,50


image file: c5ra12157d-f4.tif
Fig. 4 FE-SEM images of α-Fe2O3 nanoparticles obtained from thermal decomposition of (a) 1, (b) 2 and (c) 3.

image file: c5ra12157d-f5.tif
Fig. 5 EDX analysis of α-Fe2O3 nanoparticles obtained from thermal decomposition of (a) 1, (b) 2 and (c) 3.

Photo-degradation process

Methylene-blue (MB) is one of the generally industrial used dyes in the fields of printing, textiles, etc.51 Its industrial effluents are the major source of harmful to the living organisms through the decrease of the dissolved oxygen capacity of water and by blocking sunlight. Severe exposure to MB led to permanent burn to the eyes of organisms, nausea and vomiting.52 Therefore, the treatment of effluents containing MB is one of the challenging problems in the field of environmental chemistry. Moreover, the choice of MB in this work, as a test pollutant, to explore the catalytic performance of the synthesized nanoparticles is attributed to the high solubility in water with no possibility to volatility and gas-phase pyrolysis. Its blue solution is characterized by unique band at 665 nm, which allows easy spectrophotometrically monitoring of the degradation process.53,54 Among the different techniques used to eliminate MB, the modified Fenton reaction with various types of the synthesized catalysts55 was found to be effective through the creation of highly reactive free radicals HO˙ and safe by the formation of harmless chemicals such as CO2 and H2O. The percent of degradation depends upon the formation rate of these radicals. Heterogeneous Fenton catalyst has an advantage of working over a wider pH-range compared with the traditional one, which is limited to pH < 4. However, its disadvantage is the slower degradation rate of organic pollutants compared to the classical method. Recently, the coupling of UV irradiation with Fenton-like reagent has attracted the consideration of some working groups interested in the improvement of the degradation rate of organic pollutants such as phenol56 and methylene-blue.57

To evaluate the catalytic performance of the synthesized α-Fe2O3 nanoparticles, the absorption spectra of the illuminated samples at various time intervals were monitored. The degradation process of MB was initiated by addition of H2O2 solution to the MB solution and immediately turning on the ultraviolet irradiation. For comparison, the MB degradation experiment was also done without α-Fe2O3 nanoparticles, but in presence of H2O2 and/or UV light (Fig. S7). Fig. 6 shows the time-dependent absorption electronic spectra of MB as a function of the illumination time at 365 nm in presence of H2O2/α-Fe2O3 nanoparticles (sample no. 2). Similar spectra reported for sample no. 1 are giving in Fig. S8. It can see that intensity of the characteristic band of MB at 658 nm remains nearly unchanged in the simple systems, MB + H2O2 and MB + UV for about 200 min. In other words, control experiments indicated that the degradation of MB is hardly proceeded in absence of catalyst (Fig. 7). The degradation efficiency [R = 100 × ((AtA0)/A0)] of MB was determined, where A0 and At are the initial absorbance and that obtained after the illumination time (t). Combined of H2O2 with the UV irradiation increases the percent of discolorization of MB up to about 59%. The addition of α-Fe2O3 nanoparticles to the MB + H2O2 + UV system degraded about 94% (1), 97% (2) and 98% (3) of MB exposed to the UV light within about 100 min. The obtained results are comparable to the previously published data used α-Fe2O3 nanostructured fibers as a catalyst.58 The photo-catalytic degradation of MB was found to be first order reaction [ln(Ct/C0) = kt]. Its ln(Ct/C0) plot shows a linear relationship with the illumination time (Fig. 8), where slope is the rate constant (k). The calculated rate constant for the photo-catalytic reactions was 1.26 × 10−2 (1), 1.89 × 10−2 (2) and 2.44 × 10−2 min−1 (3), while it is only about 2.00 × 10−4 (MB + UV), 1.00 × 10−4 (MB + H2O2) and 26 × 10−4 min−1 (MB + UV + H2O2) for blank samples. The t1/2 value was found to be 55 (1), 37 (2) and 28 min (3) compared with 2.5 days (MB + UV), 5 days (MB + H2O2) and 4.5 hours (MB + UV + H2O2). It is worthy to mention that almost no degradation of MB was observed in dark, which reveals that the mechanism of the reaction is photo-degradation rather than adsorption. As shown in Fig. 7, the decrease in the particle size of the synthesized hematite nanoparticles as those coming from the thermal decomposition of complex 2, compared with compound 1, led to an improvement in the catalytic performance. The reactivity of iron oxide particles has been shown to greatly increase as their dimensions are reduced.59 Decreasing the average particle size increases the specific surface area and thus increases the number of active surface sites. However, the presence of traces of iron oxyhydroxide in sample 3 may be the motivation for the high catalytic performance. The hydroxyl radical that comes from FeOOH may be the primary oxidant that initiate the degradation of the pollutant.


image file: c5ra12157d-f6.tif
Fig. 6 Time-dependent UV-Vis absorption spectra for the photocatalytic degradation of MB with α-Fe2O3 (2) [10 mL dye (14 mg L−1), 0.1 mL H2O2 (35%), 10 mg α-Fe2O3, λ = 365 nm].

image file: c5ra12157d-f7.tif
Fig. 7 Ultraviolet light photo-degradation of MB under different conditions.

image file: c5ra12157d-f8.tif
Fig. 8 Linear plots of ln(Ct/C0) vs. time for the decolorization of MB under different conditions.

Conclusion

Reaction of ciprofloxacin, lomefloxacin and norfloxacin with K3[Fe(C2O4)3] afforded complexes of the type cis-[Fe(C2O4)(FQ)(H2O)2], which were characterized by several analytical and spectral tools. The experimental studies were complemented by quantum chemical calculations. Cis-isomer is found to be more stable than the trans-complex. Electronic spectra were assigned by the aid of TD-DFT calculations. Thermal decomposition of the synthesized complexes is a simple and nontoxic method for preparation of hematite (α-Fe2O3) nanoparticles. The purity of the hematite nanoparticles was verified by EDX analysis. The prepared hematite nanoparticles exhibited good photo-catalytic activity in the degradation process of an industrial pollutant methylene-blue by means of Fenton-like reagent.

Experimental

Instruments

TGA was performed in nitrogen atmosphere (20 mL min−1) in a platinum crucible with a heating rate of 10 °C min−1 using a Shimadzu DTG-60H simultaneous DTG/TG apparatus. FT IR spectra were recorded as KBr pellets using a Jasco FTIR 460. Elemental microanalysis was performed using Elementer Vario EL III. Magnetic measurement was carried out on a Sherwood scientific magnetic balance using Gouy method,39 and Hg[Co(SCN)4] was used as a calibrant. Electronic spectra were scanned on a Shimdazu Lambda 4B spectrophotometer. A digital Jenway 4310 conductivity with a cell constant of 1.02 was used for the molar conductance study. The iron oxide nanoparticle morphologies were investigated by field emission scanning electron microscope (Quanta FEG 250) attached with energy dispersive X-ray analyses for surface analysis. Hematite nanoparticles were characterized by X-ray powder diffraction (PANalytical, X'Pert PRO) operated with copper target (Kα, λ = 1.54056 Å) at 40 kV and 25 mA.

Synthesis

One mmol of ciprofloxacin·HCl (366 mg), lomefloxacin·HCl (386 mg) and norfloxacin (319 mg) was added to one mmol of hot aqueous solution of K3[Fe(C2O4)3]·3H2O (25 mL, 491 mg) (supplied from Sigma chemical company) and heated to reflux (1–3 h), where yellow (1), orange (2) and red (3) complexes were precipitated, respectively. The pH of the reaction mixture was raised by addition of ammonia solution. Lower molar conductance values (in DMF) were reported for the studied complexes compared with the reported values45 for 1[thin space (1/6-em)]:[thin space (1/6-em)]1 (65–90 Ω−1 cm2 mol−1) and 1[thin space (1/6-em)]:[thin space (1/6-em)]2 (130–170 Ω−1 cm2 mol−1) electrolytes in DMF, indicating their non-ionic nature. α-Fe2O3 nanoparticles were prepared by heating complexes 1–3 in an electric furnace under ambient air condition before cooling to room temperature.

• Complex 1: color: yellow. Elemental analysis (%): calcd C19H21FFeN3O9: C 44.73, H 4.15, N 8.24, found C 44.68, H 4.31, N 9.04. FT IR (cm−1): 3436 ν(NH), 1714 ν(C[double bond, length as m-dash]O)ox, 1668 νass(C[double bond, length as m-dash]O)ox, 1628 ν(C[double bond, length as m-dash]O)py, 1476 νss(C[double bond, length as m-dash]O)ox, 1386, 1271, 1181, 1106, 1034. UV-Vis (DMF, 10−4, nm): 285, 320, 335, 625. Molar cond. (10−3 M, DMF, Ω−1 cm2 mol−1): 18.14. μeff (μB, 298 K): 4.86.

• Complex 2: color: orange. Elemental analysis (%): calcd C19H22F2FeN3O9: C 43.04, H 4.18, N 7.92, found C 43.45, H 4.22, N 8.39. FT IR (cm−1): 3431 ν(NH), 1714 ν(C[double bond, length as m-dash]O)ox, 1670 νass(C[double bond, length as m-dash]O)ox, 1630 ν(C[double bond, length as m-dash]O)py, 1554, νass(C[double bond, length as m-dash]O), 1523, 1459 νss(C[double bond, length as m-dash]O)ox, 1402, 1360 νss(C[double bond, length as m-dash]O), 1323, 1273, 1164, 1123, 1093, 1051. UV-Vis (DMF, 10−4, nm): 292, 325, 340, 625. Molar cond. (10−3 M, DMF, Ω−1 cm2 mol−1): 9.73. μeff (μB, 298 K): 6.14.

• Complex 3: color: red. Elemental analysis (%): calcd C18H21FFeN3O9: C 43.39, H 4.25, N 8.43, found C 43.92, H 4.19, N 8.64. FT IR (cm−1): 3442 ν(NH), 1713 ν(C[double bond, length as m-dash]O)ox, 1665 νass(C[double bond, length as m-dash]O)ox, 1627 ν(C[double bond, length as m-dash]O)py, 1548, νass(C[double bond, length as m-dash]O), 1519, 1477, 1386, 1272, 1188, 1139, 1098, 1036. UV-Vis (DMF, 10−4, nm): 285, 320, 335, 625. Molar cond. (10−3 M, DMF, Ω−1 cm2 mol−1): 10.44. μeff (μB, 298 K): 6.15.

Quantum chemical calculations

Ground state geometry optimization, natural bond orbital analysis and molecular electrostatic potential maps of complexes 1–3 were obtained at DFT/B3LYP/6-31G* level of theory using Gaussian03.60 The complexes were characterized as local minima through harmonic frequency analysis. Electronic transitions were calculated by TD-DFT.44–46 The effect of solvent (DMSO) was performed using the default polarizable continuum model.61

Catalytic activity

The photo-catalytic degradation of MB was investigated by taking 10 mL of the dye solution (14 mg L−1, pH = 5) with 0.1 mL hydrogen peroxide (35% v/v) in presence of 10 mg of nano-α-Fe2O3. The suspension was stirred to ensure that all the active sites of catalysts are in contact with the dye solution. Then the sample was illuminated with a UV hand lamp (365 nm) positioned perpendicular at a distance of 2 cm. The irradiation was interrupted in regular intervals to take electronic spectra on the UV-visible spectrophotometer until no more changes in the characteristic band of MB was observed.

References

  1. K. S. Min and M. P. Suh, J. Solid State Chem., 2000, 152, 183 CrossRef CAS; O. Castillo, I. Muga, A. Luque, J. M. Gutiérrez-Zorrilla, J. Sertucha and P. Román, Polyhedron, 1999, 18, 1235 CrossRef; Y. Akhriff, J. Server-Carrió, A. Sancho, J. García-Lozano, E. Escrivá, J. V. Folgado and L. Soto, Inorg. Chem., 1999, 38, 1174 CrossRef PubMed.
  2. K. V. Krishnamurty and G. M. Harris, Chem. Rev., 1961, 61(3), 213 CrossRef CAS; O. Schott, J. Ferrando-Soria, A. Bentama, S. Stiriba, C. Ruiz-Pérez, M. Andruh, F. Lioret and M. Julve, Inorg. Chim. Acta, 2011, 376, 358 CrossRef PubMed.
  3. L. Androš, M. Jurić, P. Planinić, D. Žilić, B. Rakvin and K. Molčaov, Polyhedron, 2010, 29, 1291 CrossRef PubMed.
  4. M. J. Schmelz, T. Miyazawa, S. Mizushima, T. J. Lane and J. V. Quagliano, Spectrochim. Acta, 1967, 9, 51 Search PubMed.
  5. J. C. Blair and E. M. Jones, in Inorganic Synthesis, ed. H. S. Booth, McGraw-Hill Book Co., New York, 1939, vol. 1, p. 35 Search PubMed.
  6. A. Saritha, B. Raju, M. Ramachary, P. Raghavaiah and K. A. Hussain, Phys. B. Condens. Matter., 2012, 407, 4208 CrossRef CAS PubMed.
  7. P. C. Junk, J. Coord. Chem., 2005, 58(4), 355 CrossRef CAS PubMed.
  8. A. Saritha, B. Raju, D. Narayana Rao, A. Roychowdhury, D. Das and K. A. Hussain, Adv. Powder Technol., 2015, 26(2), 349 CrossRef CAS PubMed.
  9. C. A. Parker and C. C. Hotchord, J. Phys. Chem., 1959, 63, 22 CrossRef CAS; C. A. Parker, Trans. Faraday Soc., 1954, 50, 1213 RSC.
  10. W. W. Wendlandt and E. L. Simmons, J. Inorg. Nucl. Chem., 1966, 28, 2420 CrossRef CAS.
  11. M. Izakovič, J. Šima and V. Brezová, J. Photochem. Photobiol., A, 2004, 167, 81 CrossRef PubMed.
  12. K. Sahul, L. V. Natarajan and Q. Anwaruddin, J. Polym. Sci., Polym. Lett. Ed., 1977, 15, 605 CrossRef CAS PubMed.
  13. A. Abbas and I. Tajuddin, J. Photochem., 1986, 35, 87 CrossRef CAS.
  14. A. S. Brar and B. S. Randhawa, J. Solid State Chem., 1985, 58, 153 CrossRef CAS; A. S. Brar and B. S. Randhawa, Radiochem. Radioanal. Lett, 1981, 47, 219 Search PubMed; A. S. Brar and B. S. Randhawa, Radiochem. Radioanal. Lett, 1980, 44, 377 Search PubMed; A. A. Medvinskii and V. I. Poskevnev, Zh. Strukt. Khim., 1974, 15, 623 Search PubMed; G. M. Bancroft, K. G. Dharmawardena and A. G. Maddock, Inorg. Chem., 1970, 9, 223 CrossRef.
  15. T. M. Muziol, G. Wrzeszcz and L. Chrzaszcz, Polyhedron, 2011, 30, 169 CrossRef CAS PubMed; R. Prasad, Sulaxna and A. Kumar, J. Therm. Anal. Calorim., 2005, 81, 441 CrossRef.
  16. P. K. Gallagher and C. R. Kurkjian, Inorg. Chem., 1965, 5, 214 CrossRef CAS; D. Broadbent, D. Dollimore and J. Dollimore, J. Chem. Soc. A, 1967, 451 RSC; W. W. Wendlandt and T. D. George, J. Inorg. Nucl. Chem., 1961, 21, 69 CrossRef.
  17. O. Paulsen, Drugs Today, 1987, 23, 269 Search PubMed.
  18. V. T. Andriole, Clin. Infect. Dis., 2005, 41, S113 CrossRef CAS PubMed; Quinolone Antimicrobial Agents, ed. D. C. Hooper and E. Rubinstein, ASM Press, Washington, DC, 3rd edn, 2003 Search PubMed; D. C. Hooper, J. S. Wolfson, E. Y. Ng and M. N. Swartz, Am. J. Med., 1987, 82(suppl. 4A), 12 Search PubMed.
  19. D. E. King, R. Malone and S. H. Lilley, Am. Fam. Physician, 2000, 61, 2741 CAS.
  20. H. K. Mayer and J. A. Ellal, Am. J. Med., 1999, 92, 58S CrossRef CAS; G. S. Ogawa and R. Hyndiuk, Int. Ophthalmol. Clin., 1993, 33, 59 CrossRef.
  21. I. Turel, Coord. Chem. Rev., 2002, 232(1–2), 27 CrossRef CAS.
  22. L. Fratini and E. E. S. Schapoval, Int. J. Pharm., 1996, 127, 279 CrossRef CAS.
  23. D. S. Lee, H. Han, K. Kim, W. B. Park, J. K. CHO and J. H. Kim, J. Pharm. Biomed. Anal., 1994, 12(2), 157 CrossRef CAS; J. Al-Mustafa and B. Tashtoush, J. Coord. Chem., 2003, 56(2), 113 CrossRef PubMed.
  24. P. B. Issopoulos, Analyst, 1989, 114, 627 RSC.
  25. D. L. Ross and C. M. Riley, Int. J. Pharm., 1992, 87, 203 CrossRef CAS.
  26. F. Gao, P. Yang, J. Xie and H. Wang, J. Inorg. Biochem., 1995, 60, 61 CrossRef CAS.
  27. S. A. Sadeek, J. Mol. Struct., 2005, 753, 1 CrossRef CAS PubMed.
  28. S. C. Wallis, L. R. Gahan, B. G. Charles and T. W. Hambely, Polyhedron, 1995, 14(20–21), 2835 CrossRef CAS.
  29. G. Psomas, J. Inorg. Biochem., 2008, 102, 1798 CrossRef CAS PubMed.
  30. H. F. Abdel-Halim, G. G. Mohamed, M. M. I. El-Dessouky and W. H. Mahmoud, Spectrochim. Acta, 2011, 82, 8 CrossRef CAS PubMed.
  31. L. C. Yu, Z. L. Tang, P. G. Yi and S. L. Liu, J. Coord. Chem., 2009, 62(6), 894 CrossRef CAS PubMed.
  32. J. L. R. Costa, G. S. Marchetti and M. C. Rangel, Catal. Today, 2002, 77, 205 CrossRef CAS; C. Huang, H. Zhang, Z. Sun, Y. Zhao, S. Chen, R. Tao and Z. Liu, J. Colloid Interface Sci., 2011, 364, 298 CrossRef PubMed.
  33. J. Zhang, S. Rama, R. S. Srivastava and R. D. K. Misra, Acta Biomater., 2008, 4, 40 CrossRef CAS PubMed; S. Ghoshal, A. A. M. Ansar, S. O. Raja, A. Jana, N. R. Bandyopadhyay, A. K. Dasgupta and M. Ray, Nanoscale Res. Lett., 2011, 6, 540 CrossRef PubMed.
  34. Z. Medarova, W. Pharm, C. Farrar, V. Petkova and A. Moore, Nat. Med., 2006, 13(3), 372 CrossRef PubMed.
  35. A. H. Lu, E. L. Salabas and F. Schüth, Angew. Chem., Int. Ed., 2007, 46, 1222 CrossRef CAS PubMed.
  36. F. Gao, P. Yang, J. Xie and H. Wang, J. Inorg. Biochem., 1995, 60, 61 CrossRef CAS.
  37. A. M. Mansour and O. R. Shehab, Spectrochim. Acta, Part A, 2014, 128, 263 CrossRef CAS PubMed.
  38. D. K. Saha, D. Deobagkar, U. Sandbhor, K. Shirisha, C. E. Anson, S. Padhye and A. K. Powell, Bioorg. Med. Chem. Lett., 2004, 14, 3027 CrossRef CAS PubMed.
  39. N. T. Abdel-Ghani, M. F. Abo El-Ghar and A. M. Mansour, Spectrochim. Acta, Part A, 2013, 104, 134 CrossRef CAS PubMed.
  40. I. Turel, I. Leban and N. Bukovec, J. Inorg. Biochem., 1994, 56, 273 CrossRef CAS.
  41. A. M. Mansour and R. R. Mohamed, RSC Adv., 2015, 5, 5415 RSC.
  42. A. Bentama, O. Schott, J. Ferrando-Soria, S. Stiriba, C. Ruiz-Pérez, M. Julve and J. Pasán, Inorg. Chim. Acta, 2012, 389, 52 CrossRef CAS PubMed; X. Chen, L. Liu, J. Ma, L. Yi, P. Cheng, D. Liao, S. Yan and Z. Jiang, J. Mol. Struct., 2005, 750, 94 CrossRef PubMed.
  43. M. Massacesi and G. Ponticelli, Spectrochim. Acta, Part A, 1981, 37(12), 1035 CrossRef.
  44. A. M. Mansour, Polyhedron, 2014, 78, 10 CrossRef CAS PubMed.
  45. A. M. Mansour, Dalton Trans., 2014, 15950 RSC.
  46. O. R. Shehab and A. M. Mansour, J. Mol. Struct., 2013, 1047, 121 CrossRef CAS PubMed.
  47. A. E. Reed, L. A. Curtiuss and F. Weinhold, Chem. Rev., 1988, 88, 899 CrossRef CAS.
  48. N. T. Abdel-Ghani and A. M. Mansour, Inorg. Chim. Acta, 2011, 373, 249 CrossRef CAS PubMed.
  49. A. M. Mansour, Inorg. Chim. Acta, 2013, 408, 186 CrossRef CAS PubMed.
  50. P. Guardia, J. Pérez-Juste, A. Labarta, X. Batlle and L. M. Marza, Chem. Commun., 2010, 46, 6108 RSC.
  51. B. Zargar, H. Parham and M. Rezazade, J. Chin. Chem. Soc., 2011, 58, 694 CrossRef CAS PubMed.
  52. M. Rafatullaha, O. Sulaimana and R. Hashim, J. Hazard. Mater., 2010, 177, 70 CrossRef PubMed.
  53. M. Amini, B. Pourbadiei, T. P. A. Ruberu and L. K. Woo, New J. Chem., 2014, 38, 1250 RSC.
  54. I. Ursachi, A. Stancu and A. Vasile, J. Colloid Interface Sci., 2012, 377, 184 CrossRef CAS PubMed.
  55. P. V. Nidheesh, RSC Adv., 2015, 5, 40552 RSC.
  56. M. A. Andrade, R. J. Carmona, A. S. Mestre, J. Matos, A. P. Carvalho and C. O. Ania, Carbon, 2014, 76, 183 CrossRef CAS PubMed; G. Thennarasu and A. Sivasamy, Powder Technol., 2013, 250, 1 CrossRef PubMed.
  57. S. Chin, E. Park, M. Kim and J. Jurng, Powder Technol., 2010, 201, 171 CrossRef CAS PubMed; E. M. Seftel, M. Niarchos, Ch. Mitropoulos, M. Mertens, E. F. Vansant and P. Cool, Catal. Today, 2015, 252, 120 CrossRef PubMed.
  58. Y. Liu, Y. Li, H. Yu, Z. Lv, Y. Yu, S. Zhan and X. Yang, J. Sol-Gel Sci. Technol., 2011, 58, 716 CrossRef CAS.
  59. E. O. Kajander and N. Ciftcioglu, Proc. Natl. Acad. Sci. U. S. A., 1998, 95, 8274 CrossRef CAS.
  60. M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, V. G. Zakrzewski, J. A. Montgomery, R. E. Stratmann, J. C. Burant, S. Dapprich, J. M. Millam, A. D. Daniels, K. N. Kudin, M. C. Strain, O. Farkas, J. Tomasi, V. Barone, M. Cossi, R. Cammi, B. Mennucci, C. Pomelli, C. Adamo, S. Clifford, J. Ochterski, G. A. Petersson, P. Y. Ayala, Q. Cui, K. Morokuma, D. K. Malick, A. D. Rabuck, K. Raghavachari, J. B. Foresman, J. Cioslowski, J. V. Ortiz, A. G. Baboul, B. B. Stefanov, G. Liu, A. Liashenko, P. Piskorz, I. Komaromi, R. Gomperts, R. L. Martin, D. J. Fox, T. Keith, M. A. Al-Laham, C. Y. Peng, A. Nanayakkara, C. Gonzalez, M. Challacombe, P. M. W. Gill, B. G. Johnson, W. Chen, M. W. Wong, J. L. Andres, M. Head-Gordon, E. S. Replogle and J. A. Pople, GAUSSIAN 03 (Revision A.9), Gaussian, Inc., Pittsburgh, 2003 Search PubMed.
  61. N. T. Abdel-Ghani and A. M. Mansour, Spectrochim. Acta, Part A, 2012, 91, 272 CrossRef CAS PubMed.

Footnote

Electronic supplementary information (ESI) available. See DOI: 10.1039/c5ra12157d

This journal is © The Royal Society of Chemistry 2015
Click here to see how this site uses Cookies. View our privacy policy here.