Stabilization of Monodisperse, Phase-Pure MgFe2O4 Nanoparticles in

Nov 11, 2017 - Monodisperse, monocrystalline magnesium ferrite (MgFe2O4) nanoparticles were synthesized phase purely by fast nonaqueous microwave-assi...
104 downloads 20 Views 1MB Size
Subscriber access provided by READING UNIV

Article 2

4

Stabilization of Monodisperse, Phase-Pure MgFeO Nanoparticles in Aqueous and Non-Aqueous Media and Their Photocatalytic Behavior Kristin Kirchberg, Anna Becker, André Bloesser, Tobias Weller, Jana Timm, Christian Suchomski, and Roland Marschall J. Phys. Chem. C, Just Accepted Manuscript • DOI: 10.1021/acs.jpcc.7b08780 • Publication Date (Web): 11 Nov 2017 Downloaded from http://pubs.acs.org on November 14, 2017

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

The Journal of Physical Chemistry C is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 15

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Stabilization of Monodisperse, Phase-pure MgFe2O4 Nanoparticles in Aqueous and Non-aqueous Media and Their Photocatalytic Behavior Kristin Kirchberg, Anna Becker, André Bloesser, Tobias Weller, Jana Timm, Christian Suchomski, and Roland Marschall* Justus-Liebig-University Giessen, Heinrich-Buff-Ring 17, 35392 Giessen, Germany ABSTRACT: Monodisperse, monocrystalline magnesium ferrite (MgFe2O4) nanoparticles were synthesized phase-purely by fast non-aqueous microwave-assisted solution phase synthesis. Colloidal stabilization of the nanocrystals in non-aqueous media was realized either in-situ during synthesis or post-synthetically by surface capping with oleylamine and oleic acid. Phase transfer to aqueous media was performed employing citric acid and betaine hydrochloride, resulting in agglomerate free dispersions of citrateor betaine-functionalized MgFe2O4 nanocrystals. Furthermore, a one-step synthesis of highly stable, water-dispersible colloids of MgFe2O4 was achieved using polyvinylpyrrolidone as stabilizer. Characterization of the as-synthesized and functionalized nanoparticles was performed employing X-ray diffraction, UV-Vis and infrared spectroscopy, thermogravimetry, dynamic light scattering and transmission electron microscopy. Special focus was laid on phase purity, which was thoroughly monitored using Raman microscopy/spectroscopy. Photocatalytic reactions were performed to evaluate the use of such highly stable ferrite colloids for solar energy conversion.

1. Introduction Magnetic nanoparticles have gained significant attention in the fields of catalysis, magnetic resonance imaging, biomedicine, and data storage.1–5 In recent years, iron oxide based materials have proven to be interesting due to their low toxicity, high magnetic susceptibility and large variety of compositions. Recently, cubic spinel-structured iron oxide nanoparticles (MFe2O4), also called ferrite nanoparticles, have attracted attention due to their magnetic and optical properties, which make them suitable for biomedical applications,1,6 sensing,7–10 heterogeneous catalysis and photocatalysis,11–14 battery technologies,15,16 electrical devices,17 and high-density information storage.18 Especially magnetite (Fe3O4) has been investigated intensively, but also magnesium ferrite (MgFe2O4) gained a lot of attention, due to its narrow band gap and band position reported, which make them attractive for application in photocatalytic processes such as hydrogen production.7,19,20 The synthesis of ferrite nanoparticles has been reported employing numerous different techniques, e.g. sol-gel,21–23 mechanochemical,24–26 solvothermal,27–33 co-precipitation,34,35 microwave-assisted24,36,37 and high-temperature38–40 routes, in some cases under inert gas atmosphere. However, only few reports take account of minor impurities of iron oxide byphases within their samples using Raman spectroscopy as a highly sensitive tool in addition to powder X-ray diffraction (PXRD) analysis. The stabilization of magnetic nanoparticles in non-aqueous and aqueous solution can improve their handling due to lower risk of exposure to fine dusts, and even make them suitable for all liquid applications. Ferrofluids have proven to be interesting for application in exclusion seals, sensors, dampers and

shock absorbers, for magnetic resonance imaging (MRI) contrast enhancement and as cell labelling agent because of their superparamagnetic behavior and easy liquid handling.41–45 For iron oxide nanoparticles there is a range of possible stabilizing methods in non-aqueous and aqueous media reported for various nanoparticle shapes and sizes.46–50 There are a number of reports about hydrophilically surfacemodified ferrite nanoparticles, mostly in the context of biomedical applications, but colloidal stability of the surface-modified nanoparticles is usually not discussed.6,43,51–53 In this context, surface-modified MFe2O4 (M = Zn, Mn, Co, Mg) were reported by Naseri et al. using poly-(vinyl pyrrolidone) (PVP) and poly(vinyl alcohol) (PVA) as capping agents, which led to controlled growth, decreased particle agglomeration during calcination, and uniform size distribution of the ferrite nanoparticles. However, a drawback of this aqueous approach is the formation of different oxide impurities such as hematite (α-Fe2O3), ZnO or MgO, respectively.52,53 Chandradass et al. reported the surface functionalization of MgFe2O4 nanoparticles with Igepal® CO-520 (branched polyoxyethylene nonylphenylether) during reversed micelle synthesis. Variation of the water-to-surfactant ratio resulted in different particle sizes of 20-30 nm with small size distribution. With a high saturation magnetization of 40 emu/g, the application of the prepared MgFe2O4 nanoparticles in hyperthermia treatments was suggested.51 Thermal decomposition of organic metal precursors (e.g., oleates, carbonyls and acetylacetonates) in the presence of oleic acid (OA) in high boiling solvents in combination with other ligands like trioctylphosphine oxide (TOPO) or oleylamine (OLA) is a common method to synthesize stable non-polar colloids of MFe2O4 nanoparticles (M = Fe, Mn, Co,

S1

ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Zn, Ca, Mg, Cu).38,40,54–57 In contrast, the phase transfer into polar solvents or the direct synthesis of hydrophilic ferrite colloids is still subject of current research. Munjal et al. successfully performed the phase transfer of OA-coated CoFe2O4 nanoparticles, which were synthesized by hydrothermal method. Chemical treatment with citric acid in a solvent mixture of toluene and dimethyl sulfoxide (DMSO) led to the precipitation of water-dispersible, citric-acid-coated, polycrystalline CoFe2O4 nanoparticles of 13 ± 2 nm.31 Regarding applications, Singh et al. investigated the adsorptive removal of crystal violet cationic dye by sodium dodecyl sulfate (SDS)-modified CoFe2O4 nanoparticles. It was observed that the nanoparticle and dye concentration, the pH value, and the contact time influence the efficiency of the dye adsorption and therefore its removal from solution.58 Zhao et al. produced citrate-stabilized ZnFe2O4 nanoparticles by thermal decomposition of metal acteylacetonate precursors with sodium citrate in triethylene glycol (TREG) at reflux temperature. They investigated the influence of the surfactant-to-precursor ratio, finding that high surfactant amounts lead to low aggregated nanoparticle dispersions with good stability and small size distribution.59 For most of the reported synthesis procedures for bare and functionalized MgFe2O4 and ZnFe2O4 colloidal solutions, no Raman spectroscopy data were provided to confirm the results from X-ray diffraction phase analysis. As ferrites, being an iron oxide modification, tend to form various by-phases, e.g. maghemite (γ-Fe2O3) or the thermodynamically stable modification hematite (α-Fe2O3), the investigation of phase purity using Raman spectroscopy is helpful to detect small amounts of impurity phases, as this technique is very sensitive in differentiation between various iron oxide species in nanocrystalline solids.60–62 Due to similar cell parameters and crystal structures of the different iron-containing materials, typical reflections in X-ray diffraction differ only slightly. In combination with a large width of the reflections because of small coherent scattering domains in nanoparticles, sole X-ray diffraction phase analysis is not sufficient to detect minor impurities. Here, we have developed a one-pot synthesis for aqueous MgFe2O4 colloids by microwave-assisted non-aqueous synthesis. The production of stable non-aqueous colloids of phasepure MgFe2O4 nanoparticles was achieved by post-synthetic functionalization with long-chain organic ligands. Phase transfer of the highly stable hydrophobic colloids into aqueous medium was performed employing citric acid and betaine hydrochloride, resulting in agglomerate free dispersions of citrate- or betaine-functionalized MgFe2O4 nanocrystals. Furthermore, first results on the photocatalytic activity of stabilized MgFe2O4 nanoparticles and the advantage of stabilization towards degradation of rhodamine B (RhB) are presented.

2. Experimental Section 2.1 Synthesis of MgFe2O4 nanoparticles

Page 2 of 15

For microwave-assisted synthesis, a Monowave 300 (f = 2.45 GHz, Anton Paar Germany GmbH) equipped with a hydraulic pressure sensor was employed. The temperature was monitored using a ruby thermometer, which was placed inside a borosilicate glass vessel (30 mL, Anton Paar Germany GmbH) sealed with a screw cap. All reaction solutions were magnetically stirred at 300 rpm during microwave heating.

2.1.1 Microwave-assisted synthesis of non-stabilized, mono-crystalline MgFe2O4 nanoparticles was per-

formed using a modified procedure reported elsewhere.37 Briefly, a mixture of 0.5 mmol magnesium acetyl-acetonate (Mg(acac)2, 98 %, Acros) was dissolved in 15 mL dry rac-1-phenylethanol (>99 %, Sigma Aldrich) using ultrasonication. Then, 1 mmol iron(III)-acetylacetonate (Fe(acac)3, >99 %, Sigma Aldrich) was added to the solution and sonicated for 5 minutes before being transferred into a 30 mL borosilicate glass vessel. The obtained solution was heated to maximum temperatures of 200, 250, 275 and 300 °C for 30 min using microwave irradiation under constant stirring (300 rpm). Next, the reaction mixture was quenched with compressed air to 55 °C. The nanoparticles were precipitated with trichloromethane, washed with distilled water and with diethyl ether. Finally, the obtained powder was allowed to dry at room temperature, resulting in a fine powder with brown color. The pH of an 8 mg/L suspension of these nanoparticles was 9.8. Calcination of as-prepared MgFe2O4 nanoparticles in air was performed in a muffle furnace. Nanoparticles were placed inside a ceramic crucible and heated to different temperatures between 400 °C and 1000 °C with a heating rate of 10 °C/min, respectively. The maximum temperature was maintained for 60 min before cooling the samples to room temperature.

2.1.2 Microwave-assisted solution phase synthesis of oleylamine-/oleic acid-functionalized MgFe2O4 nanoparticles was performed modifying a synthesis procedure reported by Sun et al.38 The synthesis was carried out using Monowave 300 instead of the batch setup reported in literature. Briefly, a mixture of 5 mmol 1,2-dodecanediol, 0.33 mmol Mg(acac)2, 0.66 mmol Fe(acac)3, 3 mL oleylamine (OLA, 70 %, Sigma Aldrich), 3 mL oleic acid (OA, 90 %, ABCR) and 10 mL dibenzyl ether (>98 %, Fluka) was sonicated until all solid components were fully dissolved. Then, the magnetically stirred mixture was heated up to 275 °C under microwave irradiation in a 30 mL borosilicate glass vessel. After 30 min, the reaction was quenched using compressed air. The nanoparticles were precipitated with ethanol (absolute, VWR). Redispersion in a mixture of n-hexane (99 %, abcr) with 0.25 mL OA and 0.25 mL OLA was followed by centrifugation at 6000 rpm to separate nondispersible components. The stable supernatant was then washed three times with ethanol before being redispered in toluene (99 %, abcr; 25 mg/mL).

2.1.3 Microwave-assisted synthesis of oleylamine/oleic acid-functionalized MgFe2O4 nanoparticles was performed by adding 4.5 mL of OA and 4.5 mL of OLA to a reaction mixture described in 2.1.1. The work up after microwave heating was identical with that described under 2.1.2.

S2

ACS Paragon Plus Environment

Page 3 of 15

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

2.1.4 Microwave-assisted solution phase synthesis of water-dispersible, monocrystalline MgFe2O4 nanoparticles was performed adapting a procedure by Abdulwahab et al. for microwave-assisted synthesis.39 Here, 0.05 mmol Mg(acac)2 were dissolved in 12 mL of triethylene glycol (TREG, 99 %, Sigma Aldrich). Then, 0.1 mmol Fe(acac)3, 150 mg polyvinylpyrrolidone (PVP, MW 10 000, Alfa-Aesar) and 3 mL TREG were added, and the mixture was sonicated until all solid components were completely dissolved. Microwave heating to 275 °C for 45 min was followed by precipitation with a mixture of diethyl ether and acetone (1:2). The precipitate was washed with diethyl ether and dried at room temperature. The obtained nanoparticles were dispersible in polar solvents, e.g. water (15 mg/mL).

2.1.5 Microwave-assisted synthesis of PVP-modified monocrystalline MgFe2O4 nanoparticles was performed for comparison by adding 1.5 g PVP to the microwaveassisted synthesis presented in 2.1.1, following reaction for 45 min at 275 °C. Particles were precipitated in a mixture of acetone and diethyl ether. The precipitate was washed with diethyl ether and then dried at room temperature. The nanoparticles were dispersible in polar solvents, e.g. water (15 mg/mL). The pH of an 8 mg/L colloidal solution of these nanoparticles was 7.5.

2.2 Post-synthetic treatment of MgFe2O4 nanoparticles Several methods were used for surface functionalization on the one hand to obtain stable colloids in polar and non-polar solvents, and ligand stripping on the other hand to obtain bare MgFe2O4 nanoparticles. Furthermore, the nanoparticle samples prepared after 2.1.1 were calcined at different temperatures.

2.2.1 Post-synthetic stabilization of MgFe2O4 to form stable non-polar colloids was performed for nanoparticles obtained from microwave-assisted synthesis described under 2.1.1. An approach described by Patil et al.63 was therefore modified. Briefly, 20 mg of MgFe2O4 nanoparticles were added to a mixture of 1 mL of OA and 1 mL OLA in 10 mL toluene. The solution was refluxed for 48 hours. Afterwards, the nanoparticles were precipitated with methanol (>99.9 %, Sigma Aldrich) and dried at room temperature before being dispersed in non-polar solvents such as toluene. Here, nanoparticle concentrations up to 25 mg/mL were possible.

2.2.2 Phase transfer of non-polar colloids into polar solutions was performed using two different phase transfer agents. At first, an approach by Patil et al.63 was adapted. To obtain stable aqueous colloids, 8 mg MgFe2O4 nanoparticles were stirred in a betaine hydrochloride solution of 2 wt.-% (>98 %, TCI) for six hours. Precipitation with acetone was followed by drying at room temperature. Nanoparticles functionalized with betaine hydrochloride were afterwards dispersible in polar solvents, e.g. water (8 mg/mL, pH = 4.6). In a second approach after Lattuada et al.,46 30 mg of OLA/OA-functionalized MgFe2O4 nanoparticles were added to a solution of 1.9 mL N,N-dimethylformamide (99 %, Acros), 1.9 mL 1,2-dichlorobenzene (>99 %, Merck), and

25 mg citric acid monohydrate (>99.5 %, Alfa-Aesar). The mixture was kept at 100 °C for 24 hours. Then the nanoparticles were precipitated using diethyl ether and washed with acetone. After drying at room temperature, the particles were dispersible in water (8 mg/mL, pH = 8.4).

2.2.3 Ligand stripping of MgFe2O4 nanoparticle surfaces was performed using a modified procedure reported by Dong et al..64 Here, a solution of 5 mg nanoparticles per mL toluene was mixed with 5 mL of triethyloxonium tetrafluoroborate solution (10 mM Et3OBF4 (>95 %, TCI)) per liter acetonitrile (99.9 %, Acros)) and stirred for 2 hours. Then, distilled water was added, which led to the transfer of MgFe2O4 nanoparticles into the aqueous phase. The particles were washed with distilled water and then dried at room temperature. The pH of an 8 mg/L suspension of these nanoparticles was 5.7.

2.3 Materials Characterization UV-Vis absorption spectra were collected from 800 nm to 200 nm on a PerkinElmer Lambda 750 UV/Vis-NIRspectrometer equipped with PrayingMantis mirror arrangement. The step width was 1 nm. Data evaluation was performed using UV Winlab software (PerkinElmer). For degradation experiments, 3 mL of the observed solutions were placed inside a 1 cm quartz glass cuvette (Hellma). Thermogravimetric analysis (TGA) was performed on a STA409PC thermos scale (Netzsch) in combination with a QMG421 quadrupole mass spectrometer (MS) from Balzers with a heating rate of 5 °C/min in synthetic air (80 % N2, 20 % O2) or argon. The ionization energy was set to 70 eV and a temperature range between 30 °C and 1000 °C was analyzed. All samples were dried at 80 °C for several hours prior to TGA-MS analysis. Powder X-ray diffraction (PXRD) measurements were performed on a PANalytical Empyrean powder diffractometer with a θ-2θ geometry equipped with a PIXcel3D detector using Cu Kα radiation (λ = 1.540598 Å). Diffractograms were recorded between 25 ° and 65 ° 2θ in a step scan mode with an emission current of 40 mA and an acceleration voltage of 40 kV. In-situ PXRD measurements were performed in a melted off glass capillary at the I15 beamline at the Diamond Light Source, Didcot with an acceleration voltage of 60 kV (λ = 0.4246 Å). The capillary was placed in the beam focus and stirred in a stable upfront position. The heating process was inducted by a Gas blower GSB 1300 from FMB Oxford, which heated the sample as fast as possible (heating rate approx. 1 °C/min). The data was collected on a mar345 image plate detector. Crystallite sizes were calculated using Scherrer equation for (400) and (440) reflections of MgFe2O4 (JCPDS card no. 36-0398). For nitrogen physisorption measurements at 77 K, an Autosorb-1-MP (Quantachrome) setup was used. The nanoparticles we degassed in vacuum at 120 °C for at least 20 hours prior to the measurement.

S3

ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Infrared (IR) spectra were collected on a Bruker IFS25 FTIR spectrometer in a frequency range between 4000 cm-1 and 400 cm-1 using KBr pellets. Dynamic light scattering (DLS) and zeta potential measurements were performed on a Zetasizer Nano ZS (Malvern) in high dilution. For DLS measurements a refractive index of 2.39 for MgFe2O4 was assumed.65 For analysis of aqueous solutions polystyrol cuvettes (d = 1cm) were used, whereas quartz glass cuvettes (d = 1cm, Hellma) were chosen for nonaqueous solutions. Three consecutive measurements were performed for each sample and the average number distribution was used for particle size distribution evaluation. Zeta potentials were measured for aqueous solutions applying the Smoluchowski equation. Every sample was measured six times using polycarbonate DTS capillaries (Malvern). All DLS and zeta potential measurements were performed at 25 °C. Transmission electron microscopy (TEM) images were collected on a CM 30-ST electron microscope (Philips) equipped with a LaB6 cathode (300 kV). Samples were suspended in either water or toluene and drop-casted on carbon sputtered copper grids (Plano GmbH) at room temperature. After drying, the grids were loaded. Raman spectroscopy was performed using a Senterra Raman spectrometer (Bruker) with OPUS 7.5 software. A Nd:YAG laser (λ = 532 nm) with a laser power of 0.2 mW and a magnification of 100 was chosen. Measurements were performed with a spectral resolution of 3-5 cm-1, 250 co-addition and 3 seconds integration time. The nanoparticle samples were dried in a drying oven at 90 °C prior to the measurements. Then, the powders were spread on a microscope slide.

Page 4 of 15

(PXRD) patterns after synthesis at different temperatures. All major reflections attributable to spinel magnesium ferrite (JCPDS card no. 36-0398) are detected after 200 °C with a crystallite size of approx. 4.9 nm. With increasing synthesis temperature, the reflections become slightly narrower with an estimated crystallite size of 6.3 nm at 250 °C, 6.7 nm at 275 °C and 9.3 nm at 300 °C, respectively. However, at 300 °C additional reflections were observed, which derive from impurities out of the PTFE-coated silicone septum of the microwave vial, small amounts can be already seen at 275 °C. Those additional reflections do not derive from iron oxide by-phases, which can be determined via Raman spectroscopy of the as-synthesized samples after the different synthesis temperatures, also shown in Figure 1. Five Raman-active modes for cubic MgFe2O4 (space group Fd3m) were observed at ≥ 250 °C, as expected.66 The sample prepared at 200 °C shows only very weak Raman signals due to lower crystallinity. No Raman bands of α-Fe2O3, other iron oxides, or MgO were observed, which demonstrates that the as-synthesized MgFe2O4 nanoparticles do not contain these by-phases.61,67 However, an additional shoulder band of the A1g mode can be seen at roughly 660 cm-1, indicating a defective inverse spinel structure in which both Fe3+ and Mg2+ ions occupy tetrahedral sites, which is in good agreement with literature results.68,69 In a defect-free magnesioferrite spinel structure, Mg2+ ions are solely located on octahedral sites.

2.4 Testing the photocatalytic activity of MgFe2O4 nanoparticles 10 mg MgFe2O4 nanoparticles were mixed with 20 mL rhodamine B solution (10 µmol L-1, resulting nanoparticle concentration 0.5 mg mL-1). Constant homogenization was performed using low-power ultrasonication (ultrasonic bath). Before irradiation, the reaction mixture was sonicated for 30 minutes in the dark to reach adsorption-desorption equilibrium of RhB with the particles. To avoid temperature increase, a glass coil flushed with temperature-controlled water was immersed into the ultrasonic bath keeping the reaction temperature constant at 27 ± 1 °C. In intervals of 30 minutes, 3 mL of the reaction solution were taken and the particles were separated with a magnet. The clear solution was measured with absorption spectroscopy in the range of 200-800 nm. Irradiation was performed with a 150 W solar simulator (Newport Sol1A) equipped with an AM 1.5G filter.

3. Results and Discussion 3.1 Structural characterization and thermal stability of MgFe2O4 nanoparticles Microwave-assisted synthesis of non-stabilized MgFe2O4 nanoparticles resulted in crystalline powder materials directly after the synthesis procedure without any additional heattreatment Figure 1 shows the powder X-ray diffraction

Figure 1. PXRD patterns with assigned reflections (JCPDS card no. 36-0398) and Raman data with assigned vibrational bands66 of as-synthesized MgFe2O4 nanoparticles after microwave-assisted synthesis at different reaction temperatures.

S4

ACS Paragon Plus Environment

Page 5 of 15

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

The direct band gap for all as-prepared nanoparticles in the given temperature range can be determined from absorption spectroscopy to be 2.43 eV (exemplarily shown in Figure S1, Supporting Information SI), which in good agreement with the direct band gap of MgFe2O4 nanoparticles reported in literature, and slightly higher than the band gap of bulk MgFe2O4.7,70 The observed blue-shift is likely due to a quantum size effect increasing the band gap caused by the small crystallite sizes below 10 nm.71 FTIR spectra of the assynthesized nanoparticles indicate residual carboxylate species on the surface of the nanoparticles (Figure S2 SI), and an increasing number of OH-groups with increasing synthesis temperature. At 300 °C, additional bands from residues of the corroded teflon foil (i.e., PTFE-coated silicone septum) was observed. Therefore, 275 °C was used as standard synthesis temperature for the stabilization experiments. The nanoparticles were found to be monocrystalline as proven by high resolution TEM analysis (HR-TEM) (Figure S3 SI).

ing to α-Fe2O3 (indicated with * in Figure 2) and MgO (indicated with # in Figure 2). However, Raman data indicate the formation of α-Fe2O3 already at 600 °C, showing that Raman spectroscopy is a very powerful tool to analyze the phase purity of ferrite spinels, especially in the case of controlling αFe2O3 by-phase formation.61,62 The crystallite size of the calcined MgFe2O4 samples increases up to 41 nm at 1000 °C, with the band gap decreasing to a bulk value of 2.2 eV (Table S1, SI). The occurrence of minor amounts of α-Fe2O3 cannot be prevented at elevated temperatures: A very small fraction of Fe2+ is formed during the microwave-assisted synthesis by partial oxidation of 1-phenylethanol with Fe3+ above 200 °C. Small quantities of acetophenone were found in the reaction solution evidenced by gas chromatography connected with mass spectrometry (data not shown). For that reason, besides MgFe2O4, small amounts of Fe3O4 (magnetite) or even Fe2+-containing 

 MgFe2O4 (as (Mg Fe  )A[Fe )]BO4-δ) could be present. Fe3O4 has an inverse spinel structure with one third Fe2+ cations occupying octahedral sites in the ccp-lattice formed by oxygen atoms. It has very similar PXRD reflections and Raman bands like MgFe2O4. Thus, its existence cannot unambiguously be proven due to the reflection broadening caused by the small nanoparticles. Structural changes induced by thermal treatment are expected at temperatures ≥ 300 °C, accompanied with oxidation of Fe2+ to Fe3+ and formation of a mixture of MgFe2O4 and small amounts of γ-Fe2O3 (= (Fe3+)8[Fe  / 1/6)]16O32, maghemite) in the presence of oxygen.72 The latter compound is stable up to 500-600 °C, and has again very similar PXRD reflections and Raman bands like MgFe2O4. Maghemite (γ-Fe2O3, JCPDS card no. 04-0755) transforms into α-Fe2O373 (JCPDS card no. 330664) above 500-600 °C, as confirmed by PXRD and Raman data, which confirms indirectly the initial existence of Fe2+ and the transformation mechanism described. Even these results underline the necessity and importance of Raman investigations regarding the phase purity of the synthesized and thermal treated nanoparticles.

Figure 2. PXRD patterns with assigned reflections (JCPDS card no. 36-0398) and Raman spectra with assigned vibrational bands66 of MgFe2O4 nanoparticles (microwave-assisted synthesis at 275 °C) calcined in air at indicated temperatures; impurity phases observed after the heat treatment are marked (* α-Fe2O3, # MgO).

In order to investigate the thermal stability of the nanoparticles, the as-synthesized sample prepared at 275 °C was calcined at different temperatures (for details please refer to the Experimental section). The respective PXRD patterns and Raman results are shown in Figure 2. As can be seen, the reflections in the PXRD patterns become much narrower after calcination indicating crystallite growth. Moreover, at temperatures above 700 °C additional reflections are found, belong-

Calcination of the as-synthesized MgFe2O4 nanoparticles in an oxygen-free environment should lead to no formation of maghemite, and accordingly to no formation of hematite. Figure S4 shows PXRD patterns and Raman data of MgFe2O4 nanoparticles (microwave-assisted synthesis at 275 °C) after calcination in oxygen-free conditions (sealed capillary) to 650 °C. As a result of the oxygen-free calcination, no hematite formation can be observed. However, since IR spectra (Figure S2) showed residual carboxylate species on the surface of the nanoparticles, those could act reductive on the spinel lattice. In order to investigate the thermal stability of the MgFe2O4 nanoparticles in more detail, in-situ PXRD with hard X-ray irradiation at the Diamond Light source, Beamline I15, in the temperature range of 20-640 °C (PXRD pattern measured every 20 °C in a closed capillary) were performed. All the reflections of the PXRD pattern of the as-synthesized sample (bottom pattern in Figure 3) could be assigned to MgFe2O4 (JCPDS card no. 36-0398).

S5

ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Additionally, the full width at half maximum (FWHM) of all reflections deriving from MgFe2O4 nanoparticles is decreasing with increasing temperature, indicating that the coherent scattering domains grow with increasing temperature. Moreover, the intensities of the reflections increase with increasing temperature, both indicating higher crystallinity of the samples after heat treatment. These trends confirm the ex-situ measurements (Figure 2 & Table S1, SI). At 600 °C, in-situ synchrotron measurements show additional reflections (between (311) and (222) reflection) and shoulders (at (400) & (440) reflection) verifying the presence of MgO (marked with # in the pattern), which could not be observed with the lab instrument (Figure S4). Finally, in the PXRD pattern which was collected after cooling the heated sample down to room temperature, reflections of MgO as a by-phase are clearly visible, no α-Fe2O3 formation is detected. The results confirm the inhibited formation of maghemite under oxygen-free conditions, resulting in no α-Fe2O3 formation. MgO is however formed due to reducing conditions present in the sealed capillary due to residual precursors on the nanoparticle surface, clearly visible with synchrotron radiation.

Page 6 of 15

hard X-ray irradiation (λ = 0.4246 Å). Further diffractograms between room temperature and 500 °C are given in Figure S5.

3.2 In-situ stabilization MgFe2O4 nanoparticles

of

microwave-derived

In order to disperse the nanoparticles without agglomeration in non-polar environments, the addition of stabilizing agents to the microwave synthesis was investigated. As mentioned in the Experimental section, the high-temperature solution phase synthesis of Sun et al. was adapted to the microwave approach. Additionally, common surfactants OA and OLA were added to the presented rac-1-phenylethanol-based microwave synthesis. In both cases, well dispersible nanoparticles with a size of 7.2 to 8.2 nm were achieved. PXRD patterns indicate phasepure MgFe2O4 nanoparticles after both syntheses (Figure 4). Moreover, TEM (Figure S6) and DLS measurements (Figure S7) show a very narrow size distribution and a low degree of agglomeration. However, Raman spectroscopy revealed that the resulting nanoparticles are not phase-pure or even MgFe2O4 but rather a mixture of iron oxides with some amounts of MgFe2O4 (Figure 4).61,62 Although the PXRD patterns indicate phase-pure samples, it has to be noted that feroxyhyte (δ-FeOOH, JCPDS card no. 98-003-8299) and maghemite (γ-Fe2O3, JCPDS card no. 04-0755) exhibit a very similar crystal structure and therefore similar XRD peak positions as MgFe2O4, thus cannot be distinguished at crystallite sizes below 10 nm due to peak broadening. Therefore, Raman spectroscopy is used as suitable tool to investigate the phase composition in detail.

Figure 3. Temperature-dependent in-situ PXRD patterns of assynthesized MgFe2O4 (275 °C, olive) prepared by microwaveassisted synthesis measured at the Diamond Light Source with

S6

ACS Paragon Plus Environment

Page 7 of 15

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry stable colloidal dispersions of MgFe2O4 nanocrystals in water were successfully prepared in a one-step approach. More details about the stabilization agents are given in the next chapter. Detailed characteristics (e.g. zeta potentials, isoelectric points) of non-stabilized and stabilized MgFe2O4 nanoparticles can be found in Table S2. The BET surface area of ligand-stripped nanoparticles was found to be 40 m²/g. This indicates an agglomeration of the nanoparticles, which goes along with the full removal of surface groups. Also, the isoelectric point compared to non-functionalized nanoparticles increases.

3.3 Post-synthetic stabilization in non-aqueous media and phase-transfer To obtain stable dispersions of phase-pure MgFe2O4 nanocrystals in non-polar media, the non-stabilized nanocrystals were post-synthetically functionalized with OA/OLA. As a result, stable dispersions in common non-polar solvents, e.g. trichloromethan or toluene, were achieved. Figure 5 shows DLS results after post-synthetic OLA/OA-functionalization of MgFe2O4 nanocrystals in toluene. The particles showed a BET surface are of 4.7 m²/g due to organic ligands strongly forcing agglomeration during the drying of the colloidal solution. The FTIR spectra of surface-functionalized MgFe2O4 nanocrystals are given in Figure 6.

Figure 4. PXRD patterns with assigned reflections (JCPDS card no. 36-0398) (above) and Raman spectra with assigned vibrational bands61,66 (below) of microwave-derived MgFe2O4 nanoparticles synthesized at 275 °C according to 2.1.1 (A), adapting a synthesis procedure reported by Abdulwahab et al. according to 2.1.4 (B), by PVP addition to the 275 °C synthesis according to 2.1.5 (C), by OLA/OA addition to the 275 °C synthesis according to 2.1.3 (D), and adapting a synthesis procedure reported by Sun et al. according to 2.1.2 (E).

For comparison, the DLS data for the in-situ PVPfunctionalized MgFe2O4 nanoparticles (PVP addition to standard synthesis, (C) in Figure 4) are shown. The OA/OLAfunctionalized MgFe2O4 nanoparticles in toluene show a much narrower size distribution, with a maximum of around 20 nm hydrodynamic diameter. A slight tailing was determined up to 40-50 nm, indicating the formation of a few small agglomerates.

For PVP functionalization, adapting the procedure by Abdulwahab et al.39 Raman spectroscopy (Figure 4) analysis also suggests the formation of iron oxide by-phases indicated by strong shoulders above 700 cm-1, which do not refer to the partly inverted spinel structure of MgFe2O4 as discussed before.68,69 By adding PVP to the presented microwave-assisted synthesis of MgFe2O4 nanocrystals, it was possible to prepare phase-pure, water-dispersible MgFe2O4 nanoparticles in a onestep microwave synthesis indicated by the similar Raman vibration pattern compared to the synthesis of bare MgFe2O4 (2.1.1). The crystallite size was determined to 5.5 and 6.2 nm, respectively. TEM images of the resulting nanoparticles are given in Figure S8. Moreover, both approaches result in stable dispersions of PVP-stabilized MgFe2O4 nanocrystals in water (Figure S9). The size distribution is, however, broader and shifted to larger hydrodynamic diameters compared to the OA/OLA-stabilized impure MgFe2O4 nanocrystals (compare Figure S7), which might be due to particle aggregation or PVP layer(s) formed around the particles. Physisorption measurements revealed a BET surface area of only 50 m2/g compared to 140 m2/g for non-functionalized nanoparticles. This confirms the assumption of PVP-coated nanoparticle agglomerates. Nevertheless,

Figure 5. DLS results of stabilized MgFe2O4 nanocrystal dispersions in different media; OLA/OA: post-synthetically functionalized nanoparticles with OLA/OA dispersed in toluene, PVP: insitu PVP-coated nanoparticles dispersed in water, BETA: Betaine hydrochloride stabilized nanoparticles dispersed in water, CIT: citrate stabilized nanoparticles dispersed in water.

In order to obtain agglomerate-free dispersions of stabilized MgFe2O4 nanoparticles in water for photocatalytic applications, nanoparticles were transferred by two approaches from

S7

ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

non-polar to polar media. On one hand, the OA/OLA ligands were exchanged with citric acid, while in a second approach the non-functionalized, as-synthesized nanoparticles were modified with betaine hydrochloride. The surface area citric acid treated nanoparticles (5.1 m²/g) is quite low, as the electrostatic stabilization of the colloidal solution breaks down during drying, which leads to heavy agglomeration. Both approaches resulted in highly dispersible, nearly agglomerate-free MgFe2O4 nanoparticles in water as proven by DLS. The zeta potential for betaine hydrochloride modified nanoparticles (8 mg/L) was 37 mV indicating the presence of positively charged surface groups (betaine cations) and the isoelectric point was found to be at pH = 7.9 (see Table S2). Figure 5 gives the size distribution for citric acid and betaine hydrochloride modified MgFe2O4 nanoparticles dispersions in water. In both cases, the size distribution is much narrower compared to OA/OLA-MgFe2O4 dispersions in toluene. Moreover, the peak hydrodynamic diameters are only slightly larger than the determined crystallite sizes of the microwave-derived MgFe2O4 nanoparticles prepared at 275 °C, which is due to the ligand and the hydration shell. These results indicate agglomerate-free dispersions. Especially the citric acid modified MgFe2O4 nanoparticle dispersion shows very narrow size distribution, with the peak of the hydrodynamic diameter at 6-7 nm, a zeta potential of -20 mV and an isoelectric point at pH = 10.6 (see Table S2). Figure S10 shows the long time stability test for the latter dispersion and for betaine-stabilized MgFe2O4 nanoparticles. Even after 4 weeks, virtually no changes in the hydrodynamic diameters were observed. Scheme 1 summarizes the different possibilities to achieve stable dispersions of MgFe2O4 nanoparticles.

Scheme 1. Pathways to stabile MgFe2O4 nanoparticle colloids developed in this study.

The changes in the ligand sphere upon phase transfer of the MgFe2O4 nanoparticles can be followed in the FTIR spectra in Figure 6. The FTIR spectrum of the unwashed particles (“UW” in Figure 6) shows a broad resonant band between 3650 cm-1 and 3110 cm-1 attributed to –O-H stretching vibrations resulting from rac-1-phenylethanol bond to the particle surface.74 Also, residual Mg(acac)2 and Fe(acac)3 on the particle surface can be identified contributing to the –C-H stretching vibration bands between 3005 cm-1 and 2825 cm-1.75 Moreover, the strong band at 1580 cm-1 indicates –C=O stretching vibrations of carboxyl groups being present in the precursor molecules.75 Additionally, the deformation vibration of aliphatic –C-H band is detectable at and 1420 cm-1.75 The broad vibration band having its peak at 586 cm-1 as well as the vibration band at 431 cm-1 can be assigned to solid MgFe2O4 as the shape and wavenumber region match well with values

Page 8 of 15

for Fe-O stretching vibrations on tetrahedral and octahedral sites, respectively, reported for magnesium ferrite in literature.5,14,21,70 A similar vibration pattern is obtained for washed, but not functionalized MgFe2O4 nanoparticles (“NF” in Figure 6). Here, the bands resulting from acetylacetonates are less intense indicating a partly removal of the adsorbed precursor species. Additionally, no aromatic bands of rac-1-phenylethanol are detected, which underlines the removal of residual reaction solvent. Instead, –O H stretching vibration in the region between 3600 cm-1 and 3050 cm-1 as well as aliphatic –C-H stretching vibrations between 3000 cm1 and 2800 cm-1 can be ascribed to adsorbed ethanol, which was used during the workup. A –C-O stretching vibration at 1100 cm-1 is also due to adsorbed ethanol. The shape and intensity of these bands matches well with reference data of ethanol.76 Ligand stripping of OLA/OA-MgFe2O4 nanocrystals with Et3OBF4 can remove all of the surface ligands, e.g. OLA/OA or carboxylate residues from the synthesis. The respective IR spectrum (STR in Figure 6) shows a broad band for –O-H stretching vibrations in the region between 3600 cm1 and 3050 cm-1 as well as –C-H stretching vibrations between 3000 cm-1 and 2800 cm-1, which can be ascribed to ethanol adsorbed on the particle surface. A weak peak at 1079 cm-1 can be assigned to BF4- anion from the stripping agent.77,78 Furthermore, all bands reported for mixed magnesium ferrites in literature are observable.21

Figure 6. FTIR spectra of as-synthesized, functionalized and ligand-stripped MgFe2O4 nanocrystals; STR: stripped, NF: nonfunctionalized (washed), UW: unwashed, OLA/OA: postsynthetically functionalized with OLA/OA, PVP: in-situ PVPcoated, BETA: Betaine hydrochloride stabilized, CIT: citrate stabilized.

While for the unwashed and washed as-synthesized MgFe2O4 nanoparticles the bands for remaining unreacted precursor molecules (carboxylate vibrations) are visible, those nearly disappear after OA/OLA functionalization (see “OLA/OA” in Figure 6). In contrast, bands for asymmetric and symmetric – C-H bond stretching vibrations of aliphatic and olefinic bonds become apparent with peaks at 2919 cm-1 and 2850 cm-1 due to the long alkyl chains of OA and OLA.79 Moreover, bands for stretching vibrations –O-H groups of oleic acid and adsorbed water around 3420 cm-1 can be detected.79 At 1556 cm-1 and 1416 cm-1 two bands for asymmetric and symmetric stretching vibrations of –COO- are observable indicating an absorption of deprotonated oleic acid on the nanoparticle surface.79 The

S8

ACS Paragon Plus Environment

Page 9 of 15

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

asymmetric deformation vibration of the –NH3+ group of oleylamine appears at 1620 cm-1.79 After phase transfer with betaine hydrochloride, the vibration bands assigned to –C-H bonds nearly disappear, which is a clear indication for the ligand exchange (see “BETA” in Figure 6). A small signal remains for the methyl groups of betaine at 2940 cm-1.63 Moreover, a strong signal for –C=O stretching vibration of the carboxylic acid group of betaine is apparent at 1623 cm-1. The –C-N bending vibrations can be observed at 1396 cm-1.63 The citrate-stabilized nanoparticles show an FTIR vibration pattern matching well with reference data of iron(III) citrate.80 This indicates that not citric acid but citrate is bound directly to iron surface ions due to alkaline conditions during functionalization. This would be in agreement with the zeta potential value of -19.5 mV indicating negatively charged surface groups (citrate anions). Furthermore, the vibration band at 3426 cm-1 for –O-H stretching vibrations results from adsorbed water. Vibrations at 1617 cm-1 and 1388 cm-1 for – COOM stretching vibrations of the carboxylate group are very strong.59 At 1069 cm-1 –C-O stretching vibration signals appear.59 Also, weak signals are visible in the regions between 2900 cm-1 and 2500 cm-1 due to the –C-H stretching and deformation vibrations, respectively. Bands at 907 cm-1 and 850 cm-1 indicate the deformation vibrations of the carboxylate group. Also, the vibration bands for solid MgFe2O4 are detectable with maxima at 598 cm-1 for Fe-O stretching vibration in the tetrahedral site and 434 cm-1 for Fe-O stretching vibration in the octahedral site, respectively.21

stripping, TG mass loss is only 8-9 %, and MS traces indicate hardly any HF+, CO2+ and CH3C=O+ fragments. The mass loss derives from water/solvent desorption and Et3OBF4 decomposition, which is underlined by TG-MS data (Figure S11). In addition to MS traces for H2O+, CO2+ and CH3C=O+ fragments, postsynthetically OA/OLA-modified MgFe2O4 nanocrystals also show MS traces of NO+ (m/z = 30) and NO2+ (m/z = 46) due to the decomposition of the amine groups of OLA (Figure S11). Overall, in the latter case the mass loss is only about 14 %, which is less than for the washed as-synthesized MgFe2O4 nanocrystals, indicating that the OA/OLA functionalization also removes some residual precursors during OA/OLA capping. This conclusion is also supported by the FTIR spectra (Figure 6) showing a reduced intensity for carboxylic groups. Similar traces can be observed for the PVP-coated nanocrystals showing a mass loss of around 22 %, and typical nitrogen-indicating fragments of NO2+ (m/z = 46) and NO+ (m/z = 30) from the decomposition of the lactone groups of PVP between 300-500 °C (Figure S11).

The FTIR spectrum of PVP-coated MgFe2O4 nanoparticles (“PVP” in Figure 6) shows typical bands for aliphatic –C-H stretching vibrations appearing between 3000 cm-1 and 2820 cm-1. The –C=O stretching vibration of the lactamcarboxylic group appears at 1659 cm-1 and 1597 cm-1.81 Bands for aliphatic –C-H bending vibrations from PVP can be detected at 1424 cm-1 and 1351 cm-1.81 The–C-N stretching vibrations of the lactam ring manifest in a band at 1287 cm-1.81 In addition, the characteristic bands for solid MgFe2O4 described earlier are detected in all surface-modified samples, too.21 Thermogravimetric investigations reveal the amounts and chemical nature of surface capping by different moieties (Figure 7). In case of the as-synthesized nanoparticles, without washing the mass loss is around 30 %, indicating a lot of unreacted precursors (which is in agreement with FTIR data from Figure 6) and adsorbed solvent molecules on the surface. After washing, the mass loss is reduced to 17 %, which is still too much for adsorbed solvent molecules only. In agreement with Figure 6, unreacted precursors indicated by residual carboxylic acid groups (from the acetylacetonate precursors) are adsorbed on the surface of the nanocrystals, decomposing during temperature treatment at around 300 °C according to TG data. MS traces for common decomposition products are given in Figure S11. For the unwashed nanocrystals, fragments typical for H2O+ (m/z = 18), CO2+ (m/z = 44) and CH3C=O+ (m/z = 43) can be detected during decomposition of surface pre70 cursors. Furthermore, MS trace of HF+ (m/z = 20) derives from decomposition products of the PTFE sealing cap. The same signals are detectable for washed, non-functionalized nanoparticles, too, with remarkably less intensity indicating a partly removal of the species adsorbed to the particles’ surface (Figure S11). After

Figure 7. Thermogravimetric analysis for different stabilized and non-stabilized MgFe2O4 nanoparticles; STR: stripped, NF: nonfunctionalized (washed), UW: unwashed, OLA/OA: postsynthetically functionalized with OLA/OA, PVP: in-situ PVPcoated, BETA: Betaine hydrochloride functionalization, CIT: citrate stabilized). After phase transfer with betaine hydrochloride, the TG analysis gives a mass loss of 20 %, due to the relatively large amount of heteroatoms and the heavy chlorine fragments. TG-MS data presented in Figure S11 show decomposition fragments for H2O+ (m/z = 18), NO+ (m/z = 30), Cl- (m/z = 35), CH3C=O+ (m/z = 43), CO2+ (m/z = 44), and NO2+ (m/z = 46), which match the functional groups being present in betaine hydrochloride. Citratestabilized MgFe2O4 nanocrystals show an even larger mass loss of

S9

ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

41 %. This is because of the large amount of oxygen from the surface capping due to the oxygen-to-carbon ratio of 7:6 in citric acid being removed at elevated temperatures. MS traces of H2O+ (m/z = 18), CH3C=O+ (m/z = 43), CO2+ (m/z = 44) and NO2+ (m/z = 46) were detected.

Figure 8 shows transmission electron microscopy (TEM) images of MgFe2O4 nanoparticles synthesized after standard protocol (2.1.1), and betaine hydrochloride functionalized and citrate functionalized nanoparticles. It is apparent, that neither shape nor size of the phase-pure nanoparticles is affected by post-synthetic treatment. Further investigation using selected electron diffraction (SAED) are in good agreement with PXRD data reported previously and literature values.24 Further SAED images of MgFe2O4 nanoparticle samples described in this article can be found in Figure S12. A high-resolution TEM image of betaine hydrochloride stabilized MgFe2O4 nanocrystals is

Page 10 of 15

tocatalytic performance. 62 % degradation can be achieved after 1.5 h with citrate-stabilized MgFe2O4. UV-Vis spectra of the as-prepared reaction solution, the dark equilibration adsorption as well as the time-dependent absorption spectra of the irradiated MgFe2O4/RhB dispersions are given in Figure S14. Maximum values after equilibration in the dark were normed to be crel = 1, and all following spectra were evaluated relative to this. The Quantitative investigations of the reaction kinetics were done by fitting the experimental data with a first-order kinetics model. The resulting kinetic constants and degradation rates are given in Table S3. The kinetic constant for non-functionalized MgFe2O4 nanoparticles is already 1.5 times larger than the one for ligand-stripped nanoparticles, which might be due to increased agglomeration of ligand-stripped nanoparticles in aqueous solution. Therefore, the particles’ surface is less accessible for the dye molecules resulting in inhibited degradation kinetics.

Figure 8. Bright field TEM and corresponding SAED images of non-functionalized (A-C), betaine hydrochloride functionalized (D-E) and citrate functionalized (G-I) nanoparticles.

given in Figure S13, indicating that the samples are still single crystals after post-synthetic functionalization.

3.4 Investigation of photocatalytic properties of MgFe2O4 nanocrystals Photocatalytic experiments with MgFe2O4 nanocrystals (assynthesized, citrate-functionalized, betaine-functionalized, ligand-stripped) have been performed for the degradation of model pollutant Rhodamine B (RhB). Figure 9 gives the concentration of RhB over time under irradiation of MgFe2O4/RhB dispersions in water. Non-functionalized and ligand-stripped MgFe2O4 nanocrystals show only little photocatalytic activity. 24 % degradation after 2 hours can be achieved with non-functionalized MgFe2O4 nanocrystals. In contrast, only 10 % degradation can be achieved with ligandstripped nanocrystals. On the other hand, betaine and citratefunctionalized MgFe2O4 nanocrystals show much better pho-

Figure 9. Photocatalytic activities for the degradation of RhB under solar light irradiation (AM 1.5G) in the presence of MgFe2O4 nanocrystals (above) and first order kinetic evaluation to calculate rate constants (below); STR: stripped, NF: nonfunctionalized (washed), BETA: Betaine hydrochloride stabilized, CIT: citrate stabilized.

Surface-functionalization with betaine hydrochloride increases the reaction constant by factor 2.5. Surface treatment with citric acid yet leads to a 5 times higher kinetic constant compared to non-functionalized nanoparticles. This underlines

S10

ACS Paragon Plus Environment

Page 11 of 15

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

the positive influence of de-agglomeration by surface functionalization. Furthermore, kinetic constants for surfacefunctionalized nanoparticles are in the range or even exceed the value reported for MgFe2O4 nanoparticles with a crystallite size of 50-80 nm (k = 0.372 h-1) reported by Yuan et al.82 Reason for the improved degradation rates of the stabilized nanocrystals is the enhanced dispersion of the nanocrystals in water providing improved access of the model pollutant to the nanoparticle surface. Moreover, the citrate-stabilization seems to help adsorbing RhB electrostatically to the surface of MgFe2O4, better compared to the betaine stabilization, resulting in improved photocatalytic degradation. It was already reported in literature that citrate ions can improve the adsorption of cations at high concentration of surface-bond citrate ions.83 The maximum reaction constant for RhB degradation in the presence of citrate-stabilized nanoparticles (0.638 h-1) is already a promising value for a single compound system. Ye et al. reported a maximum kinetic constant of 2.16 h-1 under visible light for a α-Fe2O3/g-C3N4 photocatalyst system with an improvement of factor 1.8 in comparison to pure g-C3N4.84 The higher values were attributed to improved charge carrier separation in the system due to heterojunction formation as a synergistic effect in the hybrid system. A comparable value of 1.26 h-1 for α-Fe2O3/g-C3N4 nanocomposite photocatalyst have been reported by Liu et al.85 The authors found considerably smaller rate constants for single phase photocatalysts g-C3N4 (k = 0.459 h-1) and α-Fe2O3 (k = 0.020 h-1). Li and co-workers reported similar kinetic constant for α-Fe2O3 (k = 0.044 h-1).86 These values for α-Fe2O3 are already by factors 1.9 to 4.1 smaller than our calculated kinetic constant for ligand-stripped nanoparticles. Thus, even higher rate constants for MgFe2O4 hybrid systems might be possible. More photocatalytic experiments for stabilized MgFe2O4 nanoparticles will be performed in the future, including other model compound degradation, as well as photocatalytic hydrogen evolution and water splitting experiments.

4. Conclusions Monocrystalline MgFe2O4 nanoparticles were prepared in a microwave-assisted water-free sol-gel process. The nanoparticles are phase-pure directly after synthesis in 30 minutes. Postsynthetic calcination results in by-phase formation above 600 °C, which has been identified thoroughly. Dispersions of MgFe2O4 nanoparticles in non-polar or polar solvents can be tailored by stabilization approaches. Stable MgFe2O4 nanoparticle dispersions in water can be achieved in-situ by PVP addition to the microwave-assisted synthesis. For stable dispersions in toluene, post-synthetic surface functionalization with OA and OLA can be performed. Subsequent phase transfer with betaine hydrochloride or citric acid leads to highly stable and agglomerate-free MgFe2O4 nanocrystal dispersions in water with narrow size distribution below 10 nm. Photocatalytic experiments are presented, indicating that surface stabilization can positively influence compound adsorption and enhance pollutant degradation.

SAED images, calculated kinetic constants for RhB degradation. This material is available free of charge via the Internet at http://pubs.acs.org.

AUTHOR INFORMATION Corresponding Author * [email protected]

Author Contributions The manuscript was written through contributions of all authors. All authors have given approval to the final version of the manuscript.

ACKNOWLEDGMENT We would like to thank Hubert Woerner for TG-MS measurements, Brigitte Weinl-Boulakhrouf for IR measurements, Rüdiger Ellinghaus for N2 physisorption measurements, and Prof. Bernd M. Smarsly for his support (all Justus-Liebig-University Giessen). Further thanks go to Dr. Annette Kleppe and Dr. Michael Wharmby for their help during the measurements at I15, Diamond Light Source, Didcot. K.K. and R.M. acknowledge funding by the German Research Foundation DFG under the priority program SPP 1613, project MA 5392/5-1. R.M. and T.W. gratefully acknowledge funding in the Emmy-Noether program (MA 5392/3-1) of the German Research Foundation DFG. C.S. acknowledges financial support within the SIGNO project (03SHWB073) by the German Federal Ministry for Economic Affairs and Energy (BMWi). R.M., J.T. and A. B. gratefully acknowledge financial support from the AiF within the program for promoting the Industrial Collective Research (IGF) of the German Federal Ministry of Economic Affairs and Energy (BMWi), based on a resolution of the German Parliament (project "QuinoLight", 18904N1-3).

Funding Sources German Research Foundation DFG: MA 5392/5-1; MA 5392/3-1

German Federal Ministry for Economic Affairs and Energy BMWi: 03SHWB073 BMWi-IGF: 18904N1-3

REFERENCES (1)

(2) (3)

(4)

(5)

(6)

ASSOCIATED CONTENT Supporting Information. Tauc plots for band gap determination of MgFe2O4, FTIR spectra, crystallite sizes and corresponding band gaps, in situ PXRD patterns, TEM images, additional DLS measurements, pH values, isoelectric points, BET surface areas, MS traces for TG-MS investigations, RhB absorption spectra,

(7)

(8)

Gupta, A. K.; Gupta, M. Synthesis and Surface Engineering of Iron Oxide Nanoparticles for Biomedical Applications. Biomaterials 2005, 26, 3995–4021. Reiss, G.; Hütten, A. Magnetic Nanoparticles: Applications beyond Data Storage. Nat. Mater. 2005, 4, 725–726. Lu, A. H.; Salabas, E. L.; Schüth, F. Magnetic Nanoparticles: Synthesis, Protection, Functionalization, and Application. Angew. Chemie - Int. Ed. 2007, 46, 1222–1244. Tietze, R.; Zaloga, J.; Unterweger, H.; Lyer, S.; Friedrich, R. P.; Janko, C.; Pöttler, M.; Dürr, S.; Alexiou, C. Magnetic Nanoparticle-Based Drug Delivery for Cancer Therapy. Biochem. Biophys. Res. Commun. 2015, 468, 463–470. Yan, K.; Li, H.; Wang, X.; Yi, C.; Zhang, Q.; Xu, Z.; Xu, H.; Whittaker, A. K. Self-Assembled Magnetic Luminescent Hybrid Micelles Containing Rare Earth Eu for Dual-Modality MR and Optical Imaging. J. Mater. Chem. B 2014, 2, 546–555. Hoque, S. M.; Hossain, M. S.; Choudhury, S.; Akhter, S.; Hyder, F. Synthesis and Characterization of ZnFe2O4 Nanoparticles and Its Biomedical Applications. Mater. Lett. 2016, 162, 60–63. Liu, Y. L.; Liu, Z. M.; Yang, Y.; Yang, H. F.; Shen, G. L.; Yu, R. Q. Simple Synthesis of MgFe2O4 Nanoparticles as Gas Sensing Materials. Sensors Actuators, B Chem. 2005, 107, 600– 604. Zhou, X.; Liu, J.; Wang, C.; Sun, P.; Hu, X.; Li, X.; Shimanoe,

S11

ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(9)

(10)

(11)

(12)

(13)

(14)

(15)

(16)

(17) (18)

(19)

(20)

(21)

(22)

(23)

(24)

(25)

(26)

(27)

K.; Yamazoe, N.; Lu, G. Highly Sensitive Acetone Gas Sensor Based on Porous ZnFe2O4 Nanospheres. Sensors Actuators B Chem. 2015, 206, 577–583. Godbole, R. V.; Rao, P.; Alegaonkar, P. S.; Bhagwat, S. Influence of Fuel to Oxidizer Ratio on LPG Sensing Performance of MgFe2O4 Nanoparticles. Mater. Chem. Phys. 2015, 161, 135–141. Zhang, J.; Song, J.-M.; Niu, H.-L.; Mao, C.-J.; Zhang, S.-Y.; Nanoparticles: Synthesis, Shen, Y.-H. ZnFe2O4 Characterization, and Enhanced Gas Sensing Property for Acetone. Sensors Actuators B Chem. 2015, 221, 55–62. Matsumoto, Y. Energy Positions of Oxide Semiconductors and Photocatalysis with Iron Complex Oxides. J. Solid State Chem. 1996, 126, 227–234. Casbeer, E.; Sharma, V. K.; Li, X.-Z. Synthesis and Photocatalytic Activity of Ferrites under Visible Light: A Review. Sep. Purif. Technol. 2012, 87, 1–14. Roonasi, P.; Nezhad, A. Y. A Comparative Study of a Series of Ferrite Nanoparticles as Heterogeneous Catalysts for Phenol Removal at Neutral pH. Mater. Chem. Phys. 2016, 172, 143– 149. Hoque, S. M.; Hakim, M. A.; Mamun, A.; Akhter, S.; Hasan, M. T.; Paul, D. P.; Chattopadhayay, K. Study of the Bulk Magnetic and Electrical Properties of MgFe2O4 Synthesized by Chemical Method. Mater. Sci. Appl. 2011, 2, 1564–1571. Pan, Y.; Zhang, Y.; Wei, X.; Yuan, C.; Yin, J.; Cao, D.; Wang, G. MgFe2O4 Nanoparticles as Anode Materials for Lithium-Ion Batteries. Electrochim. Acta 2013, 109, 89–94. Permien, S.; Indris, S.; Scheuermann, M.; Schürmann, U.; Mereacre, V.; Powell, A. K.; Kienle, L.; Bensch, W. Is There a Universal Reaction Mechanism of Li Insertion into Oxidic Spinels: A Case Study Using MgFe2O4. J. Mater. Chem. A 2015, 3, 1549–1561. Goldman, A. Modern Ferrite Technology; Springer: Pittsburgh, 2006. Song, Q.; Zhang, Z. J. Shape Control and Associated Magnetic Properties of Spinel Cobalt Ferrite Nanocrystals. J. Am. Chem. Soc. 2004, 126, 6164–6168. Dillert, R.; Taffa, D. H.; Wark, M.; Bredow, T.; Bahnemann, D. W. Research Update: Photoelectrochemical Water Splitting and Photocatalytic Hydrogen Production Using Ferrites (MFe2O4) under Visible Light Irradiation. APL Mater. 2015, 3, 104001:115. Taffa, D. H.; Dillert, R.; Ulpe, A. C.; Bauerfeind, K.; Bredow, T.; Bahnemann, D.; Wark, M. Photoelectrochemical and Theoretical Investigations of Spinel Type Ferrites (MxFe3−xO4) for Water Splitting: A Mini-Review. J. Photonics Energy 2016, 7, 1–24. Pradeep, A.; Chandrasekaran, G. FTIR Study of Ni, Cu and Zn Substituted Nano-Particles of MgFe2O4. Mater. Lett. 2006, 60 (2006), 371–374. Yang, H.; Yan, J.; Lu, Z.; Cheng, X.; Tang, Y. Photocatalytic Activity Evaluation of Tetragonal CuFe2O4 Nanoparticles for the H2 Evolution under Visible Light Irradiation. J. Alloys Compd. 2009, 476, 715–719. Hussein, S. I.; Elkady, A. S.; Rashad, M. M.; Mostafa, A. G.; Megahid, R. M. Structural and Magnetic Properties of Magnesium Ferrite Nanoparticles Prepared via EDTA-Based Sol–gel Reaction. J. Magn. Magn. Mater. 2015, 379, 9–15. Chen, D.; Zhang, Y.; Tu, C. Preparation of High Saturation Magnetic MgFe2O4 Nanoparticles by Microwave-Assisted Ball Milling. Mater. Lett. 2012, 82, 10–12. Šepelák, V.; Feldhoff, A.; Heitjans, P.; Krumeich, F.; Menzel, D.; Litterst, F. J.; Bergmann, I.; Becker, K. D. Nonequilibrium Cation Distribution, Canted Spin Arrangement, and Enhanced Magnetization in Nanosized MgFe2O4 Prepared by a One-Step Mechanochemical Route. Chem. Mater. 2006, 18, 3057–3067. Berchmans, L. J.; Myndyk, M.; Da Silva, K. L.; Feldhoff, A.; Šubrt, J.; Heitjans, P.; Becker, K. D.; Šepelák, V. A Rapid OneStep Mechanosynthesis and Characterization of Nanocrystalline CaFe2O4 with Orthorhombic Structure. J. Alloys Compd. 2010, 500, 68–73. Verma, S.; Joy, P. A.; Khollam, Y. B.; Potdar, H. S.; Deshpande, S. B. Synthesis of Nanosized MgFe2O4 Powders by Microwave Hydrothermal Method. Mater. Lett. 2004, 58, 1092–

(28)

(29)

(30)

(31)

(32)

(33)

(34)

(35)

(36)

(37)

(38)

(39)

(40)

(41)

(42)

(43)

(44) (45)

(46)

Page 12 of 15

1095. Sasaki, T.; Ohara, S.; Naka, T.; Vejpravova, J.; Sechovsky, V.; Umetsu, M.; Takami, S.; Jeyadevan, B.; Adschiri, T. Continuous Synthesis of Fine MgFe2O4 Nanoparticles by Supercritical Hydrothermal Reaction. J. Supercrit. Fluids 2010, 53, 92–94. Karunakaran, C.; SakthiRaadha, S.; Gomathisankar, P.; Vinayagamoorthy, P. Nanostructures and Optical, Electrical, Magnetic, and Photocatalytic Properties of Hydrothermally and Sonochemically Prepared CuFe2O4/SnO2. RSC Adv. 2013, 3, 16728–16738. Ilhan, S.; Izotova, S. G.; Komlev, A. A. Synthesis and Characterization of MgFe2O4 Nanoparticles Prepared by Hydrothermal Decomposition of Co-Precipitated Magnesium and Iron Hydroxides. Ceram. Int. 2015, 41, 577–585. Munjal, S.; Khare, N.; Nehate, C.; Koul, V. Water Dispersible CoFe2O4 Nanoparticles with Improved Colloidal Stability for Biomedical Applications. J. Magn. Magn. Mater. 2016, 404, 166–169. Diodati, S.; Pandolfo, L.; Caneschi, A.; Gialanella, S.; Gross, S. Green and Low Temperature Synthesis of Nanocrystalline Transition Metal Ferrites by Simple Wet Chemistry Routes. Nano Res. 2014, 7, 1027–1042. Antonello, A.; Jakob, G.; Dolcet, P.; Momper, R.; Kokkinopoulou, M.; Landfester, K.; Muñoz-Espí, R.; Gross, S. Synergy of Miniemulsion and Solvothermal Conditions for the Low-Temperature Crystallization of Magnetic Nanostructured Transition-Metal Ferrites. Chem. Mater. 2017, 29 (3), 985–997. de Vicente, J.; Delgado, A. V.; Plaza, R. C.; Durán, J. D. G.; González-Caballero, F. Stability of Cobalt Ferrite Colloidal Particles. Effect of pH and Applied Magnetic Fields. Langmuir 2000, 16, 7954–7961. Diodati, S.; Nodari, L.; Natile, M. M.; Caneschi, A.; De Julián Fernández, C.; Hoffmann, C.; Kaskel, S.; Lieb, A.; Di Noto, V.; Mascotto, S.; et al. Coprecipitation of Oxalates: An Easy and Reproducible Wet-Chemistry Synthesis Route for TransitionMetal Ferrites. Eur. J. Inorg. Chem. 2014, 5, 875–887. Dom, R.; Subasri, R.; Hebalkar, N. Y.; Chary, A. S.; Borse, P. H. Synthesis of a Hydrogen Producing Nanocrystalline ZnFe2O4 Visible Light Photocatalyst Using a Rapid Microwave Irradiation Method. RSC Adv. 2012, 2, 12782–12791. Suchomski, C.; Breitung, B.; Witte, R.; Knapp, M.; Bauer, S.; Baumbach, T.; Reitz, C.; Brezesinski, T. Microwave Synthesis of High-Quality and Uniform 4 Nm ZnFe2O4 Nanocrystals for Application in Energy Storage and Nanomagnetics. Beilstein J. Nanotechnol. 2016, 7, 1350–1360. Sun, S.; Zeng, H.; Robinson, D. B.; Raoux, S.; Rice, P. M.; Wang, S. X.; Li, G. Monodisperse MFe2O4 (M = Fe, Co, Mn) Nanoparticles. J. Am. Chem. Soc. 2004, 126, 273–279. Abdulwahab, K. O.; Malik, M. A.; O’Brien, P.; Timco, G. A. A Direct Synthesis of Water Soluble Monodisperse Cobalt and Manganese Ferrite Nanoparticles from Iron Based Pivalate Clusters by the Hot Injection Thermolysis Method. Mater. Sci. Semicond. Process. 2014, 27, 303–308. Xu, Y.; Sherwood, J.; Qin, Y.; Holler, R. A.; Bao, Y. A General Approach to the Synthesis and Detailed Characterization of Magnetic Ferrite Nanocubes. Nanoscale 2015, 7, 12641-12649. Groman, E. V.; Bouchard, J. C.; Reinhardt, C. P.; Vaccaro, D. E. Ultrasmall Mixed Ferrite Colloids as Multidimensional Magnetic Resonance Imaging, Cell Labeling, and Cell Sorting Agents. Bioconjug. Chem. 2007, 18, 1763–1771. Arndt, D.; Zielasek, V.; Dreher, W.; Bäumer, M. Ethylene Diamine-Assisted Synthesis of Iron Oxide Nanoparticles in High-Boiling Polyolys. J. Colloid Interface Sci. 2014, 417, 188– 198. Ashwini, K.; Anantha Raju, K. S.; Adinarayan Reddy, P.; Sharma, S. C.; Nagabhushana, H.; Raghava Reddy, K. SurfaceModified Ferrite Nanoparticles as Magnetic Resonance Imaging T2 Contrast Agents. 2015, 1, 1–6. Raj, K.; Moskowitz, B.; Casciari, R. Advances in Ferrofluid Technology. J. Magn. Magn. Mater. 1995, 149, 174–180. Odenbach, S. Colloidal Magnetic Fluids : Basics, Development and Application of Ferrofluids; Springer: Berlin Heidelberg, 2009. Lattuada, M.; Hatton, T. A. Functionalization of Monodisperse Magnetic Nanoparticles. Langmuir 2007, 23, 2158–2168.

S12

ACS Paragon Plus Environment

Page 13 of 15 (47)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(48)

(49)

(50)

(51)

(52)

(53)

(54)

(55)

(56)

(57)

(58)

(59)

(60)

(61)

(62)

(63)

(64)

(65)

The Journal of Physical Chemistry Garzón-Manjón, A.; Solano, E.; de la Mata, M.; Guzmán, R.; Arbiol, J.; Puig, T.; Obradors, X.; Yáñez, R.; Ricart, S.; Ros, J. Induced Shape Controllability by Tailored Precursor Design in Thermal and Microwave-Assisted Synthesis of Fe3O4 Nanoparticles. J. Nanoparticle Res. 2015, 17, 291. Kharisov, B. I.; Dias, H. V. R.; Kharissova, O. V.; Vázquez, A.; Peña, Y.; Gómez, I. Solubilization, Dispersion and Stabilization of Magnetic Nanoparticles in Water and Non-Aqueous Solvents: Recent Trends. RSC Adv. 2014, 4, 45354–45381. Wu, W.; He, Q.; Jiang, C. Magnetic Iron Oxide Nanoparticles: Synthesis and Surface Functionalization Strategies. Nanoscale Res. Lett. 2008, 3, 397–415. Wei, W.; Zhaohui, W.; Taekyung, Y.; Changzhong, J.; WooSik, K. Recent Progress on Magnetic Iron Oxide Nanoparticles: Synthesis, Surface Functional Strategies and Biomedical Applications. Sci. Technol. Adv. Mater. 2015, 16, 23501 1-43. Chandradass, J.; Jadhav, A. H.; Kim, H. Surfactant Modified MgFe2O4 Nanopowders by Reverse Micelle Processing: Effect of Water to Surfactant Ratio (R) on the Particle Size and Magnetic Property. Appl. Surf. Sci. 2012, 258, 3315–3320. Goodarz Naseri, M.; Saion, E. B.; Kamali, A. An Overview on Nanocrystalline ZnFe2O4, MnFe2O4, and CoFe2O4 Synthesized by a Thermal Treatment Method. ISRN Nanotechnol. 2012, 2012 (Artcile ID 604241), 11 pages. Goodarz Naseri, M.; Ara, M. H. M.; Saion, E. B.; Shaari, A. H. Superparamagnetic Magnesium Ferrite Nanoparticles Fabricated by a Simple, Thermal-Treatment Method. J. Magn. Magn. Mater. 2014, 350, 141–147. Carta, D.; Casula, M. F.; Floris, P.; Falqui, A.; Mountjoy, G.; Boni, A.; Sangregorio, C.; Corrias, A. Synthesis and Microstructure of Manganese Ferrite Colloidal Nanocrystals. Phys. Chem. Chem. Phys. 2010, 12, 5074–5083. Maiti, D.; Saha, A.; Devi, P. S. SI Surface Modified Multifunctional ZnFe2O4 Nanoparticles for Hydrophobic and Hydrophilic Anti-Cancer Drug Molecule Loading. Phys. Chem. Chem. Phys. 2016, 18, 1439–1450. Zeng, H.; Rice, P. M.; Wang, S. X.; Sun, S. Shape-Controlled Synthesis and Shape-Induced Texture of MnFe2O4 Nanoparticles. J. Am. Chem. Soc. 2004, 126, 11458–11459. Cabrera, L. I.; Somoza, Á.; Marco, J. F.; Serna, C. J.; Puerto Morales, M. Synthesis and Surface Modification of Uniform MFe2O4 (M = Fe, Mn, and Co) Nanoparticles with Tunable Sizes and Functionalities. J. Nanoparticle Res. 2012, 14, 873:114. Singh, M.; Dosanjh, H. S.; Singh, H. Surface Modified Spinel Cobalt Ferrite Nanoparticles for Cationic Dye Removal: Kinetics and Thermodynamics Studies. J. Water Process Eng. 2016, 11, 152–161. Zhao, H.; Liu, R.; Zhang, Q.; Wang, Q. Effect of Surfactant Amount on the Morphology and Magnetic Properties of Monodisperse ZnFe2O4 Nanoparticles. Mater. Res. Bull. 2016, 75, 172–177. Chourpa, I.; Douziech-Eyrolles, L.; Ngaboni-Okassa, L.; Fouquenet, J.-F.; Cohen-Jonathan, S.; Soucé, M.; Marchais, H.; Dubois, P. Molecular Composition of Iron Oxide Nanoparticles, Precursors for Magnetic Drug Targeting, as Characterized by Confocal Raman Microspectroscopy. Analyst 2005, 130, 1395– 1403. Hanesch, M. Raman Spectroscopy of Iron Oxides and (Oxy)hydroxides at Low Laser Power and Possible Applications in Environmental Magnetic Studies. Geophys. J. Int. 2009, 177, 941–948. de Faria, D. L. a.; Silva, S. V.; de Oliveira, M. T. Raman Microspectroscopy of Some Iron Oxides and Oxyhydroxides. J. Raman Spectrosc. 1997, 28, 873–878. Patil, R. M.; Shete, P. B.; Thorat, N. D.; Otari, S. V.; Barick, K. C.; Prasad, A.; Ningthoujam, R. S.; Tiwale, B. M.; Pawar, S. H. Non-Aqueous to Aqueous Phase Transfer of Oleic Acid Coated Iron Oxide Nanoparticles for Hyperthermia Application. RSC Adv. 2014, 4, 4515–4522. Dong, A.; Ye, X.; Chen, J.; Kang, Y.; Gordon, T.; Kikkawa, J. M.; Murray, C. B. SI A Generalized Ligand-Exchange Strategy Enabling Sequential Surface Functionalization of Colloidal Nanocrystals. J. Am. Chem. Soc. 2010, 133, 998–1006. Batsanov, S. S.; Ruchkin, E. D.; Poroshina, I. A. Refractive

(66)

(67)

(68)

(69)

(70)

(71)

(72)

(73)

(74)

(75)

(76)

(77)

(78)

(79)

(80)

(81)

(82)

(83)

Indices of Solids; Springer: Pittsburgh, 2016. Wang, Z.; Lazor, P.; Saxena, S. K.; St, H.; Neill, C. O. High Pressure Raman Spectroscopy of Ferrite MgFe2O4. Mater. Res. Bull. 2002, 37, 1589–1602. Ishikawa, K.; Fujima, N.; Komura, H. First-Order Raman Scattering in MgO Microcrystals. J. Appl. Phys. 1985, 57, 973– 975. Nakagomi, F.; Silva, S. W.; Garg, V. K.; Oliveira, A. C.; Morais, P. C.; Jr, A. F. Journal of Solid State Chemistry Influence of the Mg-Content on the Cation Distribution in Cubic MgXFe3- XO4 Nanoparticles. J. Solid State Chem. 2009, 182, 2423–2429. Šepelák, V.; Bergmann, I.; Menzel, D.; Feldhoff, A.; Heitjans, P.; Litterst, F. J.; Becker, K. D. Magnetization Enhancement in Nanosized MgFe2O4 Prepared by Mechanosynthesis. J. Magn. Magn. Mater. 2007, 316, 764–767. Köferstein, R.; Walther, T.; Hesse, D.; Ebbinghaus, S. G. Preparation and Characterization of Nanosized Magnesium Ferrite Powders by a Starch-Gel Process and Corresponding Ceramics. J. Mater. Sci. 2013, 48, 6509–6518. Weller, T.; Specht, L.; Marschall, R. Single Crystal CsTaWO6 Nanoparticles for Photocatalytic Hydrogen Production. Nano Energy 2017, 31, 551–559. Cuenca, J. A.; Bugler, K.; Taylor, S.; Morgan, D.; Williams, P.; Bauer, J.; Porch, A. Study of the Magnetite to Maghemite Transition Using Microwave Permittivity and Permeability Measurements. J. Phys. Condens. Matter 2016, 28, 106002:110. Goto, Y. The Effect of Squeezing on the Phase Transformation and Magnetic Properties of γ-Fe2O3. Jpn. J. Appl. Phys. 1964, 3, 739–744. Shin-ya, K.; Sugeta, H.; Shin, S.; Hamada, Y.; Katsumoto, Y.; Ohno, K. Absolute Configuration and Conformation Analysis of 1-Phenylethanol by Matrix-Isolation Infrared and Vibrational Circular Dichroism Spectroscopy Combined with Density Functional Theory Calculation. J. Phys. Chem. A 2007, 111, 8598–8605. Tayyari, S. F.; Bakhshi, T.; Mahdizadeh, S. J.; Mehrani, S.; Sammelson, R. E. Structure and Vibrational Assignment of Magnesium Acetylacetonate: A Density Functional Theoretical Study. J. Mol. Struct. 2009, 938, 76–81. Doroshenko, I.; Pogorelov, V.; Sablinskas, V.; Doroshenko, I.; Pogorelov, V.; Sablinskas, V. Infrared Absorption Spectra of Monohydric Alcohols. Dataset Pap. Chem. 2013, 2013 (Article ID 329406), 6 pages. Rosen, E. L.; Buonsanti, R.; Llordes, A.; Sawvel, A. M.; Milliron, D. J.; Helms, B. A. Exceptionally Mild Reactive Stripping of Native Ligands from Nanocrystal Surfaces by Using Meerwein’s Salt. Angew. Chem. Int. Ed. 2012, 124, 703– 713. Dong, A.; Ye, X.; Chen, J.; Kang, Y.; Gordon, T.; Kikkawa, J. M.; Murray, C. B. A Generalized Ligand-Exchange Strategy Enabling Sequential Surface Functionalization of Colloidal Nanocrystals. J. Am. Chem. Soc. 2011, 133, 998–1006. Devaraj, M.; Saravanan, R.; Deivasigamani, R.; Gupta, V. K.; Gracia, F.; Jayadevan, S. Fabrication of Novel Shape Cu and Cu/Cu2O Nanoparticles Modified Electrode for the Determination of Dopamine and Paracetamol. J. Mol. Liq. 2016, 221, 930–941. Tsimbler, S. M.; Shevchenko, L. L.; Grigor’eva, V. V. The IR Absorption Spectra of the Tartrate and Citrate Complexes of Nickel, Cobalt, and Iron. J. Appl. Spectrosc. 1969, 11, 1096– 1101. Saravanan, P.; Venkata Ramana, G.; Srinivasa Rao, K.; Sreedhar, B.; Vinod, V. T. P.; Chandrasekaran, V. Structural and Magnetic Properties of Self-Assembled Sm–Co Spherical Aggregates. J. Magn. Magn. Mater. 2011, 323, 2083–2089. Yuan, X.; Wang, H.; Wu, Y.; Chen, X.; Zeng, G.; Leng, L.; Zhang, C. A Novel SnS2-MgFe2O4/reduced Graphene Oxide Flower-like Photocatalyst: Solvothermal Synthesis, Characterization and Improved Visible-Light Photocatalytic Activity. Catal. Commun. 2015, 61, 62–66. Redden, G.; Bargar, J.; Bencheikh-Latmani, R. Citrate Enhanced Uranyl Adsorption on Goethite: An EXAFS Analysis. J. Colloid Interface Sci. 2001, 244, 211–219.

S13

ACS Paragon Plus Environment

The Journal of Physical Chemistry (84)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(85)

Ye, S.; Qiu, L.-G.; Yuan, Y.-P.; Zhu, Y.-J.; Xia, J.; Zhu, J.-F. Facile Fabrication of Magnetically Separable Graphitic Carbon Nitride Photocatalysts with Enhanced Photocatalytic Activity under Visible Light. J. Mater. Chem. A 2013, 1, 3008–3015. Liu, X.; Jin, A.; Jia, Y.; Jiang, J.; Hu, N.; Chen, X. Facile Synthesis and Enhanced Visible-Light Photocatalytic Activity of Graphitic Carbon Nitride Decorated with Ultrafine Fe2O3

(86)

Page 14 of 15

Nanoparticles. RSC Adv. 2015, 5, 92033–92041. Li, S.; Hu, S.; Zhang, J.; Jiang, W.; Liu, J. Facile Synthesis of Fe2O3 Nanoparticles Anchored on Bi2MoO6 Microflowers with Improved Visible Light Photocatalytic Activity. J. Colloid Interface Sci. 2017, 497, 93–101.

S14

ACS Paragon Plus Environment

Page 15 of 15

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

TOC Graphic

ACS Paragon Plus Environment S15