MnO and NiO nanoparticles: synthesis and magnetic properties

Similar documents
Multiply twinned Pt Pd nanoicosahedrons as highly active electrocatalyst for methanol oxidation

Supplementary Information

Supporting Information

Shape-selective Synthesis and Facet-dependent Enhanced Electrocatalytic Activity and Durability of Monodisperse Sub-10 nm Pt-Pd Tetrahedrons and Cubes

Supporting Information

Supporting Information. Synthesis and Upconversion Luminescence of BaY 2

Synthesis and Characterization of Iron-Oxide (Hematite) Nanocrystals. Z.H. Lee

Synthesis of Uniform Hollow Oxide Nanoparticles. through Nanoscale Acid Etching

Fabrication and characterization of poly (ethylene oxide) templated nickel oxide nanofibers for dye degradation

Use of Ionic Liquids in the Synthesis of Nanocrystals and Nanorods of Semiconducting Metal Chalcogenides

Supplementary Information for

Supporting Information:

Rapid, Efficient Phase Pure Synthesis of Ca 2 AlNO 3 Layered Double Hydroxide

Magnetic nanoparticle-supported proline as a recyclable and recoverable ligand for the CuI catalyzed arylation of nitrogen nucleophiles

Shape Assisted Fabrication of Fluorescent Cages of Squarate based Metal-Organic Coordination Frameworks

Supporting Information for: Emulsion-assisted synthesis of monodisperse binary metal nanoparticles

Synthesis of Colloidal Au-Cu 2 S Heterodimers via Chemically Triggered Phase Segregation of AuCu Nanoparticles

Room-temperature method for coating ZnS shell on semiconductor quantum dots

Shape Effect of Ag-Ni Binary Nanoparticles on Catalytic Hydrogenation Aided by Surface Plasmon

Pt-Ni alloyed nanocrystals with controlled archtectures for enhanced. methanol oxidation

Supporting Information

Facile synthesis of polymer and carbon spheres decorated with highly dispersed metal nanoparticles

Electronic supplementary information

Electronic Supplementary Information

Synthesis of 2 ) Structures by Small Molecule-Assisted Nucleation for Plasmon-Enhanced Photocatalytic Activity

Hidden Role of Anion Exchange Reactions in Nucleation of Colloidal Nanocrystals

enzymatic cascade system

Supporting Information. Solution-Based Growth of Monodisperse Cube-Like BaTiO 3 Colloidal Nanocrystals

Magnetic properties in a partially oxidized nanocomposite of Cu-CuCl

Synthesis of nano-sized anatase TiO 2 with reactive {001} facets using lamellar protonated titanate as precursor

Supporting Information Detailed Experiments Materials: All the reagents were analytical grate and used without further purification.

Supporting Information

Supporting Information

Supporting Information. Capping Nanoparticles with Graphene Quantum Dots for Enhanced Thermoelectric Performance

Supporting Information. for. Advanced Materials, adma Wiley-VCH 2006

Electronic Supplementary Information

Nanoporous TiO 2 Nanoparticle Assemblies with Mesoscale Morphologies: Nano-Cabbage versus Sea-Anemone

Electronic Supplementary Information (ESI)

Electronic Supplementary Information (ESI) Green synthesis of shape-defined anatase TiO 2 nanocrystals wholly exposed with {001} and {100} facets

Electronic supplementary information. A longwave optical ph sensor based on red upconversion

Supplementary Information for Self-assembled, monodispersed, flowerlike γ-alooh

Synthesis of Oxidized Graphene Anchored Porous. Manganese Sulfide Nanocrystal via the Nanoscale Kirkendall Effect. for supercapacitor

Electronic Supplementary Information. Facile Synthesis of Germanium-Graphene Nanocomposites. and Their Application as Anode Material for Lithium Ion

Supplementary Information

Electronic Supplementary Information (ESI) From metal-organic framework to hierarchical high surface-area hollow octahedral carbon cages

Strategic use of CuAlO 2 as a sustained release catalyst for production of hydrogen from methanol steam reforming

Room-temperature ferromagnetism in nanoparticles of Superconducting materials

Structural and magnetic properties of Ni doped CeO 2 nanoparticles

Supporting Information

A Highly Efficient Double-Hierarchical Sulfur Host for Advanced Lithium-Sulfur Batteries

Magnetization reversal and ferrimagnetism in Pr 1 x Nd x MnO 3

Electronic Supplementary Material. Methods. Synthesis of reference samples in Figure 1(b) Nano Res.

Dry-gel conversion synthesis of Cr-MIL-101 aided by grinding: High surface area high yield synthesis with minimum purification

Supplementary Information

Supporting information A Porous Zr-cluster-based Cationic Metal-Organic Framework for Highly Efficient Cr 2 O 7

Graphene composites containing chemically bonded metal oxides

Supporting Information

Effects of interaction of electron-donor and accepter molecules on the electronic structure of graphene

Supporting information. Enhanced photocatalytic degradation of methylene blue and adsorption of

Achiral CdSe quantum dots exhibit optical activity in the visible region upon post-synthetic ligand exchange with D- or L-cysteine

Supplementary Information

Supporting Information for

Supporting Information

Urchin-like Ni-P microstructures: A facile synthesis, properties. and application in the fast removal of heavy-metal ions

Supplementary Information

Supplementary Material for. Zinc Oxide-Black Phosphorus Composites for Ultrasensitive Nitrogen

Supporting Information:

SYNTHESIS OF CADMIUM SULFIDE NANOSTRUCTURES BY NOVEL PRECURSOR

Supporting Information

Tailoring the shapes of Fe x. Pt 100 x. nanoparticles. Home Search Collections Journals About Contact us My IOPscience

ABSTRACT I. INTRODUCTION II. BACKGROUND OF STUDY

Electronic Supplementary Information

Novel magnetic properties of graphene: Presence of both. ferromagnetic and antiferromagnetic features and other aspects

Supporting Information. Rh-doped Pt-Ni octahedral nanoparticles: understanding the correlation between elemental distribution, ORR and shape stability

Supplementary Information

Electronic Supplementary Information

Self-rearrangement of silicon nanoparticles. high-energy and long-life lithium-ion batteries

Growth of 2D sheets of a MOF on graphene surfaces to yield composites with novel gas adsorption characteristics

A soft-templated method to synthesize sintering-resistant Au/mesoporous-silica core-shell nanocatalysts with sub-5 nm single-core

Supporting Information. Zinc Ferrite Nanoparticles as MRI Contrast Agents. J. Ping Liu, b and Jinming Gao a,*

Supporting Information

Supplementary Data. Size-controlled synthesis of MIL-101(Cr) nanoparticles with. enhanced selectivity for CO 2 over N 2

Supporting Information. Size and shape evolution of highly magnetic iron nanoparticles from successive growth reactions

Justin G. Railsback, Aaron C. Johnston-Peck, Junwei Wang, and Joseph B. Tracy DOI: Sreeprasad T. S

Metal-organic frameworks (MOFs) as precursors towards TiO x /C. composites for photodegradation of organic dye

Enhanced photocurrent of ZnO nanorods array sensitized with graphene. quantum dots

Preparation of Nanofibrous Metal-Organic Framework Filters for. Efficient Air Pollution Control. Supporting Information

J. Am. Chem. Soc. 2009, 131,

Vertical Alignment of Reduced Graphene Oxide/Fe-oxide Hybrids Using the Magneto-Evaporation Method

The CdS and CdMnS nanocrystals have been characterized using UV-visible spectroscopy, TEM, FTIR, Particle Size Measurement and Photoluminiscence.

A Tunable Process: Catalytic Transformation of Renewable Furfural with. Aliphatic Alcohols in the Presence of Molecular Oxygen. Supporting Information

Electronic Supplementary Information

Electronic Supplementary Information

Supporting Information

Efficient Co-Fe layered double hydroxide photocatalysts for water oxidation under visible light

An inorganic-organic hybrid supramolecular nanotube as high-performance anode for lithium ion batteries

Controlled Synthesis of MFe 2 O 4 (M = Mn, Fe, Co, Ni and Zn) Nanoparticles and Their Magnetic Properties.

Supporting Information. for. A Sustainable Protocol for the Spontaneous Synthesis of Zinc-Glutamate. Wet Conditions

In situ formation of metal Cd x Zn 1-x S nanocrystals on graphene surface: A novel method to synthesis sulfide-graphene nanocomposites

Supporting information. Mechanical Properties of Microcrystalline Metal-Organic Frameworks. (MOFs) Measured by Bimodal Amplitude Modulated-Frequency

Transcription:

PAPER www.rsc.org/materials Journal of Materials Chemistry MnO and NiO nanoparticles: synthesis and magnetic properties Moumita Ghosh, ab Kanishka Biswas, ab A. Sundaresan a and C. N. R. Rao* ab Received 22nd August 2005, Accepted 22nd September 2005 First published as an Advance Article on the web 13th October 2005 DOI: 10.1039/b511920k Nanoparticles of MnO with average diameters in the 6 14 nm range have been prepared by the decomposition of manganese cupferronate in the presence of TOPO, under solvothermal conditions. Nanoparticles of NiO with average diameters in the 3 24 nm range have been prepared by the decomposition of nickel cupferronate or acetate under solvothermal conditions. The nanoparticles have been characterized by X-ray diffraction and transmission electron microscopy. Both MnO and NiO nanoparticles exhibit supermagnetism, accompanied by magnetic hysteresis below the blocking temperature (T B ). The T B increases with the increase in particle size in the case of NiO, and exhibits the reverse trend in the case of MnO. Introduction Nanoparticles of transition metal oxides have been investigated by several workers in the last few years. Besides their structural aspects, magnetic properties of the oxide nanoparticles are of particular interest. Thus, it would be of value to know if the nanoparticles of antiferromagnetic oxides generally show evidence for ferromagnetic interaction at low temperatures, a behaviour that has been reported by a few workers. 1,2 A recent study of CoO nanoparticles has shown that small particles of less than 16 nm diameter exhibit magnetic hysteresis below a blocking temperature of y10 K. 3 We were interested to investigate nanoparticles of the binary metal oxides MnO and NiO, which are antiferromagnetic with Neél temperatures of 122 K and 523 K respectively. 4 There have been a few reports in the literature on MnO nanoparticles. Lee et al. 5 prepared MnO nanoparticles by the decomposition of Mn 2 (CO) 8, but the product always contained Mn 2 O 3 as an impurity. These workers, however, found ferromagnetic behaviour in the particles with an average diameter of around 5 nm. Yin and O Brien 6 synthesized MnO nanocrystals capped with organic ligands by the decomposition of manganese acetate in a mixture of trioctylamine and oleic acid. These workers did not study the magnetic properties of the nanoparticles. Colloidal MnO nanoparticles, prepared by Seo et al. 7 by the decomposition of manganese acetylacetonate in oleylamine, showed divergence in the magnetization measured under zero-field-cooled (ZFC) and field-cooled (FC) conditions. Park et al. 8,9 prepared MnO nanospheres with diameters in the 5 40 nm range and nanorods of 7 10 nm diameter by the thermal decomposition of Mn-surfactant complexes in trioctylphosphine. These workers also observed divergence in the magnetization recorded under ZFC and FC conditions. We have prepared MnO nanoparticles by the decomposition of manganese cupferronate, Mn(cup) 2, in a Chemistry and Physics of Materials Unit and CSIR Centre of Excellence in Chemistry, Jawaharlal Nehru Centre for Advanced Scientific Research, Jakkur P. O., Bangalore-560064, India. E-mail: cnrrao@jncasr.ac.in; Fax: (+91) 80 22082760 b Solid State and Structural Chemistry Unit, Indian Institute of Science, Bangalore-560012, India toluene in the presence of trioctylphosphine oxide (TOPO) under solvothermal conditions and obtained nanoparticles with diameters in the 6 14 nm range. The particles so-prepared had an organic coating and enabled a study of magnetic properties without any surface oxidation. Small particles of NiO have been known to be superparamagnetic for some time. 10 Magnetic anomalies, such as low temperature hysteresis below a blocking temperature and divergence between the FC and ZFC magnetization data, on NiO nanoparticles have been reported. 2,11,12 The anomalies in the magnetic properties are attributed to uncompensated surface spins causing a change in the magnetic order in the nanoparticles. 1,12 NiO nanoparticles have been prepared by the decomposition of the hydroxide. 10,2,13 and by the oxidation of Ni nanoparticles 14 or by the decomposition of nickel oxalate. 15 We have prepared the nanoparticles of NiO of different diameters in the 3 24 nm range by the decomposition of nickel cupferronate, Ni(cup) 2, or nickel acetate in organic solvents under solvothermal conditions. The as-prepared particles have a protective organic coating. We have examined the magnetic properties of MnO and NiO nanoparticles, after characterizing them by X-ray diffraction and transmission electron microscopy. Experimental Manganese cupferronate, Mn(C 6 H 5 N 2 O 2 ) 2, or Mn(cup) 2 was synthesized as follows. 16 1 g (5.05 mmol) of MnCl 2?4H 2 O was dissolved in 100 ml of milli-q water and 1.56 g (10.05 mmol) of cupferron was dissolved in 50 ml of milli-q water by sonication. The two solutions were cooled at 0 uc and the cupferron solution added dropwise to the MnCl 2 solution under vigorous stirring. After a few minutes, a pale yellow coloured product was obtained. It was filtered, washed thoroughly with milli-q water and dried at room temperature overnight. MnO nanoparticles were prepared by the decomposition of Mn(cup) 2 at 325 uc in toluene in the presence of trioctylphosphine oxide (TOPO) as the capping agent. In a typical reaction, 0.2 g (0.60 mmol) of Mn(cup) 2 and 0.95 g (2.45 mmol) of TOPO were taken in 10 ml of toluene and sealed in a 106 J. Mater. Chem., 2006, 16, 106 111 This journal is ß The Royal Society of Chemistry 2006

stainless steel Swagelok autoclave of 20 ml capacity. The autoclave was kept inside a preheated tube-furnace at 325 uc for 2 h, and was then allowed to cool to room temperature. A brownish black solution was obtained as the product. To this solution, excess ethanol was added and kept overnight. Black coloured TOPO-capped MnO nanoparticles (14 nm diameter) obtained as a solid residue were washed several times with ethanol and acetone. To vary the particle size of MnO, the concentration of Mn(cup) 2, and/or the concentration of TOPO as well as the reaction time were varied, keeping the amount of toluene and the filling fraction of the autoclave constant. The nanoparticles of different diameters were obtained as follows: 6 nm, 0.005 g (0.015 mmol) of Mn(cup) 2, 0.045 g (0.012 mmol) of TOPO, 325 uc, 2 h 30 min; 8.5 nm, 0.025 g (0.075 mmol) of Mn(cup) 2, 0.18 g (0.046 mmol) of TOPO, 325 uc, 2 h; 10 nm, 0.01 g (0.03 mmol) of Mn(cup) 2, 0.09 g (0.024 mmol) of TOPO, 340 uc, 2 h; 14 nm, 0.2 g (0.60 mmol) of Mn(cup) 2, 0.95 g (2.45 mmol) of TOPO, 325 uc, 2 h. The nanoparticles could be readily redispersed in toluene by sonication. Nickel cupferronate, Ni(C 6 H 5 N 2 O 2 ) 2, or Ni(cup) 2, was prepared as follows. 2 g of Ni(OAc) 2?4H 2 O was dissolved in 150 ml of milli-q water (0.05 M), and 2.5 g of cupferron dissolved in 100 ml of milli-q water (0.16 M) by sonication. The two solutions were cooled to 0 uc and the cupferron solution added dropwise to the Ni(OAc) 2 solution under vigorous stirring. After a few minutes, the solution became turbid, indicating the formation of complex. It took another 30 min for completion of the reaction. Before filtration, the product was kept at 0 uc for 1 h. The product was filtered and washed with milli-q water and dried at room temperature. NiO nanoparticles were synthesized in toluene under solvothermal conditions as follows. In a typical reaction, 0.05 g (0.15 mmol) of Ni(cup) 2 were taken in 48 ml of toluene and subjected to sonication for 10 min. The resulting green coloured solution was sealed in a Teflon-lined stainless steel autoclave of 80 ml capacity (allowing 70% filling fraction) and the autoclave was placed inside a air-oven preheated at 240 uc for 8 h. The autoclave was allowed to cool to room temperature. A brownish coloured solid dispersed in toluene was obtained as the product. Addition of methanol to this toluene dispersion allowed the solid to settle down. The product was washed with methanol by sonication followed by centrifugation. The sample was dried in an air-oven at 50 uc for 1 h. The average diameter of the particle obtained under these preparative conditions was 3 nm. In order to prepare 7 nm NiO nanoparticles, we used 0.4 g (1.2 mmol) of Ni(cup) 2 in 48 ml of toluene and carried out the reaction for 24 h, keeping the temperature and filling fraction of the autoclave as constant. In addition to the NiO nanoparticles described above, trioctylphosphine oxide (TOPO)-capped NiO nanoparticles of 3 nm average diameter were prepared by taking 0.5 g (1.5 mmol) of Ni(cup) 2 in 1.5 mmolar toluene solution of TOPO. Fig. 1 XRD patterns of the MnO nanoparticles with average diameters of (a) 6, (b) 8.5, (c) 10 and (d) 14 nm along with the Rietveld fits. Difference patterns are shown below the observed patterns. Fig. 2 TEM images of (a) 6 nm and (b) 14 nm particles of MnO. Insets in (a) show the SAED pattern and the size distribution histogram. Insets in (b) show the single particle HREM image and the size distribution. This journal is ß The Royal Society of Chemistry 2006 J. Mater. Chem., 2006, 16, 106 111 107

We have also prepared NiO nanoparticles by the solvothermal decomposition of Ni(OAc) 2 using toluene as the solvent in a stainless steel Swagelok autoclave of 20 ml capacity. To prepare 12 nm NiO particles, 0.1 g (0.4 mmol) of Ni(OAc) 2?4H 2 O was taken in 10 ml of toluene and the reaction mixture sealed in a Swagelok autoclave of 20 ml capacity. The autoclave was kept inside a preheated tube furnace at 250 uc. The reaction was carried out for 3 h. A moss green coloured, toluene-insoluble solid residue was obtained as the product. NiO particles of 24 nm diameter were prepared by taking 0.5 g (2.01 mmol) of Ni(OAc) 2?4H 2 O in 10 ml toluene, and carrying out the reaction at 335 uc for 1 h. The nanoparticles were characterized by powder X-ray diffraction (XRD) using a Phillips X Pert diffractometer employing the Bragg Brentano configuration using CuKa radiation with the scan rate of 0.16u 2h min 21. The data were then rebinned into 0.05u steps. For transmission electron microscopy (TEM), toluene dispersions of the samples were dropped onto the holey carbon-coated Cu grids, and the grids were allowed to dry in the air. The grids were examined using a JEOL (JEM3010) microscope operating with an accelerating voltage of 300 kv. Thermogravimetric analysis (TGA) was carried out using a Mettler Toledo Thermal Analyser. Powder samples of the as-prepared nanoparticles were subjected to magnetic characterization using the VSM in PPMS (Physical Property Measurement System). Results and discussion MnO In Fig. 1, we show the X-ray diffraction patterns of MnO nanoparticles of four different sizes. Fig. 1 also gives the Rietveld profile fits along with the difference patterns obtained by using the Rietveld XND code. 17 The patterns could be indexed with the Fm3m space group (JCPDS no 07-0230). By making the use of the line widths from the Rietveld fits, the average particle sizes were estimated by using the Scherrer formula. The particle sizes are indicated in Fig. 1. The cubic lattice parameter of the 8.5 and 14 nm particles are 4.429 and 4.428 Å respectively. The TEM images of the MnO nanoparticles with average diameters of 6 and 14 nm are shown in Fig. 2(a) and (b), respectively. An indexed selected area electron diffraction (SAED) pattern of the 6 nm sample is shown in the inset of Fig. 2(a). The particle size distributions are shown as insets in Fig. 2. Most of the particles in Fig. 2(a) and (b) have diameters of 6 and 14 nm respectively, which compare well with the values calculated from the XRD Fig. 3 TGA curves of (a) 14 nm TOPO-capped MnO nanoparticles and (b) 3 nm uncapped NiO nanoparticles. The experiments were carried out in a nitrogen atmosphere at a heating rate of 10 uc min 21. Fig. 4 Magnetization vs. temperature curves of (a) 6 nm and (b) 14 nm MnO particles under FC and ZFC conditions using a magnetic field of 100 Oe. The insets show inverse susceptibility vs. temperature curves. 108 J. Mater. Chem., 2006, 16, 106 111 This journal is ß The Royal Society of Chemistry 2006

patterns. We show an HREM image of a 14 nm MnO nanoparticle as an inset in Fig. 2(b). The lattice spacing of 2.56 Å in the image corresponds to the interplanar distance between (111) planes. The images of the particles are not as good as one would desire because of the organic coating on the particles. The presence of such an organic coating is evidenced from the TGA curve shown in Fig. 3(a), for the 14 nm MnO nanoparticles. The organic coating is removed around 400 uc. The presence of the organic coating helps to avoid oxidation of the MnO nanoparticles. In Fig. 4(a) and (b), we show the magnetization data of two samples with average diameters of 6 and 14 nm respectively. The FC and ZFC data show divergence at low temperatures as indeed found in the case of other oxide nanoparticles. The magnetization in the FC data shows a marked increase at low temperatures, where a shoulder like feature is also evident. The ZFC data clearly show maxima corresponding to blocking temperatures (T B ) of the nanoparticles. The T B of the 6 nm particle is higher than that of the 14 nm particles indicating an inverse relationship. A similar dependence has been noticed earlier. 7 The insets in Fig. 4(a) and (b) represent the inverse susceptibility (x 21 ) vs. temperature data. The magnetic moments estimated from the x 21 data in the high temperature data are around 4 m B, a value close to that expected for MnO. The Curie Weiss temperatures (H p ) obtained from the extrapolation of the high-temperature inverse susceptibility data are 2202 and 2365 K respectively for the 6 nm and 14 nm samples. The negative sign of H p indicates predominant antiferromagnetic interaction. The value of H p decreases with the increase in particle size, probably because the ferromagnetic interaction decreases with the increase in particle size. Accordingly, the value of the magnetization of the 6 nm particles is always higher than that of the 14 nm particles. In Fig. 5(a), we show the variation of magnetization with field at 5 K and 50 K for the 6 nm particles to reveal the occurrence of hysteresis at low temperatures, the H c value being 1750 Oe at 5 K. At high temperatures (T. T B ), the hysteresis disappears. NiO In Fig. 6, we show the X-ray diffraction patterns of four different sizes of NiO nanoparticles. The Rietveld fitting of the profiles was carried out using the Rietveld XND code. 17 The diffraction profiles could be indexed on the Fm3m space group (JCPDS no 47-1049). The particles sizes were estimated using the Scherrer formula by taking the average of three main linewidths obtained from the Rietveld fitted profiles. The lattice parameters obtained from the Rietveld fits are 4.178 Å, 4.182 Å, 4.192 Å and 4.211 Å respectively for the 24, 12, 7 and 3 nm nanoparticles, reflecting a slight increase in the lattice parameter with the decrease in particle size. The presence of a passive organic coating on the surfaces of the particles is evidenced from the TGA curve shown in Fig. 3(b). The coating is removed around 400 uc, accompanied by the partial reduction of NiO to Ni metal. In Fig. 7(a) and (b), we show the TEM images of the NiO nanoparticles with average diameters of 3 and 7 nm respectively, along with the size distributions. Fig. 5 Field dependence of magnetization of the (a) 6 nm MnO nanoparticles and (b) 3 nm NiO nanoparticles at 5 K and 50 K. Notice the absence of saturation and the disappearance of hysteresis at high temperatures. Fig. 6 XRD patterns of the NiO nanoparticles with average diameters of (a) 24, (b) 12, (c) 7 and (d) 3 nm along with the Rietveld fits. The difference patterns are shown below the corresponding observed patterns. This journal is ß The Royal Society of Chemistry 2006 J. Mater. Chem., 2006, 16, 106 111 109

Fig. 7 TEM images of (a) 3 and (b) 7 nm NiO nanoparticles. The insets show the histograms of the particle size distributions (upper panels). The lower panel in (a) shows the SAED pattern. The HREM image of a single NiO nanoparticle of 6 nm is shown in the lower panel in (b). The average particle sizes estimated from the TEM are consistent with the values obtained from the XRD patterns. The indexed SAED pattern shown in Fig. 7(a) suggests the particles to be crystalline. The pattern could be indexed on the basis of the Fm3m space group. The HREM image of a 6 nm NiO particle, shown as an inset in Fig. 7(b), shows a lattice spacing of 2.09 Å, corresponding to the interplannar distance between the (100) planes. We show the temperature dependence of magnetization curves of 3 and 7 nm NiO nanoparticles under ZFC and FC conditions in Fig. 8(a) and (b) respectively. We observe a divergence between the ZFC and FC data, as with the MnO nanoparticles. The ZFC data show maxima corresponding to the blocking temperature (T B ). The T B increases with the particle size, the values being 10 K and 15 K respectively for the particles with the average diameters of 3 nm and 7 nm. Such a proportionality between T B and particle size is expected and has been noticed earlier in some cases. 7,14 We show the inverse susceptibility (x 21 ) data in the inset of Fig. 8(a). The magnetic moments estimated from the inverse susceptibility data in the high temperature regime are around 3 m B, which is close to the value expected for Ni 2+. We notice that the Curie Weiss temperature is close to zero for the 3 nm particles and Fig. 8 The temperature dependence of dc magnetization of (a) 3 nm and (b) 7 nm NiO nanoparticles under ZFC and FC conditions (H 5 100 Oe). The inset shows the inverse susceptibility vs. temperature curves of the (a) 3 nm and (b) 7 nm particles; (c) shows the data for TOPO-capped 3 nm NiO nanoparticles. negative for the 7 nm particles. The value of the magnetization of the 3 nm particles is also higher than that of the 7 nm particles. The magnetic hysteresis curves for the 3 nm NiO particles are shown in Fig. 5(b). The hysteresis curve at 5 K is associated with a H c value of 350 Oe. The hysteresis is absent at 50 K. We have measured the magnetic susceptibility of the 3 nm NiO particles capped with TOPO as well. These capped nanoparticles exhibit a lower magnetization than the NiO nanoparticles prepared in the absence of TOPO. The Curie Weiss temperature of the TOPO-capped nanoparticles is near zero. Conclusions MnO and NiO nanoparticles (both in the Fm3m space group) of different sizes have been prepared by the decomposition of the metal cupferronate or the acetate in organic solvents under solvothermal conditions. Magnetic properties of the well-characterized nanoparticles of these antiferromagnetic oxides reveal the presence of superparamagnetism (often mentioned as ferromagnetic interactions in the literature), as evidenced by the increasing magnetization with decreasing size as well as the magnetic hysteresis at low temperatures. 110 J. Mater. Chem., 2006, 16, 106 111 This journal is ß The Royal Society of Chemistry 2006

Interestingly, the blocking temperature (T B ) increases with the increase in particle size as expected from theory in the case NiO nanoparticles, but shows the opposite relation in the case of MnO nanoparticles. It is not clear why such an inverse relation between T B and particle size manifests itself. It is noteworthy that supermagnetism at low temperatures is the general feature of small nanoparticles of transition metal oxides, which are otherwise antiferromagnetic with fairly high Néel temperatures. Acknowledgements The authors thank Professor E. V. Sampthkumaran for preliminary magnetic measurements. References 1 R. H. Kodama, J. Magn. Magn. Mater., 1999, 200, 359. 2 S. A. Makhlouf, F. T. Parker, F. E. Spada and A. E. Berkowitz, J. Appl. Phys., 1997, 81, 5561. 3 M. Ghosh, E. V. Sampatkumaran and C. N. R. Rao, Chem. Mater., 2005, 17, 2348. 4 C. N. R. Rao and B. Raveau, Transition Metal Oxides, 2nd edn, Wiley-VCH, Germany, 1995. 5 G. H. Lee, S. H. Huh, J.W. Jeong, B. J. Choi, S. H. Kim and H.-C. Ri, J. Am. Chem. Soc., 2002, 124, 12094. 6 M. Yin and S. O Brien, J. Am. Chem. Soc., 2003, 125, 10180. 7 W. S. Seo, H. H. Jo, K. Lee, B. Kim, S. J. Oh and J. T. Park, Angew. Chem., Int. Ed., 2004, 43, 1115. 8 J. Park, E. Kang, C. J. Bae, J.-G. Park, H.-J. Noh, J.-Y. Kim, J.-H. Park, H. M. Park and T. Hyeon, J. Phys. Chem. B, 2004, 108, 13594. 9 J. Park, K. An, Y. Hwang, J. G. Park, H. J. Noh, J. Y. Kim, J. H. Park, N. M. Hwang and T. Hyeon, Nature Mater., 2004, 3, 891. 10 J. T. Richardson, D. I. Yiagas, B. Turk, K. Forster and M. V. Twigg, J. Appl. Phys., 1991, 70, 6977. 11 F. Bødker, M. F. Hansen, C. B. Koch and S. Mørup, J. Magn. Magn. Mater., 2000, 221, 32. 12 R. H. Kodama, S. A. Makhlouf and A. E. Berkowitz, Phys. Rev. Lett., 1997, 79, 1393. 13 C. L. Carnes, J. Stipp and K. J. Klabunde, Langmuir, 2002, 18, 1352. 14 B. J. Park, E. Kang, S. Uk. Son, H. M. Park, M. K. Lee, J. Kim, K. W. Kim, H.-J. Noh, J.-H. Park, C. J. Bae, J.-G. Park and T. Hyeon, Adv. Mater., 2005, 17, 429. 15 W. Xiong, J. Song, L. Gao, J. Jin, H. Zheng and Z. Zhang, Nanotechnology, 2005, 16, 37. 16 K. Tamaki and N. Okabe, Acta Crystallogr., Sect. C, 1996, 52, 1612. 17 J.-F. Bérar and P. Garnier, NIST Spec. Publ., 1992, 846, 212. This journal is ß The Royal Society of Chemistry 2006 J. Mater. Chem., 2006, 16, 106 111 111