Size-Dependent Chemical and Magnetic Ordering in L1 0 -FePt Nanoparticles**

Similar documents
Magnetic Properties of FexPtyAu100_X_y Nanoparticles

Thermal Effects in Magnetic Recording Media

Ceramic Processing Research

Size and Shape Control of Monodisperse FePt Nanoparticles

Title: Magnetic chains of metal formed by assembly of small nanoparticles

Crystallographic ordering studies of FePt nanoparticles by HREM

Structural Effect on the Oxygen Evolution Reaction in the Electrochemical Catalyst FePt

Magnetic properties of spherical fcc clusters with radial surface anisotropy

Mechanism of c-axis orientation of nanostructured FePt/B2O3 thin films. Citation PHYSICAL REVIEW B (2008), 77(9)

Tuning the magnetic properties of Co nanoparticles by Pt capping

EFFECT OF MnO SHELL TO PREVENT SINTERING OF FePt NANOPARTICLES DURING THE ANNEALING PROCESS

Thickness dependence of magnetic properties of granular thin films with interacting particles

Jahresbericht 2003 der Arbeitsgruppe Experimentalphysik Prof. Dr. Michael Farle

ANGULAR DEPENDENCE OF MAGNETIC PROPERTIES IN Co/Pt MULTILAYERS WITH PERPENDICULAR MAGNETIC ANISOTROPY

Supporting Information

voltage measurement for spin-orbit torques"

Relaxor characteristics of ferroelectric BaZr 0.2 Ti 0.8 O 3 ceramics

Role of diffused Co atoms in improving effective exchange coupling in Sm-Co/ Fe spring magnets

International Journal of Scientific & Engineering Research, Volume 5, Issue 3, March-2014 ISSN

CHAPTER 4. PREPARATION AND CHARACTERIZATION OF Cr-DOPED, Co-DOPED AND Fe-DOPED NIO NANOPARTICLES

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

Structural and magnetic properties of Ni doped CeO 2 nanoparticles

STRUCTURE AND MAGNETIC PROPERTIES OF SiO 2 COATED Fe 2 NANOPARTICLES SYNTHESIZED BY CHEMICAL VAPOR CONDENSATION PROCESS

Defense Technical Information Center Compilation Part Notice

Dense-plasma-driven ultrafast formation of FePt organization on silicon substrate

SUPPLEMENTARY INFORMATION. Observation of tunable electrical bandgap in large-area twisted bilayer graphene synthesized by chemical vapor deposition

Electron Microscopy Images Reveal Magnetic Properties

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

Structure and Curie temperature of Y 2 Fe 17 x Cr x

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

Supporting Information

Structure and magnetic properties of SiO 2 -coated Co nanoparticles

Structure Optimization of FePt Nanoparticles of Various Sizes for Magnetic Data Storage

Planar Hall Effect in Magnetite (100) Films

Perpendicular exchange bias and magnetic anisotropy in CoOÕpermalloy multilayers

The Magnetic Properties of Superparamagnetic Particles by a Monte Carlo Method

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

Magnetic anisotropy of La 0.7 Sr 0.3 MnO 3 nanopowders

Supplementary information

, Fax: , Inframat Corp., Willington, CT 06279

Oriented growth of CoPt nanoparticles by pulsed laser deposition

Gold-Coated Iron Nanoparticles: Synthesis, Characterization, and Magnetic Field-Induced Self-Assembly

Effects of ion-beam irradiation on the L1 0 phase transformation and their magnetic properties of FePt and PtMn films (Invited)

Mechanical behavior and magnetic separation of quasi-one-dimensional SnO 2 nanostructures: A technique for achieving monosize nanobelts/nanowires

Layered Double Hydroxide Nanoplatelets with Excellent Tribological Properties under High Contact Pressure as Water-based Lubricant Additives

Sacrifical Template-Free Strategy

Magnetostatic Coupling in CoFe 2 O 4 /Pb(Zr 0.53 Ti 0.47 )O 3 Magnetoelectric Composite Thin Films of 2-2 Type Structure

Exchange Coupled Composite Media for Perpendicular Magnetic Recording

Zinc-Blende CdS Nanocubes with Coordinated Facets for Photocatalytic Water Splitting

Sub-10-nm Au-Pt-Pd Alloy Trimetallic Nanoparticles with. High Oxidation-Resistant Property as Efficient and Durable

Exchange bias in core/shell magnetic nanoparticles: experimental results and numerical simulations

Supporting Information

ABSTRACT I. INTRODUCTION II. BACKGROUND OF STUDY

Preparation, Structural Characterization, and Dynamic Properties Investigation of Permalloy Antidot Arrays

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

J 12 J 23 J 34. Driving forces in the nano-magnetism world. Intra-atomic exchange, electron correlation effects: Inter-atomic exchange: MAGNETIC ORDER

performance electrocatalytic or electrochemical devices. Nanocrystals grown on graphene could have

Anisotropy properties of magnetic colloidal materials

Department of Physics, University of Texas at Arlington, Arlington, TX 76019, USA

Positive and negative exchange bias effects in the simple perovskite manganite NdMnO3

Supporting Information

Available online at Universal Research Publications. All rights reserved

Supporting Information

Preparation of Fe 3 O 2 Nanostructures via Inverse Micelle Method and Study of Their Magnetic Properties for Biological Applications

Adouble-layered (DL) perpendicular anisotropy system

General Synthesis of Graphene-Supported. Bicomponent Metal Monoxides as Alternative High- Performance Li-Ion Anodes to Binary Spinel Oxides

A Simple Precipitation Method for Synthesis CoFe 2 O 4 Nanoparticles

Specific Heat and Electrical Transport Properties of Sn 0.8 Ag 0.2 Te Superconductor

Superconductivity in oxygen-annealed FeTe 1-x S x single crystal

Supporting Information

Carrie Schindler 1. Department of Physics, University of Florida, Gainesville, FL Summer REU, 28 July 2009

and their Maneuverable Application in Water Treatment

Characterisation of Nanoparticles using Advanced Electron Microscopy

Giant exchange bias in a single-phase magnet with two magnetic

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

Coupled perpendicular magnetization in Fe/Cu/Fe trilayers

Facile Synthesis and Catalytic Properties of CeO 2 with Tunable Morphologies from Thermal Transformation of Cerium Benzendicarboxylate Complexes

SYNTHESIS IN SUPERCRITICAL AMMONIA AND CHARACTERIZATION OF NANOSTRUCTURED NICKEL OXINITRIDE

Magnetic properties of dipolar interacting single-domain particles

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

Keywords: superconductivity, Fe-based superconductivity, FeTe, alcohol, wine

MAGNETIC-PROPERTY ENHANCEMENT OF SIZED CONTROLLED COBALT-GOLD CORE-SHELL NANOCRYSTALS

High-Performance Flexible Asymmetric Supercapacitors Based on 3D. Electrodes

Supporting Information

Supporting Information

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

Supporting Information

Inorganic Nanoparticles Unique Properties and Novel Applications

Supporting Information

High Salt Removal Capacity of Metal-Organic Gel Derived. Porous Carbon for Capacitive Deionization

SUPPLEMENTARY MATERIAL

Direct measurement of giant electrocaloric effect in BaTiO 3 multilayer thick film structure beyond theoretical prediction

Application of Nano-ZnO on Antistatic Finishing to the Polyester Fabric

Room-temperature ferromagnetism in nanoparticles of Superconducting materials

Novel fungus-titanate bio-nano composites as high performance. absorbents for the efficient removal of radioactive ions from.

High tunable dielectric response of Pb 0.87 Ba 0.1 La 0.02 (Zr 0.6 Sn 0.33 Ti 0.07 ) O 3 thin film

Influence of Size on the Properties of Materials

Phenomenology and Models of Exchange Bias in Core /Shell Nanoparticles

Magnetism and Hall effect of the Heusler alloy Co 2 ZrSn synthesized by melt-spinning process

Defense Technical Information Center Compilation Part Notice

Transcription:

DOI: 10.1002/adma.200601904 Size-Dependent Chemical and Magnetic Ordering in L1 0 -FePt Nanoparticles** By Chuan-bing Rong, Daren Li, Vikas Nandwana, Narayan Poudyal, Yong Ding, Zhong Lin Wang, Hao Zeng, and J. Ping Liu* [*] Dr. J. P. Liu, Dr. C.-b. Rong, D. Li, V. Nandwana, N. Poudyal Department of Physics University of Texas at Arlington Arlington, TX 76019 (USA) E-mail: pliu@uta.edu Dr. Y. Ding, Dr. Z. L. Wang School of Materials Science and Engineering Georgia Institute of Technology Atlanta, GA 30332 (USA) Dr. H. Zeng Department of Physics, University at Buffalo The State University of New York Buffalo, NY 14260 (USA) [**] This work was supported by US DoD/MURI grant N00014-05-1-0497, DARPA through ARO under Grant No DAAD 19-03-1-0038 and NSF DMR-0547036. FePt nanoparticles have great application potential in advanced magnetic materials such as ultrahigh-density recording media and high-performance permanent magnets. [1 3] The key for applications is the very high uniaxial magnetocrystalline anisotropy of the L1 0 -FePt phase, which is based on crystalline ordering of the face-centered tetragonal (fct) structure, described by the chemical-ordering parameter S. [4] Higher chemical ordering results in higher magnetocrystalline anisotropy. Unfortunately, as-synthesized FePt nanoparticles take a disordered face-centered cubic (fcc) structure that has low magnetocrystalline anisotropy. Heat-treatment is necessary to convert the fcc structure to the ordered fct structure. Several previous theoretical and experimental investigations have been reported on the size-dependent chemical ordering of FePt nanoparticles. [5 9] It has been observed that the degree of ordering decreases with decreasing particle size of the sputtered FePt nanoparticles. [5,6] Theoretical simulation predicted that the ordering would not take place when the particle size is below a critical value. [8,9] However, there have not been systematic experimental studies on quantitative size dependence of chemical ordering of FePt nanoparticles due to the lack of monodisperse L1 0 -FePt nanoparticles with controllable sizes. There are also few studies reported to date on the quantitative particle size dependence of magnetic properties, including the Curie temperature, coercivity, and magnetization of the L1 0 -FePt phase, although it has been well accepted that there is a size effect on the ferromagnetism of any low-dimensional magnets. [10,11] Additionally, the magnetic properties of FePt ferromagnets, as observed in thin-film samples, [4,12] are affected by the degree of chemical ordering, which is in turn size dependent. It is therefore highly desirable to understand the size and chemical-ordering effects, and their influence on the magnetic properties of the nanoparticles. A major hurdle in obtaining the particle size dependence of structural and magnetic properties of the L1 0 phase is particle sintering during heat-treatments that convert the fcc phase to the fct phase. [13,14] This long-pending problem has been solved recently by adopting the salt-matrix annealing technique. [15,16] With this technique, particle aggregation during the phase transformation has been avoided so that the true size-dependent properties of the fct phase can be measured. In this paper, we report results on quantitative particle size dependence of the chemical-ordering parameter S and selected magnetic properties, including the Curie temperature, T c, magnetization, M s, and coercivity, H c, with the particle size varying from 2 to 15 nm. Figure 1 shows the transmission electron microscopy (TEM) images of the FePt nanoparticles with different sizes before and after annealing in a salt matrix at 973 K for 4 h. The images, from left to right, show nanoparticles with nominal diameters of 2, 4, 6, 8, and 15 nm, respectively. The upper and lower rows are images of as-synthesized and salt-matrixannealed nanoparticles, respectively. As shown in Figure 1, the particle size is retained well upon annealing. Both the assynthesized and annealed nanoparticles are monodisperse with a standard deviation of 5 10 % in diameter. TEM observations also revealed that when the particle size is smaller than or equal to 8 nm, the fct nanoparticles are monocrystalline, whereas the 15 nm fct particles are polycrystalline. [15] It is interesting to see that the L1 0 nanoparticles, tiny ferromagnets at room temperature, are dispersed very well without agglomeration despite the dipolar interaction between the particles, if a solvent with high viscosity is chosen and if the solution is diluted. Extensive TEM and X-ray diffraction (XRD) analyses have proved that the technique of salt-matrix annealing can be applied to heat-treatments of the FePt nanoparticles without leading to particle agglomeration and sintering, if a suitable salt-to-particle ratio and proper annealing conditions are chosen. [16] Figure 2 shows the XRD patterns of the 4 nm, as-synthesized, fcc-structured nanoparticles and the particles annealed in a salt matrix at 873 K for 2 h, 973 K for 2 h, and 973 K for 4 h (from bottom to top), respectively. As shown in the figure, the positions of the (111) peaks shift in the higher-an- 2984 2006 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim Adv. Mater. 2006, 18, 2984 2988

a b c d e COMMUNICATION f g h i j Figure 1. TEM images of as-synthesized FePt nanoparticles of a) 2, b) 4, c) 6, d) 8, and e) 15 nm, and salt-annealed nanoparticles of f) 2, g) 4, h) 6, i) 8, and j) 15 nm. The annealing conditions for a 2 nm particle are 973 K for 8 h, while those for other particles are 973 K for 4 h. The SAED patterns are also given for 2 and 15 nm salt-annealed particles. Intensity (a.u.) 700ºC X 4h 700ºC X 2h ºC X 2h As-synthesized (001) (110) (111) 20 30 40 50 (200) (002) 700 ºC 4 hours 15nm 8nm 6nm (220) (202) 20 25 30 35 40 45 50 55 60 65 70 75 2θ (degree) Figure 2. XRD patterns of 4 nm FePt nanoparticles annealed at different parameters. The inset gives the XRD patterns of 6, 8, and 15 nm nanoparticles annealed in salt matrix at 973 K for 4 h. gle direction with increasing annealing temperature and time. This shift was caused by the change of lattice parameters during the phase transition from fcc to fct. The (001) and (110) peaks, which are the characteristic superlattice peaks of the ordered fct phase, developed with increasing annealing temperature and time. For samples annealed at 973 K for 4 h, the superlattice peaks (001) and (110) were fully developed, implying an fcc fct transition. The inset of Figure 2 also gives the XRD patterns of the converted fct phase of 6, 8, and 15 nm nanoparticles annealed at 973 K for 4 h in a salt matrix. The grain sizes estimated by the Scherrer formula are about 4.7, 6.8, 8.2, and 13.3 nm for the 4, 6, 8, and 15 nm nanoparticles, respectively, which matches the particle sizes (201) (112) very well. The relatively smaller grain size of the 15 nm nanoparticles is due to the polycrystalline morphology mentioned above. XRD does not give any diffraction peaks for the 2 nm FePt particles, largely because of the lower crystallinity, strong surface effects, and/or peak broadening due to the small size. Selected area electron diffraction (SAED) was therefore also used to identify the crystalline structure of the particles, especially for the smaller particles. The SAED patterns showed that the annealed nanoparticles are L1 0 -fct phase if the particle size is equal to or larger than 4 nm. However, the SAED pattern of 2 nm particles, which is given in the inset of Figure 1f, does not show (001) and (110) rings (in contrast to the SAED pattern of the fct nanoparticles shown in Fig. 1j). This indicates that the 2 nm FePt particles cannot be transformed to the L1 0 -fct structure. We extended the annealing time to 8 h and the results remained the same. This direct experimental observation is consistent with the theoretical predictions and is fundamentally important as well as being significant for specific applications. [8,9] It can be concluded that when the particle size is smaller than a critical value of around 3 nm, chemical ordering will not take place. To evaluate the effect of chemical ordering of the L1 0 phase in a quantitative way, the following equation was used to calculate the ordering parameter S [17 19] S 0:85 I 1=2 001 1 I 002 where I 001 and I 002 are the integrated intensities of the (001) and (002) diffraction peaks from the XRD patterns. The longrange S value of fct-fept nanoparticles with different sizes, annealed under the same conditions (973 K, 4 h), was calcu- Adv. Mater. 2006, 18, 2984 2988 2006 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.advmat.de 2985

lated according to Equation 1. Figure 3 shows the correlation between S and particle size. Each experimental point was obtained by measuring at least four different samples and the error bar is the standard deviation. As expected, S decreases with reducing particle size. In the case of 2 nm FePt particles, S vanishes because no superlattice structure can be found. S 1.0 0.9 0.8 0.7 leads to reduced magnetocrystalline anisotropy. [20] The exchange coupling can be further seen from the relatively high remanence ratio M r /M s > 0.5, where M r and M s are remanence and saturation magnetization, respectively (we used the magnetization values at 7 T as an approximation of M s ). The reduced coercivity may also be attributed to multiple c-axes within a nanoparticle, which can lead to a significantly reduced effective anisotropy. [21] Figure 4 shows that the M s values decrease with decreasing particle size and are lower than the value for the bulk L1 0 phase (1140 emu cm 3 (1 emu cm 3 = 1000 A m 3 ) at room temperature). [22] Figure 5 gives M s at 300 K plotted against 1/d. A linear correlation was obtained. However, the linearly extrapolated M s is lower than the experimental value since the hysteresis curve in this work is not saturated for the applied field of 70 koe. The main reason for this linear reduction may be related to the reduced magnetization on the surfaces of the nanoparticles. [23 25] 0.6 4 6 8 10 12 14 16 d (nm) 1000 Figure 3. The dependence of the long-range ordering parameter on particle size. The dashed line is a guide to the eye. Figure 4 shows the hysteresis loops at 300 K of the fct nanoparticles with different sizes. It shows that the coercivity, H c, increases with increasing particle size (d) asd 8 nm. Giant coercivity values of up to 30 koe (1 Oe = 79.57 A m 1 ) have been achieved from the monodisperse hard magnetic nanoparticles, which give the particles great potential for applications. The enhancement of H c with increased particle size can be attributed partly to the size-dependent chemical ordering discussed above. However, larger particles of 15 nm have lower H c than the 8 nm particles. This may be caused by the polycrystalline structure, as intergranular exchange coupling M (emu/cm 3 ) 400 200 0-200 -400 - - 15 nm 8 nm 6 nm 4 nm -60-40 -20 0 20 40 60 H (koe) Figure 4. Hysteresis loops of salt-annealed 4, 6, 8, and 15 nm nanoparticles. M s (emu/cm 3 ) 400 0.00 0.05 0.10 0.15 0.20 0.25 d -1 (nm -1 ) Figure 5. The relation between M s and particle size. It is of fundamental interest to understand the correlation between particle size and Curie temperature of high-anisotropy nanoparticles. Most studies on low-dimensional ferromagnets reported so far are on granular thin films and nanocomposites where grain polydispersity and existence of matrix materials mask the genuine size effect. [26] The magnetic-ordering temperatures of ferromagnets with reduced size in all three dimensions have been little studied because of a lack of monodisperse, ferromagnetic nanoparticles with small sizes. This challenge has been addressed by applying the salt-matrix-annealing technique, with which ferromagnetic nanoparticles with high anisotropy at room temperature are obtained. Figure 6 shows the dependence of the Curie temperature on the fct-fept particle size (open squares). It should be noted that S for these points corresponds to the values in Figure 3. We measured the magnetization M versus T in an applied field of 1 koe. Each point was obtained by averaging data from at least four different samples with two M T measure- 2986 www.advmat.de 2006 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim Adv. Mater. 2006, 18, 2984 2988

T c (K) 780 760 740 720 700 680 660 Direct experiment After correction matrix-annealing time and temperature. Figure 7 gives the dependence of the Curie temperature on the ordering parameters of the 8 nm particles. It shows that T c increases linearly with increasing S. The data can be fitted well in the region 0.6 S 1.0 by a linear relation T c = T c0 + as, where T c0 = (570 ± 18) K and a = (190 ± 20) K are fitting parameters. COMMUNICATION 640 750 620 ments for each sample. The error bar is based on the statistical standard deviation. To prevent sintering and agglomeration of the particles during the measurements, we kept the nanoparticles in the salt matrix. T c was determined from the intersection of extrapolations of the greatest slope and flat region above T c. As shown in Figure 6, T c decreases remarkably with declining particle size, especially when d<6 nm. For the 2 nm nanoparticles, we could not obtain T c for the fct phase since the phase transition cannot be realized. To understand the strong size dependence of the Curie temperature, the finite-size-scaling theory was used to fit the behavior of T c. [10,11] This theory predicts that the shift of T c from the value for bulk materials depends on the dimension in the following manner T c T c d T c 4 6 8 10 12 14 16 d (nm) Figure 6. The dependence of Curie temperature on the particle size. The dashed line is a guide to the eye and the solid lines are made by fitting the data. ˆ d 1=v 2 d 0 where T c (d) is the Curie temperature as a function of d, T c ( ) is the bulk Curie temperature, d 0 is a constant, and v is the critical exponent of the correlation length. By fitting the data in Figure 6 (the open squares), T c ( ) = (795 ± 10) K, d 0 = (0.84 ± 0.05) nm, and v = (0.91 ± 0.10) were obtained. The value of d 0, which is comparable to the lattice constant, and is the microscopic length scale, is reasonable. However, the fitting value of v is out of the range v = (0.65 ± 0.07) to (0.73 ± 0.02) predicted by the isotropic three-dimensional Heisenberg model. [27] In addition, T c ( ) is remarkably higher than the known T c of bulk FePt materials. [22,27] The discrepancies may be due to the influence of chemical ordering on the Curie temperature. To distinguish the effects of particle size and chemical ordering on the Curie temperature, we performed T c measurements with fixed particle size of 8 nm. Samples with a series of ordering parameters were obtained by adjusting the salt- T c (K) 700 650 Experiments Fitting line 0.6 0.7 0.8 0.9 1.0 S Figure 7. The correlation between the ordering parameters and the Curie temperature with fixed particle size of 8 nm. This linear correlation has also been obtained for the fct particles of 4 nm. It is striking to see such strong correlation between the magnetic-ordering temperature and the chemicalordering parameter. Previous studies based on sputtered thin films showed that the disordered fcc-fept phase has a T c lower than that of the ordered fct-fept phase but with a weak dependence on the ordering. [12,28] Further work is needed to fully understand the strong correlation in the nanoparticles. By correcting for the ordering effect, we can obtain the real particle size effect on the Curie temperature of fully ordered fct-fept nanoparticles, as represented by the solid squares in Figure 6. The best fits of the data give T c ( ) = (775 ± 20) K, v = (0.67 ± 0.11), and d 0 = (0.90 ± 0.10) nm. The fitted value of v is consistent with the known value of v for magnetic systems, which is in the range (0.65 ± 0.07). [27] The fitted value of d 0 is a little more than twice the unit-cell size and is reasonable. The fitted value of T c ( ) of (775 ± 20) K is slightly higher than the reported 750 K for bulk fct-fept material but the deviation is comparable to the experimental error. The Fe-rich composition could contribute to this deviation. Thus, the dependence of the Curie temperature on the particle size is generally consistent with finite-size-scaling theory. Although in such a model only nearest-neighbor interactions are considered, in FePt systems both long-range exchange interactions and exchange mediated by Pt should be taken into account. Modeling efforts including these longer-range interactions and the surface effect are underway. In summary, we have studied the size dependence of the chemical-ordering parameter and magnetic-ordering proper- Adv. Mater. 2006, 18, 2984 2988 2006 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.advmat.de 2987

ties including coercivity, magnetization, and Curie temperature of fct-fept nanoparticles. Based on the newly developed salt-matrix-annealing technique, direct experimental evidence has been found that the long-range ordering parameter S decreases with decreasing particle size d. The chemical ordering does not occur for d 2 nm. It is also revealed that magnetic properties including Curie temperature, magnetization, and coercivity are strongly dependent on both the particle size and the chemical ordering. The interplay between the size effect and the chemical-ordering effect defines the magnetic behavior. The shift in T c can generally be described by the finite-size-scaling theory after considering the chemical-ordering effect on T c. Experimental The fcc-structured FePt nanoparticles with different sizes (from 2 to 15 nm) were synthesized using a chemical-solution method with adjusted synthetic parameters. [1,29,30] The fcc particles were then mixed with ball-milled NaCl powder in hexane or another organic solvent with the assistance of surfactants. Dry mixtures of FePt particles and NaCl powders were obtained after the solvent was evaporated completely. The mixtures were then annealed in a forming gas (93 % Ar+7 % H 2 ) at different temperatures for different times. After the annealing, the NaCl powder was washed away by deionized water and the FePt nanoparticles were recovered and dispersed in organic solvents, such as cyclohexane or ethanol, in the presence of surfactants. The elemental composition analyses, using inductively coupled plasma-optical emission spectroscopy (ICP-OES), showed that there was negligible NaCl contamination in the salt-matrix-annealed FePt nanoparticles and the particle composition was Fe 52 Pt 48. A transmission electron microscope was used to analyze the morphology and crystalline structures. X-ray diffraction (XRD) was used to determine the phase transition, the long-range ordering parameters, the grain size, and the particle size. The magnetic hysteresis loops were measured with a magnetic properties measurement system (MPMS) from specimens of a mixture of epoxy and the magnetic nanoparticles. Curie temperatures were measured by a physical properties measurement system (PPMS) with high-temperature and high-vacuum vibrating sample magnetometer. Received: August 21, 2006 Revised: September 12, 2006 [1] S. H. Sun, C. B. Murray, D. Weller, L. Folks, A. Moser, Science 2000, 287, 1989. [2] H. Zeng, J. Li, J. P. Liu, Z. L. Wang, S. H. Sun, Nature 2002, 420, 395. [3] C. B. Rong, H. W. Zhang, X. B. Du, J. Zhang, S. Y. Zhang, B. G. Shen, J. Appl. Phys. 2004, 96, 3921. [4] R. A. Ristau, K. Barmak, L. H. Lewis, K. R. Coffey, J. K. Howard, J. Appl. Phys. 1999, 86, 4527. [5] Y. K. Takahashi, T. Ohkubo, M. Ohnuma, K. Hono, J. Appl. Phys. 2003, 93, 7166. [6] T. Miyazaki, O. Kitakami, S. Okamoto, Y. Shimada, Z. Akase, Y. Murakami, D. Shindo, Y. K. Takahashi, K. Hono, Phys. Rev. B: Condens. Matter Mater. Phys. 2005, 72, 144 419. [7] Z. Y. Jia, S. S. Kang, S. F. Shi, D. E. Nikles, J. W. Harrell, J. Appl. Phys. 2005, 97, 10J 310. [8] R. V. Chepulskii, W. H. Butler, Phys. Rev. B: Condens. Matter Mater. Phys. 2005, 72, 134 205. [9] B. Yang, M. Asta, O. N. Mryasov, T. J. Klemmer, R. W. Chantrell, Scr. Mater. 2005, 53, 417. [10] K. Binder, Physica 1972, 62, 508. [11] M. Farle, K. Baberschke, U. Stetter, A. Aspelmeier, F. Gerhardter, Phys. Rev. 1993, 47, 11 571. [12] S. Okamoto, N. Kikuchi, O. Kitakami, T. Miyazaki, Y. Shimada, K. Fukamichi, Phys. Rev. B: Condens. Matter Mater. Phys. 2002, 66, 024 413. [13] H. Zeng, S. H. Sun, T. S. Vedantam, J. P. Liu, Z. R. Dai, Z. L. Wang, Appl. Phys. Lett. 2002, 80, 2583. [14] T. Thomson, S. L. Lee, M. F. Toney, C. D. Dewhurst, F. Y. Ogrin, C. J. Oates, S. Sun, Phys. Rev. B: Condens. Matter Mater. Phys. 2005, 72, 064 441. [15] K. Elkins, D. Li, N. Poudyal, V. Nandwana, Z. Jin, K. Chen, J. P. Liu, J. Phys. D: Appl. Phys. 2005, 38, 2306. [16] D. Li, N. Poudyal, V. Nandwana, Z. Jin, K. Elkins, J. P. Liu, J. Appl. Phys. 2006, 99, 08E 911. [17] B. E. Warren, X-Ray Diffraction, Dover Publications, New York 1990, Ch. 12. [18] A. Cebollada, D. Weller, J. Sticht, G. R. Harp, R. F. C. Farrow, R. F. Marks, R. Savoy, J. C. Scott, Phys. Rev. B: Condens. Matter Mater. Phys. 1994, 50, 3419. [19] J. A. Christodoulides, P. Farber, M. Daniil, H. Okumura, G. C. Hadjipanayis, V. Skumryev, A. Simopoulos, D. Weller, IEEE Trans. Magn. 2001, 37, 1292. [20] G. Herzer, IEEE Trans. Magn. 1989, 25, 3327. [21] W. Scholz, J. Fidler, T. Schrefl, D. Suess, H. Forster, R. Dittrich, V. Tsiantos, J. Magn. Magn. Mater. 2004, 272-276, 1524. [22] T. Klemmer, D. Hoydick, H. Okumura, B. Zhang, W. A. Soffa, Scr. Metall. Mater. 1995, 33, 1793. [23] A. E. Berkowitz, W. J. Schuele, P. J. Flanders, J. Appl. Phys. 1987, 39, 1261. [24] Z. X. Tang, C. M. Sorensen, K. J. Klabunde, G. C. Hadjipanayis, Phys. Rev. Lett. 1991, 67, 3602. [25] J. P. Chen, C. M. Sorensen, K. J. Klabunde, G. C. Hadjipanayis, Phys. Rev. B: Condens. Matter Mater. Phys. 1996, 54, 9288. [26] H. Zeng, R. Sabirianov, O. Mryasov, M. L. Yan, K. Cho, D. J. Sellmyer, Phys. Rev. B: Condens. Matter Mater. Phys. 2002, 66, 184 425. [27] S. N. Kaul, J. Magn. Magn. Mater. 1985, 53, 5. [28] M. Watanabe, T. Nakayama, K. Watanabe, T. Hirayama, A. Tonomura, Mater. Trans., JIM 1996, 37, 489. [29] K. E. Elkins, T. S. Vedantam, J. P. Liu, H. Zeng, S. Sun, Y. Ding, Z. L. Wang, Nano Lett. 2003, 3, 1647. [30] M. Chen, J. P. Liu, S. Sun, J. Am. Chem. Soc. 2004, 126, 8394. 2988 www.advmat.de 2006 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim Adv. Mater. 2006, 18, 2984 2988