Download as pdf or txt
Download as pdf or txt
You are on page 1of 5

Materials Research Bulletin 41 (2006) 525–529

www.elsevier.com/locate/matresbu

Synthesis and magnetic properties of Fe3O4 nanoparticles


Yuan-hui Zheng, Yao Cheng, Feng Bao, Yuan-sheng Wang *
The State Key Laboratory of Structural Chemistry, Fujian Institute of Research on the Structure of Matter,
Chinese Academy of Sciences, Graduate School of Chinese Academy of Sciences,
Yangqiao West Road 155#, Fuzhou, Fujian 350002, China
Received 26 April 2005; received in revised form 25 August 2005; accepted 19 September 2005
Available online 4 October 2005

Abstract
Ferromagnetic Fe3O4 nanoparticles with diameter of 27 nm were prepared by a hydrothermal route in the presence of a
surfactant, sodium bis(2-ethylhexyl)sulfosuccinate (AOT). The as-synthesized product was characterized by X-ray diffraction
(XRD), transmission electron microscopy (TEM) and high-resolution transmission electron microscopy (HRTEM). The hysteresis
loops of the iron oxide nanoparticles were measured using a physical property measuring system (PPMS), and the results showed a
superparamagnetic behavior at room temperature.
# 2005 Elsevier Ltd. All rights reserved.

Keywords: A. Oxides; B. Chemical synthesis; C. Electron microscopy; D. Magnetic properties

1. Introduction

In the past decade, various nanomaterials have been extensively pursued for their catalytic, optical, electrical,
optoelectronic, mechanical, thermodynamic and magnetic properties, which are quite different from those of their
bulk counterparts, and thus, wide range of potential applications in nanodevices [1,2]. Among all the above-mentioned
properties, the magnetism is the one dramatically dependent on the size of nanophase. Ferromagnetic materials are
made up of domains that are groups of spins all pointing in the same direction, separated by domains walls. As the
particle size decreases toward a critical value, the formation of domains walls becomes energetically unfavorable, and
the particles consist of single magnetic domain, thus, results in the superparamagnetic phenomenon and quantum
tunneling of magnetization occurred [3]. Superparamagnetic nanoparticles can offer a great potential applications in
different areas such as ferrofluids, color imaging, magnetic refrigeration, detoxification of biological fluids,
magnetically controlled transport of anti-cancer drugs, magnetic resonance imaging (MRI) and magnetic cell
separation [4–7].
To date, many approaches including reverse micelles method [8] and thermal decomposition route [9–14]
have been developed for the preparation of iron oxide nanoparticles. They usually lead to complicated process or
require relatively high temperatures. Herein, we presented a simple sodium bis(2-ethylhexyl)sulfosuccinate
(AOT) assisted hydrothermal method to synthesize the superparamagnetic Fe3O4 nanoparticles with relative high
yield.

* Corresponding author. Tel.: +86 591 8370 5402; fax: +86 591 8370 5402.
E-mail address: yswang@fjirsm.ac.cn (Y.-s. Wang).

0025-5408/$ – see front matter # 2005 Elsevier Ltd. All rights reserved.
doi:10.1016/j.materresbull.2005.09.015
526 Y.-h. Zheng et al. / Materials Research Bulletin 41 (2006) 525–529

2. Experimental

In a typical hydrothermal synthesis procedure, 0.404 g (1 mmol) Fe(NO3)39H2O and 0.222 g (0.5 mmol) sodium
bis(2-ethylhexyl)sulfosuccinate were dissolved in 10 ml de-ionized water followed by adding 2 ml of 50% hydrazine
hydrate with agitating. The above mixture was then turned into a Teflon-lined stainless steel autoclave of 20 ml
capacity. The sealed tank was heated at a rate of 5 8C/min to 160 8C and maintained at this temperature for 10 h in an
oven, and then cooled down to room temperature naturally. The black precipitates were collected by filtration and
washed with de-ionized water and ethanol for several times, and finally dried in the air.
The X-ray diffraction (XRD) pattern of the product was recorded by step scan in the RIGAKU-DMAX2500 X-ray
diffractometer with Cu Ka radiation (l = 0.1541 nm) at 40 kV and 100 mA. The scan range (2u) was from 58 to 858
with the step of 0.058 and the resolution of 0.018. The morphologies and microstructures of the as-synthesized sample
were characterized by JEOL-2010 transmission electron microscope equipped with energy dispersive spectroscope
(EDS) and operated at 200 kV. All the above procedures were conducted at room temperature. The hysteresis loops of
the as-synthesized nanoparticles were measured using a physical property measuring system (PPMS) at 300 and 10 K,
respectively.

3. Results and discussion

The XRD pattern of the as-synthesized sample shown in Fig. 1 can be indexed to the cubic Fe3O4 phase (ICSD 20-
596) and FeO(OH) phase (ICSD 71-810). The ratio of the integral intensity from all the diffraction peaks of FeO(OH)
to that of Fe3O4 is about 0.05, revealing that Fe3O4 is the major product and the proportion of FeO(OH) is less than 5%
in the whole sample [16]. The calculated lattice constant of Fe3O4 extracted from the XRD data is a = 0.840 nm, in
good agreement with the literature data. IR spectrum shown in Fig. 2 demonstrated the hybrid chemical nature of the
nanocrystals. The absorption bands at 580 cm1 were related to the vibrations of Fe–O functional group, and the other
adsorption peaks at 1621, 1398, and 1042, 886, 790 cm1 were attributed to C O and (CH2)n vibrations,
respectively, which indicated that AOT molecules were chemically bonded to the surface of Fe3O4 nanoparticles.
The mean particle size calculated from Fe3O4 diffraction peaks according to the Debye–Scherrer equation was
27 nm, in good agreement with the result observed by transmission electron microscopy (TEM) shown in Fig. 3A.
TEM micrograph of the figure also revealed the quasi-spherical morphology of the synthesized Fe3O4 nanoparticles.
The selected area electron diffraction (SAED) pattern (inset in Fig. 3A) taken from the area consisting of many
particles presented Fe3O4 polycrystalline diffraction rings, in accordance with the XRD result. The high-resolution
transmission electron microscopy (HRTEM) image presented in Fig. 3B provided an indication of the morphology of
the sample in detail and exhibited that Fe3O4 nanoparticles were single-crystalline.

Fig. 1. XRD pattern of the sample prepared at 160 8C for 10 h.


Y.-h. Zheng et al. / Materials Research Bulletin 41 (2006) 525–529 527

Fig. 2. IR spectrum of the sample prepared through hydrothermal route at 160 8C for 10 h.

The hysteresis loops of the as-prepared sample measured at 10 and 300 K, respectively, are shown in Fig. 4. At
300 K, the saturation magnetization of the Fe3O4 nanoparticles is 3.69 emu/g, much smaller than that (68.7 emu/g)
of the Fe3O4 nanoparticles sized about 70 nm prepared through a hydrothermal method without any surfactants
[17]. Previous studies have shown that the saturation magnetization of 10 nm Fe3O4 nanoparticles was only
1.25 emu/g at room temperature [13], and the noncollinear spin structure, which originated from the pinning of
the surface spins and the surface-coated surfactant, was responsible for the reduction of magnetic moment in such
nanoparticles [18]. Thus, it is reasonable that the small saturation magnetization in our case is most likely
attributed to the much smaller size of Fe3O4 nanoparticles, their surface spins and the existence of surfactants on
the surface. As shown in the inset of Fig. 4A, the saturation magnetization and coercive force of the as-synthesized
nanoparticles at 10 K were 4.11 emu/g and 320 Oe, respectively. However, the latter was negligible at room
temperature (inset of Fig. 4B), which was a typical characteristic of superparamagnetic materials. The emerging of
superparamagnetism at room temperature was due to that the Fe3O4 particle size was smaller than a single domain
(54 nm) [19].

Fig. 3. (A) TEM micrograph and the corresponding SAED pattern of as-synthesized Fe3O4 nanoparticles, and (B) HRTEM image and the Fourier
transform pattern from an individual Fe3O4 nanoparticle.
528 Y.-h. Zheng et al. / Materials Research Bulletin 41 (2006) 525–529

Fig. 4. Hysteresis curves of the as-synthesized product measured at (A) 10, and (B) 300 K. The insets are the enlargement of the center part of the
curves.

Since the raw materials employed in the synthesis were only Fe(NO3)39H2O and hydrazine in aqueous solution, it
is believed that the disproportionate decomposition of hydrazine proceeded before the hydrothermal process. The
whole reaction process could be expressed as following:
3N2 H4 ¼ N2 þ 4NH3 (1)

NH3 þ H2 O ¼ NH4 þ þ OH ; N2 H4 þ H2 O ¼ N2 H5 þ þ OH (2)

FeðOHÞ3 ¼ FeOðOHÞ þ H2 O (3)

12FeOðOHÞ þ N2 H4 ¼ 4Fe3 O4 þ 6H2 O þ N2 (4)


where the OH anion is an important intermediate product, which increased the pH value of the solution to form
Fe(OH)3. At the early stage, Fe(OH)3 was probably dehydrated into the steady FeO(OH). However, when the system’s
pressure reached a certain value (about 20 atm [20]), most of FeO(OH) was transformed to Fe3O4 phase. Compared
with the experiments of solution syntheses without any surfactants [13–15], the usage of organic surfactant in our
hydrothermal process had important roles in slowing the nucleation rate and preventing the particles aggregation.
Y.-h. Zheng et al. / Materials Research Bulletin 41 (2006) 525–529 529

4. Conclusions

In summary, the Fe3O4 nanoparticles with diameter of about 27 nm were prepared through a hydrothermal route in
the presence of AOT. The magnetic properties of the nanoparticles exhibited a superparamagnetic behavior at room
temperature. The hydrazine played important roles not only in increasing pH value to form Fe(OH)3, but also in the
transformation of Fe(OH)3 into ferromagnetic Fe3O4 nanoparticles.

Acknowledgement

This work was supported by grants from the Natural Science Foundation of Fujian Province China (Project No
A0320001), the Ministry of Science and Technology of China (Project No. 2003BA323C) and the State Key
Laboratory of Structural Chemistry of China (Project No. 050005).

References

[1] J.F. Wang, M.S. Gudiksen, X.F. Duan, Y. Cui, C.M. Lieber, Science 293 (2001) 1445.
[2] H. Kind, H.Q. Yan, B. Messer, M. Law, P.D. Yang, Adv. Mater. 14 (2002) 158.
[3] J.L. Dormann, in: J.L. Dorman, D. Fiorani (Eds.), Magnetic Properties of Fine Particles, North-Holland, Amsterdam, 1992.
[4] J. Popplewell, L. Sakhnini, J. Magn. Magn. Mater. 149 (1995) 72.
[5] K. Raj, B. Moskowitz, R. Casciari, J. Magn. Magn. Mater. 149 (1995) 174.
[6] D.K. Kim, Y. Zhang, J. Kehr, T. Klason, B. Bjelke, M. Muhammed, J. Magn. Magn. Mater. 225 (2001) 256.
[7] D.K. Kim, Y. Zhang, W. Voit, K.V. Rao, J. Kehr, T. Klason, B. Bjelke, M. Muhammed, Scripta Mater. 44 (2001) 1713.
[8] Y.J. Lee, J.W. Lee, C.J. Bae, J.G. Park, H.J. Noh, J.H. Park, T. Hyeon, Adv. Funct. Mater. 15 (2005) 503.
[9] R.S. Sapieszko, E. Matijevic, J. Colloid Interface Sci. 74 (1980) 405.
[10] J. Rockenberger, E.C. Scher, P.A. Alivisatos, J. Am. Chem. Soc. 121 (1999) 11595.
[11] T. Hyeon, S.S. Lee, J. Park, Y. Chung, H.B. Na, J. Am. Chem. Soc. 123 (2001) 12798.
[12] S.H. Sun, H. Zeng, J. Am. Chem. Soc. 124 (2002) 8204.
[13] R. Vijayakumar, Y. Koltypin, I. Felner, A. Gedanken, Mater. Sci. Eng. A286 (2000) 101.
[14] T. Fried, G. Shemer, G. Markovich, Adv. Mater. 13 (2001) 1158.
[15] Y.D. Li, H.W. Liao, Y.T. Qian, Mater. Res. Bull. 33 (1998) 841.
[16] L.E. Copeland, R.H. Bragg, Anal. Chem. 30 (1958) 196.
[17] J. Wang, Q.W. Chen, C. Zeng, B.Y. Hou, Adv. Mater. 16 (2004) 137.
[18] R.H. Kodama, A.E. Berkowitz, E.J. McNiff, S. Foner, Phys. Rev. Lett. 77 (1996) 394.
[19] A. Aharoni, J.P. Jakubovics, IEEE Trans. Magn. 24 (1988) 1892.
[20] X. Ni, X. Su, H. Zheng, D. Zhang, D. Yang, Q. Zhao, J. Cryst. Growth 275 (2005) 548.

You might also like