Download as docx, pdf, or txt
Download as docx, pdf, or txt
You are on page 1of 11

One-step synthesis of soft magnetic iron-based bimetallic

FeM (M= Co, Cu, Ni, and Ag) nanoalloy

T. M. Freire1, R. M. Freire2, 3, J. C. Denardim3, F. G. S. Oliveira4, I. F. Vasconcelos4,


P. L. Neto5, P. B. A. Fechine1

1
Group of Chemistry of Advanced Materials (GQMat)-Department of Analytical Chemistry and Physical-
Chemistry, Federal University of Ceará-UFC, Campus do Pici, CP 12100, CEP 60451-970 Fortaleza-CE,
Brazil
2
Institute of Applied Chemical Sciences, Universidad Autónoma de Chile, 8910060, Santiago, Chile
3
Department of Physical and CEDENNA, University of Santiago de Chile, USACH, Av. Ecuador 3493,
Santiago, Chile
4
Departamento de Engenharia Metalúrgica e de Materiais, Universidade Federal do Ceará, Campus do Pici,
60440-900 Fortaleza-CE, Brazil
5
Group of Electrochemistry and Corrosion (GELCORR)-Department of Analytical Chemistry and Physical-
Chemistry, Federal University of Ceará-UFC, Campus do Pici, CP 12100, CEP 60451-970 Fortaleza-CE,
Brazil

Abstract
Iron-based magnetic nanoalloys (FeM@OLA, M= Co, Cu, Ni, and Ag; OLA= oleylamine)
have been successfully synthesized via oleylamine reduction of metal salts method. FeCo is
arranged in bcc unit cell, while FeNi, FeAg and FeCu are in fcc structure. All the samples
have different morphologies with average size between 6.4 and 21.7 nm, depending on the
bimetallic composition. The nanoalloys have ferromagnetic behavior with low coercive
field and high saturation magnetization values at room temperature. In addition, it has been
shown that FeAg fit better in the solid solution category. The bimetallic nanoparticles
synthesized here could be ideal candidates for magnetic hyperthermia, magnetic imaging
contrast agents, catalisys and hydrogen evolution reaction.
Keywords: Bimetallic nanoparticles; Nanolloys; Thermal Decomposition; Magnetic
nanoparticles;

Bimetallic nanoalloys have unique characteristics that arise by the coupling of physico-
chemical properties of two different metals [1]. In this regard, iron-based bimetallic alloys
(IBBA) have attracted attention researchers due to its magnetic properties and its great
potential in the field of the catalyst [2], water treatment [3, 4] and biomedical [5]. In
general, these applications depend on the size and composition of the nanoparticles [6].
Thus, the development of facile and robust synthesis process of IBBAwith the capacity to
control the size and composition of the nanoparticles are of great interest since they can
provide a way to tune their properties and better their efficiency [7].

In the actual context of development of new technologies, soft magnetic materials such as
FeNi, FeCo, FeCu and FeAg have been widely applicated in data storage [8], catalysts [9]
and nanomedicine [10]. IBBA can be synthesized by diverse approaches, such as
electrochemical deposition [11], chemical reduction [12], polyol [13], sputtering process
[14] and mechanical alloying [15]. However, these techniques often provide a wide size
distribution, poor colloidal stability and aggregation [1]. On the other hand, oleylamine
(OLA) reduction method has been conducted to synthesize monodisperse bimettalic alloy
with narrow size distribution and good colloidal stability [16]. In addition, this method is
based in the decomposition of OLA under mild conditions, which decrease the cost of the
synthesis process.
In this sense, the present work shows a facile, fast and robust method to produce FeM (M=
Ag, Co, Cu and Ni) nanoalloys based in OLA reduction of metal salt. Besides, it is
performed a detailed study of the structure and magnetic properties of these systems. To the
best of our knowledge, this is the first work that reports the synthesis of FeCu and FeAg
nanoalloys by this synthetic route. Nanoalloys were synthesized by OLA reduction method
[16], where 0.25mmol of iron(III) acetylacetonate, cobalt(II) acetylacetonate and 10 ml of
OLA were mixed at room temperature and under N2 flow. After the formation of a
homogeneous solution, the system was heated at 100°C for 15 min. Then, the solution was
heated up to 300°C and kept for 1h. The dispersion was cooled down and the product was
isolated by magnetic separation and washed many times with a mixture of hexane/ethanol
solvents and dispersed in hexane. The synthesis of the other nanoalloys was performed in
the same way. However, for FeNi, FeCu, and FeAg samples the precursors were Nickel(II)
acetylacetonate, Copper(II) acetate, and Silver nitrate, respectively. Figure 1 shows the
procedures scheme and possible structure for FeM nanoalloy.
Figure 1: Scheme of the FeM nanoalloy synthesis by OLA reduction method.

The crystalline structure of the nanoalloys were investigated using a X-ray powder
diffraction (XRD) in a Bruker D2 mode X-ray diffractometer with cobalt Kα1 radiation (λ =
1.78890100 Å) operating in the range 10 ° to 100° with a rate of 0.02° min -1. The Perkin-
Elmer 2000 spectrophotometer was used to carry out Fourier-transform infrared
spectroscopy (FT-IR) analysis. For this purpose, the samples were dried, grounded and
pressed in KBr pellet. High-resolution transmission electron microscopy (HRTEM) was
performed on a HITACHI HT7700 operating at an accelerating voltage of 120 kV .
Previously, the nanoalloys were dispersed in hexane and one drop was placed on 300 mesh
carbon-coated copper grid and dried under ambient conditions . Mössbauer spectroscopy
were recorded at room temperature with a FAST (ConTec) Mössbauer system spectrometer
57
using transmission geometry and a ❑ Co radiative source. Magnetic properties were
investigated using a Vibrating Sample Magnetometer (VSM) Mini 5 Tesla from Cryogenic
Ltd.

The diffractograms of the synthesized FeM nanoalloys are shown in Fig 1a. For FeAg,
diffraction peaks can be assigned to the fcc phase (JCPDS 65-8448). The Braag reflections
were found to shift to lower 2θ angles, which suggests the formation of AgFe alloy
structures rich in Ag [17, 18]. For the samples FeNi, FeCo, and FeCu, the peaks confirm
the target nanostructures, since these diffraction patterns match with those in the standard
card of fcc FeNi (JCPDS 12-0736), bcc FeCo (JCPDS 65-4131) and fcc FeCu 4 (JCPDS 65-
7002). The additional diffraction peaks for FeCu samples were attributed to copper ferrite
(JCPDS 77-0010). Its presence suggests that iron and copper metals have been oxidized,
forming the spinel phase. Once the copper ferrite is not obtaiend in low temperature
conditions [19], it is plausible to infer the influence of substances from the during the
degradation OLA process. ss, iron and copper metals have been oxidized, since the copper
ferrite is not formed in low temperature [19].

Figure 2: XRD (a) and FTIR (b) analysis of the samples AgFe, FeNi, FeCo, and FeCu.
To confirm that OLA is on the nanoalloys surface, FT-IR spectra were acquired and the
results are shown in Figure 1b. For all the samples, it was observed characteristic
vibrational modes of alkyl chains from OLA (2920 and 2848 cm -1). They are assigned to
the methylene asymmetric and symmetric C – H stretching, while the band at 1627 cm -1 can
be assigned to C = C stretch mode, while the band at 1456 cm -1 corresponds to C – H
bending mode [20]. Moreover, it was also found signals related to amine group: a broad
band at 3440 cm-1 due to N – H stretching of the primary amine, a mode at 1554 cm -1 due to
–NH2 scissoring, and the mode around 1382 cm -1 attributed to C – N stretching [21, 22].
Further, it was observed a broad band around 580 cm -1, corresponding to Fe – O stretching
and suggesting a surface passivation through the formation of iron oxide [22]. Indeed, it is
reasonable to consider that nanoalloys were capped by OLA (FeM@OLA).
TEM image was carried out to analyze the morphology and size distribution of the
synthesized bimetallic nanoalloy (Fig. 3 and S1 in Supplementary Material (SM)). For
FeNi (Fig. 3a and b) and FeCo (Fig. 3c and d), the micrographs revealed these nanoalloys
have more than one morphology. It was observed nanoparticles with spherical or irregular-
shaped morphology for FeNi sample, while for FeCo was seen nanoparticles with cubic,
rod and spherical shapes. In addition, the inset in Figure 3c shows a cubic nanoparticle of
FeCo with a shell thickness of 1.8 nm, which can be assigned to the formation of iron oxide
on the surface of the nanoparticles. The histograms shown in the Figure S1 c and f reveal
that both bimetallic alloys have a narrow size distribution with an average size of 21.7 ∓
5.1 and 15.1∓ 2.3 nm for FeNi and FeCo, respectively. For FeAg sample (Fig. 3e and f),
TEM images showed a diversified morphology with the formation of triangular plates,
spherical and octahedral type structures. In addition, this nanoalloy has a bimodal size
distribution with average sizes of 6.5 ∓ 1.2 and 18.8 ∓ 2.5 nm (Figure S1 i, SM). The inset
of the Fig. 3f shows the lattice fringes of 2.06 Å , corresponding to the (200) planes of the
AgFe phase observed in XRD. Indeed, this result suggests that the smaller nanoparticles in
the TEM images are not isolated Ag nanoparticles. FeCu sample shows a bimodal size
distribution, wherein the average sizes are 6.4 ∓ 1.0 and 20.8 ∓ 4.1 nm (Figure S1 l, SM).
Besides, the larger nanoalloys have a shell thickness of 2.5 nm. The bimodal size
distribution for FeAg and FeCu system may be assigned to the difference between the
reduction potential of these metals, which a higher rate of reduction and nucleation for Ag
and Cu compared to Fe [1].

Figure 3: TEM image of FeNi (a and b), FeCo (c and d) FeAg (e and f) and FeCu (g and
h).
Figure 3a and 3b show the magnetization versus applied magnetic field curves at room
temperature (300K) and 5K for the synthesized nanoalloys. The saturation magnetization
(Ms) at room temperature for FeCo, FeNi, FeCu, and FeAg samples were found to be equal
to 80.66, 65.09, 17.86, and 6.96 emu g -1, respectively. Furthermore, the nanoalloys exhibit
ferromagnetic behavior with a hysteresis loop not reversible and a coercive field (Hc) of 98,
168, 36.1, and 65.1 Oe for FeCo, FeNi, FeCu, and FeAg, respectively. These results
suggest that all nanoalloys have soft magnetic characteristics since the curves are narrow,
with low values of the coercivity and high value of the Ms. On the other hand, the
nanoalloy exhibit a large hysteresis at low temperature (5K), with Hc values (1325.6, 429.2,
517.1 and 446.8 Oe for FeCo, FeNi, FeCu and FeAg, repsctively) higher than those found
at 300K, deducing that there is an increase of nanoparticle blocking in this temperature
[23]. In addition, it is important to note that Hc for FeCo at 5K is larger than room
temperature by about 1227 Oe, suggesting the presence of high anisotropy phase such as
CoFe2O4 [24, 25]. On the other hande, the increase for FeCu is not significant and can be
attributed to the presence of CuFe2O4 [26].
Figure 3: (a) and (b) Magnetization versus applied magnetic field curves at room
temperature (300 K) and 5K for the synthesized nanoalloys. (c) and (d) are room
temperature Mössbauer spectroscopy and magnetic hyperfine field distribution,
respectively, for samples FeAg, FeCo, FeCu and FeNi.

Figure 3c and 3d present the Mössbauer spectra for FeCo, FeNi, FeCu, and FeAg samples
and the fit by hyperfine fields (B hf) distributions, respectively. For FeCo, a sextet
distribution was identified with maximum at approximately 34.6 T, which suggests the
presence of Co atoms in the bcc Fe structure [27, 28]. Furthermore, the obtained values of
the isomer shift (δ) = 0.03 mm s -1 and quadrupole splitting (Δ) = 0.01 mm s -1 (see Table 1)
indicates the formation of the FeCo phase with bcc structure, which is consistent with the
analysis of XRD. However, it is also observed the presence of a doublet with δ of 0.30 mm
s-1 and Δ of 0.91 mm s-1, which can be assigned to ferric ions from the Fe-oxide surface
shell of FeCo nanoparticles [25, 29]. Futhermore, TEM images and the superparamagnetic
character due to the reduced size of the particles are in aggrement with this result. Finally,
the third contribution of the spectrum is a singlet (violet line) with a high-line width value
that can be associated with a superparamagnetic incomplete contribution [29]. The
Mössbauer spectrum for FeNi exhibit two contributions. The first is a B hf distribution
centered on 29.7 T that corresponds to the formation of fcc γ-(Fe, Ni) phase atomically
disordered [30], which according to Johnson et al., starts to form at Ni concentrations above
28% [28]. Besides, it was observed a doublet with δ of 0.37 mm s -1, which can be attributed
to the superparamagnetic behavior due to the small size of the particles as well as the
presence of Fe3+ from the Fe-oxide layer [31]. Furthermore, it is interesting to note the
dominance of the field distribution in samples FeCo and FeNi with 56.3, and 90.5 %
relative spectral area (A). Indicating that Fe atoms are found mainly in ferromagnetic sites.
Therefore, the Mössbauer results for the FeCo and FeNi samples is in accordance with the
results from the magnetization measurements (Fig. 3 a and b) wherein both nanoalloys
showed low magnetic hysteresis at room temperature that can be related to the
superparamagnetic behavior identified.

Table 1: Mössbauer parameters of the FeM nanoalloyings.


Mössbauer parameters
Sample Fiting <Bhf> / T δ / mm/s ∆ / mm/s A/ %
Distribution (Green) 17.4 0.02 0.05 29.5
FeCu
Dublet (Blue) __ 0.33 0.82 70.5
Distribution (Green) 34.6 0.03 0.01 56.3
FeCo Doublet (Blue) 0.30 0.91 9.7
Singlet (Orange) 0.37 34.0
Distribution (Green) 29.7 0.05 0.03 90.5
FeNi
Doublet (Blue) 0.37 -0.95 9.5
Distribution (Green) 20.0 0.12 0.06 73.8
FeAg
Doublet (Blue) 0.36 1.02 26.1
The effect of diamagnetic atoms, such as Cu and Ag on the FeM nanoalloys was also
investigated by Mössbauer spectroscopy. Fig. 3c shows the spectrum for FeCu samples,
revealing two contributions, a doublet that can be assigned to fcc FeCu rich phase and the
superparamagnetic phase of CuFe2O4, with Fe3+ ions in tetrahedral sites [32, 33]. The sextet
fitting by distribution between 0 and 25 T, suggests that Fe atoms are fully incorpored into
the Cu lattice formatting a second fcc FeCu nanoalloy more rich in Fe [33]. Indeed, the
Mössbaues analyses for FeCu are in accordance with XRD and VSM analysis. For FeAg,
the distribution of Bhf centered on 20 T is characteristic of the Fe-rich bcc structure, in
which the presence of Ag in the structure of α-Fe decrease the Bhf due to the diamagnetic
characteristic of the silver [34]. However, the doublet with high Δ values indicates the
presence of an electric field gradient (EFG) that can be due to the presence of Ag atoms
with greater electronic density around Fe atoms [34]. Thus, the doublet suggests the
formation of a solid FeAg fcc structure rich in Ag [35]. The low miscibility between Ag
and Fe suggests that fcc structure arise from Fe diffusion in Ag nanoparticles, while bcc
structure can be devivated from Ag diffusion in Fe nanoparticles [1, 36] [37]. In addition,
to the best of our knowledge, this is the first work that report to formation of FeAg solid
solution by co-reduction method.

In summary, we have successfully synthesized the FeM (M = Ag, Co, Cu, and Ni)
nanoalloys with a mixture of the morphologies, but with narrow size distribution. XRD,
TEM, and Mössbauer analyses confirm the obtention of bimetallic alloys. FeCo is arranged
in bcc structure, FeNi and FeCu are in fcc structure and FeAg is arranged in two different
structure being a bcc and other fcc. FTIR reveals that all nanoparticles are coated by OLA
(FeM@OLA). Magnetic characterization of the nanoalloys showed that it have a
ferrimagnetic behavior, with high saturation magnetization and low coercive field. In
addition, for FeAg this synthesis method open a real possibility to control the morphology
in nanoscale, since that this alloy was formed without the use of auxiliary reagents to
control shape. Indeed, leave account all aspects presented in this work such as the variated
morphology and formation of solid solution, the nanoalloys synthesized using the
simplified method are sustentable for various application such as hydrogen evolution
reaction (HER), catalyst, magnetic separators and electro-magnetic energy transformation.
This work was supported by Brazilian agencies: CAPES (Finance Code 001), CNPq
(408790/2016-4), Funcap (PNE-0112-00048.01.00/16), and Chilean agencies Basal
CEDENNA AFB180001, ANID/CONICYT/Fondecyt 1200782.

References

[1] K.D. Gilroy, A. Ruditskiy, H.-C. Peng, D. Qin, Y. Xia, Chemical Reviews 116(18) (2016) 10414-
10472.
[2] B. Liu, Y. Xu, S. Zhang, X. Lv, Y. Ling, Z. Zhou, J. Liu, Materials Letters 239 (2019) 124-127.
[3] C.-p. Tso, D.T.F. Kuo, Y.-h. Shih, Chemosphere 250 (2020) 126155.
[4] G.I. Gunarani, A.B. Raman, J. Dilip Kumar, S. Natarajan, G.B. Jegadeesan, Materials Letters 247
(2019) 90-94.
[5] Y. Liu, P.-C. Wu, S. Guo, P.-T. Chou, C. Deng, S.-W. Chou, Z. Yuan, T.-M. Liu, Photoacoustics 19
(2020) 100179.
[6] J. Kwon, X. Mao, J. Lee, Current Applied Physics 17(8) (2017) 1066-1078.
[7] L. Wu, A. Mendoza-Garcia, Q. Li, S. Sun, Chemical Reviews 116(18) (2016) 10473-10512.
[8] M. Méndez, V. Vega, S. González, R. Caballero-Flores, J. García, V.M. Prida, Nanomaterials 8(8)
(2018).
[9] H. Zhang, Y. Feng, Y. Cheng, M.D. Baró, A. Altube, E. García-Lecina, F. Alcaide, E. Pellicer, T.
Zhang, J. Sort, ACS Omega 2(2) (2017) 653-662.
[10] Z. Chen, G. Hong, H. Wang, K. Welsher, S.M. Tabakman, S.P. Sherlock, J.T. Robinson, Y. Liang,
H. Dai, ACS Nano 6(2) (2012) 1094-1101.
[11] Y. Sun, G. Zangari, Inorganic Chemistry 59(8) (2020) 5405-5417.
[12] A. Ahmad, A.S. qureshi, L. Li, J. Bao, X. Jia, Y. Xu, X. Guo, Colloids and Surfaces B: Biointerfaces
143 (2016) 490-498.
[13] F.F. Barbosa, S.B.C. Pergher, T.P. Braga, Journal of Alloys and Compounds 772 (2019) 625-636.
[14] L. Cabral, F.H. Aragón, L. Villegas-Lelovsky, M.P. Lima, W.A.A. Macedo, J.L.F. Da Silva, ACS
Applied Materials & Interfaces 11(1) (2019) 1529-1537.
[15] H. Moumeni, S. Alleg, J.M. Greneche, Journal of Alloys and Compounds 386(1) (2005) 12-19.
[16] Y. Yu, W. Yang, X. Sun, W. Zhu, X.Z. Li, D.J. Sellmyer, S. Sun, Nano Letters 14(5) (2014) 2778-
2782.
[17] C. Jin, Z. He, Y. Zhao, Y. Pan, W. Wu, X. Wang, G. Tong, CrystEngComm 20(14) (2018) 1997-
2009.
[18] R.M. Sarhan, G.A. El-Nagar, A. Abouserie, C. Roth, ACS Sustainable Chemistry & Engineering
7(4) (2019) 4335-4342.
[19] N. Masunga, O.K. Mmelesi, K.K. Kefeni, B.B. Mamba, Journal of Environmental Chemical
Engineering 7(3) (2019) 103179.
[20] S. Mourdikoudis, L.M. Liz-Marzán, Chemistry of Materials 25(9) (2013) 1465-1476.
[21] L. Zhang, L. Wang, Z. Jiang, Z. Xie, Nanoscale Research Letters 7(1) (2012) 312.
[22] M. Chen, Y.-G. Feng, X. Wang, T.-C. Li, J.-Y. Zhang, D.-J. Qian, Langmuir 23(10) (2007) 5296-
5304.
[23] H. Sharifi Dehsari, K. Asadi, The Journal of Physical Chemistry C 122(51) (2018) 29106-29121.
[24] S. MOURDIKOUDIS, K. SIMEONIDIS, M. ANGELAKERIS, I. TSIAOUSSIS, O. KALOGIROU, C.
DESVAUX, C. AMIENS, B. CHAUDRET, Modern Physics Letters B 21(18) (2007) 1161-1168.
[25] P. Rajesh, J.-M. Greneche, G.A. Jacob, T. Arun, R.J. Joseyphus, physica status solidi (a) 216(18)
(2019) 1900051.
[26] B. Mondal, M. Kundu, S.P. Mandal, R. Saha, U.K. Roy, A. Roychowdhury, D. Das, ACS Omega
4(9) (2019) 13845-13852.
[27] M.F. Casula, G. Concas, F. Congiu, A. Corrias, A. Falqui, G. Spano, The Journal of Physical
Chemistry B 109(50) (2005) 23888-23895.
[28] C.E. Johnson, M.S. Ridout, T.E. Cranshaw, Proceedings of the Physical Society 81(6) (1963)
1079-1090.
[29] B. Babić-Stojić, V. Jokanović, D. Milivojević, Z. Jagličić, D. Makovec, N. Jović, M. Marinović-
Cincović, Journal of Nanomaterials 2013 (2013) 741036.
[30] V.A.P. Rodríguez, C. Rojas-Ayala, J.M. Medina, P.P. Cabrera, J. Quispe-Marcatoma, C.V.
Landauro, J.R. Tapia, E.M. Baggio-Saitovitch, E.C. Passamani, Materials Characterization 149 (2019)
249-254.
[31] V. Mancier, J.-L. Delplancke, J. Delwiche, M.-J. Hubin-Franskin, C. Piquer, L. Rebbouh, F.
Grandjean, Journal of Magnetism and Magnetic Materials 281(1) (2004) 27-35.
[32] D.H. Choi, I.-B. Shim, C.S. Kim, Journal of Magnetism and Magnetic Materials 320(20) (2008)
e575-e577.
[33] P. Crespo, N. Menéndez, J.D. Tornero, M.J. Barro, J.M. Barandiarán, A. Garcı́a Escorial, A.
Hernando, Acta Materialia 46(12) (1998) 4161-4166.
[34] G. Rixecker, Solid State Communications 122(6) (2002) 299-302.
[35] F. Spizzo, E. Angeli, D. Bisero, A. Da Re, F. Ronconi, P. Vavassori, Journal of Magnetism and
Magnetic Materials 272-276 (2004) 1169-1170.
[36] K. Santhi, T.A. Revathy, V. Narayanan, A. Stephen, Applied Surface Science 316 (2014) 491-
496.
[37] R. Ferrando, J. Jellinek, R.L. Johnston, Chemical Reviews 108(3) (2008) 845-910.

You might also like