Brought to you by:
Paper

From core/shell to hollow Fe/γ-Fe2O3 nanoparticles: evolution of the magnetic behavior

, , , , and

Published 17 September 2015 © 2015 IOP Publishing Ltd
, , Citation Z Nemati et al 2015 Nanotechnology 26 405705 DOI 10.1088/0957-4484/26/40/405705

0957-4484/26/40/405705

Abstract

High quality Fe/γ-Fe2O3 core/shell, core/void/shell, and hollow nanoparticles with two different sizes of 8 and 12 nm were synthesized, and the effect of morphology, surface and finite-size effects on their magnetic properties including the exchange bias (EB) effect were systematically investigated. We find a general trend for both systems that as the morphology changes from core/shell to core/void/shell, the magnetization of the system decays and inter-particle interactions become weaker, while the effective anisotropy and the EB effect increase. The changes are more drastic when the nanoparticles become completely hollow. Noticeably, the morphological change from core/shell to hollow increases the mean blocking temperature for the 12 nm particles but decreases for the 8 nm particles. The low-temperature magnetic behavior of the 12 nm particles changes from a collective super-spin-glass system mediated by dipolar interactions for the core/shell nanoparticles to a frustrated cluster glass-like state for the shell nanograins in the hollow morphology. On the other hand for the 8 nm nanoparticles core/shell and hollow particles the magnetic behavior is more similar, and a conventional spin glass-like transition is obtained at low temperatures. In the case of the hollow nanoparticles, the coupling between the inner and outer spin layers in the shell gives rise to an enhanced EB effect, which increases with increasing shell thickness. This indicates that the morphology of the shell plays a crucial role in this kind of exchange-biased systems.

Export citation and abstract BibTeX RIS

1. Introduction

Magnetic nanoparticles have been studied for some time now because their unique properties lead to a broad range of applications [15]. Recently, focus has shifted to improving the chemical and physical properties of magnetic nanoparticles to optimize their functionality introducing a secondary phase [6, 7]. For instance, Fe–Fe oxide core–shell nanoparticles are popular candidates for biomedical applications because the high magnetization of the core and the chemical stability and biocompatibility of the shell lead to better overall properties than either material alone [810]. This type of composite approach has led to a growing number of studies on magnetic core–shell nanoparticles [1116].

An interesting byproduct of these systems is the exchange coupling across the core–shell interface that is frequently seen in the form of exchange bias (EB), a horizontal shift in the hysteresis loop accompanied by an increase in coercivity after cooling in a magnetic field [17, 18], a well-known phenomenon observed in different nanostructures, such as ferromagnet−antiferromagnet (FM-AFM), ferromagnet/spin glass, ferrimagnet/ferrimagnet, etc [19]. The exploration of EB on the core–shell nanoparticles has been proposed to be a promising approach to overcoming the superparamagnetic (SPM) limit in nanoparticles, a critical bottleneck for magnetic data storage applications [20]. The EB has been reported in a large class of core–shell nanoparticles of Fe–C [11], Co–CoO [20, 21], CrO2–Cr2O3 [22], FeO–Fe3O4 [12, 13, 23], MnO–Mn3O4 [24], Fe/Fe3O4 [25, 26], Fe/γ-Fe2O3 [16, 27], CoO/γ-Fe2O3 [28], and Fe3O4/γ-Fe2O3 [29]. However, the high surface-to-volume ratio of nanoparticles resulting in a 'shell' of disordered surface spins has also been shown to be responsible for inducing EB in single component magnetic nanoparticles (NiFe2O4 [30], γ-Fe2O3 [31], CoFe2O4 [32]). The origin of this behavior is the fraction of surface spins with decreased co-ordination (and thus weaker bonding) increasing with decrease in particle size. These disordered spins can take on a number of configurations, one of which can be chosen by field-cooling the particle to induce an EB [3032]. An example of this is the case of γ-Fe2O3 nanoparticles, where the degree of disorder of the surface spins of the shell layer to which the ferrimagnetically ordered spins of the central layer are coupled has been shown to be crucial for achieving EB [29]. As a result, effects of both uncompensated interface and surface spins must be taken into account for the EB effects recently observed in Fe/γ-Fe2O3 [16, 27], CoO/γ-Fe2O3 [28], and Fe3O4/γ-Fe2O3 [29] core/shell nanoparticle systems. In an attempt to decouple collective contributions of the interface and surface spin effects to the EB in such core/shell systems, we have recently performed a comparative study of the magnetic properties and EB effect in Fe/γ-Fe2O3 core–shell nanoparticles with the same thickness of the γ-Fe2O3 shell (∼2 nm) while varying the diameter of the Fe core from 4 to 11 nm [27]. Our study has shown that there exists a critical particle size (mean size, ∼10 nm), above which the interface spin effect contributes mainly to the EB, but below which the surface spin effect is dominant. This finding points to an importance of the finite-size effect. It has been noted that in a core/shell nanoparticle composed of two different materials the oxidization-driven migration of metal atoms from the core to the shell is likely to occur (via the so-called Kirkendall effect), thus producing vacancies at the core/shell interface that gradually coalesce into voids [25, 26]. A more detailed description of the void formation can be found in [25, 26]. The Kirkendall effect has recently been utilized for transforming core/shell nanoparticles into hollow nanostructures, both of which hold great potential for applications in memristors, hyperthermia therapy in nanomedicine, and targeted drug delivery [3335]. On the other hand, it is suggested that the presence of voids at the core/shell interface influences the coupling between interface spins and hence the magnetic properties [25, 33]. However, no detailed study has been performed to address this hypothesis. In the case of hollow nanoparticles, the presence of additional inner surfaces has also been suggested to contribute to the enhanced spin disorder which gives rise to a higher surface anisotropy and consequently an increased EB effect [33, 35]. The collective magnetic behavior and EB effect (below the blocking temperature) have been reported to differ largely between the core/shell and hollow nanoparticle systems, but the reason for this has remained an open question [25, 27, 35]. Ong et al [26] explicitly compared changes in the EB effect between Fe/Fe3O4 core/shell and Fe3O4 hollow nanoparticles with a size of 14–16 nm, but no attempt was made to systematically study the spin dynamics, the effect of size variation, or the intermediate core-void-shell structure. To address these important issues, it is essential to investigate how the magnetic properties of a core/shell nanoparticle system are modified when the core/shell morphology transforms into the core/void/shell and the hollow structure.

While our previous studies focused on either the effect of surface and interface spins on the static magnetic properties of Fe/γ-Fe2O3 core/shell and γ-Fe2O3 hollow nanoparticles [9, 13, 16] or the dynamic magnetic properties of a Fe/γ-Fe2O3 core/shell nanoparticle system with one particular size (9 nm) [16], in this article we report the first systematic study of the evolution of the intraparticle- and interparticle-influenced magnetic properties of these nanoparticles, during their 'hollowfication' process, for two different size regimes (8 and 12 nm). The particle sizes were chosen to be above and below a critical value (d ∼ 10 nm) [27], in order to probe finite-size effects on the magnetism of a particle system with various morphologies. We demonstrate that a transformation in morphology from the core/shell to hollow nanoparticles resulted in a strong modification in the magnetization dynamics and EB effect. A distinctly different magnetic behavior has been observed for the 12 and 8 nm particles. These are attributed to the formation of voids in the core/void/shell structures, the appearance of additional inner disordered surface spins in the hollow structures, and the finite-size effect related to the ferrimagnetic character of maghemite and the unbalanced number of spins in an antiparallel arrangement. Since the diagnostic and therapeutic advantages of the Fe/γFe2O3 core/shell and γFe2O3 hollow nanoparticles stem from their static and dynamic magnetic properties along with their ability to impact cell-specific functionality [810], our study not only resolves the diverse results reported in previous works [2529, 3335], but also paves the way for tuning magnetic anisotropy in exchange-coupled magnetic nanostructures for applications in advanced hyperthermia [36, 37] and spintronics [20].

2. Experiment

As we noted above in the Introduction, the formation of the hollow nanoparticles is related to the nanoscale Kirkendall effect, in which the oxidation-driven migration of metal atoms from the core to the shell produces vacancies at the interface that gradually coalesce into voids. A qualitative depiction of how a core/shell nanoparticle transforms into the core/voice/shell and the hollow morphology has been reported by Ong et al [25] and by Jaffari et al [33]. In the present study, Fe/γ-Fe2O3 core/shell nanoparticles were synthesized by thermal decomposition of organometallic compounds, details of which have been reported elsewhere [9, 27]. Briefly, a three neck flask was charged with oleylamine, 70%, and 1-octadecene, 90%, and the mixture was stirred at 140 °C under a mixture of 95% Ar + 5% H2 gases for 2 h. The temperature was raised subsequently to 220 °C and iron pentacarbonyl, Fe (CO2)5, was injected at 220 °C and refluxed for 20 min. After injection, the iron pentacarbonyl immediately decomposed into iron fragments which are the onset for nanoparticle formation (black precipitate), acetone and/or CO gas formed in the reaction vessel (white smoke), and the reaction temperature raised a few degrees because of its exothermic nature. The sample was cooled down to room temperature and a sufficient amount of nanoparticles was removed for characterization. The average particle size of the core/shell nanoparticles was varied by varying injection temperature. After 4 weeks, the samples presented a discernible core/void/shell morphology. Hollow nanoparticles are often produced by further oxidizing their core/shell counterparts which became hollow via the so-called Kirkendall effect [38, 39]. To create the 12 nm hollow particles, the core/shell sample was annealed at 180 °C for one hour under a flow of oxygen. For the case of the 8 nm particles, however, the core/shell structures naturally became hollow at room temperature without annealing just after four weeks. Both core/shell and hollow nanoparticles were washed with a mixture of 3 ml hexane, 95%, and 97 ml ethanol, ≥99.5%. The structural and microstructural studies were performed using TECNAI F20 transmission electron microscope (TEM) along with high-resolution TEM (HRTEM) and selected area electron diffraction (SAED). The magnetic properties were measured using a quantum design physical properties measurement system (PPMS) with vibrating sample magnetometer (VSM) and ac-susceptometer options over a range of temperatures between 5 and 300 K, and applied fields up to 50 kOe.

3. Results and discussion

3.1. Structural and morphological characterization

Figures 1(a)–(c) show conventional bright-field TEM images of the 12 nm core/shell, core/void/shell, and hollow nanoparticles (scale bars are 20 nm), along with a representative histogram of the particle size population as observed from the TEM image of the core/shell nanoparticles (inset of figure 1(a)). The size distribution results for the other nanoparticles are very similar. Contrast variation in the center of the core/shell nanoparticles clearly indicates that with increasing time, the core of the nanoparticles becomes progressively smaller until it finally disappears. A clear transformation in the morphology from core/shell to core/void/shell and to hollow is seen in the TEM images of figures 1(a)–(c). The average particle diameter, core diameter, and shell thickness of the 12 and 8 nm core/shell nanoparticles before and after transforming into the core/void/shell and hollow structures are listed in table 1. It is worth noting herein that both the core size and the shell thickness were altered during the core/shell to hollow transformation process. For both the 8 and 12 nm nanoparticles, the average particle diameter increases considerably when transforming from the core/shell to hollow morphology. Under this transformation, the shell thickness increases significantly for the case of the 12 nm particles, while it remains almost unchanged for the case of the 8 nm particles. These morphological changes are shown below to significantly influence the static and dynamic magnetic properties of the nanoparticles.

Figure 1.

Figure 1. Bright-field TEM images of the 12 nm Fe/γ-Fe2O3 (a) core/shell, (b) core/void/shell, and (c) hollow nanoparticles; inset of figure 1(a) shows a histogram of the particle size populations for the 12 nm core/shell nanoparticles and inset of figure 1(c) shows SAED pattern of hollow nanoparticles. HRTEM images of (d) core/shell, (e) core/void/shell and (f) hollow nanoparticles. The scale bar is 20 nm in figures (a)–(c) and is 5 nm in figures (d)–(f). The discontinuous lines in '(f)' show grain boundaries of nanograins in hollow nanoparticles.

Standard image High-resolution image

Table 1.  Nanoparticle diameter, core diameter and shell thickness, as determined by TEM for the two batches of samples analyzed.

  D (nm) Core (nm) Shell (nm)
C/S (12 nm) 11.6(5) 7.1(4) 2.2(1)
C/V/S (12 nm) 12.3(6) 4.7(2) 2.1(1)
H (12 nm) 14.7(7) 3.2(2)
C/S (8 nm) 7.5(4) 3.0(1) 2.2(1)
H (8 nm) 9.4(5) 1.9(1)

The SAED pattern of the 12 nm hollow nanoparticles is presented in the inset to figure 1(c), and is indexed to be fcc iron-oxide. HRTEM images of core/shell, core/void/shell, and hollow nanoparticles are respectively displayed in figures 1(d)–(f). These HRTEM images reveal the crystalline structure of both core and shell with lattice spacing of 2.02 Å for the core and 2.52 Å for the shell corresponding to (110) planes of bcc iron and (311) planes of fcc iron oxide phase, respectively. The Fe core is single crystalline, however, the shell of γ-Fe2O3 is composed of small crystallites which are oriented randomly, as described in figure 1(f). In addition, we have recently performed x-ray absorption near edge spectroscopy (XANES) on these samples, and the fitting results (not shown here) corroborate that the shell is mostly formed by maghemite, while the core is composed of Fe.

3.2. Magnetic properties

In this section we first present and discuss the magnetic data of the 12 nm core/shell, core/void/shell, and hollow nanoparticles. The temperature dependence of magnetization (MT) was measured under the zero-field-cooled (ZFC) and field-cooled (FC) protocols in a field of 100 Oe for all samples investigated. Figure 2(a) shows the ZFC MT curves measured for the three samples. These curves have been normalized to Mmax for comparison. Clearly, the ZFC MT curves exhibit a typical peak, TP-ZFC, which displaces towards a lower temperature when the morphology changes from core/shell to core/void/shell and hollow (table 2). During this process, the mass normalized magnetization progressively decreases, and the ZFC and FC magnetization branches become more separated, with the irreversibility, Tirr, taking place at increasingly higher temperatures (see figure 2(b) and its inset). At low temperatures T < TP-ZFC, the FC MT curves show different behaviors; for the core/shell sample the FC magnetization first decreased with lowering temperature just below TP-ZFC and then became almost unchanged, while for the hollow sample the FC magnetization first increased and then remained constant. These results clearly indicate that the thermal dependent magnetic behavior of the 12 nm particles was largely altered as the core/shell morphology became progressively hollow.

Figure 2.

Figure 2. (a) Normalized ZFC MT curves for the 12 nm core/shell, core/void/shell, and hollow nanoparticles; (b) ZFC and FC MT curves for the core/shell and hollow (see inset) nanoparticles, together with the derivative curve that defines 〈TB〉; (c) inverse susceptibility curves with their corresponding Curie–Weiss fits at high temperatures; (d) the fit of the ZFC MT curve for the core/shell nanoparticles using an expression based on the Stoner–Wolhfarth model.

Standard image High-resolution image

Table 2.  ZFC peak position, average blocking temperature, effective anisotropy and Curie temperature as determined from the ZFC/FC curves for the 12 nm nanoparticles.

  TP-ZFC (K) TB〉 (K) K (erg cm−3) TC (K)
C/S 111(5) 73(5) 1.5(2) × 106 59(5)
C/V/S 94(5) 65(5) 4.1(8) × 106 11(5)
H 60(5) 47(5) 9(2) × 106 −320(9)

It is generally accepted that for an ensemble of non-interacting, monodisperse nanoparticles, the peak in a ZFC MT curve (TP-ZFC) is referred as to the mean blocking temperature (TB), and is defined as the temperature at which the relaxation time of the nanoparticles, τ, is equal to the measurement time of the system, tm. For SPM non-interacting nanoparticles, the relaxation time is given by the Néel expression τ = τ0 exp(KV/kBT), being K the anisotropy constant of the nanoparticles, V the volume of each nanoparticle, and ${{\rm{\tau }}}_{0}$ is related to the gyromagnetic precession (10−9 − 10−10 s). In the particular case of a VSM, the measurement time can be considered to be tm ∼ 100 s and therefore TB ≈ KV/25kB. In this kind of system, Tirr often takes place near TP-ZFC, and the FC magnetization continuously increases with decreasing temperature. This clearly differs from the behavior observed for our nanoparticles (figure 2), suggesting that the energy barrier is being significantly affected by the different nature of the core and the shell, inter-particle and intra-particle interactions, surface disorder, and finite-size distribution [16]. Therefore, TP-ZFC may not represent the true blocking temperature of the presently studied nanoparticles. To be more precise, we have defined the mean blocking temperature, 〈TB〉, as the temperature at which the maximum number of nanoparticles enters a blocked state as temperature decreases. Taking this into account, 〈TB〉 can be easily determined from the peak position in the d(MFC-MZFC)/dT [40], as illustrated in figure 2(b). The values of 〈TB〉 of the samples are summarized in table 2. According to the Neel–Arrhenius magnetic relaxation model, 〈TB〉 of a particle system is proportional to the average anisotropy and volume of the nanoparticles, 〈TB〉 = KV/25kB. This would lead to a general expectation that 〈TB〉 increases as the nanoparticles transform from the core/shell to hollow morphology, since the total volume effectively increases (from 800 to 1400 nm3). But, unlike the case of nanoparticles made up of one kind of material, for our nanoparticles we have to consider the individual core (Fe) and the nanograins (γ-Fe2O3) in the shell as different magnetic entities with different blocking temperatures [16, 26]. Since the nanograins in the shell are not spherical and possess a considerable size distribution (figure 1(d)), an accurate determination of the average grain size is not easy [26, 34, 35]. However, we can suppose, in a first approximation, that the average grain size is close to the shell thickness. Using these values and the previously obtained 〈TB〉, the effective anisotropy of the γ-Fe2O3 hollow nanoparticles is estimated to be ∼9.5×106 erg cm−3, which is two orders of magnitude higher than that of bulk maghemite (4.5×104 erg cm−3) and compares well with the values reported for these kinds of systems [35, 41]. It is also noted that the anisotropy axis in each individual crystallite can lead to a formation of multiple magnetic domains in the shell. For the core/shell and core/void/shell nanoparticles, the analysis becomes even more cumbersome, because of the presence of the core, with its own anisotropy constant, that will also affect the effective anisotropy of the shell due to core/shell interactions. For the core/shell and core/void/shell nanoparticles, if we assume that, due to the higher magnetization of the Fe core, we are mostly seeing the blocking of the Fe cores at 〈TB〉, the effective anisotropy is determined to be ∼1.5×106 erg cm−3 and ∼4.1×106 erg cm−3, respectively, which is one order of magnitude larger than that of bulk Fe (∼5×105 erg cm−3). Since we are not considering the contribution of the shell grains to the anisotropy, these values should be regarded as purely qualitative estimations.

As we noted above, there are some distinct differences in the FC MT behavior between the core/shell and hollow nanoparticles (figure 2(b)). For the hollow nanoparticles, a large separation between the ZFC and FC magnetization branches can be related to the size distribution of the nanograins inside the shell, the presence of large exchange anisotropy, and the effects of inter-particle and intra-particle interactions. For the core/shell nanoparticles, the FC magnetization abruptly decreases (ΔFC = 23%) from TP-ZFC down to 50 K, below which it remains almost constant. This decrease of the FC magnetization in systems of interacting solid nanoparticles at low temperatures has been attributed to the onset of a collective glassy behavior and/or to the freezing of surface spins [31, 42, 43]. On the other hand, for the hollow nanoparticles the FC magnetization keeps increasing below TP-ZFC and reaches a maximum around 40 K, below which it experiences a much smaller decrease (ΔFC = 0.5%). This behavior is more similar to that reported for non-interacting SPM systems.

The strength and nature of magnetic interactions in the present nanoparticles can be further analyzed by fitting the ZFC−FC curves for T > TP-ZFC to a Curie–Weiss model. It is well known that in the SPM regime (T > 〈TB〉), the magnetic susceptibility of a particle system can be typically described by a Curie–Weiss law:

Equation (1)

where C is a constant, which depends on the magnetic moment of the nanoparticles, and TC is a temperature whose absolute value can be considered as a measure of the strength of interparticle interactions. As can be seen in figure 2(c), reasonable fits have been obtained for all three samples in the high temperature range. The TC values are determined to be 59, 11, and −320 K, for the core/shell, core/void/shell, and hollow nanoparticles, respectively. The change in sign of TC could suggest some considerable change in the magnetic behavior of the nanoparticles when the core/shell morphology was changed into the hollow structure. For the core/shell nanoparticles, the positive sign indicates that ferromagnetic (FM) interactions are established between the nanoparticles, probably of dipolar nature since the nanoparticles are not in contact due to the surfactants on their surfaces, and mainly mediated by the Fe cores. Since the size of the core diminishes for the core/void/shell nanoparticles, the strength of these interactions also diminishes, consistent with the observation that TC decreased from 59 K for the core/shell nanoparticles to 11 K for the core/void shell nanoparticles. For the hollow nanoparticles, however, we have obtained a high negative value of TC (−320 K). Given the very small size and high anisotropy of the randomly oriented nanograins inside the shell, it is reasonable to attribute this to the relative orientation of the anisotropy axes of the shell nanograins with respect to the magnetic field [44], rather than to the existence of very strong antiferromagnetic (AFM) inter-grain interactions, although the latter cannot be completely ruled out.

To further probe this, we have compared our experimental ZFC results with the theory. For simulations, we have used an expression derived from the Stoner–Wohlfarth model [40]. In principle this model was designed for non-interacting uniaxial nanoparticles, but, in a first approximation, it can be considered that the presence of interparticle interactions will mainly affect the effective anisotropy constant, K. According to this model, the ZFC magnetization is given by:

Equation (2)

where the first and second terms correspond to the un-blocked SPM (V < VC) and the blocked nanoparticles (V > VC), respectively, being:

Equation (3)

L(x) corresponds to a Langevin function, L(x) = cotanh(x)-1/x, Mo is the saturation magnetization of the system of nanoparticles (given by the MH loops), and MS is the saturation magnetization of each nanoparticle according to their composition (e.g. MS = 392 emu cm−3 for bulk maghemite). The function f(V) or f(D) is the particle size distribution; in our case a log-normal function defined by two parameters ($\alpha $ and $\beta )$ has been used, as shown below:

Equation (4)

In this distribution the mean diameter and standard deviation are given by:

Equation (5)

As observed in figure 2(d), a good fit with this model has been obtained only for the core/shell nanoparticles. The main problem with the fittings for the core/void/shell and hollow nanoparticles is the slower decay of the magnetization above TP-ZFC, which can be related to the broader distribution of energy barriers of the shell nanograins. For the core/shell nanoparticles, the average size is estimated to be D = 7.0 nm, which is close to the core size determined from the TEM image (figure 1(a)), and the obtained value of K = 1.5 × 106 erg cm−3 is also close to that estimated above. This indicates that at T ∼ 55 K, the magnetic behavior of the core/shell system can be understood in terms of an ensemble of interacting nanoparticles, with an effective size close to that of the Fe core, which starts to become collectively blocked at T ∼ 100 K. Since the interactions are mainly of dipolar nature, we can estimate the dipolar temperature, Tdip = Edip/kB, for the core/shell nanoparticles. The value of the dipolar energy Edip is given by:

Equation (6)

Since the nanoparticles in our study are packed together, though not in direct contact due to the presence of surfactants (see figure 1), L can be considered to be close to the diameter of the particle (L ∼ 12–13 nm). Accordingly, the value of Tdip for the core/shell nanoparticles is calculated to be ∼95 K, which is close to TP-ZFC (111 K), corroborating the dipolar nature of the interactions in this system. For the hollow nanoparticles, however, Tdip is close to 0 K, suggesting that as the nanoparticles become hollow, dipolar interparticle interactions become less relevant.

To probe the magnetic field evolution of the magnetization in the core/shell, core/void/shell, and hollow morphologies, we have measured and analyzed the magnetic hysteresis (MH) loops at low (5 K) and high (300 K) temperatures. As can be seen in figure 3(a), the MH curves measured at 5 K under ZFC protocol show a clear hysteresis and a non-saturating behavior. The loops have been normalized to M50kOe for comparison. The overall shape of the MH loops changes slightly when going from the core/shell to core/void/shell morphology, with a small increase of the high field slope and a decrease of the saturation magnetization. The decrease of the coercivity when going from the core/shell to core/void/shell morphology has been attributed to the detachment of the core from the shell [45]. For the hollow nanoparticles, however, the shape of the MH loops changes drastically, presenting an elongated shape with a pronounced increase of the coercivity and the high field slope. As summarized in table 3, as the nanoparticles become hollow, the normalized remanence decreases, while the coercivity greatly increases. The MH loop for the hollow nanoparticles, especially at low temperatures, resembles those of frustrated and disordered random anisotropy magnets, and the same features have also been observed in similar hollow maghemite nanoparticles with strong shell anisotropy [35]. This suggests that there is a higher frustration for the spins in the shell of the hollow nanoparticles in comparison with the core/shell and core/void/shell ones. On the other hand, for the core/shell and core/void/shell nanoparticles, the presence of the Fe core and the reduced volume of the shell (Vsh ∼ 650 nm3 for the core/shell and core/void/shell nanoparticles, while Vsh ∼ 1400 nm3 for the hollow nanoparticles) give a behavior more similar to the one expected in interacting nanoparticle systems.

Figure 3.

Figure 3. (a) Normalized hysteresis MH loops measured at 5 K under ZFC protocol, and at 300 K for the 12 nm (b) core/shell, (c) core/void/shell and (d) hollow nanoparticles, together with their fits to the Langevin (SPM) +paramagnetic (PM) function.

Standard image High-resolution image

Table 3.  Coercivity, Hc, and normalized remanence, Mr/Ms, at 5 K; and estimated average size, D, and size distribution, σ, SPM and PM contributions, as obtained from the fittings of the hysteresis loops at 300 K for the 12 nm nanoparticles.

  Hc 5K (Oe) Mr/Ms 5K D (nm) 300K σ (nm) 300K %SPM 300K %PM 300K
C/S 690(20) 0.18(4) 7.3(5) 0.7(1) 93(4) 7.0(3)
C/V/S 540(20) 0.10(2) 5.5(3) 1.1(2) 70(3) 30(2)
H 1870(20) 0.12(2) 3.5(2) 2.4(4) 68(3) 32(2)

To understand the relative contributions of the disordered surface and interface spins to the magnetization of the nanoparticles, we have analyzed the MH loops measured at 300 K. As observed in figures 3(b)–(d), all the MH curves present no hysteresis, meaning null coercivity and remanence, and a linear contribution at high fields that tends to increase when going from the core/shell to hollow morphology. This linear contribution can be associated with the presence of an increasing amount of surface and interface spins which behave in a paramagnetic (PM) way. To quantify this, we have fitted the MH loops using two contributions: (i) a SPM contribution corresponding to the magnetization of the cores and shell grains, and (b) a PM contribution associated with uncompensated spins:

Equation (7)

where χPM is the PM susceptibility. As can be seen in figures 3(b)–(d), good fits have been obtained for the three samples, indicating that the average estimated size decreases when the morphology changes from core/shell to hollow, but the size distribution becomes bigger. This agrees with our previous analysis on the ZFC MT curves. For the core/shell and core/void/shell nanoparticles, the estimated size is similar to the core of the nanoparticles, indicating that the SPM behavior originates mainly from the Fe cores. For the core/shell nanoparticles, the Fe core occupies 23% of the total volume of the particle, but since the magnetization of Fe (1710 emu cm−3) is much higher than γ-Fe2O3 (392 emu cm−3), the magnetic contribution (emu) per particle represents nearly a 60%. For the hollow nanoparticles, however, the average size is close to the shell thickness and can be associated with the grains inside the shell, that present a considerable size distribution. In addition, the PM contribution, which is minimal for the core/shell nanoparticles, 7%, becomes much more important for the core/void/shell nanoparticles, 30%, due to the detachment of the core from the shell, which originates from the apparition of additional uncompensated spins at the inner and outer surfaces of the shell and the core, respectively. On the other hand, for the hollow nanoparticles, taking into account the disappearance of the core and the increase in the shell thickness, the relative number of uncompensated spins, and thereby the PM contribution, would be expected to diminish. Instead, it remains almost the same ∼30%. Together with the MH data at 5 K, this suggests that for the hollow nanoparticles, the number of uncompensated spins must increase and be related to the increase in the thickness of the shell observed for these nanoparticles. This can be understood if we consider that each shell nanograin is composed of central and surface spins, following the model presented by Cabot et al [35]. According to this model, in the hollow nanoparticles (thickness ∼3.2 nm) we have an exchange coupling between the irreversible outer (surface) spins and the ferrimagnetically ordered and reversible inner (core) spins. This can result in an additional source of frustration which is not observed when the shell is thinner, such as in the case of the core/void/shell nanoparticles (at a shell thickness ∼2.2 nm).

A striking consequence of the different orientation of the surface and core spins, as well as the large number of uncompensated spins at the shell of the nanoparticles is the observation of EB (i.e. a horizontal shift in the MH loop and an increase in coercivity after cooling in a magnetic field). Figure 4(a) shows the MH loops measured at 5 K under the FC (50 kOe) protocol. As can be seen in this figure, the shapes of these MH curves are similar to those measured under the ZFC protocol, but with higher coercivity and the MH loops being shifted along both the horizontal and vertical directions. It is noted that for the hollow nanoparticles the maximum applied field is smaller than the irreversibility field, so the shift observed in the MH loop may not represent a real EB effect, but just a minor hysteresis loop. Nevertheless, we will also be referring to it as EB. Another interesting feature is the presence of a sharp change or a jump in the magnetization at low fields, ΔM, in the MH loop as indicated by an arrow, which appears both in the core/shell and hollow nanoparticles. A noticeable difference in this jump is observed between these two samples. For the core/shell nanoparticles, this jump only appeared at 5 K after FC and during the negative field sweep. For the hollow nanoparticles, however, the jump was observed below 25 K in both the positive and negative field sweeps and under both the FC and ZFC protocols. For the core/shell nanoparticles, the jump can be easily related to the unidirectional alignment of frozen interfacial spins with FC, which provides a maximum exchange coupling between the core and the shell. As the field drops, this core/shell coupling is overcome by the random crystalline anisotropies in the shell, resulting in a demagnetization of the shell [26]. This would also explain why no jump was observed for the core/void/shell nanoparticles. However, for the hollow nanoparticles (with the absence of interface spins between the core and the shell), the appearance of a jump below 30 K must be related to the coupling between the central layer and surface spins in the nanograins, as explained before. Moreover, for these hollow samples an asymmetrical magnetization reversal has been observed in the FC MH loops measured below ∼30 K; the descending branch of the MH loops shows a much slower approach to saturation than the ascending one, which has been attributed to the competing anisotropy and spin orientation in the shell [46, 47]. This again suggests that ∼30 K marks the onset for a spin freezing phenomenon at the shell nanograins of the hollow nanoparticles.

Figure 4.

Figure 4. (a) Normalized hysteresis MH loops measured at 5 K under FC protocol, and the temperature dependence of the (b) coercivity, (c) the horizontal shift or the EB field and the vertical shift, for the 12 nm core/shell, core/void/shell and hollow nanoparticles. In the inset of (a) a zoom-in of the hysteresis loops region exhibiting the 'jump' is presented.

Standard image High-resolution image

Figures 4(b) and (c) show the temperature dependence of the coercivity, HC, and the EB field, HEB, for the three samples after field cooling in 50 kOe. It can be observed that for all the samples both HC and HEB progressively decrease with increasing temperature. HEB becomes close to zero above ∼30 K, while HC is still non-zero above this temperature. This indicates that for the three samples, below ∼30 K, surface and interface atoms at the shell start to freeze, as supported by the increase in coercivity, behaving as pinning centers for the development of EB. Above this temperature, these 'melted' spins no longer contribute to the EB, but they are still exchange coupled with the nearby atoms from the core and/or the shell grains, thus contributing to the coercivity [25]. It can be seen that while the HC values of the core/shell and core/void/shell nanoparticles are similar, a slight increase in HEB is observed for the core/void/shell sample. As discussed before, for the core/void/shell sample the number of uncompensated spins greatly increased in comparison with the core/shell ones, due to the detachment and decrease in size of the core. This is also confirmed by the increase in the vertical shift of the hysteresis loops (see figure 4(d)), which has been defined as:

Equation (8)

This value of Mvert is proportional to the number of frozen spins that cannot be reversed by the magnetic field [27], and, in principle, HEB should depend linearly on the ratio of number of frozen spins to reversible spins:

Equation (9)

Hence, the increase in HEB for the core/void/shell sample can be interpreted in terms of an increase in the number of frozen spins at the inner and outer surface of the shell, which act as pinning centers for the EB phenomenon. However, although hollow nanoparticles present a similar relative number of uncompensated spins as core/void/shell nanoparticles (∼30%), HEB is much larger for these, reaching values of ∼7 kOe at low temperatures. This indicates that the coupling between the central and surface atoms in each nanograin, together with the strong anisotropies and big size distribution of the nanograins in this shell, give rise to a remarkable EB effect that is not observed in the case of the core/shell and core/void/shell nanoparticles having a thinner shell. This indicates that the hollow nanoparticles are very promising for EB-based spintronics applications and that the shell thickness is crucial in order to enhance the EB effect in these systems.

To probe the spin dynamics of these nanoparticles, ac susceptibility measurements were systematically performed on the samples by applying a 10 Oe ac field within the frequency range 10 Hz to 10 kHz. By changing the frequency, we are deliberately changing the probe time, τ = 1/ω = 1/2πf, which allows us to probe the relaxation of the magnetic moments in different time windows. Figure 5 shows the real, χ', and imaginary, χ'', components of the ac susceptibility. It can be observed that for all the samples a clear maximum is obtained in both components. The position of χ' maximum, Tm', has been typically associated with the collective freezing temperature of the system. With increasing frequency, this peak becomes smaller and its position displaces towards higher temperatures, as it is typical in disordered systems such as spin glasses [48]. Above this maximum, the χ'(T) curves overlap and there is a continuous decay of the susceptibility. On the other hand, the imaginary part, χ'', has been associated with the appearance of dissipative processes in the system, with its position given by Tm''. It can be seen that there are some interesting differences in the evolution of χ'' for the core/shell on one hand, and the core/void/shell and hollow nanoparticles on the other hand. For the core/shell nanoparticles, χ'' increases with increasing frequency, as is conventional in systems of magnetic nanoparticles, but there is a change below ∼80 K, where it decreases. On the other hand, for the core/void/shell and hollow nanoparticles, χ'' slightly decreases with increasing frequency. This behavior indicates a more frustrated state for the core/void/shell and hollow nanoparticles (see for example, Frey et al [49]). In addition, the change in the evolution of χ'' for the core/shell nanoparticles reveals two different freezing processes taking place, as already hinted by dc ZFC−FC measurements; one starting around 100 K associated with the collective freezing of the nanoparticles and mediated by dipolar inter-particle interactions, and the second one at lower temperatures, which for its similarity with the core/void/shell and hollow nanoparticles, must be associated with the freezing of the nanograins inside the shell. Above the maximum, χ'' falls towards zero in the three cases, indicating that at high temperatures the system enters into a SPM-like state [50].

Figure 5.

Figure 5. Temperature dependent ac susceptibilities of the 12 nm core/shell, core/void/shell and hollow nanoparticles. The arrows indicate the shift of the curve peaks as f increases.

Standard image High-resolution image

In order to obtain more quantitative information about the different collective magnetic behaviors at low temperatures, we have analyzed the displacement of the χ' peak as a function of the frequency. The peak shift can be quantified by

Equation (10)

The obtained values of Γ (table 4) are close to those of spin glass-like systems (Γ < 0.06). Γ increases as the morphology changes from core/shell to hollow, which can be related to the weakening of the interactions [51]. An attempt to fit the Tm' versus f dependence using a Neel–Arrehnius expression, typical of SPM systems, yielded unphysical values. Thus, we have extended our analysis to the Vogel–Fulcher model, which is actually a phenomenological modification of the Neel–Arrhenius model to include the effect of the interactions in the form of a change to the energy barrier, as given by an additional temperature, TVF. According to this model, the relaxation of the magnetic moments in our systems is assessed by:

Equation (11)

where Ea = KV is the anisotropy energy barrier and τVF is the relaxation time of each magnetic nanoparticle. TVF is in principle related to the strength of the interactions. As can be seen in figure 6(a), good fits have been obtained in all the cases. The obtained τVF values are within the typical limits of the relaxation times of individual nanoparticles or clusters (10−8 − 10−12 s). As the nanoparticles become hollow, the energy barrier progressively increases. The decrease in TVF is related to the decrease in 〈TB〉, which suggests a weakening of the inter-particle interactions for the core/void/shell and hollow nanoparticles.

Table 4.  Analysis of the relaxation times for all three samples, on the basis of the Vogel–Fulcher and power law descriptions of the spin dynamics for the 12 nm nanoparticles.

  Γ τVF (s) E/kB (K) TVF (K) τC (s) zv Tg (K)
C/S 0.019(2) 3.7(2) × 10−12 260(20) 92(9) 5(1) × 10−11 4.2(6) 95(9)
C/V/S 0.039(3) 7(2) × 10−12 440(40) 56(6) 1.0(4) × 10−10 10(2) 65(7)
H 0.050(4) 2.1(9) × 10−11 510(50) 45(5) 7.9(6) × 10−9 11(2) 52(5)
Figure 6.

Figure 6. Fits of τ versus T using the (a) Vogel–Fulcher and (b) critical slowing down (double-logarithmic plot) models for the 12 nm core/shell, core/void/shell and hollow nanoparticles.

Standard image High-resolution image

Since the Γ values and the FC MT features suggested the presence of a glassy freezing at low temperatures, we have employed a critical power law, which is mostly associated with spin-glass systems:

Equation (12)

Here Tg marks the onset of the collective glassy behavior, τC corresponds again with the relaxation time of each magnetic nanoparticle, and zv is a critical exponent related to the correlation length that diverges at Tg. Again, very good fits have been obtained for the three samples (figure 6(b)). The obtained values of Tg are close to those of 〈TB〉 for the core/void/shell and hollow nanoparticles, and close to TP-ZFC for the core/shell nanoparticles. This indicates that we are probing the freezing of the shell grains in the first two cases and the collective blocking of the nanoparticles in the last one. As the particles become hollow, both zv and τC increase. Higher zv values have been reported in the case of cluster-glass systems and can be attributed to the increased disorder of the spins in the shell.

In short, we have observed a clear change in the magnetic behavior of the 12 nm core/shell nanoparticles as they become hollow. The core/shell nanoparticles behave like an ensemble of interacting nanoparticles, with a collective freezing of spins into a SSG-like state below 100 K, mediated by inter-particle dipolar interactions. The shell becomes frozen at around 30 K and the EB effect is observed below this temperature, due to the freezing of core–shell interface spins. As the core detaches from the shell and the morphology transforms into core/void/shell, the magnetization of the system decays, inter-particle interactions become weaker, and the effective anisotropy increases due to an increase in the number of uncompensated spins both at the inner and outer surfaces of the shell and the core, respectively. A slight increase of the EB is achieved as a result. The changes are more drastic when the nanoparticles become completely hollow. In this case, the magnetic behavior is only associated with the spins in the nanograins forming the polycrystalline shell. Due to the increase of the shell thickness, a coupling between the ferrimagnetic central spins layer and the disordered surface spins is established inside these nanograins. This gives rise to a highly frustrated magnetic state, which greatly increases the EB effect, indicating that the morphology of the shell plays a crucial role in this kind of exchange-biased systems. Below the freezing temperature of the shell, 30 K, the system presents a spin glass like behavior.

In addition the effect of disordered surface spins, the finite-size effect has been reported to be important in ferrimagnetic nanoparticle systems [27]. This effect arises mainly from an unbalanced number of spins in an antiparallel arrangement, which usually differs from the bulk material and is expected to vary significantly as the particle size is reduced below a critical size. In comparison with the case of the 12 nm nanoparticles (with particle size larger than the critical size, 10 nm), we have also studied the magnetic properties of 8 nm nanoparticles, that is, below the critical size. The main results are shown in figures 7 and 8. Figure 7(a) shows the ZFC−FC MH curves measured in a field of 100 Oe for the 8 nm core/shell and hollow nanoparticles. It can be observed that the curves are very similar for both samples, except for a slight displacement of the maximum position, TP-ZFC. Here we note that the Fe core only occupies 6% of the whole volume of the nanoparticle, which is nearly four times less than in the case of the 12 nm nanoparticles. As a result, a similar magnetic behavior can be expected for both the core/shell and hollow nanoparticles. Weaker dipolar inter-particle interactions are also expected for these nanoparticles, as compared to the 12 nm nanoparticles. Above TP-ZFC, the magnetization decays progressively, approaching zero at high temperatures. The values of 〈TB〉 are estimated using the same method employed before for the 12 nm nanoparticles. It is interesting to note that in contrast to the case of the 12 nm nanoparticles, 〈TB〉 increases when the morphology changes from core/shell to hollow for the case of the 8 nm nanoparticles: from 45 K for the core/shell morphology to 54 K for the hollow ones. This increase in 〈TB〉 cannot be solely attributed to the enhancement in the surface anisotropy in the particles with hollow morphology, as suggested in the previous work [34], but must be also related to the finite-size effect of the nanoparticles. This finding points towards the importance of the finite-size effect in these systems, which fully agrees with our previous observation of a critical particle size (mean size, ∼10 nm) for the Fe/γ-Fe2O3 core/shell nanoparticles, above which the interface spin effect contributes mainly to the EB, but below which the surface spin effect is dominant. On the other hand, both samples show a small and similar decrease in the FC magnetization below TP-ZFC, which can be attributed to the presence of a SG-like state of the shell at low temperatures. It is noted that we no longer obtain the high temperature irreversibility that was observed in the case of the 12 nm hollow nanoparticles' ZFC−FC magnetization curves (see the inset to figure 2(b)). This suggests a smaller size distribution of the nanograins in the 8 nm hollow nanoparticles than in the 12 nm hollow nanoparticles. Except at very high temperatures, a strong deviation of the Curie–Weiss law is observed for the 8 nm samples, which has been typically found in SG systems [48].

Figure 7.

Figure 7. (a) ZFC/FC MT curves for the 8 nm core/shell and hollow nanoparticles; (b) normalized hysteresis loop measured at 5 K and (c), (d) at 300 K together with the fits using the Langevin (SPM) +paramagnetic (PM) function. In the inset of (d), thermal evolution of the coercivity and the EB field for the hollow nanoparticles.

Standard image High-resolution image
Figure 8.

Figure 8. (a) Temperature dependent ac susceptibility of the 8 nm hollow nanoparticles. Fits of τ versus T using (b) the Vogel–Fulcher model and (c) the critical slowing down (double-logarithmic plot) models.

Standard image High-resolution image

We have also measured the MH loops at 5 and 300 K for the 8 nm core/shell and hollow nanoparticles. As can be seen in figure 7(b), at 5 K the MH loops are very similar for both samples. The HC of both 8 nm nanoparticles is ∼1250 Oe, which is smaller than that of the 12 nm hollow nanoparticles (∼1870 Oe) but larger than that of the 12 nm core/shell nanoparticles (690 Oe). On the other hand, the room-temperature MH loops present a Langevin-type non-hysteretic behavior (figures 7(c) and (d)), with a PM contribution of 27% and 36% for the core/shell and hollow nanoparticles, respectively. These results indicate that by decreasing the particle size from 12 to 8 nm, the number of uncompensated spins increases. In particular, if we compare the volume occupied by inner and outer surface spins, given the same thickness of approximately 1 atomic layer (0.8 nm), then we obtain that for the 12 nm nanoparticles the relative volume occupied by the atoms on the surfaces of the shell is 45%, while for the 8 nm samples it is about 80%.

To examine if such an increase of the volume occupied by surface spins gives rise to EB in the 8 nm nanoparticles, the thermal evolution of the EB and coercivity after FC at 50 kOe were studied, as shown in the inset to figure 7(d) for the hollow nanoparticles. HC at low temperatures is ∼3500 Oe, which keeps practically constant up to 15 K and then starts to decrease at higher temperatures. On the other hand, HEB rapidly increases below 25 K, and reaches the highest value of ∼5000 Oe at 5 K. Similar results have been obtained for the 8 nm core/shell nanoparticles. As the shell thickness in both cases is almost equal (∼2 nm), the morphology of the crystallites inside the shell should not vary appreciably. However, these values of HEB are still smaller than those obtained for the 12 nm hollow nanoparticles (∼7000 Oe at 5 K). This can be understood by considering the fact that the origin of the EB in hollow nanoparticle systems is also associated with the exchange coupling between the irreversible surface spins and the reversible central spins in the shell nanograins. As the particle size of the hollow nanoparticles decreases from 12 to 8 nm, the volume occupied by the central spins is significantly reduced, leading to the weakening of the exchange coupling. As a result, a smaller EB effect is obtained for the 8 nm hollow nanoparticles. This is fully consistent with the MT data (figure 7) and analysis that suggest a low temperature SG-like behavior for the 8 nm hollow particles.

To further elucidate this intriguing feature, we have performed ac susceptibility measurements for the 8 nm core/shell and hollow samples. The main results are presented in figure 8 and the fit parameters summarized in table 5. The evolution of χ' and χ'' is similar to the one obtained for the 12 nm hollow nanoparticles: both χ' and χ'' displace towards higher temperatures with increasing frequency, but the amplitude only changes noticeably for χ'. The obtained values of Γ are of the same order of magnitude of SG-like systems. Again, fits to the Neel–Arrhenius model give unphysical parameters. However, the Vogel–Fulcher model yields a similar anisotropy barrier for both samples, and indicates that the intrinsic relaxation time is appreciably smaller for the hollow nanoparticles than for the core/shell nanoparticles. A similar result is also obtained from the critical power law analysis. It can be seen however that the value of the relaxation time is always smaller for the hollow nanoparticles, and that the critical exponent increases. This can be related to the absence of the core and the increased number of uncompensated spins, which gives rise to a more spin glass-like behavior in the case of the hollow nanoparticles. Similar τC and values have been reported in the case of spin glass systems [50]. A final note is that the relaxation time, determined from both the Vogel–Fulcher model and the critical exponent law, is much smaller for the 8 nm hollow nanoparticles than for the 12 nm nanoparticles. This once again suggests a more conventional glassy behavior for the 8 nm particles.

Table 5.  Analysis of the relaxation times for all three samples, on the basis of the Vogel–Fulcher and power law descriptions of the spin dynamics for the 8 nm core/shell and hollow nanoparticles.

  Γ τVF (s) E/kB (K) TVF (K) τC (s) zv Tg (K)
C/S 0.043(4) 1.5(5) × 10−10 230(20) 50(5) 6(1) × 10−10 6.0(6) 59(6)
H 0.016(2) 7(2) × 10−13 200(20) 59(6) 1.6(5) × 10−12 10(1) 64(6)

4. Conclusions

The influence of the morphology, surface and finite-size effects on the magnetic properties of Fe/γ-Fe2O3 core/shell, core/void/shell, and hollow nanoparticles has been studied. We find that for the 12 nm particles the mean blocking temperature decreases as the morphology changes from core/shell to hollow, while an opposite trend is observed for the 8 nm particles. The 12 nm core/shell particles behave like an ensemble of interacting nanoparticles, with a collective freezing of spins into a super spin glass-like state below 100 K, mediated by dipolar inter-particle interactions. The shell becomes frozen below 30 K, resulting in the enhanced EB effect. As the morphology transforms into core/void/shell, the magnetization of the system decays, inter-particle interactions become weaker, and the effective anisotropy and hence the EB increases. The changes are more drastic when the nanoparticles become completely hollow. In the case of hollow nanoparticles, the morphology of the shell plays a crucial role in the low temperature magnetic behavior and EB. Below the freezing temperature of the shell, ∼30 K, the system behaves like a frustrated cluster glass. In addition, for these 12 nm hollow nanoparticles, the increased thickness of the shell allows an exchange coupling between outer irreversible spins and the ferrimagnetically ordered reversible spins in the inner layer of the shell, giving rise to an enhanced EB effect at low temperatures. Meanwhile, both the 8 nm core/shell and hollow particle systems exhibit a spin glass-like behavior at low temperatures. Our study provides deeper insights into the morphology, surface and finite-size effects in magnetic nanoparticle systems, knowledge of which is the key to manipulating novel magnetic nanostructures for biomedical and spintronics applications.

Acknowledgments

Research was supported by the US Department of Energy, Office of Basic Energy Sciences, Division of Materials Sciences and Engineering under Award #DE-FG02-07ER46438 (magnetic and structural studies). PM and HS also acknowledge support from the Center for Integrated Functional Materials through grant USAMRMC W81XWH-10-2-0101 (samples' synthesis). JA acknowledges the financial support provided through a postdoctoral fellowship from Basque Government. HS also acknowledges support from Bizkaia Talent Program, Basque Country (Spain).

Please wait… references are loading.