Effect of particle size distribution on magnetic behavior of nanoparticles with uniaxial anisotropy
Ali S Rizwan1, †, Naz Farah1, Akber Humaira1, Naeem M1, Ali S Imran2, Basit S Abdul3, Sarim M3, Qaseem Sadaf1
Department of Physics, Federal Urdu University of Arts, Science and Technology, Karachi, Pakistan
Department of Applied Chemistry and Chemical Technology, University of Karachi, Karachi, Pakistan
Department of Computer Science, Federal Urdu University of Arts, Science and Technology, Karachi, Pakistan

 

† Corresponding author. E-mail: rizwan@fuuast.edu.pk

Abstract

The effect of particle size distribution on the field and temperature dependence of the hysteresis loop features like coercivity (HC), remanence (MR), and blocking temperature (TB) is simulated for an ensemble of single domain ferromagnetic nanoparticles with uniaxial anisotropy. Our simulations are based on the two-state model for T < TB and the metropolis Monte–Carlo method for T > TB. It is found that the increase in the grain size significantly enhances HC and TB. The presence of interparticle exchange interaction in the system suppresses HC but causes MR to significantly increase. Our results show that the parameters associated with the particle size distribution (Dd,δ) such as the mean particle size d and standard-deviation δ play key roles in the magnetic behavior of the system.

1. Introduction

With the technological advancements in the field of magnetic recording, size effects on the magnetic properties of nanostructures including spherical particles, nanodots, nanowires, nanopillars, and thin films become increasingly important.[13] Single domain magnetic nanoparticles of Co, Ni, FePt, etc. with high anisotropy and magnetic moment are of particular importance.[46] In order to attain higher recording densities with these materials, their grain size has to be small.[7] Smaller grains undergo thermally activated magnetization reversals, resulting in the loss of recorded information. However, the magnetization of small particles can be effectively stabilized against thermal fluctuations by inducing additional anisotropy in the system, e.g., exchange bias in core-shell nanoparticles[8] or by using shape anisotropy.[9] Better understanding of size related effects in magnetic nanoparticles is the key to designing and optimizing thermally stable high density recording media. For these systems, not only is the mean particle size of relevance but also the broadness in the size distribution, because the presence of very small grains results in partial thermal instabilities whereas very large grains may enhance the coercivity (HC) from a medium to an intolerable limit. Studies have shown the significance of grain size for the thermal stability of the magnetization of fine particles[10,11] but broadness in the particle size distribution has not been explicitly studied. Xu et al.[12] have shown that the shape of the hysteresis loop for non-interacting spherical particles in dilute random composites is indeed sensitive to the broadness of the size distribution.

Here, we consider a system of spherical particles having a Gaussian size distribution. These particles are assumed to be single domain with uniaxial anisotropy axes randomly distributed in space. We study the effect of broadness in the size distribution on the field and temperature dependence of magnetization by using a two-state model and the metropolis Monte Carlo (MMC) method. Magnetic properties such as the HC and blocking temperature (TB) are found to be very sensitive to size variations, whereas remanence (MR) is found to be largely independent of the broadness of the particle size distribution. We also simulate the effects of interparticle exchange interaction on the magnetic behavior of these particles. Our results show a decrease in HC and increase in MR of the system with increasing strength of exchange interaction. The rest of this paper is organized as follows. In Sections SubSection 2 and SubSection 3, we deal with the description of the model system and theoretical framework within which we perform the simulations. In Section 4, we present our results followed by the discussion of some important aspects of our results. Finally, in Section 5 we draw some conclusions from the study of this work.

2. Model system

First, we consider a system of 1200 non-interacting nanoparticles which are assumed to be well-dispersed in a nonmagnetic matrix. These particles have a Gaussian size-distribution (Dd,δ) with d as the average diameter (2 nm, 3 nm, and 4 nm) and δ as the standard deviation (1 nm, 1.5 nm, and 2 nm). Physically, δ corresponds to the dispersion or broadness in the distribution of particle size. The effects of d and δ are systematically studied for five samples with different Dd,δ, viz., D2,1, D3,1, D4,1, D4,1.5, and D4,2. Initially, we consider a system with low packing density such that the interparticle interactions can be ignored. This allows us to test our model system within the theoretical framework of the Stoner–Wohlfarth (SW) model.[13,14] The atomic moments within each nanoparticle are assumed to be uniformly oriented so that their resultant moment can be treated as a point dipole located at the center of the particle. Each particle has uniaxial anisotropy with anisotropy constant K1 ∼ 1.0 × 106 J/m3, which is slightly greater than that of bulk fcc Co.[15,16] The measurement time τm is kept consistent with the experimental value, i.e., 100 s.[17] Finally, we extend this model to the case where the effects of interparticle exchange interaction are taken into consideration. Here, the simulations are run for different relative exchange strength values (J/k) ranging from 0 to 0.4, where J is the exchange energy and k = K1V is the anisotropy energy barrier of a particle of volume V. In the experimental measurements, the value of J is generally not defined.[18] Thus, in our simulations we treat it as a free parameter and its value is taken to be volume dependent.

3. Simulations

We use a quasi static process in our MMC simulations in which the energy of the system is evaluated at each change in the external parameter. The moments are reoriented on the basis of energy minimum criteria. Figure 1 shows a schematic representation of our model system. The Hamiltonian of the system as suggested by the SW model is given by[1921]

where θo is the angle between the easy axis and the applied magnetic field H, and θ is the angle between the moment μ and the applied magnetic field (Fig. 1). For each particle, μ is defined by its V and specific magnetization Mo (taken as 1.43 × 106 A · m−1 for Co). In general, the state of each particle is described by the temperature T of the system and τm. Keeping τm constant, TB of the particles is given by
where kB is the Boltzmann constant, K1V is the energy barrier height in the absence of applied field, and τo is the characteristic relaxation time (∼10−9 s).[22,23] Here, T over TB defines the state of the system. For T < TB, the particles are in a blocked state and possess pure ferromagnetic features, which can be simulated by the two-state model. In this model, there is a finite probability that the moment may switch between the two energy minimum states. Numerical calculations[24] have shown that EB is a function of K1V and H and is expressed as
where HA is given by 2K1hS/Mo and γ = 0.86 + 1.14 hS, whereas hS is defined by hS(θo) = [(cosθo)2/3 + (sin θo)2/3]−3/2.[23] For each particle, the orientation of its moment is driven by the interplay between its local HC and applied H,[23] where HC = 0.48 HA (1 − (T/TB)0.77). For the low applied field, i.e., H < HC, the process of magnetization reversal is simulated by calculating the transition probability from one minimum to the other by using p = 1 − exp (−τm/τ), where τ is the characteristic reversal time of the moments. On the other hand, at elevated temperatures, i.e., T > TB, the particles are in a superparamagnetic (SPM) state and the orientation of their moments is governed by the MMC technique.[22]

Fig. 1. (color online) Schematic representation of the system of nanoparticles with random orientation of the moment (left). Angle distribution showing the direction of moment, field, and easy axis (right).
4. Results and discussion

Figure 2 shows some representative M(H) loops simulated at different temperatures for the sample with size distribution D4,1. The initial state of the system is defined by setting H = 0. In this state, all the moments are distributed evenly in state 1 and state 2, thus giving zero net magnetization of the system. At small H, some of the moments, mainly those associated with smaller particles start switching to the state corresponding to the field direction. With increasing H in small steps, the moments keep on aligning with H and the magnetization gradually approaches to its saturation value Ms, when all the moments align with H. These initial magnetization processes are similar to each other for all distributions and are omitted in the hysteresis loops (± 1 T) of the samples. In the descending field branch, the magnetization follows a different path and reaches a remenant state having the value of MR as H decreases from 1 T to 0. It should be noted that the sample exhibits strong MR at 10 K. The higher values of MR at low temperature are attributed to small thermal energy (kBT) which is not sufficient for the moments to undergo thermally activated reversals even at H = 0. Thus, at low temperature, all the moments remain blocked in the ferromagnetic state. When H is applied in the reverse direction, the moments start to switch from their random and blocked state and align with the reverse field. This results in a decrease in M/Ms until it vanishes near H = −HC. The further increase in the field aligns all the moments along the reverse field direction and the magnetization achieves −Ms. The same process continues for the ascending field branch of the M(H) loop. At 50 K (down triangles), some of the smaller particles become SPM, i.e., their moments become thermally unstable and start to fluctuate between their minimum energy states. The large MR present in this loop is attributed to the fact that the fraction of small sized SPM particles is not very large at 50 K. With increasing temperature, the fraction of thermally unstable particles increases and accounts for the observed decrease in MR and HC. Near 190 K, all the particles become SPM and their magnetization loop can be described by the Langevin function. In the SPM regime, kBT overcomes the Zeeman and anisotropy energies and the moments undergo fast switching between their minimum energy states, i.e., M = 0 at H = 0.

Fig. 2. (color online) Variations of M/Ms with H for the sample with size distribution D4,1 at different temperatures.

Figure 3 shows the effect of temperature on HC for samples with different Dd,δ. For all samples, HC is found to be large at very low temperature and decreases with increasing temperature. Here, we have two different cases. Case 1: for the samples with constant d (4 nm), i.e., for samples D4,1, D4,1.5, and D4,2. In this case, the values of HC are found to be the same for all samples at 10 K (Fig. 3(a)). This is attributed to the same value of d for each sample. Beyond 10 K, the value of HC decreases sharply down to the mean TB (100 K), which corresponds to the value of d of the sample. The rapid decrease in HC is attributed to the increasing fraction of SPM particles due to temperature driven de-blocking transitions in the sample. Above 100 K, the decay rate of HC becomes relatively gradual before it vanishes at a temperature which corresponds to (TB of the largest particle in the distribution). Case 2: for the samples with constant δ (1 nm), i.e., for samples D2,1, D3,1, and D4,1, the temperature dependence of HC is shown in Fig. 3(b). Overall, it is similar to the one observed for the samples in case 1. However, in contrast to case 1 (Fig. 3(a)), all the samples of case 2 exhibit different values of HC at 10 K. Moreover, due to different values of d, each sample exhibits a different mean TB. The value of for sample D2,1 is found to be 41 K, which is in agreement with the one observed by Xu et al.[12] for a system of nanoparticles having a uniform diameter of 3 nm. For the samples D3,1 and D4,1, we observe the large values of HC at 10 K. This is because these samples contain particles which are larger than the largest particle present in sample D2,1. At higher temperature, HC for both types of samples (case 1 and case 2) vanishes at their corresponding TB,which is consistent with the results of Du et al.[25] It is evident from these results that the increase in either d or δ leads to enhanced TB for the sample.

Fig. 3. (color online) Coercivity HC as a function of temperature for distributions (a) D4,1, D4,1.5, D4,2 and (b) D2,1, D3,1, D4,1.

For a better understanding of the effects of particle size distribution on HC and MR, we compare the M(H) loops of the samples with different distributions. The simulated M(H) loops for the samples with constant d (4 nm) and the ones with constant δ (1 nm) are shown in Figs. 4(a) and 4(b), respectively. It is evident that the samples with broader size distributions (δ) or with larger mean diameter (d) exhibit larger HC. This is because large δ or d implies bigger particles in the distribution with significant volume dependent anisotropy energy. This makes their moments more stable against thermal fluctuations and external fields. A larger value of HC is thus required to switch their direction. The value of MR is found to be sensitive to the value of d for the sample. Thus, MR remains largely unaltered for samples with constant d (Fig. 4(a)) whereas it decreases with decreasing value of d (Fig. 4(b)).

Fig. 4. (color online) Hysteresis loops M(H) for distributions (a) D4,1, D4,1.5, D4,2 and (b) D2,1, D3,1, D4,1.

In nanoparticles, blocking/unblocking of the moments is characterized by the TB at which zero field cooled (ZFC) magnetization (M(T)) passes through a maximum value. Experiments have shown that TB depends sensitively on the particle size.[26] In order to study the effects of particle size distribution, we simulate ZFC and field cooled FC M(T) for the samples with different distributions. Our ZFC (FC) process involves the cooling of the system from 300 K down to 5 K in the absence (presence) of 10 kOe field. For both ZFC and FC cases, the data are recorded during the heating of the sample from 5 K to 300 K in the presence of the field. For each sample, we observe a strong irreversibility between FC and ZFC curves, indicating the onset of the blocking transition in the samples (not shown here). Figures 5(a) and 5(b) show the simulated ZFC M(T) of our samples. In all cases, the ZFC curves pass through a maximum value near the TB of the system. This TB corresponds to the temperature at which the particle with mean diameter (d) undergoes the blocking transition. With reducing the temperature below TB, the particles in the ensemble undergo blocking transitions at their size dependent local TB’s. Similarly, with increasing temperature beyond TB, the particles undergo SPM transitions. It is evident that the samples with larger δ (see Fig. 5(b)) each exhibit a rather broad ZFC peak. Additionally, with increasing values of δ and/or d, the ZFC peak shifts towards higher temperature. This behavior is consistent with the recent experimental results[27] and is attributed to the presence of larger and thermally stable particles in these samples. Thus, the size of the particles has a prominent effect on the TB of the system: TB increases as the size of the particles increases.

Fig. 5. (color online) Variations of M/Ms with T for distributions (a) D4,1, D4,1.5, D4,2 and (b) D2,1, D3,1, D4,1.

In the above discussion, we have considered systems of non-interacting nanoparticles. Next, we incorporate exchange interaction into our system. Figure 6(a) shows the effect of exchange interaction (J/k) on HC at different temperatures. At each temperature, we observe a clear decrease in HC with J/k increasing from 0 to 0.4. This behavior is attributed to the fact that the particles are coupled to their neighboring particles via exchange forces. If the moment of a particle undergoes switching, it can induce switching in the other particle even at a low external field.[19] This leads to suppression of HC with increasing strength of exchange interaction. Another factor which is of interest here is the observed increase in MR with increasing J/k for temperatures above 60 K (Fig. 6(b)). Below 60 K, MR remains largely independent of J/k because here the thermal energy (kBT) is too low to account for any change in MR at any value of J/k. Physically, J/k affects EB of the system according to

where α = (J + μH) / 2K1V. This leads to the above mentioned changes in HC and MR of the system. The magnetizations are evaluated by the two-minimum approach and the results are found to be consistent with[18,28] those obtained from the system that is treated by the MMC method.

Fig. 6. (color online) Effects of exchange interaction (J/k) on (a) coercivity HC and (b) remanence MR/Ms at different temperatures.
5. Conclusion

The magnetic field dependence of blocking temperature and the temperature dependence of the coercive field with changing particle size distribution are studied by Monte Carlo simulations with a modified metropolis algorithm. It is found that the hysteresis curve collapses at high temperature when the particles become superparamagnetic. Thus, the shape of the hysteresis loop is found to be dependent on the particle size distribution and temperature. We also extend our model to the case where the interacting magnetic nanoparticles are taken into consideration. In this case, the coercivity decreases and remanence increases with increasing strength of the interparticle exchange interaction between the particles.

Reference
[1] Bedoya-Hincapie C M Ortiz-Alvarez H H Restrepo-Parraa E Olaya-Florez J J Alfonso J E 2015 Chin. Phys. 24 117701
[2] Sheng-Nan S Chao W Zan-Zan Z Yang-Long H Venkatraman S S Zhi-Chuan X 2014 Chin. Phys. 23 037503
[3] Masrour R Bahmad R Benyoussef A 2013 Chin. Phys. 22 057504
[4] Wu X W Guslienko K Y Chantrell R W Weller D 2003 Appl. Phys. Lett. 82 3475
[5] Guimarâe Alberto P 2009 Principles of Nanomagnetism Berlin Springer Science & Business Media 38 978-3642014819
[6] Munoz-Sandoval E Torres-Heredia J J López-Urías F 2005 J. Appl. Phys. 97 10
[7] Weller D Moser A 1999 IEEE Trans. Magn. 35 6
[8] Skumryev V Stoyanov S Zhang Y Hadjipanayis G Givord D Nogués J 2003 Nature 423 850
[9] Lv D L Xu C 2010 Chin. Phys. Lett. 27 097503
[10] Sánchez R D Rivas J Vaqueiro P López-Quintela M A Caeiro D 2002 J. Magn. Magn. Mater. 247 92
[11] Hansen M F Mørup S 1999 J. Magn. Magn. Mater. 203 214
[12] Xu C Li Z Y Hui P M 2001 J. Appl. Phys. 89 3403
[13] Stoner E C Wohlfarth E P 1948 Phil. Tran. Roy. Soc. Lond. 240 599
[14] Wernsdorfer W Orozco E B Hasselbach K Benoit A Barbara B Demoncy N Loiseau A Pascard H Mailly D 1997 Phys. Rev. Lett. 78 1791
[15] Bodker F Morup S Linderoth S 1994 Phys. Rev. Lett. 72 282
[16] Respaud M 1999 J. Appl. Phys. 86 556
[17] Cullity B D Graham C D 2011 Introduction Magn. Mater. Massachusetts John Wiley & Sons 425 978-1118211496
[18] Kechrakos D Trohidou K N 2003 J. Mag. Mag. Mat. 262 107
[19] Aharoni A 2000 Introduction Theory Ferromagnetism New York Clarendon Press 92 978-0198508090
[20] Chuev M A Hesse J 2007 J. Phys.: Condens. Matter 19 506201
[21] Tannous C Gieraltowski J 2008 Eur. J. Phys. 29 475
[22] Serantes D Baldomir D 2012 Open Surf. Sci. J. 4 71
[23] Sattler K D 2011 Handbook Nanophyscis: Nanoparticles Quantum Dots Boca Raton CRC Press 228 978-1420075441
[24] Pfeiffer H 1990 Phys. Status Solidi 122 377
[25] Du H F Du A 2006 J. Appl. Phys. 99 104306
[26] Jaffari G H Ali S R Hasanain S K Güntherodt G Shah S I J. Appl. Phys. 108 063921
[27] Chesnel K Trevino M Cai Y Hancock J M Smith S J Harrison R G 2014 J. Phys.: Conf. Ser. 521 012004
[28] Verdes C Diaz B R Thompson S M R Chantrell W Stancu A 2002 Phys. Rev. 65 174417