Next Article in Journal
Variational Bayes for Regime-Switching Log-Normal Models
Next Article in Special Issue
Ab Initio and Monte Carlo Approaches For the Magnetocaloric Effect in Co- and In-Doped Ni-Mn-Ga Heusler Alloys
Previous Article in Journal
Entropy Generation in Steady Laminar Boundary Layers with Pressure Gradients
Previous Article in Special Issue
Long-Range Atomic Order and Entropy Change at the Martensitic Transformation in a Ni-Mn-In-Co Metamagnetic Shape Memory Alloy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Phase Competitions behind the Giant Magnetic Entropy Variation: Gd5Si2Ge2 and Tb5Si2Ge2 Case Studies

1
IFIMUP and IN—Institute of Nanoscience and Nanotechnology, Departamento de Física e Astronomia da Faculdade de Ciências da Universidade do Porto, Rua do Campo Alegre, 687, Porto 4769-007, Portugal
2
CFNUL—Centro de Física Nuclear da Universidade de Lisboa, Av. Prof. Gama Pinto, 2, Lisboa 1649-003, Portugal
3
Instituto de Nanociencia de Aragón, Universidad de Zaragoza, ES-50018 Zaragoza, Spain and Departamento de Física de la Materia Condensada, Universidad de Zaragoza, Zaragoza 50009, Spain
4
Fundación ARAID, Zaragoza 50018, Spain
5
Instituto de Ciencia de Materiales de Aragón, CSIC-Universidad de Zaragoza, and Departamento de Física de la Materia Condensada, Universidad de Zaragoza, Zaragoza 50009, Spain
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Entropy 2014, 16(7), 3813-3831; https://doi.org/10.3390/e16073813
Submission received: 11 April 2014 / Revised: 27 June 2014 / Accepted: 1 July 2014 / Published: 11 July 2014
(This article belongs to the Special Issue Entropy in Shape Memory Alloys)

Abstract

:
Magnetic materials with strong spin-lattice coupling are a powerful set of candidates for multifunctional applications because of their multiferroic, magnetocaloric (MCE), magnetostrictive and magnetoresistive effects. In these materials there is a strong competition between two states (where a state comprises an atomic and an associated magnetic structure) that leads to the occurrence of phase transitions under subtle variations of external parameters, such as temperature, magnetic field and hydrostatic pressure. In this review a general method combining detailed magnetic measurements/analysis and first principles calculations with the purpose of estimating the phase transition temperature is presented with the help of two examples (Gd5Si2Ge2 and Tb5Si2Ge2). It is demonstrated that such method is an important tool for a deeper understanding of the (de)coupled nature of each phase transition in the materials belonging to the R5(Si,Ge)4 family and most possibly can be applied to other systems. The exotic Griffiths-like phase in the framework of the R5(SixGe1−x)4 compounds is reviewed and its generalization as a requisite for strong phase competitions systems that present large magneto-responsive properties is proposed.
PACS Codes:
65.40.gd; 75.30.Kz; 75.30.Sg

Graphical Abstract

1. Introduction

Back in the 1980s, the simultaneous discovery of the “ozone hole” in the stratosphere and of the influence of the emission of CFC gases, widely used in every refrigeration system at that time, in the thinning of the ozone layer triggered the search for alternative refrigeration technologies in order to replace the gas-compression technology [1]. The thermoelectric and thermomagnetic technologies are ahead in this race. Here only the latter one will be discussed and in particular the nature of the effect behind this technology—the magnetocaloric effect (MCE). The MCE was first observed in the 1840s, when Joule observed a temperature variation in a sample of pure iron as it was subjected to a magnetic field change [2]. It was understood on the basis of thermodynamics by Weiss and Piccard in 1917 [2]. At that time it became clear that the MCE is a phenomenon occurring in all magnetic materials when subjected to magnetic field variations, originating from the energy/entropy exchange between the magnetic and atomic lattice reservoirs. Through the fundamental Maxwell relations, the MCE can be expressed as the adiabatic temperature (ΔTad) and the isothermal magnetic entropy ( Δ S iso m) variation that a magnetic material undergoes when varying an external magnetic field [3]:
Δ T ad ( S , Δ H ) Δ H = μ 0 H 0 H 1 ( T C ( T , H ) ) H ( M ( T , M ) T ) H d H
Δ S iso m ( T , Δ H ) Δ H = μ 0 H 0 H 1 ( M ( T , M ) T ) H d H
where T is the temperature of the material, ΔH = H1 − H0 is the magnetic field variation, C is the heat capacity and M the magnetization. Hence, in order to maximize ΔT and ΔS, one important parameter to consider is the magnetization variation as a function of temperature: this means that the MCE reaches its maximum at the magnetic transition temperatures (the Curie temperature TC, or the Néel temperature TN) for all magnetic materials and consequently the operational temperature range of each material is delimited by a small interval close to its magnetic transition temperature. Furthermore, the MCE maximization depends strongly on the value of the magnetization derivative with respect to temperature ∂M/∂T, which is specific for each material and depends on the order of the associated phase transition.
Still in the 1930s, Giauque understood that he could make use of the large magnetization change at low temperatures of a paramagnetic salt (particularly gadolinium sulphate) to reach sub-Kelvin temperatures for the first time [4] and for a long time magnetic refrigeration was only considered for low temperature applications. Forty years later, in 1973, Brown [5] used gadolinium (which has a TC around 295 K) to design a magnetic refrigerator capable of operating at room temperature [6,7]. However, the real drive towards room temperature magnetic refrigeration occurred with the aforementioned events during the 1980s [1]. Finally, in 1997, Pecharsky and Gschneidner discovered the reversible giant magnetocaloric effect (GMCE) in the compound Gd5Si2Ge2 [8], renewing the optimism in this technology and attracting the attention of materials scientists from different areas to its study. They indirectly measured an entropy change of 18 J Kg−1 K−1 at 278 K for a field change of 0–5 T [8]. Later, Morellon and co-workers [9] discovered that the origin of such a giant effect is the simultaneous magnetic and structural (magnetostructural) transition that Gd5Si2Ge2 undergoes at 278 K and that can also be triggered by the application of a magnetic field. The magnetostructural transition ensures a first-order nature of the magnetic transition and it is the reason for the giant ∂M/∂T value and consequently for the maximization of ΔSm. Moreover, since a magnetostructural transition occurs, it is important to consider the additional entropy change contribution due to the structural transition, ΔSstr [10]. Since ΔSm accounts for the purely magnetic driven processes, the total entropy variation is: ΔSt = ΔSm + ΔSstr. Because of the reasons mentioned above it became clear that first-order magnetostructural transitions are a requirement for the GMCE, and very rapidly several materials were found that fulfill this requirement, such as NiMnGa [11]; NiMnInCo [12]; MnCoGeB [13], MnFePGe [14] and MnAs [15]. There are other materials presenting GMCE, that do not exhibit any structural transition (in the sense that their symmetry group does not change), yet still undergo large simultaneous changes in their unit cell volume and in their magnetization, as for instance La(FeSi)13 [16], La(FeSi)13H [17] and LaCaMnO3 [18]. It was found that materials undergoing magnetostructural/magnetovolume transitions provide other giant magneto-responsive properties with technological interest besides the giant magnetocaloric effect, such as giant magnetostriction [19] and magnetoresistance [20]. It is important to remark that for these materials the magnetostriction is not a consequence of the reorientation of magnetic domains, as in the classical sense of magnetostriction, but instead it is a result of the structural transition induced by a magnetic field, which gives rise to sharp changes in the lattice parameters and consequently to macroscopic shape and/or volume changes of the material.
Thus, magnetostructural transitions have been attracting an increasing number of scientists, from both experimental and theoretical backgrounds, due to the multifunctionality arising from these effects. In systems exhibiting magnetostructural transitions there is a strong competition between the magnetic and crystallographic degrees of freedom involved in the transition, which is very sensitive to extrinsic parameters—such as temperature, magnetic field, pressure—and intrinsic properties—disorder (chemical and strain) and microstructure [1,2124]. It is very difficult to theoretically predict the influence of intrinsic properties, but the effect of external parameters, such as magnetic field and temperature, can be predicted. In 2006, Paudyal and co-workers [25] developed a pioneering work based on the fundamental thermodynamic equations and on results from first principles calculations, which allowed the theoretical estimation of the temperature dependencies of several magnetic properties (including the MCE) of the Gd5Si2Ge2 compound. They were able to plot the free energies of both structures in competition as a function of temperature and from the behavior of these curves they were able to estimate whether there is a magnetostructural transition, as well as the temperature at which it occurs. Recently, the same approach was extended to the Tb5(Si,Ge)4 system allowing a better understanding on the (de)coupled nature of the magnetic and structural transitions [26,27]. Moreover, first principles calculations have also shown to be a fundamental tool to understand the correlation between electronic and magnetic interactions within the unit cell of the R5(Si,Ge)4compounds [26,28].
It is noteworthy that whether there is a magnetostructural transition or not is a critical feature both from the scientific point of view, in order to understand the origins of such large magnetic responses such as the magnetocaloric, magnetoresistive and magnetostrictive effects, but also technologically, since the occurrence of a magnetostructural transition promotes a giant enhancement on the magnitude of all these effects and specially the MCE. This statement is very patent in the Gschneidner and co-workers report were they show that ΔSstr can account for more than half of the total entropy variation for applied magnetic fields below 2 T in Gd5(Si,Ge)4 compounds and its contribution can be even higher for other compounds. This means that for materials where no magnetostructural transition is observed, in principle it should be possible to greatly optimize their MCE through the strengthening of their magnetic and structural coupling. Simultaneously, it also means that for materials where a magnetostructural transition is observed, the path to optimize their MCE should not be focused in increasing their magnetic and structural coupling.
In this review we will focus on the nature of a first order magnetic transition from the point of view of a competition between two crystallographic phases. The thermodynamic framework will be introduced in combination with first principles calculations. Afterwards we will discuss the importance of knowing the magnetic transition temperatures of both phases involved and a general method to estimate their values through simple experimental measurements. The practical cases of Gd5Si2Ge2 and Tb5Si2Ge2 will be presented to explain the (de)coupling of the magnetic and structural phase transitions in both compounds. By incorporating the aforementioned external parameters, the plots of the free energy as a function of temperature will be presented and the effect of an applied magnetic field will be discussed. Finally, the correlation between this phase competition and one exotic phenomenon—the Griffiths-like phase—will also be discussed and presented as a fingerprint of systems with strong competition between phases.

2. Thermodynamics of First Order Phase Transitions

2.1. General Approach

According to the Ehrenfest classification, a first order phase transition must exhibit a discontinuity in the entropy variation and in the order parameter which arises from a crossover between two free energies, resulting in two distinct states: a low temperature (LT) and a high temperature (HT) state. It is the change between these two states that leads to the discontinuity of the entropy and the maximization of the entropy variation, leading to a giant magnetocaloric effect in the case of magnetic materials. Besides temperature, it is crucial that the application/removal of the external parameters (pressure, magnetic field, etc.) can be made near the phase transition region to be able to modify the system. To understand this process thermodynamically, let us consider the free energy for a simple magnetic material [25]:
F ( T , σ ) = U T S = U lattice H g J σ 3 2 J J + 1 k B T C σ 2 T S m T S lattice
where σ is the reduced magnetization, the first term (Ulattice) is the non-magnetic internal energy at T = 0 K, the second term is the interaction energy between a localized magnetic moment J and the applied magnetic field H (Zeeman term), the third one is the magnetic ion-ion exchange interaction energy and the last two terms are the entropic contributions (magnetic and lattice, respectively) to the system’s free energy.
With the help of first principles calculations, one can estimate the free energy of any structure and, in particular, of any two structures (say 1 and 2) involved in a structural transition 1→2 (where 1 and 2 correspond to the HT and LT magnetization states, respectively): ΔF = F[1] − F[2]. This will be equal to the difference between the Ulattice of both structures at T = 0 K and H = 0 T, which are estimated from first principles calculations. Considering the mean field theory, the magnetic entropy per magnetic ion (SM) for a general J can be estimated from:
S M ( σ ) = k B ln ( sinh ( 2 J + 1 2 J B J 1 ( σ ) ) sinh ( 1 2 J B J 1 ( σ ) ) ) σ B J 1 ( σ ) ~ k B [ ln ( 2 J + 1 ) 3 2 J J + 1 σ 2 + O ( σ 4 ) ]
where B J 1is the inverse Brillouin function. The magnetic equation of state (σ(T,H)) can be obtained by finding the zero of the derivative of Equation (3) with respect to σ [leading to σ = σ(T,H)], which corresponds to the condition of minimum free energy. This minimum free energy is then given by Fmin(T,H) = F(T, σ(T,H)). From Equation (3) one can see that the free energy difference between structures 1 and 2 arises from a) the non-magnetic contribution (Flattice= Ulattice − TSlattice) and b) from the exchange and magnetic entropy terms (2nd, 3rd and 4th terms in Equation (3)), given the hypothetically different TC of the two structures. Therefore it becomes critically important to know the TC of each structure involved in order to be able to plot the free energies of both structures as a function of temperature. Such a plot will allow one to understand whether or not there is a structural transition occurring and, if so, at what temperature it should occur. The first obstacle in this approximation is the determination of the magnetic ordering temperatures ( T C 1 and T C 2) of both structures involved in a phase transition.

2.2. How to Inspect the “Hidden” Magnetic Ordering Temperatures

In standard materials exhibiting second-order magnetic transitions, their magnetic long range ordering sets in at a temperature that can be easily assessed by measuring their magnetization as a function of temperature at constant magnetic field, analysing its derivative and signalling the temperature at which a maximum/minimum occurs. Unfortunately, such a simple approach is not appropriate in systems presenting magnetostructural transitions, since in these systems the structural transition occurs somewhere in between TC1 and TC2, i.e., TC1 < TS < TC2. Let us consider the M(T) measurement on cooling of such an ideal system. Starting with a low magnetization value associated with the paramagnetic state of structure 1, there will be no significant changes on M(T) when cooling through TC2. Then, when T = TS, a sharp increase towards a much higher magnetization value will be observed, which is already associated with the ferromagnetic behavior of structure 2. While continuing on cooling down the system, no major change will be observed in the M(T) curve, not even when crossing the lower Curie temperature of structure 1, TC1, because the system is already transformed into the structure 2. In conclusion, the TC1 and TC2 will be invisible to the M(T) measurement and that is why we have to figure out an alternative method to estimate TC1 and TC2.
In order to be direct and clear, two practical examples will be considered. Both are materials presenting giant magnetocaloric effect: Gd5Si2Ge2—with a simultaneous magnetic and structural transition—and Tb5Si2Ge2—presenting non-simultaneous magnetic and structural transitions. Particularly for these two materials, the structures 1 and 2 mentioned above correspond to the monoclinic (M) and orthorhombic [O(I)] structures, whereas the two magnetic states considered are paramagnetic (PM) and ferromagnetic (FM). The remaining relevant information is summarized in Table 1.
Although these compounds were already thoroughly studied [8,23,32,33], in order to exemplify the broad range of application of this procedure, we will only take into account the fact that both materials undergo structural transitions under temperature and/or magnetic field changes. We will also assume that for both structures involved, the FM state is the most stable at low temperatures. Hence, with these assumptions, there are four different possible states: [M;PMM], [M;FMM], [O(I);PMO(I)] and [O(I);FMO(I)] at any temperature and field value. Now, let us consider that the temperature of the material (Tb5Si2Ge2 or Gd5Si2Ge2) is slightly above its structural transition temperature TS, where the material stabilizes at its [M;PMM] state. At this temperature it is possible to induce the magnetostructural transition, i.e., the transition from the [M;PMM] state to the [O(I);FMO(I)] state through the application of a strong enough magnetic field (critical field, HC). Hence, for high magnetic fields (H > HC) it is possible to inspect the magnetic behavior of the O(I) structure, even at T > TS. In fact, as expected, the critical field HC depends on the temperature of the material and conversely the structural transition temperature TS depends on the applied magnetic field intensity: TS(H) ≠ TS ≡ TS(H = 0). Experimentally, such behavior can be more clearly understood in the isothermal magnetization curves as a function of applied field for several different temperatures (below and above TS). Two clearly distinct types of curves will be obtained: for T < TS isotherms, a typical FM steep increase in the magnetization up to the saturation field is observed, whereas for T > TS the magnetization exhibits two regimes: a low magnetization (LM) state at low fields undergoes a sharp metamagnetic transition into a high magnetization (HM) state at H > HC as a consequence of the very different magnetic susceptibilities of structures M and O(I). The data are then represented in the form of Arrott plots [34]—H/M as a function of M2-, as can be seen in Figures 1a,b for Gd5Si2Ge2 and Tb5Si2Ge2, respectively. This representation is based on the Landau theory of phase transitions and follows from the magnetic state equation [34]:
H M = α ( T T C ) + BM 2
From this expression one recognizes that when plotting the data in the H/M vs. M2 representation, linear curves are expected, intercepting the positive y-axis when T > TC (in the PM state) and the positive x-axis when T < TC (in the FM state). In this representation, the critical isotherm (T = TC) is a straight line that goes through the origin. The intercept with the positive y-axis, α(T-TC), is the inverse susceptibility χ−1(T). Let us consider the Gd5Si2Ge2 case. As can be seen in Figure 1a, the curves in the green region exhibit two different linear regimes. This is a signature that they belong to the TS(H) > T >TS(H = 0) temperature range, where the magnetic field applied is able to induce the structural transition [M,PMM]→[O(I),FMO(I)]. M has lower magnetic susceptibility (smoother slope) than the O(I) structure (steeper slope) (inset of Figure 1a). Now, the squared spontaneous magnetization M S 2 [ O ( I ) ] of the O(I) structure (or HM state) can be estimated as the interception of the linear extrapolation of the high field magnetization data with the abscissa axis at a fixed temperature. In the white region—T < Ts temperature range—only one linear regime is observed because the system is already in the [O(I), FMO(I)] state. By following the same procedure, the M S 2 [ O ( I ) ] values at lower temperatures are estimated. By combining the M S 2 [ O ( I ) ] values obtained for T < TS and for T > TS, it is possible to plot M s 2 [O(I)] as a function of temperature as depicted in Figure 2a (blue dots). Simultaneously, the inverse paramagnetic susceptibility of the M structure (LM state) can also be determined from the low field isothermal curves (T > Ts) and their intercept with the positive y-axis, yielding χ−1[M](T), which is also displayed in Figure 2a) (red dots). Finally, by fitting the M s 2 [O(I)](T) and the χ−1[M](T) curves with Brillouin and linear functions, respectively, it is possible to estimate the Curie temperatures of both M and O(I) phases. In this example the fits result in the following critical temperatures for O(I) and M structures: T C O ( I ) ~ 308 K and T C M ~ 251 K, in accordance with previous results [25,27,35,36]
The Arrott plots for Tb5Si2Ge2 are shown in Figure 1b. As in the Gd5Si2Ge2 plots, the green area highlights the temperature range where the magnetic field is able to induce the M to O(I) phase transition. At T = 124 K (inset of Figure 1b), the behavior of the curves signals the magnetostructural transition from the LM [M, PMM] to the HM [O(I), FMO(I)] state. Following the same procedure as in the Gd5Si2Ge2 case, the data analysis allows extracting the temperature dependence of both the inverse paramagnetic susceptibility of the M structure (red dots in Figure 2b) and the spontaneous magnetization of the O(I) structure (blue dots in Figure 2b). Again, by fitting these data, the extracted values are: T C O ( I ) ~ 200 K and T C M ~ 112 K, where the latter one is coincident with the TC ∼ 112 K obtained from the maximum derivative of the experimentally measured M(T) curves [29].
In summary, the general procedure to extract the “hidden” magnetic ordering temperatures of structures 1 and 2 is as follows:
(1)
Measure the magnetization as a function of magnetic field for several temperatures in the range where TS is expected (a broader interval will provide more points and hence more accurate fits), ensuring that for some isotherms the field is strong enough to induce the metamagnetic transition;
(2)
Replot the data in the Arrott form (H/M as a function of M2) and identify the two different linear regimes associated with LM (1) and HM (2) states for T > TS;
(3)
Perform linear fits to these two different regimes and estimate their interceptions with the positive y-axis (for the LM state) and with the positive x-axis (for the HM state) in order to extract, respectively, the inverse susceptibility of the LM (1) state—χ−1(LM)—and the squared spontaneous magnetization of the HM (2) phase—M2S (HM).;
(4)
Plot χ−1(1) and M s 2 (2) as a function of temperature;
(5)
Through linear and Brillouin fits to these curves, the magnetic ordering temperatures of both 1 and 2 structures are extracted.
To the best of our knowledge, this procedure has only been applied to the R5(Si,Ge)4 family of compounds. However, it could be useful for studying several other materials displaying magnetic field-induced structural transitions, hence with extremely sensitive magneto-responsive properties that are attractive for technological applications. Examples of such materials are the multifunctional magnetic shape memory alloy Ni55Mn20Ga25 [11], the MnNiFeGe family [12], MnCoGeB [13], MnFePGe [14], and the materials undergoing giant volume changes within the same crystallographic structure: La-Fe-Si [16] and LaCaMnO3 [18].

2.3. Free Energy Crossings of Real Systems

Once the TC of both structures is estimated from the Arrott plots using the method described above, their free energies can be calculated as a function of temperature. Then, the temperature at which the free energy curves of both structures intersect will define the structural transition temperature TS. As a first approximation, we neglect the lattice entropy term and consider the magnetic contribution to the free energy in Equation (3), as well as the result of first principles calculations that yields F min lattice ( T = 0 K ) = U lattice (the first term in Equation (3)). In order to normalize the energy range to zero, the value of the free energy at 0 K of structure 2 is subtracted from the free energy F min lattice of both structures, resulting in: Δ F min ( T ) [ 1 , 2 ] = F min ( T ) [ 1 , 2 ] F min lattice ( T = 0 K ) [ 2 ]. Let us start with a zero magnetic field, meaning that the second term in Equation 3 is also zero. The temperature dependence of the magnetic free energies ΔFmin[1] and ΔFmin[2] (that start at zero because of the subtraction mentioned above) are represented for the two examples Gd5Si2Ge2 and Tb5Si2Ge2 in Figures 3a,b, respectively. We recall that structure 1 corresponds to the monoclinic (M) crystallographic phase and structure 2 corresponds to orthorhombic O(I) phase. The first principles calculations were performed using Wien2K code, via the augmented plane wave method with local orbitals (APW + lo) and using the local spin-density approximation method allowing a correct estimation of the free energies of the M and O(I) phase at T = 0 K.
It can be clearly seen that in both cases there is an intersection of the ΔFmin[O(I)] and ΔFmin[M] curves at a temperature T = TS, i.e., the structural transition temperature for each system. In the Gd5Si2Ge2 case (Figure 3a), TS ∼ 265 K lies right between the two magnetic transition temperatures, TCM < TS < TC°(I), which means that upon cooling the system undergoes simultaneously a magnetic—paramagnetic to ferromagnetic [PMM]→[FMO(I)]—and a structural transition from a monoclinic to an orthorhombic O(I) structure [M]→[O(I)]. It is worth noting that, magnetically, the system goes from the monoclinic paramagnetic state to the orthorhombic ferromagnetic state, i.e., there is a large discontinuity on both the atomic and lattice parameters and the magnetization values and these two transitions are coupled, meaning that a magnetostructural transition [M, PMM]→[O(I), FMO(I)] takes place. In the case of Tb5Si2Ge2, TS < TCM < TCO(I), meaning that on cooling the system first undergoes the [PMM]→[FMM] magnetic transition at TCM ∼ 112 K, closely followed by the M→O(I) structural transition a few kelvins below, at TS ∼ 95 K, i.e., the two transitions are decoupled. In contrast with Gd5Si2Ge2, here the magnetic transition is from the PM to the FM state of the monoclinic structure. The TS values estimated with this theoretical model corroborate the values obtained from experimental measurements (particularly X-ray diffraction as a function of temperature) for these two compounds [29,30]. This demonstrates that such theoretical procedure can predict structural transition temperatures and their coupled/decoupled nature. As mentioned in the introduction section, this information is of key importance, since it allows evaluating the potential of the material regarding the optimization of its MCE. In particular, from the results of the two cases presented here it becomes clear that: (1) the MCE in Tb5Si2Ge2 can still be greatly increased, whereas (2) Gd5Si2Ge2 MCE cannot be improved much further by the further enhancement of its magnetic and structural coupling. Two of the most common ways to enhance this coupling is by applying pressure and by alloying. So far, our predictions have been confirmed by several works: concerning: (1), an MCE enhancement of more than 40% was reported by promoting the fully coupled magnetic and structural transitions, both by alloying with Fe [37] and by applying pressure [24], whereas regarding (2), the MCE was not improved (in fact it was smaller) when the same alloying and pressure were applied [38,39].
The T C M < T S < T C O ( I ) scenario, or more generally TC1 < TS < TC2, is more attractive from the magnetic entropy change maximization point of view. As previously discussed by Liu [40], Pecharsky, Gschneidner and co-workers [21,41], the magnetic entropy change peak ( Δ S max m) will be higher for systems where simultaneously Δ T C = T C 2 T C 1 is maximized and TS is just slightly above T C 1. In this scenario, the magnetization will jump (ΔM) from a nearly zero value to almost the maximum value of the magnetization of structure 2 (ΔM will be proportional to the difference Δ T C S = T C 2 T S), hence maximizing the ∂M/∂T at TS and consequently the Δ S max m.
Now, let us turn on the magnetic field. As can be seen in Equation (3), the second term (HgJσ) is always positive, but it is preceded by a minus sign. Therefore, the effect of a non-zero magnetic field is to lower the free energy. However, due to the different TC on each crystal structure (that critically influence the magnetization σ(TC)), the free energy of the structure with higher TC (O(I), in both materials) decreases more rapidly than the one with lower TC (M). In this way the application of a magnetic field tends to increase the temperature at which the free energy crossover occurs—i.e., where ΔFmin[O(I)] = ΔFmin[M]—as can be confirmed in Figure 4. Therefore, the structural transition temperature TS(H) is increased. This was confirmed experimentally and a linear dependence of the critical magnetic field as a function of temperature HC(T) was observed, as shown in the insets of Figure 2a,b.
So far, the effect of Flattice = TSlattice was still not considered in this analysis. To analyze the contribution of the entropic lattice term, the Debye approximation is used. The entropy per ion is given by the following expression [40]:
S lattice ( Θ ) = 3 k B ln ( 1 e Ω T ) + 12 k B ( T Θ ) 3 0 Θ T x 3 e x 1 d x ~ 3 k B ln ( Θ T ) + 3 40 k B ( Θ T ) 2 + O [ ( Θ T ) 3 ]
where Θ is the (cut-off) Debye temperature. From this expression one can infer that the high temperature phase entropic contribution to the lattice free energy (−TSlattice) is lower for the structure with the lower Θ. In fact, the Debye temperature for the M phase is lower than of the O(I) phase. This can be confirmed from the relation Θ M Θ O ( I ) = ( 1 η Δ V V ), where η is the Grüneisen parameter (typically around 10 for Gd5Si2Ge2) and Δ V V is the relative volume increase when the system goes through a O(I) to M structural change (typically ∼1%). An estimation of ΘM considering the experimentally detected velocity of sound in Gd5Si2Ge2 resulted in ΘM ∼ 250 K [42] and consequently Θ O ( I ) = Θ M ( 1 η Δ V V ) ~ 278 K. The lower Θ for the M phase was also experimentally confirmed in the case of Y5(SixGe1−x)4 system [43]. Furthermore, with the estimated Θ for each structure of Gd5Si2Ge2M and ΘO(I)) it is possible to plot the lattice contribution from each structure to the free energy. This is displayed in Figure 4b. It can be seen in this figure that at high temperature the entropic lattice contribution also tends to stabilize the M phase, as shown for the Y5(SixGe1−x)4 system in [44]. Therefore it is clear, at least for these two examples, that including the lattice contribution (Flattice = TSlattice) will not drastically affect the temperature dependence of the total free energy of each structure and, consequently, the main findings about the (de)coupled transition temperature discussed above.
However, the entropy change associated with the structural transition (ΔSstr) cannot be ignored, as it plays a major role on the total entropy variation (ΔST) of the majority of the systems undergoing magnetic field-induced structural transitions. As Pecharsky and Gschneidner [41] noticed, the lattice contribution to the total entropy change, ΔSstr, varies from system to system because it depends on the entropy differences between both the initial and final structures. They also show that ΔSstr is proportional to the volume difference between the two structures. Two examples are given: ΔSstr is around 50% of ΔST for the Gd5Si2.09Ge1.91 compound and about 93% of ΔST for MnCoGeB0.2 under a magnetic field change of 5 T [41]. Furthermore, whereas the magnetic entropy change ( Δ S max m) increases with H2/3 [45], the relation between ΔSstr and H is harder to unveil. However, for these two systems a magnetic field change of 5 T is enough to complete the structural transition and hence ΔSstr reaches its saturation value.

3. Is the Griffiths-like Phase a Requisite for a Giant Magnetocaloric Effect?

Together with the strong magneto-responsive properties, magnetic materials presenting strong coupling between structural and magnetic properties exhibit other exotic effects, such as the appearance of Griffiths-like phases. After the model proposed by Griffiths in the 1960s, the presence of Griffiths-like phases in several real materials has been recently reported [46]. In his model, Griffiths came upon with a diluted Ising ferromagnetic system with a well-defined magnetic ordering temperature TC which was obviously lower than the magnetic ordering temperature of the undiluted (or pure) ferromagnetic system (TG), i.e., TC < TG. He was able to show that for TC < T < TG it was possible that the reciprocal magnetic susceptibility of the system could deviate from its expected Curie-Weiss linear behavior. In fact, this deviation would start, on cooling, exactly at TG, as if the diluted system somehow “remembered” that its magnetic ordering temperature was TG “before” it became diluted [33]. Therefore, the concept of a Griffiths phase, also named Griffiths singularity, is based on the emergence of a new magnetic system within the paramagnetic matrix state, which evidences ferromagnetic short-range ordering [4648]. Although these short-range clusters would not reveal themselves in the typical M(T) curves, they would be responsible for slight deviations from the linear behavior of the reciprocal magnetic susceptibility as a function of temperature.
In the past few years several R5T4 compounds (R = rare earth: Gd, Tb, Dy, and Ho; T = group 14 elements) have shown a Griffiths-like behavior [33,4850]. Such phenomenon was firstly reported in the Tb5Si2Ge2 compound, when it was associated with the experimentally observed deviation from the expected linearity (from Curie-Weiss law) of the χ−1 (T) curve in the paramagnetic region, T > TC [33].
Later, different works, reported by Ouyang, Tian, Pereira and co-workers identified the Griffiths-like phase in other compounds of the R5T4 family [23,4850]. All these compounds presented anomalies featuring their reciprocal susceptibility curve: a “stair-like” behavior occurring in an intermediate temperature region, TC < T < TG. Later on, our group has completed the phase diagrams of the R5(SixGe1−x) (R = Gd, Tb, Dy and Ho) compounds with their corresponding Griffiths-like phase [50]. It was observed that despite the different main temperatures along their composition for each R compound (TG, TC and TN), the phase diagrams were very similar in shape and two main conclusions could immediately be drawn: (1) the Griffiths singularity behavior appears only in the M and O(II) phases for compositions below a characteristic Si concentration that was defined as xp and (2) the O(II) structure is stabilized when the ratio τTC,N/TG is ∼ 0.5, regardless of the rare-earth element. The third and more general conclusion was reached when these diagrams were replotted as τ as a function of (x−xp) and the collapse of all four phase diagrams into a single curve was observed, as represented in Figure 5. The physical mechanism behind the appearance of the Griffiths-like phase is the same, regardless of the rare-earth element, and it is deeply associated with the Si/Ge ratio. In fact, the Ge substitution of Si makes a perfect analogy for the original undiluted/diluted system Griffiths idealized, i.e., when no Ge is present or its concentration is such that x > xp (xp will be defined below), the system stabilizes in the O(I) structure where all interslab bonds are formed. A good review on the particular structural features of this system is given by Miller [1]. In the O(I) structure there is a strong and collinear ferromagnetic state with the highest TC for each phase diagram—the system can be thought of as undiluted/pure. When the Ge substitution leads to a Si critical concentration (xp is defined as this Si concentration value) the interslab bonds start to break (it is hypothesized that it is when x = xp that the Ge atoms start substituting Si atoms at the interslab sites) and the O(II) or M structures become the most stable structures, with a simultaneous decrease in TC. In these conditions the system can be thought of as diluted. It is well-known from previous studies that the magnetism in these compounds exists from the competition of ferromagnetic (within each slab) and antiferromagnetic interactions (mostly between slabs). More accurately, at the interslab region, the main magnetic exchange mechanism is the Ruderman-Kittel-Kasuya-Yosida (RKKY) exchange. The RKKY magnetic exchange energy parameter JRKKY is a Bessel function which starts at a positive value (ferromagnetic interaction) but decays fast with increasing distance between the atomic moments, crossing zero and becoming negative (antiferromagnetic interaction) at a critical distance. Therefore, when Ge starts replacing Si at the interslab sites (at xp), and since it has a higher atomic radius, it promotes an increase in the distance between consecutive slabs. Because of the RKKY interaction, this leads to a decrease of the magnetic exchange energy and consequently to decrease in TC. When a critical Ge concentration limit, xC, is reached, the RKKY interaction becomes negative and hence the AFM state becomes the most stable one. The xC is the ferromagnetic percolation limit for these diluted systems, below which it is not possible to achieve long range ferromagnetic order. Finally, the complete analogy with the Griffiths diluted ferromagnetic system is schematized in Figure 5b.
With this set of comprehensive works the extent of the Griffiths-like behavior on the R5(Si,Ge)4 compounds is shown, highlighting their magnetic nature as that of a diluted ferromagnetic system where the Ge substitution is the dilution mechanism—and hence demonstrating why the R5(Si,Ge)4 are an ideal case study for the Griffiths phase phenomenon. In addition it was found that in all the R5(Si,Ge)4 materials presenting a Griffiths-like phase [Gd5(Si,Ge)4; Tb5(Si,Ge)4; Dy5(Si,Ge)4 and Ho5(Si,Ge)4], TG is systematically equal to the TC of the O(I) phase. This has led to the conclusion that in the R5(Si,Ge)4 compounds, the low temperature phase can easily be assessed through simple measurements of magnetization as a function of temperature at constant magnetic field. In summary, it can be stated that that the R5(Si,Ge)4 compounds exhibiting Griffiths-like phase are most likely to present strong competition between two phases and consequently the Griffiths-like phase constitutes a fingerprint for strong magneto-responsive properties in this family of compounds. In fact, a rapid review of the literature allows us to identify several other strongly-responsive magnetic materials, such as transition metal oxides (particularly manganites [51,52], cuprates [52] and ACr2X4 spinels [53]) which exhibit colossal magnetoresistance and magnetoelasticity [54], magnetic semiconductors [55] and giant magnetocaloric compounds [33], all of which exhibit the Griffiths-like phase. This outcome allows us to suggest that the Griffiths-like phase can be a consequence of the strong correlation between magnetic and atomic lattices and of the strong competition between states in other materials beside the R5(Si,Ge)4 compounds. However, further investigation should be done in order to confirm this suggestion.

4. Conclusions

In summary, in this work we have demonstrated the importance of the strong competition between phases on the origin of several giant magneto-responsive effects, particularly the magnetocaloric effect. Such competition is responsible for an additional contribution to the entropy change that can be studied thoroughly through a combination of detailed measurement/analysis of the materials magnetic properties and first principles calculations. A method to analyze the Arrott plots in order to extract the hidden magnetic ordering temperatures is reviewed and generalized. Such parameters are shown to be crucial as inputs for the simple model of the free energy as a function of temperature and for a deeper understanding on the nature of the (de)coupled magnetic and structural phase transitions. Practical examples are given, namely by explaining the coupling of magnetic and structural transitions in Gd5Si2Ge2 or their decoupling in the Tb5Si2Ge2 case. Finally, the Griffiths-like phases are reviewed in the R5(Si,Ge)4 families and a correlation between their universal phase diagram and the large magneto-responsive properties is discussed.

Acknowledgments

The authors acknowledge to FCT for financial support in the project PTDC/CTM-NAN/115125/2009, EXPL/EMS-ENE/2315/2013 and Projeto Norte-070124-FEDER-000070. We would also like to thank Erik Kampert, José Manue Moreira, Uli Zeitler, João Nuno Gonçalves, João Amaral, Vítor Amaral and João Bessa Sousa for the important previous discussions related with this work.

Author Contributions

First two authors contributed equally to this work, particularly on the writing and on the bibliographic review. André Miguel Pereira and João Pedro Araújo were responsible for this work and supervised the manuscript. Isabel T. Gomes’ help was fundamental for a more clear and improved manuscript. Armandina Maria Lima Lopes, Luis Morellón, Cesar Magen, Pedro Antonio Algarabel, Manuel Ricardo Ibarra provided important inputs on the scientific analysis of the work. All authors have read and approved the final manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Miller, G.J. Complex rare-earth tetrelides, RE5(SixGe(1−x))4: New materials for magnetic refrigeration and a superb playground for solid state chemistry. Chem. Soc. Rev 2006, 35, 799–813. [Google Scholar]
  2. Smith, A. Who discovered the magnetocaloric effect? Eur. Phys. J. H 2013, 38, 507–517. [Google Scholar]
  3. Gschneidner, K.A., Jr.; Pecharsky, V.K.; Tsokol, A.O. Recent developments in magnetocaloric materials. Rep. Prog. Phys 2005, 68, 1479–1539. [Google Scholar]
  4. Giauque, W.; MacDougall, D. Attainment of Temperatures Below 1° Absolute by Demagnetization of Gd2(SO4)3·8H2O. Phys. Rev 1933, 43, 768–768. [Google Scholar]
  5. Brown, G.V. Magnetic heat pumping near room temperature. J. Appl. Phys 1976, 47, 3673. [Google Scholar]
  6. Gschneidner, K.A.; Pecharsky, V.K. Thirty years of near room temperature magnetic cooling: Where we are today and future prospects. Int. J. Refrig 2008, 31, 945–961. [Google Scholar]
  7. Brown, D.R.; Dirks, J.A.; Fernandez, N.; Stout, T.B. The Prospects of Alternatives to Vapor Compression Technology for Space Cooling and Food Refrigeration Applications. Energ. Eng 2012, 109, 7–20. [Google Scholar]
  8. Pecharsky, V.K.; Gschneidner, K.A. Giant Magnetocaloric Effect in Gd5(Si2Ge2). Phys. Rev. Lett 1997, 78, 3–6. [Google Scholar]
  9. Morellon, L.; Algarabel, P.A.; Ibarra, M.R.; Blasco, J.; Garcia-Landa, B.; Arnold, Z.; Albertini, F. Magnetic-field-induced structural phase transition in Gd5(Si1.8Ge2.2). Phys. Rev. B 1998, 58, 721–724. [Google Scholar]
  10. Pecharsky, V.K.; Holm, A.P.; Gschneidner, K.A.; Rink, R. Massive Magnetic-Field-Induced Structural Transformation in Gd5Ge4 and the Nature of the Giant Magnetocaloric Effect. Phys. Rev. Lett 2003, 91, 197204. [Google Scholar]
  11. Pasquale, M.; Sasso, C.; Lewis, L.; Giudici, L.; Lograsso, T.; Schlagel, D. Magnetostructural transition and magnetocaloric effect in Ni55Mn20Ga25 single crystals. Phys. Rev. B 2005, 72, 094435. [Google Scholar]
  12. Liu, J.; Gottschall, T.; Skokov, K.P.; Moore, J.D.; Gutfleisch, O. Giant magnetocaloric effect driven by structural transitions. Nat. Mater 2012, 11, 620–626. [Google Scholar]
  13. Trung, N.T.; Zhang, L.; Caron, L.; Buschow, K.H.J.; Brück, E. Giant magnetocaloric effects by tailoring the phase transitions. Appl. Phys. Lett 2010, 96, 172504. [Google Scholar]
  14. Trung, N.T.; Ou, Z.Q.; Gortenmulder, T.J.; Tegus, O.; Buschow, K.H.J.; Brück, E. Tunable thermal hysteresis in MnFe(P,Ge) compounds. Appl. Phys. Lett 2009, 94, 102513. [Google Scholar]
  15. Wada, H.; Tanabe, Y. Giant magnetocaloric effect of MnAs1−xSbx. Appl. Phys. Lett 2001, 79, 3302. [Google Scholar]
  16. Shen, B.G.; Sun, J.R.; Hu, F.X.; Zhang, H.W.; Cheng, Z.H. Recent Progress in Exploring Magnetocaloric Materials. Adv. Mater 2009, 21, 4545–4564. [Google Scholar]
  17. Lyubina, J.; Nenkov, K.; Schultz, L.; Gutfleisch, O. Multiple Metamagnetic Transitions in the Magnetic Refrigerant La(Fe,Si)13Hx. Phys. Rev. Lett 2008, 101, 177203. [Google Scholar]
  18. Ibarra, M.R.; Algarabel, P.A.; Marquina, C.; Blasco, J.; García, J. Large Magnetovolume Effect in Yttrium Doped La-Ca-Mn-O Perovskite. Phys. Rev. Lett 1995, 75, 3541–3544. [Google Scholar]
  19. Hadimani, R.L.; Bartlett, P.A.; Melikhov, Y.; Snyder, J.E.; Jiles, D.C. Field and temperature induced colossal strain in Gd5(SixGe1−x)4. J. Magn. Magn. Mater 2011, 323, 532–534. [Google Scholar]
  20. Morellon, L.; Stankiewicz, J.; García-Landa, B.; Algarabel, P.A.; Ibarra, M.R. Giant magnetoresistance near the magnetostructural transition in Gd5Si1.8Ge2.2. Appl. Phys. Lett 1998, 73, 3462–3464. [Google Scholar]
  21. Pecharsky, V.K.; Gschneidner, K.A.; Mudryk, Y.; Paudyal, D. Making the most of the magnetic and lattice entropy changes. J. Magn. Magn. Mater 2009, 321, 3541–3547. [Google Scholar]
  22. Pecharsky, V.K.; Gschneidner, K.A., Jr. Gd5(SixGe1−x)4: An Extremum Material. Adv. Mater 2001, 13, 683–686. [Google Scholar]
  23. Belo, J.H.; Pereira, A.M.; Araújo, J.P.; de la Cruz, C.; dos Santos, A.M.; Gonçalves, J.N.; Amaral, V.S.; Morellon, L.; Ibarra, M.R.; Algarabel, P.A.; et al. Tailoring the magnetism of Tb5Si2Ge2 compounds by La substitution. Phys. Rev. B 2012, 86, 014403. [Google Scholar]
  24. Morellon, L.; Arnold, Z.; Magen, C.; Ritter, C.; Prokhnenko, O.; Skorokhod, Y.; Algarabel, P.A.; Ibarra, M.R.; Kamarad, J. Pressure Enhancement of the Giant Magnetocaloric Effect in Tb5Si2Ge2. Phys. Rev. Lett 2004, 93, 137201. [Google Scholar]
  25. Paudyal, D.; Pecharsky, V.; Gschneidner, K.; Harmon, B. Electron correlation effects on the magnetostructural transition and magnetocaloric effect in Gd5Si2Ge2. Phys. Rev. B 2006, 73, 144406. [Google Scholar]
  26. Paudyal, D.; Mudryk, Y.; Pecharsky, V.K.; Gschneidner, K.A. Electronic structure, magnetic properties, and magnetostructural transition in Tb5Si2.2Ge1.8 from first principles. Phys. Rev. B 2011, 84, 014421. [Google Scholar]
  27. Pereira, A.M.; Kampert, E.; Moreira, J.M.; Zeitler, U.; Belo, J.H.; Magen, C.; Algarabel, P.A.; Morellon, L.; Ibarra, M.R.; Gonçalves, J.N.; et al. Unveiling the (De)coupling of magnetostructural transition nature in magnetocaloric R5Si2Ge2 (R = Tb, Gd) materials. Appl. Phys. Lett 2011, 99, 132510. [Google Scholar]
  28. Mudryk, Y.; Paudyal, D.; Pecharsky, V.K.; Gschneidner, K.A.; Misra, S.; Miller, G.J. Controlling Magnetism of a Complex Metallic System Using Atomic Individualism. Phys. Rev. Lett 2010, 105, 066401. [Google Scholar]
  29. Morellon, L.; Ritter, C.; Magen, C.; Algarabel, P.; Ibarra, M. Magnetic-martensitic transition of Tb5Si2Ge2 studied with neutron powder diffraction. Phys. Rev. B 2003, 68, 024417. [Google Scholar]
  30. Morellon, L.; Blasco, J.; Algarabel, P.A.; Ibarra, M.R. Nature of the first-order antiferromagnetic-ferromagnetic transition in the Ge-rich magnetocaloric compounds Gd5(SixGe1−x)4. Phys. Rev. B 2000, 62, 1022–1026. [Google Scholar]
  31. Levin, E.M.; Pecharsky, V.K.; Gschneidner, K.A. Unusual Magnetic Behavior in Gd5(Si1.5Ge2.5) and Gd5(Si2Ge2). Phys. Rev. B 2000, 62, 625–628. [Google Scholar]
  32. Casanova, F.; Batlle, X.; Labarta, A.; Marcos, J.; Mañosa, L.; Planes, A. Entropy change and magnetocaloric effect in Gd5(SixGe1−x)4. Phys. Rev. B 2002, 66, 100401. [Google Scholar]
  33. Magen, C.; Algarabel, P.A.; Morellon, L.; Araújo, J.P.; Ritter, C.; Ibarra, M.R.; Pereira, A.M.; Sousa, J.B. Observation of a Griffiths-like Phase in the Magnetocaloric Compound Tb5Si2Ge2. Phys. Rev. Lett 2006, 96, 167201. [Google Scholar]
  34. Arrott, A. Criterion for Ferromagnetism from Observations of Magnetic Isotherms. Phys. Rev 1957, 261, 2–4. [Google Scholar]
  35. Hadimani, R.L.; Melikhov, Y.; Schlagel, D.L.; Lograsso, T.A.; Jiles, D.C. Study of the Second-Order “Hidden” Phase Transition of the Monoclinic Phase in the Mixed Phase Region of Gd5(SixGe1−x)4. IEEE Trans. Magn 2012, 48, 4070–4073. [Google Scholar]
  36. Hadimani, R.L.; Melikhov, Y.; Snyder, J.E.; Jiles, D.C. Estimation of second order phase transition temperature of the orthorhombic phase of Gd5(SixGe1−x)4 using Arrott plots. J. Appl. Phys 2008, 103, 033906. [Google Scholar]
  37. Pereira, A.M.; dos Santos, A.M.; Magen, C.; Sousa, J.B.; Algarabel, P.A.; Ren, Y.; Ritter, C.; Morellon, L.; Ibarra, M.R.; Araújo, J.P. Understanding the role played by Fe on the tuning of magnetocaloric effect in Tb5Si2Ge2. Appl. Phys. Lett 2011, 98, 122501. [Google Scholar]
  38. Provenzano, V.; Shapiro, A.J.; Shull, R.D. Reduction of hysteresis losses in the magnetic refrigerant Gd5Ge2Si2 by the addition of iron. Nature 2004, 429, 853–857. [Google Scholar]
  39. Carvalho, A.M.G.; Alves, C.S.; de Campos, A.; Coelho, A.A.; Gama, S.; Gandra, F.C.G.; von Ranke, P.J.; Oliveira, N.A. The magnetic and magnetocaloric properties of Gd5Ge2Si2 compound under hydrostatic pressure. J. Appl. Phys 2005, 97. [Google Scholar] [CrossRef]
  40. Liu, G.J.; Sun, J.R.; Lin, J.; Xie, Y.W.; Zhao, T.Y.; Zhang, H.W.; Shen, B.G. Entropy changes due to the first-order phase transition in the Gd5SixGe4−x system. Appl. Phys. Lett 2006, 88, 212505. [Google Scholar]
  41. Gschneidner, K.A.; Mudryk, Y.; Pecharsky, V.K. On the nature of the magnetocaloric effect of the first-order magnetostructural transition. Scr. Mater 2012, 67, 572–577. [Google Scholar]
  42. Svitelskiy, O.; Suslov, A.; Schlagel, D.; Lograsso, T.; Gschneidner, K.; Pecharsky, V. Elastic properties of Gd5Si2Ge2 studied with an ultrasonic pulse-echo technique. Phys. Rev. B 2006, 74, 184105. [Google Scholar]
  43. Pecharsky, A.; Gschneidner, K.; Pecharsky, V.; Schlagel, D.; Lograsso, T. Phase relationships and structural, magnetic, and thermodynamic properties of alloys in the pseudobinary Er5Si4-Er5Ge4 system. Phys. Rev. B 2004, 70, 144419. [Google Scholar]
  44. Pecharsky, A.O.; Pecharsky, V.K.; Gschneidner, K.A. Phase relationships and low temperature heat capacities of alloys in the Y5Si4–Y5Ge4 pseudo binary system. J. Alloy. Compd 2004, 379, 127–134. [Google Scholar]
  45. Belo, J.H.; Amaral, J.S.; Pereira, A.M.; Amaral, V.S.; Araujo, J.P. On the Curie temperature dependency of the magnetocaloric effect. Appl. Phys. Lett 2012, 242407, 3–6. [Google Scholar]
  46. Griffiths, R.B. Nonanalytic Behavior Above the Critical Point in a Random Ising Ferromagnet. Phys. Rev. Lett 1969, 23, 17–19. [Google Scholar]
  47. Bray, A. Nature of the Griffiths phase. Phys. Rev. Lett 1987, 59, 586–589. [Google Scholar]
  48. Zou, J.D.; Liu, J.; Mudryk, Y.; Pecharsky, V.K.; Gschneidner, K.A. Ferromagnetic ordering and Griffiths-like phase behavior in Gd5Ge3.9Al0.1. J. Appl. Phys 2013, 114, 063904. [Google Scholar]
  49. Ouyang, Z.W. Griffiths-like behavior in Ge-rich magnetocaloric compounds Gd5(SixGe1−x)4. J. Appl. Phys 2010, 108, 033907. [Google Scholar]
  50. Pereira, A.M.; Morellon, L.; Magen, C.; Ventura, J.; Algarabel, P.A.; Ibarra, M.R.; Sousa, J.B.; Araújo, J.P. Griffiths-like phase of magnetocaloric R5(SixGe1−x)4 (R = Gd, Tb, Dy, and Ho). Phys. Rev. B 2010, 82, 172406. [Google Scholar]
  51. Turcaud, J.A.; Pereira, A.M.; Sandeman, K.G.; Amaral, J.S.; Morrison, K.; Berenov, A.; Daoud-Aladine, A.; Cohen, L.F. Spontaneous magnetization above Tc in La0.7Ca0.3MnO3 and La0.7Ba0.3MnO3 polycrystalline materials. Phys. Rev. B 2014, in press.. [Google Scholar]
  52. Burgy, J.; Mayr, M.; Martin-Mayor, V.; Moreo, A.; Dagotto, E. Colossal Effects in Transition Metal Oxides Caused by Intrinsic Inhomogeneities. Phys. Rev. Lett 2001, 87, 277202. [Google Scholar]
  53. Oliveira, G.N.P.; Pereira, A.M.; Lopes, A.M.L.; Amaral, J.S.; dos Santos, A.M.; Ren, Y.; Mendonça, T.M.; Sousa, C.T.; Amaral, V.S.; Correia, J.G.; et al. Dynamic off-centering of Cr3+ ions and short-range magneto-electric clusters in CdCr2S4. Phys. Rev. B 2012, 86, 224418. [Google Scholar]
  54. Michelmann, M.; Moshnyaga, V.; Samwer, K. Colossal magnetoelastic effects at the phase transition of (La0.6Pr0.4)0.7Ca0.3MnO3. Phys. Rev. B 2012, 85, 014424. [Google Scholar]
  55. Guo, S.; Young, D.; Macaluso, R.; Browne, D.; Henderson, N.; Chan, J.; Henry, L.; DiTusa, J. Discovery of the Griffiths Phase in the Itinerant Magnetic Semiconductor Fe1−xCoxS2. Phys. Rev. Lett 2008, 100, 017209. [Google Scholar]
Figure 1. Arrott plots for Gd5Si2Ge2 (a) in the temperature range 77–350 K; Inset: T = 300 K, and for Tb5Si2Ge2 (b) in the temperature range of 77–305 K; Inset: T = 124 K.
Figure 1. Arrott plots for Gd5Si2Ge2 (a) in the temperature range 77–350 K; Inset: T = 300 K, and for Tb5Si2Ge2 (b) in the temperature range of 77–305 K; Inset: T = 124 K.
Entropy 16 03813f1
Figure 2. Spontaneous magnetization (blue dots) and reciprocal susceptibility (red dots) curves of: (a) Gd5Si2Ge2 and (b) Tb5Si2Ge2, obtained from the Arrott plots. The dark green dotted lines are the Brillouin fits to the spontaneous magnetization of the O(I) structure for each compound which enable the estimation of their TC: T C O ( I ) = 308 K for Gd5Si2Ge2 and T C O ( I ) = 200 K for Tb5Si2Ge2. The red dashed lines are the linear fits to the reciprocal susceptibility data, which allowed estimating the TCM for each compound: T C M = 251 K for Gd5Si2Ge2 and T C M = 112 K for Tb5Si2Ge2, whereas the purple dotted lines represent the Brillouin curves assuming the obtained T C M. The gray zones represent the FM regions and their boundaries (represented by the light-green curves) schematically represent the experimentally measured M(T) curves. Insets: temperature dependence of the critical magnetic fields (HC(T)) obtained from the maximum of dM/dH.
Figure 2. Spontaneous magnetization (blue dots) and reciprocal susceptibility (red dots) curves of: (a) Gd5Si2Ge2 and (b) Tb5Si2Ge2, obtained from the Arrott plots. The dark green dotted lines are the Brillouin fits to the spontaneous magnetization of the O(I) structure for each compound which enable the estimation of their TC: T C O ( I ) = 308 K for Gd5Si2Ge2 and T C O ( I ) = 200 K for Tb5Si2Ge2. The red dashed lines are the linear fits to the reciprocal susceptibility data, which allowed estimating the TCM for each compound: T C M = 251 K for Gd5Si2Ge2 and T C M = 112 K for Tb5Si2Ge2, whereas the purple dotted lines represent the Brillouin curves assuming the obtained T C M. The gray zones represent the FM regions and their boundaries (represented by the light-green curves) schematically represent the experimentally measured M(T) curves. Insets: temperature dependence of the critical magnetic fields (HC(T)) obtained from the maximum of dM/dH.
Entropy 16 03813f2
Figure 3. Free energy as a function of temperature considering the magnetic entropy and the first principles calculations (Ulattice) for both M and O(I) structures in the (a) Gd5Si2Ge2 and (b) Tb5Si2Ge2 materials. TS is defined as the temperature at which the crossing ΔFmin[M] = ΔFmin[O(I)] occurs.
Figure 3. Free energy as a function of temperature considering the magnetic entropy and the first principles calculations (Ulattice) for both M and O(I) structures in the (a) Gd5Si2Ge2 and (b) Tb5Si2Ge2 materials. TS is defined as the temperature at which the crossing ΔFmin[M] = ΔFmin[O(I)] occurs.
Entropy 16 03813f3
Figure 4. (a) Schematic free energy of M and O(I) structures as a function of temperature for the zero (upper pair of curves) and non-zero (lower pair of curves) magnetic field; (b) Lattice contribution to the free energy of M and O(I) structures of Gd5Si2Ge2 compound as a function of temperature.
Figure 4. (a) Schematic free energy of M and O(I) structures as a function of temperature for the zero (upper pair of curves) and non-zero (lower pair of curves) magnetic field; (b) Lattice contribution to the free energy of M and O(I) structures of Gd5Si2Ge2 compound as a function of temperature.
Entropy 16 03813f4
Figure 5. (a) Universal x,T phase diagram of the R5(SixGe1−x)4 compounds with R = Gd (triangles), Tb (squares), Dy (circles), and Ho (hexagons); (b) Adapted T vs. (x − xp) phase diagram of a dilute FM system for the R5(SixGe1−x)4 system [50].
Figure 5. (a) Universal x,T phase diagram of the R5(SixGe1−x)4 compounds with R = Gd (triangles), Tb (squares), Dy (circles), and Ho (hexagons); (b) Adapted T vs. (x − xp) phase diagram of a dilute FM system for the R5(SixGe1−x)4 system [50].
Entropy 16 03813f5
Table 1. Relevant information about Tb5Si2Ge2 and Gd5Si2Ge2 compounds.
Table 1. Relevant information about Tb5Si2Ge2 and Gd5Si2Ge2 compounds.
Tb5Si2Ge2Gd5Si2Ge2
Simultaneous transition×
Structure 1/2M/O(I)M/O(I)
Curie temperatures, Tc (K)112/200 [27,29]208/305 [25,27]
Structural transition temperature, TS(K)95 [29]270 [30,31]

Share and Cite

MDPI and ACS Style

Pires, A.L.; Belo, J.H.; Lopes, A.M.L.; Gomes, I.T.; Morellón, L.; Magen, C.; Algarabel, P.A.; Ibarra, M.R.; Pereira, A.M.; Araújo, J.P. Phase Competitions behind the Giant Magnetic Entropy Variation: Gd5Si2Ge2 and Tb5Si2Ge2 Case Studies. Entropy 2014, 16, 3813-3831. https://doi.org/10.3390/e16073813

AMA Style

Pires AL, Belo JH, Lopes AML, Gomes IT, Morellón L, Magen C, Algarabel PA, Ibarra MR, Pereira AM, Araújo JP. Phase Competitions behind the Giant Magnetic Entropy Variation: Gd5Si2Ge2 and Tb5Si2Ge2 Case Studies. Entropy. 2014; 16(7):3813-3831. https://doi.org/10.3390/e16073813

Chicago/Turabian Style

Pires, Ana Lúcia, João Horta Belo, Armandina Maria Lima Lopes, Isabel T. Gomes, Luis Morellón, Cesar Magen, Pedro Antonio Algarabel, Manuel Ricardo Ibarra, André Miguel Pereira, and João Pedro Araújo. 2014. "Phase Competitions behind the Giant Magnetic Entropy Variation: Gd5Si2Ge2 and Tb5Si2Ge2 Case Studies" Entropy 16, no. 7: 3813-3831. https://doi.org/10.3390/e16073813

Article Metrics

Back to TopTop