Next Article in Journal
Covalent Organic Frameworks as Nanocarriers for Improved Delivery of Chemotherapeutic Agents
Next Article in Special Issue
Influence of Nd Substitution on the Phase Constitution in (Zr,Ce)Fe10Si2 Alloys with the ThMn12 Structure
Previous Article in Journal
Effects of Yttrium Doping on Erbium-Based Hydroxyapatites: Theoretical and Experimental Study
Previous Article in Special Issue
Some Thermomagnetic and Mechanical Properties of Amorphous Fe75Zr4Ti3Cu1B17 Ribbons
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Entalpy of Mixing, Microstructure, Structural, Thermomagnetic and Mechanical Properties of Binary Gd-Pb Alloys

1
Department of Physics, Częstochowa University of Technology, Armii Krajowej 19, 42-200 Czestochowa, Poland
2
Department of Mechanics, Materials and Biomedical Engineering, Wroclaw University of Science and Technology, Smoluchowskiego 25, 50-370 Wroclaw, Poland
3
Institute of Experimental Physics, Slovak Academy of Sciences, Watsonova 47, 040 01 Kosice, Slovakia
*
Author to whom correspondence should be addressed.
Materials 2022, 15(20), 7213; https://doi.org/10.3390/ma15207213
Submission received: 24 September 2022 / Revised: 12 October 2022 / Accepted: 13 October 2022 / Published: 16 October 2022

Abstract

:
The aim of the present work is to study the phase composition, microstructure and magnetocaloric effect of binary Gd100−xPbx (where x = 5, 10, 15 and 20) alloys. The XRD and SEM/EDX analysis confirmed a biphasic structure built by Gd(Pb) and Gd5Pb3 phases. The analysis of M vs. T curves showed the evolution of the Curie point of recognized phases. The temperature dependences of magnetic entropy change revealed two maxima corresponding to the recognized phases. The analysis of the exponent n (ΔSMmax = C(Bmax)n) confirmed the multiphase composition of the produced alloys. The same behavior was also observed in investigations of mechanical properties.

1. Introduction

An important factor related to the use of conventional refrigerators is their disadvantageous impact on the natural environment. Here, the application of greenhouse gases as refrigerants in home-used refrigerators is one of the most important arguments for a search for alternative methods of cooling. Nowadays, a well-known temperature-lowering technique with the highest efficiency is magnetic cooling. Magnetic cooling is based on the magnetocaloric effect (MCE) discovered more than one hundred years ago by Warburg [1,2]. A natural magnetocaloric material (MCM) is pure gadolinium with the Curie temperature at 294 K and magnetic entropy change ΔSM ~10 J (kg K)−1 under the change of external magnetic field ~5T [3]. The MCE phenomenon is still a hot topic in the field of magnetic materials, since the discovery of the giant magnetocaloric effect (GMCE) by Pecharsky and Gscheidner Jr. (in 1997) in Gd5Si2Ge2 alloys [4]. Next to the Gd5Si2Ge2 alloy, such materials as La(Fe,Si)13 [5,6] or heusler alloys [7,8,9] also exhibit GMCE. An MCM with a relatively high probability of practical application should characterize a wide temperature working range. It could be realized by using bi- or multiphase material [6,10]. Law et al. showed the possibility a one-stage production of biphase material based on pure Gd with inclusions of second phase GdZn3 characterized by different Curie temperatures in comparison to pure Gd [10]. Mo et al. investigated the MCE in Gd55Co35Mn10 alloy ribbons [11]. They identified a coexistence of two phases with different Curie temperatures, which allowed to reach the table-like shape of the temperature dependence of magnetic entropy change. On the other hand, Jayaraman et al. presented successive studies of partial substitution of Gd by Mn [12]. The XRD studies carried out in [12] revealed an occurrence of only Gd, where the structure contracted with an increase of Mn content. Moreover, a monotonic decrease of the ΔSM was detected with the rise of Mn addition. The studies carried out by Huo et al. in [13] concerning the MCE in binary GdPd alloys revealed an almost two times higher value of the ΔSM compared with two pure Gd. It is well-known that Mn and Pd have a nonzero magnetic moment, which could modify the magnetic structure of a material. References [14,15] showed a possibility of producing biphasic material which caused a significant increase of refrigeration capacity (RC). Moreover, Xu and coworkers in [16] showed that Gd-based alloys are still intensively studied, taking into account their magnetocaloric properties. As was presented by Palenzona and Cirafici in [17], Gd mixes very well with Pb in the whole range of composition. Lead has a nonzero magnetic moment equal to 0.59219 μB and, similar to Pd, could improve magnetic moment alloys. Based on these results [14,17], we investigated a phase constitution and the MCE of binary Gd100−xPbx.

2. Sample Preparation and Experimental Details

The ingot samples with nominal composition Gd100−xPbx (where x = 5, 10, 15, 20) were produced by arc-melting of the high-purity constituent elements (at least 3N). Samples were remelted several times to ensure their homogeneity. The X-ray diffraction studies were carried out using Bruker D8 Advance diffractometer with CuKα radiation and semiconductor detector LynxEye. The XRD investigation was supported by the Rietveld analysis realized using PowderCell 2.4 software [18]. Microstructure of prepared samples was observed using scanning electron microscope SEM JEOL JSM 6610LV equipped with the energy dispersive X-ray spectrometer (EDX). Magnetic measurements were carried out using a vibrating sample magnetometer (VSM) installed on a cryogen-free superconducting magnet (Cryogenic Limited, London, UK) working in a wide temperature range (2–320 K) and magnetic fields up to 18 T. The MCE was measured indirectly based on M vs. H curves collected in a wide range of temperatures.
The mechanical properties of the Gd95Pb5, Gd90Pb10 and Gd85Pb15 alloys, performed with the help of the Nanoindentation Tester (NHT2, CSM Instruments) equipped with Berkovich tip, were studied for the map of 15 × 15 imprints with 15 μm gaps covered and an area of 240 μm × 240 μm. The load-controlled nanoindentation tests were carried out with a maximum load of 100 mN. The Continuous Multi Cycles (CMC) mode was used to investigate the effect of penetration depth on the hardness and Young’s modulus. The maximum load for these measurements was 150 mN. Numerical analysis of the recorded load-displacement curves allowed the estimation of both the plastic and elastic properties of the materials, including hardness (HIT) and Young’s modulus (EIT), as well as elastic to total deformation energy ratio (nIT) during indentation.

3. Results and Discussion

As was shown in reference [16], Gd and Pb mix very well in the whole range of composition. In order to confirm these observations, a semi-empirical Miedema’s model [19,20,21,22] was used to determine the enthalpy of mixing of the Gd-Pb system. Miedema’s model treats the atom as a block with the Wigner–Seitz cell. During alloying, element A is solved in a matrix built by element B, which causes changes in the value of enthalpy. Description of this effect is possible, taking into account three main quantities: (1) the molar volume given as V; (2) potential, which is close to the electron work function; and (3) density at the boundary of Wigner–Seitz cell nWS. To start consideration concerning on enthalpy of mixing, it is extremely important to determine interfacial enthalpy Hinter(AinB) for dissolving one mole of A atoms in B matrix. The interfacial enthalpy is described by the following formula:
Δ H i n t e r ( A i n B ) = V A 2 / 3 1 2 ( 1 n W S A 1 / 3 + 1 n W S A 1 / 3 ) { P ( Δ ϕ ) 2 + Q ( Δ n W S 1 / 3 ) 2 }
where P and Q are empirical constants related to alloying elements.
The formation enthalpy of the solid solution could be defined as:
Δ H s s = Δ H c h e m + Δ H e l a s t + Δ H s t r u c t
where ΔHchem is the chemical part due to the mixing of components and can be written as:
Δ H c h e m = c A c B ( c B s Δ H i n t e r ( A i n B ) + c A s Δ H i n t e r ( B i n A ) )
where cA and cB mean fractions of A and B elements, whereas cAs and cbs are the surface factors given by the following dependences:
c A s = c A V A 2 / 3 c A V A 2 / 3 + c B V B 2 / 3
c B s = c B V B 2 / 3 c A V A 2 / 3 + c B V B 2 / 3
Elastic part of enthalpy (see Equation (2)) ΔHelast is related to the atom size mismatch and is written as:
Δ H e l a s t = c A c B ( c B Δ H e l a s t ( A i n B ) + c A Δ H e l a s t ( B i n A ) )
The elastic misfit energy of A atoms solved in the excess of B atoms is described by the following relation:
Δ H e l a s t ( A i n B ) = 2 K A G B ( W A i n B W B i n A ) 2 4 G B W A i n B + 3 K A W B i n A
where K means bulk modulus, G is shear modulus and WA/B are molar volumes corrected by electron transfer.
In Equation (2), the structural part of enthalpy ΔHstruct exists. However, this contribution, originated from the valence and crystal structure of solvent and solute atom, has minimal effect and it is difficult to calculate. As a result, the structural part of enthalpy can be neglected [19,20].
Δ H s t r u c t 0
In Figure 1, the formation enthalpy of binary Gd-Pb alloy is presented. It is clearly seen that in the whole range, the sign of enthalpy is negative, which suggests that both elements mix very well. The minimum is observed for almost equiatomic composition. This means that the alloying of Gd with Pb does not depend on the percentage content of any element.
In Figure 2, the XRD patterns of all studied samples are depicted. The analysis of the XRD pattern revealed a biphasic structure, which was built by hexagonal Gd(Pb) (space group P63/mmc no. 194) and hexagonal Gd5Pb3 (Mn5Si3- type, space group P63/mcm no. 193) phases. Moreover, a small amount of the cubic GdO1.5 (space group Fm3m no. 225) phase was detected, the presence of which was probably caused by oxidation of the sample surface during measurements. The results of the Rietveld analysis are collected in Table 1.
The lattice parameters for pure Gd are a = b = 3.636 Å and c = 5.783 Å, while for the Gd5Pb3, they are a = b = 9.12 Å and c = 6.668 Å. The values of lattice parameters are less than for nominal phases; however, some increase of them is visible with an increase of Pb in the alloy composition. The atomic radius of Gd is 188 pm, while the atomic radius of Pb is 180 pm. Taking into account that Pb dissolves very well in the Gd matrix, it is expected to observe a lower lattice parameter of Gd(Pb). A similar effect was described in Gd5Si2-xGe2-xPb2x alloys by Zhuang et al. in [23]. An increase of Pb causes the formation of Gd5Pb3 and less dissolving in the Gd matrix. Moreover, the coexistence of Gd(Pb) and Gd5Pb3 phases in the studied range were observed in [16,24]. Demel and Gschneidner in reference [24] revealed the evolution of the microstructure during the alloying of Gd by Pb. The SEM/EDX results are shown in Figure 3 for all studied samples in the as-cast state, which confirms the results presented in reference [24]. In the case of a sample of Gd95Pb5 alloy, the microstructure is almost homogenized with some small precipitations. These precipitations contain the Gd5Pb3 phase, while the other part is built by the Gd(Pb) phase. An increase of Pb up to 10 at. % changed the microstructure and “islands” of the Gd(Pb) phase formed surrounded by the Gd5Pb3 phase, similar to the observations in [14]. For the highest contents of Pb in alloy composition (15 and 20 at. %), the eutectic Gd(Pb) + Gd5Pb3 is clearly seen, which confirms the results delivered in [24]. In Table 2, the EDX results for all samples are collected. It is noticeable that the solubility of Pb in Gd decreases with an increase of Pb in the alloy composition. Taking into account the XRD results, it is clear that Pb is used in the formation of the Gd5Pb3 phase, whereby the content rises with an increase of Pb in the alloy composition.
In Figure 4a, the temperature dependences of magnetization were depicted. Based on the XRD investigation and microstructure analysis, the step shape of the collection was expected. In order to reveal the Curie temperature, the first derivative (dM/dT) of M vs. T curves were calculated and shown in Figure 4b. The characteristic minima and visible additional humps in dM/dT curves suggested an occurrence of two different Curie temperatures in each sample, corresponding to the Gd(Pb) and Gd5Pb3 phases. Determined Curie points are collected in Table 3. The sample of Gd95Pb5 revealed two Curie temperatures at 243 and 258 K, corresponding to Gd5Pb3 and Gd(Pb) phases, respectively. In the case of the Gd90Pb10 alloy, the analysis led to the determination of temperatures of 252 (Gd5Pb3) and 263 K (Gd(Pb)), while for the Gd85Pb15 alloy sample it was 274 (Gd5Pb3) and 285 K (Gd(Pb)). The Curie points revealed for the Gd80Pb20 were 275 Gd5Pb3 and 293 K (Gd(Pb)). As was shown by Dan’kov and coworkers in [25], the Curie temperature of Gd is strongly dependent on the purity and other elements dissolved in the Gd matrix. They showed that the Curie point of Gd is observed in the range of temperatures 289–295 K. In the present work, we observed changes in the range 258–293 K. However, it is worth highlighting that in our case Pb dissolved in the Gd matrix and changed the lattice parameters and probably magnetic structure. The results of the TC analysis suggest that substitution of Gd by Pb in the unit cell of Gd caused the weakening of magnetic interactions. The magnetic moment of Gd is 7.94 μB, while for Pb it is 0.59219 μB, which results in the decrease of the Curie temperature of the Gd(Pd) phase. An increase of the TC with the rise of Pb in alloy composition is also noticeable; however, the XRD analysis revealed an increase of lattice constant and EDX showed a decrease of Pb content in the Gd(Pb) phase. The lowering of Pb in the Gd matrix increased the TC. Marcinkova et al. in [26] presented results concerning the magnetic response of Gd5Pb3. The Curie temperature revealed in this paper [26] was 275 K obtained for stoichiometric composition. In our case, the stoichiometric composition of the Gd5Pb3 phase was achieved in the Gd80Pb20 alloy. Moreover, the delivered lattice constants of this phase were lower than for the nominal phase. Accordingly, the weakening of magnetic interactions was detected, resulting in a decrease in the TC to 243 K in the Gd95Pb5 alloy sample. In contradiction to the Gd(Pb) phase, an increase of Pb content in the Gd5Pb3 phase until the almost stoichiometric composition was revealed by the EDX technique. It probably caused an increase of the Curie point of the Gd5Pb3 phase from 243 to 274 K.
The magnetocaloric effect was measured indirectly based on M vs. μ0H curves collected in a wide range of temperatures. The calculations of magnetic entropy change ΔSM were performed based on the thermomagnetic Maxwell relation [27]:
Δ S M ( T , Δ H ) = μ 0 0 H ( M ( T , H ) T ) H d H ,
where T is temperature, μ0 is magnetic permeability, H is magnetic field strength and M is magnetization.
The calculated temperature dependences of magnetization are shown in Figure 5. Broad peaks are clearly visible. In the case of samples Gd95Pb5, Gd90Pb10 and Gd80Pb20, the second maximum is also noticeable. The highest value of ΔSM (6.36 J/(kg K)) was observed for the Gd95Pb5 alloy sample and related to the Gd(Pb) phase. A characteristic shift of peak toward higher temperatures was detected with an increase of Pb in the alloy. Moreover, the intensity of each peak was changed, this one corresponding to Gd(Pb) decrease, while this one corresponding to Gd5Pb3 growth. Such observations are in agreement with the results delivered by XRD and M(T) measurements. The maximum values of ΔSM of each alloy are collected in Table 4.
For good MCM, next to the magnetic entropy change ΔSM, the refrigerant capacity (RC) is also extremely important. The RC gives us information about the amount of energy produced by the magnetocaloric substance. In order to calculate the RC, the Wood–Potter relation should be used [28]:
R C ( δ T , H M A X ) = T c o l d T h o t Δ S M ( T , H M A X ) d T
where RC is refrigerant capacity, ΔT = ThotTcold is the temperature range of the thermodynamic cycle (ΔT corresponds to the full width at half maximum of magnetic entropy change peak) and HMAX is the maximum value of the external magnetic field.
The results of calculations of RC are collected in Table 4 and visualization is depicted in Figure 6. The highest RC values were determined for the sample Gd95Pb5 alloy. In contradiction to results delivered for Gd100−xPdx alloys [14], the increase of the RC with an increased volume fraction of the second phase was not observed. In the studied alloys, the peaks of the ΔSM are separated by relatively low temperatures and they are overlapping. Moreover, the value of ΔSM for samples of Gd95Pb5 and Gd90Pb10 are mainly related to the Gd(Pb) phase. In the case of the Gd85Pb15 alloy sample, the peaks of ΔSM corresponding to recognized phases overlapped and their separation is not possible. As a result, the enhancement of the RC is not observed. The obtained values are less than for pure Gd. Obtained results showed that it is possible to produce biphasic materials with quite a large temperature range of magnetic cooling. It is extremely important to take into account the practical application as an active magnetic regenerator. The calculated values are almost two times lower than those measured for pure Gd, which achieves 5.4 and 10.6 J/(kg K), for the change of external magnetic field 2 and 5T, respectively [25].
In Table 5, revealed values of the magnetic entropy change are compared to other magnetocaloric materials. The values are slightly less or comparable with most Gd-based alloys.
The magnetic entropy change ΔSM strongly depends on, next to the temperature, the external magnetic field and it is clearly seen in Figure 5. Franco and coworkers in references [31,32] developed the phenomenological temperature dependence of magnetic entropy change based on the power law, which is written in the following relation:
Δ S M max = C ( B M A X ) n
where C is a constant depending on temperature and n is the exponent related to the magnetic state of the material.
In order to calculate the exponent n, the Equation (11) should be rewritten in the following form [33]:
ln Δ S M max = ln C + n ln ( B max )
Such simple modifications allowed the use of linear fitting due to the fact that the slope of linear dependence is directly n exponent.
As was shown in [31,32], the n exponent is strongly dependent on the magnetic state of the sample. Taking into account that material obeys the Curie–Weiss law, the exponent n = 1 in the ferromagnetic state, 2 in paramagnetic state, and in critical temperature (TC) is given by the following equation:
n = 1 + 1 δ ( 1 1 β )
where β and Δ are critical exponents.
In mean field theory, the values of critical exponents equal β = 0.5, γ = 1 and Δ = 3. Based on these values, the n exponent at the Curie temperature is 0.67. In Figure 7, the temperature dependences of exponent n are presented. Presented curves are typical for samples manifested by the second order phase transition. The characteristic shift of all curves toward higher temperatures is visible. Moreover, wide minima are seen in all figures, which is related to the multiphase composition of the produced samples. The relatively broad minima were observed for samples Gd90Pb10 and Gd85Pb15, which was caused by overlapping minima corresponding to phases recognized in material. The values of n below TC (ferromagnetic state) are higher than 1, while above TC (paramagnetic state), they never achieve 2. Relatively broad minima could not allow the determination of minima corresponding to each phase. Moreover, the values of minima were higher than the theoretical value of 0.67. The above observations are caused by the multiphase composition of the produced alloys.
The mechanical properties of materials are one of the most important aspects that determine their potential application. Below, the results of hardness distributions (HVIT), Young’s modulus (EIT), as well as the ratio of elastic to total energy (nIT) recorded during deformation of samples obtained from numerical analysis of the recorded load-displacement curves (performed according to the Oliver and Pharr protocol [34,35]) are presented. To visualize the obtained results, the same x-axis scale of HVIT, EIT and nIT on all histograms was used.
Figure 8 presents the hardness distributions as histograms (left side) and mapping graphs (right side) for the Gd100−xPbx (where x = 5, 10 and 15) alloys. To analyze the above-presented histograms, statistical methods for the Gaussian distribution were used. It can be well-seen that the bimodal distribution was observed for all investigated materials. This behavior is connected to the microstructure of the materials. The two-phase nature of the investigated materials was also recognized in X-ray studies (Table 2). The results presented above clearly show that with increasing Pb content from 5 to 15 at. % in the sample, a decrease in hardness was observed. In the Gd95Pb5 sample, two phases with hardnesses of 230.48 and 350.07 Vickers for the phase Gd95.73Pb4.27 and Gd88.06Pb11.94, respectively, were distinguished. The crystalline Gd96.71Pb3.29 and Gd82.48Pb17.52 phases observed in the Gd90Pb10 alloy are characterized by a hardness equal to 222.07 and 296.45 Vickers, respectively. The lowest hardness values were determined in the sample with the highest Pb content. In this case, the first component assigned to the Gd70.38Pb29.62 phase achieved a maximum at 98 Vickers, whereas the second component attributed to the Gd96.86Pb3.14 phase at 226 Vickers. These results clearly show that with increasing Pb content in the phase, the hardness decreases.
Figure 9 shows Young’s modulus distributions plotted as histograms (left side) and mapping graphs (right side) for the Gd95Pb5, Gd90Pb10 and Gd85Pb15 alloys. The two-phase nature of the investigated samples is also well-seen in these graphs. The Young’s modulus for the Gd-rich phase is equal to 65.27, 66.33 and 69.52 GPa for the Gd95.73Pb4.27, Gd96.71Pb3.29 and Gd96.86Pb3.14 phase, respectively. The Young’s modulus, determined from the decomposition of histograms presented in Figure 9 (left side), equals to 61.15, 58.93 and 77.67 GPa for the Gd88.06Pb11.94, Gd82.48Pb17.52 and Gd70.38Pb29.62, respectively. The obtained values of EIT clearly show that with increasing Pb content from 11.94 to 29.62 at. % in GdPb phases, Young’s modulus decreases.
Figure 10 presents the elastic to the total deformation energy ratio (nIT) calculated for the Gd95Pb5, Gd90Pb10 and Gd85Pb15 alloys. It is seen that the well-visible two-modal distribution with well-separated components was observed for the Gd90Pb10 alloy. On the other hand, the Gd95Pb5, and Gd85Pb15 alloys also show multimodal character. In this case, the intensity of the components differs significantly. The detailed results of the analysis of nIT histograms are presented in Table 6.
The multimodal behavior of mechanical parameters presented in Table 6 related to the two-phase structure of the Gd95Pb5, Gd90Pb10 and Gd85Pb15 alloys is depicted in Figure 11. Two load-displacement curves (Fn(Pd)) representing the data shown in Table 5 were selected for the Gd100−xPbx (where x = 5, 10 and 15) alloys from the set of 15 × 15 recorded curves. It is well-seen that depth penetration indicated as Pd for the same load of 100 mN is different for the individual phases observed in the examined samples. The most visible difference in Pd is for the Gd85Pb15 alloy, in which the highest change in the chemical composition of the phases occurs (Gd96.86Pb3.14 and Gd70.38Pb29.62). The disturbed character of the curves at the initial stages of Pd (as it is shown in Figure 11) is related to the occurrence of gadolinium oxides GdO1.5 (Figure 2, Table 1) on the surface of the samples.
Figure 12 presents the depth profile investigations of Young’s modulus and hardness for the Gd95Pb5, Gd90Pb10 and Gd85Pb15 alloys performed by nanoindentation in Continuous Multi cycles (CMC) mode. All investigations were carried out only for the Gd-rich phases detected in the Gd100−xPbx (x = 5, 10 and 15) alloys.
It is seen that hardness HVIT decreases exponentially with penetration depth Pd and for the load of 150 mN reach about HVIT = 63 GPa for all investigated materials. Young’s modulus EIT also decreases with Pd as is seen in Figure 12. The highest value of about EIT = 72 GPa was obtained for the Pb-rich alloy, whereas the lowest one was for the Gd-rich alloy. These results are in good agreement with the data presented in Figure 9 and Table 6.

4. Conclusions

In this paper, the phase composition and magnetocaloric properties of Gd100−xPbx (where x = 5, 10, 15 and 20) alloys were studied. The theoretical calculations of enthalpy of mixing, based on the semi-empirical Miedema’s model, confirmed the dissolving of Pb in the Gd matrix in the whole range. The XRD studies revealed the phase structure of the produced alloys and the coexistence of hexagonal Gd(Pb), hexagonal Gd5Pb3 and cubic GdO1.5 phases. This was confirmed by observations of microstructure and EDX analysis. The evolution of the microstructure with an increase of Pb in alloy composition was observed. Moreover, the changes in the value of the Curie temperature were revealed for each recognized phase. The temperature dependences of ΔSM showed two peaks and their decrease with an increase of Pb. Furthermore, the decrease of RC was determined with an increase of Pb in the alloy composition. The temperature evolution of n exponent confirmed the multiphase composition of the produced alloys. The investigations of mechanical properties by nanoindentation tests also confirmed the multiphase nature of the studied alloys and allowed us to determine the mechanical parameters for every single phase discovered in the Gd100−xPbx (where x = 5, 10 and 15) alloys. It was shown that the hardness, Young modulus and elastic to total deformation energy ratio vary with the chemical composition of the phases. The Gd-rich phases observed in the studied alloys show comparable hardnesses, while their Young’s modulus increases with increasing Pb content in the chemical composition of the alloys and phases.

Author Contributions

Conceptualization, P.G.; Data curation, P.G., M.H., J.K. and M.R.; Formal analysis, M.H.; Funding acquisition, P.G.; Investigation, P.G., M.H., J.K. and M.R.; Methodology, P.G.; Software, J.K.; Supervision, P.G.; Validation, P.G.; Visualization, P.G. and M.H.; Writing—original draft, P.G. and M.H. All authors have read and agreed to the published version of the manuscript.

Funding

This study was funded by the research Project no. 2019/03/X/ST5/00066 in 2019–2020 financed by the Polish National Science Centre. The work was also funded by projects of Slovak Grant Agency APVV 19-0369 and 2/0011/20 and Slovak Research and Development Agency under the Contract Number APVV-18-0160.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Acknowledgments

This study was supported by the research Project no. 2019/03/X/ST5/00066 in 2019–2020 financed by the Polish National Science Centre. P.G. would like to thank the Rector of Częstochowa University of Technology—Norbert Sczygiol for financial support. The work was also supported by projects of Slovak Grant Agency APVV 19-0369 and 2/0011/20 and Slovak Research and Development Agency under the Contract Number APVV-18-0160.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Warburg, E. Magnetische untersuchungen. Ann. Phys. 1881, 249, 141–164. [Google Scholar] [CrossRef] [Green Version]
  2. Tishin, A.M.; Spichkin, Y.I. The Magnetocaloric Effect and Its Application; Institute of Physics Series in Condensed Matter Physics; CRC Press: Boca Raton, FL, USA, 2003. [Google Scholar]
  3. Gschneidner, K.A., Jr.; Pecharsky, V.K. Magnetic refrigeration in Rare Earths. In Science Technology and Applications, 3rd ed.; Bautista, R.G., Bounds, C.O., Ellis, T.W., Killbourn, B.T., Eds.; The Minerals, Metals and Material Society: Warrendale, PA, USA, 1997; p. 209. [Google Scholar]
  4. Pecharsky, V.K.; Gschneidner, K.A., Jr. Tunable magnetic regenerator alloys with a giant magnetocaloric effect for magnetic refrigeration from ~20 to ~290 K. Appl. Phys. Lett. 1997, 70, 3299–3301. [Google Scholar] [CrossRef]
  5. Liu, X.B.; Altounian, Z. Effect of Co content on magnetic entropy change and structure of La(Fe1xCox)11.4Si1.6. J. Magn. Magn. Mater. 2003, 264, 209–213. [Google Scholar] [CrossRef]
  6. Gębara, P.; Pawlik, P. Broadening of temperature working range in magnetocaloric La(Fe,Co,Si)13-based multicomposite. J. Magn. Magn. Mater. 2017, 442, 145–151. [Google Scholar] [CrossRef]
  7. Yang, Z.; Cong, D.Y.; Huang, L.; Nie, Z.H.; Sun, X.M.; Zhang, Q.Z.; Wang, Y.D. Large elastocaloric effect in a Ni-Co-Mn-Sn magnetic shape memory alloy. Mater. Des. 2016, 92, 932–936. [Google Scholar] [CrossRef]
  8. Li, Y.W.; Zhang, H.; Tao, K.; Wang, Y.X.; Wu, M.L.; Long, Y. Giant magnetocaloric effect induced by reemergence of magnetostructural coupling in Si-doped Mn0.95CoGe compounds. Mater. Des. 2017, 114, 410–415. [Google Scholar] [CrossRef]
  9. Tegus, O.; Brück, E.; Buschow, K.H.J.; de Boer, F.R. Transition-metal-based magnetic refrigerants for room-temperature applications. Nature 2002, 415, 150–152. [Google Scholar] [CrossRef]
  10. Law, J.Y.; Moreno-Ramirez, L.M.; Blazquez, J.S.; Franco, V.; Conde, A. Gd + GdZn biphasic magnetic composites synthesized in a single preparation step: Increasing refrigerant capacity without decresing magnetic entropy change. J. Alloys Compd. 2016, 675, 244–247. [Google Scholar] [CrossRef]
  11. Mo, H.Y.; Zhong, X.C.; Jiao, D.L.; Liu, Z.W.; Zhang, H.; Qiu, W.Q.; Ramanujan, R.V. Table-like magnetocaloric effect and enhanced refrigerant capacity in crystalline Gd55Co35Mn10 alloy melt spun ribbons. Phys. Lett. A 2018, 382, 1679–1684. [Google Scholar] [CrossRef]
  12. Jayaraman, T.V.; Boone, L.; Shield, J.E. Magnetocaloric effect and refrigerant capacity in melt-spun Gd–Mn alloys. J. Magn. Magn. Mater. 2013, 345, 153–158. [Google Scholar] [CrossRef]
  13. Huo, J.J.; Du, Y.S.; Cheng, G.; Wu, X.F.; Ma, L.; Wang, J.; Xia, Z.; Rao, G. Magnetic properties and large magnetocaloric effects of GdPd intermetallic compound. J. Rare Earths 2018, 36, 1044–1049. [Google Scholar] [CrossRef]
  14. Gębara, P.; Diaz-Garcia, A.; Law, J.Y.; Franco, V. Magnetocaloric response of binary Gd-Pd and ternary Gd-(Mn, Pd) alloys. J. Magn. Magn. Mater. 2020, 500, 166175. [Google Scholar] [CrossRef]
  15. Diaz-Garcia, A.; Law, J.Y.; Gębara, P.; Franco, V. Phase deconvolution of multiphasic materials by the universal scaling of the magnetocaloric effect. JOM 2020, 72, 2845–2852. [Google Scholar] [CrossRef]
  16. Xu, P.; Hu, L.; Zhang, Z.; Wang, H.; Li, L. Electronic structure, magnetic properties and magnetocaloric performance in rare earths (RE) based RE2BaZnO5 (RE = Gd, Dy, Ho, and Er) compounds. Acta Mater. 2022, 236, 118114. [Google Scholar] [CrossRef]
  17. Palenzona, A.; Cirafici, S. The gd-pb (gadolinium-lead) system. J. Phase Equilibria 1991, 12, 686–690. [Google Scholar] [CrossRef]
  18. Kraus, W.; Nolze, G. PowderCell 2.0 for Windows. Powder. Diffr. 1998, 13, 256–259. [Google Scholar]
  19. de Boer, F.R.; Boom, R.; Mattens, W.C.M.; Miedema, A.R.; Niessen, A.K. Cohesion in Metals. In Cohesion and Structure; de Boer, F.R., Pettifor, D.G., Eds.; North Holland Physics: Amsterdam, The Netherlands, 1988; Volume 1, pp. 1–758. [Google Scholar]
  20. Bakker, H. Enthalpies in Alloys: Miedema’s Semi-Empirical Model; Materials Science Foundations 1; Trans Tech Publications Ltd.: Wollerau, Switzerland, 1998; pp. 1–78. [Google Scholar]
  21. Basu, J.; Murty, B.S.; Ranganathan, S. Glass forming ability: Miedema approach to (Zr, Ti, Hf)–(Cu, Ni) binary and ternary alloys. J. Alloys Compd. 2008, 465, 163–172. [Google Scholar] [CrossRef]
  22. Sun, S.P.; Yi, D.Q.; Liu, H.Q.; Zang, B.; Jiang, Y. Calculation of glass forming ranges in Al–Ni–RE (Ce, La, Y) ternary alloys and their sub-binaries based on Miedema’s model. J. Alloys Compd. 2010, 506, 377–387. [Google Scholar] [CrossRef]
  23. Zhuang, Y.H.; Li, J.Q.; Huang, W.D.; Sun, W.A.; Ao, W.Q. Giant magnetocaloric effect enhanced by Pb-doping in Gd5Si2Ge2 compound. J. Alloys Compd. 2006, 421, 49–53. [Google Scholar] [CrossRef]
  24. Demel, J.T.; Gschneidner, K.A., Jr. The gadolinium—lead system. J. Nucl. Mater. 1969, 29, 111–120. [Google Scholar] [CrossRef]
  25. Dan’kov, S.Y.; Tishin, A.M.; Pecharsky, V.K.; Gschneidner, K.A., Jr. Magnetic phase transitions and the magnetothermal properties of gadolinium. Phys. Rev. B 1998, 57, 3478–3940. [Google Scholar] [CrossRef]
  26. Marcinkova, A.; de la Cruz, C.; Yip, J.; Zhao, L.L.; Wang, J.K.; Svanidze, E.; Morosan, E. Strong magnetic coupling in the hexagonal R5Pb3 compounds (R=Gd–Tm). J. Magn. Magn. Mater. 2015, 384, 192–203. [Google Scholar] [CrossRef] [Green Version]
  27. Pecharsky, V.K.; Gschneidner, K.A., Jr. Magnetocaloric effect and magnetic refrigeration. J. Magn. Magn. Mater. 1999, 200, 44–56. [Google Scholar] [CrossRef]
  28. Wood, M.E.; Potter, W.H. General analysis of magnetic refrigeration and its optimization using a new concept: Maximization of refrigerant capacity. Cryogenics 1985, 25, 667–683. [Google Scholar] [CrossRef]
  29. la Roca, P.; Lopez-Garcia, J.; Sanchez-Alacros, V.; Recarte, V.; Rodriguez-Velamazan, J.A.; Perez-Landazabal, J.I. Room temperature huge magnetocaloric properties in low hysteresis ordered Cu-doped Ni-Mn-In-Co alloys. J. All. Compd. 2022, 922, 166143. [Google Scholar] [CrossRef]
  30. Tekgül, A.; Acet, M.; Scheibel, F.; Farle, M.; Unal, N. The reversibility of the inverse magnetocaloric effect in Mn2−xCrxSb0.95Ga0.05. Acta Mater. 2017, 124, 93–99. [Google Scholar] [CrossRef]
  31. Franco, V.; Blazquez, J.S.; Conde, A. The influence of Co addition on the magnetocaloric effect of Nanoperm-type amorphous alloys. J. Appl. Phys. 2006, 100, 064307. [Google Scholar] [CrossRef]
  32. Franco, V.; Conde, C.F.; Blázquez, J.S.; Conde, A.; Švec, P.; Janičkovič, D.; Kiss, L.F. A constant magnetocaloric response in Fe Mo Cu B amorphous alloys with different Fe/B ratios. J. Appl. Phys. 2007, 101, 093903. [Google Scholar] [CrossRef]
  33. Świerczek, J. Superparamagnetic behavior and magnetic entropy change in partially crystallized Fe–Mo–Cu–B alloy. Phys. Status Solidi A 2014, 211, 1567–1576. [Google Scholar] [CrossRef]
  34. Oliver, W.C.; Pharr, G.M. An improved technique for determining hardness and elastic modulus using load and displacement sensing indentation experiments. J. Mater. Res. 1992, 7, 1564–1583. [Google Scholar] [CrossRef]
  35. Oliver, W.C.; Pharr, G.M. Measurement of hardness and elastic modulus by instrumented indentation: Advances in understanding and refinements to methodology. J. Mater. Res. 2004, 19, 3–20. [Google Scholar] [CrossRef]
Figure 1. The formation enthalpy of solid solution Gd-Pb binary alloy (a) and selected range used in the present paper (b).
Figure 1. The formation enthalpy of solid solution Gd-Pb binary alloy (a) and selected range used in the present paper (b).
Materials 15 07213 g001
Figure 2. X-ray diffraction pattern for the Gd100−xPbx (where x = 5, 10, 15, 20) alloys.
Figure 2. X-ray diffraction pattern for the Gd100−xPbx (where x = 5, 10, 15, 20) alloys.
Materials 15 07213 g002
Figure 3. SEM microstructure of: (a) Gd95Pb5, (b) Gd90Pb10, (c) Gd85Pb15 and (d) Gd80Pb20. The results of the EDX analysis of the Gd95Pb5 sample (e,f), while (g,h) are for the Gd90Pb10 alloy.
Figure 3. SEM microstructure of: (a) Gd95Pb5, (b) Gd90Pb10, (c) Gd85Pb15 and (d) Gd80Pb20. The results of the EDX analysis of the Gd95Pb5 sample (e,f), while (g,h) are for the Gd90Pb10 alloy.
Materials 15 07213 g003aMaterials 15 07213 g003b
Figure 4. M vs. T curves (a) and corresponding dM/dT dependences for all studied samples (b).
Figure 4. M vs. T curves (a) and corresponding dM/dT dependences for all studied samples (b).
Materials 15 07213 g004
Figure 5. Temperature dependences of magnetic entropy change ΔSM calculated for: Gd95Pb5 (a), Gd90Pb10 (b), Gd85Pb15 (c) and Gd80Pb20 (d) alloys.
Figure 5. Temperature dependences of magnetic entropy change ΔSM calculated for: Gd95Pb5 (a), Gd90Pb10 (b), Gd85Pb15 (c) and Gd80Pb20 (d) alloys.
Materials 15 07213 g005
Figure 6. Magnetic field dependencies of refrigeration capacity of the investigated samples.
Figure 6. Magnetic field dependencies of refrigeration capacity of the investigated samples.
Materials 15 07213 g006
Figure 7. The temperature dependences of exponent n determined for: Gd95Pb5 (a), Gd90Pb10 (b), Gd85Pb15 (c) and Gd80Pb20 (d) alloys.
Figure 7. The temperature dependences of exponent n determined for: Gd95Pb5 (a), Gd90Pb10 (b), Gd85Pb15 (c) and Gd80Pb20 (d) alloys.
Materials 15 07213 g007
Figure 8. Distributions of hardness HVIT with corresponding 2D mapping graphs obtained from nanoindentations test for the Gd95Pb5 (a), Gd90Pb10 (b) and Gd85Pb15 (c) alloys. The dots marked on the map distributions (left side) correspond to the locations of measurements.
Figure 8. Distributions of hardness HVIT with corresponding 2D mapping graphs obtained from nanoindentations test for the Gd95Pb5 (a), Gd90Pb10 (b) and Gd85Pb15 (c) alloys. The dots marked on the map distributions (left side) correspond to the locations of measurements.
Materials 15 07213 g008
Figure 9. Distributions of Young’s modulus EIT with corresponding 2D mapping graphs obtained from nanoindentations test for the Gd95Pb5 (a), Gd90Pb10 (b) and Gd85Pb15 (c) alloys.
Figure 9. Distributions of Young’s modulus EIT with corresponding 2D mapping graphs obtained from nanoindentations test for the Gd95Pb5 (a), Gd90Pb10 (b) and Gd85Pb15 (c) alloys.
Materials 15 07213 g009
Figure 10. Distributions of elastic to total deformation energy nIT with corresponding 2D mapping graphs obtained from nanoindentations test for the Gd95Pb5 (a), Gd90Pb10 (b) and Gd85Pb15 (c) alloys.
Figure 10. Distributions of elastic to total deformation energy nIT with corresponding 2D mapping graphs obtained from nanoindentations test for the Gd95Pb5 (a), Gd90Pb10 (b) and Gd85Pb15 (c) alloys.
Materials 15 07213 g010
Figure 11. The example of load-displacement curves Fn—Pd recorded for Gd-rich and Gd-poor phases present in Gd95Pb5 (a), Gd90Pb10 (b) and Gd85Pb15 (c) alloys reflecting the values exemplified in Table 5.
Figure 11. The example of load-displacement curves Fn—Pd recorded for Gd-rich and Gd-poor phases present in Gd95Pb5 (a), Gd90Pb10 (b) and Gd85Pb15 (c) alloys reflecting the values exemplified in Table 5.
Materials 15 07213 g011
Figure 12. Continuous Multi Cycle load (Fn) versus indentation penetration depth (Pd) curve, and corresponding Young’s modulus (EIT) and hardness (HIT) for the Gd-rich phases discovered in Gd95Pb5 (a), Gd90Pb10 (b) and Gd85Pb15 (c) alloys.
Figure 12. Continuous Multi Cycle load (Fn) versus indentation penetration depth (Pd) curve, and corresponding Young’s modulus (EIT) and hardness (HIT) for the Gd-rich phases discovered in Gd95Pb5 (a), Gd90Pb10 (b) and Gd85Pb15 (c) alloys.
Materials 15 07213 g012
Table 1. The results of the Rietveld analysis.
Table 1. The results of the Rietveld analysis.
AlloyRecognized PhasesContent [vol.%]Lattice Parameters [Å]
Gd95Pb5Gd(Pb)67a = 3.599
c = 5.742
Gd5Pb327a = 9.129
c = 6.675
GdO1.56a = 5.307
Gd90Pb10Gd(Pb)58a = 3.614
c = 5.758
Gd5Pb333a = 9.122
c = 6.665
GdO1.57a = 5.304
Gd85Pb15Gd(Pb)45a = 3.623
c = 5.769
Gd5Pb347a = 9.119
c = 6.661
GdO1.58a = 5.309
Gd80Pb20Gd(Pb)34a = 3.629
c = 5.784
Gd5Pb358a = 9.113
c = 6.657
GdO1.58a = 5.308
Table 2. Chemical composition of recognized phases in all studied alloys.
Table 2. Chemical composition of recognized phases in all studied alloys.
AlloyPhasesWeight Fraction [wt.%]Atomic Fraction [at.%]
Gd95Pb5Gd(Pb)Gd—94.45
Pb—5.55
Gd—95.73
Pb—4.27
Gd5Pb3Gd—84.84
Pb—15.16
Gd—88.06
Pb—11.94
Gd90Pb10Gd(Pb)Gd—95.59
Pb—4.41
Gd—96.71
Pb—3.29
Gd5Pb3Gd—78.13
Pb—21.87
Gd—82.48
Pb—17.52
Gd85Pb15Gd(Pb)Gd—95.9
Pb—4.1
Gd—96.86
Pb—3.14
Gd5Pb3Gd—64.33
Pb—35.67
Gd—70.38
Pb—29.62
Gd80Pb20Gd(Pb)Gd—96.32
Pb—3.68
Gd—97.18
Pb—2.82
Gd5Pb3Gd—60.83
Pb—39.17
Gd—67.17
Pb—32.83
Table 3. The Curie temperature values of recognized phases in the studied alloys.
Table 3. The Curie temperature values of recognized phases in the studied alloys.
AlloyPhasesTC [K]
Gd95Pb5Gd(Pb)258
Gd5Pb3243
Gd90Pb10Gd(Pb)263
Gd5Pb3252
Gd85Pb15Gd(Pb)285
Gd5Pb3274
Gd80Pb20Gd(Pb)293
Gd5Pb3275
Table 4. Values of the ΔSM and RC determined for studied alloys.
Table 4. Values of the ΔSM and RC determined for studied alloys.
AlloyΔ(μ0H)
[T]
ΔSM [J (kg K)−1]RC [J kg−1]
Gd95Pb511.6159
23.04130
34.32189
45.41315
56.36372
Gd90Pb1011.0354
22.12115
33.11192
44.04252
54.81336
Gd85Pb1511.2146
22.2990
33.14149
43.93190
54.58239
Gd80Pb2010.7426
21.4554
32.1288
42.69116
53.25164
Table 5. The magnetic entropy change for the Gd100−xPbx (where x = 5, 10, 15, 20) compared with some other magnetocaloric materials under Δ(μ0H) = 0–2 T.
Table 5. The magnetic entropy change for the Gd100−xPbx (where x = 5, 10, 15, 20) compared with some other magnetocaloric materials under Δ(μ0H) = 0–2 T.
MaterialΔSM [J/(kg K)]Ref.
Gd95Pb53.04This work
Gd90Pb102.12This work
Gd85Pb152.29This work
Gd80Pb201.45This work
Gd95Pd55.25[14]
Gd90Pd104.51[14]
Gd85Pd154.23[14]
Gd80Pd203.15[14]
Pure Gd4.98[10]
Gd75Zn253.29[10]
Gd65Zn353.11[10]
Gd55Zn453.42[10]
Gd50Zn503.25[10]
Ni44.5Mn35.5In13.5Co4Cu2.511.22[29]
Mn1.87Cr0.13Sb0.95Ga0.052.98[30]
Table 6. Hardness (HVIT), Young’s modulus (EIT) and elastic to total energy deformation ratio (nIT) for the individual phases observed in the Gd95Pb5, Gd90Pb10 and Gd85Pb15 alloys.
Table 6. Hardness (HVIT), Young’s modulus (EIT) and elastic to total energy deformation ratio (nIT) for the individual phases observed in the Gd95Pb5, Gd90Pb10 and Gd85Pb15 alloys.
Gd95Pb5Gd90Pb10Gd85Pb15
HVIT (Vickers)1st component (red)
(Phase)
230.48 ± 2.56
(Gd95.73Pb4.27)
222.07 ± 1.25
(Gd96.71Pb3.29)
98.86 ± 8.64
(Gd70.38Pb29.62)
2nd component (green)
(Phase)
350.07 ± 2.98
(Gd88.06Pb11.94)
296.45 ± 2.42
(Gd82.48Pb17.52)
226.49 ± 3.41
(Gd96.86Pb3.14)
EIT (GPa)1st component (red)
(Phase)
65.27 ± 0.15
(Gd95.73Pb4.27)
61.15 ± 0.11
(Gd82.48Pb17.52)
58.93 ± 7.61
(Gd70.38Pb29.62)
2nd component (green)
(Phase)
77.67 ± 0.08
(Gd88.06Pb11.94)
66.33 ± 0.21
(Gd96.71Pb3.29)
69.52 ± 0.51
(Gd96.86Pb3.14)
nIT (%)1st component (red)
(Phase)
19.22 ± 0.18
(Gd88.06Pb11.94)
20.98 ± 0.29
(Gd96.71Pb3.29)
21.69 ± 0.04
(Gd96.86Pb3.14)
2nd component (green)
(Phase)
21.51 ± 0.24
(Gd95.73Pb4.27)
24.73 ± 0.28
(Gd82.48Pb17.52)
25.12 ± 0.17
(Gd70.38Pb29.62)
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Gębara, P.; Hasiak, M.; Kovac, J.; Rajnak, M. Entalpy of Mixing, Microstructure, Structural, Thermomagnetic and Mechanical Properties of Binary Gd-Pb Alloys. Materials 2022, 15, 7213. https://doi.org/10.3390/ma15207213

AMA Style

Gębara P, Hasiak M, Kovac J, Rajnak M. Entalpy of Mixing, Microstructure, Structural, Thermomagnetic and Mechanical Properties of Binary Gd-Pb Alloys. Materials. 2022; 15(20):7213. https://doi.org/10.3390/ma15207213

Chicago/Turabian Style

Gębara, Piotr, Mariusz Hasiak, Jozef Kovac, and Michal Rajnak. 2022. "Entalpy of Mixing, Microstructure, Structural, Thermomagnetic and Mechanical Properties of Binary Gd-Pb Alloys" Materials 15, no. 20: 7213. https://doi.org/10.3390/ma15207213

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop