Next Article in Journal
Influence of Microstructure and Shot Peening Treatment on Corrosion Resistance of AISI F55-UNS S32760 Super Duplex Stainless Steel
Next Article in Special Issue
Green Preparation of Straw Fiber Reinforced Hydrolyzed Soy Protein Isolate/Urea/Formaldehyde Composites for Biocomposite Flower Pots Application
Previous Article in Journal
High-Temperature Tolerance in Multi-Scale Cermet Solar-Selective Absorbing Coatings Prepared by Laser Cladding
Previous Article in Special Issue
New Magnetostrictive Transducer Designs for Emerging Application Areas of NDE
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Giant Enhancement of Magnetostrictive Response in Directionally-Solidified Fe83Ga17Erx Compounds

1
College of Engineering, Northeastern University, Boston, MA 02115, USA
2
Institute of Experimental Physics, Saarland University, 66123 Saarbrucken, Germany
3
Baotou Research Institute of Rare Earths, Baotou 014010 China
*
Authors to whom correspondence should be addressed.
Materials 2018, 11(6), 1039; https://doi.org/10.3390/ma11061039
Submission received: 19 May 2018 / Revised: 7 June 2018 / Accepted: 8 June 2018 / Published: 19 June 2018
(This article belongs to the Special Issue Magnetostrictive Composite Materials)

Abstract

:
We report, for the first time, correlations between crystal structure, microstructure and magnetofunctional response in directionally solidified [110]-textured Fe83Ga17Erx (0 < x < 1.2) alloys. The morphology of the doped samples consists of columnar grains, mainly composed of a matrix phase and precipitates of a secondary phase deposited along the grain boundary region. An enhancement of more than ~275% from ~45 to 170 ppm is observed in the saturation magnetostriction value (λs) of Fe83Ga17Erx alloys with the introduction of small amounts of Er. Moreover, it was noted that the low field derivative of magnetostriction with respect to an applied magnetic field ( i . e . ,   d λ s / d H a p p for Happ up to 1000 Oe) increases by ~230% with Er doping ( d λ s / d H a p p , FeGa = 0.045 ppm/Oe; d λ s / d H a p p , FeGaEr = 0.15 ppm/Oe). The enhanced magnetostrictive response of the Fe83Ga17Erx alloys is ascribed to an amalgamation of microstructural and electronic factors, namely: (i) improved grain orientation and local strain effects due to deposition of Er in the intergranular region; and (ii) strong local magnetocrystalline anisotropy, due to the highly anisotropic localized nature of the 4f electronic charge distribution of the Er atom. Overall, this work provides guidelines for further improving galfenol-based materials systems for diverse applications in the power and energy sector.

Graphical Abstract

1. Introduction

Functional materials systems that demonstrate large magnetostriction play an important role in a wide array of commercial applications, ranging from acoustic sensors and linear actuators to electromechanical energy harvesters and sonar transducers [1,2,3]. One of the most successful magnetostrictive materials hitherto is the rare-earth compound, (Dy0.7Tb0.3)Fe2 (also known as Terfenol D) [4]. These alloys demonstrate a cubic C15 laves crystal structure and exhibit large room temperature magnetic-field-induced strains up to 2000 ppm [4]. It is well known that Terfenol D has several major drawbacks that constrain its use in commercial devices, particularly: (i) high cost and global shortage of the rare-earth elements, Tb and Dy [5]; (ii) low mechanical integrity (high brittleness, low yield stress, low magneto-mechanical coupling) [6]; and (iii) high fields required for magnetic saturation [7]. To this end, a promising alternative to (Dy0.7Tb0.3)Fe2 is the rare-earth-free, inexpensive and corrosion resistant Fe1−xGax alloy (commercially known as Galfenol), which exhibits moderate magnetostriction (up to approximately 400 ppm) under a very low magnetic field of 100 Oe and a high tensile strength of 500 MPa in the temperature range from −20 to 80 °C [8].
An intriguing characteristic of Fe1−xGax is that its functional response is closely correlated with its microstructural and crystallographic properties [8]. The binary phase diagram of Fe1−xGax indicates that the single-phase terminal solid solution possesses a chemically disordered body-centered cubic (bcc) crystal structure that extends to Ga concentrations of 11 at.% at room temperature and to as much as 35 at.% Ga at 1050 °C [9]. In the composition range (~27–28 at.% Ga), the alloy also exists as Fe3Ga, and it exhibits a chemically-ordered cubic L12 crystal structure that undergoes polymorphic transformations to the ordered hexagonal D019 and cubic D03 phase upon heating [10,11]. Further, a B2 ordered cubic phase variant is also noted at high temperatures for compositions exceeding 32 at.% Ga [10]. Depending upon the sample synthesis and the processing technique employed, significant amounts of Ga (well in excess of the solubility limit) can be retained in a metastable disordered bcc solid solution at room temperature [8,10,12]. High magnetostriction values ranging from 250 to 400 ppm have been reported in single crystals of bcc alloys of Fe1−xGax,, where x ranges from 15 to 25 [13].
Over the last twenty years, attempts have been made to improve the magnetostrictive response of galfenol by addition of a wide variety of elements into its crystal lattice. In almost all cases, ternary additions of 3d and 4d transition metals (V, Cr, Mn, Co, Ni, Rh and Mo) decrease the magnetostriction values, relative to that of the parent binary FeGa alloy [14,15,16,17,18]. While tiny amounts of small interstitial atoms (C, B or N) have a slight but favorable effect on the magnetostriction of FeGa (particularly at high atomic compositions of Ga) [18,19]. More recent studies indicate that the magnetostriction of this compound can be significantly increased by adding small amounts of rare-earth elements ranging from La to Lu [20,21,22,23,24,25,26,27,28]. A phenomenological model based on the rare-earth crystal field interaction, proposed by He et al., suggests that the best trace dopants are the light rare earths, Ce and Pr (up to 0.2 at.% doping), which give a transverse magnetostriction of up to 800 ppm [24]. However, to date, very little attention has been given to Er-doped FeGa compounds.
To add to the FeGa literature, here, we present, for the first time, an experimental study that aims at investigating the crystallographic, microstructural, magnetic and magnetostrictive properties of a series of Er-doped polycrystalline alloys of composition, Fe83Ga17Erx (0 < x < 1.2). Results obtained in this research effort represent an exceptional increase (~275%) in the magnetostriction coefficient of [110]-textured FeGa alloys. The optimal composition for the best magnetostrictive response was found to be Fe83Ga17Erx (x = 0.6). The origin of the enhanced magnetostrictive effect in this materials system is discussed in the context of the microstructure, as well as the electronic structure of samples. Overall, this work provides pathways for enhancing the functional response of FeGa alloys.

2. Materials and Methods

Polycrystalline samples of nominal composition Fe83Ga17Erx (0 < x < 1.2) were synthesized from constituent elements of 99.9% purity, using vacuum arc-melting and directional-solidification techniques. The bulk ingots were subsequently placed in an Ar atmosphere and annealed at 900 °C for two hours to obtain the desired phase and microstructure. The arc-melted charges were then sliced into cuboid-shaped slabs (dimensions: 0.01 m × 0.01 m × 0.001 m) using a low-speed diamond saw for characterization of structural, magnetic and magnetostrictive attributes.
Microstructural analysis was carried out on mechanically polished sample slices using an Electron Backscatter Diffractometer, consisting of a JEOL 700F SEM microscope (JEOL, Welwyn Garden City, UK) equipped with a TSL OIM analysis unit (AMETEK, Leicester, UK) and a laboratory CuKα X-ray diffractometer (Rigaku Ultima III, Wilmington, MA, USA). Bragg peaks obtained from the X-ray diffraction pattern were least-squares fit to a Pseudo-Voigt function to estimate lattice parameters of the Fe83Ga17Erx alloys [29]. During electron back-scattering diffraction (EBSD) measurements, Kikuchi patterns were generated using an acceleration voltage of 20 kV, and recorded by means of a DigiView camera system (AMETEK, Leicester, UK) at a recording speed of the order of 0.1 s/pattern. Slightly longer time was required for multi-phase analysis. The working distance was 20 mm, and the step-size of the EBSD system was chosen to be 20 nm. More details regarding the measurement procedures may be found in References [30,31]. The results of the EBSD measurements are presented in the form of maps, the most important of which are the inverse pole figure (IPF) maps that indicate crystallographic orientation of individual foci.
Magnetic characterization was carried out using a vibrating sample magnetometer (Lake Shore, Model 7400, Westerville, OH, USA) in magnetic fields up to Happ = ±1.2 T and in the temperature range (300 K ≤ T ≤ 1000 K). The magnetic transition temperature (Tt) was determined as the inflection point in the derivative of magnetization (M) as a function of temperature (T) at an applied magnetic field of μ0H = 1 T. During magnetostriction measurements, the strain gauge was bonded longitudinally to the Galfenol samples in a quarter-bridge configuration to measure the magnetostrictive coefficient along the direction of growth of the samples. A Vishay Micro-measurement P3 strain Indicator was employed to measure the magnetostrictive strain, as the magnetic field was swept from 0 to 1 T at room temperature (~300 K).

3. Experimental Results & Discussion

3.1. Structural Attributes: Crystallographic and Microstructural Properties

X-ray diffraction data of the Fe83Ga17Erx alloys, obtained by scanning the sample plane perpendicular to the growth direction, as shown in Figure 1a, indicates the presence of a single phase having the bcc crystal structure for all samples of composition x < 0.6 (lattice parameter, a = 2.905 ± 0.005 Å). An additional Bragg peak corresponding to a minor secondary phase is observed at ~42°, as Er dopant concentration (x) is increased to x > 0.6. Overall, all the samples were found to be polycrystalline in character with a preferred orientation along the [110] growth direction—a feature attributed to the thermal gradient imposed by directional solidification. The dependence of relative X-ray intensity of the Bragg planes (110) and (200) with Er content is illustrated in Figure 1b. At all Er doping concentrations, the (110) plane retains dominance. Nonetheless, it is interesting to note that the (200) plane increases in intensity and displays the largest value for x = 0.6. It is hypothesized that Er favors occupation of the (200) planes at x < 0.6. These results are essential to further understanding the following measurements of magnetostriction as a function of Er content, as will be discussed in subsequent paragraphs.
The SEM images of select Fe83Ga17Erx alloys (0 < x < 1) are shown in Figure 2. The microstructure of the parent Fe83Ga17 alloy consists of a single solid solution with equiaxed grains of dimensions ~200 µm (Figure 2a). Conversely, the morphology of the Er-doped FeGa alloys demonstrates columnar grains mainly composed of a matrix phase (i.e., gray area) and precipitates of a secondary phase (i.e., white area) deposited along the grain boundary. The average size of the precipitates was approximately 1–3 μm, and it is observed that the fraction of precipitates increases with Er dopant concentration. Formation of the secondary phase in the Fe83Ga17Erx alloys is attributed to the realization that the atomic size of Er (175 pm) is significantly larger than that of Fe (140 pm) and Ga (135 pm). This difference in atomic dimensions leads to low solid solubility of Er in the FeGa matrix, and thus, Er exists mostly in the precipitates. This characteristic dual-phase microstructure has been observed previously in other rare-earth doped FeGa alloys [20,21,22,26,27], and it is considered to be advantageous for improving the magnetostrictive response and mechanical properties of this materials system.
To determine the distribution of elements, the sample Fe83Ga17Er0.6 was analyzed by EDXS. Figure 3 shows SEM morphology and the determination of the constituent elements at spot “a” (grain boundary) and spot “b” (grain). It was found that the grains are composed primarily of Fe and Ga, while the grain boundary consists of Fe and Ga, as well as a relatively high percentage of Er. Stoichiometric determination using EDXS indicates that the Er atoms accumulate more in the grain boundary region and form an intermetallic secondary phase, possibly Ga6Er. Presence of the secondary phase in samples with relatively high Er concentrations is consistent with results obtained by X-ray diffraction analysis (see Figure 1).
Information regarding the texture of the Fe83Ga17Erx alloys (0 < x < 1.2) was obtained using EBSD analysis. Kikuchi patterns obtained using this technique were indexed, using material files pertaining to FeGa (based on bcc α-Fe structure; Pearson symbol: cI2) and Ga6Er (based on tetragonal crystal structure; Pearson Symbol: tP14). Here, it should be noted that we could observe the Ga6Er phase from the sample with x = 0.2. This demonstrates the importance of a highly spatially resolved measurement. Figure 4 shows the IPF maps along the [001] direction (i.e., perpendicular to the sample surface) and corresponding pole figures for the following compositions: Fe83Ga17 (i.e., the parent compound), Fe83Ga17Er0.6 and Fe83Ga17Er1.2. The IPFs shown here provide the crystallographic orientation of the grains, according to the stereographic triangles for each phase. In confirmation with results obtained by X-ray diffraction and SEM-EDS, no Ga6Er phase was observed in the parent Fe83Ga17 compound (Figure 4a). In Fe83Ga17Er0.6, the Ga6Er precipitates were present, as indicated as small spots scattered over the scan area (Figure 4b). Conversely, significant aggregation of Ga6Er was also found in Fe83Ga17Er1.2 (Figure 4c). The IPF maps and the pole figures indicate that while the Fe83Ga17 parent compound demonstrates [001] texture, Fe83Ga17Er0.6 and Fe83Ga17Er1.2 samples do not exhibit preferred grain orientations. It is important to realize that grain size of both the FeGa and the Ga6Er phases in the Fe83Ga17Erx samples increased as a function of the Er concentration. As the dopant concentration was increased, Ga6Er formed a secondary phase in the grain boundary region (Figure 4b,c), and the orientation of the grains is found to be in [001] and [101] directions. When the Ga6Er is located within the matrix, the grains have the same orientation as the matrix—a feature attributed to the coherent relationship between the two phases. As the grain size of the FeGa and the Ga6Er phases increases, the crystallographic orientation between these phases became significantly different (see pole figures in Figure 4c), and thus, it is inferred that the secondary phase precipitates were randomly oriented in the matrix phase in the large grained Fe83Ga17Erx alloys.
Addition of Er has noteworthy effects on the magnetic properties of [110]-textured polycrystalline FeGa alloys. The magnetization behavior of Fe83Ga17Erx alloys (0 < x < 1.5) as a function of a magnetic field at room temperature (T = 300 K), shown in Figure 5a, indicates that doping with Er increases the saturation magnetization (Ms) by approximately 10%, as compared to the Fe83Ga17 parent alloy. As shown in the inset of Figure 5a, Ms initially increases from 153.8 to 168.2 emu/g, as the Er content increased from 0 to 0.6. Further increase in Er doping decreases Ms slightly—a feature attributed to the significant presence of the secondary Ga6Er phase. The temperature-dependent magnetization behavior of the Fe83Ga17Erx samples at an applied magnetic field of H = 1000 Oe is shown in Figure 5b. Consistent with previous reports, the Curie temperature of Fe83Ga17 was found to be approximately 990 K [8]. Overall, the Tc of the Fe83Ga17Erx samples was found to be independent of the Er concentration.

3.2. Enhanced Functional Response: Magnetostriction Measurements at Room Temperature

The magnetostriction strain (λ) along the direction of growth of the Fe83Ga17Erx alloys is plotted as a function of an applied magnetic field in Figure 6a. In all samples, λ increases with applied magnetic field until a saturation magnetostriction value (λs) is obtained. In agreement with previous studies on directionally solidified [110]-textured FeGa alloys [21], the λs of Fe83Ga17 was found to be ~45 ppm. Change in λs with Er content (x) for the Fe83Ga17Erx alloys is shown in Figure 6b. Overall, λ increased with Er doping till the maximum of 170 ppm was achieved at x = 0.6. These results represent a record enhancement of more than ~275% in λs of Fe83Ga17Erx alloys with the introduction of small amounts of Er. For operation in actuators and sensors in low loss magnetoelectric and multiferroic devices, such as those described in References [32,33,34,35,36], magnetostrictive materials must be operated under mechanical and magnetic bias conditions to achieve the maximum strain per unit magnetic field. Thus, permeability (µ) as realized from the derivative of magnetization/magnetostriction with respect to applied magnetic field ( d M / d H a p p or d λ / d H a p p   ) is a desired figure of merit in magnetostrictive materials. To this end, it is critical to observe from Figure 6a that the low field derivative of magnetostriction with respect to the applied magnetic field ( d λ / d H a p p for Happ up to 1000 Oe) increases by ~230% with Er doping ( d λ s / d H a p p , FeGa = 0.045 ppm/Oe; d λ s / d H a p p , FeGaEr = 0.15 ppm/Oe).
For FeGa single crystals, a maximum in magnetostriction is reported along the easy magnetic axis, i.e., along the <100> crystallographic direction. Assuming the approximation of only dipole–dipole interactions within the material, the magnetostriction values for a [110]-textured polycrystalline materials may be calculated using the expression [8]:
λ 110 = 1 4 λ 100 + 3 4 λ 111
where λ100 and λ111 are the saturation magnetostriction when the crystal is magnetized and the strain is measured along the <100> and <111> directions, respectively. In the absence of compressive pre-stress, the calculated values of (3/2)λ100 and (3/2)λ111 for Fe83Ga17 are ~311 ppm and −20 ppm, respectively [37]. Note, that the factor 3/2 arises from the definition of magnetostriction as a deformation from a demagnetized state [37]. The theoretical (3/2)λ110 value for a polycrystalline Fe83Ga17 sample was in approximate agreement with our experimentally determined magnetostriction value of λ110 = 45 ± 5 ppm. It is critical to note that the use of Equation (1) for the Fe83Ga17Erx alloys was based on the broad postulation that the texturing degree of the doped samples is similar to that of the pure Fe83Ga17 sample.
The (3/2)λ110 results of [110]-textured rare-earth doped FeGa alloys is summarized in Table 1 with relevant references. It is important to note that with an exception of one report concerning Fe83Ga17Ce0.8 [22], the Fe83Ga17Er0.6 sample demonstrates higher magnetostriction than any other [110]-textured directionally-solidified FeGa alloy synthesized to date.
At present, the enhanced magnetostrictive response of the doped Fe83Ga17Erx alloys is attributed to a combination of electronic and microstructural effects. Previous studies on binary FeGa alloys, conducted by Clark et al., suggest that the large magnetostriction in this intermetallic alloy originates from local magnetocrystalline anisotropy, induced by local short-range interactions between the Ga atoms along specific crystallographic directions in the disordered bcc α-Fe structure [37]. It is likely that a large number of 4f valence electrons and aspherical charge cloud distributions observed in Er atoms lead to enhanced magnetic anisotropy due to crystalline electric field effects. Considering that Er possesses a larger atomic radius (178 pm) relative to Fe (127 pm) and Ga (140 pm), it is possible that the strain due to local lattice distortions influence the magnetic properties of the Fe83Ga17Erx. In related compounds, namely Tb- and Ce-doped directionally solidified FeGa systems, an increase in magnetostriction has been linked to improved grain orientation and morphology [23,24]. Based on experimental and computational studies conducted by He et al., it is surmised that the giant magnetostriction in rare-earth doped FeGa alloys may be ascribed to the presence of nano-heterogeneties in the samples [28]. The dopants tend to enter the nano-heterogeneities, creating a larger tetragonal distortion of the matrix, as well as increased magnetocrystalline anisotropy. A mesoscopic model developed using phase field simulations shows that the bulk tetragonal distortion arises mainly from those nano-heterogeneities with fixed Ga–Ga pairs parallel to the applied magnetic field [28].
It is worth discussing the experimental results obtained in this study in the context of the phenomenological model reported by He et al. to predict magnetostrictive trends in FeGa alloys doped with rare-earth elements [24]. According to this model, the elements (i.e., Ce, Pr and Tb) have the greatest impact on magnetostriction of FeGa among all the rare-earth elements, due to the negative rare-earth quadrupole moment and local lattice tetragonal distortion of the matrix, which is triggered by the heterogeneous nanostructure with local tetragonal distortion [24]. If this were indeed true, the magnetostrictive behavior of the directionally solidified [110]-textured Fe83Ga17Tbx compounds investigated by Fitchorov et al. [34] and Jiang et al. [20] would be greater than that of the Fe83Ga17Erx samples examined in the current study. However, our experimental results suggested otherwise. Insight into a potential explanation for this contradiction may be obtained by comparing structure-magnetic property correlations between Fe83Ga17Erx and Fe83Ga17Tbx alloys. In both materials systems, only trace amounts of rare-earth dopants were required for optimal magnetostrictive performance in the Fe83Ga17Erx and Fe83Ga17Tbx systems. Excessive doping destroys magnetostriction, due to limited solid solubility of the rare-earth element in the FeGa lattice and subsequent formation of the intergranular secondary phases. The maximum magnetostriction in the [110]-textured directionally solidified Fe83Ga17Erx and related Fe83Ga17Tbx system was observed at x = 0.6 and 0.2 respectively, and thus, it is speculated that the solid solubility of Er in FeGa may be slightly more than that of Tb. It is further hypothesized that enhancement of magnetostriction may be achieved either by the application of a compressive pre-stress or by increasing the solid solubility of Er in the bcc FeGa matrix through quenching during the cooling phase of the sample fabrication technique. Indeed, previous studies in the FeGa literature demonstrate that the saturation magnetostriction of Fe83Ga17 alloys can be remarkably increased by the melt spinning method [22,38]. It is, however, critical to note that it is difficult to experimentally measure the magnetostriction of melt-spun ribbons, as the grains in the sample usually grow perpendicular to the direction of sample growth [22]. Moreover, due to large demagnetizing effects, shape anisotropy in FeGa ribbon samples typically leads to high magnetic saturation fields [38]. From the perspective of user inspired research, the directional solidification technique is more amenable for commercial production of rare-earth doped FeGa alloys.

4. Conclusions

In summary, in this work we present the effects of Er additives upon the microstructure, magnetic and microstructural properties of Fe83Ga17Erx alloys prepared by vacuum arc-melting and directional solidification methods. Data obtained in this experimental study indicate a room temperature magnetostriction value of 170 ppm in a [110]-textured polycrystalline sample of nominal composition Fe83Ga17Er0.6—a value that is ~275% higher than that of the corresponding parent Fe83Ga17 compound. Overall, addition of small amounts of Er into the FeGa lattice results in an increase in saturation magnetization and magnetostriction and a reduction in the saturation field. These characteristics are beneficial for practical applications, such as actuators in multiferroic magnetic field generators that use a converse magnetoelectric effect or high-sensitivity magnetic field sensors that operate based on the direct magnetoelectric effect without the need for a bias DC field. The enhanced magnetostrictive response of the Fe83Ga17Erx alloys is ascribed to an amalgamation of electronic and microstructural factors, namely: (i) strong local magnetocrystalline anisotropy due to the large spin-orbit coupling and the highly anisotropic localized nature of the 4f electronic charge distribution of the Er atom, (ii) improved grain orientation and morphology due to deposition of Er in the intergranular region and (iii) local strain effect that may arise due to incorporation of Er into the FeGa lattice. Excessive Er doping destroys the improved magnetostriction in Fe83Ga17Erx alloys, due to formation of an undesirable secondary phase, which is identified as the intermetallic compound, Ga6Er. Overall, these results highlight the potential for modifying the functional response of FeGa alloys by addition of tiny amounts of the rare-earth element, Er. To further understand the origin of the superior functional response of rare-earth doped FeGa systems, future work involving computational modeling of the magnetostrictive behavior of these compounds is desired.

Author Contributions

R.B. analyzed experimental data, prepared the manuscript and compiled the graphical images, P.T. performed experiments and prepared the initial draft of the manuscript, A.K.-V. and M.R.K. performed EBSD experiments, Y.C. helped design the study, L.J. synthesized the samples used in the study and finally V.G.H. designed the study, helped analyze the data and approved the final manuscript.

Funding

This research was funded by the U.S. Army under grant W911NF-10-2-0098, subaward 15-215456-03-00 is gratefully acknowledged.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Parsons, M.J.; Datta, S.; Mudivarthi, C.; Na, S.M.; Flatau, A. Torque sensing using rolled galfenol patches. In Proceedings of the SPIE Smart Sensor Phenomena, Technology, Networks, and Systems, San Diego, CA, USA, 7 April 2008; Volume 6933, p. 693314. [Google Scholar]
  2. Weng, L.; Wang, B.; Dapino, M.J.; Sun, Y.; Wang, L.; Cui, B. Relationships between magnetization and dynamic stress for Galfenol rod alloy and its application in force sensor. J. Appl. Phys. 2013, 113, 17A917. [Google Scholar] [CrossRef]
  3. Berbyuk, V. Vibration energy harvesting using Galfenol-based transducer. In Proceedings of the SPIE: Active and Passive Smart Structures and Integrated Systems, San Diego, CA, USA, 10 April 2013; Volume 8688, p. 86881F. [Google Scholar]
  4. Clark, E.; Teter, J.P.; McMasters, O.D. Magnetostriction “jumps” in twinned Tb0.3Dy0.7Fe1.9. J. Appl. Phys. 1988, 63, 3910. [Google Scholar] [CrossRef]
  5. Humphries, M. Rare Earth Elements: The Global Supply Chain; Diane Publishing Company: Collingdale, PA, USA, 2010. [Google Scholar]
  6. Olabi, A.G.; Grunwald, A. Design and application of magnetostrictive materials. Mater. Des. 2008, 29, 469–483. [Google Scholar] [CrossRef] [Green Version]
  7. Kellogg, R.A. Development and Modelling of Iron-Gallium Alloys. Ph.D. Thesis, Iowa State University, Ames, IA, USA, 2003. [Google Scholar]
  8. Atulasimha, J.; Flatau, A.B. A review of magnetostrictive iron-gallium alloys. Smart Mater. Struct. 2011, 20, 043001. [Google Scholar] [CrossRef]
  9. Xing, Q.; Du, Y.; McQueeney, R.J.; Lograsso, T.A. Structural investigations of Fe-Ga alloys: Phase relations and magnetostrictive behavior. Acta Mater. 2008, 56, 4536–4546. [Google Scholar] [CrossRef]
  10. Lograsso, T.A.; Summers, E.M. Detection and quantification of D03 chemical order in Fe-Ga alloys using high resolution X-ray diffraction. Mater. Sci. Eng. A 2006, 416, 240–245. [Google Scholar] [CrossRef]
  11. Srisukhumbowornchai, N.; Guruswamy, S. Influence of ordering on the magnetostriction of Fe–27.5 at.% Ga alloys. J. Appl. Phys. 2002, 92, 5371. [Google Scholar] [CrossRef]
  12. Taheri, P.; Barua, R.; Hsu, J.; Zamanpour, M.; Chen, Y.; Harris, V.G. Structure, magnetism, and magnetostrictive properties of mechanically alloyed Fe81Ga19. J. Alloys Compd. 2016, 661, 306–311. [Google Scholar] [CrossRef]
  13. Atulasimha, J.; Flatau, A.B.; Cullen, J.R. Analysis of the effect of gallium content on the magnetomechanical behavior of single-crystal FeGa alloys using an energy-based model. Smart Mater. Struct. 2008, 17, 025027. [Google Scholar] [CrossRef]
  14. Hattrick-Simpers, J.R.; Hunter, D.; Craciunescu, C.M.; Jang, K.S.; Murakami, M.; Cullen, J.; Wuttig, M.; Takeuchi, I.; Lofland, S.E.; Bendersky, L.; et al. Combinatorial Investigation of Magnetostriction in FeGa and FeGa-Al. Appl. Phys. Lett. 2008, 93, 8–13. [Google Scholar] [CrossRef]
  15. Mungsantisuk, P.; Corson, R.P.; Guruswamy, S. Influence of Be and Al on the magnetostrictive behavior of FeGa alloys. J. Appl. Phys. 2005, 98, 1–7. [Google Scholar] [CrossRef]
  16. Dai, L.; Cullen, J.; Wuttig, M.; Lograsso, T.; Quandt, E. Magnetism, elasticity, and magnetostriction of FeCoGa alloys. J. Appl. Phys. 2003, 93, 8627–8629. [Google Scholar] [CrossRef]
  17. Restorff, J.B.; Clark, A.E.; Lograsso, T.A.; Ross, A.R. Magnetostriction of ternary Fe-Ga-X alloys (X = Ni, Mo, Sn, Al). J. Appl. Phys. 2002, 91, 8225. [Google Scholar] [CrossRef]
  18. Clark, A.E.; Restorff, J.B.; Wunfogle, M.; Hathaway, K.B.; Lograsso, T.A. Magnetostriction of ternary Fe-Ga-X (X = C, V, Cr, Mn, Co, Rh) alloys. J. Appl. Phys. 2016, 507, 99–102. [Google Scholar] [CrossRef]
  19. Huang, M.; Lograsso, T.A.; Clark, A.E.; Restorff, J.B.; Wun-Fogle, M. Effect of interstitial additions on magnetostriction in Fe-Ga alloys. J. Appl. Phys. 2008, 103, 1–4. [Google Scholar] [CrossRef]
  20. Jiang, L.; Yang, J.; Hao, H.; Zhang, G.; Wu, S.; Chen, Y.; Obi, O.; Fitchorov, T.; Harris, V.G. Giant enhancement in the magnetostrictive effect of FeGa alloys doped with low levels of terbium. Appl. Phys. Lett. 2013, 102, 222409. [Google Scholar] [CrossRef]
  21. Jin, T.; Wu, W.; Jiang, C. Improved magnetostriction of Dy-doped Fe83Ga17 melt-spun ribbons. Scr. Mater. 2014, 74, 100–103. [Google Scholar] [CrossRef]
  22. Yao, Z.; Tian, X.; Jiang, L.; Hao, H.; Zhang, G.; Wu, S. Influences of rare earth element Ce-doping and melt-spinning on microstructure and magnetostriction of Fe83Ga17 alloy. J. Alloys Compd. 2015, 637, 431–435. [Google Scholar] [CrossRef]
  23. Jiheng, L.I.; Ximing, X.; Chao, Y.; Xuexu, G.A.O.; Xiaoqian, B.A.O. Effect of yttrium on the mechanical and magnetostrictive properties of Fe83Ga17 alloy. J. Rare Earths 2015, 33, 1087–1092. [Google Scholar]
  24. He, Y.; Jiang, C.; Wu, W.; Wang, B.; Duan, H.; Wang, H.; Zhang, T.; Wang, J.; Liu, J.; Zhang, Z.; et al. Giant heterogeneous magnetostriction in FeGa alloys: Effect of trace element doping. Acta. Mater. 2016, 109, 177–186. [Google Scholar] [CrossRef]
  25. Meng, C.; Jiang, C. Magnetostriction of a Fe83Ga17 single crystal slightly doped with Tb. Scr. Mater. 2016, 114, 9–12. [Google Scholar] [CrossRef]
  26. Golovin, I.S.; Balagurov, A.M.; Palacheva, V.V.; Emdadi, A.; Bobrikov, I.A.; Churyumov, A.Y.; Cheverikin, V.V.; Pozdniakov, A.V.; Mikhaylovskaya, A.V.; Golovin, S.A. Influence of Tb on structure and properties of Fe-19% Ga and Fe-27% Ga alloys. J. Alloys Compd. 2017, 707, 51–56. [Google Scholar] [CrossRef]
  27. Wei, W.; Jiang, C. Improved magnetostriction of Fe83Ga17 ribbons doped with Sm. Rare Met. 2017, 36, 18–22. [Google Scholar]
  28. He, Y.; Ke, X.; Jiang, C.; Miao, N.; Wang, H.; Coey, J.M.D.; Wang, Y.; Xu, H. Interaction of Trace Rare-Earth Dopants and Nanoheterogeneities Induces Giant Magnetostriction in Fe-Ga Alloys. Adv. Funct. Mater. 2018, 28, 1800858. [Google Scholar] [CrossRef]
  29. Novak, G.A.; Colville, A.A. A practical interactive least-squares cell-parameter program using an electronic spreadsheet and a personal computer. Am. Miner. 1989, 74, 488–490. [Google Scholar]
  30. Koblischka-Veneva, A.; Koblischka, M.R.; Simon, P. Electron backscatter diffraction study of polycrystalline YBa2Cu3O7−δ. Phys. C: Supercond. 2002, 382, 311–322. [Google Scholar] [CrossRef]
  31. Koblischka-Veneva, A.; Muchlich, F.; Koblischka, M.R.; Babu, N.H.; Cardwell, D.A. Crystallographic Orientation of Y2Ba4CuMO. J. Am. Ceram. Soc. 2007, 90, 2582–2588. [Google Scholar] [CrossRef]
  32. Chen, Y.; Gillette, S.M.; Fitchorov, T.; Jiang, L.; Hao, H.; Li, J.; Gao, X.; Geiler, A.; Vittoria, C.; Harris, V.G. Quasi-one-dimensional miniature multiferroic magnetic field sensor with high sensitivity at zero bias field. J. Appl. Phys. 2011, 99, 042505. [Google Scholar] [CrossRef]
  33. Geiler, A.L.; Gillette, S.M.; Chen, Y.; Wang, J.; Chen, Z.; Yoon, S.D.; He, P.; Gao, J.; Vittoria, C.; Harris, V.G. Multiferroic heterostructure fringe field tuning of meander line microstrip ferrite phase shifter. Appl. Phys. Lett. 2010, 96, 053508. [Google Scholar] [CrossRef]
  34. Fitchorov, T.; Chen, Y.; Hu, B.; Gillette, S.M.; Geiler, A.; Vittoria, C.; Harris, V.G. Tunable fringe magnetic fields induced by converse magnetoelectric coupling in a FeGa/PMN-PT multiferroic heterostructure. J. Appl. Phys. 2011, 110, 123916. [Google Scholar] [CrossRef] [Green Version]
  35. Stephan, M.; Jahns, R.; Greve, H.; Quandt, E.; Knöchel, R.; Wagner, B. MEMS magnetic field sensor based on magnetoelectric composites. J. Micromech. Microeng. 2012, 22, 6. [Google Scholar]
  36. Dong, S.; Zhai, J.; Li, J.; Viehland, D.; Summers, E. Strong magnetoelectric charge coupling in stress-biased multilayer-piezoelectric/magnetostrictive composites. J. Appl. Phys. 2007, 101, 124102. [Google Scholar] [CrossRef]
  37. Clark, A.E.; Hathaway, K.B.; Wun-Fogle, M.; Restorff, J.B.; Lograsso, T.A.; Keppens, V.M.; Petculescu, G.; Taylor, R.A. Extraordinary magnetoelasticity and lattice softening in bcc Fe-Ga alloys. J. Appl. Phys. 2003, 93, 8621. [Google Scholar] [CrossRef] [Green Version]
  38. Wu, W.; Liu, J.; Jiang, C.; Xu, H. Giant magnetostriction in Tb-doped Fe83Ga17 melt-spun ribbons. Appl. Phys. Lett. 2013, 103, 262403. [Google Scholar] [CrossRef]
Figure 1. (a) X-ray diffraction pattern of directional solidified bulk alloys of Fe83Ga17Erx (0 < x < 1). An additional Bragg peak corresponding to a minor secondary phase is observed in samples where the Er dopant concentration (x) is higher than 0.6; (b) dependence of the relative intensity of the Bragg peaks corresponding to (110) and (200) planes on Er content (x).
Figure 1. (a) X-ray diffraction pattern of directional solidified bulk alloys of Fe83Ga17Erx (0 < x < 1). An additional Bragg peak corresponding to a minor secondary phase is observed in samples where the Er dopant concentration (x) is higher than 0.6; (b) dependence of the relative intensity of the Bragg peaks corresponding to (110) and (200) planes on Er content (x).
Materials 11 01039 g001
Figure 2. A microstructure of Er-doped (Fe0.83Ga0.17)100−xErx (x = 0, 0.2, 0.4, 0.6).
Figure 2. A microstructure of Er-doped (Fe0.83Ga0.17)100−xErx (x = 0, 0.2, 0.4, 0.6).
Materials 11 01039 g002
Figure 3. (a,b) SEM morphology of a sample of composition, Fe83Ga17Er0.6.; (c,d) EDXS profile of Fe, Ga and Er in the grain boundary region marked as “+” and in the intragranular region marked as “□”.
Figure 3. (a,b) SEM morphology of a sample of composition, Fe83Ga17Er0.6.; (c,d) EDXS profile of Fe, Ga and Er in the grain boundary region marked as “+” and in the intragranular region marked as “□”.
Materials 11 01039 g003
Figure 4. Inverse pole figure maps and corresponding pole figures for FeGa and the Ga6Er phases in samples of the following compositions: (a) Fe83Ga17 (parent compound); (b) Fe83Ga17Er0.6 and (c) Fe83Ga17Er1.2. The inverse pole figures shown here give the crystallographic orientation of the grains, according to the stereographic triangles for each phase.
Figure 4. Inverse pole figure maps and corresponding pole figures for FeGa and the Ga6Er phases in samples of the following compositions: (a) Fe83Ga17 (parent compound); (b) Fe83Ga17Er0.6 and (c) Fe83Ga17Er1.2. The inverse pole figures shown here give the crystallographic orientation of the grains, according to the stereographic triangles for each phase.
Materials 11 01039 g004
Figure 5. (a) Magnetization hysteresis loops of Fe83Ga17Erx (x = 0, 0.2, 0.6, 1). The inset shows the saturation moment as a function of the Er doping amount; (b) Magnetization (emu/g) of (Fe0.83Ga0.17)100−xErx at H = 10 k Oe as a function of temperature for different amounts of Er doping.
Figure 5. (a) Magnetization hysteresis loops of Fe83Ga17Erx (x = 0, 0.2, 0.6, 1). The inset shows the saturation moment as a function of the Er doping amount; (b) Magnetization (emu/g) of (Fe0.83Ga0.17)100−xErx at H = 10 k Oe as a function of temperature for different amounts of Er doping.
Materials 11 01039 g005
Figure 6. (a) Magnetostriction of the (Fe0.83Ga0.17)100−xErx (0 < x < 1) alloys as a function of the applied magnetic field. An enhanced saturation magnetostriction value (λs) of more than 250% is measured relative to the parent compound; and (b) change in λs as a function of the Er dopant concentration (x).
Figure 6. (a) Magnetostriction of the (Fe0.83Ga0.17)100−xErx (0 < x < 1) alloys as a function of the applied magnetic field. An enhanced saturation magnetostriction value (λs) of more than 250% is measured relative to the parent compound; and (b) change in λs as a function of the Er dopant concentration (x).
Materials 11 01039 g006
Table 1. Magnetostriction coefficients of [110]-textured rare-earth doped FeGa alloys synthesized via the directional solidification technique.
Table 1. Magnetostriction coefficients of [110]-textured rare-earth doped FeGa alloys synthesized via the directional solidification technique.
AlloyFabrication Techniqueλ110 (ppm)ConditionReferences
Fe83Ga17Directional solidification45Bulk; Pre-stressedCurrent work
Fe83Ga17Directional solidification68Bulk; Pre-stressedL. Jiang et al. [20]
Fe81Ga19Tb0.3Directional solidification85Bulk; Pre-stressedT.I. Fitchorov et al. [34]
Fe83Ga17Y0.64Directional solidification133Bulk; Compressed under 15 MPaL. Jiheng et al. [23]
Fe83Ga17Tb0.2Directional solidification160Bulk; Pre-stressedL. Jiang et al. [20]
Fe83Ga17Er0.6Directional solidification170Bulk; Pre-stressedCurrent work
Fe83Ga17Ce0.8Directional solidification200Bulk; Pre-stressed; Sample not at saturationZ. Yao et al. [22]

Share and Cite

MDPI and ACS Style

Barua, R.; Taheri, P.; Chen, Y.; Koblischka-Veneva, A.; Koblischka, M.R.; Jiang, L.; Harris, V.G. Giant Enhancement of Magnetostrictive Response in Directionally-Solidified Fe83Ga17Erx Compounds. Materials 2018, 11, 1039. https://doi.org/10.3390/ma11061039

AMA Style

Barua R, Taheri P, Chen Y, Koblischka-Veneva A, Koblischka MR, Jiang L, Harris VG. Giant Enhancement of Magnetostrictive Response in Directionally-Solidified Fe83Ga17Erx Compounds. Materials. 2018; 11(6):1039. https://doi.org/10.3390/ma11061039

Chicago/Turabian Style

Barua, Radhika, Parisa Taheri, Yajie Chen, Anjela Koblischka-Veneva, Michael R. Koblischka, Liping Jiang, and Vincent G. Harris. 2018. "Giant Enhancement of Magnetostrictive Response in Directionally-Solidified Fe83Ga17Erx Compounds" Materials 11, no. 6: 1039. https://doi.org/10.3390/ma11061039

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