Next Article in Journal
PIN-PMN-PT Single Crystal 1-3 Composite-based 20 MHz Ultrasound Phased Array
Previous Article in Journal
Mechanically Enabled Two-Axis Ultrasonic-Assisted System for Ultra-Precision Machining
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Magnetostrictive Performance of Electrodeposited TbxDy(1−x)Fey Thin Film with Microcantilever Structures

1
Department of Mechanical Systems Engineering, Tohoku University, Sendai 980-8579, Japan
2
Micro-Nanomachining Research Education Center, Tohoku University, Sendai 980-8579, Japan
3
Micro System Integration Center (µSiC), Tohoku University, Sendai 980-8579, Japan
*
Authors to whom correspondence should be addressed.
Micromachines 2020, 11(5), 523; https://doi.org/10.3390/mi11050523
Submission received: 30 March 2020 / Revised: 16 May 2020 / Accepted: 17 May 2020 / Published: 21 May 2020
(This article belongs to the Section D:Materials and Processing)

Abstract

:
The microfabrication with a magnetostrictive TbxDy(1−x)Fey thin film for magnetic microactuators is developed, and the magnetic and magnetostrictive actuation performances of the deposited thin film are evaluated. The magnetostrictive thin film of TbxDy(1−x)Fey is deposited on a metal seed layer by electrodeposition using a potentiostat in an aqueous solution. Bi-material cantilever structures with the Tb0.36Dy0.64Fe1.9 thin-film are fabricated using microfabrication, and the magnetic actuation performances are evaluated under the application of a magnetic field. The actuators show large magnetostriction coefficients of approximately 1250 ppm at a magnetic field of 11000 Oe.

1. Introduction

Magnetostriction is a useful property of ferromagnetic materials that causes strain during the process of magnetization. The strain of the magnetostriction materials can be controlled by a magnetic field [1,2,3,4,5,6]. Also, the strain itself can generate a magnetic field, referred to as inverse-magnetostrictive effect or Villari effect [1,2,3]. Magnetostriction can be quantified by the magnetostrictive coefficient which can be positive or negative and is defined as the generated strain when a magnetic field causing magnetization saturation is applied. For example, Fe, Ni, and Co are well known as magnetic materials, which show small magnetostrictive coefficients of −14, −50, and −93, respectively [1,2,3]. It is known that Fe alloys containing a rare earth material exhibit a large magnetostrictive coefficient at room temperature, and those alloys are referred to as “giant magnetostriction material” [1,2,3,4]. Among those, Terfenol-D (Tb0.3Dy0.7Fe2) exhibits a large magnetostrictive coefficient up to 1400 ppm at a magnetic field of ~2 kOe [1,2,3,4,7]. In addition, Galfenol, Fe0.8Ga0.2, is known as a material, which shows a large magnetostriction up to 400 ppm [5,6] and CoFe shows 260 ppm as well [8,9,10,11,12]. Since these discoveries, those materials have emerged as a smart material for microdevices, including actuators [13,14], wireless sensors, biosensors [15,16], energy harvesting devices [17], and atomic force microscopy [18].
Most of the devices based on the giant magnetostriction materials are made from bulk materials. However, the importance of the thin-film technology for those materials have gained for realizing miniaturized smart actuators and devices. Many methods have been reported for thin-film preparation, including pulsed laser deposition [7], sputtering [10,11,19,20,21,22,23,24,25,26,27,28], and electrochemical deposition [5,6,9,29]. For the film deposition of Terfenol-D, the sputtering method has been reported because of the simple approach, high film uniformity, and low roughness. However, the substrate must be heated at a temperature higher than 400 °C for crystallization, because the sputter-deposited Terfenol-D films are amorphous state at low temperatures, which show a low magnetostrictive performance [23,24,25]. Those sputtered films exhibit a magnetostriction coefficient approximately 1/3 (540 ppm) of the bulk value without annealing and 2/3 (920 ppm) of the bulk value with annealing at 450 °C [19,20,21]. Electrochemical deposition has advantages for simplicity, low cost, and compatibility with batch fabrication, etc., and some of the researches have been reported regarding magnetostrictive thin films including Galfenol [5,6], CoFe [9], and TbFe2 [29].
Rare earth atoms including Tb and Dy are generally difficult to deposit by electrodeposition using an aqueous solution because those materials have reduction potential <−2 V (eg., Tb+3 + 3e = Tb: −2.28 V, Dy3+ + 3e = Dy: −2.6 V); therefore, hydrogen evolution makes the aqueous solution unstable [26]. Chemical additives add to reducing electrochemical deposition potential and improve film quality [29,30,31,32,33]. Gong et al. demonstrated the deposition of TbFe2 film in an aqueous solution with a rare earth metal complex [29].
In this work, Terfenol-D films are deposited by electroplating and the performances of the deposited films are evaluated using energy-dispersive X-ray spectroscopy (EDX) and vibrating-sample magnetometer (VSM) analysis. Microcantilever bi-material structures are fabricated, and the magnetostriction performances are evaluated.

2. Experimental

A TbxDy(1−x)Fey thin film was deposited by electrodeposition using a potentiostat with three conventional electrodes: working, counter, and reference electrodes. This aqueous electrolyte was prepared from rare earth sulfate salts and iron salts. In order to reduce the deposition potential, those ions are chelated by citric acid and tartaric acid. Thus, the electrolyte is formed by mixing deionized 150 mL water with following chemicals, i.e., Tb2(SO4)3 0.3 g, Dy2(SO4)3 0.7 g, FeCl3 0.5 g, FeSO4 1.8 g, tartaric acid 3 g, citric acid 0.5 g, KCl 33 g, NaOH 1.1 g [34]. Electrodeposition was proceeded on a 300 nm-thick Cr-Cu seed layer deposited on a silicon on insulator (SOI) wafer with a 100 nm-thick Si/300 nm-thick buried oxide layer/ 750 µm-thick Si handling layer. The electrodeposition was performed at 40 °C for 2.5 h at a working electrode potential of (−925)–(−950 mV). The typical deposition rate of the TbxDy(1−x)Fey film is approximately 100 nm/h. However, the deposition speed was varied by deposition area, the resistance of the seed layer, the distance between working and counter electrode, and temperature of the electrolyte. The compositions of the deposited films are analyzed by energy-dispersive X-ray spectroscopy (EDX). The magnetization properties for the in-plane direction of the film are measured using a vibrating sample magnetometer (VSM). For the analysis of surface morphology, atomic force microscopy (AFM) and magnetic force microscopy (MFM) are used. The magnetostriction coefficients are measured from the bi-material cantilever actuation using an optical microscope with a special resolution of 4.4 µm corresponding to one imaging pixel. The sample preparation for the TbxDy(1−x)Fey cantilevers is performed by microfabrication. The details of the fabrication process are described later.
Figure 1 shows the typical EDX result of a 300 nm-thick TbxDy(1−x)Fey thin film prepared at −930 mV of potential, which formed on the SOI wafer with the Cr-Cu seed layer. For the analysis, the substrate and seed layer components are ignored. Generally, the composition of electrodeposited alloys can be adjusted by the applied potential because of the reduction potential difference of each component. However, applicable potential has a limited window; for the potential V > −700 mV, Fe atoms do not deposit. For the case of the potential V < −1055 mV, hydrogen evolution happens, and the electrolyte is degraded. To maximize the magnetostriction performance, the atomic concentration ratio of rare earth and iron atoms must satisfy 1:2 [1,2,3,4]. Thus, the ideal weight percentage of Fe atoms is approximately 40%, and the atomic percentage is approximately 66%. The Fe concentration can be controlled by the working electrode potential, as shown in Figure 2, which shows approximately 1% to 2% of the atomic percentage change of Fe atoms with 1 mV of potential variation. The optimal electrochemical potential can be found at −930 mV for the film deposition on the Cu seed layer. The composition fractions of Tb, Dy, and Fe were analyzed to be approximately 12.6, 22 and 65.4 atomic %, respectively; thus, the film composition is approximated to be Tb0.36Dy0.64Fe1.9.
The in-plane magnetization analysis of the 200 nm-thick Tb0.36Dy0.64Fe1.9 sample with the SOI wafer was proceeded using VSM, as a result is shown in Figure 3. It is found that the coercive magnetic field is approximately 285 Oe. The magnetization is saturated at approximately 5000 Oe, and magnetization starts to decrease. Thus, the effective magnetic field for magnetostrictive actuation can be regarded to be in the range of 285–5000 Oe. It is reported that the bulk Terfenol-D shows 63 Oe of coercive magnetic field [27], 2000 Oe of saturation magnetic field and 1 T of saturation magnetization [1]. Compared with bulk value, the saturation magnetization of the electrodeposited film is 40% lower than that of the bulk value. There are several suspected reasons for this degradation. One is that the film composition ratio is slightly different from the ideal value of Terfenol-D. The different composition ratio shows different magnetization characteristics [1]. Another reason is the magneto crystalline anisotropy effect [35]. The magneto crystalline anisotropy plays important roles in magnetic domain rotation, magnetization and magnetostriction. An electrodeposited magnetostriction GaFe film shows uncontrolled lattice orientation as reported [6]. These random lattice structures affect low magnetization and magnetostriction characteristics [6,9]. The possible other reason for this is the oxide impurity of the film, which pins the magnetic domains [21], decreases saturation magnetizations and increases the coercive magnetic field [21,22,26].
Figure 4 and Figure 5 show the atomic force microscopy and magnetic force microscopy images of the deposited film with a thickness of 300 nm. In the atomic force microscopy image, the Tb0.34Dy0.65Fe1.9 film has a fine grain microstructure with a diameter of ~170 nm, which is larger than the grain size of reported sputtered films ~50–55 nm [19,20]. Generally, the grain size of electrodeposited films is much bigger than that of the as-deposited sputtered film. From the magnetic force microscopy images, the Tb0.34Dy0.65Fe1.9 film shows a large magnetic domain and grain boundary. This is one of the evidences that electrodeposited Tb0.34Dy0.65Fe1.9 film possesses a polycrystalline structure. Sputtered Terfenol-D films without annealing shows amorphous state crystallinity. The amorphous Terfenol-D film show maze shape magnetic domain image from magnetic force microscopy. Polycrystalline state Terfenol-D, however, shows magnetic domain structure sillier to crystal grain structure [21].
The magnetostriction coefficient is generally defined by generated strain under the application of a magnetic field. However, in the case of thin films, it is difficult to measure the strain of the films directly. Thus, an analytical model using the displacements of bi-material cantilevers is employed to evaluate the magnetostriction coefficients [9,36,37], in which the effective magnetostriction coefficient value λ e f f can be calculated from the displacements of the bi-material cantilever structures with an application of magnetic fields in parallel and perpendicular against the longitudinal direction of the cantilever, as given by [37],
λ e f f = 2 ( D D ) E s t s 2 ( 1 + v f ) 9 l 2 E f t f ( 1 + v s )
where D is the displacement in parallel to the magnetic field, D is the displacement in perpendicular to the magnetic field, E s and v s are Young’s modulus and Poisson ratio of the substrate material, respectively, E f and v f are the Young’s modulus and Poisson ratio of the magnetostrictive material, t f and t s are the film thicknesses of the magnetostrictive layer and the substrate, respectively, and l is the length of the cantilever. The spring constant kcantilever of the cantilever structure, and the force F generated by the magnetostrictive film can be approximated as given by following equations [38],
k c a n t i l e v e r = 4 + 6 t f t s + 4 ( t f t s ) 2 + E f E s ( t f t s ) 3 + E f E s t s t f ,
F = k c a n t i l e v e r D
where D is the displacement of the cantilever. The elastic properties of thin-films and microstructures are generally almost same with that of the bulk [39]; thus, for this calculation, the Young’s modulus of each layer is supposed to be the bulk value of silicon and Terfenol-D, i.e., 179 × 10 9   P a and 50 × 10 9   P a , respectively, and the Poisson ratios of the silicon substrate and the Tb0.34Dy0.65Fe1.9 film are supposed to be 0.22 and 0.3, respectively [28,35].
Table 1 shows the typical dimensions of the fabricated cantilever. From Table 1 and Equation (2), the effective spring constant of the composite cantilever is calculated to be 26.4 N/m.
The fabrication process of the bi-material cantilever is shown in Figure 6. The Tb0.34Dy0.65Fe1.9 (Terfenol-D) film is deposited on the SOI wafer with a Cu seed layer by electrodeposition and patterned by ion beam milling with a photoresist mask. To prevent the oxidation of the Tb0.36Dy0.64Fe1.9 film, a 25 nm-thick Si3N4 thin film is deposited on the magnetostrictive film by sputtering. After etching the handling Si layer from the backside, the cantilever structures are released by etching the buried oxide using vapor HF etching. The SEM images of the fabricated bi-material microcantilever structures are shown in Figure 7. Owing to the stress of the films, the cantilevers are slightly bent upward.

3. Result and Discussion

Using an electromagnet, a magnetic field of 0–11 kOe is applied to the fabricated cantilevers along the parallel direction of the cantilever. The actuation is observed by a microscope, as the typical result is shown in Figure 8, where the magnetic field was applied along horizontal direction in parallel to the cantilever length direction. The magnetostrictive film on the Si cantilever has 91 MPa tensile stress as observed from the initial bending. Theoretically, Terfenol-D is known as positive magnetostriction material. The Terfenol-D material will extend toward the magnetic field direction. When a magnetic field is applied to the cantilever, the cantilever will be bent downward because of the magnetostriction effect.
With an optical microscope, this displacement could be observed, as shown in Figure 8. An application of the magnetic field actuates the cantilever downward with displacement D . Figure 9 shows the observed displacements as a function of applied magnetic field for three cantilevers with different lengths, and Figure 10 shows the magnetostriction coefficients calculated using Equation (1). Figure 11 shows the generated forces of each cantilever calculated from the cantilever deflection under various magnetic fields using Equation (3). The maximum force can be estimated to be approximately 65 mN.
The magnetostriction coefficient λ e f f can be calculated using Equation (1), as shown in Figure 10, where Dis supposed to be negligible. The actuation is saturated at approximately 5000 Oe. This actuation characteristic seems to be reasonable with the VSM result. At 11000 Oe, the Tb0.36Dy0.64Fe1.9 film shows a magnetostrictive coefficient of approximately 1250 ppm in strain, which is comparable to 1400 ppm of the magnetostriction coefficient of the bulk Terfenol-D [1,2,7]. This is the highest value among reported magnetostriction coefficients of the TbxDy(1−x)Fey films [19,20,21,22,23,24,25,26,27,28]. Also, from the displacement data and deflection data, the energy density of the thin film actuator can be calculated from the stored elastic energy W e l in the cantilever [37]. The stored elastic energy is given by
W e l = E s 1 v s ( 1 R ) 2 z 2 L l d z  
where R is the radius of the curvature of the cantilever, L is the length of the cantilever, l is the width of the cantilever and E s , v s are the Young’s modulus and Poisson ratio of the cantilever, respectively. Specific dimensions and parameters used are shown in Table 1. The radius R of the curvature R and deflection D of the cantilever is given by
R 1 = β β 2 β + 1 3 σ i n t h f h s 2 1 v s E s = 6 σ i n t h f h s 2 1 v s E s
D = 3 σ i n t h f h s 2 1 v s E s L 2
where σ i n t is the initial stress of the cantilever, and β is the constant of the neutral plane of the cantilever. The Young’s modulus of silicon is approximately three times larger than that of Terfenol-D, also the thickness of the Tb0.34Dy0.65Fe1.9 film is very thin in comparison with the Si layer; thus β = 1/2 is the proper assumption in this model. As a consequence, the stored elastic energy in the cantilever is given by
W e l = ( β 1 ) h s β h s E s 1 v s ( 1 R ) 2 z 2 L l d z = E s 1 v s ( 1 R ) 2 z 2 L l h s 3 ( β 2 β + 1 3 )
The energy density Edensity of the magnetostrictive film is obtained by dividing the stored elastic energy by the volume Vf of the magnetostrictive film, as given by
Edensity = Wel/Vf
In the actual calculation, the maximum energy density is calculated from the actuated cantilever deflection using Equation (8). The calculated and reported energy densities of the film and bulk [1,6] are summarized in Table 2. The variation of the calculated energy density seems to be large for three cantilevers, it may come from the cantilever dimension errors (possibly ±50 µm) caused by alignment error and side etching in microfabrication.
Figure 12 shows the comparison of the energy density of the Tb0.36Dy0.64Fe1.9 film with another actuator. It is found that this magnetostrictive film can produce very high energy density for actuation. It is found that the energy density can be higher than that of piezoelectric material (PZT) that is widely used for microelectromechcanical devices.
In Table 3, the magnetostriction coefficient of this Tb0.36Dy0.64Fe1.9 film is compared with that of bulk materials and reported magnetostrictive thin films. The electrodeposited Galfenol and CoFe films show a relatively low performance in comparison with the sputtered films. It is considered that this low performance possibly comes from random lattice orientation [6,9]. The sputtered and annealed Terfenol-D and CoFe show better than non-annealed film [11,19,20]. This better performance comes from improved crystallinity and grain size. It can be concluded that the electrodeposited Tb0.36Dy0.64Fe1.9 film shows excellent magnetostriction performance than that of other types of magnetostrictive thin films and has high potential ability for the application to microelectromechanical systems (MEMS) including magnetic actuators, energy harvesters, and microsensors.

4. Conclusion

This paper reported the performance of a Tb0.36Dy0.64Fe1.9 film deposited by electrodeposition at 40 °C. The deposited Tb0.36Dy0.64Fe1.9 film shows the coercive magnetic field 285 Oe and the saturation magnetic field 5000 Oe. From AFM and MFM analysis, the film has ~170 nm grain size. At 11 kOe magnetic field, the Tb0.36Dy0.64Fe1.9 film shows approximately 1250 ppm of magnetostriction coefficient. Moreover, the energy density of the film is calculated to be 100,000~165,000 J/m3. These performances are almost the same to those of bulk Terfenol-D. As a consequence, the electrodeposited Tb0.36Dy0.64Fe1.9 film has a high potential ability for magnetic actuator, energy harvesting, and sensor applications.

Author Contributions

Conceptualization, T.O. and H.S.; support to film deposition, K.S.; validation, formal analysis, investigation, and data curation, H.S.; writing—original draft preparation, H.S.; writing—review and editing, T.O.; supervision, N.I., M.T., N.V.T., and T.O.; funding acquisition, T.O. All authors have read and agreed to the published version of the manuscript.

Funding

A part of this work was funded by The New Energy and Industrial Technology Development Organization, NEDO.

Acknowledgments

This work is partly done at Micro/nanomachining education center, Tohoku University, and Micro System Integration Center (µSiC), Tohoku University. This work was partly supported by Cabinet Office, Government of Japan, Cross-ministerial Strategic Innovation Promotion Program (SIP).

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

References

  1. Engdahl, G.; Mayergoyz, I.D. Handbook of Giant Magnetostrictive Materials; Academic Press: Cambridge, MA, USA, 2000. [Google Scholar]
  2. Clark, A.E. Handbook of Ferromagnetic Material; North-Holland Publishing Company: Amsterdam, The Netherlands, 1980; Volume 1, pp. 531–583. [Google Scholar]
  3. Buschow, K.H.J.; Wohlfarth, E.P. Handbook of Ferromagnetic Material; Elsevier Science Publishers B.V: Amsterdam, The Netherlands, 1990; Volume 5, pp. 1–132. [Google Scholar]
  4. Mnyukh, Y. The True Cause of Magnetostriction. Am. J. Condens. Matter Phys. 2014, 4, 57–62. [Google Scholar]
  5. McGary, P.D.; Reddy, K.S.; Haugstad, G.D.; Stadler, B.J. Combinatorial Electrodeposition of Magnetostrictive Fe1−xGax. J. Electrochem. Soc. 2010, 157, D656–D665. [Google Scholar] [CrossRef]
  6. Ng, J.H.G.; Record, P.M.; Shang, X.; Wlodarczyk, K.L.; Hand, D.P.; Schiavone, G.; Abraham, E.; Cummins, G.; Desmulliez, M.P. Optimised co-electrodeposition of Fe–Ga alloys for maximum magnetostriction effect. Sens. Actuators A 2015, 223, 91–96. [Google Scholar] [CrossRef] [Green Version]
  7. Houqing, Z.; Jianguo, L.; Xiurong, W.; Yanhong, X.; Hongping, Z. Applications of Terfenol-D in China. J. Alloys Compd. 1997, 258, 49–52. [Google Scholar] [CrossRef]
  8. Thang, P.D.; Rijnders, G.; Blank, D.H.A. Stress-induced magnetic anisotropy of CoFe2O4 thin films using pulsed laser deposition. J. Magn. Magn. Mater. 2007, 3, 2621–2623. [Google Scholar] [CrossRef]
  9. Özkale, B.; Shamsudhin, N.; Bugmann, T.; Nelson, B.J.; Pané, S. Magnetostriction in electroplated CoFe alloys. Electrochem. Commun. 2017, 76, 15–19. [Google Scholar] [CrossRef]
  10. Yamaura, S.I.; Nakajima, T.; Satoh, T.; Ebata, T.; Furuya, Y. Magnetostriction of heavily deformed Fe–Co binary alloys prepared by forging and cold rolling. Mater. Sci. Eng. B 2015, 193, 121–129. [Google Scholar] [CrossRef] [Green Version]
  11. Hunter, D.; Osborn, W.; Wang, K.; Kazantseva, N.; Hattrick-Simpers, J.; Suchoski, R.; Takahashi, R.; Young, M.L.; Mehta, A.; Bendersky, L.A.; et al. Giant magnetostriction in annealed Co1–xFex thin-films. Nat. Commun. 2011, 2, 518. [Google Scholar] [CrossRef] [Green Version]
  12. Yamazaki, T.; Yamamoto, T.; Furuya, Y.; Nakao, W. Magnetic and magnetostrictive properties in heat-treated Fe-Co wire for smart material/ device. Jpn. Soc. Mech. Eng. 2018, 5, 17–00569. [Google Scholar] [CrossRef] [Green Version]
  13. Tiercelin, N.; Youssef, J.B.; Preobrazhensky, V.; Pernod, P.; Le Gall, H. Giant magnetostrictive superlattices: From spin reorientation transition to MEMS. Static and dynamical properties. J. Magn. Magn. Mater. 2002, 249, 519–523. [Google Scholar] [CrossRef]
  14. Lim, S.H.; Han, S.H.; Kim, H.J.; Choi, Y.S.; Choi, J.W.; Ahn, C.H. Prototype microactuators driven by magnetostrictive thin films. IEEE Trans. Magn. 1998, 34, 2042–2044. [Google Scholar] [CrossRef]
  15. Fu, L.; Li, S.; Zhang, K.; Chen, I.; Petrenko, V.; Cheng, Z. Magnetostrictive Microcantilever as an Advanced Transducer for Biosensors. Sensors 2007, 7, 2929–2941. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Fu, L.; Li, S.; Zhang, K.; Chen, I.H.; Barbaree, J.M.; Zhang, A.; Cheng, Z. Detection of Bacillus anthracis Spores Using Phage-Immobilized Magnetostrictive Milli/Micro Cantilevers. IEEE Sens. J. 2011, 11, 1684–1691. [Google Scholar] [CrossRef]
  17. Wang, L.; Yuan, F.G. Vibration energy harvesting by magnetostrictive material. Smart Mater. Struct. 2008, 17, 045009. [Google Scholar] [CrossRef]
  18. Kawashima, K.; Mineta, T.; Makino, E.; Kawashima, T.; Shibata, T. Self-Align Fabrication of Narrow-Gapped Dual AFM Tip Using Si Trench Refilling with SOG and Magnetostrictive Film Stacked Dual Cantilever Formation. Electron. Commun. Jpn. 2015, 98, 90–95. [Google Scholar] [CrossRef]
  19. Panduranga, M.K.; Lee, T.; Chavez, A.; Prikhodko, S.V.; Carman, G.P. Polycrystalline Terfenol-D thin films grown at CMOS compatible temperature. AIP Adv. 2018, 8, 056404-1-5. [Google Scholar] [CrossRef] [Green Version]
  20. Mohanchandra, K.P.; Prikhodko, S.V.; Wetzlar, K.P.; Sun, W.Y.; Nordeen, P.; Carman, G.P. Sputter deposited Terfenol-D thin films for multiferroic applications. AIP Adv. 2015, 5, 097119-1-9. [Google Scholar] [CrossRef] [Green Version]
  21. Kerrigan, C.A.; Ho, K.K.; Mohanchandra, K.P.; Carman, G.P. Sputter deposition and analysis of thin film Nitinol/Terfenol-D multilaminate for vibration damping. Smart Mater. Struct. 2009, 18, 015007. [Google Scholar] [CrossRef]
  22. Loveless, M.; Guruswamy, S. Texture in magnetic annealed Terfenol-D films. J. Appl. Phys. 1996, 79, 6222–6224. [Google Scholar] [CrossRef]
  23. Schatz, F.; Hirscher, M.; Schnell, M.; Flik, G.; Kronmiiller, H. Magnetic anisotropy and giant magnetostriction of amorphous TbDyFe films. J. Appl. Phys. 1994, 76, 5380–5382. [Google Scholar] [CrossRef]
  24. Quandt, E. Multitarget sputtering of high magnetostrictive Tb-Dy-Fe films. J. Appl. Phys. 1994, 75, 5653–5655. [Google Scholar] [CrossRef]
  25. Liu, M.; Li, S.; Zhou, Z.; Beguhn, S.; Lou, J.; Xu, F.; Jian Lu, T.; Sun, N.X. Electrically induced enormous magnetic anisotropy in Terfenol-D/lead zinc niobatelead titanate multiferroic heterostructures. J. Appl. Phys. 2012, 112, 063917-1-4. [Google Scholar] [CrossRef]
  26. Loveless, M.; Guruswamy, S.; Shield, J.E. Crystallization Behavior of Amorphous Terfenol-D Thin Films. IEEE Trans. Magn. 2012, 33, 3937–3939. [Google Scholar] [CrossRef]
  27. Jerzy, K.; Rafal, M.; Mariusz, H. Magnetomechanical properties of Terfenol-D based composites. In Proceedings of the 6th International Conference on Mechanics and Materials in Design Ponta Delgada/Azores, Ponta Delgada, Portugal, 26–30 July 2015; pp. 683–690. [Google Scholar]
  28. Meníc, J.; Quandt, E.; Munz, D. Elastic modulus of TbDyFe films a comparison of nanoindentation and bending measurements. Thin Solid Films 1996, 287, 208–213. [Google Scholar]
  29. Gong, J.; Podlaha, E.J. Electrodeposition of Fe-Tb Alloys from an Aqueous Electrolyte. Electrochem. Solid State Lett. 2000, 3, 422–425. [Google Scholar] [CrossRef]
  30. Zarkadas, G.M.; Stergiou, A.; Papanastasiou, G. Influence of citric acid on the silver electrodeposition from aqueous AgNO3 solutions. Electrochim. Acta 2005, 50, 5022–5031. [Google Scholar] [CrossRef]
  31. Aaboubi, O.; Douglade, J.; Abenaqui, X.; Boumedmed, R.; VonHoff, J. Influence of tartaric acid on zinc electrodeposition from sulphate bath. Electrochim. Acta 2011, 56, 7885–7889. [Google Scholar] [CrossRef]
  32. Fashu, S.; Gu, C.D.; Zhang, J.L.; Huang, M.L.; Wang, X.L.; Tu, J.P. Effect of EDTA and NH4Cl additives on electrodeposition of Zn−Ni films from choline chloride-based ionic. Trans. Nonferrous Met. Soc. China 2015, 25, 2054–2064. [Google Scholar] [CrossRef]
  33. Wang, T.; Chen, Y.N.; Chiang, C.C.; Hsieh, Y.K.; Li, P.C.; Wang, C.F. Carbon-Coated Hematite Electrodes with Enhanced Photoelectrochemical Performance Obtained through an Electrodeposition Method with a Citric Acid Additive. ChemElectroChem 2016, 3, 966–975. [Google Scholar] [CrossRef]
  34. Shim, H.; Sakamoto, K.; Inomata, N.; Toda, M.; Van Toan, N.; Song, Y.; Ono, T. Magnetostrictive performance of electrodeposited TbxDy1-xFey thin film evaluated from microactuator. In Proceedings of the 2019 20th International Conference on Solid-State Sensors, Actuators and Microsystems & Eurosensors XXXIII (TRANSDUCERS & EUROSENSORS XXXIII), Berlin, Germany, 23–27 June 2019; pp. 1698–1700. [Google Scholar]
  35. Yang, Y.V.; Huang, Y.Y.; Jin, Y.M. Effects of magnetocrystalline anisotropy constant K2 on magnetization and magnetostriction of Terfenol-D. Appl. Phys. Lett. 2011, 9, 012503. [Google Scholar] [CrossRef]
  36. Klokholm, E. The measurement of magnetostriction in ferromagnetic thin films. IEEE Trans. Magn. 1976, 12, 819–821. [Google Scholar] [CrossRef]
  37. De Lacheisserie, E.D.T.; Peuzin, J.C. Magnetostriction and internal stresses in thin films: The cantilever method revisited. J. Magn. Magn. Mater. 1994, 136, 189–196. [Google Scholar] [CrossRef]
  38. Budynas, R.G.; Young, W.C.; Sadegh, A. Roark’s Formulas for Stress and Strain, 8th ed.; Mc Garw Hill: New York, NY, USA, 2012; pp. 165–169. [Google Scholar]
  39. Antunes, J.M.; Fernandes, J.V.; Sakharova, N.A.; Oliveira, M.C.; Menezes, L.F. On the determination of the Young’s modulus of thin films using indentation tests. Int. J. Solids Struct. 2007, 44, 8313–8334. [Google Scholar] [CrossRef] [Green Version]
  40. Krulevitch, P.; Lee, A.P.; Ramsey, P.B.; Trevino, J.C.; Hamilton, J.; Northrup, M.A. Thin film shape memory alloy microactuators. J. Microelectromech. Syst. 1996, 5, 270–282. [Google Scholar] [CrossRef]
Figure 1. Energy-dispersive X-ray spectroscopy (EDX) spectrum of the TbxDy(1−x)Fey film formed at an electrochemical potential of −930 mV.
Figure 1. Energy-dispersive X-ray spectroscopy (EDX) spectrum of the TbxDy(1−x)Fey film formed at an electrochemical potential of −930 mV.
Micromachines 11 00523 g001
Figure 2. Composition dependence on the applied working electrode potential on a Cu seed layer.
Figure 2. Composition dependence on the applied working electrode potential on a Cu seed layer.
Micromachines 11 00523 g002
Figure 3. In-plane magnetization measurement of the TbxDy(1−x)Fey film using vibrating sample magnetometer (VSM).
Figure 3. In-plane magnetization measurement of the TbxDy(1−x)Fey film using vibrating sample magnetometer (VSM).
Micromachines 11 00523 g003
Figure 4. Atomic force microscopy image of the deposited film.
Figure 4. Atomic force microscopy image of the deposited film.
Micromachines 11 00523 g004
Figure 5. Magnetic force microscopy the deposited film.
Figure 5. Magnetic force microscopy the deposited film.
Micromachines 11 00523 g005
Figure 6. Fabrication process of the magnetostrictive bi-material cantilevers.
Figure 6. Fabrication process of the magnetostrictive bi-material cantilevers.
Micromachines 11 00523 g006
Figure 7. SEM images of fabricated Tb0.34Dy0.65Fe1.9 bi-material cantilevers. (a) Low magnification of bi-material cantilevers; (b) High magnification of bi-material cantilevers.
Figure 7. SEM images of fabricated Tb0.34Dy0.65Fe1.9 bi-material cantilevers. (a) Low magnification of bi-material cantilevers; (b) High magnification of bi-material cantilevers.
Micromachines 11 00523 g007
Figure 8. Optical images of side view for the typical magnetostrictive actuation of the Si- Tb0.34Dy0.65Fe1.9 bi-material cantilever for the cases without magnetic field and with a magnetic field of 11 kOe along the cantilever direction.
Figure 8. Optical images of side view for the typical magnetostrictive actuation of the Si- Tb0.34Dy0.65Fe1.9 bi-material cantilever for the cases without magnetic field and with a magnetic field of 11 kOe along the cantilever direction.
Micromachines 11 00523 g008
Figure 9. Observed displacements of the cantilevers with different lengths (700, 1000, 1100 µm).
Figure 9. Observed displacements of the cantilevers with different lengths (700, 1000, 1100 µm).
Micromachines 11 00523 g009
Figure 10. Magnetostriction coefficients of the Tb0.36Dy0.64Fe1.9 film obtained from three cantilevers with lengths 700, 1000, 1100 µm.
Figure 10. Magnetostriction coefficients of the Tb0.36Dy0.64Fe1.9 film obtained from three cantilevers with lengths 700, 1000, 1100 µm.
Micromachines 11 00523 g010
Figure 11. Generated forces of Tb0.36Dy0.64Fe1.9 obtained from three cantilevers with lengths 700, 1000, 1100 µm.
Figure 11. Generated forces of Tb0.36Dy0.64Fe1.9 obtained from three cantilevers with lengths 700, 1000, 1100 µm.
Micromachines 11 00523 g011
Figure 12. Comparison of energy densities for Tb0.36Dy0.64Fe1.9 and another types of actuators [40].
Figure 12. Comparison of energy densities for Tb0.36Dy0.64Fe1.9 and another types of actuators [40].
Micromachines 11 00523 g012
Table 1. Typical dimensions of the fabricated cantilevered structure.
Table 1. Typical dimensions of the fabricated cantilevered structure.
Cantilever Length500 µm to 1.1 mm; 100 µm Step
Cantilever width100 µm
SOI wafer dimension (Si/SiO2/Si)1.5 µm/3 µm/550 µm
Cu Seed layer thickness300 nm
TbxDy(1−x)Fey thickness250 nm
Table 2. Energy density of bulk Terfenol-D and electrochemical deposited Tb0.36Dy0.64Fe1.9.
Table 2. Energy density of bulk Terfenol-D and electrochemical deposited Tb0.36Dy0.64Fe1.9.
MaterialsPower Density
Bulk Terfenol-D5000 to 25,000 J/m3
Tb0.36Dy0.64Fe1.9 (700 µm-long cantilever at 11 kOe)129,000 J/m3
Tb0.36Dy0.64Fe1.9 (1000 µm-long cantilever at 11 kOe)169,000 J/m3
Tb0.36Dy0.64Fe1.9 (1100 µm-long cantilever at 11 kOe)100,000 J/m3
Table 3. Comparison of the magnetostrictive coefficients among bulk values, sputtered, and electrochemical deposited films.
Table 3. Comparison of the magnetostrictive coefficients among bulk values, sputtered, and electrochemical deposited films.
MaterialsMagnetostriction Coefficient (ppm)Refs
Bulk Terfenol-D1400[1,2,7]
Electrodeposited Tb0.36Dy0.64Fe1.9 at 11 kOe1250This work
Sputtered Terfenol-D at 6 kOe450[24]
Sputtered Terfenol-D at 10 kOe540[23]
Sputtered Terfenol-D annealed 400 °C at 740 emu/cc910[20]
Sputtered Terfenol-D annealed 450 °C at 700 emu/cc880[19]
Electrodeposited Galfenol at 628 Oe96[6]
Bulk Galfenol320~400[5,6]
Sputtered Co0.66Fe0.34 annealed 800 °C260[11]
Electrodeposited Co0.65Fe0.351.5[9]
Bulk TbFe22630[1,2]

Share and Cite

MDPI and ACS Style

Shim, H.; Sakamoto, K.; Inomata, N.; Toda, M.; Toan, N.V.; Ono, T. Magnetostrictive Performance of Electrodeposited TbxDy(1−x)Fey Thin Film with Microcantilever Structures. Micromachines 2020, 11, 523. https://doi.org/10.3390/mi11050523

AMA Style

Shim H, Sakamoto K, Inomata N, Toda M, Toan NV, Ono T. Magnetostrictive Performance of Electrodeposited TbxDy(1−x)Fey Thin Film with Microcantilever Structures. Micromachines. 2020; 11(5):523. https://doi.org/10.3390/mi11050523

Chicago/Turabian Style

Shim, Hang, Kei Sakamoto, Naoki Inomata, Masaya Toda, Nguyen Van Toan, and Takahito Ono. 2020. "Magnetostrictive Performance of Electrodeposited TbxDy(1−x)Fey Thin Film with Microcantilever Structures" Micromachines 11, no. 5: 523. https://doi.org/10.3390/mi11050523

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