Sensing behavior of ferromagnetic shape memory Ni-Mn-Ga

Similar documents
Magnetomechanical characterization and unified actuator/sensor modeling of ferromagnetic shape memory alloy Ni-Mn-Ga

SMASIS2008 SMASIS

A continuum thermodynamics model for the sensing effect in ferromagnetic shape memory Ni Mn Ga

Voltage generation induced by mechanical straining in magnetic shape memory materials

CHARACTERIZATION AND MODELING OF THE FERROMAGNETIC SHAPE MEMORY ALLOY Ni-Mn-Ga FOR SENSING AND ACTUATION

Dynamic strain-field hysteresis model for ferromagnetic shape memory Ni-Mn-Ga

3D-FEM-Simulation of Magnetic Shape Memory Actuators

Fully-Coupled magnetoelastic model for Galfenol alloys incorporating eddy current losses and thermal relaxation

The Use of Multiphysics Models in the Design and Simulation of Magnetostrictive Transducers. Dr. Julie Slaughter ETREMA Products, Inc Ames, IA

CONSTITUTIVE MODELING OF MAGNETIC SHAPE MEMORY ALLOYS WITH MAGNETO-MECHANICAL COUPLING ABSTRACT

Computation of magnetic field in an actuator

Characterization and finite element modeling of Galfenol minor flux density loops

Stress-strain relationship in Terfenol-D ABSTRACT. Keywords: Terfenol-D, magnetostriction, transducer, elastic modulus, Delta-E effect

Lecture 19. Measurement of Solid-Mechanical Quantities (Chapter 8) Measuring Strain Measuring Displacement Measuring Linear Velocity

Linear Magnetostrictive Models in Comsol

IEEE TRANSACTIONS ON MAGNETICS, VOL. 36, NO. 3, MAY transducer. Marcelo J. Dapino, Ralph C. Smith, and Alison B. Flatau.

3D Dynamic Finite Element Model for Magnetostrictive Galfenol-based Devices

APPLICATION OF THE NONLINEAR HARMONICS METHOD TO CONTINUOUS

This is an electronic reprint of the original article. This reprint may differ from the original in pagination and typographic detail.

based on HBLS Smart Materials

DESIGN OF A HIGH-EFFICIENCY MAGNETORHEOLOGICAL VALVE

6.4 A cylindrical specimen of a titanium alloy having an elastic modulus of 107 GPa ( psi) and

Design of a magnetostrictive (MS) actuator

Piezoelectric Composites as Bender Actuators

Modeling of 3D Magnetostrictive Systems with Application to Galfenol and Terfenol-D Transducers

Stress-Strain Behavior

MAGNETIC FLUX LEAKAGE INVESTIGATION OF INTERACTING DEFECTS: COMPETITIVE EFFECTS OF STRESS CONCENTRATION AND MAGNETIC SHIELDING

Lecture 20. Measuring Pressure and Temperature (Chapter 9) Measuring Pressure Measuring Temperature MECH 373. Instrumentation and Measurements

SURFACE BARKHAUSEN NOISE INVESTIGATIONS OF STRESS AND LEAKAGE FLUX

Characterization of residual stresses in ferrous components by magnetic anisotropy measurements using a hall effect sensor array probe

Sensors and Actuators A: Physical

Magnetic Shape Memory Alloys as smart materials for micro-positioning devices

An Energy-Based Hysteresis Model for Magnetostrictive Transducers

ABSTRACT. Associate Professor, Alison B. Flatau, Department of Aerospace Engineering

Project PAJ2 Dynamic Performance of Adhesively Bonded Joints. Report No. 3 August Proposed Draft for the Revision of ISO

Surface Magnetic Non-Destructive Testing

STRAIN GAUGES YEDITEPE UNIVERSITY DEPARTMENT OF MECHANICAL ENGINEERING

Sensor Measurements For Diagnostic Equipment

10 Measurement of Acceleration, Vibration and Shock Transducers

A REVIEW AND COMPARISON OF HYSTERESIS MODELS FOR MAGNETOSTRICTIVE MATERIALS

Open Loop Nonlinear Optimal Tracking Control of a Magnetostrictive Terfenol-D Actuator

FINITE ELEMENT MODELING OF THE BULK MAGNITIZATION OF RAILROAD WHEELS TO IMPROVE TEST CONDITIONS FOR MAGNETOACOUSTIC

The science of elasticity

The performance of a magnetorheological fluid in squeeze mode

1.103 CIVIL ENGINEERING MATERIALS LABORATORY (1-2-3) Dr. J.T. Germaine Spring 2004 LABORATORY ASSIGNMENT NUMBER 6

Transducer based measurements of Terfenol-D material properties. Frederick T. Calkins Alison B. Flatau

Validation of High Displacement Piezoelectric Actuator Finite Element Models

PIEZOELECTRIC TECHNOLOGY PRIMER

Detection of Wear in Oilwell Service Tubings using Magnetic Flux Leakage

Slide 1. Temperatures Light (Optoelectronics) Magnetic Fields Strain Pressure Displacement and Rotation Acceleration Electronic Sensors

Electrical to mechanical - such as motors, loudspeakers, charged particle deflection.

DEVELOPMENT OF A NON-CONTACTING STRESS MEASUREMENT SYSTEM DURING TENSILE TESTING USING THE ELECTROMAGNETIC ACOUSTIC TRANSDUCER FOR A LAMB WAVE

Module 6: Smart Materials & Smart Structural Control Lecture 33: Piezoelectric & Magnetostrictive Sensors and Actuators. The Lecture Contains:

Mapping of power consumption and friction reduction in piezoelectrically-assisted ultrasonic lubrication

NEGATIVE MAGNETOSTRICTIVE DELAY LINES USED IN SENSING APPLICATIONS

Module 2 Mechanics of Machining. Version 2 ME IIT, Kharagpur

Mathematical Modelling and Simulation of Magnetostrictive Materials by Comsol Multiphysics

ABSTRACT. Frank Graham, M.S., Professor, Alison Flatau, Aerospace Engineering

INTRODUCTION TO PIEZO TRANSDUCERS

Module I Module I: traditional test instrumentation and acquisition systems. Prof. Ramat, Stefano

Pre-stressed Curved Actuators: Characterization and Modeling of their Piezoelectric Behavior

ME 243. Mechanics of Solids

DEVELOPMENT OF DROP WEIGHT IMPACT TEST MACHINE

FINITE ELEMENT MODELLING OF COMPOSITES USING PIEZOELECTRIC MATERIAL

Review of Basic Electrical and Magnetic Circuit Concepts EE

Investigation of magneto-mechanical properties of Galfenol for coil-less magnetic force control using inverse magnetostrictive effect

Application of the HES in Angular Analysis

Vibration control of a beam using linear magnetostrictive actuators

Fully coupled discrete energy-averaged model for Terfenol-D

3D-FEM-Simulation of Magnetic Shape Memory Actuators

Numerical Modelling of Dynamic Earth Force Transmission to Underground Structures

PROPERTY STUDY ON EMATS WITH VISUALIZATION OF ULTRASONIC PROPAGATION

SOUTH AFRICAN NATIONAL STANDARD. Modulus of elasticity and modulus of rupture in static bending of fibreboards Amdt 1

STANDARD SAMPLE. Reduced section " Diameter. Diameter. 2" Gauge length. Radius

Piezo materials. Actuators Sensors Generators Transducers. Piezoelectric materials may be used to produce e.g.: Piezo materials Ver1404

Last Revision: August,

Optical Method for Micro Force Measurement. Yusaku FUJII Gunma University

Analysis and Experiments of the Linear Electrical Generator in Wave Energy Farm utilizing Resonance Power Buoy System

Mechatronics II Laboratory EXPERIMENT #1: FORCE AND TORQUE SENSORS DC Motor Characteristics Dynamometer, Part I

1439. Numerical simulation of the magnetic field and electromagnetic vibration analysis of the AC permanent-magnet synchronous motor

Apparent stress-strain relationships in experimental equipment where magnetorheological fluids operate under compression mode

Investigating the Hall effect in silver

1.105 Solid Mechanics Laboratory Fall 2003 Experiment 3 The Tension Test

Resistance of a commercial magnetorheological fluid to penetration

CHAPTER 5 QUASI-STATIC TESTING OF LARGE-SCALE MR DAMPERS. To investigate the fundamental behavior of the 20-ton large-scale MR damper, a

Dynamic Mechanical Analysis of Solid Polymers and Polymer Melts

E102. Study of the magnetic hysteresis

Magneto-Mechanical Modeling and Simulation of MEMS Sensors Based on Electroactive Polymers

Bending Load & Calibration Module

DEVELOPMENT OF AUTOMATIC CONTROL OF MULTI-STAGE TRIAXIAL TESTS AT THE UNIVERSITY OF MISKOLC

Osaka University, 2-1, Yamada-oka, Suita, Osaka , Japan. Naka-gun, Ibaraki , Japan

Electromagnetic Formation Flight Progress Report: July 2002

WE address the modeling of the magnetomechanical behavior

Non-linear Properties of Piezoelectric Ceramics

Magnetism & Electromagnetism

MAGNETIC MATERIAL CHARACTERIZATION BY OPEN SAMPLE MEASUREMENTS

magneticsp17 September 14, of 17

Mechatronics II Laboratory EXPERIMENT #1 MOTOR CHARACTERISTICS FORCE/TORQUE SENSORS AND DYNAMOMETER PART 1

Mechanical Behavior of Composite Tapered Lamina

Laboratory 4 Topic: Buckling

Transcription:

Sensing behavior of ferromagnetic shape memory Ni-Mn-Ga Neelesh N. Sarawate a and Marcelo J. Dapino a a Mechanical Engineering Department, The Ohio State University, 65 Ackerman Road, Columbus OH 4322, USA; ABSTRACT Due to their large magnetic field induced strains and fast response potential, ferromagnetic shape memory alloys have mainly been studied from the perspective of actuator applications. This paper presents characterization measurements on a commercial Ni-Mn-Ga alloy with a goal to investigate its feasibility as a deformation sensor. Experimental determination of flux density as a function of quasistatic strain loading and unloading at various fixed magnetic fields gives the bias field needed for maximum recoverable flux density change. This bias field is shown to mark the transition from irreversible (quasiplastic) to reversible (pseudoelastic) stress strain behavior. A reversible flux density change of 145 mt is observed over a range of 5.8 % strain and 4.4 MPa stress at a bias field of 368 ka/m. The alloy investigated therefore shows potential as a high-compliance, high-displacement deformation sensor. Keywords: Ferromagnetic shape memory alloy, Ni-Mn-Ga, magnetomechanical sensor 1. INTRODUCTION AND BACKGROUND Ferromagnetic shape memory alloys in the Ni-Mn-Ga system exhibit strains of up to 1 % when exposed to magnetic fields. 1 The strain in these alloys result from the rotation of martensite twin variants in response to external magnetic fields or external stresses. Due to the magnetic field activation, Ni-Mn-Ga exhibits higher frequency response than conventional shape memory alloys. 2 The large strain and extended frequency bandwidth make these materials useful for actuator applications. 3 Consequently, the existing literature on Ni-Mn-Ga is largely focused on the effect of magnetic field on strain. 4 8 Several models have been developed which quantify the reversible and irreversible strains observed for drive configurations with uniaxial stress and field applied parallelaswellasperpendicular to each other. 9 11 The effect of external mechanical input on the magnetic behavior of Ni-Mn-Ga, or sensor effect, has received only limited attention in the literature. Mullner et al. 12 experimentally studied strain-induced changes in the flux density of a single crystal with composition Ni 51 Mn 28 Ga 21 under external quasistatic loading at a constant field of.7 T (558 ka/m). Suorsa et al. 13 reported magnetization measurements conducted on stoichiometric Ni 2 MnGa material for various discrete strain and field intensities respectively ranging between 6 % and 5 12 ka/m. Straka and Heczko 14, 15 reported superelastic response of a Ni 49.7 Mn 29.1 Ga 21.2 single crystal with 5M martensitic structure for fields higher than 239 ka/m and established the interconnection between magnetization and strain. Heczko 16 further investigated this interconnection and proposed a simple energy model. The paper presents experimental measurements on the dependence of flux density with deformation, stress and magnetic field in a commercially-available Ni-Mn-Ga alloy, with a view to determining the bias field needed for obtaining maximum reversible deformation sensing as well as the associated strain and stress ranges. A reversible flux density change of 145 mt is observed over a range of 5.8 % strain and 4.4 MPa stress at a bias field of 368 ka/m. Further author information: (Send correspondence to M.J.D.) N.N.S.: E-mail: sarawate.1@osu.com, Telephone: 1 614 247 748 M.J.D.: E-mail: dapino.1@osu.edu, Telephone: 1 614 688 3689 Smart Structures and Materials 26: Active Materials: Behavior and Mechanics, edited by William D. Armstrong, Proc. of SPIE Vol. 617, 6171B, (26) 277-786X/6/$15 doi: 1.1117/12.65813 Proc. of SPIE Vol. 617 6171B-1

Hall probe (Flux density measurement) σ Load cell (2 lb) Electromagnet Pole piece(s) H Ni-Mn-Ga sample Pushrod(s) Figure 1. Experimental setup for quasistatic strain loading of Ni-Mn-Ga. 2. EXPERIMENTAL SETUP As shown in Figure 1, the experimental setup consists of a custom built electromagnet and a loading stage. The electromagnet is made from laminated transformer steel and can produce DC and low frequency AC fields of up to 8 ka/m. It consists of two opposing E-halves which complete the magnetic flux path. Two parallel coils of about 55 turns each are mounted on the tapered center legs of opposing E-halves. The air gap of 8 mm in between the center legs is sufficient to accommodate a Ni-Mn-Ga sample and Hall probe. The variation of magnetic field in the air gap is less than 2% around the area of the pole faces. The electromagnet is powered by a 1 VA amplifier. A6x6x2mm 3 single crystal Ni-Mn-Ga sample (AdaptaMat Ltd.) is placed in the center gap of the electromagnet. The external uniaxial quasistatic strain is applied using an MTS machine with Instron controller. The sample exhibits a maximum magnetic field induced deformation of 5.8%. Initially, the sample is converted to a single field-preferred variant by applying a transverse DC field of 72 ka/m with no mechanical loading. The sample is subsequently compressed at a fixed displacement rate of.254 mm/sec, and unloaded at the same rate. The flux density inside the material is measured using a Walker Scientific MG-4D Gaussmeter with a transverse Hall probe with active area 1 x 2 mm 2 placed in the gap between the magnet pole and the face of the sample. The accuracy of the method was confirmed by FEMM software. The small air gap ensures that the flux density inside the sample and that acting on the probe are equal. The compressive force is measured by a 2 lb load cell, and the displacement is measured by an LVDT. The externally applied field is obtained from the calibration curve of the electromagnet as a function of the measured current. This process is repeated under varying magnitudes of bias fields ranging from -479 ka/m. 3. RESULTS 3.1. Stress-Strain Behavior Figure 2 shows stress vs. strain curves at varied bias fields, in which the expected magnetoelastic behavior of 12, 14 Ni-Mn-Ga is observed. The applied transverse field results in orientation of crystals with their magnetically easy c-axis in the transverse direction, which causes the sample to elongate in the longitudinal direction. This is a consequence of growth of field-preferred martensite variants with the c-axis aligned with the external magnetic field. Application of compressive stresses favors the growth of stress-preferred twin variants with the c-axis in the longitudinal direction. Thus, in this configuration magnetic fields and stresses create competing effects. Proc. of SPIE Vol. 617 6171B-2

12 1 Compressive Stress (MPa) 8 6 4 2 94 ka/m 173 ka/m 251 ka/m 368 ka/m 479 ka/m 2.1.2.3.4.5.6 Compressive Strain Figure 2. Stress-strain curves of Ni-Mn-Ga under different magnetic fields. Because the twin variants are not mobile during initial compression below the detwinning stress (3 MPa in this case), the sample exhibits a relatively high stiffness in this region. With further application of stress, the rearrangement of twin variants, i.e., growth of stress-preferred variants at the expense of field preferred variants, continues resulting in a low stiffness region. This rearrangement continues until the sample is converted to one variant preferred by stress. When the sample is loaded further, it follows the steeper region indicating high stiffness after the twin rearrangement has been completed. The stress-strain behavior varies with applied magnetic field. As the effect of stress is opposite that of the applied field, the detwinning stress increases with increasing bias fields. The external stress has to do more work at higher applied fields to initiate the rearrangement of twin variants. The detwinning stress is a characteristic of the specimen, and is an important parameter for model development. 17 During unloading, the stress-strain curves show reversible or irreversible behavior depending on the magnitude of bias field. At low fields, the sample does not return to its original shape. The stress-induced deformation in the longitudinal direction remains almost unchanged. This is because the magnetic energy is not high enough to initiate the redistribution of twin variants. This irreversible behavior is also termed as quasiplastic behavior. This effect is analogous to the actuation effect under no or small load, when the field induced strain in the sample remains even after the field is removed. In that case, the stress is not strong enough to initiate growth of stress-preferred variants to bring the sample to its original length. At high bias fields, the sample exhibits reversible behavior known as magnetic field induced superelasticity or pseudoelasticity. The magnetic energy is high enough to nucleate the growth of field-preferred variants when the sample is unloaded. Again, this is analogous to actuation under moderate stress when the sample returns to its original dimensions after every cycle resulting in reversible actuation. For the intermediate bias fields, the material exhibits partial recovery. The field is strong enough to initiate growth of field-preferred variants but not strong enough to achieve complete strain recovery. 3.2. Flux Density Dependence on Strain and Stress The flux density plots shown in Figures 3 4 are of interest for sensing applications. The absolute value of flux density decreases with increasing compressive stress. As the sample is compressed from its initial field-preferred variant state, the stress-preferred variants are nucleated at the expense of field-preferred variants. Due to the high magnetocrystalline anisotropy of Ni-Mn-Ga, the nucleation and growth of stress-preferred variants occurs in concert with rotation of magnetization vectors into the longitudinal direction, which causes a reduction of the permeability and flux density in the transverse direction. Proc. of SPIE Vol. 617 6171B-3

.9 Flux Density (Tesla).8.7.6.5.4.3.2.1 479 ka/m 47 ka/m 368 ka/m 33 ka/m 251 ka/m 173 ka/m 133 ka/m 94 ka/m 55 ka/m.1.2.3.4.5.6 Compressive Strain Figure 3. Flux density vs compressive strain as a function of varied bias fields..9 Flux Density (Tesla).8.7.6.5.4.3.2.1 479 ka/m 47 ka/m 368 ka/m 33 ka/m 251 ka/m 173 ka/m 133 ka/m 94 ka/m 55 ka/m 1 2 3 4 5 Compressive Stress (MPa) Figure 4. Flux density vs compressive stress as a function of varied bias fields. Proc. of SPIE Vol. 617 6171B-4

top pushrod H b bottom pushrod ε, σ ε max, σ max ε, σ = (a) (b) (c) (d) (e) Figure 5. Schematic of loading and unloading at low magnetic fields. 4. DISCUSSION 4.1. Magnetic Field Induced Stress and Flux Density Recovery There is a close correlation between Figure 2 and Figures 3-4 regarding the reversibility of the magnetic and elastic behaviors. The change in flux density relative to the initial field-preferred single variant is directly associated with growth of stress-preferred variants. Thus, the flux density value returns to its initial value only if the stress vs. strain curve exhibits magnetic field induced pseudoelasticity. This occurs for this alloy at bias fields of 368 ka/m and higher. At high bias fields, the magnetic energy is high enough to initiate and complete the redistribution of variants relative to the single stress-preferred variant formed at maximum compression, during unloading. During this redistribution the magnetization vectors rotate into the transverse direction, resulting in recovery of flux density to its original value along with pseudoelastic recovery. At fields of 94 ka/m or lower, the magnetic field energy is not strong enough to initiate redistribution of variants. Hence the flux density remains unchanged while the sample is unloaded. Correspondingly the stress vs. strain curve also shows irreversible (quasiplastic) behavior. Figures 5 and 6 illustrate this mechanism in more detail. Figure 5 illustrates the compression of a simplified, two-variant FSMA structure at low bias fields. Before the compression cycle commences, a high transverse field is applied to transform the sample to a single field-preferred variant. In this configuration, all magnetization vectors align themselves in the direction of the field. When a low bias field is applied, the magnetization vectors reorient to form 18-degree stripe magnetic domains which results in lower net flux density. The magnetization vectors remain in the transverse direction, and since no external stress is applied, the field-preferred variant configuration remains intact. The compression starts at this maximum sample length, with comparatively low net flux density, panel (a). With increasing compression, the stress-preferred variants nucleate and grow. The variant nucleation is associated with rotation of magnetization vectors into the longitudinal direction, as they are attached to c-axis due to high magnetocrystalline anisotropy. This results in reduction in flux density in transverse direction, panel (b). The sample is entirely converted to stress-preferred state, but few magnetization vectors remain in the horizontal (hard) direction depending on the field strength, panel (c). When the sample is unloaded, the magnetic field energy is not high enough to initiate redistribution of variants into a single field-preferred variant state, panel (d). Hence, there is little or no change in the flux density value after unloading, panel (e), which corresponds with the fact that the stress-strain and flux density plots do not show any recovery for fields lower than 94 ka/m. Proc. of SPIE Vol. 617 6171B-5

top pushrod H b bottom pushrod ε, σ ε max, σ max ε, σ ε =, σ = (a) (b) (c) (d) (e) Figure 6. Schematic of loading and unloading at high magnetic fields. Figure 6 illustrates the effect of stress loading and unloading at high bias fields. The initial net flux density is high when the sample is at its maximum length, panel (a). As in the earlier case, there is a reduction in the transverse flux density with increasing compression, panel (b). When the sample is converted to single stresspreferred variant state, some magnetization vectors remain in the transverse direction as the bias field is large enough to force magnetic moments to break away from the c-axis, panel (c). When the unloading starts, the available magnetic energy is high enough to cause nucleation and growth of field-preferred variants, while forcing the magnetization vectors to rotate into the transverse direction. Thus, the sample starts elongating again, and the expanding sample tries to force on the pushrods resulting in increasing compressive stress, panel (d). When the sample is near zero deformation, the field is high enough to induce complete variant rearrangement, the sample returns to its original structure thus exhibiting pseudoelastic behavior, and the original value of flux density is also recovered, panel (e). Thus, the magnetic field induced pseudoelasticity occurs in concert with the recovery of flux density. This correlation can also be realized from the Figure 4, where it can be seen that the flux density-stress curves bear a resemblance to the conventional magnetic field induced strain curves. 1 Under low stresses, the strain-field plots show irreversible behavior, whereas at higher stresses the behavior is reversible. If the stress is very high (higher than the blocking force), the material shows no field induced deformation as the applied stress will be too high to allow variant rearrangement. Similarly, if the applied bias field is very high (higher than the saturation field), there will not be any change in flux density even when the sample is completely compressed. This is because the magnetic field will be too high to allow rotation of magnetization vectors in a direction perpendicular to it. An optimum compressive stress is needed to achieve maximum field induced deformation for actuation applications. Similarly, an optimum bias field is required to achieve maximum flux density change for sensing applications. 4.2. Optimum Bias Field for Sensing The flux density starts changing when the initial detwinning stress is reached and continues to change until the final detwinning stress is reached, with all variants being transformed to one variant preferred by stress. The magnitude of total change in flux density during compression can be defined as sensitivity. The sensitivity is found to initially increase with increasing bias fields and then decrease after reaching a maximum at 173 ka/m (Figure 7). However, the reversible behavior required for sensing applications is observed at bias fields of 368 ka/m and higher. Thus, 368 ka/m can be defined as optimum field for sensing applications. Proc. of SPIE Vol. 617 6171B-6

Flux Density Change (Tesla).3.25.2.15.1.5 Irreversible/ Quasiplastic Behavior Reversible/ Pseudoelastic Behavior 1 2 3 4 5 Bias Field (ka/m) Figure 7. Flux density change as function of bias field. This behavior can be explained from the easy-axis and hard-axis magnetization curves for this alloy shown in Figure 8. The easy-axis curve refers to magnetization of material along its easy axis (c-axis). It was obtained by first converting the sample to a single field-preferred variant and subsequently exposing it to a.5 Hz sinusoidal transverse field while leaving it mechanically unconstrained. The easy axis magnetization curve has a steeper slope, and it tends to saturate at low fields, about 12 ka/m in this case. The hard-axis curve refers to magnetization of material along its hard axis (other than the c-axis). To obtain the hard-axis curve, the sample was first converted to a single stress-preferred variant. Then the sample having all variants with c-axis in longitudinal direction was subsequently exposed to a.5 Hz sinusoidal field while being prevented from expanding. This means that the sample was magnetized along an axis other than the c-axis i.e. the hard axis. The hard axis magnetization curve has a lower slope with higher saturation field, 64 ka/m for this alloy. To magnetize the sample along hard axis, the externally applied field has to overcome the anisotropy energy to rotate the magnetization vectors away from the c-axis which is perpendicular to field. The mechanical constrains ensure that the field-preferred variants do not nucleate, thus maintaining the c-axis along longitudinal axis. At maximum elongation for a given bias field, the flux density value is that corresponding to the easy axis value. When the sample is compressed at a constant field the induction value changes from the corresponding easy axis value to the corresponding hard axis value. Compression at constant field corresponds to a straight line starting at easy axis curve and ending at the hard axis curve. Therefore at maximum compression, with all variants being stress-preferred, the hard axis value is the lowest flux density at given bias field. Hence, the maximum flux density change occurs when the two curves are at maximum vertical distance from each other. A large flux density change of 23 mt is observed at a bias field of 173 ka/m. However, the optimum sensing range for reversible sensing behavior occurs when the two curves are at maximum distance from each other and the sample shows pseudoelastic behavior. At a bias field of 368 ka/m, a reversible flux density change of 145 mt is obtained over a range of 5.8 % strain and 4.4 MPa stress. 5. CONCLUDING REMARKS A reversible flux density change of 145 mt is observed over a range of 5.8 % strain and 4.4 MPa stress at a bias field of 368 ka/m. By way of comparison, Terfenol-D exhibits a higher maximum sensitivity of.4 T at a lower bias field of 16 ka/m and higher stress range of 2 MPa 18 However, the associated deformation is only.1 % due to the higher Terfenol-D stiffness. The Ni-Mn-Ga alloy investigated here therefore shows potential for high-compliance, high-displacement deformation sensors. Proc. of SPIE Vol. 617 6171B-7

1.5 Flux Density (Tesla) 1.5.5 1 Hard Axis Start points during compression Easy Axis End points during compression 1.5 1 5 5 1 Applied Field (ka/m) Figure 8. Easy and hard axis flux density curves of Ni-Mn-Ga. REFERENCES 1. A. Sozinov, A. Likhachev, N. Lanska, and K. Ullakko, Giant magnetic field induced strain in Ni-Mn-Ga seven-layered martensitic phase, Applied Physics Letters, 8 (1), March 22. 2. L. Faidley, M. Dapino, G. Washington, T. Lograsso, Dynamic response in the low-khz range and delta-e effect in ferromagnetic shape memory Ni-Mn-Ga, Proceedings of IMECE (43198), 23. 3. I. Suorsa, E. Pagounis, K. Ullakko, Magnetic shape memory actuator performance, Journal of Magnetism and Magnetic materials, 272-276, pp. 229-23, 24. 4. K. Ullakko, J. Huang, C. Kantner, R. O Handley, and V. Kokorin, Large magnetic-field-induced strains in Ni2MnGa single crystals, Applied Physics Letters, 69, pp. 1966-1968, September 1996. 5. R. O Handley, Model for strain and magnetization in magnetic shape-memory alloys, Journal of Applied Physics, 83 (6), pp. 3263-327, March 1998. 6. S. Murray, M. Marioni, S. Allen, R. O Handley, and T. Lograsso, 6% magnetic field induced strain by twin boundary motion in ferromagnetic Ni-Mn-Ga, Applied Physics Letters, 77 (6), pp. 886-888, August 2. 7. C. Henry, D. Bono, J. Feuchtwanger, S. Allen, and R. O Handley, AC field induced actuation of single crystal Ni-Mn-Ga, Journal of Applied Physics, 91 (1), pp. 781-7811, May 22. 8. L. Faidley, M. Dapino, G. Washington, T. Lograsso, and R. Smith, Analytical and experimental issues in Ni-Mn-Ga transducers, Proceedings of SPIE Smart Structures and Materials 21, 4327, pp. 521-532, Newport Beach, CA, 4-8 March 21. 9. S. Murray, M. Marioni, A. Kukla, J. Robinson, R. O Handley, and S. Allen, Large field induced strain in single crystalline Ni-Mn-Ga ferromagnetic shape memory alloy, Journal of Applied Physics, 87 (9), pp. 5774-5776. 1. A. Likhachev, K. Ullakko, Magnetic field controlled twin boundaries motion and giant magneto-mechanical effects in Ni-Mn-Ga shape memory alloy, Applied Physics Letters, 275, pp. 142-151, October 2. 11. L. Faidley, M. Dapino, G. Washington, Strain model for Ni-Mn-Ga with collinear field and stress, Proceedings of IMECE 25, Orlando, FL, November 25. 12. P. Mullner, V. Chernenko, and G. Kostorz, Stress-induced twin variant rearrangement resulting in a Ni- Mn-Ga ferromagnetic martensite, Scripta Materialia, 49, pp. 129-133, 23. 13. I. Suorsa, E. Pagounis, and K. Ullakko, Magnetization dependence on strain in the Ni-Mn-Ga magnetic shape memory material, Applied Physics Letters, 84 (23), pp. 4658-466, June 24. 14. L. Straka, and O. Heczko, Superelastic response of Ni-Mn-Ga martensite in magnetic fields and a simple model, IEEE Transactions on Magnetics, 39 (5), pp. 342-344, September 23. Proc. of SPIE Vol. 617 6171B-8

15. L. Straka, and O. Heczko, Reversible 6 % strain of Ni-Mn-Ga martensite using opposing external stress in static and variable magnetic fields, Journal of Magnetism and Magnetic Materials, 29-291 (2), pp. 829-831, 25. 16. O. Heczko, Magnetic shape memory effect and magnetization reversal, Journal of Magnetism and Magnetic Materials, 29-291, pp. 787-794, 25. 17. R. Couch, I. Chopra, A quasi-static model for Ni-Mn-Ga magnetic shape memory alloy, Proceedings of SPIE Smart Structures and Materials 25. 18. R. Kellog, The Delta-E effect in terfenol-d and its applications in a tunable mechanical resonator, M.S. Thesis, Iowa State University, Ames, 2. Proc. of SPIE Vol. 617 6171B-9