A Vertical Electrostatic Actuator with Extended Digital Range via Tailored Topology

Similar documents
GENERAL CONTACT AND HYSTERESIS ANALYSIS OF MULTI-DIELECTRIC MEMS DEVICES UNDER THERMAL AND ELECTROSTATIC ACTUATION

Platform Isolation Using Out-of-plane Compliant Mechanism

Design of Electrostatic Actuators for MOEMS Applications

Simulation of a Micro-Scale Out-of-plane Compliant Mechanism

Deformation and Structural Stability of Layered Plate Microstructures Subjected to Thermal Loading

Y. C. Lee. Micro-Scale Engineering I Microelectromechanical Systems (MEMS)

CHAPTER 4 DESIGN AND ANALYSIS OF CANTILEVER BEAM ELECTROSTATIC ACTUATORS

December 1999 FINAL TECHNICAL REPORT 1 Mar Mar 98

EE C247B / ME C218 INTRODUCTION TO MEMS DESIGN SPRING 2014 C. Nguyen PROBLEM SET #4

CHAPTER 5 FIXED GUIDED BEAM ANALYSIS

An Accurate Model for Pull-in Voltage of Circular Diaphragm Capacitive Micromachined Ultrasonic Transducers (CMUT)

NONLINEAR ANALYSIS OF PULL IN VOLTAGE IN MICRO- CANTILEVER BEAM

SINGLE-STEP ASSEMBLY OF COMPLEX 3-D MICROSTRUCTURES

Optimizing micromechanical force detectors for measuring. magnetization at high magnetic fields

Electrostatic Micromechanical Actuator with. Extended Range of Travel

Solder Self-assembly for MEMS

2D BEAM STEERING USING ELECTROSTATIC AND THERMAL ACTUATION FOR NETWORKED CONTROL ABSTRACT

Curvature of a Cantilever Beam Subjected to an Equi-Biaxial Bending Moment. P. Krulevitch G. C. Johnson

Introduction to Microeletromechanical Systems (MEMS) Lecture 9 Topics. MEMS Overview

1. Narrative Overview Questions

DESIGN, FABRICATION AND ELECTROMECHANICAL CHARACTERISTICS OF A MEMS BASED MICROMIRROR

Optimizing the Performance of MEMS Electrostatic Comb Drive Actuator with Different Flexure Springs

2.76/2.760 Multiscale Systems Design & Manufacturing

MECH 466. Micro Electromechanical Systems. Laboratory Manual Laboratory #3: Stiction of MEMS and Strength of MEMS Materials

MICROELECTROMECHANICAL system (MEMS)

Tunable MEMS Capacitor for RF Applications

Proceedings of IMECE' ASME International Mechanical Engineering Congress and Exposition November 5-11, 2005, Orlando, Florida

Modeling and simulation of multiport RF switch

EE C245 ME C218 Introduction to MEMS Design

DESIGN AND SIMULATION OF ACCELEROMETER SPRINGS

Dynamic Capacitance Extraction of A Triaxial Capacitive Accelerometer

Simulation of a Polyimide Based Micromirror

Piezoelectric Resonators ME 2082

EE C247B ME C218 Introduction to MEMS Design Spring 2016

EE C245 ME C218 Introduction to MEMS Design

MEMS PARALLEL PLATE ACTUATORS: PULL-IN, PULL-OUT AND OTHER TRANSITIONS

Four Degrees-of-Freedom Micromachined Gyroscope

Modeling, Simulation and Optimization of the Mechanical Response of Micromechanical Silicon Cantilever: Application to Piezoresistive Force Sensor

Introduction to Actuators PK

The Pull-In of Symmetrically and Asymmetrically Driven Microstructures and the Use in DC Voltage References

Compliant MEMS Motion Characterization by Nanoindentation

EE C247B / ME C218 INTRODUCTION TO MEMS DESIGN SPRING 2016 C. NGUYEN PROBLEM SET #4

Design of a MEMS Capacitive Comb-drive Accelerometer

NIST ELECTROSTATIC FORCE BALANCE EXPERIMENT

International Journal of ChemTech Research CODEN (USA): IJCRGG ISSN: Vol.7, No.2, pp ,

Coupled electrostatic-elastic analysis for topology optimization using material interpolation

EE C247B / ME C218 INTRODUCTION TO MEMS DESIGN SPRING 2014 PROBLEM SET #1

Mechanics of wafer bonding: Effect of clamping

EE C245 - ME C218 Introduction to MEMS Design Fall Today s Lecture

Institute for Electron Microscopy and Nanoanalysis Graz Centre for Electron Microscopy

Abstract. 1 Introduction

Chapter 1 Introduction

EECS C245 ME C218 Midterm Exam

New Die Attach Adhesives Enable Low-Stress MEMS Packaging

Active elastomer components based on dielectric elastomers

EE C245 ME C218 Introduction to MEMS Design

Simple piezoresistive accelerometer

How Does a Microcantilever Work?

EE C245 ME C218 Introduction to MEMS Design Fall 2007

INTRODUCTION TO PIEZO TRANSDUCERS

Design and Simulation of Comb Drive Capacitive Accelerometer by Using MEMS Intellisuite Design Tool

MEMS tunable gratings with analog actuation

EE C245 ME C218 Introduction to MEMS Design Fall 2010

Research Article Dynamic Analysis of Cracked Cantilever, Electrostatic Microactuators Using Radial Basis Functions

Sensors & Transducers 2016 by IFSA Publishing, S. L.

DEVELOPMENT OF A MEMS REPULSIVE ACTUATOR FOR LARGE OUT-OF-PLANE FORCE

Basic Laboratory. Materials Science and Engineering. Atomic Force Microscopy (AFM)

AC : MEMS FABRICATION AS A MULTIDISCIPLINARY LABORATORY

Finite Element Analysis of Piezoelectric Cantilever

Foundations of MEMS. Chang Liu. McCormick School of Engineering and Applied Science Northwestern University. International Edition Contributions by

ELECTROTHERMAL MEMS MICROENGINE CAPABLE OF BI-DIRECTIONAL MOTION

Vibration Combined High Temperature Cycle Tests for Capacitive MEMS Accelerometers

ABSTRACT. Keywords: MEMS, Electrostatic, Repulsive force, Cantilever, IR Sensor, Casimir Force, Finite element analysis 1.

CAPACITIVE MICRO PRESSURE SENSORS WITH UNDERNEATH READOUT CIRCUIT USING A STANDARD CMOS PROCESS

MCE603: Interfacing and Control of Mechatronic Systems

Precise open-loop control of MEMS deformable mirror shape

Microstructure cantilever beam for current measurement

Mechanics of Microstructures

Natural vibration frequency of classic MEMS structures

Design and Simulation of A MEMS Based Horseshoe Shaped Low Current Lorentz Deformable Mirror (LCL-DM).

874. The squeeze film effect on micro-electromechanical resonators

Validation of High Displacement Piezoelectric Actuator Finite Element Models

Non-Linear Finite Element Modeling of THUNDER Piezoelectric Actuators

LARGE DISPLACEMENT LINEAR ACTUATOR

MICROELECTROMECHANICAL systems (MEMS) devices

Computers and Mathematics with Applications

An Analytical Electrothermal Model of a 1-D Electrothermal MEMS Micromirror

Design and Analysis of Various Microcantilever Shapes for MEMS Based Sensing

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

Variable Capacitance Accelerometers: Design and Applications

MODELING AND CONTROL OF MEMS MICROMIRROR ARRAYS WITH NONLINEARITIES AND PARAMETRIC UNCERTAINTIES

DESIGN AND OPTIMIZATION OF BULK MICROMACHINED ACCELEROMETER FOR SPACE APPLICATIONS

A Comparison of Pull-in Voltage Calculation Methods for MEMS-Based Electrostatic Actuator Design

Thickness Optimization of a Piezoelectric Converter for Energy Harvesting

Through-Wafer Interconnects for High Degree of Freedom MEMS Deformable Mirrors

MODELING OF T-SHAPED MICROCANTILEVER RESONATORS. Margarita Narducci, Eduard Figueras, Isabel Gràcia, Luis Fonseca, Joaquin Santander, Carles Cané

ANALYSIS AND NUMERICAL MODELLING OF CERAMIC PIEZOELECTRIC BEAM BEHAVIOR UNDER THE EFFECT OF EXTERNAL SOLICITATIONS

Final Project Report. Contract/Grant Title: Development of High-Fill-Factor Large-Aperture Micromirrors for Agile Optical Phased Arrays

Design and Optimization of Piezoelectric Dual-Mode Micro-Mirror

Efficient Multi-Physics Transient Analysis Incorporating Feedback-Dependent Boundary Conditions

Transcription:

A Vertical Electrostatic Actuator with Extended Digital Range via Tailored Topology Yanhang Zhang and Martin L. Dunn Department of Mechanical Engineering University of Colorado at Boulder Boulder, CO 80309 Abstract We describe the design, fabrication, and testing of an electrostatic vertical actuator that exhibits a range of motion that covers the entire initial gap between the actuator and substrate and provides controllable digital output motion. This is obtained by spatially tailoring the electrode arrangement and the stiffness characteristics of the microstructure to control the voltage-deflection characteristics. The concept is based on the electrostatic pull down of bimaterial beams, via a series of electrodes attached to the beams by flexures with tailored stiffness characteristics. The range of travel of the actuator is defined by the post-release deformed shape of the bilayer beams, and can be controlled by a post-release heattreat process combined with a tailored actuator topology (material distribution and geometry, including spatial geometrical patterning of the individual layers of the bilayer beams). Not only does this allow an increase in the range of travel to cover the entire initial gap, but it also permits digital control of the tip of the actuator which can be designed to yield linear displacement - pull in step characteristics. We fabricated these actuators using the MUMPs surface micromachining process, and packaged them in-house. We measured, using an interferometric microscope, full field deformed shapes of the actuator at each pull in step. The measurements compare well with companion simulation results, both qualitatively and quantitatively. Keywords: electrostatic actuator, digital movement, MEMS, range of travel, bilayer beam, pull in step 1. Introduction Actuators are a principal component of a large class of microelectromechanical systems (MEMS). Micromechanical electrostatic actuation is very attractive because of the relative ease of fabrication via thin-film technology. However, electrostatic actuators that are able to generate relatively large vertical displacements (normal to the substrate) are difficult to realize. The controllable range of travel is usually limited to one third to one half of the initial gap. Many applications, especially optical MEMS, 1 would benefit greatly by the availability of vertical actuators that extend the range of motion. Several methods have been proposed to extend this range of motion. Legtenberg et al. 2 described an electrostatic actuator consisting of a laterally compliant cantilever beam and a fixed curved electrode where the beam is bent around the fixed electrode. Chan and Dutton 3 presented a design that employs a series of capacitors to extend the range of travel to 60% of the initial gap. Hung and Senturia 4 described a design with leverage bending. The above efforts have been focused on overcoming the conventional pull in limit; to the best of our knowledge no work has been reported to provide controllable digital output motion. Corresponding author, Phone: (303) 492-2523, Fax: (303) 492-3498, email: yanhang.zhang@colorado.edu. Smart Structures and Materials 2002: Smart Electronics, MEMS, and Nanotechnology, Vijay K. Varadan, Editor, Proceedings of SPIE Vol. 4700 (2002) 2002 SPIE 0277-786X/02/$15.00 147

In this paper we prescribe a unique microstructure that exhibits a range of motion that covers the entire initial gap (between the actuator and substrate) and provides controllable digital output motion. This is obtained by spatially tailoring the electrode arrangement and the stiffness characteristics of the microstructure to control the voltage-deflection characteristics, while at the same time providing digital output. 2. Concept The range of travel of the actuator is defined by the post-release deformed shape of the bilayer beams, and can be controlled by a post-release heat-treat process combined with a tailored actuator topology (material distribution and geometry, including spatial geometrical patterning of the individual layers of the bilayer beams). The electrostatic actuator concept is based on the electrostatic pull down of bimaterial beams, via a series of electrodes attached to the beams by flexures with tailored stiffness characteristics. A schematic diagram of one realization of the actuator is shown in Fig. 1 (a). Two bilayer beams are cantilevered from an anchor above the substrate and are connected to each other by flexures and plates which serve as electrodes. In this design, eight flexures connect four plates to the two bilayer beams; the plates are numbered from one to four from the cantilevered end of the beam. Underneath each plate, a polysilicon electrode is deposited on the substrate to which voltage will be applied for actuation. Of course, the number of plates (electrodes) and their shapes, sizes, and locations are design variables. In the design to be fully described later, the bilayer beams (350 µm long 25 µm wide) consist of gold (0.5 µm thick) and polysilicon (1.5µm) layers. Due to internal stresses induced during fabrication, including those due to thermal expansion mismatch between gold and polysilicon, the bilayer beams bend up from the substrate after release; we refer to this as a positive curvature. The magnitude of this deformation, which we will characterize by either the curvature or tip deflection of the beam can be controlled by a post-release heat-treat process, and the spatial geometrical patterning of the gold film on the polysilicon film, which will be discussed in the next section. This initial deformation provides the initial gap, and thus defines the range of travel of the actuator. The flexures and plates are made of a single layer of polysilicon (1.5 µm thick). The flexures consist of very thin (about 4 µm wide) bending strips. To actuate, we first apply a voltage to the electrode beneath the first plate (that closest to the cantilevered end). At a critical applied voltage the first plate pulls in completely to the substrate and pulls the bilayer beams down via the flexures, decreasing the tip height of the actuator. Similar actuation processes occur when voltages are subsequently applied to the other electrodes in sequence from two to four, and the plates are pulled down sequentially. The result is that four pull-in steps are realized, which result in digital output of the displacement at the tip of the actuator. Figure 1 (b) shows the side view of the actuator at the initial and each pull in steps. Stepwise pull-in allows the range of travel to increase to the entire initial gap of the vertical electrostatic actuator. Bimorph beam Flexure Plate 1 2 3 4 Electrode (a) Figure 1: Digital electrostatic vertical actuator designed to increase the controllable range of motion and to provide a controllable digital movement output: (a) schematic diagram illustrating components; (b) side view of the actuator at the initial and subsequent pull-in steps. (b) 148 Proc. SPIE Vol. 4700

3. Design, Analysis, and Fabrication 3.1 Initial Gap/Range of Travel The range of travel of the actuator is defined by the initial deformation of the bilayer beam and can be controlled by a post-release heat-treat process and the spatial geometrical patterning of the gold film on the polysilicon film. To illustrate, Fig. 2 shows the measurements of curvature vs. temperature for gold (0.5 µm thick) / polysilicon (1.5 µm thick) 300 µm 50 µm beams upon uniform heating and cooling cycles (a heat treat process) over a temperature range from room temperature to 100 C 275 C. 5 This temperature range encompasses the temperature excursions that are expected during many post-processing and packaging steps. During operation there are two key parameters that are strongly impacted by this cyclic thermal behavior: the linearity of the curvature developed as a function of temperature, and the curvature at room temperature (or any other specified temperature) which sets the reference state for the device. As shown in Fig. 2, the linearity can be ensured to a large degree over a certain temperature range by cycling the microstructure to a temperature higher than the maximum intended use temperature. For example, if the microstructure has been cycled to 200 o C, it can be expected to subsequently deform linearly over a temperature range with a maximum temperature less than 200 o C. Although beyond the scope of this study, the complete characterization of the deformationtemperature behavior requires the consideration of the rate of temperature change, and in particular, the length of time the microstructure is held at the elevated temperature. 5, 6 1000 Curvature (m -1 ) 800 600 400 200 Gold (0.5 µm) Polysilicon (1.5 µm) dκ/dt = -5.56 m -1 / o C 0-200 0 50 100 150 200 250 300 Temperature ( o C) Figure 2: Curvature vs. temperature for gold (0.5 µm thick) / poly2 (1.5 µm thick) 340µm 50 µm beam microstructures. The room-temperature deformed shape sets the initial gap of the electrostatic actuator. It can be tailored by a combination of the structural geometry and the heat-treat process parameters. To illustrate, consider the gold/polysilicon beam microstructure in the actuator. We will describe how one can tailor the heat-treat process and the topology (material distribution and geometry) to meet two different initial gap requirements. For the purposes of this discussion we use the data in Figs. 2 and 3. The generalization to other gap requirements follows along the same lines. Proc. SPIE Vol. 4700 149

Assume that we want the initial tip deflection to be 25 µm, which is larger than the tip deflection after release. In order to obtain the desired tip deflection, first let us calculate the curvature required to yield a 25 µm tip deflection for the L = 350 µm gold (0.5 µm thick) / polysilicon (1.5 µm thick) beam. For a beam with uniform curvature, the tip deflection δ is related to the curvature κ simply as δ = ½ κl 2, where L is the length of the beam. For L = 350 µm and δ = 25 µm, κ = 408 m -1. Now from Fig. 2, in order to obtain a room-temperature curvature of 408 m -1, the beam must be heated to about 120 C and cooled to room temperature. After this heat-treat process, the deformation as a function of temperature will then be linear as long as the temperature does not exceed 120 C. Suppose that the actuator must undergo a post-release process such as wire bonding. If the wire bonding temperature is below 120 C, then no additional unwanted deformation will occur during wire bonding. The same argument applies to any other packaging process. Thus, the upper temperature of the heat-treat should be chosen to be above that of any subsequent packaging processes so that the deformation experienced during the packaging processes will be linear thermoelastic and result in no permanent deformation. Suppose, however, that the device application requires that after packaging, the beam have a curvature of 200 m -1 (corresponding to the tip deflection of 12 µm), and that the packaging process exposes the actuator to a temperature of 120 o C. If one simply carried out the process as described above, upon return to room temperature the curvature would be about 400 m -1, rendering the device unusable. The challenge is then to design the geometry, for example the layout of the gold film on the polysilicon, so that the room-temperature curvature after packaging is still acceptable. In the present example this could be accomplished by covering only part of the polysilicon beam with gold. The results of such an approach are shown in Fig. 3 where the thermoelastic slope dκ/dt is plotted as a function of the fraction of the beam that is covered with gold. Both analytical predictions and measurements are shown. In principle, one could use the results of Figs. 2 and 3 to design the shape of the cantilever. For the example here, suppose we want to ensure linear thermoelastic deformation between 140 o C and room temperature, so the temperature change is 25 o C 140 o C = -115 o C. Thus dκ/dt = 200 m -1 /115 o C 1.74 m -1 / o C. According to Fig. 3, the coverage of gold will be approximately 20% of the polysilicon. 6 5 dκ/dt (m -1 / o C) 4 3 2 y 1 w g x wp 0 0 0.2 0.4 0.6 0.8 1 w g / w p Figure 3: Slope of the curvature vs. temperature change as a function of gold film width for a strip patterned beam. Filled circles are measurements, open circles are finite element calculations, and the solid line is an analytical estimate 5 based on linear plate theory. 150 Proc. SPIE Vol. 4700

3.2 Electrostatic Actuation Consider a conventional cantilever beam electrostatic actuator. The basic idea is that a potential difference is applied across the gap between the bottom of the beam and an electrode on a substrate beneath it. This generates an electrostatic force that pulls the beam toward the substrate. The actuator characteristics (voltage vs. deflection) are highly nonlinear due to the competing effects of the nonlinear electrostatic forces and the linear mechanical restoring forces. The former arise because the electrostatic force acting on the beam varies as 1/δ(x) 2 where δ(x) is the distance between the substrate and the point on the beam at the lateral position x. As the beam deforms, δ(x) decreases and the electrostatic force increases nonlinearly. At some critical applied voltage the system becomes unstable and the beam pulls in completely to the substrate. Almost all electrostatic actuators of this type have been designed based on simple perturbations of the simple cantilever. In addition to the highly nonlinear voltage-deflection characteristics, the actuator is analog in nature. Our objective is to achieve a step-wise pull in, leaving the tip up and pulling in the actuator from the fixed end. This is not possible for a single solid beam. Our approach to achieve such pull-in characteristics is to balance the nonlinear electrostatic forces with nonlinear mechanical restoring forces by spatially tailoring the distribution of both the structural material (and thus stiffness) and the electrodes. With our actuator design, both the restoring electrostatic and mechanical forces are impacted by the design variables. The electrostatic forces are modified by tailoring the distribution of the electrodes and the overlap area between the plates and electrodes. The mechanical restoring forces are modified by the tailored topology, i.e., the flexure geometry and layout, and the geometry of the bilayer beams. To illustrate how the stiffness/flexibility are modified, a simplified design is shown in Fig. 4 (a), where each plate is connected via two flexures to the bilayer beams. We assume that for each plate, the spring constant is only related to the flexures connected to it and the bilayer beams. The electrostatic force acting on the plate is transferred to the two flexures connected to the plate, and via the flexures, the electrostatic force is transferred to the cantilever bilayer beams as an intermediate load. A schematic diagram is shown in Fig. 4 (b), where F i (i = 1, 2, 3, and 4) represents the electrostatic force acting on the plate that is transferred to the flexures and the bilayer beams. Therefore for each plate, a spring constant K equiv can be calculated as shown in Fig. 4(b) where K flexure represents the spring constant of the flexure and K beam represents the spring constant of the bilayer beam with intermediate load. K flexure will vary for different flexure designs, and K beam will vary as the intermediate load position changes, i.e., as the beam is pulled in. Polysilicon plate F 1 Bimorph beam F 2 F 3 F 4 Polysilicon flexure F 1 F 2 F 3 F 4 1 2 3 4 Bimorph beam K beam K flecture K beam K flecture equiv K = 2 K (a) (b) Figure 4: (a) Top view of the actuator design; (b) schematic showing the dominant stiffness characteristics of the actuator. K beam beam K + K flexure flexure 3.3 Other Features of the Actuator Besides the features discussed in section 3.1 and 3.2, the actuator presented in this paper has the following additional features: Proc. SPIE Vol. 4700 151

As the plates are sequentially pulled down, they pull down the bilayer beam via the flexures and thus decrease the tip deflection. Without this tailored topology, it is impossible to achieve the stepwise pull in for a solid beam. By simply modifying the stiffness of the bilayer beam, for example, by making some part of the beam flexible, stepwise pull in can be achieved to some degree, but the controllability is very limited and the freedom to modify the tip response characteristics is compromised. The topology can be tailored to yield desired digital output characteristics. For example, the topology can be tailored to yield a linear tip displacement voltage relationship. Once the tip displacement response characteristics have been identified for a particular device or grouping of components based upon the overall MEMS design criteria, an actuator can be constructed to achieve those characteristics by spatially tailoring the topology, i.e., the material geometry and layout. Although the microstructures described so far were designed to have four pull-in steps, the number of the pull in steps, which is determined by the real application, can be easily changed. 3.4 Simulated Response We simulated the coupled electromechanical behavior of the actuator using a fully coupled simulator (MEMCAD 4.8) that couples boundary element electrostatics and finite element structural mechanics. Both the gold and polysilicon are modeled as linear thermoelastic with isotropic material properties. Input parameters are E p =163 GPa, ν p =0.22 (in line with many measurements over many MUMPS runs), 7 E g =78 GPa, ν g =0.42. 8 The thermal expansion coefficients of the materials are assumed constant and their average values over the temperature range considered are α p =2.8 10-6 / o C, and α g =14.4 10-6 / o C. 8 Here and throughout the subscripts p and g denote polysilicon and gold, respectively. Typical finite element meshes for the actuator microstructures contained elements with a characteristic dimension of about 10 µm, a size that was chosen after a careful convergence study with mesh size. The complete calculations consist of two parts. First, the thermomechanical deformation with uniform temperature change is computed using the MemMech solver. The second step of the calculations is the electrostatic pull in using the CoSolve_EM solver with the voltage applied to the appropriate electrodes at each pull in step. The starting point for the second step is the end point of the first step. Results from simulations as just described were used to tailor the layout and geometry to yield specified actuator deformation response characteristics. Specifically, we attempted to obtain a linear pull-in vs. tip displacement response. 3.5 Fabrication We fabricated actuator designed as described above using the commercially available Multi-User MEMS Process (MUMPs). 9 Specifically they are fabricated using the MUMPs 41 surface micromachining process, and then packaged them in-house. The maximum temperature experienced by the actuator during packaging was about 75 C. We then measured, using an interferometric microscope, full field deformed shapes of the actuator at each pull in step. Details regarding the measurements are described elsewhere. 10 4.1 Single Actuator Performance 4. Results and Discussion Figure 5 shows a scanning electron image of an actuator designed to provide linear tip deflection pull-in behavior. Figure 6 shows the measured full field deformed shapes of the actuator at the initial position after packaging, and at each 152 Proc. SPIE Vol. 4700

subsequent pull in step. The measured pull in voltages for each pull in step are 46.6v, 38.2v, 9.4v, and 11.2v. Compared with the simulation, the measured pull in voltages are about 10 to 15 volts lower. 100 µm Figure 5: SEM image of the electrostatic actuator designed to provide linear digital of tip deflection vs. pull-in step characteristics. (a) (b) (c) (d) Figure 6: Measured deformed shapes of the actuator at the (a) initial, (b) first, (c) second, (d) third, and (e) fourth pull-in step, respectively. (e) From the measurements of the full-field deformed shapes, we determined the tip deflection relative to the fixed end; the results are plotted in Fig. 7 vs. the pull-in step. Figure 7 also shows the initial tip deflection and that at the first three pull in steps from the simulation. The measurements compare well with the companion simulation results. The tip deflection Proc. SPIE Vol. 4700 153

at the fourth pull in step is not plotted in Fig. 7 because it was not designed to follow the linear relationship exhibited during the previous steps. A linear fit to the measurements shows the excellent linearity of the response. 20 Tip deflection (µm) 15 10 5 0 Measurement Simulation Linear fit to measurement 0 1 2 3 Pull in steps Figure 7: Tip deflection vs. pull-in steps from both the measurements and simulations of the actuator shown in Figure 5. In Fig. 8 we plot the surface profile relative to the height of the fixed end along one of the bilayer beams at the initial, fist, second, and third pull in steps. The solid lines are measurements and the open circles are simulation result. The two significant bumps in each experiment result are due to the conformal surface micromachining process, where two wires connected to two electrodes lie beneath the bilayer beam. This can be seen clearly in both Figs.5 and 6. At the initial position, the bilayer beam deforms due to the internal stresses only, therefore the curvature is uniform along the beam. When the first plate (that closest to the cantilever end) is pulled down, the bilayer beam deforms due to the combination of the internal stress and the electrostatic force. As a result, the curvature is not uniform over the part of the beam extending from the fixed end to the first electrode; that part of the beam, about 45µm from the fixed end, is relatively flat. Similarly, in the second and third pull in steps, the length of the relatively flat part increases to about 120 µm and 190µm, respectively. For all pull-in steps, the measurements agree well with the simulation results. 154 Proc. SPIE Vol. 4700

20 15 Deflection (µm) 10 5 0-5 0 50 100 150 200 250 300 350 Position along the beam (µm) Figure 8: Surface profile along one bilayer beam from both the experiment and simulation results at the initial, first, second, and third pull-in steps from top to bottom, respectively. Solid lines are measurements and open circles are simulation results. 4.2 Actuator Assemblies that Provide 2-D and 3-D Adjustments Micromirror Applications The controllable and digital character of this actuator offers the ability to operate it in an open-loop control mode, a significant advantage, particularly in the common situation where arrays of multiple actuators exist. For example, this digital electrostatic vertical actuator can be combined with a mirror or a mirror array with applications in optical switching. Work along these lines is underway and will be presented elsewhere. 5. Summary We designed and tested an electrostatic vertical actuator that exhibits a range of motion that covers the entire initial gap between the actuator and substrate and provides controllable digital output motion. The range of travel of the actuator is defined by the initial deformation of the bilayer beam, and can be controlled by a post-release heat-treat process, and the spatial geometrical patterning of the gold film on the polysilicon film. Not only is the range of travel increased to the entire initial gap of the vertical electrostatic actuator, but also the movement of the tip is controlled to yield a linear displacement - pull in step relationship by modifying the actuator topology (material distribution and geometry). We fabricated these actuators using the MUMPs surface micromachining process, and packaged them in-house. We measured full field deformed shapes of the actuator at each pull in step. The measurements compare well with the companion simulation results, both qualitatively and quantitatively. Proc. SPIE Vol. 4700 155

6. Acknowledgments We gratefully acknowledge support by the Defense Advanced Research Projects Agency (DARPA) and the Air Force Research Laboratory, Air Force Materiel Command, USAF, under agreement number F30602-98-1-0219, and the Air Force Office of Scientific Research. 7. References 1. Helmbrecht, M. A., Srinivasan, U., Rembe, C., Howe, R. T., and Muller, R. S., Micromirrors for Adaptive-Optics Arrays, Transducers 01, Germany, June 2001. 2. Legtenberg, R., Gilbert, J., Senturia, S. D., and Elwenspoek, M., 1997, Electrostatic Curved Electrode Actuators, Journal of Microelectromechanical Systems, v. 6, pp. 257-265. 3. Chan, E. K., and Dutton, R. W., 2000, Electrostatic Micromechanical Actuator with Extended Range of Travel, Journal of Microelectromechanical Systems, v. 9, pp. 321-328. 4. Hung, E. S., and Senturia, S. D., 1998, Leveraged Bending for Full-Gap Positioning with Electrostatic Actuation, Solid State Sensor and Actuator Workshop, Hilton Head Island, South Carolina, June, pp. 83-86. 5. Zhang, Y. and Dunn, M. L., 2002, Deformation of Multilayer Thin Film Microstructures During Post-Release Cyclic Thermal Loading: Implications for MEMS Design, Packaging, and Reliability, submitted for publication. 6. Zhang, Y. and Dunn, M.L., 2001, Stress Relaxation of Gold/Polysilicon Layered MEMS Microstructures Subjected to Thermal Loading, in Proceedings of the MEMS Symposium, ASME International Mechanical Engineering Congress and Exposition. 7. Sharpe, W. N., 2001, Mechanical Properties of MEMS Materials, a chapter in the CRC Handbook of MEMS, M. Gad-el-Hak, ed., in press. 8. King, J. A., 1988, Materials Handbook for Hybrid Microelectronics, Teledyne Microelectronics, Los Angeles, California. 9. Koester, D. A., Mahadevan, R., Hardy, B., and Markus, K. W., 2001, MUMPs Design Rules, Cronos Integrated Microsystems, A JDS Uniphase Company, http://www.memsrus.com/cronos/svcsrules.html. 10. Dunn, M. L., Zhang, Y., and Bright, V. M., 2002, Deformation and Structural Stability of Layered Plate Microstructures Subject to Thermal Loading, Journal of Microelectromechanical Systems, in press. 156 Proc. SPIE Vol. 4700