Structural mechanics necessitates the development of the following concepts to obtain a basic understanding of the subject for purposes of MEMS design: • Structural material models • Mod
Trang 111 T.G Szymanski, C.J Van Wyk, Space efficient algorithms for VLSI artwork
analysis, Proc 20th Design Automation Conf., 734–739, June 1983.
12 H.S Baird, Fast algorithms for LSI artwork analysis, Proc 14th Design
Automa-tion Conf., 303–311, 1977.
13 http://www.memscap.com/ memsrus/crmumps.html.
14 R.A Buser, N.F de Rooij, ASEP: a CAD program for silicon anisotropic etching
(micromechanical structure), Sensors Actuators A, A28(1), 71–78, 1991.
15 C.H Sequin, Computer simulation of anisotropic etching, Proc 6th Int Conf.
Solid-State Sensors Actuators (Transducers ’91), San Francisco, CA, 28–28 June,
1991, 801–806.
16 T.J Hubbard, E.K Antonsson, Emergent faces in crystal etching, J
Microelectro-mech, Syst., 3, 19–28, 1994.
17 J Fruhauf, K Trautmann, J Wittig, D Zielke, A simulation tool for orientation
dependent etching, J Micromech Microeng., 3(3), 113–115, 1993.
18 D Dietrich, J Fruhauf, Computer simulation of the development of dish-shaped
deepenings by orientation-dependent etching of (100) silicon, Sensors Actuators
A, A39(3), 261–262, 1993.
19 U Heim, A new approach for the determination of the shape of etched devices,
J Micromech Microeng., 3(3), 116–117, 1993.
20 D Zielke, J Fruhauf, Determination of rates for orientation-dependent etching,
Sensors Actuators A, A48(a), 151–156, 1995.
21 H Camon, A Moktadir, M Djafari-Rouhani, New trends in atomic scale
simu-lation of wet chemical etching of silicon with KOH, Mater Sci Eng B, B37(1–3),
23 S Yamamato, T Kure, M Ohgo, T Matsuzama, S Tachi, H Sunami, A
two-dimensional etching profile simulator: ESPRIT, IEEE Trans Computer-Aided
Design, CAD-6(3), 417–422, 1987.
24 J Pelka, K.P Muller, H Mader, Simulation of dry etch processes by
COMPOS-ITE, IEEE Trans Computer-Aided Design, 7(2), 154–159, 1988.
25 C Hedlund, C Strandman, I.V Katardjiev, Y Backlund, S Berg, H.O Blom,
Method for the determination of the angular dependence during dry etching, J.
Vac Sci Tech B, 14(5), 3239–3243, 1996.
26 M Fujinaga, N Kotani, 3-D topography simulator (3-D MULSS) based on a
physical description of material topography, IEEE Trans Electron Devices, 44,
226–238, 1997.
27 G.M Koppelman, OYSTER, a three-dimensional structural simulator for
micro-electromechanical design, Sensors Actuators, 20(1/2), 179–185, 1989.
28 P.M Osterberg, S.D Senturia, Membuilder: an automated 3D solid-model
con-struction program for microelectromechanical structures, Proc Transducers ’95,
Stockholm, Sweden, June 1995, 2, 21–24.
29 N.R Lo, K.S.J Pister, 3D µV — a MEMS 3-D visualization package, Proc SPIE
Micromachined Devices Components, Austin, TX, Oct 23–24, 1995, 290–295.
30 C.R Jorgensen, V.R Yarberry, A 3D geometry modeler for the SUMMiT V MEMS
designer, Proc Modeling Simulation Microsyst Conf., 606–609, 2001.
Trang 231 C.R Jorgensen, V.R Yarberry, A 2D visualization tool for SUMMiT V designs,
Proc Modeling Simulation Microsyst Conf., 594–597, 2001
32 S.D Senturia, R.M Harris, B.P Johnson, S Kim, K Nabors, M.A Shulman, J.K White, A computer-aided design system for microelectromechanical systems
(MEMCAD), J Microelectromech Syst., 1(1), 3–13, 1992.
33 J.R Gilbert, P.M Osterberg, R.M Harris, D.O Ouma, X Cai, A Pfajfer, J White, S.D Senturia, Implementation of MEMCAD system for electrostatic and mechan-
ical analysis of complex structures from mask descriptions, Proc IEEE Micro
Electro Mechanical Syst., Fort Lauderdale, FL, Feb 7–10, 1993, 207–212.
34 Coventer: http://www.coventor.com.
35 S.D Senturia, Simulation and design of microsystems: a 10-year perspective,
Sensors Actuators A, 67, 1–7, 1998.
36 S.D Senturia, CAD challenges for microsensors, microactuators, and
microsys-tems, Proc IEEE, 86(8), 1611–1626, 1998.
37 S.D Senturia, N Aluru, J White, Simulating the behavior of MEMS devices:
computational methods and needs, IEEE Computational Sci Eng., 30–43,
Jan.–Mar., 1997.
38 R.M Kirby, G.E Karniadakis, O Mikulchenko, K Mayaram, An integrated
simulator for coupled domain problems in MEMS, J Microelectromech Syst.,
10(3), 379–391, 2001.
39 SUGAR: http://bsac.berkeley.edu/ cadtools/sugar/sugar/#overview.
40 J.V Clark, D Bindel, W Kao, E Zhu, A Kuo, N Zhou, J Nie, J Demmel, Z Bai, S Govindjee, K.S.J Pister, M Gu, A Agogino, Addressing the needs of
complex MEMS design, Proc MEMS 2002, Las Vegas, January 20–24, 2002.
41 J.V Clark, N Zhou, K.S.J Pister, Modified nodal analysis for MEMS with
multi-energy domains, Int Conf Modeling Simulation Microsyst., Semiconductors,
Sen-sors Actuators, San Diego, CA, March 27–29, 2000, 31–34.
42 R.E Rudd, The atomic limit of finite element modeling in MEMS: coupling of
length scales, Analog Integrated Circuits Signal Process., 29, 17–26, 2001.
43 MEMSCAP: http://www.memscap.com.
44 ANSYS: http://www.ansys.com.
45 V Venkataraman, R Sarma, S Ananthasuresh, Part to art: basis for a systematic
geometric design tool for surface micromachined MEMS, Proc ASME Design
Eng Tech Conf., 1–14, 2000.
46 A Perrin, V Ananthakrishnan, F Gao, R Sarma, G.K Ananthasuresh,
Voxel-based hetrogeneous geometric modeling for surface micromachined MEMS, Proc.
Modeling Simulation Microsys Conf., 136–139, 2001.
47 R Schiek, R Schmidt, A new, topology driven method for automatic mask
generation from three-dimensional models, NanoTech 2003, San Francisco, CA,
Feb 2003.
Trang 3MEMS devices invariably involve engineering of multiphysics designs to attain
a design objective The two physical domains most frequently utilized in MEMSdevices are structural and electrical dynamics Regardless of the design objec-tive, a structure invariably needs to be designed to support, contain, or possiblydeflect to perform a function An electrical system is needed to sense themechanical motion of the structure At the microscale, damping due to viscouslosses of the device to the surrounding atmosphere greatly influences the dynam-ics of the system
For example, a MEMS accelerometer requires the suspended seismic mass
to have a preferred mode of vibration in the sensitive axis at a specific resonantfrequency This device would also have an electrical sense interface to transducethe motion of the seismic mass and, possibly, electrical force feedback to maintainthe position of the accelerometer sense mass at a neutral position The damping
of the accelerometer seismic mass will greatly influence the dynamics of thesystem and needs to be considered in the design
This chapter will present an overview of the important topics in structuralmechanics, damping, and electrical circuit elements Due to space limitations, anin-depth treatment of these topics is not possible; however, the topics relevant tothe design of MEMS devices will be presented References 1 through 4 provide
a more complete background in structural mechanics
Structural mechanics necessitates the development of the following concepts
to obtain a basic understanding of the subject for purposes of MEMS design:
• Structural material models
• Models of the basic structural elements (bending, torsion, axial rods,columns)
• Combining the basic structural elements
Damping mechanisms for vertical and laterally moving MEMS devices will
be presented The basic electrical circuit elements and models for them will bepresented along with methods for combining them to form a circuit The set ofequations that describe the electrical circuit elements will also be developed
Trang 46.1 STRUCTURAL MECHANICS
6.1.1 MATERIAL MODELS
The atomic structure of materials — broadly classified as crystalline,
polycrys-talline , and amorphous — is illustrated in Figure 2.2. A crystalline material has
a large-scale, three-dimensional atomic structure in which the atoms occupyspecific locations within a lattice (e.g., epitaxial silicon, diamond) The atomicpacking may be in one of seven main crystal patterns with orientations measured
via the Miller indices (discussed in Chapter 2)
A polycrystalline material consists of a matrix of grains, which are small
crystals of material; the interface material between adjacent grains is called thegrain boundary Most metals, such as aluminum and gold as well as polycrystal-line silicon, are examples of this material structure A noncrystalline material that
exhibits no large-scale structure is called amorphous Silicon dioxide and other
glasses are examples of this material structure
The material type greatly influences fundamental structure and completeness
of interatomic bonds This basic material structure affects a number of materialproperties, such as the electrical and thermal conductivities, chemical reactivity,and mechanical strength For example, the metallurgical processes of cold work-ing and annealing greatly affect the material grains and grain boundary and theresulting material properties of strength, hardness, ductility
The characteristics of a material that first come to mind in connection withthe design of a structure are strength, elasticity, and ductility These characteristicsrelate to the ability of the material to resist mechanical forces and how the materialwill fail In order to establish a meaningful way to design with these consider-ations, it is necessary first to define some commonly used engineering terms.Given a bar of material loaded with a uniform force distribution across the
cross-sectional area, A, as shown in Figure 6.1, a quantity, stress σ, is defined as
the total force, F, per unit cross-sectional area A (Equation 6.1) The applied load
will deform the material, which will require the definition of a metric to describethe extent of deformation The metric for localized deformation of a material,
strain, is a dimensionless quantity defined as the change in length, δ, per length,
L (Equation 6.2)
(6.1)(6.2)
When an experiment is performed on the specimen of Figure 6.1 in whichthe load is increased in a controlled manner, stress vs strain can be plotted (Figure6.2) The material shown in this figure exhibits elastic strain The material deformsunder load as indicated by strain, but the deformation is not permanent Whenthe load is removed, the stress and strain return to zero
σ = F A/
ε δ= / L
Trang 5If the load on the material is increased further, the material will plasticallydeform or fail abruptly Figure 6.2 shows a material that deforms elastically until
it abruptly fails The stress at failure is known as the ultimate strength of the material, S u This type of material failure is known as brittle Figure 6.3 shows
a material that deforms elastically until the material yields at a stress known as
the yield stress, S y This is the elastic limit of the material Increasing the stress (by increasing the load) beyond the S y will induce plastic strain, which is a
permanent deformation of the material Unloading a material that has been
stressed beyond S y will cause a different path to be followed on the stress–straincurve upon unloading
When the material is unloaded, a permanent deformation has been induced
in the material as shown by a nonzero deformation existing at zero load If thestress in the material (load on the specimen) is increased past the yield stressuntil the material eventually fails, the stress at failure is the ultimate strength of
the material S u The shape of the stress–strain curve for different ductile materialsstressed beyond the elastic limit can vary due to large changes in the materialcross-section during plastic deformation Some material will exhibit a distinct
FIGURE 6.1 Loaded material specimen.
FIGURE 6.2 Elastic stress–strain relationship.
Trang 6change in slope or distinct plastic deformation at the yield point, but others will
be more subtle When the yield point is not distinct (Figure 6.4), the yield point
is generally defined as that stress, which induces 0.2% (0.002) plastic strain Most engineering applications will not intentionally stress a material past theyield strength The system will be designed to operate within the elastic region
of the material The slope of the elastic region of the stress–strain curve is a
widely used engineering property of a material known as Young’s modulus, E,
which has units of force per area and is a measure of material stiffness Appendix
E lists typical values of Young’s modulus for a number of materials frequentlyused in MEMS devices
A frequently used material model for operation within the elastic region of
a material is Hooke’s law, which states that the stress in a material is proportional
to the strain that produced it This is merely the mathematical relationship forthe material operating within the elastic portion of the stress–strain curve:
(6.3)
FIGURE 6.3 Plastic stress–strain relationship.
FIGURE 6.4 Plastic stress–strain relationship.
σ = Eε
Trang 7The discussion thus far has centered on a material loaded normally to thecross-section of the bar, as shown in Figure 6.1 The load could be in tension orcompression Alternatively, a material could be loaded in shear (Figure 6.5) Inthis case, the load is in the plane of the loaded cross-section Shear stress, τ, isdefined as the load divided by the cross-sectional area, which is similar to thedefinition for normal stress, σ However, shear strain, γ, is defined as the change
in angle of a unit cube of the material shown in Figure 6.5 The development ofshear stress and shear strain is similar to that presented for normal stress loading.Hooke’s law for shear loading is shown in Equation 6.4 The constant of propor-
tionality, G, is known as the modulus of rigidity or the shear modulus E and G
represent fundamental properties of a material, and they have units of force per
area squared E and G are measures of the stiffness or rigidity of a material for
normal and shear loading, respectively
(6.4)
It has also been observed that a material placed in tension also exhibits lateralstrain in addition to axial strain Poisson demonstrated that these two strains areproportional to each other within the elastic region modeled by Hooke’s law The
proportionality constant is known as Poisson’s ratio, ν (Equation 6.5) The son ratio is dimensionless and typically has a value between 0 and 0.5 A solidwith ν = 0.5 does not undergo a change volume when strained uniaxially Forexample, rubber is a material with ν = 0.5 The common situation for most solid
Pois-FIGURE 6.5 Planar unit element of material loaded with normal and shear stress.
(a) Normal Stresses and Strain
(b) Shear Stresses and Strain
τ= Gγ
Trang 8materials is for the volume to expand under uniaxial loading, which corresponds
(6.7)
However, for crystalline materials, the material properties will frequently be
a function of the spatial orientation An orthotropic material has three planes ofmaterial property symmetry To describe this spatial material property, dependencyfor an orthotropic material requires nine independent material properties — anincrease over the three independent material properties required for an isotropic
material There will be a Young’s modulus for each axis (E x , E y , E z); modulus of
rigidity for the three shear planes (G xy , G yz , G xz); and a Poisson ratio (νxy, νyz, νxz)for each axis Equation 6.8 shows the orthotropic stress–strain relations
ν = − lateral strainaxial strain
E 2G= ( )1+ν
εεεγγγ
ν
x y z xy yz zx
σστττ
Trang 9An anisotropic material is the most general material that requires 21 materialproperties to model its behavior Equation 6.9 shows the stress–strain relationsthat would model an anisotropic material Isotropic and orthotropic materialmodels are special cases of an anisotropic material model
(6.9)
FIGURE 6.6 Unit cube with three-dimensional stresses.
εεεγγγ
ν
x y z xy yz zx
yz z xz
x y z xy yyz zx
x y z xy yz zx
στττ
Trang 106.1.2 THERMAL STRAINS
When the temperature of an unconstrained elastic member is increased, themember expands in all directions The normal strain produced in the material is
called thermal strain and is proportional to the temperature increase, ∆T, shown
in Equation 6.10 The proportionality constant, α, is a material property called
the coefficient of thermal expansion [8] Appendix E gives representative valuesfor the coefficient of thermal expansion of a number of commonly used MEMSmaterials If the elastic member is unconstrained, the temperature increase pro-duces thermal strain; however, no stress is induced in the material
(6.10)
However, if a uniform rod is constrained at each end so that the material cannotexpand when subjected to a temperature increase, a compressive stress (Equation6.11) will be induced because of the constraint
(6.11)
Thermal strains can also be developed in devices incorporating materials withdifferent coefficients of thermal expansion For example, a beam with aluminumdeposited on silicon, as shown in Figure 6.7, will flex out of plane when exposed
to a uniform temperature increase due to the different coefficients of thermalexpansion of the materials
In the discussion thus far, only the case in which a uniform temperatureincrease in a material produces a thermal strain has been considered Anothercommon and interesting situation is produced when temperature gradients due
to nonuniform temperature distributions in the material exist Temperature dients can be due to thermal transients and nonuniform heat generation or heat
gra-deposition within the material A thermal stress is developed due to a temperature
gradient in a body
during transient heat transfer The heat flux on the bottom surface of the substrate
FIGURE 6.7 Aluminum–silicon cantilever beam.
Al Al= 25 × 10-6 1/Co
Si Si= 3 × 10-6 1/Co
εx = = = ∆εy εz α T
σ ε= E=( )α∆T E
Trang 11produces a temperature gradient; the bottom surface is the hottest The ature gradient causes the material at different depths and temperature to expanddifferently, thus resulting in a thermal strain that causes out-of-plane deflections
temper-of the substrate
The effect of thermally induced stresses on precision MEMS sensors willhave a direct impact on performance and thus needs to be considered in the systemdesign Thermal stress is also utilized as a MEMS actuation mechanism (discussed
in Chapter 8)
6.1.3 AXIAL ROD
The axial rod shown in Figure 6.1, which was discussed during the development
of material models, is a fundamental structural element A rod is an idealized dimensional structural element subjected only to axial forces, which can be tensile
one-or compressive Figure 6.9 is a schematic of the displacement of an element of
material in the rod In this schematic, the element of material, dx, has an applied force, F Due to this loading, the material at position x undergoes a displacement
FIGURE 6.8 Thermal stress during a transient heat transfer.
FIGURE 6.9 Loaded one-dimensional axial rod material element.
substrate
deflected
shape
MEMS device
Transient Heat Flux - Q
∂
∂+
x
u u
∂
∂ +
ρ, A
dx
Trang 12u , and the material at position x + dx undergoes a displacement Itcan be seen that the element of material has changed length by the amount
Therefore, the strain is , which is a mathematical partial
deriv-ative expression for strain (i.e., change in length ∂u per length ∂x) Using Hooke’s
law, Equation 6.12 is obtained upon substituting for σ and ε
(6.12)
Differentiating F with respect to x yields Equation 6.13, which is an
expres-sion for the rate of change of force in the material
(6.13)
Newton’s law of motion is now applied for the element and the net elasticforce in the material element is equated with the inertial force of the materialelement, where ρ is the density of the rod:
(6.14)
Substituting for yields the one-dimensional wave in Equation 6.15,which is a common governing equation for a number of physical phenomena inaddition to axial vibration of a rod (e.g., string under tension, acoustics):
(6.15)
The quantity c is the velocity of propagation of displacement or stress in a
rod This metric includes information regarding the stiffness of the material aswell as its density
(6.16)
x dx+ ∂∂
2 2
net elastic force mass acceleration
2 2
ux
1c
ut
c= Eρ
Trang 13The stiffness coefficient, K, is another useful design metric that can be
obtained for the axial rod Using Equation 6.12 and recognizing that the strain
is the change in length, δ, divided by the original length, L, yields the stiffness coefficient, K The stiffness coefficient for the axial rod is the ratio of the force per unit deflection and has units of force divided by length Whereas E and G were metrics for stiffness of a particular material, K is a measure of the stiffness
of a particular structural element with a specified loading situation
(6.17)
6.1.4 TORSION ROD
Another common structural element involved in structural design, torsion rodsmay be used as flexures to couple or suspend structures The torsion rod is alength of material loaded with an applied torque that will produce an angulardisplacement, θ, and shear stress, τ, in the structural element (Figure 6.10) Thetheoretical development for torsion bars involves the following assumptions:
• The torsion bar is straight and of uniform circular cross-section (solid
or concentrically hollow)
• The torsion bar is loaded by and opposite torques
• The torsion bar is not stressed beyond it elastic limit
The torsion bar twists as torque is applied where plane cross-sections remainplane and radii remain straight Torsion produces a shear stress at any point inthe cross-section, which is a maximum at the outer surface; the shear stress isproportional to the distance from the center Equation 6.18 defines the shear stress
due to torsion as a function of the distance from the center of the torsion bar, r.
J is the area polar moment of inertia The shear stress is a maximum at the outer
surface, r = R (Equation 6.19) This fact explains why many torsion drive shafts
are annular tubes The material near the shaft center carries little stress in pure
torsion loading J for circular cross-sections is given in Appendix G, Table A.G.1
FIGURE 6.10 Torsion rod.
L
= =δ
Trang 14(6.19)
The development of the governing equation for the torsion rod is similar tothat for the axial rod Once again, the governing equation is the one-dimensional
wave equation (Equation 6.20) involving a speed of propagation, c, for torsional
displacement and stress (Equation 6.21) A stiffness coefficient for a torsion rodthat relates the applied torque to the angular displacement can also be developed
(Equation 6.22) The stiffness coefficient, K, has units of torque per radian These equations involve the coefficient of rigidity, G, and material density, ρ, as well
as the length of the torsion rod, L.
(6.20)
(6.21)
(6.22)where,
(6.23)
MEMS applications rarely encounter circular cross-sections Rectangular andsquare cross-sections are more common due to the restrictions of MEMS fabri-cation techniques When a torsion load is applied to a noncircular cross-section,some basic assumptions for torsion are violated The cross-sections becomewarped; the greatest stress occurs at a point on the perimeter nearest the axis oftwist and the corners of the rectangular and square cross-sections have no stress.The analysis of torsion for these cross-sections becomes complex and has beenstudied for years [4,5] Table A.G.1 in Appendix G gives values for the torsional
constant, J, that account for the effects of noncircular cross-sections; these can
be used in calculations for torsional stiffness
τ = TrJτmax = TRJ
=
Trang 156.1.5 BEAM BENDING
Lateral beam bending is a one-dimensional element in which the loading isperpendicular to the axis of the beam The loading may be distributed along thelength of the beam or concentrated at a specific location; it can also be a com-bination of these situations A basic formulation of lateral beam bending is called
the Euler–Bernouli beam The assumptions involved in the development of the
Euler–Bernouli beam model are:
• The material is homogenous and isotropic and obeys Hooke’s law
• The beam is initially straight with a constant cross-section
• The beam is subjected to pure bending (i.e., no torsion or axial loads)
• Plane cross-sections (y–z plane) remain plane during bending
• The beam has an axis of symmetry
The lateral deflection of a Euler–Bernouli beam, shown in Figure 6.11, is
due to a bending moment, M, that bends the beam in a curve with a radius of
curvature, ρ The radius of curvature is assumed to be related to the bending
moment, as shown in Equation 6.24, where E is Young’s modulus and I is the
area moment of inertia about the axis of bending (z axis in Figure 6.11a) The
EI product is the proportionality constant in Equation 6.24, which is frequently
called the beam cross-section stiffness The stiffness of the entire beam will
involve other information, such as the beam length and the beam end conditions.The radius of curvature, ρ, is also shown in Equation 6.24 to be approximately
equal to second derivative of y with respect to x This approximation used in the
FIGURE 6.11 Beam bending.
M
M
(a) Beam loading
Neutral axis
− + Compressive
Tensile