/home/journal/dvi/SMO531-of-SMO.dviTopology optimization of
electrostatically actuated microsystems
Received: 11 July 2004 / Revised manuscript received: 15 March 2005
/ Published online: 28 July 2005 Springer-Verlag 2005
Abstract This study addresses the design of electrostati- cally
actuated microelectromechanical systems by topology optimization.
The layout of the structure and the elec- trode are simultaneously
optimized. A novel, continuous, material-based description of the
interface between the structural and electrostatic domains is
presented that allows the optimization of the interface topology.
The resulting top- ology optimization problem is solved by a
gradient-based algorithm. The electromechanical system response is
deter- mined by a coupled high-fidelity finite element model and a
staggered solution procedure. An adjoint formulation of the coupled
electromechanical design sensitivity analysis is introduced, and
the global sensitivity equations are solved by a staggered method.
The proposed topology optimization method is applied to the design
of mechanisms. The opti- mization results show the significant
advantages of varying the interface topology and the layout of the
electrode ver- sus conventional approaches optimizing the
structural layout only.
Keywords Microelectromechanical systems · Topology optimization ·
Electrostatic-mechanical coupling · Sensitiv- ity analysis ·
Adjoint formulation
1 Introduction
Since the manufacture of the first polysilicon surface mi-
cromachined devices by Bustillo et al. (1998) in the early 1980s,
microelectromechanical systems (MEMS) have be- come essential
components of numerous applications, such
M. Raulli Villanova University, Dept. of Mechanical Engineering,
800 Lancaster Avenue, Villanova, PA 19085, USA E-mail:
[email protected]
K. Maute (B) University of Colorado, Dept. of Aerospace Engineering
Sciences, Campus Box 429, Boulder, CO 80309, USA E-mail:
[email protected]
as medical devices, automobiles, locking systems, and com-
munications devices, among others. MEMS are used primar- ily for
sensing and actuation purposes and allow for minia- turizing at low
cost.
The design of MEMS is a challenging task since the fun- damental
actuation and sensing mechanisms are often based on the interaction
of multiphysics phenomena, such as elec- trothermomechanical
coupling and electrostatic-mechanical interaction. Conventionally,
these interaction phenomena have been accounted for in the design
process by low- fidelity models, which negatively affect the
performance and reliability of the system.
More recently, numerical optimization methods with high-fidelity
simulation models have been applied to the de- sign of MEMS. In
particular, topology optimization is an appealing approach to the
design of MEMS for two dis- tinct reasons: (a) MEMS are often used
in new contexts and applications demanding conceptually new designs
and (b) the fabrication process allows for arbitrary planar geome-
tries without increasing the manufacturing costs. The latter aspect
is particularly noteworthy as topology optimization often generates
designs that are too complex to manufacture in a cost-effective
manner for conventional structural appli- cations. This common
issue does not apply to the design of MEMS since it is resolved by
their distinct manufacturing techniques.
The potential of topology optimization for the design of MEMS has
been studied before. The bulk of this work is concerned with the
design of elastic compliant mechanisms for force and displacement
amplification purposes (Bruns and Tortorelli 1998; Pedersen et al.
2001). Most often, the physical actuation mechanism is simplified
and replaced by a fixed, design-independent, external load. More
advanced studies account for thermomechanical and electrothermo-
mechanical actuation (Sigmund 1998, 2001a,b).
Most MEMS devices currently in use, however, are elec-
trostatically actuated. A topology optimization approach for this
class of devices has been lacking and is presented in this study.
Considering this type of interaction in the formulation and
solution of the topology optimization problem adds sig- nificant
complexity. In contrast to thermal and electrother-
M. Raulli, K. Maute
mal actuation, the electrostatic pressure is not an “internal”
force but acts on the interface between the structural and
electrostatic domains. A major challenge is the appropriate
treatment of the electrostatic pressure, as it is governed by the
electrostatic field equations and acts on a surface that is not
known in advance but evolves in the optimization process. The
proposed approach is based on a high-fidelity simulation model of
the coupled problem and analytical sen- sitivity analysis embedded
into a gradient-based optimiza- tion procedure. The formulations
and solution procedures of the topology optimization, analysis, and
sensitivity analysis problems are presented and applied to the
design of 2D and 3D electrostatically actuated mechanisms.
For this purpose, the paper is arranged in the follow- ing manner.
The basic formulations of the electromechanical topology
optimization and analysis problems are presented in Sect. 2. In
Sect. 3, the developments necessary for top- ology optimization
without restriction on the conducting in- terface are discussed.
Relevant optimization results are pre- sented in Sect. 4 followed
by a summary in Sect. 5.
2 Electromechanical topology optimization
Recent years have seen the application of topology opti- mization
to increasingly complex engineering design prob- lems (Eschenauer
and Olhoff 2001). Most often the topology optimization problem is
formulated as a material distribu- tion problem in a given design
space. Among numerous material interpolation schemes, the Solid
Isotropic Material with Penalization (SIMP) model in combination
with fil- tering techniques has become popular in the engineering
design community (Bendsøe 1989; Zhou and Rozvany 1991; Sigmund
1994). This approach leads to solid-void material distributions and
can be applied to the interpolation of var- ious material
parameters. For example, the elastic modulus (E) is interpolated as
follows:
Fig. 1 Fixed and free interfaces in electromechanical topology
optimization
E(ek) = s pE k E0 ; pE > 1 (1)
where ek refers to the kth element in a finite element mesh and sk
is an optimization variable that represents the solid- void state
of finite element k. The reader is referred to the textbook by
Bendsøe and Sigmund (2002) for a thorough in- troduction into
topology optimization. The proposed work follows and extends the
SIMP-based material topology op- timization approach for
electromechanical applications.
Based on the material formulation, topology optimiza- tion has been
extended for various multiphysics problems. For example, Rodriques
and Fernandes (1995) considered thermal loading, and Sigmund (1998,
2001a,b) studied the design of electrothermo-actuated MEMS, as did
Yin and Ananthasuresh (2002), including the use of multiple ma-
terials. These problems have in common that all physical fields
involved are defined for the entire design space and the coupling
results in mechanical forces due to thermal strains only. The
structural, thermal, and all other fields evolve au- tomatically
along with the material distribution in the course of the
optimization process.
Multiphysics topology optimization problems with loads acting on
the surface of the structure, however, are less adaptable to
material topology optimization as the surface is not known in
advance. The surface evolves in the course of the optimization
process and is not explicitly defined using the material
formulation of the topology optimization problem. Topology
optimization with design-dependent sur- face loads has been studied
by Hammer and Olhoff (2000), Chen and Kikuchi (2001), Bourdin and
Chambolle (2003), and Du and Olhoff (2004a,b). However, in these
studies the magnitude of the surface load is given rather than be-
ing governed by other physical fields. Though the issues of
changing location and magnitude are addressed, the present study
proposes to address them in a manner more suitable for modeling
multiphysics problems with a varying interface topology. The fully
coupled multiphysics problem is not simplified, as in previous
design-dependent loading stud-
Topology optimization of electrostatically actuated
microsystems
ies. Additionally, the use of parametric surfaces (Hammer and
Olhoff 2000; Du and Olhoff 2004a,b) to represent the topology of
the interface is limiting. The methodology pro- posed herein allows
for arbitrary interface locations and maintains the fully coupled
multiphysics solution of the electromechanical system. The downside
is, of course, a more complicated methodology and longer
computation times.
Maute and Allen (2004) applied topology optimization to the design
of aeroelastic structures. The aerodynamic loads act on the
fluid–structure interface, that is, the skin of the wing. For
practical purposes, only the layout of the internal structure of
the wing was determined by topology optimization. The external
shape of the structure (the skin) was fixed. While this limitation
makes sense for the design of wings, it might considerably restrict
the design space for other multiphysics applications with surface
loads, such as the design of electrostatically actuated MEMS.
The allowance of a free interface that evolves in the top- ology
optimization process leads to several difficulties in the
formulation of the optimization problem and the mod- eling and
evaluation of the system response. Figure 1 gives a conceptual idea
of fixed versus free interface topology op- timization problems.
The fundamental difficulties of a free interface are due to the
fact that the spatial extent of the elec- trostatic domain changes,
leading to a need to change the computational domain. Also, as
indicated by the arrows, the electrostatic force changes in
magnitude and location. Fi- nally, the voltage differential is
applied at different locations. The change in application of the
voltage affects the bound- ary conditions for the electrostatic
problem.
The objective and constraints of the optimization prob- lem consist
of parameters in the electromechanical problem, such as
displacements, stresses, electrostatic forces, etc., and are termed
“criteria” herein. The evaluation of these criteria is directly
linked to the location and extent of the conducting interface. If
the interface is allowed to change, as in Fig. 1b, this has a
significant impact on the evaluation of criteria. The evaluation of
the electromechanical system response is dis- cussed in Sect.
2.1.
Material topology optimization leads to a large num- ber of
optimization variables. Gradient-based algorithms can efficiently
handle a large number of variables. The coupled electromechanical
sensitivity analysis is presented in Sect. 2.2. Using
gradient-based methods, however, re- quires that the optimization
criteria and the system response be sufficiently smooth functions
of the optimization vari- ables. Therefore, since the interface is
free to evolve, the changes in Fig. 1b must be handled in a
continuous fash- ion. The modeling of the free interface is
presented in detail in Sect. 3.
2.1 Electromechanical analysis
A broad range of analysis methods are applied to the de- sign of
electromechanical MEMS devices. Many studies focus on the design of
geometrically simple MEMS and use analytical equations for the
analysis of the electrome- chanical response, for example, Abdalla
et al. (2003) for
microbeams and Bochobza-Degani and Nemirovsky (2004) for torsional
actuators. There have also been investigations into the use of a
combination of higher- and lower-fidelity- solution techniques.
Younis et al. (2003) proposed the use of reduced-order models for
simplifying parametric studies of the electromechanical response. A
methodology that uses 2D finite element models for structural
analysis and a sim- plified electrostatic loading has been proposed
by Chen et al. (2004).
High-fidelity solvers for electromechanical analysis have been
developed in the past decade. Senturia et al. (1992) presented a
high-fidelity finite element–boundary element solver, for the
mechanical and electrostatic analysis, re- spectively (Senturia et
al. 1992). A coupled sequential elec- tromechanical solver that
uses finite elements for both the mechanical and electrostatic
analysis has been developed by Zhulin et al. (2000).
This study uses finite element discretizations for both the
structural and electrostatic subproblems. Though boundary elements
are the standard for electrostatic analysis, the use of finite
elements is accurate for electrostatic simulations (Shaul and
Sumner 2004) and takes advantage of the exist- ing finite element
software. Additionally, when using a fi- nite element
discretization, topology optimization removes the material
contribution of nonoptimal elements. In order to achieve a smooth
representation of the geometry, a suf- ficiently fine mesh is
required that eliminates the use of boundary elements for topology
optimization of the full electromechanical domain. Also, since a
boundary element method only discretizes the interface, the
topology of the electrostatic field cannot be altered. See Sect.
3.2 for further discussion of this idea.
The structural domain is coupled to the electrostatic field because
of the electrostatic pressure exerted on the struc- ture. The
geometry of the electrostatic domain changes as the structural
shape changes due to deformation and opti- mization. The choice of
finite elements for the electrostatic discretization necessitates
the inclusion of a method that pre- vents the degeneration of the
electrostatic mesh due to the shape changes of the
electrostatic–structure interface. For boundary element techniques
only a surface mesh of the electrostatic domain is necessary and
only the geometry of this surface mesh needs to be updated when the
structure changes shape. Using a finite element mesh for the elec-
trostatic domain, however, requires that the structural shape
changes are propagated throughout the entire electrostatic mesh,
not just the interface. Subsequently, the mesh up- dating procedure
of the electrostatic mesh is referred to as mesh-motion.
The mesh-motion is treated as a ficticious physical field that is
coupled to the structural field via deformations and to the
electrostatic field by the determination of the elec- trostatic
mesh configuration. The state variables of the mesh-motion field
are the displacements of the electrostatic mesh due to the
structural deformations of the conduct- ing interface. This leads
to a three-field formulation of the electrostatic-mechanical
problem, similar to the one intro- duced by Farhat et al. (1998)
for fluid–structure interaction problems.
M. Raulli, K. Maute
The three-field formulation consists of the following residuals,
where u is structural displacements, v is electro- static voltages,
x is electrostatic mesh displacements, and s is the vector of
optimization variables:
S(s, u, v, x) = 0 Structure (2)
E(s, v, x) = 0 Electrostatic (3)
R(s, u, x) = 0 Electrostatic mesh (4)
The following state equations define the above residuals. It should
be noted that the electrostatic and mesh-motion equations (6)–(7)
are generally Dirichlet boundary condition problems, in which case
the only right-hand side in the re- sidual equation results from
constraining part of the solution vector. Furthermore, it is
assumed that the structure is a per- fect conductor.
S = Ku− fs(v, x) (5)
E = P v +PΓ vΓ −p ; vΓ = v on Γ v (6)
R = {
x Γ
−Tmu (7)
with K and P being the stiffness and permittivity matrices,
respectively, in a linear finite element representation. The
structural load due to mechanical and electrostatic forces is
represented by fs, and the right-hand side for the electro- static
problem due to prescribed charge is represented by p and is
generally 0. Prescribed voltage boundary conditions are represented
by v. The and Γ subscripts represent, respectively, the internal
and boundary portions of a par- titioned finite element stiffness
matrix. This separation is convenient for the solution of Dirichlet
boundary condition problems, as in (6)–(7). The matrix K represents
the ficti- cious stiffness of the electrostatic mesh. The
mesh-motion stiffness matrix K is built using elasticity finite
elements, which are topologically identical to the elements in the
elec- trostatic mesh. This is similar to the technique used for
flow problems by Johnson and Tezduyar (1994). The matrix Tm is the
transformation for passing structural displacements to the
electrostatic mesh. The transpose of this matrix trans- forms
electrostatic forces to the structural mesh. This matrix is built
with the energy-conserving matching procedure of Farhat et al.
(1998), which is applicable to matching and nonmatching
meshes.
The electrostatic force, fe, on the structure, which con- tributes
to the total structural load fs in (5), is computed as
follows:
fe = ∫
Te n dΓ E/S (8)
where Γ E/S represents the conducting interface between the
electrostatic and structure, n is the outward surface normal on the
interface, and Te is the Maxwell stress tensor, defined
below:
Te = ε
exey e2 y − 1
2e2 eyez
2e2
(9)
where ε is the permittivity of free space and e=[ex, ey, ez ]
represents the electric field, which is the spatial gradient of the
voltage.
The conducting interface is discretized with interface fi- nite
elements for evaluating (8). The interface elements are linearally
interpolated by two-node linear elements and four- node quadratic
elements for 2D and 3D applications, respec- tively. The nodes of
the interface elements exactly corres- pond with the nodes of the
electrostatic elements; therefore the electric field terms present
in the Maxwell electrostatic stress tensor (9) are taken directly
from the corresponding electrostatic nodes and used to compute the
electrostatic forces at the interface element nodes. The elemental
forces are then combined to obtain a global electrostatic force
vec- tor. Figure 4 illustrates the location of the interface
elements for fixed and free interface problems.
The overall coupled system (5)–(7) is solved using a staggered
procedure described in Maute et al. (2000) for a three-field
aeroelastic formulation, here extended to elec- tromechanical
coupling. The advantage of the staggered procedure is that
different solution techniques can be used for the three fields. All
three subproblems are solved with sequential sparse solvers. The
solution algorithm is pre- sented in Table 1. Equation (13)
assesses the convergence of the overall staggered procedure. The
electrostatic field is assumed to be converged if the structural
residual has con- verged since a direct solver is employed for the
computation of the electrostatic states.
2.2 Electromechanical sensitivity analysis
The gradients of the optimization criteria are evaluated by a
coupled electromechanical sensitivity analysis. Nu- merical
methods, such as finite differencing schemes, are frequently
applied to coupled multiphysics problems as they require no
modifications of existing analysis software (Giunta and
Sobieszczanski-Sobieski 1998). However, nu- merical schemes suffer
from added computational costs and accuracy problems. Therefore, in
this study, the gradients of the optimization criteria are computed
by solving the coupled electromechanical sensitivity
equations.
Previous work in the computation of analytical sensi- tivities for
coupled electromechanical problems has been accomplished for a
hybrid boundary element–finite element coupled solution technique
by Shi et al. (1995). The work of Shi et al. fully couples the
fields in analysis but does not fully couple them in the
sensitivity analysis. Also, the sensitivities of the electrostatic
field are computed by differ- entiating the governing equations and
then discretizing the gradients.
The analytical sensitivity approach used in this study fol- lows
the general framework for deriving the global sensitiv-
Topology optimization of electrostatically actuated
microsystems
Table 1 Electromechanical solution algorithm
Step 0: Initialize x(0) = u(0) on Γ E /S
For iteration (n):
x(n) Γ
= Tmu(n) (10)
Solve the mesh-motion equation (7) to determine x(n) and update the
electrostatic mesh configuration using x(n)
and x(n) Γ .
Step 2: Compute the internal electrostatic state vector (v ), using
(6).
Step 3: Compute force on structure from electrostatic
pressure:
fs (n) = Tm
T fe (n) (11)
The matrix Tm is the transformation matrix of (7), which transfers
electrostatic pressure to the structure when transposed (indicated
by the superscript T ). The electrostatic force vector (fe) is
computed with (8) and (9).
Step 4: Solve (5) for u, this is u(n) . Apply a relaxation factor
(θ) to u and u(n−1):
u(n) = (1−θ)u(n−1) +θu(n) (12)
Step 5: Check convergence:
S (s, u (n)
, x (0) ) (13)
where εem is a specified tolerance for the electromechanical
analysis. If (13) is satisfied, stop, otherwise go to Step 1.
ity equations of Sobieszczanski-Sobieski (1990). In general, the
derivative of an optimization criterion, qj , with respect to an
optimization variable, si , is computed as follows:
dqj
(14)
The crux of the gradient calculation in (14) is the computa- tion
of the state derivatives with respect to si .
The matrix A is the electromechanical Jacobian matrix, which is the
linearization of the electromechanical system about the equilibrium
solution. The off-diagonal terms in A
do not generally exist in analysis software. However, if they are
neglected, the sensitivities will be incorrect, leading to longer
convergence times or no convergence at all in the optimization
problem. Therefore, in this study these terms are derived by
analytically differentiating the associated dis- cretized terms in
the finite element models.
M. Raulli, K. Maute
where I is an identity matrix of appropriate size. The mesh- motion
component is divided into internal and boundary portions, since xΓ
= Tmu, not a static value, as with vΓ
and u Γ
. Only the internal portion of dv/ds is considered because the
boundary portion dvΓ /ds = 0, since the volt- age is prescribed in
this study. Two techniques for solv- ing (16), the direct and
adjoint methods, are discussed in Sect. 2.2.1.
2.2.1 Adjoint sensitivity method
The determination of the sensitivities of the optimization cri-
teria, of which there are nq , involves the computation of the
total state derivatives, obtained from solving (15), for each
optimization variable, of which there are ns:
Table 2 Adjoint method algorithm
Step 0: Compute and store ∂qj/∂u, ∂qj/∂v and ∂qj/∂x; initialize au
(0) j = 0, ax
(0) j = 0
Step 1: Compute the pseudoload for the structure and transfer it to
the structural mesh (fs).
fs = Tm T (
) (18)
Step 2: Determine a(k) uj , which is au
(k) j of (27), with fs substituted in. Apply a relaxation factor
(θ):
au (k) j = (1−θ)au
(k−1) j +θa(k)
Step 3: Transfer au (k) j to the electrostatic field.
a(k) uj
= Tmau (k) j (20)
In order to compute (28), the matrix–vector product of (
∂fs/∂v
)T au (k) j must be computed. This matrix–vector product is
computed as
follows:
∂fs
Compute av (k) j
with (28) and (22).
Step 4: Compute ax (k) j by solving (29) using av
(k)
∂fs
∂x
T
Step 5: Check convergence:
j
) (24)
where Au refers to the structural portion of the adjoint global
sensitivity equations and εsa is a specified tolerance for the
sensitivity analysis.
dqj
(17)
Topology optimization problems, however, lead to a large number of
optimization variables. In this study, for example, problems with
up to 7200 optimization variables are consid- ered. Therefore, the
adjoint method is used for the solution
Topology optimization of electrostatically actuated
microsystems
of (17), since the direct method is infeasible for a large
computational problem with this many variables. The com- putational
procedure applied to the adjoint equations follows the method of
Maute et al. (2003), which was developed for fluid–structure
interaction problems.
The adjoint method involves the solution of the transpose of (17)
with the solution of the state derivative substituted in:
dqj
−aj
(25)
In this form, only nq linear systems need to be solved for each
optimization step. In this study, no more than three cri- teria are
used for a given optimization problem. The adjoint method does
present the added difficulty that the inverse transpose ofA is
required. In order to obtain the adjoint vec- tor (aj), the
following equations need to be solved:
K 0 [ Tm
T K Γ
= 0 (26)
The diagonal terms in (26) are all symmetric; therefore, they
remain unaffected by the transpose. The off-diagonal terms,
however, are not symmetric and their transpose needs to be
computed. The term K
Γ is the transpose of K
Γ . The trans-
pose of A is not factorized and stored, but the adjoint states are
computed by solving the linear system in (26), with a staggered
Gauss–Seidel procedure, for each optimization criteria. Using
Gauss–Seidel on the block matrices inA , the following three
equations are solved during the process:
au (k) j = K−1
[ Tm
T (
) − ∂qj
∂u
] (27)
− ∂qj
∂x
(29)
The algorithm used to obtain the adjoint vector is sum- marized in
Table 2. An advantage of using this staggered procedure is that it
allows the use of direct solvers for the individual domains.
Therefore, the factorized stiffness ma- trices from the analysis
can be reused with minimal compu- tational cost. The off-diagonal
derivative matrices are never stored, due to memory considerations,
but are recomputed as a matrix–vector product for each step in the
staggered pro- cedure. The matrices only need to be formed on an
element level and multiplied by an element vector. This result is
then stored in a global vector.
3 Topology optimization model for free interface
3.1 Modified SIMP model for free interface
The classical SIMP model is extended in this study to the
electrostatic domain for the purpose of electromechanical topology
optimization with a free interface. Figure 1b il- lustrates how the
voltage boundary conditions, electrostatic forces, and
electrostatic domain all change when the con- ducting interface is
allowed to evolve during the optimiza- tion process. In order to
accurately determine the electrome- chanical response when the
interface is evolving and main- tain a smooth relationship between
the optimization vari- ables and the system response, the behavior
of the electro- static domain is altered for changes in the
material distribu- tion. The classical SIMP model only addresses
how struc- tural properties are affected by redistribution of
material. In order to effectively compute the coupled
electromechanical response, the analysis of the electrostatic
domain must also be considered. The modifications of the SIMP model
affect the analysis not only of the system but of the gradients as
well. The ∂S/∂si and ∂E/∂si terms of (15) are affected by the SIMP
model since structural and electrostatic parameters are linked to
optimization variables. The consideration of the above issues is
discussed in detail below.
3.1.1 Voltage boundary condition
In the electromechanical problems of this study a voltage
differential is applied between an electrode and a conduct- ing
body (the structure). Therefore, the conducting interface, where
the structure interfaces with the electrostatic domain, is subject
to a constrained voltage. This prescribed voltage differential is
what drives the electrostatic forcing and the system response. In
the computational electrostatic prob- lem, voltage is a Dirichlet
boundary condition. Therefore,
M. Raulli, K. Maute
Fig. 2 Changing of voltage boundary condi- tions during topology
optimization
changing the location of the conducting interface changes the
location of the Dirichlet boundary conditions, as illus- trated in
Fig. 2.
The methodology employed is to enforce all potential voltage
boundary conditions throughout the topology opti- mization process.
This cannot be done by strictly enforcing Dirichlet boundary
conditions because then changes in volt- age would be in an
“on–off” manner, which would lead to discontinuities.
Therefore, this issue is addressed by enforcing all po- tential
voltage boundary conditions indirectly, or “softly.” Rather than
eliminating equations in the electrostatic system for prescribed
voltages, a voltage boundary condition (vk) is enforced by adding a
weighting term (w(ek)
v ) to the corres- ponding diagonal entry in the permittivity
matrix (Pkk) that is relatively large. This term will dominate that
particular equation, leading essentially to Pkkvk = pk. This allows
the voltage to be enforced in a nonexplicit manner, through the
values in p, as shown in the following equations:
Pkk = w(ek) v +P0
pk = wvvk (31)
The terms w (ek) v and wv are defined in (32) and (33), re-
spectively. The value of the weighting term for a given
electrostatic element is made dependent on an optimization variable
such that the enforcement of vk changes in a smooth fashion:
w(ek) v = wv(sk − smin) (32)
wv = wv0Pavg
1− smin (33)
where Pavg is a scalar representing the average value of the
entries in P and wv0 is a user-defined value that determines how
large the relative weighting factor is. The denomina- tor in wv is
used to account for the fact that smin > 0. Since voltage is a
nodal quantity, each node with a soft voltage condition is linked
to one element to which it is connected. This approximation can be
further refined in future work, but for fine meshes the effect is
not significant.
3.1.2 Electrostatic mesh
As portions of the initial structural domain become void, regions
that were previously part of the structural domain but are now void
need to be considered as part of the elec-
trostatic domain. One potential solution is a remeshing of the
complete electromechanical domain in order to rede- fine the
electrostatic and structural domains. This method is not chosen
because it would require automatic remesh- ing based on the current
material distribution, which in itself can be difficult to
determine due to the high proportion of poorly defined “gray”
regions in the initial iterations. Addi- tionally, remeshing
results in discontinuities and automatic mesh generators typically
suffer from robustness problems.
The methodology used in this study is to generate an electrostatic
mesh that covers the initial electrostatic do- main (E0) as well as
the region occupied by the potential structural domain (Eδ), as
indicated on the left-hand side of Fig. 3. In this way,
electrostatic elements become “active” as the corresponding
structural elements become void, as il- lustrated on the right-hand
side of Fig. 3. This activation is accomplished in a continuous
manner, rather than an “on– off” manner, by adjusting the
permittivity of the electrostatic elements in Eδ.
The permittivity of an element in Eδ is made dependent on an
optimization variable such that overlapped electro- static elements
that correspond to solid structural elements (sk = 1) behave like
conductors (the electrostatic equivalent of rigid
structures):
ε(ek) = εmax(sk − smin) pε + ε0 (34)
εmax = ε0(smin) −pE (35)
where pE is the exponent penalizing the elastic modu- lus
computation, see (1). The maximum permittivity term (εmax) is
computed such that conducting electrostatic elem- ents have the
same relative magnitude to the free-space elec- trostatic elements
as the solid structural elements have to the void structural
elements, generally about 1×109 more. This term is necessary since
when sk = 1, it should cause the permittivity of that element to go
to infinity, resulting in a poorly conditioned global permittivity
matrix. There- fore, εmax is limited in magnitude to avoid
numerical prob- lems while still being large enough to represent a
conductor. The exponent (pε) is used such that elements that are
close to void are penalized to more closely represent free space
rather than a conductor.
3.1.3 Electrostatic forces
The elements in Eδ also have implications for the compu- tation of
the electric field, which is used to compute the electrostatic
forces (9). In the finite element problem, the
Topology optimization of electrostatically actuated
microsystems
Fig. 3 Overlap of structural and electrostatic mesh for initial
configuration (left) and optimized configuration (right). The
structural mesh and the electrostatic mesh in Eδ are spatially
identical but shown offset for illustration purposes
global electric field at a given node is computed by aver- aging
the local electric field in all the connected elements. If an
element in Eδ corresponds to a solid structural elem- ent (sk = 1),
the electric field computation from this element should not be
considered. Therefore, a weighting term for the electric field
computation in a given element is included in the optimization
formulation:
w(ek) e = we0
1− smin ; we0 = 1.0 (36)
The denominator in (36) is used such that the elemental weight
varies between 0 and 1 even though the optimization variable varies
between smin and 1. This SIMP parameter en- sures that the correct
electric field is computed where solid and void elements
meet.
In addition to the effect of the electric field on elec- trostatic
force computation, the interface between the elec- trostatic and
structural domains is continually changing, as illustrated in Fig.
3. Since electrostatic forces are only com- puted on the conducting
interface, the locations of the inter- face finite elements, which
compute the electrostatic forces with (8), are constantly changing.
The solution to this prob- lem is handled in a similar manner to
the electrostatic elem- ents: interface elements are generated at
all element bound- aries in Eδ. Figure 4 shows a conceptual
representation of
Fig. 4 Interface elements in standard (left) and SIMP
electromechanical models
this approach. This allows for a continuous representation of the
electrostatic forces. The basic methodology employed is that solid
structural elements are conducting; therefore, the interface
elements should fully compute forces there. For interface elements
attached to void structural elements, no electrostatic forces
should be computed. The interface elem- ents are modified as
follows:
ε (ek) i = ε0(sk − smin)
pi (37)
ε0 = ε0
(38)
where pi is the penalization exponent for intermediate values of sk
and ε0 is the same as is used for building the permittivity matrix.
The term ε
(ek) i refers only to the permit-
tivity used to compute the electrostatic forces in the interface
elements, see (9).
3.2 Optimization of electrostatic domain topology
As shown in the previous sections, especially in Figs. 1–4, the
proposed electromechanical topology model always con- tains some
region that is permanently free space (electro- static domain), in
addition to the regions where there is
M. Raulli, K. Maute
the potential for structure or free space. To achieve greater
flexibility in the overall design of an electromechanical sys- tem,
the topology of the electrostatic domain can also be altered. This
can be accomplished by making the permittiv- ity of electrostatic
elements independent optimization vari- ables (sj):
ε(ej ) = εsj ; smin,ε ≤ sj ≤ 1.0 (39)
The term smin,ε indicates that there may be a different minimum
value of the variables for the electrostatic elem- ents than the
structural elements. This permittivity variation refers only to the
part of the electromechanical domain that is always free space, not
the part that overlaps the struc- ture. Conceptually, this is the
equivalent of putting insulating material in the electrostatic
domain in order to prevent the electric field from being
transmitted in certain locations. Also, the topology of an
electrode can be changed in this manner. Making the permittivity
zero in electrostatic elem- ents connected to the electrode is the
physical equivalent of removing that part of the electrode. This
latter technique is used in the example of Sect. 4.2. The topology
optimization of purely electrostatic systems is addressed by Byun
et al. (2002).
It is assumed that the changes in permittivity do not af- fect the
stiffness of the structural domain in any way; there- fore, there
is no need for any additional parameters that couple the structural
model to the independent permittivity variables, sj .
3.3 Verification of SIMP model behavior
In addition to the two standard structural properties—the elastic
modulus and the density—four electrostatic proper- ties have been
introduced for smoothly varying the elec- tromechanical interface.
The associated SIMP parameters are summarized in Table 3. As
multiple material interpo- lation schemes are used, it is important
to verify that the modified SIMP model behaves in a manner that
encour- ages a “0-1” distribution. The simple example in Fig. 5 is
used to verify the behavior of the SIMP model. The state of the
structural/electrostatic element is controlled by one optimization
variable, s. The value of pE is fixed at 3.0, a common value in
topology optimization, and the value of pi is varied. Figure 6
plots the value of the displacements at the free structural nodes
labeled 1 and 2. The displacements at the two nodes are identical
and are plotted as s is varied between 0 and 1 for the different
values of pi .
Table 3 Electrostatic SIMP parameters
w (ek ) v Controls the ‘soft’ enforcement of voltage boundary
conditions
w (ek ) e Adjusts weighting of elemental contribution to electric
field
ε(ek) Changes permittivity of overlapped electrostatic
elements
ε (ek) i Changes permittivity of overlapped interface
elements
Fig. 5 Simple mesh for exponent verification. E: electrostatic
elem- ents, S: structural elements
Fig. 6 Displacement relationship to optimization variable, for
varying pi
The plots in Fig. 6 illustrate the importance of choosing the
correct exponential values in order to encourage a “0-1”
distribution. Since the goal of the SIMP method is to pe- nalize
intermediate variables, thereby obtaining a “0-1” dis- tribution,
the curves resulting from pi ≥ 3.0 are desirable. Though the
displacement for all values of pi are equivalent at the endpoints,
lower values of pi can lead to a nonconvex
Fig. 7 Structural (top) and electrostatic meshes for 2D exact
represen- tation
Topology optimization of electrostatically actuated
microsystems
Fig. 8 Structural (left) and electrostatic meshes for 2D topology
representation
relationship between displacements and optimization vari- ables
that does not encourage a “0-1” distribution. In this study, 3.5 ≤
pi ≤ 4.0. This range is chosen over pi = 6.0 because the gradients
for s < 0.3 are small and thereby do not strongly force the
optimization problem to the lower bound. This example of the SIMP
model behavior is rel- evant for optimization objectives involving
displacements and strain energy, with mass constraints, which is
the case in this study.
Another aspect of the SIMP model that requires verifica- tion is
that it reproduces a system response consistent with one in which
the material distribution is exactly represented by the
computational meshes for completely solid and void elements.
The meshes used for an exact and a topology optimiza- tion problem
are given in Figs. 7 and 8, respectively. The material distribution
of the structural mesh in Fig. 8 is ad- justed to the solid–void
distribution that corresponds to the geometry in Fig. 7. The arrows
on the structural mesh are displacement boundary conditions. The
arrowheads on the electrostatic mesh are strictly enforced voltage
boundary conditions. The physical parameters for the problem are
given in Table 4.
In order to assess the agreement between the two prob- lems,
several physical quantities are compared visually and numerically.
Figure 9 shows the material distribution for the topology problem,
which mimics the geometry of the ex- act problem. Figure 10 shows
the voltage distribution in the exact and topology models, which
are visually similar.
The norms of the nodal vectors for electrostatic and structural
quantities, as well as the scaled norm of the dif- ference between
the exact and topology solutions, are given in Table 5. It should
be noted that the size of the exact and
Table 4 Electrostatic and structural properties for verification of
2D topology approximations
Electrostatic Permittivity Voltage Min/max air gap Vacuum
8.85×10−12 F/m 100.0 V 0.5/2.5 µm
Structure Elastic modulus Poisson ratio Plate thickness
Width/height Silicon 1.5×1011 Pa 0.17 1.0×10−7 m 6.0 µm
topology vectors are different, since there are more nodes in the
topology model than in the model based on the true geometry.
However, the additional nodes in the topology models are filtered
out such that the differences in solution vectors can be
effectively compared. The actual norms of the exact and
(unfiltered) topology models are shown to demon- strate that the
nodal values filtered out of the topology model are small in
comparison to the exact nodal values. It should be noted that the
value of the topology norm for the struc- tural displacements still
has the nodes filtered out since there are nonzero displacements in
the topology model nodes that do not correspond to nodes in the
exact model. The associ- ated elements, however, are not
significantly contributing to the stiffness. The electrostatic and
structural quantities agree well for the exact and topology
models.
Fig. 9 Material distribution for topology model
M. Raulli, K. Maute
Fig. 10 Voltage distribution for 2D exact model (left) and topology
model
Table 5 Numerical comparison of exact and topology models
Exact norm (e) Topology norm (t) Scaled difference norm (e−
t/e)
voltage 8.35048712×102 8.35049176×102 1.2662×10−6
x-electric field 4.48090547×108 4.48090344×108 3.1619×10−6
y-electric field 2.16490769×109 2.16490786×109 6.4872×10−7
x-electric force in E 1.65375485×10−9 1.65375030×10−9
2.8157×10−6
x-electric force in S 1.65375000×10−9 1.65375000×10−9
1.0761×10−2
y-electric force in E 1.41106880×10−8 1.41114835×10−8
5.6671×10−7
y-electric force in S 1.41106970×10−8 1.41114938×10−8
1.0761×10−2
x-displacement 7.67808298×10−11 7.76841956×10−11 1.2014×10−2
y-displacement 1.64765068×10−10 1.64072914×10−10 5.7870×10−3
4 Examples
The proposed topology optimization methodology is ap- plied to the
design of electrostatically actuated mechanisms. The example in
Sect. 4.1 is a 2D force inverter. Section 4.2 presents a 3D force
inverter. In both examples the optimiza- tion problems are solved
by the Method of Moving Asymp- totes (MMA) (Svanberg 1987). The
Electromechanical sys- tem response is computed by the
computational procedure described in Sect. 2.1. The gradients of
the optimization criteria are computed by the adjoint approach of
Sect. 2.2. Though the geometric dimensions in the following exam-
ples are small for MEMS devices and more appropriate for nanoscale
devices, the examples demonstrate the effective- ness of this
methodology for performing topology optimiza- tion on coupled
electromechanical systems.
4.1 Two-dimensional force inverter
The goal of this example is to create a force inverter. A simi- lar
design problem has been studied by Sigmund (1997), Pedersen et al.
(2001), and Maute and Frangopol (2003). In contrast to the above
problems, the actuation force is not
a given force but the electrostatic pressure acting on the bot- tom
edge of the structure, which pulls the structure towards the
electrode. Figure 11 gives a schematic of this optimiza- tion
example. The desired design will invert the electrostatic pressure
acting in the negative y-direction such that point C in Fig. 11
moves in the positive y-direction.
The optimization problem is formulated as follows:
maxs z(s) = uc subject to:
Mass ≤ 10% of total
(40)
where uc is the displacement at point C in Fig. 11. The cur- rent
and initial strain energy are represented by Π and Π0,
respectively. A static force of −5.0×10−13 N is applied at the same
node as the displacement objective in order to sim- ulate actuation
of a work piece. See Table 6 for the values of the SIMP parameters
used in this example.
The mass constraint is applied in order to encourage a ‘0-1’
distribution using the SIMP model. An energy con- straint is also
used in the problem formulation, for two rea- sons:
Topology optimization of electrostatically actuated
microsystems
Fig. 11 Schematic of 2D force inverter example
Table 6 SIMP parameters for Sect. 4.1
Elastic modulus penalization (pE ) 3.0 Electrostatic permittivity
penalization (pε) 6.0 Interface permittivity penalization (pi ) 4.0
Soft voltage weighting factor (wv0 ) 1.0×104
1. The optimization process seeks to maximize the up- ward
displacement. Since the electrostatic pressure is what drives the
inversion, once the appropriate mech- anisms have been determined
by the optimization pro- cess, a larger electrostatic pressure will
lead to increased upward displacement. This is good from the point
of
Fig. 12 Structural (left) and electrostatic computa- tional
meshes
view of the objective, but if the electrostatic pressure becomes
too strong, the electromechanical system will become unstable and
be pulled into contact with the elec- trode (pull-in), causing the
elements in the electrostatic mesh to collapse and the optimization
process to stop. To ensure that the optimization process converges
to an optimal design and that the design is stable, an energy
constraint is enforced such that the structure is limited in its
deformations. The value of Π0 is chosen because allowing higher
energies consistently leads to pull-in.
2. Since the energy constraint limits the overall strains of the
structure, the optimization process is forced to stiffen in order
to satisfy the constraint. This leads to a more “0-1” material
distribution.
This example is treated as purely 2D, in both the struc- tural and
electrostatic analysis. Figure 12 shows the full structural and
electrostatic meshes used in this example. Ap- plying symmetry
boundary conditions, only half of each computational domain is
analyzed and optimized. The struc- ture is discretized with
four-node quadratic plane stress elements. The electrostatic mesh
uses four-node quadratic elements. The physical properties and
computational mesh sizes for this optimization problem are given in
Tables 7 and 8, respectively.
In addition to running the problem with a free interface, the same
problem is run with a fixed interface. Both prob- lems are started
with an initial interface with a height of one element; however,
the fixed interface problem does not have the freedom to remove the
row of interface elements. The same numerical values for mass and
energy constraints are used in the fixed and free problems, in
order to effec- tively compare the methodologies. The optimization
results are summarized in Table 9. The initial displacement in the
free interface problem is less because the interface is started at
80% of solid. The reported times are for running each
M. Raulli, K. Maute
Table 7 Electrostatic and structural properties for Sect. 4.1
Electrostatic Permittivity Voltage Initial air gap Vacuum
8.85×10−12 F/m 2.0 V 0.5 µm
Structure Elastic modulus Poisson ratio Thickness Initial width
& height Silicon 1.5×1011 Pa 0.17 1.0×10−7 m 5.0 µm
Table 8 Summary of computational problem for Sect. 4.1
Nodes Total DOF Free DOF Elements
Structure 7381 14,762 14,758 7200 Electrostatic 8133 8133 8072 7920
Mesh-motion 8133 16,266 16,144 7920 Total 23,647 39,161 38,974
23,040
Table 9 Summary of optimization problem and results for Sect.
4.1
Free interface Fixed interface
# optimization variables 7200 7080 Initial uc −1.2972×10−12
−4.1089×10−12
Final uc 2.1418×10−9 6.7104×10−10
Displacement change from fixed initial value
−5.2127×104% −1.6332×104%
Energy/Mass constraint ac- tive
Yes/Yes Yes/Yes
Number of iterations 2649 5295 Total time 55.4 hours 7.57
hours
computational domain on one Pentium IV 1.7 GHz proces- sor.
The final material distribution is shown in Fig. 13 for the free
and fixed interfaces. The methodology that allows for a free
interface makes this problem much more flexible, as illustrated by
the fact that much of the bottom layer of the structure is removed
by the optimization process and the al- lowable material is used to
create the necessary mechanisms and stiffen the structure such that
the energy constraint is not
Fig. 13 Final material distribu- tion for Sect. 4.1, with a free
(left) and fixed interface
violated. In the fixed interface problem, the objective can- not
improve as much because of the need to brace the entire fixed
interface with extra supports, allowing less material to be used in
making the mechanism more efficient. Also, the need to support the
fixed interface leads to a more com- plex structure. The
computational time for the fixed problem is less because there is
no need to overlap the electrostatic domain, resulting in a much
smaller electrostatic computa- tional problem.
One minor issue is that the interface in the free problem is not
entirely solid. This is due to the fact that the prob- lem finds
mechanisms that accomplish the desired goal of a force inverter
while violating the energy constraint. Rather than changing the
mechanisms in order to satisfy the energy constraint, the
optimization algorithm reduces the material fraction at the
interface, which effectively reduces the elec- trostatic forcing.
If this piece was manufactured, this inter- face could be made
solid and the voltage reduced, achieving the same behavior. Forcing
the elements at the interface to be completely solid for force
inverter problems should be addressed in future work.
4.2 Three-dimensional force inverter
The goal of this example is also to create a force inverter. In
addition to the optimization of the structural topology, the
permittivity in the layer of electrostatic elements above the
electrode is also optimized, essentially determining the top- ology
of the electrode. Figure 14 gives a schematic of this optimization
example. The desired design inverts the elec- trostatic pressure
acting in the negative y-direction such that point t in Fig. 11
moves in the positive y-direction. This can- tilevered example does
not have the advantage of opposing
Topology optimization of electrostatically actuated
microsystems
Fig. 14 Schematic of 3D force inverter example
supports to form mechanisms around, as does the 2D ex- ample in
Sect. 4.1.
The optimization problem is run as follows. 1. The optimization
process is started with the following
initial formulation:
Mass ≤ 10% of total
(41)
where ut is the displacement at point t in Fig. 14. The su-
perscripts S and E indicate optimization variables relat- ing to
solid–void status in the structural and electrostatic domains,
respectively. The constraint that prevents the strain energy from
decreasing serves the purpose of forc- ing the problem away from
the local optimum of a stiff structure. The optimization problem
was attempted with- out this constraint, and the results were a
stiffer structure with a marginal decrease in the negative
y-displacement of point t. With the energy constraint, the
optimization algorithm gradually finds the desired optimum of
invert- ing the displacement. The mass constraint is used for
clarity of the final design. See Table 10 for the values of the
SIMP parameters used in this example.
2. The optimization process slowly minimizes the negative
y-displacement until it finally inverts the displacement, indicated
by ut becoming positive. The positive displace- ment grows, and
eventually the structure reaches pull-in because there is no
constraint to prevent the increase of the structural displacements,
causing the simulation pro-
Table 10 SIMP parameters for Sect. 4.2
Elastic modulus penalization (pE ) 3.0 Interface permittivity
penalization (pi ) 3.5 Soft voltage weighting factor (wv0 )
1.0×104
cedure to fail. At this point, the optimization process is
restarted, with an additional constraint added to the op-
timization problem.
maxs z(s) = uc subject to:
Mass ≤ 10% of total
(42)
where Π = 4.0 × 10−19, which is approximately one quarter the value
of the strain energy at the last stable it- eration before pull-in,
Πlast1 = 1.7391×10−18, in step 1. In this case, the energy
constraint is used to prevent the system from becoming unstable.
This optimization for- mulation is run until convergence.
Varying the permittivity independently in the lower layer of
electrostatic elements allows the optimization algorithm greater
flexibility in determining the electrostatic pressure on the
structure. If the permittivity in the electrostatic do- main is not
varied, then there will be a large voltage gra- dient everywhere
that there is structural material, resulting in significant
electrostatic forces. By varying the permit- tivity beneath a
location where there is structural material, it is possible to have
solid structure without creating elec- trostatic pressure. The
voltage is not transmitted through the insulating layer, leading to
negligible voltage gradients and therefore negligible electrostatic
pressure. For compar- ison purposes, the optimization problem was
run without the electrostatic permittivity variations, yielding
only an in- significant improvement in the objective.
This example is fully 3D, in both the structural and elec-
trostatic analyses. Figure 15 shows the full structural and
electrostatic meshes used in this example. The structure is
discretized with three-node triangular ANDES plate elem- ents
(Militello and Felippa 1991). The electrostatic mesh uses
eight-node hexahedron elements. The physical prop- erties and
computational mesh sizes for this optimization problem are given in
Tables 11 and 12, respectively.
Fig. 15 Structural and electrostatic mesh for Sect. 4.2
M. Raulli, K. Maute
Table 11 Electrostatic and structural properties for Sect.
4.2
Electrostatic Permittivity Voltage Initial air gap Vacuum
8.85×10−12 F/m 5.0 V 0.5 µm
Structure Elastic modulus Poisson ratio Thickness Init. width &
height Silicon 1.5×1011 Pa 0.17 1.0×10−7 m 5.0 µm
Table 12 Summary of computational problem for Sect. 4.2
Nodes Total DOF Free DOF Elements
Structure 1681 10,086 9840 6400 Electrostatic 8405 8405 6724 6400
Mesh-motion 8405 25,215 20,172 6400 Total 18,491 43,706 36,736
19,200
Fig. 16 Final material distribution for Sect. 4.2
The optimization results are summarized in Table 13. The
computational time is for running each computational domain on one
Pentium IV 1.7 GHz processor. The final structural topology is
shown in Fig. 16. A schematic of the
Fig. 18 Permittivity distribution (left) and voltage distribution
in displaced mesh
Fig. 17 Schematic of 3D force inverter
final design is given in Fig. 17 to illustrate its functional- ity.
The center part of the device, connected to the support, serves as
a fulcrum. The part of the device attached to the side of the
support is pulled downward, by the electrostatic forces, allowing
the curved lever arm, which is connected to node t, to rotate
around the center support.
Figure 18 shows the permittivity distribution in the elec-
trostatic mesh, which varies from the permittivity of vac- uum
(black) to zero (white). Figure 18 also shows the elec- trostatic
voltage distribution in the deformed electrostatic mesh. The
voltage varies between zero volts (white) and five volts (black).
In the permittivity distribution, the front corner
Topology optimization of electrostatically actuated
microsystems
Table 13 Summary of optimization problem and results for Sect.
4.2
# optimization variables 1600 (S ); 1600 (E )
Initial ut −2.4799×10−9
Final ut 1.7226×10−6
Displacement change −6.9461×104% Number of iterations 306 Total
time 193.5 hours
corresponds to the front corner of the voltage distribution. The
other corner in the permittivity plot with the solid patch
corresponds to the back corner along the left edge of Fig. 16. The
perspective is different because the electrostatic mesh in the
permittivity plot has been flipped upside down to show the layer of
elements on the electrode. In effect, the electrode has been
reduced to two small circular patches, only one of which
significantly contributes to the force inverter.
5 Summary
The design of MEMS is continually evolving, with chang- ing
parameters and applications. The conceptual design of MEMS in an
automatic fashion through the use of high- fidelity topology
optimization is a powerful tool for design- ing new devices. This
study has presented a methodology for performing topology
optimization of MEMS that are electrostatically actuated, without
limitation on the interface between the structural and
electrostatic computational do- mains, allowing for greater freedom
in the generation of op- timal topologies for various design
objectives. This method- ology requires a fully coupled sensitivity
analysis of the electromechanical response in addition to the fully
coupled analysis. Additionally, the classical SIMP model is
modified for electromechanical problems. The voltage boundary con-
ditions are enforced in an indirect manner in order to allow a
flexible interface. Two numerical examples of force invert- ers
were presented to show the applicability of the developed
methodology.
The results illustrated the advantages of varying the in- terface
topology and the layout of the electrode versus con- ventional
approaches optimizing the internal structural lay- out only. In
this study, constraints on the strain energy were introduced to
prevent pull-in instabilities. In future stud- ies, constraints on
pull-in instabilities should be directly accounted for, requiring
appropriate prediction and sensi- tivity analysis capabilities of
this phenomenon, which are currently lacking.
Acknowledgement The first author would like to acknowledge the
support of Sandia National Laboratory under the direction of Jim
Allen. Both authors acknowledge the support by the National Science
Foundation under Grant DMI-0300539 and the Air Force Office of
Scientific Research under Grant F49620-02-1-0037. The opinions and
conclusions presented in this paper are those of the authors and do
not necessarily reflect the views of the sponsoring agencies.
References
Abdalla MM, Reddy CK, Faris W, Gurdal Z (2003) Optimal design of an
electrostatically actuated microbeam for maximum pull-in voltage.
In: AIAA/ASME/AHS/AISC 43rd conference on struc- tures, structural
dynamics and materials, Norfolk, VA
Bendsøe MP (1989) Optimal shape design as a material distribution
problem. Struct Optim 1:193–202
Bendsøe MP, Sigmund O (2002) Topology optimization, theory, methods
and applications. Springer, Berlin Heidelberg New York
Bochobza-Degani O, Nemirovsky Y (2004) Experimental verification of
a design methodology for torsion actuators based on a rapid pull-in
solver. J Microelectromech Syst 13(1):121–130
Bourdin B, Chambolle A (2003) Design-dependent loads in topology
optimization. ESAIM Control Optimisat Calculus Variat 9:19–48
Bruns TE, Tortorelli DA (1998) Topology optimization of geometri-
cally nonlinear structures and compliant mechanisms. In: Pro-
ceedings of the 7th AIAA/USAF/NASA/ISSMO symposium on
multidisciplinary analysis and optimization, St. Louis, MO, pp
1874–1882
Bustillo JM, Howe RT, Muller RS (1998) Surface micromachining for
microelectromechanical systems. Proc IEEE 86(8):1552–1574
Byun JK, Park IH, Hahn SY (2002) Topology optimization of elec-
trostatic actuator using design sensitivity. IEEE Trans Magnet
38(2):1053–1056
Chen B-C, Kikuchi N (2001) Topology optimization with design-
dependent loads. Finite Elements Anal Des 37:57–70
Chen K-S, Ou K-S, Li L-M (2004) Development and accuracy assess-
ment of simplified electromechanical coupling solvers for mems
applications. J Micromech Microeng 14:159–169
Du J, Olhoff N (2004) Topological optimization of continuum struc-
tures with design-dependent surface loading—Part I: new com-
putational approach for 2d problems. Struct Multidisc Optim
27:151–165
Du J, Olhoff N (2004) Topological optimization of continuum struc-
tures with design-dependent surface loading—Part II: algorithm and
examples for 3d problems. Struct Multidisc Optim 27:166– 177
Eschenauer HA, Olhoff N (2001) Topology optimization of continuum
structures: a review. Appl Mech Rev 54(4):331–389
Farhat C, Degand C, Koobus B, Lesoinne M (1998) Torsional springs
for two-dimensional dynamic unstructured fluid meshes. Comput
Methods Appl Mech Eng 163:231–245
Farhat C, Lesoinne M, LeTallec P (1998) Load and motion trans- fer
algorithms for fluid/structure interaction problems with non-
matching discrete interfaces: momentum and energy conservation,
optimal discretization and application to aeroelasticity. Comput
Methods Appl Mech Eng 157:95–114
Giunta AA, Sobieszczanski-Sobieski J (1998) Progress towards using
sensitivity derivatives in a high-fidelity aeroelastic analysis of
a supersonic transport. In: AIAA 98–4763, 7th AIAA/USAF/ NASA/ISSMO
symposium on multidisciplinary analysis and op- timization, St.
Louis, MO, September 1998, pp 441–453
Hammer VB, Olhoff N (2000) Topology optimization of continuum
structures subjected to pressure loading. Struct Multidisc Optim
19:85–92
Johnson AA, Tezduyar TE (1994) Mesh update strategies in parallel
finite element computations of flow problems with moving bound-
aries and interfaces. Comput Methods Appl Mech Eng 119:73–94
Maute K, Allen M (2004) Conceptual design of aeroelastic structures
by topology optimization. Struct Multidisc Optim 27:27–42
Maute K, Lesoinne M, Farhat C (2000) Optimization of aeroelastic
systems using coupled analytical sensitivities. In: AIAA 2000–
0560, 38th Aerospace Science Meeting and Exhibit, Reno, NV, 10–13
January 2000
Maute K, Frangopol D (2003) Reliability-based design of mems mech-
anisms by topology optimization. Comput Struct 81:813–824
Maute K, Nikbay M, Farhat C (2003) Sensitivity analysis and design
optimization of three-dimensional non-linear aeroelastic systems by
the adjoint method. Int J Numer Methods Eng 56(6):911–933
M. Raulli, K. Maute
Militello C, Felippa C (1991) The first andes elements: 9-dof plate
bending triangles. Comput Methods Appl Mech Eng 91:217–246
Pedersen CBW, Buhl T, Sigmund O (2001) Topology optimization of
large displacement compliant mechanism. Int J Numer Methods Eng
50:2683–2705
Rodriques H, Fernandes P (1995) A material based model for top-
ology optimization of thermoelastic structures. Int J Numer Methods
Eng 38:1951–1965
Senturia S, Harris R, Johnson B, Kim S, Nabors K, Shulman M, White
J (1992) A computer-aided design system for micro-
electromechanical systems (memcad). J Microelectromech Syst
1(1):3–13
Shaul D, Sumner T (2004) Estimating accuracy of electrostatic
finite elementmodels.CommunicationsNumerMethodsEng20:313–321
Shi F, Ramesh P, Mukherjee S (1995) Simulation methods for micro-
electro-mechanical structures (MEMS) with application to a mi-
crotweezer. Comput Struct 56:769–783
Sigmund O (1994) Design of material structures using topology opti-
mization. PhD thesis, Danish Center for Applied Mathematics and
Mechanics, Technical University of Denmark, Lyngby, Denmark
Sigmund O (1997) On the design of compliant mechanisms. Mech Struct
Mach 25:493–524
Sigmund O (1998) Topology optimization in multiphysics problems.
In: AIAA 98–4905, Proceedings of the 7th AIAA/USAF/NASA/ ISSMO
symposium on multidisciplinary analysis and optimiza- tion,
St.Louis, MO, pp 1492–1500
Sigmund O (2001) Design of multiphysics actuators using topology
optimization—Part I: One-material structures. Comput Methods Appl
Mech Eng 190(49–50):6577–6604
Sigmund O (2001) Design of multiphysics actuators using topology
optimization—Part II: Two-material structures. Comput Methods Appl
Mech Eng 190(49–50):6605–6627
Sobieszczanski-Sobieski J (1990) Sensitivity of complex, internally
coupled systems. AIAA J 28:153–160
Svanberg K (1987) The method of moving asymptotes—a new method for
structural optimization. Int J Numer Methods Eng 24:359–373
Yin L, Ananthasuresh GK (2002) A novel topology design scheme for
the multi-physics problems of electro-thermally actuated compliant
micromechanisms. Sensors Actuators A 97–98:599– 609
Younis M, Abdel-Rahman E, Nayfeh A (2003) A reduced-order model for
electrically actuated microbeam-based mems. J Microelec- tromech
Syst 12(5):672–680
Zhou M, Rozvany GIN (1991) The coc algorithm, Part II: Topological,
geometrical and generalized shape optimization. Comput Methods Appl
Mecha Eng 89(1–3):309–336