Resonant micro beam

Embed Size (px)

Citation preview

  • 7/27/2019 Resonant micro beam

    1/27

    Nonlinear Dynamics 31: 91117, 2003.

    2003 Kluwer Academic Publishers. Printed in the Netherlands.

    A Study of the Nonlinear Response of a Resonant Microbeam to

    an Electric Actuation

    M. I. YOUNIS and A. H. NAYFEHDepartment of Engineering Science and Mechanics, MC 0219, Virginia Polytechnic Institute and State

    University, Blacksburg, VA 24061, U.S.A.; E-mail: [email protected]

    (Received: 21 February 2002; accepted: 1 October 2002)

    Abstract. An investigation into the response of a resonant microbeam to an electric actuation is presented. A non-

    linear model is used to account for the mid-plane stretching, a DC electrostatic force, and an AC harmonic force.

    Design parameters are included in the model by lumping them into nondimensional parameters. A perturbation

    method, the method of multiple scales, is used to obtain two first-order nonlinear ordinary-differential equations

    that describe the modulation of the amplitude and phase of the response and its stability. The model and the results

    obtained by the perturbation analysis are validated by comparing them with published experimental results. Thecase of three-to-one internal resonance is treated.

    The effect of the design parameters on the dynamic responses is discussed. The results show that increasing

    the axial force improves the linear characteristics of the resonance frequency and decreases the undesirable fre-

    quency shift produced by the nonlinearities. In contrast, increasing the mid-plane stretching has the reverse effect.

    Moreover, the DC electrostatic load is found to affect the qualitative and quantitative nature of the frequency-

    response curves, resulting in either a softening or a hardening behavior. The results also show that an inaccurate

    representation of the system nonlinearities may lead to an erroneous prediction of the frequency response.

    Keywords: MEMS, resonator, primary resonance, forced vibration.

    1. Introduction

    Electrically actuated microbeams have been widely used and studied by the MEMS com-

    munity. They form a major component in many micromechanical devices, such as capacitive

    switches [1, 2], pull-in sensors [3], and resonant sensors [416]. In this paper, we study

    resonant sensors using electrically actuated microbeams.

    The fundamental natural frequency of a resonant microbeam is very sensitive to the axial

    strain. External loads, such as pressure, temperature, force, and acceleration, apply axial

    strains on a microbeam, leading to a shift in its natural frequencies. This shift is readily

    converted to a digital signal, which is related to the physical quantity being measured. The

    fundamental natural frequency also can be affected by other factors, such as squeeze-film

    damping [4] and the elasticity of the microbeam supports [5].

    The electric load and the mechanical restoring force govern the behavior of a microbeam.The electric load is composed of a DC polarization voltage and an AC voltage. The DC

    component applies an electrostatic force on the microbeam, thereby deflecting it to a new

    equilibrium position, while the AC component vibrates the microbeam around this equilibrium

    position. The combined electric load has an upper limit beyond which the mechanical restoring

    force can no longer resist its opposing force, thereby leading to the collapse of the microbeam.

    This structural instability phenomenon is known as pull-in, and the critical voltage associated

  • 7/27/2019 Resonant micro beam

    2/27

    92 M. I. Younis and A. H. Nayfeh

    with it is called pull-in voltage. Younis et al. [6] showed that the pull-in voltage corresponds

    to a saddle-node bifurcation. They calculated the static deflection at each DC voltage and

    found two solution branches: the larger solution is unstable and the smaller one is stable. As

    the DC voltage increases, these branches get closer to each other, coalesce , and destroy each

    other in a saddle-node bifurcation, corresponding to the pull-in voltage. The deflection at pull-

    in is approximately 57% of the airgap space [6, 7], which separates the microbeam from the

    stationary substrate.

    A microbeam restoring force is composed of three components: bending, axial force, and

    mid-plane stretching due to immovable boundaries. The mechanical restoring force tends to

    shift the natural frequencies to higher values, while the electrostatic force tends to shift the

    natural frequencies to smaller values.

    A number of studies have investigated the problem of a microbeam that is initially deflected

    by a DC electrostatic force and driven by an AC force. Zook et al. [8] reported experiment-

    ally that increasing the driving AC voltage leads to an increase in the resonance frequency

    (hardening behavior). Also, they observed a hysteresis that depends on the direction of the

    frequency sweep.

    Tilmans and Legtenberg [9] approximated the nonlinear resonance frequency as the fun-

    damental natural frequency for small values of the AC forcing amplitude. For large-amplitudeAC excitations, they approximated the dynamic problem using Rayleighs energy method,

    modified to account for electric forces and mid-plane stretching. They derived an equation that

    governs the resonance frequency as a function of the excitation amplitude. They compared the

    results obtained using this equation to their experimental results. Although they found a good

    qualitative agreement, showing a hardening effect, the agreement was poor quantitatively.

    They concluded that such a hardening behavior is more severe for high-quality factors, high

    DC voltages, and high AC voltages, whereas large axial strains can decrease it.

    Gui et al. [10] investigated the dynamic problem using Rayleighs energy method, modi-

    fied to account for electric forces and mid-plane stretching. They neglected the applied axial

    force. They derived an equation for the fundamental resonance frequency, which predicts a

    hardening behavior. They also derived a criterion that defines a region of DC and AC voltages

    for a hysteresis-free operation. They compared their theoretical and experimental results and

    found good agreement.

    Turner and Andrews [11] approximated the nonlinear resonance frequency of a microbeam

    using a perturbation method. The problem was modeled using a spring-mass system with a

    cubic nonlinearity representing mid-plane stretching. Using the method of harmonic balance,

    they derived equations describing the microbeam resonance frequency for two separate cases.

    In one case, they neglected the electrostatic force and included mid-plane stretching, and,

    in the second case, they included the electrostatic force and neglected mid-plane stretching.

    Using the equations obtained from both cases, they derived an equation that compensates for

    the increase in the resonance frequency due to mid-plane stretching and the decrease in the

    resonance frequency caused by the electrostatic force. The results showed that eliminating the

    effect of the frequency dependence on the amplitude requires a very high DC voltage, whichmay not be attainable.

    Ayela and Fournier [12] experimentally studied the response of microbeams of different

    geometric shapes to a general electric load composed of DC and AC components. They

    made several plots to show variation of the resonance frequency with the excitation amplitude

    for different axial loads. The experimental data were used to extract parameters that fit the

    equation of a spring-mass system with cubic nonlinearity and a linear electric force (i.e., the

  • 7/27/2019 Resonant micro beam

    3/27

    Nonlinear Response of Resonant Microbeams to Electric Actuations 93

    electric force is assumed to be independent of the gap distance). The experimental results

    showed that, for the operating conditions used, some devices exhibited softening behavior,

    whereas others exhibited hardening behavior. They concluded that a nonlinear behavior may

    result from many distinct phenomena and that each case has to be studied separately. They

    added that the nonlinear behavior of these devices may be investigated experimentally, but

    there is no way of predicting it analytically.

    Veijola et al. [13] modeled a microbeam using a spring-mass model with a cubic nonlin-

    earity representing mid-plane stretching. They included the effect of the driving AC voltages,

    but neglected the effect of the DC polarization voltages. They used the method of harmonic

    balance to show that the nonlinearity of the electrostatic forces can cause a softening-type

    behavior, whereas the mid-plane stretching can cause a hardening-type behavior. They con-

    ducted an experiment, which showed a hardening behavior. They used the results to extract

    values for the parameters involved in their model. They compared the experimental results

    with those obtained using their model and found them to be in good agreement. However,

    they did not give an explanation for observing a hardening rather than a softening behavior.

    From the aforementioned overview, we note that the literature lacks a comprehensive

    model that predicts the dynamic response of a microbeam to an electric excitation over general

    operating conditions. Most existing models are formulated with a pre-assumed behavior basedon experimental observations, which limits the analysis to specific design parameters only.

    For instance, the nonlinearity of the system is assumed to be cubic and positive to justify

    the observed hardening behavior whereas the nonlinearity of the electric forces, which tends

    to yield softening behavior, is ignored. In some models, the system nonlinearity is totally

    ignored, thereby restricting the analysis to small values of AC voltages and DC polarization

    voltages. Further, in many cases, the microbeam is modeled as a single-degree-of-freedom

    system, thereby neglecting its mass distribution. The electric force is included incompletely

    in some papers by neglecting one of its components. In other cases, the dependence of the

    electric forces on the variable gap is ignored. In addition, most of the previous works do not

    recognize the fact that a deflected microbeam caused by DC polarization voltages has natural

    frequencies different from those of a straight beam; thus, the models are valid only for very

    low DC polarization voltages.

    In a previous work [7], we introduced a nonlinear model for a microbeam, which accounts

    for the initial deflection due to a DC electrostatic force, mid-plane stretching, and an applied

    axial load. The model does not account for the effects of the shear deformation and rotary

    inertia. We analyzed the static deflection of the microbeam and investigated its linear natural

    frequencies and mode shapes when actuated by a DC electrostatic force. Here, we expand on

    that work by considering the microbeam response to a general electric load composed of both

    DC and AC components. We model the microbeam as a distributed-parameter system. The

    equations are nondimensionalized and the design parameters of the resonator are lumped into

    nondimensional parameters, which are then used to develop a general equation that describes

    the dynamics of the microbeam. We apply a perturbation method, the method of multiple

    scales, to the general equation of motion and its associated boundary conditions, to obtainan approximation of the response of the microbeam to a primary-resonance excitation. We

    derive equations that describe the nonlinear resonance frequency, the amplitudes of periodic

    solutions, and the stability of these solutions. We study the effect of the design parameters on

    the nonlinear resonance frequency and the effective nonlinearity of the system. We compare

    the theoretical results with available experimental data. Then, we investigate the possibility

    of activating a three-to-one internal resonance between the first and second modes, which is

  • 7/27/2019 Resonant micro beam

    4/27

    94 M. I. Younis and A. H. Nayfeh

    Figure 1. A schematic drawing of a resonant microbeam.

    the most likely case to occur during the normal operation of the resonator. The absence of

    nonlinear interactions is necessary for a stable and reliable device.

    2. Problem Formulation

    We consider a resonant microbeam (Figure 1) actuated by an electric load composed of a

    DC component (polarization voltage) Vp and an AC component v(t) and subject to a viscous

    damping c per unit length.

    We model the microbeam as a plate undergoing cylindrical bending under an applied load,which is constant along the width of the plate. Typical dimensions of resonators violate the

    basic assumptions of a beam model; the ratio of a typical microbeam length to width b is

    < 10. The plate is clamped across its width and free across its long ends. We assume that the

    transverse deflection w(x,t) of the plate is constant along the width, thus reducing the plate

    static deflection equation to [17]

    EI

    1 2 4w

    x4=

    EA

    2(1 2)

    0

    w

    x

    2dx + N

    2w

    x2, (1)

    where x is the position along the plate length, A and I are the area and moment of inertia

    of the cross section (A = bh and I = (1/12)bh

    3

    , where h is the microbeam thickness), Eis Youngs modulus, is Poissons ratio, and N is the applied tensile axial force. The plate

    boundary conditions are

    w(0, y) = w(, y) = 0,w(0, y)

    x=

    w(,y)

    x= 0. (2)

    Using the modified beam equation, we rewrite the equation of motion governing the transverse

    deflection of the microbeam under electric forces as

    EI

    1 24w

    x4+ A

    2w

    t2+ c

    w

    t=

    EA

    2(1 2)

    0

    w

    x

    2dx + N

    2w

    x2

    +1

    20r b

    Vp + v(t)

    2

    (d w)2, (3)

    where w(x,t) is the deflection of the microbeam, t is time, is the material density, d is the

    gap width, 0 is the dielectric constant of vacuum, and r is the relative dielectric constant of

    the gap medium to that of an air gap. The last term in Equation (3) represents the parallel-

    plate electric forces [18] assuming a complete overlapping area between the microbeam and

  • 7/27/2019 Resonant micro beam

    5/27

    Nonlinear Response of Resonant Microbeams to Electric Actuations 95

    the stationary electrode. The microbeam is assumed to be clamped at both ends; hence the

    boundary conditions are

    w(0, t) = w(,t) = 0,w(0, t)

    x=

    w(,t)

    x= 0. (4)

    For convenience, we introduce the nondimensional variables (denoted by stars)

    w =w

    d, x =

    x

    , t =

    t

    T, (5)

    where T is a time scale, which is defined below. Substituting Equations (5) into Equations (3)

    and (4) and dropping the stars, we obtain

    4w

    x4+

    2w

    t2+ c

    w

    t= [1(w,w) + N]

    2w

    x2+ 2

    (Vp + v(t))2

    (1 w)2, (6a)

    w(0, t) = w(1, t) = 0,w(0, t)

    x=

    w(1, t)

    x= 0, (6b)

    where

    (f1(x,t),f2(x,t)) =

    10

    f1

    x

    f2

    xdx. (6c)

    Equation (6a) is a nondimensional integral-partial-differential equation with linear and non-

    linear terms as well as external and parametric excitation terms. The parameters appearing in

    Equation (6a) are

    c =c4(1 2)

    EI T

    , 1 = 6dh

    2

    , N =N 2(1 2)

    EI

    , 2 =60r

    4(1 2)

    Eh3d3(7)

    and T is chosen as

    T =

    bh4(1 2)

    EI

    1/2.

    Typically, T is in the ranges of microseconds. The ranges of the rest of the nondimensional

    parameters are c 100300, 1 0.110, N 10100, and 2V2

    p 0.150.

    The microbeam deflection under an electric force is composed of a static component due to

    the DC voltage, denoted by ws (x), and a dynamic component due to the AC voltage, denoted

    by u(x,t); that is,

    w(x,t) = ws (x) + u(x, t). (8)

    To calculate the static deflection of the microbeam, we set the time derivatives and the AC

    forcing term in Equation (6a) equal to zero and obtain

    d4ws

    dx4= [1(ws , ws ) + N]

    d2ws

    dx2+

    2V2

    p

    (1 ws )2

    . (9a)

  • 7/27/2019 Resonant micro beam

    6/27

    96 M. I. Younis and A. H. Nayfeh

    It follows from Equations (6b) and (8) that the boundary conditions are

    ws = 0 anddws

    dx= 0 at x = 0 and x = 1. (9b)

    We generate the problem governing the dynamic behavior of the microbeam around the

    deflected shape by substituting Equation (8) into Equations (6) and using Equations (9) to

    eliminate the terms representing the equilibrium position. To third-order in u, the result is 2u

    t2+ c

    u

    t+

    4u

    x4= (1(ws , ws ) + N )

    2u

    x2+ 21(ws , u)

    d2ws

    dx2

    +22V

    2p

    (1 ws )3

    u + 1(u, u)d2ws

    dx2+ 21(ws , u)

    2u

    x2

    +32V

    2p

    (1 ws )4u2 + 1(u, u)

    2u

    x2+

    42V2

    p

    (1 ws )5u3

    +22Vp

    (1 ws )2v(t) +

    42Vp

    (1 ws )3v(t)u + , (10a)

    u(0, t) = u(1, t) = 0, u(0, t)x

    = u(1, t)x

    = 0, (10b)

    where the term involving v2(t) is dropped because typically v2(t) V2p in resonant sensors.

    We end this section with two notes regarding the choice of the boundary conditions and the

    use of the linear damping in the model. First, the fixed-fixed end conditions are idealization

    to actual cases which may have some elasticity [5]. Some structures employ step-up type

    supports, which are not perfectly rigidly fixed [19]. According to [19], the flexibility of the

    supports reduces the critical value of the buckling load to be less than that for ideal fixed-

    fixed ends. Bouwstra and Geijselaers [5] mentioned that this flexibility makes the resonance

    frequencies lower than the predicted values for ideal fixed-fixed ends. The second note relates

    to the so-called squeeze-film damping. When a resonator moves too close to a stationary

    substrate, additional damping forces, the squeeze-film damping, increases due to the pressurebuilt up in the airgap [14]. Because typically resonant sensors are placed in a vacuum cavity in

    low pressure under 0.2 mbar, the resonance frequencies are almost independent of the pressure

    change [9, 15].

    3. Linear Mode Shapes and Corresponding Natural Frequencies

    We drop the nonlinear, forcing, and damping terms in Equation (10a) and obtain the linear

    undamped eigenvalue problem

    4u

    x4 (11 + N )

    2u

    x2

    22V2

    p

    (1 ws )3u 21(ws , u)

    d2ws

    dx2=

    2u

    t2, (11a)

    u = 0 andu

    x= 0 at x = 0 and x = 1, (11b)

    where 1 = (ws , ws ). We solve Equations (11) for the undamped mode shapes and natural

    frequencies under the static deflection distribution ws (x). Assuming a harmonic motion in the

    nth mode shape

    u(x,t) = n(x) ein t, (12)

  • 7/27/2019 Resonant micro beam

    7/27

    Nonlinear Response of Resonant Microbeams to Electric Actuations 97

    where n(x) is the nth mode shape and n is the nth nondimensional natural frequency, we

    reduce Equations (11) to

    ivn (11 + N )n

    22V2

    p

    (1 ws )3n 212w

    s =

    2nn, (13a)

    n = n = 0 at x = 0 and x = 1, (13b)

    where 2 = (ws , n) and the prime denotes the derivative with respect to x.

    4. Forced Vibration: Single-Mode Approximation

    4.1. PERTURBATION ANALYSIS

    We investigate the nonlinear vibrations of a clamped-clamped microbeam subject to an electric

    load composed of DC and AC components. We analyze its nonlinear response to a primary-

    resonance excitation of its first mode because it is the case that is used in resonator applica-

    tions. Still, the analysis is general and can be used to study nonlinear responses of the beam

    to a primary resonance of any of its mode.

    There are at least two approaches for determining the response of a distributed-parameter

    system with quadratic and cubic nonlinearities. In the first approach, called discretization

    approach, one discretizes the governing integral-partial-differential equation and associated

    boundary conditions, keeps enough modes in the discretization, and uses a perturbation method

    to attack the discretized system [20]. In the second approach, called the direct approach, one

    attacks directly the distributed-parameter problem. In this paper, we attack Equations (10) dir-

    ectly using the method of multiple scales [20, 21] to determine a uniformly valid approximate

    solution. To this end, we seek a second-order uniform solution of Equations (10) in the form

    u(x,t; ) = u1(x,T0, T2) + 2

    u2(x,T0, T2) + 3

    u3(x,T0, T2) + , (14)

    where is a small nondimensional bookkeeping parameter, T0 = t, and T2 = 2t. We note

    that u(x,t) does not depend on T1 = t because, as shown below, secular terms arise at

    O( 3). In order that the nonlinearity balances the effects of the damping c and excitation v(t),

    we scale them so that they appear together in the modulation equations as follows: 2c and

    3VAC cos(t) where VAC is the magnitude of the applied AC voltage and is the excitation

    frequency. Substituting Equation (14) into Equation (10a) and equating coefficients of like

    powers of, we obtain

    order

    L(u1) = D20 u1 + u

    iv1 11u

    1 N u

    1 21(ws , u1)w

    s

    22V2p

    (1 ws )3u1 = 0; (15)

    order 2

    L(u2) = 1ws (u1, u1) + 21(ws , u1)u

    1 +

    32V2

    p

    (1 ws )4

    u21; (16)

  • 7/27/2019 Resonant micro beam

    8/27

    98 M. I. Younis and A. H. Nayfeh

    order 3

    L(u3) = 2D0D2u1 cD0u1 + 21ws (u1, u2) + 21(ws , u2)u

    1 + 21(ws , u1)u

    2

    + 1(u1, u1)u1 + 2F(x) cos(T0) +

    62V2

    p

    (1 ws )4u1u2 +

    42V2

    p

    (1 ws )5u31, (17)

    where

    F(x) =2VpVAC

    (1 ws )2.

    The boundary conditions at all orders are given by

    uj = 0 and uj = 0 at x = 0 and x = 1, j = 1, 2, 3. (18)

    The first-order problem given by Equations (15) and (18) is identical to the linear ei-

    genvalue problem, Equations (11). Because in the presence of damping the homogeneous

    solutions corresponding to all modes that are not directly or indirectly excited decay with

    time [21], the solution of Equations (15) and (18) is assumed to consist of only the directly

    excited mode. Accordingly, we express u1

    as

    u1 = A(T2) eiT0 (x) + A(T2) e

    iT0 (x), (19)

    where A(T2) is a complex-valued function that is determined by imposing the solvability

    condition at third order, the overbar denotes the complex conjugate, and and (x) are the

    natural frequency and corresponding eigenfunction of the considered mode, respectively. The

    eigenfunction (x) is normalized such that1

    02 dx = 1.

    Substituting Equation (19) into Equation (16), we obtain

    L(u2) = (A2 e2iT0 + 2AA + A2 e2iT0 )h(x), (20)

    where

    h(x) = 1(,)ws + 21

    (ws , ) + 32V

    2

    p

    (1 ws )42.

    The solution of Equations (20) and (18) can be expressed as

    u2 = 1(x)A2 e2iT0 + 22(x)AA + 1(x)A

    2 e2iT0 , (21)

    where 1 and 2 are the solutions of the boundary-value problems

    M(i , 21i ) = h(x), (22)

    j = 0 and j = 0 at x = 0 and x = 1, j = 1, 2 (23)

    and ij is the Kronecker delta and the linear differential operator M(,) is defined as

    M(,) = iv 2 + 11 N 21(ws ,)w

    s

    22V2

    p

    (1 ws )3. (24)

    To describe the nearness of the excitation frequency to the fundamental natural fre-

    quency , we introduce the detuning parameter defined by

    = + 2. (25)

  • 7/27/2019 Resonant micro beam

    9/27

    Nonlinear Response of Resonant Microbeams to Electric Actuations 99

    Substituting Equations (19), (21), and (25) into Equation (17) and keeping the terms that

    produce secular terms only, we obtain

    L(u3) =

    i (2A + cA)(x) + (x)A2A + F(x) ei T2

    eiT0 + cc + NST, (26)

    where A denotes the derivative of A with respect to T2, cc denotes the complex conjugate of

    the preceding terms, and NST stands for terms that do not produce secular terms. The function

    (x) is defined as

    (x) = Gq (x) + Gc (x) +

    Eq (x) +

    Ec (x),

    where

    Gq (x) = 21ws (1, ) + 41w

    s (2, ) +

    21

    1 + 41

    2

    (ws , )

    + 21(ws , 1) + 41(ws , 2),

    Gc (x) = 31(,),

    Eq (x) =62V

    2p

    (1 ws )4

    (22 + 1) ,

    Ec (x) =42V

    2p

    (1 ws )53.

    The subscripts q and c denote quadratic and cubic nonlinear terms and the superscripts G and

    E denote terms produced by the geometric and electric nonlinearities.

    For a uniform second-order approximation, we do not need to solve for u3. Only, we need to

    impose the solvability condition, which provides an equation for the function A. Because the

    homogeneous problem governing u3 has a nontrivial solution, the corresponding nonhomo-

    geneous problem has a solution only if the right-hand side of Equation (26) is orthogonal

    to every solution of the adjoint homogeneous problem governing u3 [20]. We note that our

    problem is self adjoint, the adjoints are given by (x) eiT0 . Multiplying the right-hand side

    of Equation (26) with (x) eiT0 and integrating the result from x = 0 to x = 1, we obtain

    the following solvability condition:

    2i(D2A + A) + 8SA2A F ei T2 = 0, (27)

    where

    S = SGq + SGc + S

    Eq + S

    Ec ,

    SGq = 1

    8

    10

    Gq dx,

    SGc = 1

    8

    10

    Gc dx,

    SEq = 1

    8

    10

    Eq dx,

  • 7/27/2019 Resonant micro beam

    10/27

    100 M. I. Younis and A. H. Nayfeh

    SEc = 1

    8

    10

    Ec dx,

    =1

    2

    1

    0

    c2 dx,

    F =

    10

    F dx.

    Here, S is the effective nonlinear coefficient of the system and Smn denotes a nonlinear coeffi-

    cient of source m (electric or geometric) and of order n (quadratic or cubic). Ifc is independent

    ofx, then = c/2. In general, c is a function ofx.

    Next, we express A in the polar form A = (1/2)a ei , where a and are real-valuedfunctions, representing, respectively, the amplitude and phase of the response. Substituting

    for A in Equation (27), separating the real and imaginary parts, and letting = T2 , we

    obtain the following modulation equations:

    a = a +F

    sin , (28)

    a = a Sa3

    +

    F

    cos . (29)

    Substituting Equations (19) and (21) into Equation (14) and setting = 1, we express, to the

    second approximation, the microbeam response to the external excitation as

    u(x,t) = a cos(t )(x) +1

    2 a2

    2(x) + cos2(t )1(x)

    + , (30)

    where a and are governed by Equations (28) and (29).

    It follows from Equation (30) that periodic solutions of Equations (10) correspond to

    constant a and ; that is, the fixed points (a0, 0) of Equations (28) and (29). Thus, letting

    a = = 0 in Equations (28) and (29) and eliminating 0 from the resulting equations, we

    obtain the following frequency-response equation:

    a20

    2 +

    Sa 20

    2=

    F2

    2(31)

    Equation (31) is an implicit equation for the amplitude a of the periodic response as a func-tion of the detuning parameter (which is a representation of the excitation frequency), the

    effective nonlinear coefficient S, the damping coefficient , and the amplitude of excitation

    F. Solving Equation (31) for , we obtain

    = 1

    a0

    F2

    2 2a20

    1/2+

    Sa 20

    . (32)

  • 7/27/2019 Resonant micro beam

    11/27

    Nonlinear Response of Resonant Microbeams to Electric Actuations 101

    Recalling that = 2, setting = 1, and noting that the amplitude a is maximum

    when the expression inside the square root vanishes, we obtain the following equation for the

    nonlinear resonance frequency:

    r = +SF2

    32. (33)

    The parameters is related to the quality factor Q by

    =

    2Q(34)

    Substituting Equation (34) into Equation (33), we obtain

    r = +4SQ2

    5F2. (35)

    Equation (35) relates the nonlinear resonance frequency r to the effective nonlinearity S

    of the system, the amplitude F of the AC forcing, and the quality factor Q. Equation (35) can

    be used to predict the resonance frequency for excitation frequencies much less than twice the

    fundamental natural frequency. This because when 2 a principal parametric resonanceand a subharmonic resonance of order one-half will be activated.

    The stability of the fixed points (a0, 0) is determined by examining the eigenvalues of the

    Jacobian matrix of Equations (28) and (29) evaluated at the corresponding fixed point [22].

    The characteristic equation is

    2 + 2 +

    2 +

    3Sa 20

    Sa 20

    = 0. (36)

    For asymptotically stable solutions, all of the eigenvalues must be in the left-half plane.

    4.2. RESULTS

    To describe the dynamic response of the microbeam, we need to determine the natural fre-

    quency , the excitation amplitude F, the effective nonlinearity of the system S, and the

    damping coefficients or Q.

    As a first step, we solved the boundary-value problem, Equations (9), for the deflection

    due to the DC electrostatic force. We did this iteratively using a shooting method to de-

    termine the value of the integral (ws , ws ). Using the converged solution ws , we solved

    the boundary value problem, Equations (13), for the fundamental natural frequency and its

    corresponding eigenfunction. We used a shooting method to iterate on 2, , and until they

    converged to within a predefined tolerance. Next, we solved the boundary-value problems,

    Equations (2224) to evaluate the functions 1 and 2 using a method similar to that used to

    solve Equations (13). Finally, we evaluated the s, Ss, and F.

    We begin by comparing the results obtained by using the present model with the experi-mental results of Tilmans and Legtenburg [9] and Gui et al. [10]. Results from other papers,

    which are mentioned in the introduction, are not used because either they were produced from

    another device configuration [11, 12], for which our model does not apply, or the authors did

    not give enough data about the specifications of their devices [8, 13].

    In Figure 2, we compare the normalized nonlinear resonance frequency r / obtained

    using Equation (35) with those obtained theoretically and experimentally by Tilmans and

  • 7/27/2019 Resonant micro beam

    12/27

    102 M. I. Younis and A. H. Nayfeh

    Figure 2. Comparison of the normalized nonlinear resonance frequency r / calculated using our model (solid

    lines) with those obtained theoretically (dashed lines) and experimentally by Tilmans and Legtenburg [9] for two

    microbeams of lengths 210 m (diamonds) and 310 m (circles). The thin solid lines are based on the quality

    factors of Tilmans and Legtenburg [9] and the thick solid lines are based on quality factors estimated using our

    model.

    Legtenburg [9] for two microbeams of lengths 210 and 310 m with h = 1.5 m, b =

    100 m, d = 1.18 m, and subject to an axial load of 0.0009 Newton. The reported pull-

    in voltages for the 210 and 310 m microbeams are 28 and 13.8 V, respectively. The DC

    polarization voltage Vp = 1 V for data points of the first microbeam except for the first two

    points where Vp = 2 V. For the second microbeam, Vp = 2 V for all data points. The theory

    of Tilmans and Legtenburg [9] is based on Rayleighs energy method modified to account for

    electric forces and mid-plane stretching. The latter effect however was misrepresented, as we

    show in Appendix A.

    We show two sets of calculated results for each microbeam. The first set, shown as thin

    solid lines, uses the same values of Q reported by Tilmans and Legtenburg [9]. They extracted

    the values of Q analytically using an equation derived from their model. The parameters

    of this model were obtained by fitting the predicted frequency-response curve at a low DC

    voltage to the one obtained experimentally. These quality factors are Q = 592 and 151 for

    the 210 and 310 m microbeams, respectively. As shown in Figure 2, these values give poor

    results. Legtenburg and Tilmans [16] mentioned that the quality factor for the design of their

    device varies across the wafer due to variations in the sealing pressure of the microbeam

    encapsulation. Because such variations can lead to a wrong measurement ofQ and due to thedifficulty of measuring the system damping, in general, we determined the quality factors by

    matching r / obtained using Equation (35) at VAC = 0.6 V to the experimental value of

    Tilmans and Legtenburg [9].

    We obtained Q = 816.6 and 197 for the 210 and 310 m microbeams, respectively. The

    results obtained using these values, shown as thick solid lines, are in excellent agreement with

    the experimental results.

  • 7/27/2019 Resonant micro beam

    13/27

    Nonlinear Response of Resonant Microbeams to Electric Actuations 103

    Figure 3. Comparison of the hysteretic points obtained using our model (solid lines) with those obtained theoret-

    ically (dashed line) and experimentally (circles) by Gui et al. [10]. The thin solid line is based on the quality factor

    Q = 900 reported by Gui et al. [10] and the thick solid line is based on a quality factor Q = 1000 estimated using

    our model.

    In Figure 3, we show another comparison for the hysteretic points (points correspond to

    the largest AC and DC voltages below which the response is single-valued and above which it

    is multi-valued) predicted by our model with those obtained theoretically and experimentally

    by Gui et al. [10] for a microbeam of length 210 m and h = 1.5 m, b = 100 m, d =

    1 m, and Q = 900. The theoretical results of Gui et al. [10] were obtained using a modified

    Rayleighs energy method that accounts for mid-plane stretching and electric forces. Mid-

    plane stretching however was misrepresented, as we show in Appendix A.

    In Figure 3, we show also two curves obtained using our theory corresponding to two dif-

    ferent values of Q: the thick solid line was obtained using Q = 1000, which was determined

    by matching our result at Vp = 4 V to the experimental result of Gui et al. [10]. The thin solid

    line was obtained using Q = 900, which is the value used by Gui et al. [10]. We estimated the

    axial force to be N = 0.00011 Newton by fitting the theoretically obtained natural frequency

    to the experimental value obtained by Gui et al. [10]. There is excellent agreement between

    our results and the experimental results.

    Figures 4 and 5 show the frequency-response curves (amplitude and phase responses)corresponding to Vp = 8 V in Figure 3. In both figures, thin lines are for VAC = 0.007 V

    and thick lines are for VAC = 0.03 V. The unstable regions in the latter case are shown as

    dashed lines. The results are obtained by solving Equations (31) and (36).

    Next, we study the effect of the nondimensional design parameters on the nonlinear reson-

    ance frequency. As indicated by Equation (35), the nondimensional design parameters affect

    the nonlinear resonance frequency in two ways: changing the effective nonlinearity of the

  • 7/27/2019 Resonant micro beam

    14/27

    104 M. I. Younis and A. H. Nayfeh

    Figure 4. The amplitude of the dynamic response versus the frequency of excitation corresponding to Vp = 8 V

    in Figure 3.

    Figure 5. The phase of the dynamic response versus the frequency of excitation corresponding to Vp = 8 V in

    Figure 3.

  • 7/27/2019 Resonant micro beam

    15/27

    Nonlinear Response of Resonant Microbeams to Electric Actuations 105

    Figure 6. Variation of the normalized nonlinear resonance frequency r / with VAC for various values of the

    nondimensional axial load N. The values of1, 2V2

    p , and Q are 3.7, 5.5, and 796, respectively.

    system S and the fundamental natural frequency . We refer to Younis et al. [7] for the pull-in

    voltages of the numerical data used in the following figures.

    In Figure 6, we show the effect of varying the driving voltage AC on r / for various val-

    ues of the nondimensional axial load N. Increasing the driving voltage amplitude AC increase

    r /. On the other hand, increasing the axial force N at a constant VAC decreases r / and

    also increases its linear behavior. We note that negative values of N, which correspond to

    compressive loads, have a greater effect on r /.

    In resonant sensors, residual stresses are typically of the tensile type. They are inducedduring the fabrication process. Legtenberg and Tilmans [16] mentioned that such stresses are

    desirable to reduce the possibility of buckling. However, because resonant sensors may exhibit

    temperature variations, which may induce compressive residual stresses, operation away from

    the critical buckling load should be ensured. As an example, we calculated a critical buckling

    load Pcr = 13.3037 for the microbeam in Figure 2 of length 210 m using the formula

    Pcr =4 Eh3

    12L2(1 2)

    for a fixed-fixed beam.

    In Figure 7, we show the effect of varying the driving voltage amplitude AC on r / for

    various values of 1. It can be seen that, for values of 1 greater than 0.05, increasing ACincreases r /. In contrast, for values of1 less than 0.05, increasing AC leads to a decrease

    in r /. We also note that increasing 1 increases the nonlinear behavior of r / and its

    value at a constant value of AC.

    In Figure 8, we show the effect of varying AC on r / for various values of 2V2

    p . In-

    creasing 2V2

    p at a constant VAC increases r / up to a value of 2V2

    p 35. Above this

    value, the increase in r / becomes smaller until it is reversed completely at 2V2

    p 41,

  • 7/27/2019 Resonant micro beam

    16/27

    106 M. I. Younis and A. H. Nayfeh

    Figure 7. Variation of the normalized nonlinear resonance frequency r / with VAC for various values of 1.

    The values ofN, 2V2

    p , and Q are 8.7, 5.5, and 796, respectively.

    Figure 8. Variation of the normalized nonlinear resonance frequency r / with VAC for various values of2V2

    p .

    The values of1, N, and Q are 3.7, 8.7, and 796, respectively.

  • 7/27/2019 Resonant micro beam

    17/27

    Nonlinear Response of Resonant Microbeams to Electric Actuations 107

    Figure 9. Variation of the nonlinear coefficients with the axial load N. The values of 1 and 2V2

    p are 3.7 and

    5.5, respectively.

    where the nonlinear resonance frequency becomes highly sensitive to changes in the DC

    polarization voltage. In fact, a slight increase in 2V2

    p beyond this value leads to a sharp drop,

    which can even reach zero, in the nonlinear resonance frequency with increasing VAC. We

    note that such a strange behavior occurs before reaching the pull-in limit, which in this case

    occurs at 2V2

    p 90. This result shows that the nondimensional parameter 2V2

    p changes

    the dynamic response of the microbeam from a hardening to a softening response. This is

    because, for this range of2V2

    p , the electrostatic force, which tends to lower the resonance fre-

    quency, drastically dominates the mid-plane stretching, which tends to increase the resonancefrequency.

    4.3. DISCUSSION

    In order to better understand how the nondimensional parameters N, 1, and 2V2

    p affect

    the effective nonlinearity of the system, we study the influence of each parameter on the

    nonlinear coefficients SGq , SGc , S

    Eq , S

    Ec , and S. In Figure 9, we show variation of these nonlinear

    coefficients with increasing the axial load N. It shows that increasing N in the positive range

    leads only to a slight linear decrease in the mid-plane stretching represented by SGc , and hence

    in the effective nonlinear coefficient S. However, this is not the major factor that leads to the

    decrease in r / with increasing N, which is shown in Figure 6. Increasing N also increases

    , which according to Equation (35) decreases the shift in r /. This agrees with the resultsof Figure 6, where the ratio r / approaches unity for large N.

    In Figure 10, we show variation of the nonlinear coefficients with 1. Increasing 1 in-

    creases linearly the mid-plane stretching coefficient SGc , without affecting the other nonlinear

    coefficients. The increase in SGc , and so in S, increases r /, as indicated by Equation (35).

    On the other hand, increasing 1 increases , which tends to decrease r /. The effect of

    increasing S however dominates the later effect. This explains the increase in the normalized

  • 7/27/2019 Resonant micro beam

    18/27

    108 M. I. Younis and A. H. Nayfeh

    Figure 10. Variation of the nonlinear coefficients with 1. The values ofN and 2V2

    p are 8.7 and 5.5, respectively.

    nonlinear resonance frequency with increasing 1 at a constant VAC, which is seen in Figure 7.

    The effective nonlinear coefficient S starts with a negative value for very small values of 1,

    which explains the qualitative change observed in Figure 7 near this range of values.

    In Figure 11, we show variation of the nonlinear coefficients with 2V2

    p . For small values

    of2V2

    p , the effective nonlinear coefficient Sdecreases with increasing 2V2

    p up to values near

    40 where the quadratic electric term SEc starts to decrease sharply, thereby dominating all other

    nonlinear coefficients. At 2V2

    p

    41, the sign of S changes from positive, corresponding to

    a hardening behavior, to negative, corresponding to a softening behavior, which explains the

    qualitative change in the dynamic behavior observed in Figure 10. This is a significant result;

    it shows that using a spring-mass model with a cubic nonlinearity, which is the model usually

    used in the literature, is inaccurate and might lead to wrong results. Such models neglect the

    quadratic nonlinearity, which is due to the electrostatic force and the static deflection. As

    clearly shown in Figure 11, this nonlinearity becomes dominant and governs the dynamic

    behavior beyond a critical value of2V2

    p . Further, we note that assuming a single value for the

    effective nonlinearity in the model may introduce another source of error because, as shown

    in Figure 11, the effective nonlinearity is a strong function of the electrostatic force.

    We note that the DC voltage needed to achieve a linear behavior in the nonlinear resonance

    frequency (corresponding to S = 0) is relatively high (Vp 19V for the data of Figure 2)

    and may not be attainable in resonant sensor applications. This result is in agreement with thatfound by Turner and Andrews [11].

    Next, we examine the sensitivity of the dynamic response of the microbeam to the quality

    factor. In Figures 12 and 13, we show variation of the amplitude of the response a with the

    detuning parameter for a microbeam with 1 = 3.7 and N = 8.7 and quality factors ofQ =

    200, Q = 500, and Q = 796. We use 2V2

    p =15 and 45 in Figures 12 and 13, respectively.

    Both figures show that increasing the quality factor amplifies the effect of the nonlinearity and

  • 7/27/2019 Resonant micro beam

    19/27

    Nonlinear Response of Resonant Microbeams to Electric Actuations 109

    Figure 11. Variation of the nonlinear coefficients with 2V2

    p . The values of1 and N are 3.7 and 8.7, respectively.

    Figure 12. Variation of the response amplitude a of a microbeam with the detuning parameter for 1 = 3.7,

    N = 8.7, and 2V2

    p =15.

  • 7/27/2019 Resonant micro beam

    20/27

    110 M. I. Younis and A. H. Nayfeh

    Figure 13. Variation of the response amplitude a of a microbeam with the detuning parameter for 1 = 3.7,

    N = 8.7, and 2V2

    p =45.

    can change the dynamic state from a single-valued response, as in the Q = 200 curves, to a

    multi-valued response, as in the Q = 500 and 796 curves. As can be noted from Figure 11,

    the response curves exhibit a hardening behavior in Figure 12 and a softening behavior in

    Figure 13.

    5. Three-to-One Internal Resonance

    5.1. PERTURBATION ANALYSIS

    In this section, we consider modal interactions among the microbeam modes involving the

    first mode. It follows from Figure 14 that 1 is away from (1/2)n for any n. Hence, a two-

    to-one internal resonance between 1 and n cannot be activated. However, 1 (1/3)2 for

    some range of Vp, and hence we study the possibility of activating a 1:3 internal resonance

    between the first and second modes when the first mode is excited with a primary resonance.

    We apply the method of multiple scales directly to Equations (10). The analysis presen-

    ted here is general and applicable to any two modes whose frequencies are in the ratio of

    three-to-one. We seek a solution of Equations (10) in the form of Equation (14) and obtain

    Equations (1518). Because in the presence of damping all modes that are not directly or

    indirectly excited decay with time [16], the solution of Equations (15) and (18) is assumed toconsist of the two interacting modes; that is,

    u1 = An(T2) einT0 n(x) + Am(T2) e

    imT0 m(x) + cc, (37)

  • 7/27/2019 Resonant micro beam

    21/27

    Nonlinear Response of Resonant Microbeams to Electric Actuations 111

    Figure 14. Variation of the first four natural frequencies, calculated by Younis et al. [7], of a microbeam with

    2V2

    p for 1 = 3.7 and various applied axial loads.

    where m and n denote the two modes being considered and the j are normalized such that10

    2j dx = 1. Substituting Equation (37) into Equation (16) yields

    L(u2) = A2n e

    2inT0 h1n(x) + A2m e

    2imT0 h1m(x) + AnAnh1n(x) + AmAmh1m(x)

    + (AnAm ei(n+m)T0 + AnAm e

    i(nm)T0 )Hnm(x) + cc, (38)

    where

    h1i(x) = 1(i , i )ws + 21

    i (ws , i ) +

    32V2

    p

    (1 ws )42i ,

    Hij(x) = 21(i , j)ws + 41

    i (ws , j) +

    62V2

    p

    (1 ws )4i j,

    and the functional is defined in Equation (6c).

    We express the solution of Equations (38) and (18) as

    u2 = 1n(x)A2

    n

    e2inT0 + 1m(x)A2

    m

    e2imT0 + 3(x)AnAm ei(n+m)T0

    + 4(x)AnAm ei(nm)T0 + 2n(x)AnAn + 2m(x)AmAm + cc, (39)

    where the j and ij are the solutions of the boundary-value problems

    M(1i , 2i ) = h1i (x), (40a)

    M(2i , 0) = h1i (x), (40b)

  • 7/27/2019 Resonant micro beam

    22/27

    112 M. I. Younis and A. H. Nayfeh

    M(3, n + m) = Hnm(x), (40c)

    M(4, n m) = Hnm(x). (40d)

    The operator M(,) is defined in Equation (24). The boundary conditions for the j and

    ij are

    = 0 and = 0 at x = 0 and x = 1. (41)

    Substituting Equations (37) and (39) into Equation (17) and considering the case n 3mand either n or m, we obtain

    L(u3) =

    in(2An + cAn)n(x) + 1n(x)A

    2nAn + nm(x)AnAmAm

    ein T0

    +

    im(2Am + cAm)m(x) + 1m(x)A

    2mAm + mn(x)AmAnAn

    eimT0

    + 5A3m e

    3imT0 + 6AnA2m e

    i(n2m)T0 + F(x) eiT0 + cc + NST, (42)

    where Aj is the derivative ofAj with respect to T2 and

    1i = 21(1i , i )ws + 41(2i , i )w

    s + 21(ws , 1i )

    i + 41(ws , 2i )

    i

    + 31(i , i )i + 21(ws , i )

    1i + 41(ws , i )

    2i +

    122V2

    p

    (1 ws )5

    3i

    +62V

    2p

    (1 ws )4(ii i + 2ii ),

    ij = 21(3, j)ws + 21(4, j)w

    s + 41(i , 2jw

    s + 21(ws , 3)

    j

    + 21(ws , 4)j + 41(ws , 2j)

    i + 21(ws , j)

    3 + 21(ws , j)

    4

    + 41(ws , i )2j) + 21(j, j)

    i + 41(i , j)

    j +

    242V2

    p

    (1 ws )5

    i 2j

    +62V

    2p

    (1 ws )4(3j + 4j + 22ji ),

    5 = 21(1m, m)ws + 21(1m, ws )

    m + 21(ws , m)

    1m + 1(m, m)

    m

    +62V

    2p

    (1 ws )41mm,

    6 = 21(4, m)ws + 21(1m, n)w

    s + 21(4, ws )

    m + 21(1m, ws )

    n

    + 21(ws , m)4 + 21(n, ws )

    1m + 1(m, m)

    n + 21(n, m)

    m

    +62V

    2p

    (1 ws )44m +

    62V2

    p

    (1 ws )41mn +

    122V2

    p

    (1 ws )5n

    2m.

    To describe the nearness ofn to 3m and to either n or m, we introduce the detuning

    parameters 1 and 2 defined by

    n = 3m + 21 and = i +

    22. (43)

  • 7/27/2019 Resonant micro beam

    23/27

    Nonlinear Response of Resonant Microbeams to Electric Actuations 113

    Because the homogeneous part of Equations (42) and (18) has a nontrivial solution, the non-

    homogeneous problem has a solution only if the right-hand side of Equation (42) is orthogonal

    to every solution of the adjoint homogeneous problem governing u3 [20]. We note that the

    problem is self-adjoint, and hence the adjoints are given by j(x) eijT0 . Multiplying the

    right-hand side of Equation (42) by n(x) ein T0 and m(x) e

    im T0 , respectively, integrating

    the results from x = 0 to x = 1, and using Equations (43), we obtain the following solvability

    conditions:

    2in(D2An + nAn) = 8SnnA2nAn + 8SnmAnAmAm + 8nA

    3m e

    i1 T2

    + fn(x)in ei2T2 , (44)

    2im(D2Am + mAm) = 8SmmA2mAm + 8SmnAmAnAn + 8mAnA

    2m e

    i1T2

    + fm(x)im ei2T2 , (45)

    where

    i = 12

    10

    c2i dx,

    fi =

    10

    F i dx,

    Sii =1

    8

    1o

    1i i dx, Sij =1

    8

    1o

    iji dx, i = j,

    n =1

    8

    1o

    5n dx, m =1

    8

    1o

    6m dx.

    The Sij are nonlinear coefficients due to electric and geometric sources, and the i are the

    nonlinear interaction coefficients between the nth and mth modes.

    Expressing the Ai in polar form and separating real and imaginary parts in Equations (44)

    and (45), we obtain the modulation equations

    an = nan na

    3m

    nsin 1 +

    fnin

    nsin 2, (46)

    an

    n

    = Snna

    3n

    n

    Snmana2m

    n

    na3m

    n cos

    1

    fnin

    n cos

    2,

    (47)

    am = mam mana

    2m

    msin 1 +

    fmim

    msin 2, (48)

    amm =

    Smma3m

    m

    Smnama2n

    m

    mana2m

    mcos 1

    fmim

    mcos 2, (49)

  • 7/27/2019 Resonant micro beam

    24/27

    114 M. I. Younis and A. H. Nayfeh

    where 1 and 2 are defined as

    1 = 1T2 3m + n, 2 = 2T2 in n im m.

    Consequently, the microbeam dynamic response to second order can be expressed as

    u(x,t) = an cos(nt + n)n(x) + am cos(mt + m)m(x)

    +1

    2a2n [cos 2(nt + n)in(x) + 2n(x)]

    +1

    2a2m [cos 2(mt + m)im (x) + 2m(x)]

    +1

    2anam[cos((n + m)t + n + n)3(x)

    + cos((n m)t + n m)4(x)] + , (50)

    where an, am, n, and m are governed by Equations (4649).

    5.2. RESULTS

    We considered the case in which the first mode is directly excited with a primary resonance

    (fn = 0) when 2 31. We solved the boundary-value problem, Equations (9), for the static

    deflection. Then using this deflection, we solved Equations (13) to determine the first and

    second natural frequencies and mode shapes. Using these results, we solved Equations (40)

    and (41) for the s and then evaluated the s. Finally, we calculated the Sij, i , and fm.

    We considered three cases. The first case corresponds to 1 = 3.70, 2V2

    p = 35.30, N = 0,

    and the first and the second natural frequencies are 1 = 20.35 and 2 = 61. The second case

    corresponds to 1 = 3.70, 2V2

    p = 5.50, N = 10, and the natural frequencies are 1 =

    19.12 and 2 = 57.70. The last case corresponds to 1 = 3.70, 2V2

    p = 56.80, and N = 10,and the natural frequencies are 1 = 21.60 and 2 = 64.20. For all three cases, the numerical

    results show that the nonlinear interaction coefficients i vanish identically, which precludesthe possibility of activating this internal resonance. Hence, although the ratio between the

    natural frequencies is close to three-to-one, the first and second modes do not exchange energy.

    This is because the first mode is symmetric, the second mode is antisymmetric, and the static

    deflection is symmetric. And hence, the interaction coefficient m and n are identically zero.

    6. Conclusions

    We studied the nonlinear dynamic response of a microbeam that is actuated by a general

    electric load subject to an applied axial load, accounting for mid-plane stretching. We used

    a model that assumes operation at low pressures where the effect of squeeze-film damping

    on the resonance frequencies can be neglected. The model does not account for the effectsof shear deformation and rotary inertia. We used the method of multiple scales to determine

    the response to a primary resonance excitation of the first mode and obtained two nonlinear

    first-order ordinary-differential equations governing the amplitude and phase of the response.

    We derived an equation that describes the nonlinear resonance frequency of the microbeam

    as a function of the damping and the effective nonlinearity coefficient. This equation shows

    that increasing the AC forcing and/or decreasing the damping leads to either an increase

  • 7/27/2019 Resonant micro beam

    25/27

    Nonlinear Response of Resonant Microbeams to Electric Actuations 115

    or a decrease in the nonlinear resonance frequency, depending on the sign of the effective

    nonlinearity coefficient. We compared the nonlinear resonance frequency computed using our

    theory to the experimental results available in the literature. We found excellent agreement.

    We studied the effect of the nondimensional design parameters N, 1, and 2V2

    p on the

    normalized nonlinear resonance frequency r /. The results show that increasing N, the

    axial force, improves the linear characteristics of r / and decreases the frequency shift.

    In contrast, increasing 1, the mid-plane stretching, has the reverse effect on r /. On the

    other hand, 2V2

    p affects the qualitative and quantitative nature of the effective nonlinearity

    coefficient of the system. The sign of S changes from positive, corresponding to a hardening

    behavior, to negative corresponding to a softening behavior. This is because the electric non-

    linearity, mainly the quadratic, drastically increases in magnitude and overcomes the influence

    of the geometric nonlinearity. We note that most of the models used in the literature neglect

    the effect of the electric nonlinearity and particularly the quadratic one. Instead, they assume

    the nonlinearity of the system to be solely cubic and positive, which predicts a harding beha-

    vior rather than the correct softening behavior. Therefore, failure to correctly account for the

    nonlinearities in the system may lead to erroneous results. It is interesting to note that, for the

    studied case, this reverse in behavior occurs at a value of 2V2

    p , which is about 46% of the

    pull-in limit. Because at this value the coefficient S is equal to zero, the dynamic response istheoretically linear. However, the simulation results show that r / becomes highly sensitive

    to any slight changes in 2V2

    p . This indicates that, although it is possible to compensate for the

    effect of the electrostatic force with mid-plane stretching, operations near such conditions are

    unstable and impractical. We believe more experimental work is needed to better understand

    the system behavior near this region.

    We applied the method of multiple scales to investigate possibility of activating a three-to-

    one internal resonance between the first and second modes, which if exists can adversely affect

    the performance of the resonator. The analysis shows that these two modes are nonlinearly

    uncoupled, and hence this internal resonance cannot be activated. This result enhances the

    reliability of such a device.

    In conclusion, the present nonlinear model provides an accurate prediction of the dynamic

    behavior of microbeams, which linear models fail to explain. Unlike existing models in the

    literature, the present nonlinear model is capable of simulating the mechanical behavior of

    microbeams for general operating conditions and for a wider range of applied electric loads.

    Further, by using the perturbation solution, one can easily derive analytical expressions that

    present a clearer picture of the influence of various design parameters, which is critical for

    improving and optimizing designs.

    Appendix A: Note on Energy Methods

    We consider the microbeam shown in Figure 1 with zero AC and DC forces. To obtain the

    equation that governs the microbeam motion, we set Vp and v(t) equal to zero in Equation (3),

    assume the same boundary conditions as in Equation (4), and obtainEI

    1 24w

    x4+ bh

    2w

    t2+ c

    w

    t

    =

    EA

    2(1 2)

    0

    w

    x

    2dx + N

    2w

    x2, (A.1a)

  • 7/27/2019 Resonant micro beam

    26/27

    116 M. I. Younis and A. H. Nayfeh

    w(0, t) = w(,t) = 0 andw

    x

    (0,t )

    =w

    x

    (,t)

    = 0. (A.1b)

    Equations (A.1) can be derived using the EulerLagrange equations from the Lagrangian

    L = T U, (A.2)

    where the potential U and kinetic T energies of the system are given by

    U =EI

    2(1 2)

    0

    2w

    x2

    2dx +

    1

    2N

    0

    w

    x

    2dx

    +EA

    8(1 2)

    0

    w

    x

    2dx

    2

    , (A.3a)

    T =

    1

    2 bh

    0

    wt

    2dx. (A.3b)

    The total energy of the system is given by

    H = U + T . (A.4)

    Studying the same problem, Tilmans et al. [14] misrepresented the potential energy due to

    mid-plane stretching by expressing it as

    Umid =EA

    8

    0

    w

    x

    4dx.

    The correct representation however is the last term in Equation (A.3a). The theories of Tilmans

    and Legtenburg [9] and Gui et al. [10] are based on this incorrect representation of the potential

    energy.

    References

    1. Goldsmith, C. L., Yao, Z., Eshelman, S., and Denniston, D., Performance of low-loss RF MEMS capacitive

    switches, IEEE Microwave and Guided Wave Letters 8, 1998, 269271.

    2. Chan, E. K., Garikipati, K., and Dutton, W. R., Characteristics of contact electromechancis through

    capacitance-voltage measurements and simulations, Journal of Microelectromechanical Systems 8, 1999,

    208217.

    3. Gupta, R. K. and Senturia, S. D., Pull-in time dynamics as a measure of absolute pressure, in Proceedingsof the IEEE Tenth Annual International Workshop on Microelectromechanical Systems: MEMS 97, Nagoya,

    Japan, IEEE, New York, 1997, pp. 5159.

    4. Andrews, M. K. , Turner, G. C., Harris, P. D., and Harris, I. M., A resonant pressure sensor based on a

    squeezed film of gas, Sensors and Actuators A36, 1993, 219226.

    5. Bouwstra, S. and Geijselaers, B., On the resonance frequencies of microbridges, in Proceedings of the 6th

    International Conference on Solid-State Sensors and Actuators (TRANSDUCERS 91), San Francisco, CA,

    IEEE, New York, 1997, Vol. 2, pp. 11411144.

  • 7/27/2019 Resonant micro beam

    27/27

    Nonlinear Response of Resonant Microbeams to Electric Actuations 117

    6. Younis, M. I., Abdel-Rahman, E. M., and Nayfeh, A. H., A reduced-order model for electrically actuated

    microbeam-based MEMS, Journal of Microelectromechanical Systems, to appear.

    7. Younis, M. I., Abdel-Rahman, E. M., and Nayfeh, A. H., Static and dynamic behavior of an electrically ex-

    cited resonant microbeam, in Proceedings of the AIAA 43rd Structures, Structural Dynamics, and Materials

    Conference, Denver, CO, 2002, AIAA Paper No. 2002-1305.

    8. Zook, J. D., Burns, D. W., Guckel, H., Sniegowski, J. J., Engelstad, R. L., and Feng, Z., Characteristics of

    polysilicon resonant microbeams, Sensors and Actuators A35, 1992, 290294.

    9. Tilmans, H. A. and Legtenberg, R., Electrostatically driven vacuum-encapsulated polysilicon resonators.

    Part II. Theory and performance, Sensors and Actuators A45, 1994, 6784.

    10. Gui, C., Legtenberg, R., Tilmans, H. A., Fluitman, J. H., and Elwenspoek, M.,Nonlinearity and hysteresis

    of resonant strain gauges, Journal of Microelectromechanical Systems 7, 1998, 122127.

    11. Turner, G. C. and Andrews, M. K., Frequency stabilization of electrostatic oscillators, in Digest of the 8th

    International Conference on Solid-State Sensors and Actuators, Vol. 2, Stockholm, Sweden, S. Middelhoek

    and K. Cammann (eds.), Elsevier, Amsterdam, 1995, Vol. 2, pp. 624626.

    12. Ayela, F. and Fournier, T., An experimental study of anharmonic micromachined silicon resonators,

    Measurement, Science and Technology 9, 1998, 18211830.

    13. Veijola, T., Mattila, T., Jaakkola, O., Kiihamki, J., Lamminmki, T., Oja, A., Ruokonen, K., Sep, H.,

    Seppl, P., and Tittonen, I., Large-displacement modeling and simulation of micromechanical electrostat-

    ically driven resonators using the harmonic balance method, in the IEEE MTT-S International Microwave

    Symposium Digest, Boston, MA, T. Perkins (ed.), IEEE, New York, 2000, Vol. 1, pp. 99102.

    14. Tilmans, H. A., Elwespoek, M., and Fluitman, J. H., Micro resonant force gauges, Sensors and ActuatorsA30, 1992, 3553.

    15. Gui, C., Legtenberg, R., Elwenspoek, M., and Fluitman, J. H., Q-factor dependence of one-port encapsulated

    polysilicon resonator on reactive sealing pressure, Journal of Micromechanics and Microengineering 5,

    1995, 183185.

    16. Legtenberg, R. and Tilmans, H. A., Electrostatically driven vacuum-encapsulated polysilicon resonators.

    Part I. Design and fabrication, Sensors and Actuators A45, 1994, 5766.

    17. Nayfeh, A. H., Nonlinear Interactions, Wiley, New York, 2000.

    18. Griffiths., D. J., Introduction to Electrodynamics, Prentice Hall, Englewood Cliffs, NJ, 1981.

    19. Meng, Q., Mehregany, M., and Mullen, R., Theoretical modeling of microfabricated beams with elastically

    restrained supports, Journal of Microelectromechanical Systems 2, 1993, 128137.

    20. Nayfeh, A. H., Introduction to Perturbation Techniques, Wiley, New York, 1981.

    21. Nayfeh, A. H. and Mook, D. T., Nonlinear Oscillations, Wiley, New York, 1979.

    22. Nayfeh, A. H. and Balachandran B., Applied Nonlinear Dynamics, Wiley, New York, 1995.