modeling, analysis, and experimental validation of a...

10
Proceedings of the ASME 2011 International Design Engineering Technical Conferences & Computers and Information in Engineering Conference IDETC/CIE 2011 August 28 - 31, 2011, Washington, DC, USA DETC2011-48199 MODELING, ANALYSIS, AND EXPERIMENTAL VALIDATION OF A BIFURCATION-BASED MICROSENSOR Vijay Kumar, J. William Boley, Yushi Yang, Hendrik Ekowaluyo, Jacob K. Miller, George T.-C. Chiu, Jeffrey F. Rhoads * School of Mechanical Engineering, Birck Nanotechnology Center, and Ray W. Herrick Laboratories Purdue University West Lafayette, Indiana 47907 Email: [email protected] ABSTRACT Microelectromechanical chemical and biological sensors have garnered significant interest over the past two decades due to their ability to selectively detect very small amounts of added mass. Today, most resonant mass sensors utilize chemomechanically-induced shifts in the linear natural fre- quency for detection. In this paper, an alternative, amplitude- based sensing approach, which exploits dynamic transitions across saddle-node bifurcations that exist in a microresonator’s nonlinear frequency response, is investigated. In comparison to their more traditional, linear counterparts, these bifurcation- based sensors have the ability to provide improved sensor met- rics, eliminate power-consuming hardware from final sensor im- plementations, and operate effectively at smaller (e.g. nano) scales. The present work details the ongoing development of a bifurcation-based mass sensor founded upon the near-resonant response of piezoelectrically-actuated microcantilevers. Specif- ically, the work details the modeling and analysis of these de- vices, their functionalization, and proof-of-concept mass sensing experiments which not only validate the proposed technique, but allow for the direct evaluation of pertinent sensor metrics. * Address all correspondence to this author. 1 INTRODUCTION Chemical and biological sensors founded upon microelec- tromechanical devices have garnered significant research inter- est over the past two decades due to their ability to selectively detect small amounts of added mass (see, for example, [1–3]). This capability, in addition to the advantages of comparatively small size, low power consumption metrics, and favorable cost, has led to the consideration of microscale sensors in applica- tions ranging from medical diagnostics and environmental mon- itoring to national security and public safety [4–7]. Conven- tional resonant mass sensors utilize chemomechanically-induced shifts in the linear natural frequency of the system, arising from the adsorption/absorption of an analyte onto/into a selectively- functionalized surface, for mass detection. These changes in fre- quency can be directly correlated to an analyte detection event or can be used to determine the relative concentration of an analyte in a given test environment [8]. While there is unquestionable utility associated with sens- ing approach described above, recent results indicate that im- proved sensor metrics may be achieved by exploiting a sensor’s near-resonant, nonlinear behavior. The benefits of tracking a sys- tem’s nonlinear resonant frequency for mass detection was high- lighted by Dai and collaborators [9]. Likewise, Younis, Zhang and their respective collaborators highlighted the potential ben- efits of mass sensors based on amplitude shifts resulting from 1 Copyright c 2011 by ASME Proceedings of the ASME 2011 International Design Engineering Technical Conferences & Computers and Information in Engineering Conference IDETC/CIE 2011 August 28-31, 2011, Washington, DC, USA DETC2011-4

Upload: others

Post on 08-Jul-2020

2 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Modeling, Analysis, and Experimental Validation of a ...jfrhoads/Publications/IDETC_2011_Kumar.pdfVijay Kumar, J. William Boley, Yushi Yang, Hendrik Ekowaluyo, Jacob K. Miller, George

Proceedings of the ASME 2011 International Design Engineering Technical Conferences &Computers and Information in Engineering Conference

IDETC/CIE 2011August 28 - 31, 2011, Washington, DC, USA

DETC2011-48199

MODELING, ANALYSIS, AND EXPERIMENTAL VALIDATION OF ABIFURCATION-BASED MICROSENSOR

Vijay Kumar, J. William Boley, Yushi Yang, Hendrik Ekowaluyo,Jacob K. Miller, George T.-C. Chiu, Jeffrey F. Rhoads∗

School of Mechanical Engineering,Birck Nanotechnology Center,

and Ray W. Herrick LaboratoriesPurdue University

West Lafayette, Indiana 47907Email: [email protected]

ABSTRACTMicroelectromechanical chemical and biological sensors

have garnered significant interest over the past two decadesdue to their ability to selectively detect very small amountsof added mass. Today, most resonant mass sensors utilizechemomechanically-induced shifts in the linear natural fre-quency for detection. In this paper, an alternative, amplitude-based sensing approach, which exploits dynamic transitionsacross saddle-node bifurcations that exist in a microresonator’snonlinear frequency response, is investigated. In comparisonto their more traditional, linear counterparts, these bifurcation-based sensors have the ability to provide improved sensor met-rics, eliminate power-consuming hardware from final sensor im-plementations, and operate effectively at smaller (e.g. nano)scales. The present work details the ongoing development of abifurcation-based mass sensor founded upon the near-resonantresponse of piezoelectrically-actuated microcantilevers. Specif-ically, the work details the modeling and analysis of these de-vices, their functionalization, and proof-of-concept mass sensingexperiments which not only validate the proposed technique, butallow for the direct evaluation of pertinent sensor metrics.

∗Address all correspondence to this author.

1 INTRODUCTIONChemical and biological sensors founded upon microelec-

tromechanical devices have garnered significant research inter-est over the past two decades due to their ability to selectivelydetect small amounts of added mass (see, for example, [1–3]).This capability, in addition to the advantages of comparativelysmall size, low power consumption metrics, and favorable cost,has led to the consideration of microscale sensors in applica-tions ranging from medical diagnostics and environmental mon-itoring to national security and public safety [4–7]. Conven-tional resonant mass sensors utilize chemomechanically-inducedshifts in the linear natural frequency of the system, arising fromthe adsorption/absorption of an analyte onto/into a selectively-functionalized surface, for mass detection. These changes in fre-quency can be directly correlated to an analyte detection event orcan be used to determine the relative concentration of an analytein a given test environment [8].

While there is unquestionable utility associated with sens-ing approach described above, recent results indicate that im-proved sensor metrics may be achieved by exploiting a sensor’snear-resonant, nonlinear behavior. The benefits of tracking a sys-tem’s nonlinear resonant frequency for mass detection was high-lighted by Dai and collaborators [9]. Likewise, Younis, Zhangand their respective collaborators highlighted the potential ben-efits of mass sensors based on amplitude shifts resulting from

1 Copyright c© 2011 by ASME

Proceedings of the ASME 2011 International Design Engineering Technical Conferences & Computers and Information in Engineering Conference

IDETC/CIE 2011 August 28-31, 2011, Washington, DC, USA

DETC2011-48199

Page 2: Modeling, Analysis, and Experimental Validation of a ...jfrhoads/Publications/IDETC_2011_Kumar.pdfVijay Kumar, J. William Boley, Yushi Yang, Hendrik Ekowaluyo, Jacob K. Miller, George

Normalized Frequency

Am

plitu

de Time

Am

plitu

deA

A’

B

FIGURE 1. FREQUENCY RESPONSE OF A DUFFING-LIKERESONATOR WITH SOFTENING NONLINEARITY. POINTS AAND B REPRESENT SADDLE-NODE BIFURCATIONS. NOTETHAT AS THE SYSTEM TRANSITIONS FROM POINT A’ TOPOINT A, THERE IS A SUDDEN JUMP IN RESPONSE AM-PLITUDE (THE INSET SHOWS THE TIME RESPONSE WHENTHE SYSTEM MOVES ACROSS THE BIFURCATION POINT -IGNORING TRANSIENT DYNAMICS). THIS TRANSITION CANBE INDUCED BY CHEMOMECHANICAL SHIFTS IN THE RES-ONATOR’S NATURAL FREQUENCY AND CAN BE DIRECTLYCORRELATED TO A MASS DETECTION EVENT.

operation near electrostatic pull-in and a parametric instabilityrespectively [10, 11]. Building upon these prior endeavors, thepresent work seeks to develop an amplitude-based mass sens-ing approach that utilizes dynamic transitions across sub-criticalsaddle-node bifurcations that exist in the nonlinear frequency re-sponse of a prototypical microscale device.

Microscale resonators operating in a nonlinear regime typ-ically exhibit a Duffing-like frequency response characteristicnear the first natural frequency. This frequency response ischaracterized by the presence of multiple steady-state solutions,hysteretic behavior and two saddle-node bifurcations. Figure 1shows the steady-state response of a representative resonator,plotted against the normalized excitation frequency (the exci-tation frequency normalized by the linear natural frequency ofthe resonator). A small change in the normalized excitation fre-quency of the resonator (caused by either sweeping the excitationfrequency or the adsorption of an analyte onto the surface), whenoperating near the lower saddle-node bifurcation point causes asudden change in the response amplitude (i.e. the system jumpsfrom the lower amplitude solution branch to the higher ampli-tude solution branch). This abrupt change in the near-resonantamplitude of the resonator can be exploited for mass sensing pur-poses, much like a shift in the resonant frequency of the system.

Note that an amplitude-based sensor offers the potential to elim-inate frequency tracking hardware components, which are typi-cally required in a linear resonant mass sensor, from final deviceimplementations.

Piezoelectrically-actuated microcantilevers were selected asa test bed to investigate the proposed bifurcation-based sensingapproach. These devices are well suited for mass sensing dueto their self-sensing capabilities [12] and maturity in force sens-ing and in scanning probe microscopy contexts [13]. In orderto fully describe the dynamics of these probes, a comprehensivenonlinear equation of motion has been developed accounting forthe geometric, inertial and the inherent material nonlinearities ofthe composite system, as described in Section 2. Section 3 de-scribes the analysis of the aforementioned model and includes anevaluation of the sensor’s performance metrics. Section 4 detailsthe experimental characterization of the proposed bifurcation-based sensor, including device functionalization, frequency re-sponse characterization, and proof-of-concept mass sensing ex-periments. The paper then concludes with a brief summary ofresults and an overview of ongoing research.

2 MODELINGFigure 2 highlights, both schematically and pictorially,

the piezoelectrically-actuated microbeam of interest. The mi-crobeam consists of a silicon cantilever and an integrated ZnOlayer, which is sandwiched between two Au/Ti electrodes. Fig-ure 3 shows the beam element, with associated variables, usedfor modeling.

Employing classical energy techniques, the equation of mo-tion governing the transverse vibrations of the system is derived.Defining the axial and transverse deformations of the beam asu(s, t) and v(s, t) respectively, and ˙(•) and (•)′ as the derivativeoperators associated with the time and the arc length variables, the kinematic constraint relating the transverse and longitu-dinal deformations to the beam’s angular deflection ψ can beexpressed as

tanψ =v′

1+u′. (1)

The axial strain associated with the microbeam ε11 is then givenby

ε11 =−(y− yn)ψ ′ for s < l1,

=−yψ′ for l1 < s < l,

(2)

where yn is the location of the neutral axis from the midplane of

2 Copyright c© 2011 by ASME

Page 3: Modeling, Analysis, and Experimental Validation of a ...jfrhoads/Publications/IDETC_2011_Kumar.pdfVijay Kumar, J. William Boley, Yushi Yang, Hendrik Ekowaluyo, Jacob K. Miller, George

wt

wb

wp

Piezoelectric Layer

Silicon Cantilever

l

l1l2

tp

tb

FIGURE 2. SCHEMATIC AND PICTORIAL REPRESENTA-TIONS OF THE PIEZOELECTRICALLY-ACTUATED MICROCAN-TILEVER OF INTEREST.

FIGURE 3. SCHEMATIC DIAGRAM OF A BEAM ELEMENT.

the silicon cantilever for the region s < l1 and is given by

yn =Epwptp(tp + tb)

2(Eptpwp +Ebtbwb). (3)

Note that Eb and Ep are the elastic modulus of the beam andthe elastic modulus of the laminate stack consisting of the Au/Tielectrodes and the piezoelectric material respectively [14], w andt represent the width and the thickness of different sections of thedevice, with the subscripts p and b representing the piezoelectricstack and the silicon cantilever respectively, and l1, l2 and l aredefined as in Fig. 2.

Using classical constitutive equations, the relationships be-tween the induced stress and the strain in the silicon cantileverand the piezoelectric patch can be defined. For the silicon can-tilever, the constitutive equation is given by σb

11 = Ebεb11, while,

for the piezoelectric patch, the constitutive relations are given

by [15, 16]

σp11 = Epε

p11−h31Q3 +

α1

2(ε p

11)2 +

α2

2(Q3)2−α3ε

p11Q3

D3 = h31εp11 +β31Q3 +

α3

2(ε p

11)2 +

α4

2(Q3)2−α2ε

p11Q3,

(4)

where the superscripts b and p denote the beam and the piezo-electric layer respectively, and σ11, Q3, D3, h31 and β31 representthe stress, applied electric field, electrical displacement, effec-tive coupling coefficient that relates the electrical and mechan-ical displacements and the permittivity coefficient respectively.αi (i = 1,2,3,4) are nonlinear material coefficients. Note thatan assumption that the electrical displacement is linear is madehere. Also note that the electric field is the gradient of the appliedpotential (φ ) with respect to the thickness variable (y):

Q3 =−∂φ

∂y. (5)

The absence of free charges inside the piezoelectric layer im-plies, D3,3 = 0 [17]. This, along with Eqns. (2) and (4) leads tothe conclusion that the potential must be a quadratic function iny. Using the aforementioned relationships along with the theoryof classical elasticity, the kinetic energy T of the resonator canbe expressed as

T =12

∫ l

0m(s)

[u2 + v2]ds, (6)

where m(s) represents the mass density of the composite sys-tem. The potential energy U associated with the system is thenexpressed as

U =12

∫ l1

0

∫∫A

σp11ε

p11dAds+

12

∫ l1

0

∫∫A

σb11ε

b11dAds

− 12

∫ l1

0

∫∫A

Q3D3dAds+12

∫ l2

l1

∫∫A

σb11ε

b11dAds

+12

∫ l

l2

∫∫A

σb11ε

b11dAds,

(7)

where A represents the cross-sectional area of the composite sys-tem.

Using these as a basis for study and assuming that the beamhas negligible rotational inertia, the specific Lagrangian of thegiven system can be defined as L = T −U . Application of ex-tended Hamilton’s principle results in the following variational

3 Copyright c© 2011 by ASME

Page 4: Modeling, Analysis, and Experimental Validation of a ...jfrhoads/Publications/IDETC_2011_Kumar.pdfVijay Kumar, J. William Boley, Yushi Yang, Hendrik Ekowaluyo, Jacob K. Miller, George

equation for the system:

δH = 0 = δ

∫ t2

t1

∫ l

0

L+

12

λ[1− (1+u′)2− (v′)2]dsdt

+∫ t2

t1

∫ l

0(Quδu+Qvδv)dsdt,

(8)

where λ is a Lagrange multiplier introduced to maintain theinextensibility constraint and Qu and Qv are the external non-conservative forces in the longitudinal and transverse directionsrespectively. Note that the only external non-conservative forceacting on the system is the force resulting from damping andthus, Qu = 0 and Qv =−cv, where c is the specific viscous damp-ing coefficient. Using third-order Taylor series approximationsfor ψ and integrating Eqn. (8) successively by parts yields theequations governing the longitudinal and transverse vibrationsof the system. Solving the equation governing the longitudinalvibration for the Lagrange multiplier and substituting the resultin the equation for the transverse vibration, along with the ex-pression for u resulting from the inextensibility constraint, yieldsa single equation governing the transverse vibrations of the sys-tem. Nondimensionalizing this result by introducing a scaling ofthe displacement, arc length, and time variables yields a scaled,distributed parameter model for the system, which is not givenexplicitly here for the sake of brevity.

The complex distributed parameter model noted above canbe reduced to a system of ordinary differential equations usingstandard modal projection techniques. Thus, an expansion forthe transverse displacement of the system v(s, t) of the form

v(s, t) =∞

∑i=0

wi(t)Φi(s), (9)

where Φi(s) is the modeshape of the individual modes and wi(t)is the displacement in the respective modes, is introduced. Sincethe response of the system is dominated by the first mode, thesystem is projected onto the first bending mode. Note thatthe beam geometry is discontinuous and hence the appropriatemodeshape has to be formulated. To derive the modeshape ofthe discontinuous beam, the beam is split into three regions, withregion 1 spanning from 0 to l1, region 2 spanning from l1 to l2and region 3 spanning from l2 to l.

Solving the linear, unforced model for each region, themodeshape for each region can be obtained. Assembling themodeshape solutions together with appropriate boundary andcontinuity conditions renders the modeshape of the entire beam.Note that the modeshape is normalized here such that the ampli-tude at the tip of the microcantilever is equal to 1. The first andthe second bending modes of a Veeco DMASP probe are shownFig. 4.

0.2 0.4 0.6 0.8 1.0Position

0.2

0.4

0.6

0.8

1.0

Amplitude

(a)

0.2 0.4 0.6 0.8 1.0Position

0.2

0.4

0.6

0.8

1.0

Amplitude

(b)

FIGURE 4. (a) FIRST AND (b) SECOND MODESHAPES OF AVEECO DMASP PROBE, DERIVED USING A DISCONTINUOUSBEAM MODEL.

Using the above formulation for the modeshape, projectiononto the first mode of the beam yields the final, lumped-massequation of motion for the system:

w+ ε2cw+(ω2

n + ε2∆φλ1 + ε

2∆φ

2γ1)w

+ εk2w2 + ε2∆φλ2w2 + ε

2(k3 +∆φλ3 +∆φ2γ3)w3

+ ε2χ(ww2 +w2w) = ε

2η1∆φ .

(10)

Note that ∆φ = f cos(Ωt), where f is the amplitude of voltageexcitation and Ω is the normalized excitation frequency. Thenondimensional coefficients in Eqn. (10) are strong functionsof the system’s geometry, resonant modeshapes and constitutivematerial properties. The effective linear and the cubic stiffnesses(ω2

n , k3), and the nonlinear term (χ) vary strongly with system’sgeometry and elastic material properties and exhibit a weak de-pendence on material nonlinearities, while the other terms inthe expression are strong functions of the linear and nonlinearpiezoelectric material coefficients. Detailed expressions for thenondimensional coefficients in Eqn. (10) are omitted here for thesake of brevity. Note that the coefficient scaling employed herewas developed such that physically-relevant terms remain in asecond-order perturbation analysis of the system. Also note thatif linear constitutive relationships are used in place of their non-linear counterparts, the resulting system model resembles a clas-sical Duffing resonator.

4 Copyright c© 2011 by ASME

Page 5: Modeling, Analysis, and Experimental Validation of a ...jfrhoads/Publications/IDETC_2011_Kumar.pdfVijay Kumar, J. William Boley, Yushi Yang, Hendrik Ekowaluyo, Jacob K. Miller, George

3 ANALYSISTo study the behavior of the system near its primary reso-

nance, a second-order multiple scales analysis is performed. Tosimplify the analysis of the equation of motion, all material non-linearities except α1 [in Eqn. (4)] are assumed to be zero (Notethat ongoing experimental efforts are aimed at validating this as-sumption) [16]. Thus, Eqn. (10) is transformed to:

w+ω2n w+ ε

2cw+ εk2w2

+ ε2 [k3w3 +λ2∆φw2 + χ(ww2 +ww2)

]= ε

2η1∆φ .

(11)

To facilitate the analysis, multiple time scales are introduced asfollows

t = T0 + εT1 + ε2T2 + . . .

ddt

=d

dT0+ ε

ddT1

+ ε2 d

dT2+ . . .

= D0 + εD1 + ε2D2 + . . .

(12)

Likewise, w(t) is expanded according to

w(t;ε) = w0(T0,T1,T2)+ εw1(T0,T1,T2)+ ε2w2(T0,T1,T2)+ . . .

(13)Substituting Eqns. (12) and (13) into Eqn. (11) and equating

coefficients of like powers of ε yields

(ε0) :D20w0 +ω

2n w0 = 0,

(ε1) :D20w1 +ω

2n w1 =−2D0D1w0− k2w2

0,

(ε2) :D20w2 +ω

2n w2 =−cD0w0−2D0D2w0

−2D0D1w1−D21w0−2k2w0w1− k3w3

0

−χ[w0(D0w0)2 +w2

0(D20w0)

]−λ2w2

0∆φ +η1∆φ .

(14)

Solving this set of equations using higher-order multiple scalemethods, and introducing the following quantities

σ =Ω−ωn

ε2ωn,

Ne f f =9k3ω2

n −10k22−6χω4

n

24,

(15)

where σ is a detuning parameter used to describe the nearness ofthe excitation frequency to the natural frequency and Ne f f is aneffective nonlinear coefficient, results in a relationship governingthe steady-state frequency response of the system:

16a20c2ω2

n

f 2(−4η1 +a2

0λ2)2 +

64(a3

0Ne f f −a0σω4n)

2

f 2(4η1−3a2

0λ2)2

ω4n

= 1. (16)

-5 0 50

10

20

30

40

50

60

70

Detuning parameter, σ

Out

put A

mpl

itude

Neff = -0.01Neff = -0.005Neff = 0.005

FIGURE 5. FREQUENCY RESPONSE OF REPRESENTATIVEMICRORESONATOR WITH VARIOUS VALUES OF THE EFFEC-TIVE NONLINEAR COEFFICIENT (Ne f f ) EXCITED BY AN AP-PLIED EXCITATION VOLTAGE OF f = 10 V. STABLE SOLU-TIONS ARE REPRESENTED USING SOLID LINES, WHILE THEUNSTABLE SOLUTIONS ARE REPRESENTED WITH DASHEDLINES. NOTE THAT AS THE VALUE OF THE EFFECTIVENONLINEAR COEFFICIENT INCREASES, THE SYSTEM SHIFTSFROM SOFTENING TO HARDENING BEHAVIOR.

Here, a0 is the steady-state near-resonant amplitude of the sys-tem. Solving Eqn. (16) for a0 and plotting the solutions as afunction of the detuning parameter σ reveals the steady-statefrequency response structure of the system. Note that the pa-rameters c, ωn, η1 and λ2 are known functions of the beam’s ge-ometry and material properties, while the effective nonlinear co-efficient Ne f f is a function of the material nonlinearities of thepiezoelectric material and hence, generally unknown.

To simulate the behavior of the system, the frequency re-sponse of the system is plotted for an excitation amplitude of10 V for different values of the effective nonlinear coefficientNe f f . The output amplitude is plotted against the detuning pa-rameter σ , which can be directly correlated to the excitation fre-quency (Fig. 5). The plot shows the presence of multiple so-lutions, with the stable solutions being represented by the solidlines and the unstable solution being represented by the dashedlines. The values of the detuning parameter (and hence, thefrequency) at which the stable and unstable solutions meet arethe saddle-node bifurcation frequencies. The frequency responseshows the presence of two saddle-node bifurcation frequencies,one where the lower amplitude stable solution and the unstablesolution meet, and another where the higher amplitude stable so-lution and the unstable solution meet (not shown in figure). Fornegative values of the effective nonlinear coefficient, the systemshows a softening behavior while for positive values, the systemshows hardening frequency response characteristics.

Motivated by the above results, the excitation voltage is plot-

5 Copyright c© 2011 by ASME

Page 6: Modeling, Analysis, and Experimental Validation of a ...jfrhoads/Publications/IDETC_2011_Kumar.pdfVijay Kumar, J. William Boley, Yushi Yang, Hendrik Ekowaluyo, Jacob K. Miller, George

-1.3 -1.2 -1.1 -1 -0.9 -0.8 -0.7 -0.610

12

14

16

18

20

Value of detuning parameter at bifurcation, σcr

Exci

tatio

n V

olta

ge, V

pp

FIGURE 6. THE BIFURCATION FREQUENCY (CRITICALVALUE OF THE DETUNING PARAMETER σcr) PLOTTED AS AFUNCTION OF EXCITATION VOLTAGE FOR A REPRESENTA-TIVE VALUE OF THE EFFECTIVE NONLINEAR COEFFICIENT(Ne f f = -0.01). THE PLOT SHOWS THAT THE BIFURCATIONFREQUENCY HAS A NEAR-LINEAR RELATIONSHIP WITH THEEXCITATION VOLTAGE OVER THE EXCITATION RANGE OF IN-TEREST.

ted against the lower saddle-node bifurcation frequency for agiven value of the effective nonlinear coefficient (Fig. 6). Theplot shows that the bifurcation frequency has a near-linear rela-tionship with the excitation voltage over the excitation range ofinterest. To accurately estimate the bifurcation frequency, an es-timate of the effective nonlinear coefficient Ne f f is needed. Ne f fis a function of the material nonlinear coefficient α1. An estimateof the nonlinear material coefficient can be achieved by design-ing a simple parametric identification routine and using it in con-junction with experimental results. Illustrations of typical para-metric identification routines in piezoelectrically-actuated mi-croscale resonators can be found in [16] and [18]. Employinga similar identification scheme will provide a better estimate forthe bifurcation frequency than that detailed utilized here, and willalso allow for improved estimates of the sensor’s metrics.

3.1 SENSOR METRICSTo compare the effectiveness of a given mass sensor against

other comparable platforms, it is important to define the metricsof the device both analytically and experimentally. One metric ofparticular interest is the minimum detectable mass or mass sen-sitivity of the device. For linear resonance based mass sensors,mass sensitivity is a function of the mass responsivity of the sen-sor and frequency resolution, and is given by [19]

δm≈ R−1δω, (17)

where δm is the mass sensitivity of the sensor, R is the massresponsivity of the sensor and δω is the frequency resolution ofthe system. The mass responsivity R is given by

R =∂ωn

∂me f f, (18)

where ωn is the natural frequency of the system and me f f is theeffective mass of the system, accounting for the added mass dueto the adsorption of analyte molecules onto the functionalizedsurface. Since R is a deterministic quantity fixed by sensor selec-tion, the mass sensitivity of a given device is commonly a strongfunction of the frequency resolution. In the presence of noise anduncertainty, the frequency resolution (and hence, the minimumdetectable mass) is strongly governed by the minimum resolu-tion that could be achieved from the experimental setup.

For an amplitude-shift based mass sensor, the chain rule ofdifferentiation can be employed to derive an expression for masssensitivity from Eqn. (17)

δm∗ = R−1S−1δa, (19)

where δm∗ is the mass sensitivity of an amplitude-based sensor,R is the mass responsivity as defined in Eqn. (18), S is the changein output amplitude as a function of the excitation frequency andδa is the amplitude resolution, that is, the minimal detectablechange in amplitude as limited by the presence of noise and un-certainty:

S =∂a∂ω

. (20)

Note that the sensor operates in a frequency range where mul-tiple stable steady-state solutions are present. In the absence ofnoise and external perturbations, the system remains on the solu-tion branch dictated by initial conditions until a bifurcation pointis crossed, whereat the rate of change of amplitude as a func-tion of the frequency tends to infinity. The minimal detectabledifference in amplitude between the two stable solutions at thesaddle-node bifurcation dictates the amplitude resolution of thesensor. For the given system, this quantity is quite high (as shownin Fig. 5 and later verified experimentally) and therefore, the sys-tem is not practically limited by the amplitude resolution. Thus,a bifurcation-based sensor, operating at a saddle-node bifurca-tion frequency in the absence of noise has a minimum detectablemass tending to zero.

The effects of noise on the bifurcation frequencies of a non-linear system have been investigated by a number of researchers,with notable efforts emphasizing saddle-node bifurcations being

6 Copyright c© 2011 by ASME

Page 7: Modeling, Analysis, and Experimental Validation of a ...jfrhoads/Publications/IDETC_2011_Kumar.pdfVijay Kumar, J. William Boley, Yushi Yang, Hendrik Ekowaluyo, Jacob K. Miller, George

reported in [20, 21]. As shown in [20], for a typical Duffing-likeresonator, noise blurs the separatrix between the basins of attrac-tion of the two different stable solutions. Thus, when choosingthe operating point for the nonlinear sensor, the excitation fre-quency should be selected such that noise does not cause the sys-tem to crossover into the other basin of attraction. This placesconstraints on the minimum detectable mass of the sensor (or,the exposure time needed for the sensor to detect the presenceof analyte in the case of a threshold sensor). While a number ofprevious efforts have addressed the control and stabilization ofthe dynamic response of microresonators near bifurcations (see,for example [22]), the study of such control techniques, as wellas stochastic nature of the bifurcation phenomenon, is largely be-yond the scope of the present work and hence, is not addressedin detail here.

4 EXPERIMENTAL CHARACTERIZATIONTo validate the operating principles of the proposed

bifurcation-based sensors, a series of experimental investigationswere initiated. As a first step, Veeco DMASP probes werefunctionalized with a polymer capable of absorbing/adsorbingthe target molecules of interest. After establishing the nonlin-ear frequency response characteristics of a representative micro-cantilever, the sensor was subjected to mass sensing trials in acarefully-controlled chemical environment.

Although there exists a wide variety of methods for func-tionalizing microcantilevers, inkjet printing technology hasgained significant research attention recently due to its abilityto precisely deposit small volumes of liquid onto any planar ornon-planar surface [23]. Compared to conventional fabricationtechniques, inkjet technology provides high throughput, has lowproduction cost and is relatively clean [24]. In this work, thebeams were functionalized with Poly 4-Vinyl Pyridine to facili-tate methanol sensing [25].

A micro-inkjet printing process with thermal actuation en-abled drop formation was used to functionalize the sensors [26].In this process, electrical current is passed through a resistor thatis placed within a reservoir containing the polymer. Resistiveheating causes the temperature to rise until a vapor bubble formsand subsequently collapses. This induces a wave in the fluid,which, in turn, results in the formation of a drop [27]. Each ofthese cantilevers were functionalized with 110 pL drops in lessthan 1 second. To preserve the integrity of the sensor’s surfacechemistry prior to testing, the beams were stored in a nitrogen en-vironment. A scanning electron micrograph of a representativefunctionalized beam is shown in Fig. 7 [28].

The experimental setup used in this work consisted of acustom-designed chemical test setup, a scanning Laser DopplerVibrometer and associated hardware (Fig. 8). The chemical testsetup was used to precisely control the chemical makeup of theenvironment in which the sensors were tested. The chemical test

FIGURE 7. A SCANNING ELECTRON MICROGRAPH OF AFUNCTIONALIZED MICROCANTILEVER [28].

N2

LabVIEW Interface

MFC 2

MFC 1 Test Chamber

Methanol Bubbler

SLDV

FIGURE 8. SCHEMATIC REPRESENTATION OF THECOMPUTER-CONTROLLED EXPERIMENTAL SETUP.

setup consisted of multiple analyte bubblers, which were con-nected to precision mass flow controllers and a carrier gas sup-ply. By controlling the pressure of the carrier gas and tempera-ture of the analyte in the bubbler and the mass flow rates in themass flow controllers, a wide range of concentration levels wereproduced. The temperature of the analyte bubblers were moni-tored and controlled through Resistance Temperature Detectors(RTDs) and heat pads. The mixture was then diverted to an iso-lated test chamber which housed the sensors. The test chamber isoptically-accessible from the top and has a complete temperaturecontrol similar to that of the analyte bubblers. Sealed electricaloutlets are available at the base of the chamber to provide inputto the sensor and to provide electrical access to the heat pads andRTDs. The exhaust from the chamber was routed to a condensingsetup where the target analyte was recovered. A LabVIEW inter-face controlled the entire setup, providing an input to the sensor,recovering the output from the vibrometer and controlling thechemical concentration of the test chamber (which includes thecontrol of heat pads, RTDs and the mass flow controllers).

The nonlinear frequency responses of various Veeco

7 Copyright c© 2011 by ASME

Page 8: Modeling, Analysis, and Experimental Validation of a ...jfrhoads/Publications/IDETC_2011_Kumar.pdfVijay Kumar, J. William Boley, Yushi Yang, Hendrik Ekowaluyo, Jacob K. Miller, George

5.95 6 6.05x 104

0.2

0.4

0.6

0.8

1

Frequency, Hz

Nor

mal

ized

Am

plitu

de

Forward SweepReverse Sweep

FIGURE 9. FREQUENCY RESPONSE OF THE SYSTEM, ASRECOVERED VIA FORWARD AND REVERSE FREQUENCYSWEEPS, FOR AN EXCITATION VOLTAGE OF 18 Vpp AND A 160sSWEEP PERIOD.

DMASP probes were recovered using the Polytec MSA-400Laser Doppler Vibrometer and the LabVIEW interface. Afterestablishing the natural frequencies of a given device, frequencysweeps were performed near the first natural frequency to char-acterize the nonlinear frequency response behavior. Frequencysweeps were conducted at a variety of actuation voltage levelsand a variety of sweep rates, ranging from quasi-static to highlynon-stationary, to characterize these devices. Figure 9 showsthe amplitude response of a typical Veeco DMASP probe fora sweep period of 160 seconds and an excitation voltage of 18Vpp, a representative set of parameters which yield repeatableand predictable bifurcation behavior. The response shows thepresence of saddle-node bifurcations in both the forward and re-verse sweeps.

Although the analysis above clearly shows the presence ofa saddle-node bifurcation frequency during a frequency sweep, aprecise estimation of the jump frequency requires careful consid-eration of a number of factors, including (i) Noise in the response– Most microsystems are inherently noisy and, as shown by [29]for example, exhibit observable frequency response characteris-tics which have a strong dependence on sweep rate; and (ii) Re-sistive heating – The piezoelectrically-actuated sensor is proneto resistive heating which causes the material to soften, and thebifurcation frequency to decrease. Therefore, it is necessary tolet the system achieve (near) thermal equilibrium before makinga measurement.

4.1 MASS SENSING EXPERIMENTSIn an effort to validate the performance of the functionalized

microcantilevers and the test chamber, a series of conventional,linear mass sensing trials were performed. To this end, the beam

0 5000 10000 150006.15

6.16

6.17

6.18

6.19x 104

Time, sec

Bifu

rcat

ion

freq

uenc

y, H

z

0 5000 10000 15000-5

0

5

10

15

Flow

rate

of M

etha

nol,

sccm

FIGURE 10. THE SENSOR’S BIFURCATION FREQUENCY,AND THE FLOW RATE OF METHANOL, PLOTTED AS A FUNC-TION OF TIME. NOTE THAT AS THE MIXTURE IS SUPPLIED TOTHE CHAMBER THE BIFURCATION FREQUENCY DECREASES.AS DESORPTION OCCURS, THE BIFURCATION FREQUENCYINCREASES.

was excited with a pseudorandom signal of 2 V amplitude, a fastfourier transform of the beam’s response was recovered, and thefirst natural frequency was tracked as a function of time. An an-alyte/carrier gas mixture (Methanol/Nitrogen) was supplied forvarious periods of time and the system was purged periodicallywith nitrogen to ensure the successive adsorption and desorptionof methanol molecules onto the sensor’s functionalized surface.As the methanol molecules adsorbed on the surface, the naturalfrequency decreased and as the molecules desorbed, the naturalfrequency increased, thus indicating a mass-dominated chemo-mechanical process [28].

To realize an amplitude-based mass sensor, an estimate ofthe change in bifurcation frequency as a function of the addedanalyte mass is required. For this trial, the beam was subjected toa series of adsorption/desorption cycles and the saddle-node bi-furcation frequency was identified at each step. Figure 10 showsa representative plot of the saddle-node bifurcation frequency asa function of time, plotted in conjunction with the flow rate of thetarget analyte. Here, a concentration setting of 1.18% by mass ofanalyte in the carrier gas mixture, as specified at the mass flowcontrollers, was used. As evident from the figure, the bifurcationfrequency decreased as the analyte adsorbed onto the surface, ata rate of 0.125 Hz/s and increased as the analyte desorbed fromthe surface at a rate of 0.076 Hz/s. This trend is similar to thatseen with the linear natural frequency.

Motivated by the above results, an operating point was se-lected for each device based on that particular device’s bifurca-tion frequency. The operating point was typically selected nearthe lower saddle-node bifurcation frequency, but sufficiently faraway that stochastic switching was not observed. The sensor was

8 Copyright c© 2011 by ASME

Page 9: Modeling, Analysis, and Experimental Validation of a ...jfrhoads/Publications/IDETC_2011_Kumar.pdfVijay Kumar, J. William Boley, Yushi Yang, Hendrik Ekowaluyo, Jacob K. Miller, George

0 500 1000 15001

2

3

4

Time, sec

Out

put a

mpl

itude

, V

0 500 1000 1500-2

0

2

4

Flow

rate

of M

etha

nol,

sccm

FIGURE 11. THE SENSOR’S RESPONSE AMPLITUDE, ANDTHE FLOW RATE OF THE TARGET ANALYTE, PLOTTED AS AFUNCTION OF TIME. AS MASS ADSORBS ONTO THE SENSOR,THE BIFURCATION FREQUENCY DECREASES PROVIDING ASUDDEN JUMP IN RESPONSE AMPLITUDE, WHICH SIGNALSA DETECTION EVENT.

then excited at constant frequency and the dynamic response wasrecorded. Figure 11 shows the results of a representative masssensing trial carried out with a concentration of 0.67% by massof the analyte in carrier gas (as specified at the mass flow con-trollers), with the operating frequency set at 30 Hz below the bi-furcation frequency. As the analyte/carrier gas mixture was sup-plied to the chamber, the bifurcation frequency decreased belowthe operating frequency of the sensor, and the response amplitudeshowed a sudden jump to the higher amplitude solution branch,signaling a positive detection event. The results presented hereinhave been repeated with multiple devices and at various concen-tration levels, with very few, if any, false positives and negativesbeing observed to date.

5 CONCLUSIONS AND FUTURE DIRECTIONSThe results presented herein detail the modeling, analysis,

experimental characterization, and ultimately demonstrate the vi-ability, of bifurcation-based resonant mass sensors based upondynamic transitions across saddle-node bifurcations. Currentefforts are focused on further characterizing the sensitivity ofthese devices, optimizing sensor metrics and exploiting their self-sensing nature for implementation in an integrated architecture.

AcknowledgementsThe work was generously supported by the School of Me-

chanical Engineering, Purdue University, and the Purdue Re-search Foundation.

REFERENCES[1] Thundat, T., Oden, P. I., and Warmack, R. J., 1997. “Mi-

crocantilever sensors”. Microscale Thermophysical Engi-neering, 1(3), pp. 185–199.

[2] Ekinci, K. L., Huang, X. M. H., and Roukes, M. L.,2004. “Ultrasensitive nanoelectromechanical mass detec-tion”. Applied Physics Letters, 84(22), pp. 4469–4471.

[3] Naik, A. K., Hanay, M. S., Hiebert, W. K., Feng, X. L., andRoukes, M. L., 2009. “Towards single-molecule nanome-chanical mass spectrometry”. Nature Nanotechnology, 4,pp. 445–450.

[4] Lang, H. P., Berger, R., Battiston, F., Ramseyer, J.-P.,Meyer, E., Andreoli, C., Brugger, J., Vettiger, P., Despont,M., Mezzacasa, T., Scandella, L., Guntherodt, H.-J., Ger-ber, C., and Gimzewski, J. K., 1998. “A chemical sensorbased on a micromechanical cantilever array for the identi-fication of gases and vapors”. Applied Physics A: MaterialScience & Processing, 66(Supplement 1), pp. S61–S64.

[5] Su, M., Li, S., and Dravid, V. P., 2003. “Microcantileverresonance-based DNA detection with nanoparticle probes”.Applied Physics Letters, 82(20), pp. 3562–3564.

[6] Boisen, A., Thaysen, J., Jensenius, H., and Hansen, O.,2000. “Environmental sensors based on micromachinedcantilevers with integrated read-out”. Ultramicroscopy,82(1-4), pp. 11–16.

[7] Senesac, L., and Thundat, T. G., 2008. “Nanosensors fortrace explosive detection”. Materials Today, 11(3), pp. 28–36.

[8] Thundat, T., Wachter, E. A., Sharp, S. L., and Warmack,R. J., 1995. “Detection of mercury vapor using resonat-ing microcantilevers”. Applied Physics Letters, 66(13),pp. 1695–1697.

[9] Dai, M. D., Eom, K., and Kim, C.-W., 2009. “Nanome-chanical mass detection using nonlinear oscillations”. Ap-plied Physics Letters, 95, 203104.

[10] Younis, M. I., and Alsaleem, F., 2009. “Exploration of newconcepts for mass detection in electrostatically-actuatedstructures based on nonlinear phenomena”. Journal ofComputational and Nonlinear Dynamics, 4, 021010.

[11] Zhang, W., and Turner, K. L., 2005. “Application of para-metric resonance amplification in a single-crystal siliconmicro-oscillator based mass sensor”. Sensors and Actua-tors A: Physical, 122(1), pp. 23–30.

[12] Gurjar, M., and Jalili, N., 2007. “Toward ultrasmallmass detection using adaptive self-sensing piezoelectricallydriven microcantilevers”. IEEE/ASME Transactions onMechatronics, 12(6), pp. 680–688.

[13] Wolf, K., and Gottlieb, O., 2002. “Nonlinear dynamics ofa noncontacting atomic force microscope cantilever actu-ated by a piezoelectric layer”. Journal of Applied Physics,91(7), pp. 4701–4709.

[14] Li, H., 2006. “Oscillations of microscale composite struc-

9 Copyright c© 2011 by ASME

Page 10: Modeling, Analysis, and Experimental Validation of a ...jfrhoads/Publications/IDETC_2011_Kumar.pdfVijay Kumar, J. William Boley, Yushi Yang, Hendrik Ekowaluyo, Jacob K. Miller, George

tures with applications to microresonators”. PhD thesis,University of Maryland.

[15] Joshi, S. P., 1992. “Non-linear constitutive relations forpiezoceramic materials”. Smart Materials and Structures,1(1), pp. 80–83.

[16] Mahmoodi, S. N., Jalili, N., and Daqaq, M. F., 2008. “Mod-eling, nonlinear dynamics, and identification of a piezo-electrically actuated microcantilever sensor”. IEEE/ASMETransactions on Mechantronics, 13(1), pp. 58–65.

[17] Basak, S., Raman, A., and Garimella, S. V., 2005. “Dy-namic response optimization of piezoelectrically excitedthin resonant beams”. Journal of Vibration and Acoustics,127(1), pp. 18–27.

[18] Dick, A. J., Balachandran, B., DeVoe, D. L., and Mote Jr.,C. D., 2006. “Parametric identification of piezoelectric mi-croscale resonators”. Journal of Micromechanics and Mi-croengineering, 16(8), pp. 1593–1601.

[19] Ekinci, K. L., Yang, Y. T., and Roukes, M. L., 2004. “Ul-timate limits to inertial mass sensing based upon nano-electromechanical systems”. Journal of Applied Physics,95(5), pp. 2682–2689.

[20] Kozinsky, I., Postma, H. W. C., Kogan, O., Husain, A., andRoukes, M. L., 2007. “Basins of attraction of a nonlinearnanomechanical resonator”. Physical Review Letters, 99,207201.

[21] Stambaugh, C., and Chan, H. B., 2006. “Noise-activatedswitching in a driven nonlinear micromechanical oscilla-tor”. Physical Review E, 73, 172302.

[22] Alsaleem, F. M. and Younis, M. I., 2010. “Stabilization ofelectrostatic MEMS resonators using a delayed feedbackcontroller”. Smart Materials and Structures, 19, 035016.

[23] Calvert, P., 2001. “Inkjet printing for materials and de-vices”. Chemistry of Materials, 13(10), pp. 3299–3305.

[24] Boley, W., Bhuvana, T., Hines, B., Sayer, R., Chiu, G.-C.,Fisher, T., Bergstrom, D., Reifenberger, R., and Kulkarni,G., 2009. “Inkjet printing involving palladium alkanethi-olates and carbon nanotubes functionalized with single-strand DNA”. In Proceedings of DF2009: Digital Fabri-cation Processes Conference, pp. 824–827.

[25] DeMartini, B. E., Rhoads, J. F., Zielke, M. A., Owen, K. G.,Shaw, S. W., and Turner, K. L., 2008. “A single input-singleoutput coupled microresonator array for the detection andidentification of multiple analytes”. Applied Physics Let-ters, 93, 054102.

[26] Post, N. J., 2007. “Precision micro-deposition of functionallayers using inkjet drop-on-demand and applications to thefunctionalization of microcantilever sensors”. MS thesis,Purdue University.

[27] Lee, E., 2003. Microdrop Generation, 1st ed. CRC Press.[28] Kumar, V., Boley, J. W., Ekowaluyo, H., Miller, J. K.,

Marvin, G. C., Chiu, G. T.-C., and Rhoads, J. F., 2010.“Linear and nonlinear mass sensing using piezoelectrically-

actuated microcantilevers”. In Proceedings of The 2010SEM Annual Conference and Exposition on Experimentaland Applied Mechanics, 11th International Symposium onMEMS and Nanotechnology.

[29] Miller, N., Burgner, C., Dykman, M., Shaw, S., and Turner,K., 2010. “Fast estimation of bifurcation conditions usingnoisy response data”. In Proceedings of SPIE: Sensors andSmart Structures Technologies for Civil, Mechanical, andAerospace Systems, 7647, 76470O.

10 Copyright c© 2011 by ASME