vibrational control of mathieu's equation

6
Vibrational Control of Mathieu’s Equation I. P. M. Wickramasinghe, Student Member, IEEE, and Jordan M. Berg, Fellow, ASME, Senior Member, IEEE Abstract— The vertically driven inverted pendulum— sometimes called the “Kapitza pendulum”—is a well-known example of an unstable system that can be stabilized by oscillatory forcing. Averaging methods and asymptotic stability results can be applied to develop a general framework for designing suitable inputs. Linearizing the Kapitza pendulum yields Mathieu’s equation, which is also extensively studied for its stability characteristics.In this paper, results from these two bodies of work are compared, from the point of view of open-loop control of mechanical systems. The averaging approaches applied to Mathieu’s equation are seen to access only a very limited portion of the stability map. Furthermore, stabilizing input signals exist that may be found from the linear stability map, but are not found by averaging methods. The results suggest that the linear stability map is a powerful and underutilized tool for design and analysis of open-loop oscillatory control. I. I NTRODUCTION AND MOTIVATION The results of this paper are motivated by studies of side pull-in of MEMS comb drive actuators [22], [23], [24]. Side pull-in is an instability that occurs in constant-gap elec- trostatic devices, when the drive voltage exceeds a critical value. Side pull-in occurs in the direction orthogonal to the desired actuator travel, as opposed to standard pull-in, which occurs along the direction of desired actuator travel. Those studies were attempting to prevent or delay side pull-in, and thus allow operation of the device at higher drive voltages, corresponding to increased stroke or force. As in many MEMS control problems, the need for sensing, actuation, and/or computation must be balanced against complexity of design and fabrication. For the side pull-in problem, the nominal plant is uncontrollable and unstable beyond the pull-in point. Additional sensors and actuators are needed to stabilize those operating points using feedback control methods. The goal of those studies was to stabilize operation beyond the pull-in voltage without redesigning the device— that is, without the use of feedback. In 1951, the Russian physicist P. L. Kapitza published an analysis of the stability of the inverted pendulum with a verti- cally oscillating base [10]. At that time it was already known that the upright equilibrium point may be stabilized for forcing with sufficiently high frequency and sufficiently low amplitude [18]. Kapitza applied successive approximation to obtain an “effective moment” that captured the average effect of the base motion. Kapitza’s analysis is an iconic application of averaging theory to oscillatory stabilization; the verti- cally driven inverted pendulum is now sometimes called the I. P. M. Wickramasinghe and Jordan M. Berg are with the Depart- ment of Mechanical Engineering and the Nano Tech Center, Texas Tech University, Lubbock, TX, 79409, USA, e-mail: {m.wickramasinghe, jor- dan.berg.}@ttu.edu. Kapitza pendulum. Approaches to oscillatory stabilization based on averaging have received extensive attention across many disciplines, including control theory. As observed in an influential 1980 paper by S. M. Meerkov [13], this form of stabilization is appealing from a control perspective because it requires no measurement of the system states or outputs, nor does it require feedforward of commands or disturbances. Thus it would seem ideally suited to address side pull-in of comb drives. Meerkov named this strategy “vibrational control,” and classified it as a separate category of control, alongside feedback and feedforward methods. In the years since Meerkov’s seminal publication, control researchers have further developed open-loop oscillatory stability and control of linear and nonlinear systems using averaging, time-scale separation, and asymptotic stability theorems [2], [3], [4], [1], [21], [6]. It has previously been reported that bidirectional forcing is required for the successful use of averaging techniques in MEMS devices [15], [16], [17], [19]. These methods were successfully applied to stabilize a voltage-controlled gap- closing electrostatic actuator. A novel device design was required to overcome the unidirectional nature of capacitively coupled Coulomb forces [16], [19]. This result is a major advance over feedback control, despite the added complexity of the redesigned device, since it eliminates the need for sensing. However the goal of the constant-gap problem reported in [22], [23], [24] was feedback-free stabilization of the nominal device, with no special design features at all. Therefore averaging methods were adapted to the comb drive model. However, it was found that while stabilizing solutions could be found, they did not significantly delay the side pull-in bifurcation. However, it was also observed that linearization of the transverse comb drive dynamics yields Mathieu’s equation. Mathieu’s equation has been extensively analyzed using Floquet theory—which does not require av- eraging or time-scale separation—and is known to exhibit both parametric resonance and oscillatory stabilization [9], [5]. The linearized model can be stabilized well beyond the side pull-in point through consideration of the resulting stability map, and simulations show that the resulting input signal also stabilizes the nonlinear system. Furthermore, the result includes an explicit three-way trade-off between the magnitude of the input frequency, the amount by which the stable operating range is extended, and the precision with which the stabilizing forcing signal is applied. This is in contrast to the averaging methods, which guarantee stability only in the limit as the forcing frequency increases. This is an advantage of the Floquet approach that may have important practical consequences. We note that the linearization of the 2013 IEEE/ASME International Conference on Advanced Intelligent Mechatronics (AIM) Wollongong, Australia, July 9-12, 2013 978-1-4673-5320-5/13/$31.00 ©2013 IEEE 686

Upload: others

Post on 15-Jan-2022

2 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Vibrational Control of Mathieu's Equation

Vibrational Control of Mathieu’s Equation

I. P. M. Wickramasinghe, Student Member, IEEE, and Jordan M. Berg, Fellow, ASME, Senior Member, IEEE

Abstract— The vertically driven inverted pendulum—sometimes called the “Kapitza pendulum”—is a well-knownexample of an unstable system that can be stabilized byoscillatory forcing. Averaging methods and asymptotic stabilityresults can be applied to develop a general framework fordesigning suitable inputs. Linearizing the Kapitza pendulumyields Mathieu’s equation, which is also extensively studiedfor its stability characteristics.In this paper, results from thesetwo bodies of work are compared, from the point of viewof open-loop control of mechanical systems. The averagingapproaches applied to Mathieu’s equation are seen to accessonly a very limited portion of the stability map. Furthermore,stabilizing input signals exist that may be found from thelinear stability map, but are not found by averaging methods.The results suggest that the linear stability map is a powerfuland underutilized tool for design and analysis of open-looposcillatory control.

I. INTRODUCTION AND MOTIVATION

The results of this paper are motivated by studies of sidepull-in of MEMS comb drive actuators [22], [23], [24]. Sidepull-in is an instability that occurs in constant-gap elec-trostatic devices, when the drive voltage exceeds a criticalvalue. Side pull-in occurs in the direction orthogonal to thedesired actuator travel, as opposed to standard pull-in, whichoccurs along the direction of desired actuator travel. Thosestudies were attempting to prevent or delay side pull-in, andthus allow operation of the device at higher drive voltages,corresponding to increased stroke or force. As in manyMEMS control problems, the need for sensing, actuation,and/or computation must be balanced against complexity ofdesign and fabrication. For the side pull-in problem, thenominal plant is uncontrollable and unstable beyond thepull-in point. Additional sensors and actuators are neededto stabilize those operating points using feedback controlmethods. The goal of those studies was to stabilize operationbeyond the pull-in voltage without redesigning the device—that is, without the use of feedback.

In 1951, the Russian physicist P. L. Kapitza published ananalysis of the stability of the inverted pendulum with a verti-cally oscillating base [10]. At that time it was already knownthat the upright equilibrium point may be stabilized forforcing with sufficiently high frequency and sufficiently lowamplitude [18]. Kapitza applied successive approximation toobtain an “effective moment” that captured the average effectof the base motion. Kapitza’s analysis is an iconic applicationof averaging theory to oscillatory stabilization; the verti-cally driven inverted pendulum is now sometimes called the

I. P. M. Wickramasinghe and Jordan M. Berg are with the Depart-ment of Mechanical Engineering and the Nano Tech Center, Texas TechUniversity, Lubbock, TX, 79409, USA, e-mail: m.wickramasinghe, [email protected].

Kapitza pendulum. Approaches to oscillatory stabilizationbased on averaging have received extensive attention acrossmany disciplines, including control theory. As observed in aninfluential 1980 paper by S. M. Meerkov [13], this form ofstabilization is appealing from a control perspective becauseit requires no measurement of the system states or outputs,nor does it require feedforward of commands or disturbances.Thus it would seem ideally suited to address side pull-inof comb drives. Meerkov named this strategy “vibrationalcontrol,” and classified it as a separate category of control,alongside feedback and feedforward methods. In the yearssince Meerkov’s seminal publication, control researchershave further developed open-loop oscillatory stability andcontrol of linear and nonlinear systems using averaging,time-scale separation, and asymptotic stability theorems [2],[3], [4], [1], [21], [6].

It has previously been reported that bidirectional forcingis required for the successful use of averaging techniques inMEMS devices [15], [16], [17], [19]. These methods weresuccessfully applied to stabilize a voltage-controlled gap-closing electrostatic actuator. A novel device design wasrequired to overcome the unidirectional nature of capacitivelycoupled Coulomb forces [16], [19]. This result is a majoradvance over feedback control, despite the added complexityof the redesigned device, since it eliminates the need forsensing. However the goal of the constant-gap problemreported in [22], [23], [24] was feedback-free stabilizationof the nominal device, with no special design features at all.

Therefore averaging methods were adapted to the combdrive model. However, it was found that while stabilizingsolutions could be found, they did not significantly delay theside pull-in bifurcation. However, it was also observed thatlinearization of the transverse comb drive dynamics yieldsMathieu’s equation. Mathieu’s equation has been extensivelyanalyzed using Floquet theory—which does not require av-eraging or time-scale separation—and is known to exhibitboth parametric resonance and oscillatory stabilization [9],[5]. The linearized model can be stabilized well beyondthe side pull-in point through consideration of the resultingstability map, and simulations show that the resulting inputsignal also stabilizes the nonlinear system. Furthermore, theresult includes an explicit three-way trade-off between themagnitude of the input frequency, the amount by which thestable operating range is extended, and the precision withwhich the stabilizing forcing signal is applied. This is incontrast to the averaging methods, which guarantee stabilityonly in the limit as the forcing frequency increases. This is anadvantage of the Floquet approach that may have importantpractical consequences. We note that the linearization of the

2013 IEEE/ASME International Conference onAdvanced Intelligent Mechatronics (AIM)Wollongong, Australia, July 9-12, 2013

978-1-4673-5320-5/13/$31.00 ©2013 IEEE 686

Page 2: Vibrational Control of Mathieu's Equation

Kapitza pendulum is also Mathieu’s equation, but that theredoes not seem to have much overlap between the literatureon these topics. Two exceptions, neither from the controlliterature, are [5], [7]. Kapitza himself remarks that analysisof Mathieu’s equation can find the same stability result ashe obtains, but that the analysis technique is unnecessarilycomplicated, provides no new knowledge, and is limited tosmall motions. In these remarks he seems either unconcernedor unaware of the complete stability diagram that the fullanalysis of Mathieu’s equation provides [10].

It is natural to ask why the averaging approaches failto find these solutions. This paper reports the results offurther investigation, in which averaging is applied directlyto Mathieu’s equation. It ifinds that the averaging methodsaccess only a small part of the stability behavior, but that thisfact is masked by the effects of time scaling. Two previouslyreported stabilization results for Mathieu’s equation and theKapitza pendulum are revisited, and it is seen that theoutcomes in those papers have been misinterpreted [13], [1].

The Floquet-based method can use piecewise constantforcing as well as sinusoidal forcing. The approach is cur-rently formulated for second-order systems, and it is not clearhow the results may best be extended to higher-order systemswith more general structure. However we believe that theresults show at least that existing approaches to oscillatorycontrol do not fully capture the range of stability behaviorpossible, and therefore that designers should be willing tolook beyond these approaches when faced with a difficultstabilization challenge.

This paper is organized as follows: Section II presents theKapitza pendulum and its linearization. Section III showshow the stability diagram for Mathieu’s equation may beused to design a suitable stabilizing input signal, with anexplicitly stated trade-off between the location of the newbifurcation point and the frequency of the excitation. SectionIV considers the result of applying averaging theorems fromthe literature to Mathieu’s equation. Section V presents asimplified version of the side pull-in bifurcation control prob-lem,and shows why averaging methods fail while methodsbased on the stability diagram succeed. Section VI is apreliminary result showing that the stability diagram canalso be used, for example, to design a large-angle open-loop swing-up controller for the Kapitza pendulum. Finally,Section VII presents our conclusions.

II. KAPITZA’S PENDULUM AND MATHIEU’S EQUATION

The equation of motion of a one degree-of-freedomdamped inverted pendulum with vertical base excitation isof the form [10], [5]

θ + Cθ − (A+B sinωt) sin θ = 0, (1)

where ˙(•) denotes the derivative with respect to t and Ais positive when θ = 0 corresponds to the upward pointingequilibrium. Scaling time by τ = t/ω gives

θ′′ +C

ωθ′ −

(A

ω2+B

ω2sin τ

)sin θ = 0 (2)

−1 0 1 2 3 4 5 6 7 8 9 100

1

2

3

4

5

6

7

8

α

β

Fig. 1. Stability diagram for Mathieu’s equation. Green regions indicatestability. Unstable regions for α > 0 correspond to parametric resonance.Stable regions for α < 0 correspond to oscillatory stabilization.

−1 0 1 2 30

1

2

3

4

5

6

7

−0.25 −0.2 −0.15 −0.1 −0.05 0 0.05 0.1 0.15 0.20

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

β

α

β = 0 unstable β = 0 stable

lower stability limit

upper stability limit

asymptotic stabilityregion

Fig. 2. Stability diagram for Mathieu’s equation at small α. For fixedα < 0 there is a range of β values from within the blue-shaded regionthat stabilize the system. This range has both a lower and an upper stabilityboundary.

where (•)′ denotes the derivative with respect to τ .Linearization of (1) and (2) about the equilibrium at θ = 0

is immediate and yields

y + Cy − (A+B sinωt) y = 0, (3)

and

y′′ +C

ωy′ −

(A

ω2+B

ω2sin τ

)y = 0, (4)

respectively. Equation (4) is the damped Mathieu equation[20]. Setting

α = −A/ω2 (5)

β = B/ω2 (6)δ = C/ω (7)

we obtain a standard form of the damped Mathieu’s equation,

y′′ + δy′ + (α− β sin τ) y = 0. (8)

The use of sine or cosine in the time-varying coefficient doesnot affect the stability. In this paper we will use whicheveris most convenient. The undamped forms are easily obtainedby setting the damping coefficients to zero.

687

Page 3: Vibrational Control of Mathieu's Equation

The stability of the damped or undamped Mathieu’s equa-tion may be investigated in terms of parameters α and βby the use, for example, of Hill’s determinant [9]. Figure1 shows the resulting well-known stability diagram for theundamped case, where parameter values corresponding tobounded solutions are shaded green. In the right half-plane,where α is positive, the β = 0 solutions are stable, while inthe left half-plane, where α is negative, the β = 0 solutionsare unstable. The regions in the right half-plane that areunstable for some for β > 0 correspond to the phenomenonknown as parametric resonance. In this paper our focus ison the regions in the left half-plane that are stable for someβ > 0, which correspond to oscillatory stabilization.

Figure 2 expands the stability diagram for the region nearthe origin. From the figure it can be seen that for a fixedα there is a minimum β that is required for stability. Thatis, there is a minimum amplitude for stabilizing sinusoidalforcing. This is referred to here as the lower stability limit.From the figure it is also clear that there is an upper limit onthe forcing amplitude β above which the system is once againunstable. This is referred to here as the upper stability limit.In fact there is a sequence of increasing β intervals for whichthe system is stable, due to other stable branches crossingthe β axis. These intervals become increasingly narrow as βincreases. For the remainder of this paper we consider thefirst stable region in the left half-plane, which is shaded bluein Fig. 2.

Figures 1 and 2 suggest that a method for choosing a sta-bilizing amplitude and frequency for the Kapitza pendulumshould account for the upper stability bound as well as thelower stability bound. In fact, the great majority of work onthis subject finds only the lower stability bound. However,the upper bound is not an artifact of the linearization, but alsoappears in the Kapitza pendulum, as predicted by Mathieu’sequation. Furthermore, a more general nonlinear methodshould also work when applied to Mathieu’s equation. Doingso can be an enlightening exercise. In the context of (1) or(3), the amplitude of B can become large, but in the contextof the time-scaled forms (2) or (4), these methods find onlysolutions that approach the origin of the α–β parameter spacefrom within the dark blue region in Fig. 2.

III. OSCILLATORY CONTROL DESIGN USING THELINEAR STABILITY MAP

Consider the stability map shown in Fig. 2. As α movesto the left, the corresponding interval of stabilizing valuesof β becomes narrower. This will make it more difficult togenerate the precise stabilizing signal, and will make thesystem more vulnerable to modeling errors.With this in mind,consider the following design procedure:

1) Choose a point (α∗, β∗) from the stable region in Fig.2. Note that A is positive and α∗ is negative. Thelarger the value of A, the more unstable is the nominalsystem.

2) For a desired value of A, a stabilizing frequency willbe ω =

√−A/α∗. Higher frequencies are necessary

to stabilize a higher degree of static instability, but thisis a specified value, not an unspecified threshold.

3) The amplitude of the signal is then B = ω2β∗.4) Therefore the stabilizing signal can be given in terms

of the desired operating point as

B sin(ωt) =

(β∗A

α∗

)sin

√− A

α∗t (9)

Figure 3 shows the amplitude response of both the dampedMathieu’s equation (8) and the damped Kapitza pendulum(1) with C = 0.01, and A equal to 1, 10, and 100. Theamplitude and frequency of the oscillatory input have beendesigned following (9) with (α∗, β∗) = (−0.25, 0.8).

IV. VIBRATIONAL CONTROL OF MATHIEU’S EQUATION

In this section we consider side pull-in control usingaveraging methods. Applying the methods of [2], [3], [4],[1], [6] to the linear dynamics (3), gives the result that aninput of the form

B sinωt = ωB1 sinωt (10)

will stabilize the origin of (3) for some sufficiently largevalue of ω. That is,

y + Cy + (−A+ ωB1 sinωt) y = 0, (11)

will be stable for sufficiently large ω. The magnitude of thesignal scales as ω, and so such inputs have been called “largeamplitude, high frequency” [6]. However, consider the time-scaled version of (11) in the standard form of Mathieu’sequation:

y′′ +C

ωy′ +

(− A

ω2+B1

ωsin τ

)y = 0. (12)

From (12) we see that as ω becomes large, both α = −A/ω2

and β = B1/ω become arbitrarily small, and the (α, β)point approaches the origin from within the dark blue shadedregion in Fig. 2. Therefore, although the control input (10)is stabilizing, it accesses only a very limited portion of thestability diagram.

From (12) it can also be seen that an input of the form

B cosωt = ω2B2 sinωt (13)

would be required to reach these other regions of the stabilitydiagram. In fact, [1] and [13] propose controls of this form,with the claim that such a form of the input is stabilizing forsufficiently large ω. Example 2 of [1] states that the system

y −(A+ ω2B2 sinωt

)y = 0 (14)

is stable ifω2 > −2A

B22

. (15)

However, time-scaling this system gives

y′′ −(A

ω2+B2 sin τ

)y = 0. (16)

Figure 4(a) shows the inequality (15) interpreted in theα–β stability diagram. A sufficiently small choice of B2

688

Page 4: Vibrational Control of Mathieu's Equation

0 200 400 600 800 1000 1200−8

−6

−4

−2

0

2

4

6x 10

−3

time

angle (rad/pi)

(a) Mathieu’s Equation, A = 1

0 200 400 600 800 1000 1200−2

−1.5

−1

−0.5

0

0.5

1

1.5

2

2.5x 10

−3

time

angle (rad/pi)

(b) Kapitza Pendulum, A = 1

0 200 400 600 800 1000 1200−2. 5

−2

−1. 5

−1

−0. 5

0

0. 5

1

1. 5

2x 10

−3

time

angle (rad/pi)

(c) Mathieu’s Equation, A = 10

0 200 400 600 800 1000 1200−2

−1.5

−1

−0.5

0

0.5

1

1.5

2x 10

−3

time

angle (rad/pi)

(d) Kapitza Pendulum, A = 10

0 200 400 600 800 1000 1200−8

−6

−4

−2

0

2

4

6

8x 10

−4

time

angle (rad/pi)

(e) Mathieu’s Equation, A = 100

0 200 400 600 800 1000 1200−2

−1.5

−1

−0.5

0

0. 5

1

1. 5

2x 10

−3

time

angle (rad/pi)

(f) Kapitza Pendulum, A = 100

Fig. 3. Simulation results using the oscillatory input of (9) with (α∗, β∗) = (−0.25, 0.8). In all cases there is damping, with C = 0.01

corresponds to β1. In this case, satisfying (15) does guaranteestability. However, for a larger value of B2, for example theβ2 value shown in the figure, the frequency ω must also sat-isfy an additional inequality that captures the upper stabilitylimit. Thus Example 2 of [1] is not a valid application of thetheorems developed in that paper. In fact, Theorem 3, whichis invoked in Example 2, explicitly assumes an input of theform ωB1 sinωt.

Theorem 2 of [13] concerns inputs of the form u(t) =B0 sinωt. Applied to (3), that theorem yields the statementthat if

1

2

(B0

ω

)2

> A, (17)

then there exist B∗0 > 0 and ω∗ > 0 such that for all B0 >B∗0 and ω > ω∗, the system

y + Cy − (A+B sinωt) y = 0, (18)

is stable. In this case, the statement can be interpreted inthe α–β plane as shown in Fig. 4(b). This figure is obtainedby considering the case when B0 is held fixed and ω isincreased (solid lines), and the case when ω is held fixed andB0 is increased. In the first case, the relations α = −A/ω2,β = B0/ω

2 give β = −(B0/A)α. In the second case,α remains constant while β increases. From this figure itcan be seen that the theorem as stated misses the upperstability limit, since for (α∗, β∗) corresponding to any choiceof (B∗0 , ω

∗), there is a point at which the (α, β) values leavethe stability region when B0 is increased by too much. Theinequality (17) handles the cases where the (α, β) pointdrops below the lower stability limit. Thus we see fromapplication to Mathieu’s equation that two previous resultsfrom the literature obtained using averaging approaches havebeen misinterpreted, and must be modified to account for the

689

Page 5: Vibrational Control of Mathieu's Equation

a)−0.25 −0.2 −0.15 −0.1 −0.05 00

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

β

α

β2

β1

b)−0.25 −0.2 −0.15 −0.1 −0.05 00

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

β

α

(α*, β*)

Fig. 4. a) Interpretation in the α–β plane of Example 2 of [1]. Theconclusion of the Example holds for β1 but fails for β2. b) Interpretationin the α–β plane of Theorem 2 of [13] as applied to Mathieu’s equation.Solid lines show (α, β) values obtained by holding B0 > B∗

0 fixed andincreasing ω, dashed lines show (α, β) values obtained by holding ω > ω∗

fixed and increasing B0. The statement of the theorem must have a flaw,as seen by the dashed lines crossing the upper stability limit.

upper stability limit.

V. OSCILLATORY CONTROL OF SIDE PULL-IN

The problem of controlling side pull-in has special featuresthat make it impossible to treat by the standard averagingresults. These arise because the electrostatic force appearsas the square of the control voltage. The result is that theα and β terms are coupled, and cannot be independentlyspecified. To see this, consider the following linear systemwhich shows the salient elements in a simplified form:

x+ Cx+ x = u2 (19)

y + Cy + y = u2y. (20)

The x-direction is the desired actuation direction, with equi-librium x = u2. Therefore larger u corresponds to largeractuator displacement. However, for static u the lateral y-dynamics are destabilized at u = 1. This is the side pull-in point. The goal of the side pull-in control problem isto replace the constant u by a time-varying term u(t) thatwill enable the actuator to have a larger equilibrium xdisplacement without destabilizing the y direction. To avoidthe need for additional sensors, u(t) must not depend on y.

We focus on the y dynamics, and choose a control of theform u(t) = B sinωt. Neglecting damping, (20) becomes,

y +(1−B2 sin2 ωt

)y = 0, (21)

or,

y +

(1− B2

2+B2

2cos 2ωt

)y = 0. (22)

To analyze the stability of the lateral dynamics via Mathieu’sequation, we denote 2ω by Ω in (21), and apply time scaling,to obtain

y′′ +

(1

Ω2− B2

2Ω2+

B2

2Ω2cos τ

)y = 0. (23)

The problem with applying standard methods of vibra-tional control as presented in [2], [3], [4], [1], [6] to this

system is the coupling between the constant term and thezero-mean periodic term. To see this, choose B =

√2ΩB1

in (21) to obtain the standard form of the control:

y + (1− ΩB1 + ΩB1 cos Ωt) y = 0. (24)

Now the usual procedure in applying one of the methodsfrom the literature would be to increase Ω until the systemwas stabilized. However, the presence of Ω in the constantterm as well as the periodic input invalidates those results. Infact, consider the inequality that defines the lower stabilitylimit, β2/2 > −α. For α and β arising in (24), this inequalitybecomes,

B21 > 2ΩB1 − 2. (25)

Thus we see that there can be no asymptotic stabilityguarantee for large Ω. Increasing Ω increases the actuatorstroke, however the lateral dynamics will become unstable.Therefore these methods are unsuited to the side pull-inproblem.

Now consider a technique based on the stability map. From(23) we have

α =

(1

Ω2− B2

2Ω2

)(26)

β =B2

2Ω2(27)

So for some choice of stabilizing (−α∗, β∗), we have

Ω =1√

β∗ − α∗(28)

B = Ω√

2β∗ =

√2β∗

β∗ − α∗(29)

Recall that the objective of the side pull-in problem is tomaintain stability while increasing the control amplitude.Here that is accomplished by increasing B. For any B thereis a stabilizing signal with

Ω = B/√β∗ (30)

In principal, this completely eliminates side pull-in, andallows the application of arbitrarily large control voltages.Rather than use unnecessarily high forcing frequencies, thesmallest possible frequency can be chosen. Further, by choos-ing β∗ as large as possible, the necessary frequency can befurther reduced.

VI. OPEN-LOOP OSCILLATORY SWING-UP CONTROL OFTHE INVERTED PENDULUM

Examination of Fig. 1 and 2 show that there are points forwhich Mathieu’s equation at (−α, β) is stabilized by oscilla-tory stabilization, and Mathieu’s equation at (α, β) is desta-bilized by parametric resonance. For example, (0.08, 0.5) isone such point. This raises the question of whether it ispossible to swing up from the downward pointing position,to end at the stable upward-pointing position. This questionwas investigated by numerical simulation.

As seen in Fig. 5, simulations showed that the undampedpendulum departs from initial positions near the downward

690

Page 6: Vibrational Control of Mathieu's Equation

0 50 100 150 200 250–0.5

0

0.5

1

2

2.5i

anglevelocity

1.5

angl

e (r

ad/π)

time(sec)

Fig. 5. Open-loop swing-up control of the Kapitza pendulum. Nonlinear,time-varying damping is used to capture the pendulum at the uprightequilibrium point.

pointing unstable equilibrium, but does not settle at the up-ward pointing stable equilibrium. Adding constant dampingdid not improve the capture at the upward pointing point.Experimenting with different types of damping led to thechoice of viscous damping with a nonlinear, time-varyingdamping coefficient of the form δz θ with

δz = 10−6 +

e−t/20 cos(θ) > 0.9

0 otherwise(31)

This is an admittedly contrived and impractical form ofdamping. However it is still remarkable that an open-loopswing-up controller can be designed largely on the basis ofthe linearized dynamics at the two equilibrium points.

VII. CONCLUSIONS

We have applied nonlinear methods for the design ofstabilizing time-varying inputs to the Kapitza pendulum, andto its linearization, Mathieu’s equation. We have comparedthe results to the stability map obtained for Mathieu’s equa-tion using Floquet theory. The results show that the non-linear methods, which typically rely on asymptotic stabilitytheorems, address only a small part of the stability map.Design based directly on the stability map for Mathieu’sequation can access a much more extensive range of controlsignals. The consequences for design are threefold. First, theresult suggests that design for second-order systems baseddirectly on linearization and application of Floquet theory isa strategy worth further investigation. Second, the benefits ofthis method seem to warrant efforts to extend it for higher-order systems. Third, the nonlinear averaging methods thathave been developed should be re-examined in light of theseresults, to see if the limitations that have been revealed arefundamental to the methods, or can be removed by furtherdevelopment.

REFERENCES

[1] J. Baillieul, “Energy methods for stability of bilinear systems withoscillatory inputs,” International Journal of Robust and NonlinearControl, Vol. 5, No. 4, pp. 285-301, 1995.

[2] R. Bellman, J. Bentsman and S. M. Meerkov, “On vibrational stabi-lizability of nonlinear systems,” Journal of Otimization Theory andApplications, Vol. 46, No. 4, pp. 421–430, Aug. 1985.

[3] R. Bellman, J. Bentsman and S. M. Meerkov, “Vibrational controlof nonlinear systems: Vibrational stabilizability,” IEEE Tr. AutomaticControl, Vol. AC-31, No. 8, pp. 710–716, Aug. 1986.

[4] J. Bentsman, “Vibrational control of a class of nonlinear systems bynonlinear multiplicative vibrations,” IEEE Tr. Automatic Control, Vol.AC-32, No. 8, pp. 711–716, Aug. 1987.

[5] J. A. Blackburn, H. J. T. Smith and N. Gronbech-Jensen, “Stabilityand Hopf bifurcations in an inverted pendulum,” Amer. J. Phys., Vol.60, No. 10, pp. 903–908, Oct. 1992.

[6] F. Bullo, “Averaging and vibrational control of mechanical systems,”SIAM Journal on Control and Optimization, Vol. 41, No. 2, pp. 542-562, 2002.

[7] E. I. Butikov, “An improved criterion for Kapitza’s pendulum stability,”Journal of Physics A: Mathematical and Theoretical, Vol. 44, No. 29,295202, 2011.

[8] J. A. Holyst and W. Wojciechowski,“The effect of Kapitza pendulumand price equilibrium,” Physica A: Statistical Mechanics and itsApplications, Vol. 324, No. 1, pp. 388-395, 2003.

[9] D. W. Jordan and P. Smith, Nonlinear Ordinary Differential Equations,Oxford uni. press (4th edition), pp. 315–330, 2007.

[10] P. L. Kapitza, “Dynamic stability of the pendulum with vibratingsuspension point (1951),” Collected Papers of P. L. Kapitza, Vol. 2,Pergamon, pp. 714–726, 1965.

[11] M. Levi and W. Weckesser, “Stabilization of the inverted pendulum byhigh frequency vibrations,” SIAM Review, Vol. 37, No. 2, pp. 219–223,June, 1995.

[12] W. Magnus and S. Winkler, Hill’s Equation, Dover (2004 reprint of1979 edition), 1979.

[13] S. M. Meerkov, “Principle of vibrational control: Theory and applica-tions,” IEEE Trans. Automat. Control, vol. AC-25, no. 8, pp. 755–762,Aug. 1980.

[14] J. Mitchell, “Techniques for the Oscillated Pendulum and the MathieuEquation,” http://www.math.ou.edu/ npetrov/joe-report.pdf

[15] K. Nonaka, J. Baillieul, and M. Horenstein, “Open loop robustvibrational stabilization of a two wire system inside the snap-throughinstability region,” Proceedings of the CDC,Orlando,Florida, 2001.

[16] K. Nonaka, K. Sugimoto and J. Baillieul, “Bi-directional extensionof the travel range of electrostatic actuators by open loop periodicallyswitched oscillatory control,” Proceedings of the CDC, San Diego,California, 2004.

[17] K. Nonaka, K. Sakai and J. Baillieul, “ Open loop oscillatory controlfor electromagnetic actuated microgrippers,” Proceedings of the SICE2004 Annual Conference, Sapporo, Japan, 2004.

[18] A. Stephenson, “On an induced stability,” Phil. Mag., 15, 233236,1908.

[19] T. Sugimoto, K. Nonaka, and M. N. Horenstein, “Bidirectional Elec-trostatic Actuator Operated with Charge Control,” IEEE Journal ofMicroelectromechanical Systems, Vol. 14, No. 4, 718–724, August2005.

[20] J. H. Taylor and K. S. Narendra, “Stability regions for the dampedMathieu equation,” SIAM Journal on Applied Mathematics, Vol. 17,No. 2, pp. 343-352, 1969.

[21] A. R. Teel, J. Peuteman and D. Aeyels, “Semi-global practical asymp-totic stability and averaging,” Systems & control letters, Vol. 37, No.5, pp. 329-334, 1999.

[22] I. P. M. Wickramasinghe and J. M. Berg, “Lateral Stability of aPeriodically Forced Electrostatic Comb Drive,” Proceedings of the2012 American Control Conference, Montreal, Canada, June 2012.

[23] I. P. M. Wickramasinghe and J. M. Berg,“On the Periodic SymmetricElectrostatic Forcing of a Microcantilever,” Proceedings of the 2012Advanced Intelligent Mechatronics Conference, Kaohsiung, Taiwan,July 2012.

[24] I. P. M. Wickramasinghe and J. M. Berg,“Delay of Side Pull-In for anElectrostatic Comb Drive Model with Rotational Degree of Freedom,”Proceedings of the 2013 American Control Conference, to appear.

[25] H. Yabuno, “Bifurcation Control of Kapitza Pendulum,” Instite ofEngineering Mechanics and Systems, University of Tsukuba, 1369, pp.121-127, 2004.

691