microfluidic circuit dynamics and control for caterpillar-inspired...

8
Microfluidic Circuit Dynamics and Control for Caterpillar-Inspired Locomotion in a Soft Robot Helene Nguewou-Hyousse 1 , Giulia Franchi 2 , and Derek A. Paley 3 Abstract— Because caterpillars can locomote through com- plex terrain, caterpillar-inspired soft robots promise safe and reliable access for search, rescue, and exploration. Microfluidic components analogous to electronic components like resistors and capacitors further promise the principled design and fabrication of autonomous soft robots powered by microfluidic circuits. This paper presents a model for caterpillar locomotion using an oscillator network approach, in which the periodic motion of each leg is described by an oscillator, and the collection of all legs forms a network. We use tools from graph theory to model (1) a first-order system consisting of a network of RC oscillators connected to an astable multivibrator circuit serving as a Central Pattern Generator (CPG); and (2) a second-order system consisting of a network of RLC oscillators with a deCentralized Pattern Generator (dCPG) as reference control. For the RC network, we analyze the rate of convergence and the gait number, a metric that characterizes the locomotion gait. For the RLC network, we design feedback laws to stabilize consensus and traveling wave solutions. Numerical modeling and preliminary experimental results show promise for the design and control of locomotion circuitry in a caterpillar- inspired soft robot. I. INTRODUCTION Robots made from soft materials are more flexible and may be more capable of adapting to complex environments than rigid robots. As such, their applications range from search and rescue to exploration and medicine [1], [2]. Soft robots inspired by organisms without a skeleton may also be easier to design and fabricate than vertebrate-inspired robots, because of the level of abstraction offered by the simplicity of their anatomy [3]. A caterpillar is an example of a soft-bodied animal that can navigate irregular and vertical terrains [4]. Caterpillar locomotion exhibits (i) the synchronous motion of left-right pairs of legs, and (ii) a wave propagation pattern traveling from tail to head [4]–[6]. Thus, a caterpillar-inspired soft-robotic system should be capable of collective behavior such as synchronization and wave propagation. Furthermore, choosing appropriate fabrication material [3] and structure geometry should guarantee its flexibility. *This work is supported by ARO Grant No. W911NF1610244 1 Helene Nguewou-Hyousse is a graduate student in the Department of Electrical and Computer Engineering, University of Maryland, College Park, MD 20742, USA [email protected] 2 Giulia Franchi is Assistant Professor of Computer Science in the Department of Mathematics and Computer Science, Salisbury University, Salisbury, MD 21801, USA [email protected] 3 Derek A. Paley is the Willis H. Young Jr. Professor of Aerospace Engineering Education in the Department of Aerospace Engineering and the Institute for Systems Research, University of Maryland, College Park, MD 20742, USA [email protected] The motivation behind this work is the design and closed-loop control of a novel locomotion mechanism in a caterpillar-inspired soft robot. We seek to design a 3D- printed caterpillar robot with internal microfluidic circuitry that drives leg actuation.When an actual caterpillar moves, each left-right pair of legs is lifted simultaneously and goes through a stance phase and a swing phase [5], [6]. In addi- tion, a wave is repeatedly generated at the terminal proleg and travels to the front leg, creating a periodic behavior that can be modeled by an oscillator network with a cyclic topology. Electrical circuits may be used to represent microfluidic circuits using Equivalent Circuit (EC) Theory [7]. Although not all electrical components have a microfluidic analog (e.g., there are no microfluidic inductors due to low inertial effects at the millimeter scale), EC supports the design of microfluidic circuits and robots. For example, the Octobot is an octopus-inspired autonomous soft robot with microfluidic logic fabricated using soft lithography [1]. However, recent trends in microfluidics have demonstrated the viability of 3D- printed circuit components [8]; using this technique to print a microfluidic soft robot would reduce the need for circuit interconnections and may simplify design and fabrication. An astable multivibrator [1], [9], for example, is an oscillat- ing circuit that can be implemented with either electric or microfluidic components. This paper considers a coordinated network of fluidic oscillators generating and controlling periodic leg actua- tion. Previous work on modeling caterpillar-inspired loco- motion in a soft body applied inverse-dynamics to a planar extensible-link model [10], or lumped-mass dynamic model coupled with a control law for gait optimization [2], [4]. Autonomous decentralized control is a solution to generating and controlling local and adaptive locomotion in soft robots [11], [12]. In particular, Umedachi and Trimmer imple- mented an autonomous decentralized control to extend and contract segments of a 3D printed modular soft robot; the results show that implementing modular deformity resulted in a phase gradient that induced locomotion [11]. Prior work in the area of collective motion of networked oscillators used a spatial approach to look at the phases of each oscillator [13], [14]. Here we investigate feedback control of an oscillator network to generate cyclic actuation of many individual legs. We study both first- and second- order dynamical models. For the first-order model, which consists of a multivibrator circuit driving a set of RC oscil- lators in series, we introduce a gait number to characterize the locomotion pattern. For the second-order model, which

Upload: others

Post on 10-Apr-2020

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Microfluidic Circuit Dynamics and Control for Caterpillar-Inspired ...cdcl.umd.edu/papers/ccta18.pdf · tion. Previous work on modeling caterpillar-inspired loco-motion in a soft

Microfluidic Circuit Dynamics and Control for Caterpillar-InspiredLocomotion in a Soft Robot

Helene Nguewou-Hyousse1, Giulia Franchi2, and Derek A. Paley3

Abstract— Because caterpillars can locomote through com-plex terrain, caterpillar-inspired soft robots promise safe andreliable access for search, rescue, and exploration. Microfluidiccomponents analogous to electronic components like resistorsand capacitors further promise the principled design andfabrication of autonomous soft robots powered by microfluidiccircuits. This paper presents a model for caterpillar locomotionusing an oscillator network approach, in which the periodicmotion of each leg is described by an oscillator, and thecollection of all legs forms a network. We use tools fromgraph theory to model (1) a first-order system consisting of anetwork of RC oscillators connected to an astable multivibratorcircuit serving as a Central Pattern Generator (CPG); and (2) asecond-order system consisting of a network of RLC oscillatorswith a deCentralized Pattern Generator (dCPG) as referencecontrol. For the RC network, we analyze the rate of convergenceand the gait number, a metric that characterizes the locomotiongait. For the RLC network, we design feedback laws to stabilizeconsensus and traveling wave solutions. Numerical modelingand preliminary experimental results show promise for thedesign and control of locomotion circuitry in a caterpillar-inspired soft robot.

I. INTRODUCTION

Robots made from soft materials are more flexible andmay be more capable of adapting to complex environmentsthan rigid robots. As such, their applications range fromsearch and rescue to exploration and medicine [1], [2]. Softrobots inspired by organisms without a skeleton may alsobe easier to design and fabricate than vertebrate-inspiredrobots, because of the level of abstraction offered by thesimplicity of their anatomy [3]. A caterpillar is an exampleof a soft-bodied animal that can navigate irregular andvertical terrains [4]. Caterpillar locomotion exhibits (i) thesynchronous motion of left-right pairs of legs, and (ii) a wavepropagation pattern traveling from tail to head [4]–[6]. Thus,a caterpillar-inspired soft-robotic system should be capableof collective behavior such as synchronization and wavepropagation. Furthermore, choosing appropriate fabricationmaterial [3] and structure geometry should guarantee itsflexibility.

*This work is supported by ARO Grant No. W911NF16102441Helene Nguewou-Hyousse is a graduate student in the Department of

Electrical and Computer Engineering, University of Maryland, College Park,MD 20742, USA [email protected]

2Giulia Franchi is Assistant Professor of Computer Science in theDepartment of Mathematics and Computer Science, Salisbury University,Salisbury, MD 21801, USA [email protected]

3Derek A. Paley is the Willis H. Young Jr. Professor of AerospaceEngineering Education in the Department of Aerospace Engineering andthe Institute for Systems Research, University of Maryland, College Park,MD 20742, USA [email protected]

The motivation behind this work is the design andclosed-loop control of a novel locomotion mechanism ina caterpillar-inspired soft robot. We seek to design a 3D-printed caterpillar robot with internal microfluidic circuitrythat drives leg actuation.When an actual caterpillar moves,each left-right pair of legs is lifted simultaneously and goesthrough a stance phase and a swing phase [5], [6]. In addi-tion, a wave is repeatedly generated at the terminal prolegand travels to the front leg, creating a periodic behaviorthat can be modeled by an oscillator network with a cyclictopology.

Electrical circuits may be used to represent microfluidiccircuits using Equivalent Circuit (EC) Theory [7]. Althoughnot all electrical components have a microfluidic analog(e.g., there are no microfluidic inductors due to low inertialeffects at the millimeter scale), EC supports the design ofmicrofluidic circuits and robots. For example, the Octobot isan octopus-inspired autonomous soft robot with microfluidiclogic fabricated using soft lithography [1]. However, recenttrends in microfluidics have demonstrated the viability of 3D-printed circuit components [8]; using this technique to printa microfluidic soft robot would reduce the need for circuitinterconnections and may simplify design and fabrication.An astable multivibrator [1], [9], for example, is an oscillat-ing circuit that can be implemented with either electric ormicrofluidic components.

This paper considers a coordinated network of fluidicoscillators generating and controlling periodic leg actua-tion. Previous work on modeling caterpillar-inspired loco-motion in a soft body applied inverse-dynamics to a planarextensible-link model [10], or lumped-mass dynamic modelcoupled with a control law for gait optimization [2], [4].Autonomous decentralized control is a solution to generatingand controlling local and adaptive locomotion in soft robots[11], [12]. In particular, Umedachi and Trimmer imple-mented an autonomous decentralized control to extend andcontract segments of a 3D printed modular soft robot; theresults show that implementing modular deformity resultedin a phase gradient that induced locomotion [11].

Prior work in the area of collective motion of networkedoscillators used a spatial approach to look at the phasesof each oscillator [13], [14]. Here we investigate feedbackcontrol of an oscillator network to generate cyclic actuationof many individual legs. We study both first- and second-order dynamical models. For the first-order model, whichconsists of a multivibrator circuit driving a set of RC oscil-lators in series, we introduce a gait number to characterizethe locomotion pattern. For the second-order model, which

Page 2: Microfluidic Circuit Dynamics and Control for Caterpillar-Inspired ...cdcl.umd.edu/papers/ccta18.pdf · tion. Previous work on modeling caterpillar-inspired loco-motion in a soft

consists of a network of RLC circuits in a cyclic topology,we modify the network to include a deCentralized PatternGenerator (dCPG) as the decentralized control. For example,in consensus control of multiple oscillators, the dCPG setsthe collective phase, but the phase lags if the network itselfis lagging.

The contributions of this paper are (1) the design of a first-order RC oscillator network model for generating caterpillar-inspired locomotion in a microfluidic circuit, and the charac-terization of the resulting locomotion gait; (2) the design ofa second-order RLC oscillator network model, and closed-loop feedback control laws for consensus and traveling wavesolutions; and (3) the inclusion of a deCentralized PatternGenerator to regulate the collective phase of the consensussolution in the RLC network. This work demonstrates coop-erative behavior of a network of oscillator circuits compatiblewith microfluidics; it also proposes a model for the closed-loop control design of an RLC network to produce stablecyclic motion. These steps provide a theoretical approach tothe study of legged locomotion in an autonomous soft robotusing microfluidics; we also include preliminary results fromour ongoing experimental testbed.

The paper is organized as follows. Section II presentsan overview of RLC oscillators, algebraic graph theory,first- and second-order consensus dynamics, and microfluidiccircuit theory. Section III-A provides a mathematical frame-work for the RC network model and presents an analysisof its convergence properties, including a parametrizationof the locomotion gait. Section III-B contains a mathemat-ical framework for the RLC network, including open-loopstability analysis, closed-loop control design for consensusand traveling wave solutions, and the effect of the dCPG.Section IV presents preliminary experimental results fromongoing work with the RC microfluidic network. Section Vsummarizes the paper and our future work.

II. BACKGROUND

The proposed work models each leg of a caterpillar asan RLC (or RC) circuit. The network of N caterpillar legsis connected in either a circulant or a chain topology. Inthe microfluidic testbed, pressure is analogous to voltage,whereas volumetric flow rate is analogous to current. In apneumatic setting, each microfluidic leg will be actuated byinternal pressure; we use Equivalent Circuit theory to designthese pressure inputs using electrical circuits. This sectionfirst reviews RLC and RC circuits, followed by a backgroundoverview of graph theory, consensus dynamics, and the fieldof microfluidics.

A. RLC and RC Circuit Theory

Kirchoff’s loop equation for a series RLC circuit withresistance r, inductance l, capacitance c, current i, and inputvoltage V (t) is vr + vc + vl = V (t), where vl = ldi/dt,vr = ri and vc = 1

c

∫ t−∞ i(τ)d(τ) [15]. Defining state

variables v = vc and a = vc, the system dynamics have

the following linear state-space representation:d

dt

[va

]=

[0 1− 1lc − rl

] [va

]+

[01lc

]V (t).

The circuit oscillates if it is underdamped, i.e., when r2c <4l. In the absence of an inductor, i.e., in a microfluidic circuit,we have a resistor-capacitor (RC) circuit in series, which isgoverned by the first-order equation

cv +v

r=

1

rV (t).

The time constant of τ1 = rc represents the time requiredfor the voltage at the capacitor to drop to v(0)/e.

B. Algebraic Graph TheoryGraph theory is used to represent the topology of the

oscillator network in which each leg is represented as a node.The Laplacian matrix L defines the connections between thenodes of the network. Given the undirected graph G = (V, E)with vertex set V of cardinality N and edge set E , we writethe Laplacian matrix as L = D − A. The adjacency matrixA ∈ RN×N is symmetric, with entries aij = aji given by

aij =

1, if i 6= j and (i, j) ∈ E0, otherwise.

The degree matrix D ∈ RN×N has the ijth entry

dij =

∑Nj=1 aij , if i = j

0, otherwise.

The Laplacian of an undirected graph is a symmetric, pos-itive semi-definite matrix with zero row sums. For example,the Laplacian matrix of an undirected chain graph with Nedges is

L1 =

1 −1 0 · · · 0−1 2 −1 0 · · · 00 −1 2 −1 0 · · · 0...

. . . . . . . . . . . ....

.... . . . . . . . . . . .

.... . . −1 2 −1

0 0 · · · 0 −1 1

∈ RN×N ,

(1)whereas the Laplacian matrix of an undirected cyclic graphwith degree 2 is

L2 =

2 −1 0 · · · −1−1 2 −1 0 · · · 00 −1 2 −1 0 · · · 0...

. . . . . . . . . . . ....

.... . . . . . . . . . . .

.... . . −1 2 −1

−1 0 · · · 0 −1 2

∈ RN×N .

(2)Note that L2 is a circulant matrix, i.e, each row vector is acyclic permutation of the preceding row vector.

Page 3: Microfluidic Circuit Dynamics and Control for Caterpillar-Inspired ...cdcl.umd.edu/papers/ccta18.pdf · tion. Previous work on modeling caterpillar-inspired loco-motion in a soft

C. Consensus Theory

Depending on the network topology, it is often possibleto achieve collective behavior via consensus. Consensusrefers to a group of agents achieving a common decisionor behavior over time through a local distributed protocol[16]–[18]. Consider the linear consensus scheme [19]

v = −Lv,

where v = [v1, · · · , vN ]T ∈ RN and L is a graph Laplacian.Previous work [19], [20] showed that this system eventuallyachieves first-order consensus, i.e., vj = vk for all pairs j andk, if the associated communication graph G has a spanningtree. G has a directed spanning tree if and only if L has asimple zero eigenvalue (with associated eigenvector 1N ) andall other eigenvalues have positive real parts. Furthermore,e−Lt → 1NpT and vi(t) →

∑Ni=1(pivi(0)) as t → ∞,

where p = [p1, · · · , pN ]T ≥ 0 is a left eigenvector of Lassociated with the simple zero eigenvalue and

∑Ni=1 pi = 1.

Consider N coupled harmonic oscillators with local inter-action of the form [16]

d

dt

[va

]=

[0N IN−αIN −L

] [va

],

where 0 denotes the N × N zero matrix , I denotes theN ×N identity matrix, and L is a Laplacian matrix associ-ated with directed graph G. There exists right eigenvector1N satisfying L1N = 0 and left eigenvector p ∈ RNsatisfying p ≥ 0, pTL = 0, and pT1N = 1. TheN agents achieve second-order synchronization if G hasa directed spanning tree [16]. More specifically, vi(t) →cos(√αt)pT v(0) + (1/

√α) sin

(√αt)pTa(0) and ai(t) →

−√α sin

(√αt)pT v(0) + cos

(√αt)pTa(0) as t→∞.

D. Microfluidic Circuit Theory

Microfluidics refers to the precise manipulation of fluidicnetworks at the submillimiter level [21], [22]. Microfluidiccomponents include diodes, capacitors, transistors, and hy-draulic resistances. Microfluidic circuits can be analyzedusing computational fluid dynamics (CFD) analysis, althoughas the number of circuit components to be modeled increases,the CFD analysis becomes complex. Electric circuit theoryis an alternative technique that uses the analogy betweenfluidic and electronic components to describe the dynamicsof a microfluidic circuit using electrical equations [22],[7]. Pressure, volumetric flow rate, and hydraulic resistanceare equivalent to voltage, current, and electric resistance,respectively. By making the assumption that the fluidic flowis laminar, viscous, and incompressible, the use of electriccircuit theory greatly simplifies the analysis of the flowdynamics in microfluidic circuits. Microfluidic circuit theoryapplied to robotics is relevant because it promises to enablethe design and fabrication of autonomous microfluidic softrobots; moreover, using 3D-printing techniques to fabricatethe microfluidic components and integrate with a soft bodymay allow the fabrication of an autonomous fully integratedsoft robot [21].

Fig. 1: Schematic of an astable multivibrator connected toan RC network with N = 4

III. DYNAMICAL MODELS

This section contains the mathematical framework govern-ing the dynamics of two oscillator networks. Section III-Adescribes an RC network driven by a multivibrator oscillator,and Section III-B describes an RLC network with controlinputs at each node. The RLC network incorporates a dCPGas reference input to regulate the collective phase.

A. First-Order System: RC Network

Motivated by the existence of microfluidic transistors,capacitors, and resistances, this design includes an astablemultivibrator connected to N resistor-capacitor (RC) circuitsas shown in Fig. 1. To create a phase delay between each RCcircuit, the network has a chain topology in which each nodeis connected to its two nearest neighbors, with the exceptionof the first element, which is connected to the multivibratorand one neighbor, and the last element, which is connectedto only one neighbor. The multivibrator is a transistor circuitwhose output is a square wave that oscillates continuouslybetween high and low values [9]. When one transistor (Q1)is switched on, the other (Q2) is switched off. At the sametime, capacitor C1 discharges while C2 charges.

Let L1 be the Laplacian matrix associated with a chaintopology, and vi be the voltage associated with capacitor i.Let v = (v1, · · · , vN )T and v0 be the voltage at the transistorQ2, i.e., the output of the multivibrator. During the high(or low) output of the multivibrator, v0 is constant, and thedynamics of this network system are

v = − 1

rcL1v +

1

rc(v0 − v1),

where L1 is the Laplacian matrix defined in (1). Usingaugmented state v = (v0, · · · , vN )T and Laplacian L1 ∈RN+1×N+1, the dynamics become

˙v = − 1

rcL1v, (3)

Page 4: Microfluidic Circuit Dynamics and Control for Caterpillar-Inspired ...cdcl.umd.edu/papers/ccta18.pdf · tion. Previous work on modeling caterpillar-inspired loco-motion in a soft

0 0.5 1 1.5 2 2.5 3 3.5

time (s)

-1

0

1

2

3

4

5

6

Voltage (

V)

v0

v1

v2

v3

v4

Fig. 2: Evolution of N = 4 RC oscillators with CPG v0

where

L1 =

0 · · · 0−1 2 −1 0 · · · 00 −1 2 −1 0 · · · 00 0 −1 2 −1 0 · · · 0...

.... . . . . . . . . . . .

......

.... . . . . . . . . . . .

......

. . . −1 2 −10 0 0 · · · 0 −1 1

.

Equation (3) essentially includes the astable multivibra-tor as a Central Pattern Generator that drives the networkdynamics. Fig. 2 shows the results of a Matlab simulationwith v0 = 5.2V , r = 1 kΩ, R = 10 kΩ, and C1 = C2 =c = 0.1 mF . Note that, for the parameters chosen, the RCoscillators do not reach consensus, but instead show a phaselag in the rising voltage of consecutive oscillators.

We formalize this idea as follows. The period of an astablemultivibrator is τ0 = ln 2(R1C1 +R2C2) and τ1 = rc is thetime constant of the RC circuit. Define the gait number

η =τ0Nτ1

(4)

to characterize the caterpillar-inspired gait with N legs.Let C1 = C2 = c and R1 = R2 = R. Then, η =2 ln (2)R/(Nr). Observations from simulations at various ηvalues show that when η < 1, the contribution of the frontlegs to the caterpillar locomotion is negligible. On the otherhand, when η > 1, all legs contribute to the locomotion.Furthermore, as η increases, the voltage outputs of legs1 through N increases to v0(t), and the phase differencebetween the leg nodes decreases to 0.

B. Second-Order System: RLC Network

Next, consider a second-order model that represents eachleg of the caterpillar as an RLC circuit. Although this modelis currently unattainable in microfluidic circuitry, because of

Fig. 3: Network of N RLC circuits connected through anedge resistance

the lack of microfluidic inductors, it has several attractivefeatures for control design of networked oscillators. In par-ticular, the RLC model illustrates how feedback at the circuitlevel might be used to achieve caterpillar-inspired locomotionincluding consensus and traveling wave solutions.

Consider an undirected graph with a circulant topologyof N nodes. This graph describes the network of identicalRLC oscillators coupled through an edge resistance depictedin Fig. 3. Suppose the kth RLC series circuit has resistancerk = r, inductance lk = l, capacitance ck = c, inputvoltage Vk(t) and edge resistance rjk = r. In addition,the state of oscillator k is [vk ak]T , k = 1, . . . , N . Letv = [v1, . . . , vN ]T and a = [a1, . . . , aN ]T , as before, sothat the dynamics of the oscillator are

d

dt

[va

]= M

[va

]+

[01lcI

]V (t), (5)

where

M =

[0 I

− 1lcI −

rlcrL2 − rl I −

1crL2

],

V (t) = [V1(t), · · · , VN (t)]T is the collection of voltageinputs, one per node, and L2 is the circulant Laplacian matrixdefined in (2).

Let wr = [xr, yr]T be a right eigenvector of M associated

with eigenvalue λ, such that Mwr = λwr. We have

yr = λxr(−1

lcI − r

lcrL2

)xr − λ

(r

lI +

1

crL2

)xr = λ2xr,

which implies

L2xr = −crλ2 + λ rl + 1

lcrl + λ︸ ︷︷ ︸

xr. (6)

Let µ be an eigenvalue of L2 associated with the vectorχr = xr. λ is a solution of

crλ2 +(rcrl

+ µ)λ+

r

l+r

lµ = 0.

For the eigenvalue µi of L2 associated with eigenvector χri,we can therefore define an eigenvalue of M with associatedright eigenvector wri = [χri, λi±χri]

T as follows:

λi± =−( rcrl + µi)±

√( rcrl + µi)2 − 4cr( rl + r

l µi)

2cr.

Page 5: Microfluidic Circuit Dynamics and Control for Caterpillar-Inspired ...cdcl.umd.edu/papers/ccta18.pdf · tion. Previous work on modeling caterpillar-inspired loco-motion in a soft

In a similar way, if we let wz = [xz, yz]T be the left

eigenvectors of M associated with the eigenvalue λ suchthat wT

zM = λwTz , we have

yTz

(−1

lcI − r

lcrL2

)= λxTz

xzT − yz

T

(r

lI +

1

crL2

)= λyz

T ,

which implies

yzTL2 = cr

λ2 + λ rl + 1lc

rl + λ

yzT , (7)

Note that (7) is similar to (6). Let χz = yz bethe left eigenvector of L2 associated with eigenvalueµ. Thus, the corresponding left eigenvector of M iswzi=[1/λ(−1/(lc)I − r/(lcr)L2)χzi χzi]

T . We have thefollowing result.

Lemma 1: λi± has negative real part, which implies theorigin of the open-loop system (5) is exponentially stable.

Proof: Let ∆ = ( rcrl + µi)2 − 4cr( rl + r

l µi).case: ∆ > 0

µmax <rcr

l− 2r

√c

l

µmin >rcr

l+ 2r

√c

l

Using the results of Section II-A, we conclude that this caseis impossible, since rcr

l < 2r√

cl and the eigenvalues of the

Laplacian are positive (µ > 0).case: ∆ < 0

µmax <rcr

l+ 2r

√c

l

µmin >rcr

l− 2r

√c

l

The condition on µmin is always satisfied since µmin > 0.In addition, Re(λi±) < 0, and hence λi± is complex andnegative.

Lemma 1 shows the origin of the open-loop system (5)is stable and the oscillations decay over time. As such, con-sensus may not be achieved before the oscillations die out.However, we seek to be able to sustain a rhythmic patternover time. In order to achieve this result, we design a voltagefeedback control that effectively cancels the resistance ofeach RLC, so as to emulate the behavior of LC circuits.

Consider the following state-feedback control, which en-sures the individual oscillators in Fig. 3 converge to acommon solution over time, corresponding to the caterpillarlegs oscillating in synchrony. Assume a circulant topologywhere every node is connected both to its predecessor andsuccessor nodes, and the head to the tail. A feedback inputto (5) of the form

V (t) =r

rL2v + rca (8)

0 0.5 1 1.5 2 2.5 3 3.5

time (s)

-1.5

-1

-0.5

0

0.5

1

1.5

Voltage (

V)

v1

v2

v3

v4

v5

1 2 3 4 5

Nodes

-1.5

-1

-0.5

0

0.5

1

1.5

Voltage (

V)

time 0 s

time 0.62 s

time 1.54 s

time 2.53 s

time 3.5 s

Fig. 4: N=5 RLC oscillators with consensus control

results in the closed-loop system

d

dt

[va

]=

[0 I− 1lcI − 1

crL2

]︸ ︷︷ ︸

,M1

[va

]

Let ω20 = 1/(lc). According to [16, Theorem 3.1], the

system reaches consensus for large values of t. Specifically,

vk(t)→ 1

N[cos(ω0t)1

T v(0) +1

ω0sin(ω0t)1

Ta(0)]

ak(t)→ 1

N[−ω0 sin(ω0t)1

T v(0) + cos(ω0t)1Ta(0)],

where (1/N)1 = p ∈ RN satisfies p ≥ 0, pTL2 = 0,and pT1N = 1. Each oscillator evolves at its own phase andamplitude until the network reaches a consensus (see Fig. 4).The consensus amplitude is a weighted average of the initialconditions, and the phase is arbitrary. Matlab simulationswere computed using r = 0.5 Ω, c = 0.1 F , l = 0.1 H ,and r = 0.5 Ω. Note the consensus phase is determined byinitial conditions.

Page 6: Microfluidic Circuit Dynamics and Control for Caterpillar-Inspired ...cdcl.umd.edu/papers/ccta18.pdf · tion. Previous work on modeling caterpillar-inspired loco-motion in a soft

To regulate the phase of the output voltages under the con-sensus control we invoke a deCentralized Pattern Generator(dCPG). The dCPG is a rhythmic pattern generator designedas a virtual LC circuit labeled k=0, that takes weak feedbackfrom the network. Node k = 0 is connected to node k = 1.Let r10 = r denote the coupling strength from node 0 to node1, whereas r01 = εr is the coupling strength from node 1 tonode 0. The dynamics of node 0 are

v0 +ε

lc(v0 − v1) =

1

lcV0(t).

The modified network dynamics, including the dCPG node,are described using the (N + 1 )-dimensional Laplacian ma-trix

L2 =

ε −ε 0 0 0 · · · 0 0−1 3 −1 0 0 · · · 0 −10 −1 −2 −1 0 · · · 0 0...

. . . . . . . . . . . ....

.... . . . . . . . . . . .

......

. . . . . . . . . . . ....

.... . . −1 2 −1

0 −1 0 · · · −1 2

Let v = [v0, · · · , vN ]T , a = [a0, · · · , aN ]T , as before, andr = diag(0, r, · · · , r)T . The dynamics of the N + 1 networkof oscillators with dCPG becomes

d

dt

[va

]= M

[va

]+

[01lc I

]V (t), (9)

where

M =

[0 I

− 1lc I −

rlcr L2 − rl I −

1cr L2

],

V (t) = [V0(t), · · · , VN (t)]T is the collection of voltageinputs, and 0 and I are the corresponding zero and identitymatrices.

Consider a feedback input similar to (8), i.e.,

V (t) =r

rL2v + rca,

which results in the closed-loop system

d

dt

[va

]=

[0 I

− 1lc I − 1

cr L2

]︸ ︷︷ ︸

,M2

[va

].

Under the closed-loop control with dCPG, the networkreaches consensus with node k=0. If ε > 0, the networkof oscillators influence the dCPG, whereas if ε = 0,(vk(t), ak(t)) converge to (v0(t), a0(t)) as t→∞ (see Fig.5), and

v0(t) = cos(ω0t)v0(0) +1

ω20

sin(ω0t)a0(0)

a0(t) = −ω0 sin(ω0t)v0(0) + cos(ω0t)a0(0).

0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5

time (s)

-1.5

-1

-0.5

0

0.5

1

1.5

Voltage (

V)

v0 at =1

v1

v2

v3

v4

v5

v0 at =0

Fig. 5: Evolution of N = 5 RLC oscillators with dCPG andconsensus feedback control

Next, consider a feedback controller that achieves a stabletraveling wave, i.e., the output propagates around the oscil-lator network like the traveling wave observed in caterpillarlocomotion. The feedback input to (5)

V (t) = v + rca +l

rL2a,

yields v = a, and a = −r/(lcr)L2v, corresponding to theclosed-loop system

d

dt

[va

]=

[0 I

−rlcrL2 0

]︸ ︷︷ ︸

,M3

[va

](10)

Because the network is undirected and circulant, theLaplacian L2 is the finite-difference approximation of a2nd-order spatial derivative. Specifically, at node k, vk =d2vk/dt

2 and L2,kv = −vk−1 + 2vk − vk+1, where L2,k isthe kth row of L2. Let xk denote the position of node k andv(x, t) a spatially continuous representation of v(t), e.g., forlarge N . We have

−L2,kv = vk−1 − 2vk + vk+1 ≈ ∆x2 d2v

dx2

∣∣∣∣x=xk

where ∆x is the (uniform) spacing between the legs of thecaterpillar. Thus, (10) approximates the partial differentialwave equation

∂2v

∂t2=

r

lcr∆x2︸ ︷︷ ︸,γ

∂2v

∂x2,

which has solution v(x, t) = 1/2[f(x+ γt) + f(x− γt)

]for v(x, 0) = f(x) [23]. Choosing a(x, 0) = (γ/2)(df/dx)yields a one-directional wave, as shown in Fig. 6.

Page 7: Microfluidic Circuit Dynamics and Control for Caterpillar-Inspired ...cdcl.umd.edu/papers/ccta18.pdf · tion. Previous work on modeling caterpillar-inspired loco-motion in a soft

-1 -0.8 -0.6 -0.4 -0.2 0 0.2 0.4 0.6 0.8 1

normalized position

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1pro

babili

ty d

ensity function

t=0 s

t=0 s

t=5 s

t=10 s

t=15 s

t=20s

t=25 s

Fig. 6: Propagation of wave in time

IV. EXPERIMENTAL RESULTSWe fabricated an astable multivibrator (see Fig. 1) using

3D-printed microfluidic transistors and microfluidic capac-itors [21]. For the resistors, we used various lengths ofthe connector (Tygon microbore tubing, 06420-03, ColeParmer, Vernon Hills, IL)). The internal resistors were threetimes larger than the external ones, which gave the ca-pacitors the opportunity to fully charge before switching.We used the MAESFLO system (Fluigent, Paris, France),which includes a Microfluidic Flow Control System anda FLOWELL microfluidic flow sensor, and the SensirionSLI-1000 flow sensor to regulate the input pressures whilesimultaneously monitoring the flow rates through the 3D-printed devices. We used Tygon microbore tubing and micropipe fittings to connect the circuit components. We conductedall experiments at room temperature environment (20–25

C).To examine the performance of the circuit, we connected

the input point to the Fluigent system and the output tothe flow unit. Using the Fluigent MAESFLO software, weincrementally increased the pressure within the circuit. Onceall the air bubbles were out of the circuit and the pressurestabilized at a constant value (usually around 200–250 mbar),we collected data from the flow unit, plotted in Fig. 7. Thesepreliminary results indicate the output pressure oscillates asexpected. Moreover, the results correlates that of an elec-tronic astable multivibrator with capacitance C1 = C2 = 0.1mF, resistance R1 = R2 = 10 kΩ, and a period τ0 ≈ 0.6932seconds.

V. CONCLUSIONCaterpillar locomotion enables robust, albeit slow, naviga-

tion of irregular terrains. It is initiated from the tail to thefront, and each left-right pair of legs is lifted simultaneously.We used RC and RLC oscillators to mimic the locomotorypattern of the caterpillars as a precursor to a microfluidicfabrication design. We characterized the locomotion gait inthe RC network, and designed a closed-loop control for

Fig. 7: Experimental set-up and multivibrator output pressurefrom experiment with constant input pressure 230 mbar.

consensus and traveling wave solution in the RLC network.A dCPG was successfully implemented as an input to theRLC network to regulate the collective phase. Preliminaryexperimental results for the microfluidic testbed are included.In ongoing work, we seek to implement the RC circuit designin microfluidics and design fluidic leg actuators for a soft-robotic caterpillar.

VI. ACKNOWLEDGEMENTS

We acknowledge support of Joshua Hubbard and RyanSochol for the microfluidic experiments.

REFERENCES

[1] M. Wehner, R. L. Truby, D. J. Fitzgerald, B. Mosadegh, G. M.Whitesides, J. A. Lewis, and R. J. Wood, “An integrated designand fabrication strategy for entirely soft, autonomous robots,” Nature,vol. 536, no. 7617, pp. 451–455, 2016.

[2] F. Saunders, E. Golden, R. D. White, and J. Rife, “Experimentalverification of soft-robot gaits evolved using a lumped dynamicmodel,” Robotica, vol. 29, no. 6, pp. 823–830, 2011.

[3] R. F. Shepherd, F. Ilievski, W. Choi, S. A. Morin, A. A. Stokes, A. D.Mazzeo, X. Chen, M. Wang, and G. M. Whitesides, “Multigait softrobot,” Proceedings of the National Academy of Sciences, vol. 108,no. 51, pp. 20400–20403, 2011.

[4] D. W. Schuldt, J. Rife, and B. Trimmer, “Template for robust soft-bodycrawling with reflex-triggered gripping,” Bioinspiration & biomimet-ics, vol. 10, no. 1, p. 016018, 2015.

[5] B. A. Trimmer, A. E. Takesian, B. M. Sweet, C. B. Rogers, D. C.Hake, and D. J. Rogers, “Caterpillar locomotion: A new model forsoft-bodied climbing and burrowing robots,” in 7th InternationalSymposium on Technology and the Mine Problem, vol. 1, pp. 1–10,Monterey, CA: Mine Warfare Association, 2006.

[6] L. van Griethuijsen and B. Trimmer, “Locomotion in caterpillars,”Biological Reviews, vol. 89, no. 3, pp. 656–670, 2014.

Page 8: Microfluidic Circuit Dynamics and Control for Caterpillar-Inspired ...cdcl.umd.edu/papers/ccta18.pdf · tion. Previous work on modeling caterpillar-inspired loco-motion in a soft

[7] S. Vedel, Millisecond dynamics in microfluidics: Equivalent circuittheory and experiment. PhD thesis, DTU Nanotech, Department ofMicro and Nanotechnology, 2009.

[8] R. D. Sochol, E. Sweet, C. C. Glick, S. Venkatesh, A. Avetisyan, K. F.Ekman, A. Raulinaitis, A. Tsai, A. Wienkers, K. Korner, K. Hanson,A. Long, B. J. Hightower, G. Slatton, D. C. Burnett, T. L. Massey,K. Iwai, L. P. Lee, K. S. J. Pister, and L. Lin, “3D printed microfluidiccircuitry via multijet-based additive manufacturing,” Lab on a Chip,vol. 16, pp. 668–678, 2016.

[9] A. S. Sedra and K. C. Smith, Microelectronic circuits theory andapplications. Oxford University Press.

[10] F. Saunders, B. A. Trimmer, and J. Rife, “Modeling locomotion ofa soft-bodied arthropod using inverse dynamics,” Bioinspiration &biomimetics, vol. 6, no. 1, p. 016001, 2010.

[11] T. Umedachi and B. A. Trimmer, “Autonomous decentralized controlfor soft-bodied caterpillar-like modular robot exploiting large andcontinuum deformation,” in Intelligent Robots and Systems (IROS),2016 IEEE/RSJ International Conference on, pp. 292–297, IEEE,2016.

[12] T. Umedachi, K. Takeda, T. Nakagaki, R. Kobayashi, and A. Ishiguro,“Fully decentralized control of a soft-bodied robot inspired by trueslime mold,” Biological cybernetics, vol. 102, no. 3, pp. 261–269,2010.

[13] T. Vicsek and A. Zafeiris, “Collective motion,” Physics Reports,vol. 517, no. 3, pp. 71–140, 2012.

[14] D. A. Paley, N. E. Leonard, R. Sepulchre, D. Grunbaum, and J. K.Parrish, “Oscillator models and collective motion,” IEEE ControlSystems, vol. 27, no. 4, pp. 89–105, 2007.

[15] R. C. Dorf and J. A. Svoboda, Introduction to electric circuits. JohnWiley & Sons, 2010.

[16] W. Ren, “Synchronization of coupled harmonic oscillators with localinteraction,” Automatica, vol. 44, no. 12, pp. 3195–3200, 2008.

[17] W. Ren and R. W. Beard, Distributed consensus in multi-vehiclecooperative control. Springer, 2008.

[18] R. Sepulchre, D. A. Paley, and N. E. Leonard, “Stabilization of planarcollective motion: All-to-all communication,” IEEE Transactions onAutomatic Control, vol. 52, no. 5, pp. 811–824, 2007.

[19] W. Ren, R. W. Beard, and T. W. McLain, “Coordination variablesand consensus building in multiple vehicle systems,” in CooperativeControl, pp. 171–188, Springer, 2005.

[20] R. W. Beard and V. Stepanyan, “Information consensus in distributedmultiple vehicle coordinated control,” in 42nd IEEE Conference onDecision and Control, vol. 2, pp. 2029–2034, IEEE, 2003.

[21] R. D. Sochol, E. Sweet, C. C. Glick, S. Venkatesh, A. Avetisyan, K. F.Ekman, A. Raulinaitis, A. Tsai, A. Wienkers, K. Korner, K. Hanson,A. Long, B. J. Hightower, G. Slatton, D. C. Burnett, T. L. Massey,K. Iwai, L. P. Lee, K. S. J. Pister, and L. Lin, “3D printed microfluidiccircuitry via multijet-based additive manufacturing,” Lab on a Chip,vol. 16, no. 4, pp. 668–678, 2016.

[22] K. W. Oh, K. Lee, B. Ahn, and E. P. Furlani, “Design of pressure-driven microfluidic networks using electric circuit analogy,” Lab on aChip, vol. 12, no. 3, pp. 515–545, 2012.

[23] S. J. Farlow, Partial differential equations for scientists and engineers.Courier Corporation, 1993.