A Passivity-Based Approach for Trajectory Tracking and ...

8
A Passivity-Based Approach for Trajectory Tracking and Link-Side Damping of Compliantly Actuated Robots Manuel Keppler 1 , Dominic Lakatos 1 , Christian Ott 1 , Alin Albu-Sch¨ affer 2 Abstract— This paper presents a control method to imple- ment trajectory tracking and disturbance rejection charac- teristics for the link-side dynamics of compliantly actuated robots with nonlinear spring characteristics. This is achieved by introducing new motor coordinates reflecting the damping and feedforward terms and shaping the dynamics of the motor such that it structurally equals the dynamics in the original coordinates. Thus, the approach achieves the control goal while changing the original plant dynamics only to a minimum extent. Passivity, stability, and convergence properties of the closed- loop dynamics are proven. The performance of the control approach has been experimentally evaluated on the variable stiffness robot arm DLR Hand Arm System, where the stiffness in each of the joints is highly nonlinear. To our best knowledge, this is the first experimentally validated tracking controller for compliantly actuated robots with nonlinear elastic elements. I. I NTRODUCTION To handle impacts and unknown contact forces, robot design recently evolved from rigid towards compliant actu- ators having real, generally nonlinear springs in their power train [1]. The elastic elements acting between motor and link inertias together with the intrinsic damping properties of any physical system lead to a dynamical behavior of a low-pass filter on external loads. As a consequence high force peaks resulting, for instance, from unexpected envi- ronmental collisions act attenuated on gear box and motor parts preventing the destruction of these parts. Therefore, the elastic elements can significantly increase the mechanical robustness on the one side, but they also introduce oscillatory dynamics [2] into the plant on the other side. Especially, if the contact impedance matches according to requirements of mechanical robustness, resulting frequencies of the oscilla- tory dynamics are in a range which is undesired for accurate link motion tracking control. To benefit from the mechanical robustness properties of compliantly actuated robots while simultaneously achieving a desired tracking dynamics of the manipulator links, damping and the required feedforward terms have to be added to the system by control. Solutions to the regulation problem of link configuration variables have been proposed in [3], [4], [5], and [6]. These methods are exclusively introduced for so-called flexible joint robots (FJR), where elasticities in links and power trains are modelled as linear springs (i. e., constant stiffness), concentrated in the joints. A generalization to the case of nonlinear elasticities as appearing in many implementations 1 The authors are with the Institute of Robotics and Mechatronics, German Aerospace Center (DLR), D-82234 Oberpfaffenhofen, Germany. Contact: [email protected] 2 Alin Albu-Sch¨ affer is also with Technical University Munich, Chair of Sensor Based Robots and Intelligent Assistance Systems, Department of Informatics, D-85748 Garching, Germany. spring deflection φ [deg] -20 0 20 stiffness [Nm/rad] 0 200 400 600 800 VARIABLE STIFFNESS ACTUATOR (a) (c) σ Variable Stiffness κ(σ, φ) (b) θ Main Motor q Link Fig. 1. Fig. 1a shows the DLR Hand Arm System. Fig. 1b shows a simplified working scheme of a single VSA. Fig. 1c depicts the stiffness characteristics for stiffness adjuster positions σ =[02.55.07.5 10 ] deg, the outer curve corresponds to σ =0 deg. of variable stiffness actuators (VSA) [7] has been proposed in [8]. As all above methods consider only collocated feed- back 1 , the convergence performance depends strongly on the joint stiffness, i. e., low stiffness required for impact attenuation leads to low damping performance. Regulation controller with high damping performance considering also feedback of non-collocated variables are reported in [9], [10], [11]. While the former approach [9] (applicable to FJR) provides a comprehensive stability analysis, the latter methods [10], [11] (applicable to VSA robots) are not proven to be stable yet. The three latter controllers are link-side damping controllers for the regulation case which are of course of less performance if the set-point is not constant. Solutions to the tracking problem of robots with elastic transmissions are reported in the pioneering works [12], [13], [14]. Further relevant tracking controllers are based on cascaded structures [15], integrator backstepping [16], [17, Chap. 6.2], extensions of the well known Slotine and Li controller [18] to the flexible joint case [19], feedback linearization [12], [20], [21], and integral manifold control [12]. The above mentioned papers either derive their results only for the FJR case (except the method in [21] which applies to VSA robots) or validate their controllers only in computer simulations (except in [15] experimental results of a single FJR joint). The same is true for the comprehensive 1 Control input collocated variables, are variables for which its time derivatives form a conjugate power pair together with the control input (torque).

Transcript of A Passivity-Based Approach for Trajectory Tracking and ...

Page 1: A Passivity-Based Approach for Trajectory Tracking and ...

A Passivity-Based Approach for Trajectory Tracking and Link-SideDamping of Compliantly Actuated Robots

Manuel Keppler1, Dominic Lakatos 1, Christian Ott 1, Alin Albu-Schaffer 2

Abstract— This paper presents a control method to imple-ment trajectory tracking and disturbance rejection charac-teristics for the link-side dynamics of compliantly actuatedrobots with nonlinear spring characteristics. This is achievedby introducing new motor coordinates reflecting the dampingand feedforward terms and shaping the dynamics of the motorsuch that it structurally equals the dynamics in the originalcoordinates. Thus, the approach achieves the control goal whilechanging the original plant dynamics only to a minimum extent.Passivity, stability, and convergence properties of the closed-loop dynamics are proven. The performance of the controlapproach has been experimentally evaluated on the variablestiffness robot arm DLR Hand Arm System, where the stiffnessin each of the joints is highly nonlinear. To our best knowledge,this is the first experimentally validated tracking controller forcompliantly actuated robots with nonlinear elastic elements.

I. INTRODUCTION

To handle impacts and unknown contact forces, robotdesign recently evolved from rigid towards compliant actu-ators having real, generally nonlinear springs in their powertrain [1]. The elastic elements acting between motor andlink inertias together with the intrinsic damping propertiesof any physical system lead to a dynamical behavior of alow-pass filter on external loads. As a consequence highforce peaks resulting, for instance, from unexpected envi-ronmental collisions act attenuated on gear box and motorparts preventing the destruction of these parts. Therefore,the elastic elements can significantly increase the mechanicalrobustness on the one side, but they also introduce oscillatorydynamics [2] into the plant on the other side. Especially, ifthe contact impedance matches according to requirements ofmechanical robustness, resulting frequencies of the oscilla-tory dynamics are in a range which is undesired for accuratelink motion tracking control. To benefit from the mechanicalrobustness properties of compliantly actuated robots whilesimultaneously achieving a desired tracking dynamics of themanipulator links, damping and the required feedforwardterms have to be added to the system by control.

Solutions to the regulation problem of link configurationvariables have been proposed in [3], [4], [5], and [6]. Thesemethods are exclusively introduced for so-called flexiblejoint robots (FJR), where elasticities in links and powertrains are modelled as linear springs (i. e., constant stiffness),concentrated in the joints. A generalization to the case ofnonlinear elasticities as appearing in many implementations

1The authors are with the Institute of Robotics and Mechatronics, GermanAerospace Center (DLR), D-82234 Oberpfaffenhofen, Germany.Contact: [email protected]

2Alin Albu-Schaffer is also with Technical University Munich, Chair ofSensor Based Robots and Intelligent Assistance Systems, Department ofInformatics, D-85748 Garching, Germany.

spring deflection φ [deg]-20 0 20

stiff

ness

[Nm

/rad

]

0

200

400

600

800

VARIABLESTIFFNESSACTUATOR

(a) (c)

σ

Variable Stiffness κ(σ, φ)

(b) θ

MainMotor

q

Link

Fig. 1. Fig. 1a shows the DLR Hand Arm System. Fig. 1b shows asimplified working scheme of a single VSA. Fig. 1c depicts the stiffnesscharacteristics for stiffness adjuster positions σ = [ 0 2.5 5.0 7.5 10 ] deg,the outer curve corresponds to σ = 0 deg.

of variable stiffness actuators (VSA) [7] has been proposedin [8]. As all above methods consider only collocated feed-back1, the convergence performance depends strongly onthe joint stiffness, i. e., low stiffness required for impactattenuation leads to low damping performance. Regulationcontroller with high damping performance considering alsofeedback of non-collocated variables are reported in [9],[10], [11]. While the former approach [9] (applicable toFJR) provides a comprehensive stability analysis, the lattermethods [10], [11] (applicable to VSA robots) are not provento be stable yet. The three latter controllers are link-sidedamping controllers for the regulation case which are ofcourse of less performance if the set-point is not constant.

Solutions to the tracking problem of robots with elastictransmissions are reported in the pioneering works [12],[13], [14]. Further relevant tracking controllers are basedon cascaded structures [15], integrator backstepping [16],[17, Chap. 6.2], extensions of the well known Slotine andLi controller [18] to the flexible joint case [19], feedbacklinearization [12], [20], [21], and integral manifold control[12]. The above mentioned papers either derive their resultsonly for the FJR case (except the method in [21] whichapplies to VSA robots) or validate their controllers only incomputer simulations (except in [15] experimental results ofa single FJR joint). The same is true for the comprehensive

1Control input collocated variables, are variables for which its timederivatives form a conjugate power pair together with the control input(torque).

Page 2: A Passivity-Based Approach for Trajectory Tracking and ...

link coord.: q

new motor coord.: η

ηd

KP (η − ηd)

KD η

K(η − q) Dq

B M

link coord.: q

motor coord.: θ

K(θ − q)

B M

ACHIEVED CLOSED LOOP DYNAMICS

ORIGINAL PLANT DYNAMICS

τm

ηd . . . des. valueof new motorcoord.

OurApproach

Fig. 2. Design idea shown for the simplest form of a single joint withflexible transmission: adding feedback control τm which adds stability andlink-side damping while changing the original dynamics only to a minimalextent.

overview given in [22] and the supplementing survey givenby [23].

This paper aims at closing this gap by solving the trackingproblem for multi-joint robots with nonlinear elastic trans-mission (see Fig. 1) from a theoretical (stability analysis)and a practical (experimental validation) point of view. Theidea is to exclusively add the damping and feedforwardterms to the dynamics of the link variables without changingthe rigid body dynamics of the links nor the structureof the nonlinear springs. This is achieved by introducingnew coordinates reflecting these damping and feedforwardterms and then expressing the motor dynamics in the newcoordinates while shaping the resulting dynamics such that itstructurally equals the dynamics of the original coordinates.The desired tracking behavior (and disturbance rejection) ofthe link coordinates resulting from the original link dynam-ics with the additional desired damping and feedforwardterms are achieved by pure PD regulation control in thenew coordinates. Since the controlled motor dynamics andthe dynamics of the link variables can be represented asan interconnection of passive subsystems, a comprehensivestability statement can be concluded based on physicallymotivated Lyapunov functions. In contrast to the state ofthe art methods mentioned above, this controller achievesthe exact tracking and damping behavior by changing theoriginal plant dynamics only to a minimum extent. To ourbest knowledge, this is the first link-side tracking controllerof compliantly actuated multi-joint robots with nonlinearelastic transmissions which is theoretically founded (stabilityanalysis) and experimentally verified to perform practically.

II. DESIGN IDEA

Consider the simplest form of a robot joint with flexibletransmission as depicted in Fig. 2a. The motor inertia B

acts via a spring f = K(θ− q) with stiffness K on the linkinertia M , where θ and q are motor and link coordinatesrespectively. The control input is a force τm acting on themotor inertia B. This results in dynamics of the form:

Mq = K(θ − q) (1)

Bθ +K(θ − q) = τm (2)

Since the elasticity is introduced to exploit the energy storingcapabilities, compliant actuators are often designed such thatdamping and friction in parallel to the spring is negligible.As such mechanical robustness against impacts and unknowncontact forces is achieved at one side, but also intrinsicoscillatory dynamics result on the other side. The goal is tofind a feedback control τm which exclusively adds stabilityand a damping term Dq by changing the remaining dynamicsstructurally only to a minimum extent. This is achieved byintroducing a new motor coordinate η satisfying

K(θ − q) = K(η − q)−Dq (3)

which implements the desired damping term. It can be easilyverified that the control law

τm = u−BK−1Dq(3) −Dq (4)

results in the dynamics

Mq = K(η − q)−Dq (5)Bη +K(η − q) = u, (6)

which is visualized in Fig. 2. For simplicity we assumea constant damping gain D. Since the damping is alreadyformulated in the coordinate transformation (3), the problemreduces to regulating η to a desired value ηd, such thatq → qd. This can be achieved using a simple PD controlof the form

u = −KDη −KP (η − ηd) (7)

which results in an entire closed-loop dynamics (see Fig. 2b)

Mq = K(η − q)−Dq (8)Bη +KDη +KP (η − ηd) +K(η − q) = 0. (9)

it can be easily verified that the only equilibrium point isη = q, η = ηd and consequently q = ηd ≡ qd.

The remainder of the paper applies these basic ideas togeneral robotic systems

• with nonlinear elastic transmissions,• with multiple degrees of freedom,• that are subject to gravity

and extends this control concept to• the tracking case

III. PROBLEM STATEMENT

Consider the dynamics of a compliantly actuated robotsatisfying

M(q)q +C(q, q)q = −g(q) +ψ(θ − q) + τext (10)

Bθ +ψ(θ − q) = τm. (11)

Page 3: A Passivity-Based Approach for Trajectory Tracking and ...

In the dynamics of the link-side (10), the inertia matrixM(q) ∈ Rn×n is symmetric and positive definite, the Corio-lis matrix C(q, q) ∈ Rn×n is such that the matrix M − 2Cis skew-symmetric, the vector of generalized gravitationalforces g(q) ∈ Rn×n is derived from the potential Ug(q)such that

g(q) := −(∂Ug(q)

∂q

)T, (12)

and τext ∈ Rn are generalized, externally applied forces. Inthe dynamics of the motors (11), the motor torque τm is thecontrol input, the inertia matrix B ∈ Rn×n is assumed to beconstant, symmetric and positive definite. The structure ofcompliantly actuated robots (10)–(11) represents an under-actuated mechanical system in which only the generalizedmotor coordinates θ ∈ Rn can be directly actuated via thecontrol input τm ∈ Rn. The generalized link coordinatesq ∈ Rn can only be indirectly actuated via the generalizedelastic forces

ψ(θ − q) :=∂Us(φ)T

∂φ

∣∣∣∣φ=θ−q

∈ Rn , (13)

which are derived from the even, strictly convex and positivesemi-definite potential function Us(φ). This form of under-actuation is the major challenge in the control of the statesq, q.

Due to the properties of Us(φ), ψ is a global diffeomor-phism and thereby its inverse f−1 always exists. We alsoassume

Us(φ) = 0⇔ φ = 0. (14)

Remark 1: For VSA the elastic potential2 depends onadditional variables σ ∈ Rn which can be considered asan input to change the characteristics of the stiffness (seeFig. 1c), i.e. Us(φ,σ), such that the generalized elastic forcetakes the form

ψ(θ − q,σ) :=∂Us(φ,σ)T

∂φ

∣∣∣∣φ=θ−q

∈ Rn . (15)

This input can be altered to achieve a certain interactionbehavior intrinsically (i.e., the characteristics of the realsprings are changed). Since the focus of this work is ondamping and tracking control, we consider σ as constantparameters. This reduces the complexity of the equations.

In the following, it is assumed that the potential functionUs(φ) is at least three times continuously differentiable.Further, it is assumed that for the total potential

U(θ, q) = Ug(q) + Us(q − θ) (16)

there exists constants c1, c2, c3, c4, c5 > 0 such that the

2Condition (14) still holds true, i.e. ∀σ ∈ Rn, Us(φ,σ) = 0 ⇔ φ = 0.

inequalities

c1‖q‖2 < qT∂2U(θ, q)

∂q2q < c2‖q‖2 (17)

c3 <

∥∥∥∥∂2U(θ, q)

∂q∂θ

∥∥∥∥ < c4 (18)

c5‖θ‖2 < θT∂2U(θ, q)

∂θ2θ (19)

hold3 ∀q,θ ∈ Rn. These assumptions define the class ofsystems for which the control approach of this paper can beapplied. Physical interpretations of these inequalities can befound in [8].

In this paper we address the problem of finding a controlinput τm which allows to influence the convergence behaviorof q → 0 via a link-side damping term Dq, satisfyingqTDq > 0, ∀q 6= 0, while simultaneously introducinglink-side tracking behavior q(t) → qd(t), where qd ∈ C4.Thereby, the goal is to change the original link dynamics(10) only to a minimum extent by keeping M(q), C(q, q),g(q), and the structure of Us(q− θ) unchanged and addingonly damping and tracking terms.

IV. CONTROLLER DESIGN

A. Desired Link-Side Dynamics

We aim to find a controller τm in (11) such that thebehavior of the resulting compensated system is similar tothe following desired dynamics

M(q)¨q +C(q, q) ˙q =−D ˙q +ψ(η − q) + τext, (20)Bη +ψ(η − q) =u, (21)

u =−KPη −KDη, (22)

where where q ≡ q − qd(t) is the tracking error. D ∈Rn×n represents a symmetric, positive definite and boundeddamping matrix. Note that D can be configuration dependentunder the assumption that D(q,η) is bounded and thatD(q,η) is smooth, as will be required in sec. IV and sec.V.

As will be shown later, the PD controller (22), withthe symmetric, bounded and positive controller gain matri-ces KP ,KD ∈ Rn×n, globally stabilizes the system andachieves tracking behavior q(t) → 0, η(t) → 0 as well asenforces link-side damping such that q(t) = const.⇒ q → 0for the (partially) feedback linearized dynamics (20)–(21).The corresponding steady-state behavior is provided in V-B.Note, in contrast to the PD controller for the regulation casepresented in II, ηd = 04.

The link-side of the new, gravity-free system behaveslike the original system (10) with additional damping andtracking. The motor dynamics in the new motor coordinatesη are be structurally equal to original motor dynamics (11).

3The usual Euclidean norm for vectors and the corresponding inducedmatrix norm is assumed throughout the paper.

4The corresponding reference values ηd of controller (22) result, underthe assumptions made in III, from the closed-loop link-side dynamics byimposing q(t) = qd(t).

Page 4: A Passivity-Based Approach for Trajectory Tracking and ...

B. Coordinate Transformation

By comparing (10) and (20) we can find the relation

ψ(θ − q)− g(q) = ψ(η − q)−Dq − n0 (23)

between the new coordinates η and the original systemvariables of (10)–(11), such that

η = q +ψ−1(ψ(θ − q) +D ˙q − g(q)− n0

), (24)

where5

n0(q, q, t) ≡M(q)qd(t) +C(q, q)qd(t). (25)

Since the inverse of ψ is usually not analytically available(24) cannot be solved directly and η has to be determinednumerically. For the experiments we use fix point iteration(see VIII).

Writing the motor dynamics in the new coordinates ηrequires the second time derivative of θ as function of η.To this end, we differentiate (23) w.r.t. time, i.e.

κ(θ − q)(θ − q)− g(q)

= κ(η − q)(η − ˙q)− ddt

(D ˙q) + n1,(26)

where

n1 (q, q, q, t) ≡ ddt

(n0(q, q, t)) = M(q)qd(t)

+M(q)q(3)d (t) + C(q, q)qd(t) +C(q, q)qd(t)

(27)

and

κ(φ0) ≡ ∂2Us(φ)

∂φi∂φj

∣∣∣∣φ=φ0

=∂ψ(φ)

∂φ

∣∣∣∣φ=φ0

∈ Rn×n (28)

is the local stiffness, i.e. the Hessian of the potential Us(φ).Using the second time derivative of (23)

κ(θ − q)(θ − q) + κ(θ − q)(θ − q) =

κ(η − q)(η − ˙q) + κ(η − q)(η − ¨q)

− d2

dt2(D ˙q) + g(q) + n2,

(29)

where

κ(φ0) =ddtκ(φ(t)) =

(n∑k=1

∂κ

∂φkφk

)∣∣∣∣∣φ=φ0

∈ Rn×n

(30)

and

n2(q, q, q, q(3), t) ≡ d2

dt2n0 = M(q)qd + 2M(q)q

(3)d

+M(q)q(4)d + C(q, q)qd + 2C(q, q)qd +C(q, q)q

(3)d ,

(31)

5With better readability in mind the arguments of n0 and its timederivatives (n1, n2; see below) will be left out in the remainder of thistext, once they have been introduced.

allows us to express θ as function of η, η and η

θ =q + κ−1(θ − q)

(− κ(θ − q)(θ − q)

+ κ(η − q)(η − ˙q) + κ(η − q)(η − ¨q)

− d2

dt2(D ˙q) + g(q) + n2

).

(32)

Relations for η and η can be found in VIII.Relations for η and η as function of θ, θ and θ, θ, θ,

respectively, can be derived in analog fashion.

C. Partial Feedback Linearization of the Motor Dynamics

Substituting ψ(θ− q) and θ from (23) and (32) into (11)yields the transformed motor dynamics

Bq +B

(κ−1(θ − q)

(− κ(θ − q)(θ − q)

+ κ(η − q)(η − ˙q) + κ(η − q)(η − ¨q)− d2

dt2(D ˙q)

+ g(q) + n2

))+ψ(η − q)−D ˙q + g(q) + n0 = τm.

(33)

To achieve closed-loop motor dynamics that structurallyresemble the desired motor dynamics (21) we perform acontroller design in two steps.

First, we pre-compensate some undesired nonlinear termsvia τm = u+ u, where

u =Bq +B

(κ−1(θ − q)

(− κ(θ − q)(θ − q)

+ κ(η − q)(η − ˙q)− κ(η − q)¨q − d2

dt2(D ˙q)

+ g(q) + n2

))−D ˙q + g(q) + n0,

(34)

resulting in the following motor dynamics

Bκ−1(θ − q)κ(η − q)η +ψ(η − q) = u. (35)

with the intermediate control input u. Note, that in case ofa state-dependent damping matrix an implicit equation for uresults which always can be solved.

The difference to the desired motor dynamics (21) isthe state-dependent inertia matrix Bκ−1(θ − q)κ(η − q),compared to the constant inertia B. To achieve constantinertia B in (39), the intermediary control input u can bechosen as

u =(I −Bκ−1(θ − q)κ(η − q)B−1

)ψ(η − q)

+Bκ−1(θ − q)κ(η − q)B−1u(36)

which yields the desired motor dynamics in (21).The final controller input τm is composed of eqs. (22),

(34) and (36)

τm = u+ u(u), (37)

where u is a function of u.

Page 5: A Passivity-Based Approach for Trajectory Tracking and ...

Passive Environment

ψ(η − q)

τext

˙q

Closed-Loop MotorDynamics (39)

Closed-LoopLink Dynamics (38)

Fig. 3. Representation of the closed-loop dynamics (38)–(39) as aninterconnection of passive subsystems and its respective power ports.

V. STABILITY & PASSIVITY ANALYSIS

This section analyses the passivity properties and derivesa stability statement of the closed-loop dynamics

M(q, t)¨q +(C(q, ˙q, t) +D(q,η, t)

)˙q =ψ(η − q, t) + τext,

(38)Bη +ψ(η − q, t) =−KPη −KDη

(39)

resulting for the plant dynamics (10)-(11) together with thecontroller (37), written in tracking error variables, the time-variant behavior becomes immediately apparent.

A. Passivity

On the basis of definition [24] it will be shown that thelink dynamics (38) and the closed-loop motor dynamics (39)can be presented as an interconnection of passive subsystems[24] as depicted in Fig. 3. For the link side dynamics weconsider the time varying storage function

Sq =1

2˙qTM(q, t) ˙q. (40)

For the analysis of the controller proposed in this work, weexpress the derivative of (40) in the form,

Sq = − ˙qTD ˙q + ˙qTψ(η − q) + ˙qTτext, (41)

which corresponds to the representation of the closed-looplink dynamics given in (38)6. On the right hand side of(41) one can identify two terms, each corresponding to aninterconnection port with the link dynamics. The second termrepresents the interconnection with the controlled motor dy-namics (39) and the third term represents the interconnectionwith an external system. For the former we consider thestorage function of the form

Sη =1

2ηTBη + Uτ (η − q) +

1

2ηTKPη. (42)

The first term represents a virtual kinetic energy of the motor-side, while the second term represents the total potentialenergy stored in virtual springs and the third term representsa virtual control energy. Its derivative along the solution ofthe closed-loop motor dynamics (39) can be expressed as

Sη = −ηTKDη − ˙qTψ(η − q). (43)

6Note that (10) and (38) are equivalent.

Comparing (41) and (43), the interconnection between thelink and motor dynamics becomes directly apparent. Theabove analysis motivates the following proposition.

Proposition 1: The system (38)–(39) can be representedas a passive map from the generalized external forces τext tothe generalized velocity of the link-side tracking error ˙q.

Proof: Consider the storage function S = Sq + Sηcomprising (40) and (42) of which the derivative along thesolution of the system (38)–(39) satisfies

S = − ˙qTD ˙q − ηTKDη + ˙qTτext ≤ ˙qτext, (44)

which completes the proof according to [24].Note, for the special case of regulation control, i.e. qd =const., the closed-loop system can be seen as a passivemapping from the environment with respect to the, physicallymore intuitive, power port qT, τext.

B. Stability, Steady-State Behavior

First, we determine the steady state conditions for theclosed loop dynamics. Second, we deduce stability by con-sidering the storage function of proof of Proposition 1 asLyapunov function candidate.

By setting the state derivatives ( ˙q, η) to zero, the steadystate conditions for (38)-(39), in absence of external loads,become

˙q0 =0, (45)η0 =0, (46)

ψ(η0 − q0) =0, (47)ψ(η0 − q0)−KPη0 =0. (48)

It can be easily verified that (45)–(46) have a unique solution

q0 =0 (49)η0 =0 (50)

under assumptions in III and if KP is positive definite.Proposition 2: Under the assumptions made in III, the

closed-loop dynamics (38)–(39) is uniformly stable.Proof: We show stability by invoking Lyapunov’s

direct method for non-autonomous systems [25]. For that wehave to find a scalar, positive definite function V (x, t) withcontinuous partial derivatives such that V is negative semi-definite. Furthermore, for uniform stability, V (x, t) has to beupper bounded by a time-invariant positive definite functionV ∗(x) such that V (x, t) ≤ V ∗(x) ∀t [25]. Consider

V (x, t) = Sq(q, ˙q, t) + Sη(η, η, q, t) (51)

as an energy based, time-variant Lyapunov function, wherex = (qT,ηT, ˙qT, ηT)T. Obviously, V (x, t) is positive definitesince M , B and KP are positive definite and the springpotential is, by assumption, positive semi-definite. Let uschoose V ∗ as follows

V ∗(x) = M || ˙q||2 + ηTBη + Uτ (η − q) + ηTKPη, (52)

where M ≡ maxλ(M(q, t)

). Note, M exists and is

bounded as the inertia matrix M(q) is bounded [26]. The

Page 6: A Passivity-Based Approach for Trajectory Tracking and ...

time derivative of V along the solutions of the closed loopsystem (38)–(39) is given by7

V (q, q,η, η, t) = − ˙qTD ˙q − ηTKDη. (53)

Since D and KD are positive definite, V is negative semi-definite. This completes the proof.

By invoking Barbalat’s lemma [25]

Lemma 1: If a scalar function V (x, t) satisfies the fol-lowing conditions

• V (x, t) is lower bounded• V (x, t) is negative semi-definite• V (x, t) is uniformly continuous in time

then V (x, t)→ 0 as t→∞.

we can show asymptotic convergence of the generalizedvelocities of the tracking error, i.e. q(t) → qd(t) andη(t)→ 0, for the closed-loop dynamics (38)–(39).

Proof: To complete the proof only condition 3 is leftto show. The second time derivative of V is given by

V (x, t) = 2(

˙qTDM−1C ˙q + ˙qTDM−1D ˙q

− ˙qTDM−1ψ(η − q))− ˙qTDq

+ 2(ηTKDB

−1ψ(η − q)

+ ηTKDB−1KPη + ηTKDB

−1KDη).

(54)

From the stability prove above, we already know that q,η, ˙qand η are bounded. M , M−1, C, B are obviously boundedunder the assumptions made in III. D and D are assumed tobe bounded (see III). The sequence of bounded matrices isagain bounded. Thereby we can deduce that V is bounded,which again, is a sufficient condition for V to be uniformlycontinuous [25]. Barbalat’s lemma then states that V → 0,which, due the positive definite nature of M and KD, canonly be the case if η, ˙q → 0.

VI. EXPERIMENTAL VALIDATION AND PERFORMANCEANALYSIS

The performance of the control approach has been ex-perimentally evaluated on the first four Variable StiffnessActuators (VSA) joints of the DLR Hand Arm System [1],see also fig. 1. The first four arm joints, namely the elbowand the three shoulder joints, are implemented by FloatingSpring Joints (FSJ) [27].

First, we present one possible layout of the dampingmatrix D based on modal decomposition. Note, the controlapproach presented in this paper doesn’t restrict the user toa specific damping matrix layout. The only restrictions onD are the assumptions made in IV-A. In the latter part wepresent the experimental results.

7The symmetry of KP is exploited while transforming the equations.

A. Damping Design

The entries of the inertia matrix of robotic systems withrotational joints vary strongly with the kinematic configura-tion of the robot. Additionally, the (local) stiffness of manyvariable stiffness mechanisms varies with varying load. Toachieve similar performance in the complete workspace ofthe robot, an approach which accounts for these variationsshould be considered. A damping design which accounts forthe variation of the inertia and stiffness matrix can be derivedfrom the generalized eigenvalue problem of the linearizedsystem

M(q)q +D(q,η)q − κ(η − q) (η − q) = 0 (55)

such that

D(q,η) = 2Q(q,η)−Tdiag(ξi√λi

)Q(q,η)−1 . (56)

Herein, Q ∈ Rn×n is a non-singular matrix satisfyingQTQ = M and κ = QTΛQ, where Λ is a diagonalmatrix containing the real and positive eigenvalues λi, sinceM is a symmetric and positive definite matrix and κ isa symmetric matrix ([28]). Further, ξi > 0 are normalizeddamping factors sometimes referred to modal damping. Notethat the instantaneous linearization is only performed for thedesign of the damping matrix.

Remark 2: This damping design implies the matrixD(q,η) to be a function of the link variables q and motorvariables η. Substitution of d2

dt2

(D(q,η) ˙q

)in (34) would

result, as previously mentioned, in an algebraic loop whichcan always be solved. In addition, the first and second timederivatives of eigenvalues and eigenvectors of (56) have tobe calculated. The interested read can find more details aboutthis issue in [29] and [30].

Remark 3: Simulations have shown, that for our case,neglecting terms, containing D and D, does not significantlyimpact the controller performance. Therefore these terms aredropped for the experiments.

B. Results

Three different experiments have been performed to ana-lyze the controller performance. Throughout all experimentsthe damping design presented in VI-A has been used tolayout D, with ξqi = 0.5, i = 1, . . . , 4. KP was manuallytuned such the that the control input τm doesn’t overshootits limits and was set to KP = diag([ 1000 900 750 750 ])Nm. KD is set such that the (decoupled) motor dynamics arecritically damped (KD)ij=1...4 = 2δijξxi

√KPijBij , ξxi =

0.7 ∀i. The stiffness adjustment was set to σi = 5 deg ∀i;corresponding to the mid curve of the stiffness plot in fig. 1.

First, we show the tracking performance in absence ofexternal disturbances, see fig. 4. In this experiment the max-imum angular velocities are qT = [ 190 30 103 113 ] deg/s.Note, between each of three motions, at time points t1 and t2,qd is adjusted to the current values of q. Model and sensoruncertainties are sources for the non-zero steady state error.

In the second experiment, we show the damping perfor-mance. Figure 5.a shows the response of a classical motorPD controller. Figure 5.b shows of our proposed controller.

Page 7: A Passivity-Based Approach for Trajectory Tracking and ...

link

pos.

[deg

]

-50

0

50

100

0 1 2 3 4

trac

k.er

ror

[deg

]

-10

-5

0

5

10

q1

q2

q3

q4

q1

q2

q3

q4

t1 t2

time [s]

ref. values

Fig. 4. Experimental result: tracking performance in absence of externaldisturbances. At time points t1 and t2, qd is adjusted to the currentvalues of q. The maximum angular velocities reached reached qT =[ 190 30 103 113 ] deg/s.

In a stationary configuration of q = [ 110 61 16 − 35 ]T degan external torque was applied to system through physicalhuman interaction. The magnitude of disturbance can be bestseen on peaks in the respective joint torques maxt(τj) =[ 68 63 17 − 28 ]T Nm. The deflections in joint one and twoare close to the maximum the system can handle.

It can be seen that the convergence rate of the linkvariables in case of our controller is about 8 times highercompared to the classical PD controller.

Figure 6 show the disturbance rejection performance whilemoving on a trajectory. Again the system is disturbed throughphysical human interaction. Within 0.3 s the system con-verges into an error band of ±1 deg.

A Video showing all three experiments can be found asattachment.

VII. CONLUSION

This paper proposes a control law to implement an ex-act link-side trajectory tracking and disturbance rejectionbehavior while changing the original plant dynamics onlyto a minimum extend. This is achieved by introducingnew coordinates for the motor reflecting these dampingand feedforward terms. The motor dynamics in the newcoordinates structurally equal the original motor dynamicssuch that PD regulation control achieves the desired link-side behavior. This in turn allows for a comprehensivestability and passivity analysis based on physically motivatedLyapunov functions. The performance of the approach hasbeen experimentally evaluated on a multi-joint VSA robotarm. As for any link-side tracking controller, a theoreticallimitation of the approach seemed to be the usage of thesecond and third time derivative of the measured link positionin the control law, which however turned out not to bea practical limitation, since the time derivatives could be

computed based on an accurate model of the plant. To ourbest knowledge, this is the first link-side tracking controllerfor compliantly actuated robots with nonlinear elastic trans-missions, which is theoretically founded and reported tobe experimentally validated on a multi degrees of freedomrobot.

VIII. APPENDIX

REFERENCES

[1] A. Albu-Schaffer, O. Eiberger, M. Fuchs, M. Grebenstein, S. Had-dadin, C. Ott, A. Stemmer, T. Wimbck, S. Wolf, C. Borst, andG. Hirzinger, “Anthropomorphic soft robotics from torque con-trol to variable intrinsic compliance,” in Robotics Research, ser.Springer Tracts in Advanced Robotics, C. Pradalier, R. Siegwart, andG. Hirzinger, Eds. Springer, 2011, vol. 70, pp. 185–207.

[2] D. Lakatos and A. Albu-Schaffer, “Switching based limit cycle controlfor compliantly actuated second-order systems,” in Proceedings of the19th IFAC World Congress, 2014, pp. 6392–6399.

[3] P. Tomei, “A simple pd controller for robots with elastic joints,”Automatic Control, IEEE Transactions on, vol. 36, no. 10, pp. 1208–1213, Oct 1991.

[4] R. Ortega, R. Kelly, and A. Loria, “A class of output feedbackglobally stabilizing controllers for flexible joints robots,” Robotics andAutomation, IEEE Transactions on, vol. 11, no. 5, pp. 766–770, Oct1995.

[5] L. Zollo, A. De Luca, and B. Siciliano, “Regulation with on-linegravity compensation for robots with elastic joints,” in Robotics andAutomation, 2004. Proceedings. ICRA ’04. 2004 IEEE InternationalConference on, vol. 3, April 2004, pp. 2687–2692.

[6] C. Ott, A. Albu-Schaffer, A. Kugi, S. Stamigioli, and G. Hirzinger,“A passivity based cartesian impedance controller for flexible jointrobots-part i: Torque feedback and gravity compensation,” in Proc.IEEE Int. Conf. on Robotics and Automation, 2004.

[7] B. Vanderborght, A. Albu-Schaffer, A. Bicchi, E. Burdet, D. G. Cald-well, R. Carloni, M. Catalano, O. Eiberger, W. Friedl, G. Ganesh et al.,“Variable impedance actuators: A review,” Robotics and AutonomousSystems, vol. 61, no. 12, pp. 1601–1614, 2013.

[8] A. Albu-Schaffer, C. Ott, and F. Petit, “Constructive energy shapingcontrol for a class of euler-lagrange systems,” in 10th Int. IFACSymposium on Robot Control, 2012.

[9] A. Albu-Schaffer and G. Hirzinger, “A globally stable state feedbackcontroller for flexible joint robots,” Advanced Robotics, vol. 15, no. 8,pp. 799–814, 2001.

[10] F. Petit and A. Albu-Schaffer, “State feedback damping control for amulti dof variable stiffness robot arm,” in Proc. IEEE Int. Conf. onRobotic and Automation, 2011.

[11] I. Sardellitti, G. Medrano-Cerda, N. Tsagarakis, A. Jafari, and D. Cald-well, “Gain scheduling control for a class of variable stiffness actuatorsbased on lever mechanisms,” Robotics, IEEE Transactions on, vol. 29,no. 3, pp. 791–798, June 2013.

[12] M. W. Spong, “Modeling and control of elastic joint robots,” Trans-actions of the ASME: Journal of Dynamic Systems, Measurement, andControl, vol. 109, pp. 310–319, 1987.

[13] S. Nicosia and P. Tomei, “Design of global tracking controllers forflexible-joint robots,” Journal of robotic systems, vol. 10, no. 6, pp.835–846, 1993.

[14] A. Loria and R. Ortega, “On tracking control of rigid and flexiblejoints robots,” Appl. Math. Comput. Sci, vol. 5, no. 2, pp. 101–113,1995.

[15] C. Ott, A. Albu-Schaffer, A. Kugi, and G. Hirzinger, “Decouplingbased cartesian impedance control of flexible joint robots,” in Proc.IEEE Int. Conf. on Robotics and Automation, 2003.

[16] J. H. Oh and J. S. Lee, “Control of flexible joint robot systemby backstepping design approach,” Intelligent Automation & SoftComputing, vol. 5, no. 4, pp. 267–278, 1999.

[17] C. Ott, Cartesian Impedance Control of Redundant and Flexible-JointRobots, B. Siciliano and O. Khatib, Eds. Springer, 2008.

[18] J.-J. Slotine and L. Weiping, “Adaptive manipulator control: A casestudy,” Automatic Control, IEEE Transactions on, vol. 33, no. 11, pp.995–1003, 1988.

[19] M. W. Spong, “Adaptive control of flexible joint manipulators,”Systems & Control Letters, vol. 13, no. 1, pp. 15–21, 1989.

[20] A. De Luca and P. Lucibello, “A general algorithm for dynamicfeedback linearization of robots with elastic joints,” in IEEE Int. Conf.on Robotics and Automation, vol. 1, 1998, pp. 504–510.

Page 8: A Passivity-Based Approach for Trajectory Tracking and ...

trac

king

erro

r[d

eg]

-10

-5

0

5

-10

-5

0

5

join

tto

rque

[Nm

]

-20

0

20

40

60

-20

0

20

40

60

time [s]

0 1 2 3 4

mot

orto

rque

[Nm

]

-40

-20

0

20

40

time [s]

0 1 2 3 4

-40

-20

0

20

40

q1

q2

q3

q4

τj1

τj2

τj3

τj4

τm1

τm2

τm3

τm4

a) b)

physical human interaction physical human interaction

Fig. 5. Experimental result: damping performance after external disturbance through physical human interaction.

link

pos.

[deg

]

-50

0

50

100

trac

king

erro

r[d

eg]

-10

0

10

join

tto

rque

[Nm

]

-40-20

02040

mot

orto

rque

[Nm

]

-40-20

02040

time [s]0 1 2 3 4

q1q2q3q4

τj1τj2τj3τj4

u1

u2

u3

u4

q1q2q3q4

physical human interaction

error band: ±1 deg

Fig. 6. Experimental results: the system is externally disturbed by physicalhuman interaction while tracking a reference trajectory.

[21] G. Palli, C. Melchiorri, and A. De Luca, “On the feedback linearizationof robots with variable joint stiffness,” in IEEE Int. Conf. on Roboticsand Automation, 2008, pp. 1753–1759.

[22] B. Brogliato, R. Ortega, and R. Lozano, “Global tracking controllersfor flexible-joint manipulators: A comparative study,” Automatica,vol. 31, no. 7, pp. 941–956, Jul. 1995. [Online]. Available:http://dx.doi.org/10.1016/0005-1098(94)00172-F

[23] S. Ozgoli and H. Taghirad, “A survey on the control of flexible jointrobots,” Asian Journal of Control, vol. 8, no. 4, pp. 332–344, 2006.

[24] A. van der Schaft, L2-gain and passivity techniques in nonlinearcontrol. London New York: Springer, 2000.

[25] J.-J. Slotine and W. Li, Applied Nonlinear Control. Prentice Hall,1991.

[26] F. Ghorbel, B. Srinivasan, and M. W. Spong, “On the uniform bound-edness of the inertia matrix of serial robot manipulators,” Journal ofRobotic Systems, vol. 15, no. 1, 1998.

[27] S. Wolf, O. Eiberger, and G. Hirzinger, “The dlr fsj: Energy baseddesign of a variable stiffness joint,” in Robotics and Automation(ICRA), 2011 IEEE International Conference on, May 2011, pp. 5082–5089.

[28] R. A. Horn and C. R. Johnson, Matrix Analysis. Cambridge UniversityPress, 2012.

[29] R. B. Nelson, “Simplified calculation of eigenvector derivatives,” AIAAJournal, vol. 14, no. 9, pp. 1201–1205, sep 1976.

[30] M. I. Friswell, “Calculation of second and higher order eigenvectorderivatives,” Journal of Guidance, Control, and Dynamics, vol. 18,no. 4, pp. 919–921, jul 1995.