16
IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 59, NO. 5, MAY 2014 1115 Human-Inspired Control of Bipedal Walking Robots Aaron D. Ames Abstract—This paper presents a human-inspired control ap- proach to bipedal robotic walking: utilizing human data and output functions that appear to be intrinsic to human walking in order to formally design controllers that provably result in stable robotic walking. Beginning with human walking data, out- puts—or functions of the kinematics—are determined that result in a low-dimensional representation of human locomotion. These same outputs can be considered on a robot, and human-inspired control is used to drive the outputs of the robot to the outputs of the human. The main results of this paper are that, in the case of both under and full actuation, the parameters of this controller can be determined through a human-inspired optimiza- tion problem that provides the best t of the human data while simultaneously provably guaranteeing stable robotic walking for which the initial condition can be computed in closed form. These formal results are demonstrated in simulation by considering two bipedal robots—an underactuated 2-D bipedal robot, AMBER, and fully actuated 3-D bipedal robot, NAO—for which stable robotic walking is automatically obtained using only human data. Moreover, in both cases, these simulated walking gaits are realized experimentally to obtain human-inspired bipedal walking on the actual robots. Index Terms—Bipedal locomotion, hybrid systems, nonlinear control, robotics. I. INTRODUCTION D ESPITE the simplicity with which humans appear to walk, human locomotion is inherently complex utilizing hundreds of degrees of freedom coupled with highly nonlinear dynamics and forcing due to the 57 muscles employed during human walking [41]. There is evidence to suggest that humans control walking behaviors through a hierarchical subdivi- sion between cortical control and central pattern generators in the spinal column [11], [16], [19], [28], indicating that when humans perform motion primitives, i.e., steady-state well-practiced walking behaviors—potentially simple and characterizable control strategies are implemented. Inspired by this hierarchical control present in humans, the motivation for the ideas presented in this paper is that the essential information needed to understand walking is encoded by a simple class of functions canonical to human walking. In other words, taking the control theorist approach to understanding a complex and unknown system, we view the human walking system as a “black box,” where the “inputs” to the system are specic Manuscript received December 30, 2011; revised November 02, 2012 and October 31, 2013; accepted January 02, 2014. Date of publication January 10, 2014; date of current version April 18, 2014. This work was supported in part by the National Science Foundation under Grants CNS-0953823 and CNS-1136104 and in part by the Norman Hackerman. Advanced Research Programunder Award 00512-0184-2009. Recommended by Associate Editor E. Frazzoli. The author is with the Department of Mechanical Engineering, with a joint appointment in Electrical and Computer Engineering, Texas A&M University, College Station TX, 77840 USA (e-mail: [email protected]). Color versions of one or more of the gures in this paper are available online at http://ieeexplore.ieee.org. Digital Object Identier 10.1109/TAC.2014.2299342 walking behaviors, and we seek “outputs” of this system that characterize these behaviors. In this fashion, a low-dimensional characterization of the system is achieved. These outputs can then be utilized in the formal design of robotic controllers—the outputs of the robot can be driven to the outputs of the human, resulting in “human-like” robotic walking. Given human walking as the motivation for achieving robotic walking, this paper begins by looking at human walking data achieved through motion capture of subjects walking on at ground at a “natural” pace. By studying this data, we discover a collection of human outputs, e.g., the forward position of the hip, that appear to characterize human walking—they are mu- tually exclusive, thus providing a low-dimensional represen- tation of the system’s behavior. Moreover, we nd that these human outputs, as computed from the data, appear to be de- scribed by two very simple functions: i) a linear function of time (describing the forward position of the hip) and ii) the time so- lution to a linear spring-mass-damper system (describing the re- maining the outputs). We term ii) canonical walking functions, due to their simple yet general form, and verify that in fact these functions describe the human data by showing that they can be t to the output data with a remarkably high correlation coef- cient of essentially 1. This result both motivates the introduc- tion of the general notion of a human output combination, and provides insight into the basic mechanisms underlying human walking since we conclude that, at the most basic level, the pri- mary outputs associated with locomotion are characterized by a system of linear springs and dampers parameterized in time by the forward walking velocity. Utilizing human outputs and their time-based representa- tion given by the canonical walking functions, we construct human-inspired controllers that drive the outputs of the robot to the outputs of the human as represented by the canonical walking functions. As a specic example, this paper considers a 5 degree of freedom 2-D bipedal robot with knees and a torso, for which we dene human-inspired outputs: the difference between the outputs of the robot (computed via kinematics) and the human outputs (encoded through canonical walking functions parameterized by the forward position of the hip). In the context of both under and full actuation, human-in- spired outputs are used to construct human-inspired controllers through the use of input–output linearization that drive the human-inspired outputs to zero exponentially fast. Thus, on the corresponding zero dynamics surface, the outputs of the robot and human are in agreement. Since a bipedal robot is a hybrid system (due to the impacts that occur at foot strike), the key challenge is to determine parameters of the human-inspired controller that result in a hybrid zero dynamics surface yielding stable robotic walking for which the canonical walking func- tions best t the human output data. The main contribution of this paper is a formal method for determining the parameters of the human-inspired con- troller that provably result in stable robotic walking that is as “human-like” as possible. Beginning with the case of full 0018-9286 © 2014 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission. See http://www.ieee.org/publications_standards/publications/rights/index.html for more information.

IEEE TRANSACTIONS ON AUTOMATIC CONTROL, …edge.rit.edu/content/P15201/public/MSD II/Research Papers...IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 59, NO. 5, MAY 2014 1115 Human-Inspired

Embed Size (px)

Citation preview

Page 1: IEEE TRANSACTIONS ON AUTOMATIC CONTROL, …edge.rit.edu/content/P15201/public/MSD II/Research Papers...IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 59, NO. 5, MAY 2014 1115 Human-Inspired

IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 59, NO. 5, MAY 2014 1115

Human-Inspired Control of Bipedal Walking RobotsAaron D. Ames

Abstract—This paper presents a human-inspired control ap-proach to bipedal robotic walking: utilizing human data andoutput functions that appear to be intrinsic to human walkingin order to formally design controllers that provably result instable robotic walking. Beginning with human walking data, out-puts—or functions of the kinematics—are determined that resultin a low-dimensional representation of human locomotion. Thesesame outputs can be considered on a robot, and human-inspiredcontrol is used to drive the outputs of the robot to the outputsof the human. The main results of this paper are that, in thecase of both under and full actuation, the parameters of thiscontroller can be determined through a human-inspired optimiza-tion problem that provides the best fit of the human data whilesimultaneously provably guaranteeing stable robotic walking forwhich the initial condition can be computed in closed form. Theseformal results are demonstrated in simulation by considering twobipedal robots—an underactuated 2-D bipedal robot, AMBER,and fully actuated 3-D bipedal robot, NAO—for which stablerobotic walking is automatically obtained using only human data.Moreover, in both cases, these simulated walking gaits are realizedexperimentally to obtain human-inspired bipedal walking on theactual robots.

Index Terms—Bipedal locomotion, hybrid systems, nonlinearcontrol, robotics.

I. INTRODUCTION

D ESPITE the simplicity with which humans appear towalk, human locomotion is inherently complex utilizing

hundreds of degrees of freedom coupled with highly nonlineardynamics and forcing due to the 57 muscles employed duringhuman walking [41]. There is evidence to suggest that humanscontrol walking behaviors through a hierarchical subdivi-sion between cortical control and central pattern generatorsin the spinal column [11], [16], [19], [28], indicating thatwhen humans perform motion primitives, i.e., steady-statewell-practiced walking behaviors—potentially simple andcharacterizable control strategies are implemented. Inspired bythis hierarchical control present in humans, the motivation forthe ideas presented in this paper is that the essential informationneeded to understand walking is encoded by a simple class offunctions canonical to human walking. In other words, takingthe control theorist approach to understanding a complex andunknown system, we view the human walking system as a“black box,” where the “inputs” to the system are specific

Manuscript received December 30, 2011; revised November 02, 2012 andOctober 31, 2013; accepted January 02, 2014. Date of publication January10, 2014; date of current version April 18, 2014. This work was supportedin part by the National Science Foundation under Grants CNS-0953823 andCNS-1136104 and in part by the Norman Hackerman. Advanced ResearchProgramunder Award 00512-0184-2009. Recommended by Associate EditorE. Frazzoli.The author is with the Department of Mechanical Engineering, with a joint

appointment in Electrical and Computer Engineering, Texas A&M University,College Station TX, 77840 USA (e-mail: [email protected]).Color versions of one or more of the figures in this paper are available online

at http://ieeexplore.ieee.org.Digital Object Identifier 10.1109/TAC.2014.2299342

walking behaviors, and we seek “outputs” of this system thatcharacterize these behaviors. In this fashion, a low-dimensionalcharacterization of the system is achieved. These outputs canthen be utilized in the formal design of robotic controllers—theoutputs of the robot can be driven to the outputs of the human,resulting in “human-like” robotic walking.Given human walking as the motivation for achieving robotic

walking, this paper begins by looking at human walking dataachieved through motion capture of subjects walking on flatground at a “natural” pace. By studying this data, we discovera collection of human outputs, e.g., the forward position of thehip, that appear to characterize human walking—they are mu-tually exclusive, thus providing a low-dimensional represen-tation of the system’s behavior. Moreover, we find that thesehuman outputs, as computed from the data, appear to be de-scribed by two very simple functions: i) a linear function of time(describing the forward position of the hip) and ii) the time so-lution to a linear spring-mass-damper system (describing the re-maining the outputs). We term ii) canonical walking functions,due to their simple yet general form, and verify that in fact thesefunctions describe the human data by showing that they can befit to the output data with a remarkably high correlation coef-ficient of essentially 1. This result both motivates the introduc-tion of the general notion of a human output combination, andprovides insight into the basic mechanisms underlying humanwalking since we conclude that, at the most basic level, the pri-mary outputs associated with locomotion are characterized by asystem of linear springs and dampers parameterized in time bythe forward walking velocity.Utilizing human outputs and their time-based representa-

tion given by the canonical walking functions, we constructhuman-inspired controllers that drive the outputs of the robotto the outputs of the human as represented by the canonicalwalking functions. As a specific example, this paper considersa 5 degree of freedom 2-D bipedal robot with knees and a torso,for which we define human-inspired outputs: the differencebetween the outputs of the robot (computed via kinematics)and the human outputs (encoded through canonical walkingfunctions parameterized by the forward position of the hip).In the context of both under and full actuation, human-in-spired outputs are used to construct human-inspired controllersthrough the use of input–output linearization that drive thehuman-inspired outputs to zero exponentially fast. Thus, on thecorresponding zero dynamics surface, the outputs of the robotand human are in agreement. Since a bipedal robot is a hybridsystem (due to the impacts that occur at foot strike), the keychallenge is to determine parameters of the human-inspiredcontroller that result in a hybrid zero dynamics surface yieldingstable robotic walking for which the canonical walking func-tions best fit the human output data.The main contribution of this paper is a formal method

for determining the parameters of the human-inspired con-troller that provably result in stable robotic walking that isas “human-like” as possible. Beginning with the case of full

0018-9286 © 2014 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission.See http://www.ieee.org/publications_standards/publications/rights/index.html for more information.

Page 2: IEEE TRANSACTIONS ON AUTOMATIC CONTROL, …edge.rit.edu/content/P15201/public/MSD II/Research Papers...IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 59, NO. 5, MAY 2014 1115 Human-Inspired

1116 IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 59, NO. 5, MAY 2014

Fig. 1. AMBER, an underactuated 2-D bipedal robot, and NAO, a fully actu-ated 3-D bipedal robot.

actuation, we introduce a human-inspired optimization problemwhere the cost is the least squares fit of the outputs of the robotto the human output data subject to constraints that ensurepartial hybrid zero dynamics, i.e., constraints that ensure thatthe zero dynamics surface associated with the relative degree 2output functions are invariant through impact. This invarianceallows us to characterize the behavior of the hybrid systemmodeling a bipedal robot (which is ten-dimensional for the ex-amples considered) through a two-dimensional hybrid system.Utilizing this reduced dimensional representation, we are ableto prove the main result of this paper: the parameters thatsolve the partial human-inspired optimization problem implythe existence of an exponentially stable hybrid periodic orbit,i.e., the existence of stable walking gait. Moreover, the fixedpoint of the Poincaré map associated with this periodic orbitcan be explicitly computed from these parameters. In addition,these results are extended to the case of underactuation, whereadditional constraints to the human-inspired optimization areadded that ensure stable robotic walking, again allowing forthe fixed point to be computed in closed form. Therefore,in the case of both under and full actuation, using only thehuman data we are able to automatically generate parametersfor the human-inspired controller that imply the existence of astable walking gait, and we can explicitly compute the initialcondition to this walking gait from these parameters in the caseof both full and under actuation.To demonstrate the validity of the formal methods employed

in this paper, we apply them to two bipedal robots (shownin Fig. 1): AMBER, an underactuated 2-D bipedal robot, andNAO, a fully actuated 3-D bipedal robot for which we considera simplified 2-D sagittal representation. In both cases, we applythe constructions of this paper to automatically obtain roboticwalking in simulation utilizing only human data; the end resultis human-like robotic walking. This is quantified by comparingthe outputs of AMBER to the human output data, and showingthat it lies within one standard deviation of the norm, thusimplying “normal” human walking [37]. To further validatethe theory developed in this paper, we discuss the experimentalimplementation of the results presented on both AMBER andNAO. In both cases, we experimentally obtain human-inspiredrobotic walking on the physical robots (videos can be foundat [2] and [3]). This paper, therefore, formally and provablyobtains robotic walking from human data, and realizes theseformal results experimentally on two real-world bipedal robots.

This paper has drawn inspiration from work on roboticwalking that attempts to discover the fundamental princi-ples underlying bipedal walking, including but not limitedto: passive walking and controlled symmetries [15], [32],[47], geometric reduction [9], [21], [44] and hybrid nonlinearfeedback control [20], [54], [55]. It is also important to notethat this is not the first paper that attempts to bridge the gapbetween human and robotic walking [30], [43], [48], [49],although there have been relatively few studies in this directionwhen compared to the vast literature on robotic walking andbiomechanics (see [12]–[14], [54] and [23], [35], [52], [56],to only name a few). Of special note is the preexisting workthat studies human data in the context of both robotic walkingand optimization [48], [49], [57]; a common thread in thatwork is that the human data considered consists of simplythe joint angles over time, and this data is fit with high orderpolynomials. The results in this paper, therefore, diverge fromthis preexiting work in two important and fundamental ways:a) human output data is considered, rather than just angles,yielding a low-dimensional representation of human walking,b) this output data is described with canonical walking func-tions, giving insight into the behavior of these outputs overtime that is of the simplest, yet most general, form possible,and c) a framework is presented that is naturally applicable toboth under and fully actuated robotic systems.It is important to note that the results presented in this paper

build upon [6]–[8]. These papers considered human-inspiredcontrol of 4 and 5 link (under and fully) actuated robots in thecontext of specific output combinations. This paper generalizesand extends these constructions and results in the following sig-nificant ways: i) considers the hybrid system model of a generalrobot for which the previous models and the specific robots con-sidered in this paper are a special case, ii) introduces the generalnotion of a human output combination allowing for the freedomto choose different human outputs as long as they satisfy certainconditions, iii) extends the formal results from [6]–[8] to thisgeneral setting, and iv) introduces novel results related to thecharacterization of solutions and limit cycles in the context offull actuation—results that are essential to experimental realiza-tion. The main motivation for these generalizations is the recentresults that build upon the human-inspired control framework.These include: walking up stairs, down stairs and running insimulation through the human-inspired control framework [40],[60], walking up slopes and on uneven terrain experimentallyon AMBER 1 [36], walking on flat ground (both in simulationand experimentally) for a fully actuated 2-D robot with feet,AMBER 2 [4], [31], 3-D speed regulated robotic walking withNAO both in simulation and experimentally [39], and multi-domain locomotion on ATRIAS [24] and AMBER 2 [59], in-cluding phases of under, full and over actuation. At the core ofall of these results are the formal results presented in this paper,thus indicating the extensibility of the results presented.

II. BIPEDAL ROBOT MODELS

Bipedal walking robots naturally display continuous and dis-crete behavior throughout the course of a step—the continuousbehavior occurs when the leg swings forward and the discretebehavior occurs when the foot strikes the ground. It is, there-fore, natural to model robots of this form by hybrid systems.This section introduces the basic formalisms of hybrid systemsalong with the specific hybrid models modeling bipedal robots.A special case of these systems yields the models for the two

Page 3: IEEE TRANSACTIONS ON AUTOMATIC CONTROL, …edge.rit.edu/content/P15201/public/MSD II/Research Papers...IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 59, NO. 5, MAY 2014 1115 Human-Inspired

AMES: HUMAN-INSPIRED CONTROL OF BIPEDAL WALKING ROBOTS 1117

robots considered in this paper: AMBER and NAO (see Fig. 1for pictures of both of these robots). Specifically, we considertwo hybrid models, one for full actuation and one for under-actuation.

A. Hybrid Systems

We begin by introducing hybrid (control) systems (also re-ferred to as systems with impulsive effects or systems with im-pulse effects [20], [21]). We consider hybrid systems with onedomain because the specific biped models considered in thispaper do not have feet; the same model applies to robots withflat foot walking, while for more complex foot behavior moreelaborate hybrid systems must be considered [21], [44], [50].A (simple) hybrid control system is a tuple,

where is the domain with a smooth submanifold ofthe state space , is the set of admissible controls,

is a proper subset of called the guard or switchingsurface, is a smooth map called the reset map, and

is a control system on , i.e., in coordinates:. A hybrid system is a hybrid control system with

, e.g., any applicable feedback controllers have been applied,making the system closed-loop. In this case,

where is a dynamical system on , i.e., .

B. Periodic Orbits

Stable bipedal robotic walking corresponds to stable periodicorbits in hybrid systems. For simplicity, we focus our atten-tion on introducing periodic orbits of hybrid systems with fixedpoints on the guard (for more general definitions, see [21], [53]).Let be the solution to with initial condition

. For , we say that is periodic with periodif . A set is a periodic orbit with

fixed point if for a periodicsolution . Associated with a periodic orbit is a Poincaré map[53]. In particular, taking to be the Poincaré section, one ob-tains the Poincaré map which is a partial function:

where is the time-to-impact function [54]:

(1)

for . As with smooth dynamical systems, the stability ofthe Poincaré map determines the stability of the periodic orbit. In particular, the Poincaré map is (locally) exponentially

stable (as a discrete time system ) at the fixedpoint if and only if the periodic orbit is (locally) exponen-tially stable [33]. Although it is typically not possible to analyt-ically compute the Poincaré map, it is possible to numericallycompute its Jacobian. If the eigenvalues of the Jacobian havemagnitude less than 1, the stability of the periodic orbit hasbeen numerically verified.

C. Robotic Hybrid Models

Utilizing the formulation of hybrid systems, the remainderof this section is devoted to constructing hybrid system models

for bipedal robots. Specifically, we consider two hybrid controlsystems corresponding to full and under actuation:

(2)

(3)

The constructions of these systems will be presented in the gen-eral case of a robot with a single discrete phase of walking and,in the case of underactuation, we will assume 1 degree of under-actuation. As a result they are applicable to both 2-D and 3-Drobots in the case of full actuation. In the case of underactua-tion, due to the assumption on 1 degree of underactuation, theyare only applicable to 2-D underactuated robots or 3-D robotswith feet where the underactuation is, for example, a result ofankles with only one degree of actuation. In the case of 3-Drobots with point feet, more elaborate constructions are neededto understand the zero dynamics [22] due to the fact that theyhave higher dimensionality.Note that the specific robot models for AMBER and NAO

considered in this paper are just a special case of (2) and (3),with the type of actuation and parameters, e.g., masses, lengthsand inertias, determined by the specific robot. For example, themass and inertia properties for AMBER were obtained via aSolidWorks model and these constants were obtained for NAOthrough the published values at [1].We suppress the dependenceof the models on these parameters for notational simplicity.Continuous Dynamics: Let be the configuration space

of a robot with degrees of freedom, i.e., , withcoordinates . For the sake of definiteness, it may be nec-essary to choose to be a subset of the actual configurationspace of the robot so that global coordinates can be defined,1 i.e.,such that is embeddable in , or more simply .Calculating the mass and inertia properties of each link of therobot using the specifications of the robot allows for the con-struction of the Lagrangian of the form

(4)

Explicitly, this is done symbolically through the method oftwists and exponential maps (see [34]). The Euler–Lagrangeequations yield the equations of motion of the form

(5)

where , i.e., depends on the degree of actua-tion. In the case of full actuation, . In the caseof underactuation, we will assume one degree of underactua-tion, and therefore .Converting the equations of motion (5) to a first-order ODE

yields the affine control systems

where , i.e., we obtain control systems for bothfull and under actuation. Finally, the set of admissible values forcontrol depends on the degree of actuation and the admissiblecontrol values, i.e., and .In the specific case of the robots AMBER and NAO, we

will consider a 2-D five-link point foot model of both of theserobots. In the case of AMBER this is a valid model of the

1Note that at various points we will assume that matrix functions have fullrank; it may be necessary to carefully choose to satisfy these conditions.

Page 4: IEEE TRANSACTIONS ON AUTOMATIC CONTROL, …edge.rit.edu/content/P15201/public/MSD II/Research Papers...IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 59, NO. 5, MAY 2014 1115 Human-Inspired

1118 IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 59, NO. 5, MAY 2014

Fig. 2. Modeled robot’s configuration, mass and length distribution, and virtualconstraints (outputs). (a) Configuration. (b) Mass/length. (c) Constraints.

system, since the robot is 2-D operating on a boom while in thecase of NAO this is a simplified model which only considersthe sagittal dynamics and abstracts away the behavior of thefeet which are compensated experimentally through localcontrollers. As a result of this choice of model, there are 5degrees of freedom and the coordinates of are denotedby where, as illustrated inFig. 2, is the angle of the stance foot, is the angle ofthe stance knee, is the angle of the torso with the stancethigh, is the angle of the non-stance thigh with the torso,and is the angle of the non-stance (or swing) knee. SinceNAO is fully actuated, based upon this choice of coordinates

. Since AMBER is underactuated at the stanceankle, consists of the second through fifth column of

. For NAO and AMBER and ,respectively (since we will, for the sake of simplicity, not putany restrictions on the values of the controllers).Domain and Guard: The domain specifies the allowable con-

figuration of the system as specified by a unilateral constraintfunction ; for the bipeds considered in this paper, this func-tion specifies that the non-stance foot must be above the ground,i.e., is the height of the non-stance foot and the system is sub-ject to the unilateral constraint . Therefore, the domain

is given by

The guard is just the boundary of the domain with the additionalassumption that the unilateral constraint is decreasing as fol-lows:

where is the Jacobian of at .Discrete Dynamics: The discrete dynamics of the robot de-

termine how the velocities of the robot change when the foot im-pacts the ground, while simultaneously switching the “stance”and “non-stance” legs. In particular, the reset map is givenby

(6)

where is the relabeling matrix which switches the stanceand non-stance leg at impact (by appropriately changing the an-gles). Here, determines the change in velocity due to im-pact; a detailed discussion on its computation can be found in[6], [21], and [26]. In particular, it is computed by consideringextended coordinates that include the position of the stance footand by employing a perfectly plastic impact law that results inthe pre-impact non-stance foot being fixed post-impact whereinit becomes the stance foot.

Fig. 3. Tiles of the walking gait for a human subject, plotted directly from thehuman walking data.

III. HUMAN OUTPUT COMBINATIONS

Because the ultimate goal of this work is to develop a con-trol scheme which yields stable human-like robotic walking, webegin by looking at human locomotion for motivation. In partic-ular, we consider experimental human walking data, and fromthis data we seek a low-dimensional representation of humanwalking through human outputs and canonical walking func-tions. In particular, examination of human walking data revealsthat certain outputs of the human locomotion system, i.e., func-tions of the kinematics which we term human output functions,can be represented as second-order system responses. In otherwords, we find that certain outputs of the human are describedby the time solution to a linear mass-spring-damper system:

(7)

where is the damping ratio, is the natural frequency,is the damped natural frequency, and are de-

termined by the initial conditions of the system, and is an “ac-celeration” related constant. This section will provide evidenceof this fact by considering human walking data and showing thatfunctions of this form indeed describe human walking with highcorrelation.

A. Human Walking Data

An integral part of the human-inspired nature of the con-trol scheme presented is its motivation from human locomo-tion. This takes the form of data associated with human walkingwhich will be considered throughout this paper. This data wascollected from nine subjects: two females and seven males withages ranging from 17 to 30, heights ranging from 160.0 cm to188.5 cm, and weights ranging from 47.7 kg to 90.9 kg. Thesubjects walked three meters along a line drawn on the floor at a“natural” pace, with each subject performing 11 trials. For eachtrial, LED’s were fixed to a test subject in key locations, suchas the joints, along the lower body—as the test subject walkedforward, the spatial position of each LED sensors was measuredat 480 Hz. Details on the experiment and data analysis can befound in [6] and [10].For purposes of this paper, the mean data from all nine sub-

jects are considered. A single step, consisting of heel strike toheel strike is isolated (see Fig. 3). The end result is discretetimes, , and discrete values for the angles of the humanmapped to the five-link robot model considered in this paper[see Fig. 2(a)], denoted by , where herewith the number of data points.

B. Human Outputs

We now construct a low-dimensional representation ofhuman walking from the given experimental data through theuse of human outputs. The goal is to determine a collectionof output functions, one for each degree of freedom of therobot, that are independent (the decoupling matrix associatedwith these outputs must be nonsingular) while simultaneouslygiving intuitive insight into human walking.

Page 5: IEEE TRANSACTIONS ON AUTOMATIC CONTROL, …edge.rit.edu/content/P15201/public/MSD II/Research Papers...IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 59, NO. 5, MAY 2014 1115 Human-Inspired

AMES: HUMAN-INSPIRED CONTROL OF BIPEDAL WALKING ROBOTS 1119

Fig. 4. Mean human output data computed over one step (heel strike to heel strike) from the data of nine subjects walking (black circles), with error bands showingone standard deviation from the mean (gray bands), and the canonical walking function fits (red line). The data for the was shifted to be zero at time zero.(a) ; (b) ; (c) ; (d) ; (e) .

Definition 1: A human output combination is a tupleconsisting of a configuration space , a ve-

locity-modulating output and position-modu-lating outputs . Let be an indexing forwhereby . A set of human outputs are in-dependent if2

(8)

on , and linear if

(9)

(10)

for and .In the case of the robotic models considered in this paper, we

will utilize the following human outputs:

(11)

(12)

where

(13)

(14)

(15)

with the linearization of the -position of the hip,is the linearization of the slope of the non-stance leg , (thetangent of the angle between the -axis and the line on the non-stance leg connecting the ankle and hip), and is the angle ofthe torso from zero. It is clear that this choice of human outputsare linear and independent. We denote the indexing set for theposition-modulating outputs by .The human outputs and from (11) and

(12) are computed from the experimental human walking data,, with the results given in Fig. 4. In addition, the error

bands in the figure show one standard deviation from this meanhuman data. The simple form that the outputs take in Fig. 4 mo-tivates the consideration of a special class of time based func-tions representing these outputs: canonical walking functions.It is important to note that numerous other human outputs

are possible, even in the case of the five-link robot model con-sidered in this paper. In particular, the actual position of the hip

2Here the rank is defined to be the rank of the Jacobian.

and non-stance slope, as nonlinear functions, were considered in[6], [46] yielding almost identical results. More generally, [45]and [46] considers collections of human outputs and analyzestheir ability to correctly describe human locomotion: examplesof other velocity-modulating outputs include the position of thecenter of mass, and position-modulating outputs such as the hipangle and leg lengths.

C. Canonical Walking Functions

Visually inspecting the human outputs, computed from thehuman data and shown in Fig. 4, they all appear to be describedby two simple functions. In particular, the linearized position ofthe hip appears to be essentially a linear function of time:

(16)

The remaining human outputs, , , , appearto be described by the solution to a linear mass-spring-dampersystem.With this in mind, define the canonical walking function

(17)

This function can be related to the more standard form of thetime solution of a mass-spring-damper system given in (7) bynoting that , , , , and .In particular, with (17), for the four remaining human outputs,, the claim is that we can describe their behavior as

(18)

where in (17), and the parame-ters of all of the outputs can be combined to yield a single vector:

. If it can be verifiedthat these functions accurately fit the mean human output data,then it can be concluded that humans appear to act like linearspring-mass-damper systems for the chosen outputs.

D. Human-Data-Based Cost Function

We now show that the canonical walking functions accuratelydescribe the human data. This is achieved by simply fitting (17)to the human walking data to achieve a least squares fit. Whilethis can be done automatically with a wide variety of preex-isting software, we prefer to achieve these fits by stating themin the form of an optimization problem. This is necessitatedby two facts: i) this same cost function will later be used toachieve robotic walking through a more sophisticated optimiza-tion problem, i.e., an optimization problem with constraints thatformally guarantee stable robotic walking and ii) it will form thebasis for the notion of admissible output combinations.Given a human output combination as

given in Definition 1 with associated human data

Page 6: IEEE TRANSACTIONS ON AUTOMATIC CONTROL, …edge.rit.edu/content/P15201/public/MSD II/Research Papers...IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 59, NO. 5, MAY 2014 1115 Human-Inspired

1120 IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 59, NO. 5, MAY 2014

defined at discrete times with , define thefollowing human-data-based cost function:

(19)

which is simply the sum of squared residuals, where here rep-resents a constant desired velocity [as will be seen in (21)] and

. To determine the parameters forthe human walking functions, we need only solve the optimiza-tion problem:

(20)

which yields the least squares fit of the mean human outputdata with the canonical walking functions. Therefore, given aset of human outputs, these outputs are accurately described bythe canonical walking function if is “small.” This,coupled with the desire to utilize the human outputs in controllerdesign, motivates the following definition of admissible humanoutput combinations.Definition 2: For sufficiently small, the human output

combination is admissible if the decou-pling matrix given in (29) is nonsingular and -admissibleif the parameters solving (20) satisfy .The main idea behind admissible human output combi-

nations is that one should choose output combinations thatare accurately described by the canonical walking function.Doing so will facilitate the human-inspired control introducedin Section IV; this is where the admissibility of the outputswill be formalized through the introduction of the decouplingmatrix. The requirement that stems from thefact that because human data can vary, a perfect fit is neverpossible but it should be “good enough.” The size, , of thisvariation between the time based representation and the actualdata should be kept small; naturally, this is both a functionof the data collection and the choice of outputs. In addition,since the human-inspired optimization introduced in Section Vwill minimize (19) subject to constraints, having a small initialvalue of the cost is desirable.In the case of AMBER and NAO, the same human output

combinations, (11) and (12), are considered; it is easy toverify that this output combination is admissible. Solving theoptimization problem (20) results in the parameters stated inTable I. In this case , and thereforethis output combination is -admissible with, for example,

. The small value of indicates that the canonicalwalking function accurately fits the human data for the out-puts chosen; this is further supported by the coefficients ofcorrelation given in Table I which range between 0.9862 to0.999 and visually through Fig. 4; note that the correlations fordetrended and zero mean data are essentially the same. Thisprovides evidence that the canonical walking function appears,in fact, to be canonical in some sense due to the small numberof parameters needed to accurately describe the human datacoupled with the intuitive representation of human walkinggiven—humans appear to act like linear spring-mass-dampersystems with respect to certain outputs.It is important to note that the human outputs considered,

and the fact that humans appear to act like linear mass-spring-

TABLE IPARAMETER VALUES OF THE CANONICAL WALKING FUNCTIONSOBTAINED FROM FITTING THE MEAN HUMAN DATA, TOGETHER

WITH THE CORRELATIONS OF THE FITS

dampers for these outputs, provides evidence for many of the“human-inspired” constructions considered in the past. In par-ticular, the specific choice of human outputs was motivated bythe compass gait biped [9], [17], [18], [46]; in essence, the (non-linear) form of these outputs represent the human as a compassgait biped. Moreover, the specific choice of canonical functionis motivated by the well-studied SLIP model [25], [29], [38]in that it represents human walking as essentially being the be-havior of a linear mass-spring-damper system for the outputschosen. Finally, it is important to note that the viewpoint onhuman walking taken in this paper, and the fact that output func-tions of a very special form are considered, differentiates it fromother work that has considered human data because previouswork considers, almost exclusively, joint angles over time [48],[49], [57].

IV. HUMAN-INSPIRED CONTROL

In this section, we construct a human-inspired controller thatdrives the outputs of the robot to the outputs of the human (asrepresented by canonical walking functions). In particular, inboth the case of full and under actuation, we consider an admis-sible output combination and constructa feedback controller through the use of a parametrization oftime based upon the velocity-modulating output. The end resultis feedback controllers that will be used to obtain stable bipedalrobotic walking.

A. Output Functions

Based upon the human outputs and their time-based represen-tation given by the canonical walking functions, we define out-puts for the robot based upon our desire for the robot to have thesame output behavior as the human.With the goal of controllingthe velocity of the robot, we define the relative degree 1 outputsto be the velocity of the velocity-modulating output—hence themotivation for the name—and a desired velocity:

(21)

Similarly, with the goal of the robot tracking the canonicalwalking function (18) through the position-modulating outputs,we consider the following actual and desired outputs:

(22)

In the case when is a linear and indepen-dent output combination, it follows that

(23)

by (8)–(10).

Page 7: IEEE TRANSACTIONS ON AUTOMATIC CONTROL, …edge.rit.edu/content/P15201/public/MSD II/Research Papers...IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 59, NO. 5, MAY 2014 1115 Human-Inspired

AMES: HUMAN-INSPIRED CONTROL OF BIPEDAL WALKING ROBOTS 1121

B. Parameterization of Time

Motivated by the desire to create autonomous controllers,which are more robust than time-based controllers, we intro-duce a state-based parameterization of time in our system as iscommon practice [48], [54]. Examination of human data revealsthat the (linearized) forward position of the hip evolves in anapproximately linear manner with respect to time (16). Takingadvantage of this observation, for an admissible output combi-nation , the following parameterization oftime is considered:

(24)

where is the value of the velocity-modulating output ofthe robot at the beginning of the current step, and is, as will beseen, the desired velocity (which is just an element of the vectorof parameters ).

C. Fully Actuated Control

The goal is to drive the outputs of the robot, obtained througha human output combination, to the outputs of the human asrepresented by the canonical walking functions parameterizedby . This motivates the final form of the outputs, termedthe human-inspired outputs for full actuation, given by

(25)

(26)

These outputs can be grouped together to form a single vectorof outputs:

(27)

where and will be seen to be relative degree 1 and (vector)relative degree 2 outputs, respectively, due to the assumption ofadmissible output combinations coupled with the fact that isthe output to a mechanical system that is only dependent on .The feedback linearization controller for full actuation, de-

noted by and termed the human-inspired controllerfor full actuation due to its dependence on the human-inspiredoutputs, can now be stated as

(28)

with control gain and decoupling matrix

(29)

Note that the decoupling matrix is nonsingular by the defini-tion of an admissible output combination (see Def. 2). Moregenerally, this follows from the fact that care was taken whendefining the human output combinations so that they are inde-pendent (see Def. 1). It follows that results in dy-namics on the outputs given by

(30)

(31)

and therefore, for a control gain , the control lawrenders the output exponentially stable [42]. That is, the human-inspired outputs and exponentially at a rate of; in other words, the outputs of the robot will converge to thecanonical walking functions exponentially.Applying the feedback control law in (28) to the hybrid

control system modeling the bipedal robot being considered,as given in (2), yields a hybrid system

(32)

where , , and are defined as for , and

(33)

where the dependence of on the vector of parameters ofthe canonical walking functions, , and the control gain for theinput/output linearization control law, , has been made explicit.

D. Underactuated Control

When the robot is underactuated, there is no longer the con-trol authority to drive all of the outputs of the system to be inagreement with the human outputs. In particular, since we willno longer have control of the stance ankle (as in the case ofAMBER, due to its point feet), we can no longer directly con-trol the forward movement of the velocity modulating outputthrough the output (25). Thus, in the case of underactuation, weonly consider the relative degree 2 outputs given in (26).Motivated by these considerations related to underactuated

control, for the affine control system associated withthe robotic model (3), we define the human-inspired controllerfor underactuation:

(34)

with as defined in (26) and the underactuated decou-pling matrix: . As in the caseof full actuation, the decoupling matrix is nonsingular becauseof the choice of output functions and the fact that we assumeadmissibility of the outputs. In addition, the control lawresults in dynamics on the outputs given by (31) and thereforerenders the relative degree two outputs, , exponentially stable,i.e., exponentially at a rate of . Due to the fact that thesystem is underactuated, the output will no longer convergeto 0. Thus, the velocity of the velocity modulating output cannotbe directly controlled in the case of underactuation.Applying the feedback control law in (34) to the hybrid con-

trol system modeling and a bipedal robot with 1 degree of un-deractuation, as given in (3), yields a hybrid system

(35)

where , , and are defined as for , and

V. HUMAN-INSPIRED PARTIAL HYBRID ZERO DYNAMICS

This section considers the fully actuated model of a bipedalrobot , with the goal of formally obtaining human-likebipedal robotic walking. In particular, we determine parameters

Page 8: IEEE TRANSACTIONS ON AUTOMATIC CONTROL, …edge.rit.edu/content/P15201/public/MSD II/Research Papers...IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 59, NO. 5, MAY 2014 1115 Human-Inspired

1122 IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 59, NO. 5, MAY 2014

of the human-inspired controller that result in the best fit withthe human output data while simultaneously guaranteeingrobotic walking. To this end, we introduce a novel concepttermed partial hybrid zero dynamics (PHZD), and presentan optimization problem that determines the parameters thatprovide the best fit to the human data while ensuring PHZD. Inaddition, we prove that these parameters—as a consequence ofthe satisfaction of the PHZD constraints—will automaticallyguarantee robotic walking. The fixed point of this walkinggait is explicitly constructed, and it is formally proven thatthe periodic orbit associated with this fixed point is locallyexponentially stable.Throughout this section, and the remainder of this paper, we

make the following assumption:A1: The human output combination is

linear, independent and -admissible.

A. Full Hybrid Zero Dynamics (FHZD)

As stated in Section IV, the goal of the control law (28)is to drive the human-inspired output exponentiallyat a rate of . Therefore, for the continuous dynamics, the con-troller renders the full zero dynamics surface

(36)exponentially stable; moreover, this surface is invariant for thecontinuous dynamics of the hybrid system . Note thathere is a vector of zeros of appropriate dimension andwemakethe dependence of on the set of parameters explicit. It is atthis point that continuous systems and hybrid systems diverge:while this surface is invariant for the continuous dynamics, it isnot necessarily invariant for the hybrid dynamics. In particular,the discrete impacts in the system cause the state to be “thrown”off of the full zero dynamics surface. Therefore, a hybrid systemhas full hybrid zero dynamics (FHZD) if the full zero dynamicsare invariant through impact: . It turnsout that this condition is too strong in that—due to the differ-ences between the robotic model and the human—it is not pos-sible to obtain “good” walking that satisfies this requirement.Instead, due to these differences, it is more beneficial to allowthe velocity of the velocitymodulating output to “jump” throughimpact, while requiring the other relative degree 2 outputs to re-main on the invariant. This motivates the consideration of par-tial hybrid zero dynamics.

B. Partial Hybrid Zero Dynamics (PHZD)

While the realization of FHZD is the “best case scenario,” itis quite difficult in the case of bipedal robotic walking since itwould force the hybrid system to evolve on a one-dimensionalmanifold. Therefore, we enforce zero dynamics only for the rel-ative degree 2 outputs. We refer to this as the partial zero dy-namics surface, given by

(37)Themotivation for considering this surface is that it allows some“freedom” in the movement of the system to account for differ-ences between the robot and human models. Moreover, sincethe only output that is not included in the partial zero dynamics

surface is the output that forces the velocity of the velocity mod-ulating output to be constant, enforcing partial hybrid zero dy-namics simply means that we allow the velocity of the velocitymodulating output to compensate for the shocks in the systemdue to impact.

C. Problem Statement

The goal of human-inspired PHZD is to find parametersthat solve the following constrained optimization problem:

(38)

(PHZD)

with the cost given in (19). This is simply the optimiza-tion problem in (20) that was used to determine the parametersof the canonical walking functions that gave the best fit of thehuman walking functions to the human output data, but subjectto constraints that ensure PHZD. The formal goal of this sec-tion is to restate (PHZD) in such a way that it can be solvedpractically.

D. Partial Zero Dynamics

This section utilizes the fact that the human outputs werespecifically chosen to be linear in order to explicitly constructthe partial hybrid zero dynamics. In particular, we reformulatethe constructions in [54] in a way applicable to full-actuationand reframe them in the context of canonical walking functions.Because of the specific form of in (23) due to the linearoutput assumption, we begin by picking the following coordi-nates for the partial zero dynamics surface:

(39)

where is obtained from (9). Moreover, since is justthe velocity-modulating output which was used to parameterizetime (24), we can write . Picking thecoordinates

(40)

with as in (10), and defining

(41)

it follows that for and , the point.

As a result of the fact that we have full actuation and com-pletely linearize the dynamics with (28), it follows that the rel-ative degree 1 output (25) evolves according to .Therefore, the partial hybrid zero dynamics evolve according tothe linear ODE:

(42)

The advantage of the partial zero dynamics representationintroduced is that its linear form allows for the existence

Page 9: IEEE TRANSACTIONS ON AUTOMATIC CONTROL, …edge.rit.edu/content/P15201/public/MSD II/Research Papers...IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 59, NO. 5, MAY 2014 1115 Human-Inspired

AMES: HUMAN-INSPIRED CONTROL OF BIPEDAL WALKING ROBOTS 1123

and stability of a fixed point of the zero dynamics to bedetermined in closed form. Specifically, given a pre-impactpoint on the guard with its post-impact state

, we can compute and. From this, if (PHZD) is satisfied, the change in

and due to this impact can be determined through

(43)

where is a point that is chosen a priori and

(44)

In essence, this defines a 2-D hybrid system and, since the partialhybrid zero dynamics is 2-D, when considering the existenceand stability of a periodic orbit in this surface one need onlyconsider the restricted Poincaré map

(45)

where depends on due to (42) and

(46)In other words, the hyperplane can be chosen as thePoincaré section. The Poincaré map for the partial hybrid zerodynamics is therefore a 1-D (partial) map

, and so can be viewed as only a function ofwhich therefore defines a discrete time dynamical system:

and the stability of this 1-D discrete time dynamical system, aswill be seen in Theorem 2, completely determines the stabilityof a periodic orbit in the full dimensional system.

E. Calculation of the Fixed Point Corresponding to PartialZero Dynamics

To achieve the goal of restating (38) in a way that is indepen-dent of state variables (position and velocity), we seek to ex-plicitly construct the point that will be utilized in theconstructions outlined in (43) and (44). To this end, we can usethe outputs and guard functions to explicitly solve for the config-uration of the system on the guardin terms of the parameters . In particular, let

(47)

where is the relabeling matrix (6). Note that exists be-cause of the specific structure of the outputs, , chosen.In fact, the reason for considering at is because it impliesthat the configuration at the beginning of the step isand thus implying that:

, or there is a solution to (47) because of the simpleform that takes at .Using , we can explicitly solve for a point

. In particular, let

(48)

It follows from the definition of and in (25) and (26) that

(49)

Therefore, define

(50)

where is invertible because of the choice of outputs. The pointis essential to all of the constructions in this paper,

as it will: a) be used to remove state-dependence in the opti-mization, b) be a fixed point of a periodic orbit, and c) be usedto determine the stability of the periodic orbit by taking it to bethe point chosen a priori in (43).

F. Human-Inspired Optimization

With the notation of this section in hand, we define a human-inspired optimization problem (first introduced in [6]). This con-strained optimization uses the human data as a cost function(through the human-data-based cost (19)), but enforces con-straints that, as will be seen in the main result, ensure that thebipedal robot has a stable walking gait.Theorem 1: The parameters solving the constrained opti-

mization problem

(HIO)

(C1)

(C2)

(C3)

yield partial hybrid zero dynamics: .Proof: Let be the solution to the optimization problem

(HIO). We begin by showing thatsince this will imply that . By (C1)and (49), . Moreover, by (47) (specif-ically, the fact that ) and (C3), it follows that

. Therefore,. Now, and intersect transversally since

by (C3). Since is a 1-D submanifold of , it followsthat is a unique point. Therefore,

.To show that we need only show

that . From the form of given in(6), the requirement that is equiv-alent to the following conditions being satisfied:

(51)

(52)

By the definition of , and specifically (47), (51) is satis-fied. Moreover,

Page 10: IEEE TRANSACTIONS ON AUTOMATIC CONTROL, …edge.rit.edu/content/P15201/public/MSD II/Research Papers...IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 59, NO. 5, MAY 2014 1115 Human-Inspired

1124 IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 59, NO. 5, MAY 2014

Therefore, (52) is satisfied as a result of (C2). Thus, we haveestablished that .To complete the proof we note that

implies that by [54, Th. 5.2] sinceis a point satis-

fying (51) and (52).

G. Exact Reconstruction

One of the main advantages to considering partial hybrid zerodynamics in the case of full actuation is the simple form thatthe partial zero dynamics takes. In particular, using (42) wecan explicitly, and in closed form, reconstruct the solution ofthe full-order system. This is outlined in the following propo-sition—a result that is essential to both proving the main resultin the case of full actuation along with adding constraints thatensure the physical realizability of the resulting walking gait.Consider the point and let

and be the representation of this point in thepartial hybrid zero dynamics coordinates (39); note that from(50), . By picking in (43), we obtain

. From the constructions presented in this section, andbecause , by viewing asa 1-D space as in (46), for an initial conditionit follows from (42) that the solution to the partial hybrid zerodynamics is given by

(53)

With this notation in hand, we present the following result onreconstructing solutions for the full-order system in closed formthrough partial hybrid zero dynamics.Proposition 1: Let be the parameters solving (HIO) and

assume that

For , the time-to-impact function for the partialhybrid zero dynamics is given by

(54)with

(55)

where is the LambertW function (or product logarithm) and3

(56)

Moreover, for the hybrid system with initial condition, the time to im-

3Note that determines the change in the velocity of the velocity modulatingoutput relative to ; or, in other words, the perturbation away from the full zerodynamics surface . That is, if we were to require the stronger condition offull hybrid zero dynamics, , then fromwhich it would follow that .

pact function is given by and the solu-tion to the continuous dynamics (33) is given by

(57)

for .Proof: For , the corresponding initial

condition for the partial hybrid zero dynamics is given by thepre-impact point . Correspondingly, applying thereset map to this point for the partial hybrid zero dynamics(43) yields the initial condition forthe partial zero dynamics (42) with solution given by (53).According to the definition of the guard for the partial hybridzero dynamics (46), to establish that is the time-to-im-pact function it is necessary to determine the time, , suchthat . It is easy to verify through directcalculation that solving for yields as desired.The fact that the solution to is given by (57) for

the given initial condition follows from. In particular, the system evolves according to the par-

tial hybrid zero dynamics (42) with solution given by (53). Bythe definition of , it follows that (41) can be used to ob-tain from for all

.

H. Main Fully Actuated Result

The main result of this paper, in the context of full actuation,is that the point , determined through the inversekinematics and utilizing the parameters obtained by solving theoptimization problem in Theorem 1, is “essentially” the fixedpoint to a stable hybrid periodic orbit. Thus, the constrained op-timization problem (HIO) not only ensures partial hybrid zerodynamics, but it automatically yields a fixed point to a stablewalking gait that can be computed in closed form from the pa-rameters of the human-inspired controller. Moreover, since thecost function (19) only depends on human walking data, we au-tomatically generate a controller for a stable walking gait, its pa-rameters, a stable hybrid periodic orbit and its fixed point usingonly human data.Theorem 2: Let be the parameters solving (HIO). If

then there exists a constant such thatfor all the hybrid system has an exponen-tially stable periodic orbit, , dependent on , with fixedpoint and period which satisfy the followingproperties:

(P1)

(P2)

(P3)

where .Proof: From the constructions related to Proposition 1, the

Poincaré map (45) can be explicitly computed; in fact, it is givenby (55). To see this, for , it is given by

Page 11: IEEE TRANSACTIONS ON AUTOMATIC CONTROL, …edge.rit.edu/content/P15201/public/MSD II/Research Papers...IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 59, NO. 5, MAY 2014 1115 Human-Inspired

AMES: HUMAN-INSPIRED CONTROL OF BIPEDAL WALKING ROBOTS 1125

with and given in (54) and (56), respectively.From the explicit form of the reduced Poincaré map it followsthat

(58)

since and .To prove the existence of a periodic orbit for the partial hybrid

zero dynamics, we need only prove the existence of a fixed pointfor . Consider a ball of radius around , i.e., for

, . Then for this it follows by (58) thatthere exists a such that for all ,. Therefore, . By the Brouwer fixed-point theorem, it follows that there exists a fixed point of ,i.e., , dependent on and satisfying ,and the hybrid partial zero dynamics has a periodic orbit. Toprove the stability of this periodic orbit, we need only check thederivative of at and ensure that its magnitude is lessthan 1. The derivative can be explicitly computed as

if

otherwise.

Since as , it follows that

Therefore, there exists an such that for ,establishing the stability of the periodic orbit

for the partial hybrid zero dynamics.To establish the existence of a stable periodic orbit for

, by [54, Th. 4.5] (see also [33]), a stable fixed point forthe restricted Poincaré map implies that

(59)

is a stable fixed point of the Poincaré map for the hybridsystem for sufficiently large, i.e., for . Picking

implies the desired existence and stabilityresult.To complete the proof, note that (P1)–(P3) follow

from Proposition 1 applied to . In particular, in-spection of (54) yields (P1). Since

, (P2) follows from (59) since. Finally, (P3) follows

from the fact that for the initial condition the periodicorbit is defined by (57) with . Using (P1) coupledwith the fact that, from (53), and

yields the desired result.

I. Imposing Physical Realizability Constraints

Theorem 2 demonstrated that partial hybrid zero dynamicsimplies the existence of a stable periodic orbit that can be com-pletely characterized in the limit. Yet, for the purpose of realiza-tion on physical robots, convergence is often not enough; it isdesirable to know, in closed form, the solution of the system sothat constraints on physical realizability can imposed directly inthe optimization in closed form (e.g., constraints on maximumtorque, foot height, etc.). The importance of Proposition 1, in

Fig. 5. Walking gaits for AMBER in the case of both full (left) and under (right)actuation. The periodic orbits corresponding to the walking gaits are shown in(a) for the fully actuated case and (c) for the underactuated case. Similarly, thetorques for the fully and under actuated case are shown in (b) and (d). (a) Peri-odic Orbit (FA). (b) Torques (FA). (c) Periodic Orbit (UA). (d) Torques (UA).

addition to the role it plays in proving Theorem 2, is that it al-lows for these conditions to be enforced in closed form. To pro-vide a specific example, suppose that the maximum allowabletorque by the motors is . For a given we can enforcethis constraint by solving for such that .Using in Proposition 1, through (57) the periodic solution

to the hybrid system can be con-structed in closed form, resulting in the constraint

(C4)

which, when added to (HIO), will ensure that the maximum al-lowable torque is not violated (assuming a feasible solution tothe optimization problem is found). The same approach can beutilized to implement constraints on maximum power, angularvelocity, minimum foot height (i.e., scuffing prevention) andZMP conditions. In essence, any behavior that can be quanti-fied in terms of the solution to the dynamics can be codified inclosed form through the closed form solution obtained from par-tial hybrid zero dynamics via Proposition 1.

J. Simulation Results

It will now be demonstrated how these formal results can beapplied to models of bipedal robots, and specifically AMBERand NAO, to achieve walking in simulation. These simulationresults serve as the basis for the control of the actual NAO robot

Page 12: IEEE TRANSACTIONS ON AUTOMATIC CONTROL, …edge.rit.edu/content/P15201/public/MSD II/Research Papers...IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 59, NO. 5, MAY 2014 1115 Human-Inspired

1126 IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 59, NO. 5, MAY 2014

Fig. 6. Outputs of AMBER (blue) for fully actuated (a) and underactuated (b) walking obtained plotted against the mean human output data computed from thedata of nine subjects walking, with error bands showing one standard deviation from the mean. (a) Human Outputs and Outputs for AMBER for Fully ActuatedWalking. (b) Human Outputs and Outputs for AMBER for Underactuated Walking.

to obtain walking experimentally as will be discussed at the endof this section.AMBER: We begin with the hybrid model of AMBER,

, obtained using the mass, length and inertia values ofthis robot (in this case, we assume the robot is fully actuatedfor the sake of simulation only—we later consider the morephysically realistic case of underactuation). Utilizing the humanoutput combination specified in (11)–(15) and the associatedhuman output data, we numerically solve the optimizationproblem in Theorem 1. In particular, we utilize only the humandata (for the cost in (HIO)), and seed the optimization problemwith the parameters of the canonical walking functions that bestfit the human output data (as given in Table I). The end resultof this optimization is parameters for the human-inspiredcontroller that provably result in stable robotic walking due toTheorem 2, along with the initial condition to this walking gait:

.Comparing the parameters of the human-inspired controller,

and thus the canonical walking functions, obtained as a result ofTheorem 1 against the mean human walking data (see Fig. 6) itcan be seen that the resulting robotic walking is in fact “human-like.” In particular, in [37], “normal” human walking is de-fined to be walking that lies within one standard deviation ofmean human walking. In the case of the considered outputs, thewalking for AMBER lies almost entirely within this one stan-dard deviation band indicating “normal” human walking. Theone exception is the velocity of the hip which is modulated totake into account the physical differences between AMBER andthe humans from whom the data was collected, i.e., the dif-ference in link lengths between AMBER and the human sub-jects results in different linear velocities for the same angularvelocities. The walking gait obtained, as simulated from theinitial condition with , can be seen inFig. 5, with the corresponding periodic orbit and torques re-quired shown in Fig. 5(a) and (b). While Theorem 2 formallyguarantees stability of the periodic orbit, we verify this numer-ically by computing the eigenvalues of the Poincaré map; wefind the largest magnitude of an eigenvalue to be 0.0172, indi-cating local exponential stability (in fact, all other eigenvaluesare essentially zero as the proof of Theorem 2 predicts).NAO: We apply Theorem 1 to the hybrid model of NAO

where, to achieve practical results (with a view towards im-

Fig. 7. Periodic orbits for the simulated behavior of NAO starting from rest,i.e., a zero initial condition (a) and the fixed point of the periodic orbit (b).Starting from rest, the actual outputs of NAO converge to the desired outputsand display partial hybrid zero dynamics (c). Snapshots of the walking gaitsfrom the NAO simulation starting from zero initial conditions (d) and contin-uing from the fixed point (e). (a) Periodic Orbit from Rest. (b) Periodic Orbitfrom Fixed Point. (c) Outputs from Rest. (d) Gait Tiles from Rest. (e) Gait Tilesfrom Fixed Point.

plementation), additional constraints were enforced in this op-timization which limited the maximum joint velocity to 3 rad/sand ensured proper foot clearance. The end result is parametersfor the human-inspired controller that (due to Theorem 2) re-

sult in stable robotic walking for sufficiently large control gain. Picking and simulating the hybrid system

Page 13: IEEE TRANSACTIONS ON AUTOMATIC CONTROL, …edge.rit.edu/content/P15201/public/MSD II/Research Papers...IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 59, NO. 5, MAY 2014 1115 Human-Inspired

AMES: HUMAN-INSPIRED CONTROL OF BIPEDAL WALKING ROBOTS 1127

Fig. 8. Simulated (desired) and experimental joint trajectories in the sagittal plane for NAO. (a) Stance Ankle. (b) Stance Knee. (c) Stance Hip. (d) Non-stanceHip. (e) Non-stance knee.

Fig. 9. Simulated (desired) and experimental joint trajectories for the sagittalplane (a). Comparison of the actual (top) and simulated (bottom) walking gaitsfor NAO over one step (b).

(with the mass, lengths, and inertia values for NAO) from theinitial condition verifies that we do, in fact, havea walking gait, i.e., a periodic orbit (see Fig. 7(b) and (e)).Moreover, we can verify the fact that the chosen results in astable walking gait by checking the eigenvalues of the Poincarémap; we find that the magnitude of the maximum eigenvalue is

, thus verifying the exponential stability guaranteedby Theorem 2. Furthermore, and as indicated in Fig. 7(c), the re-sulting walking exhibits partial hybrid zero dynamics. In addi-tion to stable, periodic walking, the robustness of the human-in-spired control law allows for the robot to start from rest and con-verge to the walking periodic orbit corresponding to the walkinggait. As shown in Fig. 7(a), trajectories of the system whenstarted from rest converge to the stable limit periodic orbit pre-dicted by Theorem 2. Convergence is also seen in a plot of thehuman inspired outputs; in Fig. 7(c) the convergence of the ac-tual outputs of the robot to the desired outputs can be seen. Tilesof the first step of the walking, starting from rest, can be seenin Fig. 7(d). These simulated trajectories will be used to exper-imentally achieve walking on the real NAO robot.

K. NAO Experimental Results

Implementing the simulated trajectories of the human-in-spired walking control on NAO results in dynamically stablewalking on the actual NAO robot. In particular, the simulatedtrajectories, starting from rest, are implemented on the sagittalactuators of the robot through trajectory tracking, and a simplefeedback control law is implemented in the lateral plane (moredetails can be found in [8]). Tiles of the walking gait achievedon the NAO robot can be seen in Fig. 9 where the experimentalwalking is compared to tiles of the simulated walking takenat the same time instances showing that, in fact, there is goodagreement. To provide qualitative evidence of this, the simu-lated and experimentally observed figures are plotted in Fig. 8where very good agreement between simulation and experi-mentation is shown. Finally, the human-inspired walking thatwas obtained on NAO subjectively appears more human-likethan other walking gaits that have been achieved for NAOthrough ZMP related methods [27], [51]. We invite readersto form their own opinions by watching the video [3] of thehuman-inspired robotic walking and its comparison with thepre-existing (ZMP) walking.

VI. HUMAN-INSPIRED HYBRID ZERO DYNAMICS

Having demonstrated how human-inspired control automati-cally results in “human-like” robotic walking for fully actuatedbipedal robots, this section shows how robotic walking is alsoautomatically obtained from human data in the case of under-actuation. In particular, we now consider hybrid zero dynamics(HZD) as originally developed for the underactuated control ofbipedal robots [20], [54]. In the framework of HZD, we add con-straints to the human-inspired optimization in Theorem 1 so thatthe parameters that guarantee HZD also imply the existence ofa stable walking gait for an underactuated robot. Moreover, wecan explicitly compute the fixed point to the stable periodic orbitcorresponding to this stable walking gait. It is important to notethat formulations considered in this section utilize the results of[54]; the novelty lies in the different optimization problem usedto obtain HZD and, more importantly, the fact that the same opti-mization problem that yields walking for the fully actuated caseyields walking in the underactuated case with the addition oftwo constraints.

A. Problem Statement

The goal of the underactuated human-inspired controller (34)is to drive the outputs of the robot to the outputs of the human:

. In other words, the controller renders the zerodynamics surface

(60)exponentially stable; moreover, this surface is invariant for thecontinuous dynamics of the hybrid system defined in(35). Note that the hybrid zero dynamics surface is simply thepartial hybrid zero dynamics surface, , considered in the caseof full actuation; that is, for underactuation , and thiswas one of the original motivations for considering the partialhybrid zero dynamics surface. As in the case of full actuation,the goal of human-inspired HZD is to find parameters thatsolve the following constrained optimization problem:

(61)

(HZD)

with the cost given in (19). The formal goal of thissection is to restate (HZD) in such a way that it can be practicallysolved through the utilization of the constructions for the fullyactuated case, but suitably modified for underactuation.

B. Zero Dynamics

The zero dynamics must be constructed so as to take intoaccount the fact that the system is underactuated. In particular,the zero dynamics will no longer be linear as in the case of fullactuation. To this end we will utilize the constructions in [54],reframed in the context of canonical walking functions. Because

Page 14: IEEE TRANSACTIONS ON AUTOMATIC CONTROL, …edge.rit.edu/content/P15201/public/MSD II/Research Papers...IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 59, NO. 5, MAY 2014 1115 Human-Inspired

1128 IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 59, NO. 5, MAY 2014

of the specific choice of in (22), we begin by picking thefollowing coordinates for the zero dynamics:

(62)

where is the first row of the inertia matrix in (4).Moreover, as in the case of full actuation, we can again write

.Picking the coordinates and as in (40) and defining

it follows that for and that. Finally, the zero dynamics evolve according to the ODE

(63)

with the potential energy of the robot (4).The advantage of the low-dimensional zero dynamic repre-

sentation introduced is that it allows for the existence and sta-bility of a fixed point of the zero dynamics to be determined apriori. In the case of underactuation, this is achieved by con-sidering the energy of the zero dynamics, and in particular thepotential energy. As in the case with full actuation, we beginwith a point on the guard with its post-impactstate , we can computeand . From this, the change in due to this im-pact can be determined through

(64)

wherein .Picking the same Poincaré section as in the case of full actua-

tion (46), the stability of a periodic orbit can be characterized byone-dimensional discrete time dynamical system

, given by

where

is the potential energy of the zero dynamics (63). Energymethods can be coupled with these constructions to determinewhen the robot will take a step by considering the constant

with . Specifically, ifit will imply the existence of a limit cycle in the hybrid zero

dynamics surface, and if it will imply thestability of that limit cycle [54].

C. Main Underactuated Result

We now present the main result of this paper related to under-actuated walking. Picking and [see (47) and (50)] asthe choice of , constraints can be added to the human-

inspired optimization problem (HIO) to automatically obtain aninitial condition corresponding to stable periodic walking for anunderactuated robot.Theorem 3: The parameters solving the constrained opti-

mization problem (HIO) in Theorem 1 subject to the additionalconstraints

(C4)

(C5)

yields hybrid zero dynamics: . Moreover,there exists an such that for all the hybrid system

has a stable periodic orbit with fixed pointgiven by

Proof: From Theorem 1, it follows that conditions(C1)–(C3) imply partial hybrid zero dynamics. Since in the caseof underactuation and it follows that theseconditions imply hybrid zero dynamics: .Utilizing Theorem 5.3 in [54], conditions (C4) and (C5) implythe desired result.

D. Simulation Results

Since AMBER (Fig. 1) is an underactuated robot due to itspoint feet, we apply the results of this section to AMBER toobtain underactuated human-inspired robotic walking in simu-lation. We then discuss how this simulated walking can be usedas a basis for the controller design on the physical bipedal robotwith the end result being real-world robotic walking.We begin with the application of Theorem 3 to the under-

actuated hybrid model of AMBER. Specifically, we solve theoptimization problem in Theorem 1 subject to the additionalconstraints in Theorem 3. The end result of this optimizationis a collection of control parameters for the resulting in a hy-brid system . Moreover, the same optimization automat-ically generates a fixed point to a stable periodic orbit (shownin Fig. 5(c)); this is verified by picking and checking theeigenvalues of the linearization of the Poincaré map for whichthe maximummagnitude is 0.7761 (and hence less than 1). Tilesof the walking obtained in simulation can be seen in Fig. 5. Inaddition, it can be seen in Fig. 5(d) that the walking is very lowtorque (in fact, the torque is far lower than for the fully actuatedwalking obtained). What is especially interesting is the compar-ison of the outputs, , of the robot against the human outputdata as shown in Fig. 6. Despite the fact that the robot consistsof only 5 degrees of freedom and is underactuated at the ankle,the end result is still “human-like” walking—the outputs of therobot fall within one standard deviation of the mean human datafor a majority of the outputs.

E. Experimental Results

The underactuated walking obtained through simulationforms a basis for the controllers realized on the physical robotAMBER. The implementation details are involved (see [58]),and do not lie within the scope of this paper, but the matrix ofparameters is used as the control parameters for a voltagebased feedback proportional controller that was also shownto obtain walking in simulation (shown in Fig. 11). Finally,this feedback voltage control was implemented on the physicalrobot through National Instruments CompactRIO system. Theend result was human-inspired robotic walking as can be seen

Page 15: IEEE TRANSACTIONS ON AUTOMATIC CONTROL, …edge.rit.edu/content/P15201/public/MSD II/Research Papers...IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 59, NO. 5, MAY 2014 1115 Human-Inspired

AMES: HUMAN-INSPIRED CONTROL OF BIPEDAL WALKING ROBOTS 1129

Fig. 10. Simulated (desired) and experimental joint trajectories for AMBER.(a) Stance Knee. (b) Stance Hip. (c) Non-stance Hip. (d) Non-stance knee.

Fig. 11. Comparison of the simulated and experimental walking gait forAMBER over one step.

Fig. 12. Comparison human output data and outputs for AMBER calculatedfrom the experimentally realized walking over one step.

in Fig. 11. Good agreement between the simulated and experi-mental behavior was observed; for example, the simulated andexperimental trajectories can be seen in Fig. 10. Note that thedifferences between these trajectories are a result of the fact thatthe simulated trajectories are not tracked directly, but ratherfeedback voltage control is used, which can result in drift.Nonetheless, the end result of this approach is “human-like”robotic walking; this can best be seen by watching the actualwalking which can be found at [2], and is quantified by thecomparisons between the human output data and the outputsof the robot shown in Fig. 12. In addition, the human-inspiredwalking appears to be robust. As can be seen at [5], the robotAMBER is subjected to disturbances both on the robot and asa result of the environment. The robot is able to react to these

disturbances in such a way that walking is maintained and,more subjectively, AMBER reacts in a “human-like” manner.

VII. CONCLUSION

This paper presents a formal human-inspired approach tobipedal robotic walking, proving through Theorem 1, 2, and3 that by using only human data, parameters to the human-in-spired controller can be determined that simultaneously providethe best fit of the human data, yield (partial) hybrid zero dy-namics, imply the existence of a stable walking gait, and allowthe fixed point for this stable walking gait to be explicitlycomputed. Moreover, these results are established in the caseof both full and under actuation. To demonstrate these results,walking gaits are generated in simulation for two models ofbipedal robots: AMBER and NAO. In the case of AMBER,both full and under actuated walking is achieved that is re-markably “human-like.” In the case of NAO, the robustness ofthe walking obtained through human-inspired control allowsfor both steady state walking and walking starting from rest.Finally, the formal results presented are utilized to achieverobotic walking in experimentation on both AMBER in NAO.In particular, 3-D fully actuated bipedal robotic walking isobtained for NAO, and 2-D underactuated bipedal roboticwalking is achieved with AMBER. This provides evidence ofthe practical applicability of the formal results presented inthis paper. This hope is, therefore, that these results lay thegroundwork for a “human-inspired” framework for roboticlocomotion controller design.

ACKNOWLEDGMENT

The author would like to thank the members of AMBER Labfor their contributions in developing the framework of human-inspired control and their essential work in experimentally re-alizing the formal results presented in this paper. The authorwould also like to thank R. Bajcsy for introducing him to theworld of experimentation and J. Grizzle for the many enlight-ening discussions on hybrid zero dynamics. Finally, the authorwould like to thank the anonymous reviewers and the AssociateEditor for their insightful and constructive comments on thismanuscript.

REFERENCES[1] [Online]. Available: http://www.aldebaran-robotics.com/[2] AMBER-Lab, First Human-Inspired Robotic Walking With AMBER

[Video file], [Online]. Available: http://youtu.be/SYXWoNU8QUE[3] AMBER-Lab, First Human-Inspired Robotic Walking With NAO

[Video file], [Online]. Available: http://youtu.be/OBGHU-e1kc0[4] AMBER-Lab, Human-Inspired Robotic Walking With AMBER 2.0

[Video file], [Online]. Available: http://youtu.be/d6oM5sLI9vA[5] AMBER-Lab, Robustness Tests on the Bipedal Robot AMBER [Video

file], [Online]. Available: http://youtu.be/RgQ8atV1NW0[6] A. D. Ames, “First steps toward automatically generating bipedal

robotic walking from human data,” in Robotic Motion and Control2011, ser. LNICS. New York, NY, USA: Springer, 2012, vol. 422,pp. 89–116.

[7] A. D. Ames, “First steps toward underactuated human-inspired bipedalrobotic walking,” inProc. IEEE Int. Conf. Robot. Autom., St. Paul,MN,USA, 2012, pp. 1011–1017.

[8] A. D. Ames, E. A. Cousineau, and M. J. Powell, “Dynamically stablerobotic walking with NAO via human-inspired hybrid zero dynamics,”in Proc. Hybrid Syst.: Comput. Control, Beijing, China, 2012.

[9] A. D. Ames, R. D. Gregg, and M.W. Spong, “A geometric approach tothree-dimensional hipped bipedal robotic walking,” in Proc. 45th Conf.Decision Control, San Diego, CA, USA, 2007.

[10] A. D. Ames, R. Vasudevan, and R. Bajcsy, “Human-data based costof bipedal robotic walking,” in Proc. Hybrid Syst.: Comput. Control,Chicago, IL, USA, 2011.

[11] I. Beloozerova and M. Sirota, “The role of the motor cortex in the con-trol of accuracy of locomotor movements in the cat,” J. Physiol (Lond),vol. 461, pp. 1–25, 1993.

[12] D. J. Braun andM. Goldfarb, “A control approach for actuated dynamicwalking in bipedal robots,” IEEE TRO, vol. 25, no. 6, pp. 1292–1303,Dec. 2009.

Page 16: IEEE TRANSACTIONS ON AUTOMATIC CONTROL, …edge.rit.edu/content/P15201/public/MSD II/Research Papers...IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 59, NO. 5, MAY 2014 1115 Human-Inspired

1130 IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 59, NO. 5, MAY 2014

[13] C. Chevallereau, G. Bessonnet, G. Abba, and Y. Aoustin, BipedalRobots: Modeling, Design and Walking Synthesis. New York, NY,USA: Wiley-ISTE, Jan. 2009.

[14] C. Chevallereau, A. Formal’sky, and D. Djoudi, “Tracking a joint pathfor the walk of an underactuated biped,” Robotica, vol. 22, no. 1, pp.15–28, Jan. 2004.

[15] S. Collins and A. Ruina, “A bipedal walking robot with efficient andhuman-like gait,” in Proc. IEEE IRCA’05, Apr. 2005, pp. 1983–1988.

[16] T. Drew, “Motor cortical activity during voluntary gait modificationsin the cat I. cells related to the forelimbs,” J. Neurophysiol., vol. 70,no. 1, pp. 179–199, 1993.

[17] B. Espiau and A. Goswami, “Compass gait revisited,” in IFAC Symp.Robot Control, Capri, Italy, Sep. 1994, pp. 839–846.

[18] A. Goswami, B. Thuilot, and B. Espiau, “A study of the passive gait ofa compass-like biped robot: Symmetry and chaos,” Int. J. Robot. Res.,vol. 17, no. 12, pp. 1282–1301, Dec. 1998.

[19] M. Goulding, “Circuits controlling vertebrate locomotion: Moving ina new direction,” Annu. Rev. Neurosci., vol. 10, no. 7, pp. 507–518,2009.

[20] J. W. Grizzle, G. Abba, and F. Plestan, “Asymptotically stable walkingfor biped robots: Analysis via systems with impulse effects,” IEEETrans. Autom. Conrtol, vol. 46, no. 1, pp. 51–64, Jan. 2001.

[21] J. W. Grizzle, C. Chevallereau, A. D. Ames, and R. W. Sinnet, “3Dbipedal robotic walking: Models, feedback control, and open prob-lems,” in Proc. IFAC Symp. Nonlinear Control Syst., Bologna, Italy,Sep. 2010.

[22] J. W. Grizzle, C. Chevallereau, and C. Shih, “Asymptotically stablewalking of a five-link underactuated 3-D bipedal robot,” IEEE Trans.Robotics, vol. 25, no. 1, pp. 37–50, Feb. 2009.

[23] M. O. Heller, G. Bergmann, G. Deuretzbacher, L. Dürselen, M. Pohl,L. Claes, N. P. Haas, and G. N. Duda, “Musculo-skeletal loading con-ditions at the hip during walking and stair climbing,” J. Biomechanics,vol. 34, pp. 883–893, 2001.

[24] A. Hereid, S. Kolathaya, M. S. Jones, J. V. Why, J. W. Hurst, and A. D.Ames, “Dynamic multi-domain bipedal walking with ATRIAS throughSLIP based human-inspired control,” Hybrid Syst.: Comput. Control,2014, to be published.

[25] P. Holmes, R. J. Full, D. Koditschek, and J. Guckenheimer, “Thedynamics of legged locomotion: Models, analyses, and challenges,”SIAM Rev., vol. 48, no. 2, pp. 207–304, May 2006.

[26] Y. Hürmüzlü and D. B. Marghitu, “Rigid body collions of planar kine-matic chains with multiple contact points,” Int. J. Robot. Res., vol. 13,no. 1, pp. 82–92, Feb. 1994.

[27] S. Kajita, F. Kanehiro, K. Kaneko, K. Fujiwara, K. Harada, K. Yokoi,and H. Hirukawa, “Biped walking pattern generation by using pre-view control of zero-moment point,” in Proc. IEEE—Int. Conf. Robot.Autom., Taipei, China, 2003, pp. 1620–1626.

[28] O. Kiehn, “Locomotor circuits in the mammalian spinal cord,” Annu.Rev. Neurosci., vol. 29, pp. 279–306, 2006.

[29] D. Koepl and J. W. Hurst, “Force control for planar spring-mass run-ning,” in Proc. IEEE Int. Conf. Intell. Robots Syst. (IROS), 2011, pp.3758–3763.

[30] A. D. Kuo, “Energetics of actively powered locomotion using the sim-plest walking model,” J. Biomech. Eng., vol. 124, pp. 113–120, 2002.

[31] W. Ma, H. Zhao, S. Kolathaya, and A. D. Ames, “Human-inspiredwalking via unified PD and impedance control,” in Proc. IEEE Int.Conf. Robot. Autom., 2014, submitted for publication.

[32] T. McGeer, “Passive dynamic walking,” Int. J. Robot. Res., vol. 9, no.2, pp. 62–82, Apr. 1990.

[33] B. Morris and J. Grizzle, “A restricted Poincaré map for determiningexponentially stable periodic orbits in systems with impulse effects:Application to bipedal robots,” in Proc. IEEE Conf. Decision Control,Seville, Spain, Dec. 2005, pp. 4199–4206.

[34] R. M. Murray, Z. Li, and S. S. Sastry, A Mathematical Introduction toRobotic Manipulation. Boca Raton, FL, USA: CRC, Mar. 1994.

[35] V. M. Zatsiorsky, Kinematics of Human Motion, 1st ed. Champaign,IL, USA: Human Kinetics, Sep. 1997.

[36] M. Pasupuleti, S. N. Yadukumar, and A. D. Ames, “Human-inspiredunderactuated bipedal robotic walking with AMBER on flat-ground,up-slope and uneven terrain,” in Proc. IEEE/RSJ Int. Conf. In-tell. Robots Syst. (IROS), Vilamoura, Algarve, Portugal, 2012, pp.2478–2483.

[37] J. Perry and J. Burnfield, Gait Analysis: Normal and PathologicalFunction, 2nd ed. Thorofare, NJ, USA: Slack Incorporated, 2010.

[38] I. Poulakakis and J. W. Grizzle, “The spring loaded inverted pendulumas the hybrid zero dynamics of an asymmetric hopper,” IEEE Trans.Autom. Control, vol. 54, no. 8, pp. 1779–1793, Aug. 2009.

[39] M. Powell, A. Hereid, and A. D. Ames, “Speed regulation in 3D roboticwalking throughmotion transitions between human-inspired partial hy-brid zero dynamics,” in Proc. IEEE Int. Conf. Robot. Autom., Karl-sruhe, Germany, 2013.

[40] M. J. Powell, H. Zhao, and A. D. Ames, “Motion primitives for human-inspired bipedal robotic locomotion: Walking and stair climbing,” inProc. IEEE Int. Conf. Robot. Autom., St. Paul, MN, USA, 2012, pp.543–549.

[41] J. Rose and J. G. Gamble, Human Walking. Philadelphia, PA, USA:Lippincott Williams & Wilkins, Dec. 2005.

[42] S. S. Sastry, Nonlinear Systems: Analysis, Stability and Control. NewYork, NY, USA: Springer, Jun. 1999.

[43] T. Schaub, M. Scheint, M. Sobotka, W. Seiberl, and M. Buss, “Effectsof compliant ankles on bipedal locomotion,” in IEEE/RSJ Int. Conf.Intell. Robots Syst., 2009, pp. 2761–2766.

[44] R. W. Sinnet and A. D. Ames, “3D bipedal walking with knees andfeet: A hybrid geometric approach,” in Proc. 48th IEEE Conf. DecisionControl, Shanghai, China, 2009, pp. 3208–3213.

[45] S. Jiang, S. Patrick, H. Zhao, and A. D. Ames, “Outputs of humanwalking for bipedal robotic controller design,” in Proc. Amer. ControlConf., 2012, pp. 4843–4848.

[46] R. W. Sinnet, M. J. Powell, S. Jiang, and A. D. Ames, “Compass gaitrevisited: A human data perspective with extensions to three dimen-sions,” in Proc. 50th IEEE Conf. Decision Control and Eur. ControlConf., 2011, 2013, pp. 682–689.

[47] M. W. Spong and F. Bullo, “Controlled symmetries and passivewalking,” IEEE Trans. Autom. Control, vol. 50, no. 7, pp. 1025–1031,Jul. 2005.

[48] S. Srinivasan, I. A. Raptis, and E. R. Westervelt, “Low-dimensionalsagittal plane model of normal human walking,” ASME J. Biomech.Eng., vol. 130, no. 5, pp. 051017-1–051017-11, Oct. 2008.

[49] S. Srinivasan, E. Westervelt, and A. Hansen, “A low-dimensionalsagittal-plane forward-dynamic model for asymmetric gait and itsapplication to study the gait of transtibial prosthesis users,” ASME J.Biomech. Eng., vol. 131, pp. 031003-1–031003-9, 2009.

[50] R. Vasudevan, A. D. Ames, and R. Bajcsy, “Using persistent homologyto determine a human-data based cost for bipedal walking,” in Proc.18th IFAC World Congr., Milan, Italy, 2011.

[51] M. Vukobratovi’c and B. Borovac, “Zero-moment point-thirty-fiveyears of its life,” Int. J. Humanoid Robot., vol. 1, no. 1, pp. 157–173,Mar. 2005.

[52] M. Vukobratovi’c, B. Borovac, D. Surla, and D. Stokic, Biped Loco-motion. Berlin, Germany: Springer-Verlag, Mar. 1990.

[53] E. Wendel and A. D. Ames, “Rank properties of Poincaré maps forhybrid systems with applications to bipedal walking,” in Proc. HybridSyst.: Comput. Control, Stockholm, Sweden, 2010.

[54] E. R. Westervelt, J. W. Grizzle, C. Chevallereau, J. H. Choi, and B.Morris, Feedback Control of Dynamic Bipedal Robot Locomotion.Boca Raton, FL, USA: CRC, Jun. 2007.

[55] E. R. Westervelt, J. W. Grizzle, and D. E. Koditschek, “Hybrid zerodynamics of planar biped walkers,” IEEE Trans. Autom. Control, vol.48, no. 1, pp. 42–56, Jan. 2003.

[56] D. A. Winter, Biomechanics and Motor Control of Human Movement,2nd ed. New York, NY, USA: Wiley-Interscience, May 1990.

[57] Y. Xiang, J. Arora, H.-J. Chung, H.-J. Kwon, S. Rahmatalla, R. Bhatt,and K. Abdel-Malek, “Predictive simulation of human walking tran-sitions using an optimization formulation,” Struct. MultidisciplinaryOptimiz., vol. 45, no. 5, pp. 759–772, 2012.

[58] S. N. Yadukumar, M. Pasupuleti, and A. D. Ames, “From formalmethods to algorithmic implementation of human inspired control onbipedal robots,” in Proc. 10th Int. Workshop Algorithmic Foundat.Robot. (WAFR), Boston, MA, USA, 2012.

[59] H. Zhao, W.Ma, M. Zeagler, and A. D. Ames, “Human-inspired multi-contact locomotion with AMBER2,” in Proc. Int. Conf. Cyber-Phys.Syst., 2014, to be published.

[60] H. Zhao, S. N. Yadukumar, and A. D. Ames, “Bipedal robotic runningwith partial hybrid zero dynamics and human-inspired optimization,”in Proc. IEEE/RSJ Int. Conf. Intell. Robots Syst. (IROS), Vilamoura,Algarve, Portugal, 2012.

Aaron D. Ames received the B.S. degree in mechan-ical engineering and the B.A. degree in mathematicsfrom the University of St. Thomas, St. Paul, MN,USA, in 2001, and the M.A. degree in mathematicsand the Ph.D. degree in electrical engineering andcomputer sciences from the University of California(UC), Berkeley, CA, USA, in 2006.He is an Assistant Professor in mechanical engi-

neering at Texas A&M University, College Station,TX, USA, with a joint appointment in electrical andcomputer engineering. His research interests center

on robotics, nonlinear control, hybrid and cyber-physical systems, with specialemphasis on foundational theory and experimental realization on bipedal robots.He served as a Postdoctoral Scholar in the Control and Dynamical System De-partment, California Institute of Technology, from 2006 to 2008.Dr. Ames was the recipient of the 2005 Leon O. Chua Award for achieve-

ment in nonlinear science and the 2006 Bernard Friedman Memorial Prize inApplied Mathematics while at UC Berkeley. In 2010, he received the NSF CA-REER award for his research on bipedal robotic walking and its applications toprosthetic devices.