Online Learning of Low Dimensional Strategies for High-Level Push Recovery in Bipedal Humanoid Robots

Similar documents
Practical Bipedal Walking Control on Uneven Terrain Using Surface Learning and Push Recovery

ZMP Trajectory Generation for Reduced Trunk Motions of Biped Robots

Body Stabilization of PDW toward Humanoid Walking

Stable Upright Walking and Running using a simple Pendulum based Control Scheme

Emergent walking stop using 3-D ZMP modification criteria map for humanoid robot

Dynamically stepping over large obstacle utilizing PSO optimization in the B4LC system

Gait Pattern Generation and Stabilization for Humanoid Robot Based on Coupled Oscillators

A Walking Pattern Generation Method for Humanoid robots using Least square method and Quartic polynomial

Humanoid Robots and biped locomotion. Contact: Egidio Falotico

Stability Control of Bipedal Walking Robot

Toward a Human-like Biped Robot with Compliant Legs

Walking Experiment of Biped Robot with Antagonistic Actuation Using Non-Linear Spring

Motion Control of a Bipedal Walking Robot

arxiv: v1 [cs.ro] 28 Feb 2017

Robots With Legs. Helge Wrede

DEVELOPMENT OF A FULL-SIZED BIPEDAL HUMANOID ROBOT UTILIZING SPRING ASSISTED PARALLEL FOUR-BAR LINKAGES WITH SYNCHRONIZED ACTUATION

Reactive Balance Control in Walking based on a Bipedal Linear Inverted Pendulum Model

YAN GU. Assistant Professor, University of Massachusetts Lowell. Frederick N. Andrews Fellowship, Graduate School, Purdue University ( )

Using sensory feedback to improve locomotion performance of the salamander robot in different environments

Toward a Human-like Biped Robot with Compliant Legs

Decentralized Autonomous Control of a Myriapod Locomotion Robot

A Bio-inspired Behavior Based Bipedal Locomotion Control B4LC Method for Bipedal Upslope Walking

Compliance Control for Biped Walking on Rough Terrain

Improvement of Humanoid Walking Control by Compensating Actuator Elasticity

Faster and Smoother Walking of Humanoid HRP-2 with Passive Toe Joints *

Study of Dynamic Biped Locomotion on Rugged Terrain - Derivation and Application of the Linear Inverted Pendulum Mode -

An Energy Efficient Dynamic Gait for a Nao Robot

Sample Solution for Problem 1.a

Slip Observer for Walking on a Low Friction Floor

Learning Energy Efficient Walking Based on Ballistics

Dynamic Lateral Stability for an Energy Efficient Gait

Capturability-based Analysis of Legged Robot with Consideration of Swing Legs

Gait Analysis of a Little Biped Robot. Received May 2015; accepted July 2015

OPTIMAL TRAJECTORY GENERATION OF COMPASS-GAIT BIPED BASED ON PASSIVE DYNAMIC WALKING

BIPED TRANSFORMER. Group No. 9

Walking Gait Generation with Foot Placement Control for the HOAP-3 Humanoid Robot

The Design of ZMP Measure System for Biped Robot LI Bin 1, a, LI Zhihai 2, b, LI Jiejia 1, c, BU Chunguang 2, d

Design, Fabrication and Analysis of Microcontroller Based Bipedal Walking Robot Vaidyanathan.V.T 1 and Sivaramakrishnan.R 2

Biomechanics and Models of Locomotion

Development of Fish type Robot based on the Analysis of Swimming Motion of Bluefin Tuna Comparison between Tuna-type Fin and Rectangular Fin -

Walking Control Algorithm of Biped Humanoid Robot on Uneven and Inclined Floor

CONFIRMATION OF PHD CANDIDATURE DAMIEN KEE 2003

Compliant Attitude Control and Stepping Strategy for Balance Recovery with the Humanoid COMAN

Journal of Chemical and Pharmaceutical Research, 2016, 8(6): Research Article. Walking Robot Stability Based on Inverted Pendulum Model

Biped locomotion on the Hoap2 robot

Velocity Based Stability Margins for Fast Bipedal Walking

arxiv: v1 [cs.ro] 27 Jul 2016

Using GPOPS-II to optimize sum of squared torques of a double pendulum as a prosthesis leg. Abstract

REPORT DOCUMENTATION PAGE

RUNNING ON SOFT GROUND: SIMPLE, ENERGY-OPTIMAL DISTURBANCE REJECTION

arxiv: v2 [cs.ro] 12 Jan 2017

Learning Energy Efficient Walking with Ballistic Walking

ZSTT Team Description Paper for Humanoid size League of Robocup 2017

Simulation of the Hybtor Robot

In memory of Dr. Kevin P. Granata, my graduate advisor, who was killed protecting others on the morning of April 16, 2007.

A Study on the Human Impulse Characteristics Of the Standing Shooting Posture

The Design and Control of a Bipedal Robot with Sensory Feedback

Controlling Walking Behavior of Passive Dynamic Walker utilizing Passive Joint Compliance

Effects of Ankle Stiffness on Gait Selection of Dynamic Bipedal Walking with Flat Feet

Omnidirectional Humanoid Balance Control: Multiple Strategies for Reacting to a Push

arxiv: v1 [cs.ro] 6 May 2018

Stabilizing Walking Gaits using Feedback from Gyroscopes

DETC DESIGN OPTIMIZATION OF A NOVEL TRIPEDAL LOCOMOTION ROBOT THROUGH SIMULATION AND EXPERIMENTS FOR A SINGLE STEP DYNAMIC GAIT

Trajectory Planning for Smooth Transition of a Biped Robot

Learning to Improve Capture Steps for Disturbance Rejection in Humanoid Soccer

Shuuji Kajita and Kazuo Tani

Foot Placement in the Simplest Slope Walker Reveals a Wide Range of Walking Solutions

Toward a human-like biped robot with compliant legs

Computer Aided Drafting, Design and Manufacturing Volume 26, Number 2, June 2016, Page 53. The design of exoskeleton lower limbs rehabilitation robot

INCLINOMETER DEVICE FOR SHIP STABILITY EVALUATION

Trajectory Planning and Motion Simulation for a Hydraulic Actuated Biped Robot

Emergency Stop Algorithm for Walking Humanoid Robots

Current issues regarding induced acceleration analysis of walking using the integration method to decompose the GRF

A NEW GOLF-SWING ROBOT MODEL UTILIZING SHAFT ELASTICITY

Development of a Locomotion and Balancing Strategy for Humanoid Robots

TEN YEARS IN LOCOMOTION CONTROL RESEARCH

Swing leg retraction helps biped walking stability

Programming Self-Recovery in the humanoid Leong Ti Xean 1 Yap Kian Tiong 2

Book Review. I. Background

An investigation of kinematic and kinetic variables for the description of prosthetic gait using the ENOCH system

Experimental Realization of Dynamic Walking for a Human-Riding Biped Robot, HUBO FX-1

Proof Copy. Controlling the Walking Period of a Pneumatic Muscle Walker. Takashi Takuma Koh Hosoda. Abstract. 1. Introduction

Effects of a Passive Dynamic Walker s Mechanical Parameters on Foot- Ground Clearance

AN ISOLATED SMALL WIND TURBINE EMULATOR

Designing falling motions for a humanoid soccer goalie

RECENTLY, various humanoid robots have been

Omnidirectional Capture Steps for Bipedal Walking


PERCEPTIVE ROBOT MOVING IN 3D WORLD. D.E- Okhotsimsky, A.K. Platonov USSR

Human Pose Tracking III: Dynamics. David Fleet University of Toronto

Biorobotic Locomotion: Biped Humanoid Walking. using Optimal Control

Speed Control System Design in Bicycle Robot by Low Power Method. Abstract

Vibration Analysis and Test of Backup Roll in Temper Mill

Online Learning of a Full Body Push Recovery Controller for Omnidirectional Walking

Stabilization of a Three-Dimensional Limit Cycle Walking Model through Step-to-Step Ankle Control

Adaptive Motor Patterns and Reflexes for Bipedal Locomotion on Rough Terrain*

Skippy: Reaching for the Performance Envelope

Sensing and Modeling of Terrain Features using Crawling Robots

Neuro-Fuzzy ZMP Control of a Biped Robot

Mutual and asynchronous anticipation and action in sports as globally competitive

Transcription:

2013 IEEE International Conference on Robotics and Automation (ICRA) Karlsruhe, Germany, May 6-10, 2013 Online Learning of Low Dimensional Strategies for High-Level Push Recovery in Bipedal Humanoid Robots Seung-Joon Yi, Byoung-Tak Zhang, Dennis Hong and Daniel D. Lee Abstract Bipedal humanoid robots will fall under unforeseen perturbations without active stabilization. Humans use dynamic full body behaviors in response to perturbations, and recent bipedal robot controllers for balancing are based upon human biomechanical responses. However these controllers rely on simplified physical models and accurate state information, making them less effective on physical robots in uncertain environments. In our previous work, we have proposed a hierarchical control architecture that learns from repeated trials to switch between low-level biomechanically-motivated strategies in response to perturbations. However in practice, it is hard to learn a complex strategy from limited number of trials available with physical robots. In this work, we focus on the very problem of efficiently learning the high-level push recovery strategy, using simulated models of the robot with different levels of abstraction, and finally the physical robot. From the state trajectory information generated using different models and a physical robot, we find a common low dimensional strategy for high level push recovery, which can be effectively learned in an online fashion from a small number of experimental trials on a physical robot. This learning approach is evaluated in physics-based simulations as well as on a small humanoid robot. Our results demonstrate how well this method stabilizes the robot during walking and whole body manipulation tasks. Keywords: humanoid robot, biomechanically motivated push recovery, low-dimensional policy, online learning I. INTRODUCTION Bipedal humanoid robots are unstable to external perturbations due to their small foot sizes and high center of mass. Perturbations caused by uneven terrain may be enough to cause the robots to fall during locomotion. On the other hand, humans are adept at walking even in the presence of perturbations. Biomechanical studies of human walking have shown that humans employ three basic balancing strategies, which are denoted as ankle, hip, and step strategies [1]. The ankle strategy modulates torque at the ankle joint to keep humans upright, the hip strategy converts unforeseen linear momentum into angular motion at the torso, and the step strategy moves the landing foot in the direction of the perturbation. Recent work has focused on using such biomechanically motivated push recovery strategies on humanoid robots [2], [3], [4]. These approaches utilize simplified physical models of the robots to derive basic analytical controllers, with GRASP Laboratory, University of Pennsylvania, Philadelphia, PA 19104 {yiseung,smcgill,ddlee}@seas.upenn.edu BI Laboratory, Seoul National University, Seoul, Korea. btzhang@ceas.snu.ac.kr RoMeLa Laboratory, Virginia Tech, Blacksburg, VA 24061 dhong@vt.edu the ability to potentially reject large perturbations compared to standard zero moment point (ZMP) feedback methods. However, the implementation of these controllers on physical robots has been limited due to their assumptions about perfect state estimation and unsophisticated models. In our previous work [5], we proposed an alternative approach. Rather than relying upon the accuracy of simplified models, the approach learns how to modulate the control parameters to account for the more complex dynamics of actual humanoid robots [6], [7]. The hierarchical control architecture learns to effective switch between low-level biomechanically-motivated controllers based upon noisy state information provided by the robot sensors. However, applying the approach to physical robot with scarce training data required arbitrary simplification of the controller[8]. In this paper, we investigate the decision boundary of the high level controllers using state trajectory data generated from simulated models of the robot with different levels of abstraction, as well as the physical robot. As a result, we have found that there exists a common low dimensional decision boundary for all the models which, although quite different from the theoretical one from previous approaches, can be learned effective in an online fashion even with limited training data acquired with physical robot. Using simulations and experiments on an actual DARwin-OP humanoid robot, we show how our method can effectively stabilize the robot against a wide range of perturbations during walking and whole body manipulation. The remainder of the paper is organized as follows. Section 2 reviews three biomechanically motivated push recovery controllers and their implementations for position controlled humanoid robots. Section 3 describes how to learn the high-level strategy from repeated trials in a simulated environment, and Section 4 shows the experimental results using the DARwIn-OP humanoid robot. Finally, we conclude with a discussion of outstanding issues and potential future directions arising from this work. II. BIOMECHANICALLY MOTIVATED HIERARCHICAL PUSH RECOVERY CONTROLLER In this section we first review three biomechanicallymotivated push recovery controllers assuming simplified physical models of the robot, and explain how they can be implemented on physical humanoid robots with position controlled actuators. Finally, we describe how a high level controller that switches between three strategies can be found from the analytical stability region of each controller. 978-1-4673-5642-8/13/$31.00 2013 IEEE 1641

(a) Ankle strategy Fig. 1. (b) Hip strategy Three biomechanically motivated push recovery strategies implemented on DARwIn-OP smalll humanoid robot and their abstract models. A. The ankle push recovery controller The ankle strategy applies control torque on the ankle joints to keep the center of mass within the base of support. We can assume the abstract model in figure 1 (a), where ankle torque τankle is applied to a linear inverted pendulum model (LIPM)[9] with mass m, center of mass (COM) height z0 and COM horizontal position x from current support point. Then the resulting linearized dynamic model is τankle x = ω 2 (x ), (1) p where ω = g/z0. It can be controlled by a PD-control on x with control gains K p and Kd τankle = K p x + Kd x. (2) This requires torque control of ankle actuators, but in practice it can be approximated for position controlled actuators with proportional feedback control by directly setting the target angle bias of the ankle actuator θankle [6], [10], θankle = K p0 x + Kd0 x, where K p0 (c) Step strategy and Kd0 (3) are control gains. B. The hip push recovery controller The hip strategy uses angular acceleration of the torso and limbs to generate a backward GRF to pull the COM back towards the base of support. Abstract model in figure 1 (b) includes a flywheel with point mass at height z0 and rotational inertia I, and control torque τhip at the COM. Then the resulting linearized dynamic model is τhip x = ω 2 (x ) (4) θ hip = τhip. I (5) However we should stop the flywheel from exceeding joint limits. In this case, following bang-bang profile [2] can be used for applying hip torque to maximize the effect while satisfying the joint angle constraint MAX τhip 0 t < TH1 τhip (t) = (6) MAX τhip TH1 t < 2TH1, MAX is the maximum torque that the can be applied where τhip on torso and TH1 is the time the torso stops accelerating. This torque profile makes the hip joint accelerate and then deccelerate with the maximum torque, stopping at the final MAX at t = 2T. After then, the hip angle bias position θhip H1 should return to zero [11]. This two-phase behavior can be approximated for position controlled actuators by controlling TARGET as the hip target angle bias θhip ( MAX θhip 0 t < 2TH1 TARGET θhip = 2TH1 +TH2 t MAX θhip 2TH1 t < 2TH1 + TH2, TH2 (7) where TH2 the duration of the returning phase. C. The step push recovery controller The step strategy moves the base of support towards the direction of push by taking a step, as shown in figure 1 (c). In practice, step strategy is handled by a locomotion controller that changes next foot landing position by xcapture from initial support point. In practice, it requires a reactive locomotion controller that can replan both COM and ZMP in real time, while satisfying kinematic and timing constraints. They are 1642

not covered in this paper as they are out of scope of this work. D. The high-level push recovery controller Humans perform a combination of push recovery behaviors according to the particular situation. We use a similar hierarchical control structure, where ankle, hip and step push recovery controllers work as low-level subcontrollers and the high-level push recovery controller triggers each according to the current sensory input. For abstract models we have seen in Fig. 1, there have been studies for analytic decision boundaries of each controller based on the current state [3]. If we assume maximum ankle torque as τmax ankle, then the stability region for ankle push recovery controller is derived as ẋ ω + x < τmax ankle, (8) and following stability region for the hip strategy plus the ankle strategy ẋ ω + x < τmax ankle + τmax hip (e ωt H1 1) 2. (9) Finally, if we assume instantaneous support point transition without loss of linear momentum, we have the following stability region for using all three strategies at once: ẋ ω + x < τmax ankle + τmax hip (e ωt H1 1) 2 + x MAX capture, (10) where x MAX capture is the maximum step size available. In this case we can use two boundary conditions in (8) and (9) to select between controllers based on current state. III. LEARNING THE HIGH-LEVEL PUSH RECOVERY STRATEGIES The analytical high level decision strategy introduced in previous section assumes simplified models and idealized sensing, so it can be less effective with physical robots in uncertain environments. In this section, we propose a learning approach, where the state trajectory information gathered during experimental trials are used to learn the high level strategy that better suits the actual dynamics of the robot. A. The extended inverted pendulum model Physical robots have feet with nonzero size, and joint torque is not directly controllable in most cases. Also, COM position x is hard to measure compared to torso tilt angle θ, which can be easily measured using inertial sensors. Based on these differences from simplified models, we propose a more realistic abstract model shown in Fig. 2 (a). It is an inverted pendulum with the tilt angle θ as state, and has a foot with toe position δ + and heel position δ from the ankle joint. The ankle torque τ ankle is controlled by a PD control of θ with saturation values δ + and δ. θ = ω 2 (sin(θ) τ ankle + τ hip ) (11) f sat (x) = τ ankle = f sat (K pθ + K d ) (12) δ + x δ + x δ + < x < δ + (13) δ x δ We can linearize the stability regions in (8), (9) and consider the saturated case to get the following approximate stability regions for the ankle, hip and step strategies δ < ω + θ < δ +, (14) τmax hip (e ωt H1 1) 2 ω + θ > δ ω + θ < δ + + τmax hip (e ωt H1 1) 2, τmax hip (e ωt H1 1) 2 ω + θ > δ ω + θ < δ + + τmax hip (e ωt H1 1) 2 xmax capture + xmax capture, (15) (16) Fig. 2 (b), (c) and (d) show three phase trajectory plots acquired using the extended inverted pendulum model and three different sets of push recovery strategies from various initial positions. Parameters used are m = 2, = 0.295, δ + = 0.05, δ = 0.05, K p = 500, K d = 57.83, τmax hip = 1, T H1 = 0.3, xcapture MAX = 0.08. We see that the hip and step strategies help to enlarge the stability region, and even with the non-linear dynamic model we use, the empirical stability regions closely follows the theoretical one derived using simpler models. B. The multibody model in physically-realistic simulated environment To model a physically realistic multi-body dynamics of the DARwIn-OP robot, we use the Webots commercial robotic simulator [12] based on the Open Dynamics Engine physics library. The controller update frequency and physics simulation frequency are both set to 100Hz. We use the COM height = 0.295, step duration t ST EP = 0.50 and robot center to ankle width d stance = 0.375 for walk parameters. For the ankle strategy gain parameters, we use values of K p = 0, = 0.15 which are found to be effective in practice. The robot is pushed with impulse forces for one time step (0.01s) with different magnitudes and directions, and initial phase space trajectories for 0.03 < t < 0.3 are logged and labeled as either stable or unstable according to their outcomes. Fig. 3 (a) shows the phase space trajectories for the ankle strategy. We can see that the empirical state trajectory plots have similar shapes to those in Fig. 2 (b), but the actual boundary between stable and unstable trajectories differs significantly from theoretical ones assuming simplified models. We use the online perceptron algorithm to learn the linear decision boundary from the labeled state trajectory data, and K d get following estimated values for ˆδ and zˆ 0 for each boundary for frontal pushes ˆ δ + = 0.45, ˆ z + 0 = 1.09, ˆ δ = 0.42, and following values for the lateral pushes ˆ δ + = 0.84, ˆ z + 0 = 1.45, ˆ δ = 0.84, = 1.02, (17) = 1.45, (18) 1643

(a) An inverted pendulum model of robot with position controlled ankle joint and foot. (b) Ankle strategy (c) Ankle plus hip strategy (d) Ankle plus step strategy Fig. 2. The extended inverted pendulum model and the phase space trajectory plots generated using different push recovery strategies. White and gray regions are theoretical stable and unstable regions from (14). Darker gray regions in (c) and (d) are the increased stable region compared to using the ankle strategy alone from (15) and (16). (a) Ankle strategy (b) Ankle plus hip strategy (c) Ankle plus step strategy Fig. 3. Phase space trajectory plots generated with the multi-body model and different push recovery strategies in physically realistic simulations. Frontal pushes are applied to standing robot. White and gray regions are theoretical stable and unstable regions from (14), (15) and (16). Thick dashed lines are learned linear boundaries between stable and unstable regions. which implies following empirical stability boundaries for the ankle strategy: ( g/ z ˆ δ 0 θ + ˆ ˆ ) ( < < g/ z ˆ+ δ 0 θ + ˆ+ˆ ). (19) z 0 Unlike abstarct models in Fig. 1, there are practical limits for θhip MAX and xcapture MAX due to kinematic and velocity constraints. We set x hip = 40, T H1 = 0.15, T H2 = 0.3 and x capture = 0.08 for hip and step strategy parameters and z + 0 compare the results of the two strategies. Fig. 3 (b) and (c) show trajectory plots acquired from two sets of push recovery controllers: the ankle-plus-hip strategy and the ankle-plusstep strategy. As the inertial sensor of our robot lies in the torso and rotates when the hip strategy is triggered, we could not get a clear boundary for ankle plus hip strategy as in Fig. 2 (c), Instead, we compared the outcome of two push recovery strategies against various magnitudes of perturbation to better compare the effectiveness of the two 1644

(a) (b) (c) (d) (e) (f) Fig. 4. A comparison of the ZMP tracking controller and suggested hierarchical push recovery controller with different impulse forces. (a) ZMP tracking controller, 1.05Ns of frontal push (b) Hierarchical push recovery controller, 1.05Ns of frontal push (c) ZMP tracking controller, 1.04Ns of backward push (d) Hierarchical push recovery controller, 1.04Ns of backward push (e) ZMP tracking controller, 1.61Ns of lateral push (f) Hierarchical push recovery controller, 1.61Ns of lateral push. controllers. We have found that against frontal push, the step strategy can withstand slightly larger maximum perturbations than the hip strategy, 104.3 Ns versus 105.0 Ns for the step strategy, and step strategy has a wider region of stability than the hip strategy with fixed parameter values θhip MAX and xcapture. MAX On the other hand, the step strategy is not available for purely lateral perturbation due to kinematic constraints and we have to rely on the hip strategy for such cases. We can summarize the high level push recovery strategy as follows. We set the ankle strategy active all the time, and if the state estimate moves beyond the empirical stability boundaries in (19), the step strategy is triggered. In case the step strategy is not available due to kinematic or timing constraints, the hip strategy is triggered instead. C. Comparison with ZMP tracking controller To demonstrate the effectiveness of the hierarchical push recovery controller, we compare it to the commonly used closed-loop ZMP tracking controller. We implement the ZMP tracking controller based on [13], with a single difference that the current current state is estimated using an inertial sensor rather than joint encoders and forward kinematics. All other parameters remain unchanged. Various amounts of frontal and lateral impulses were applied, and the outcome of push recovery effort is logged for each controller. Fig. 5. The comparison of stability regions of two controllers. Fig. 4 show a comparison of two approaches. From Fig. 5, we can see that the stability region of the suggested approach is approximately 21% larger and completely encompasses that of ZMP tracking method, which demonstrate the superiority of our approach. 1645

to the front with the default standing pose, eliminating the effect of backlash for frontal pushes. From the sets of state trajectories, we obtain the following linear boundaries with estimated values for for ˆδ and zˆ 0 for frontal pushes, ˆ δ + = 1.3, ˆ z + 0 = 2.7, ˆ δ = 1.1, = 2.7, (20) and following values for lateral pushes ˆ δ + = 1.4, ˆ z + 0 = 2.7, ˆ δ = 1.4, = 2.7. (21) (a) Frontal pushes (b) Lateral pushes Fig. 6. The phase space trajectory plots acquired from the experiments using DARwIn-OP robot. White and gray regions are theoretical stable and unstable regions from (14), and thick dashed lines are learned linear boundaries between stable and unstable regions. IV. EXPERIMENTAL RESULTS In addition to the simulated environment, we have implemented the hierarchical push recovery controller on a commercially available DARwIn-OP humanoid robot and have learned the high level control strategy from experience. Learned controller is used to stabilize the robot for manipulation task, where robot gets unknown amount of perturbation due to impacts. A. Learning the High level Control Strategy for Physical Robot We use the DARwIn-OP small humanoid robot as the testing platform. The DARwIn-OP robot is 45cm tall, weighs 2.8kg, and has 3-axis accelerometer and 3-axis gyroscope for inertial sensing. It has position-controlled Dynamixel servos for actuators, which are controlled by a custom microcontroller connected to an Intel Atom-based embedded PC at a control frequency of 100Hz. To generate repeatable external perturbations, a motorized moving platform was constructed using Dynamixel servomotors. The motorized platform is attached to the same bus connecting the actuators of the robot, and is controlled by the same I/O process. To generate maximum peak acceleration, the platform is slowly accelerated in one direction and then suddenly accelerated in the opposite direction. We applied various magnitudes of perturbations to the robot from the front, back, and one side while running the ankle strategy controller and measured the inertial sensor readings to generate state trajectories of robot. Fig. 6 (a) and (b) shows the state trajectories in the frontal and lateral axis, which are filtered with a moving average filter with n = 3. For the frontal pushes we can see the trajectory plot shown in top part of Fig. 6 (a) closely follows the graph acquired using simulated multi-body model in Fig. 3 (a), showing an almost linear boundary between stable and unstable trajectories, while the slope is quite different from theoretical one from LIPM shown in gray shade. However, for backward pushes, we see the shape of boundary is non-linear at the initial part of the trajectory. This is due to mechanical backlash of the joint, and it is only noticeable for backward pushes as the robot leans slightly B. Whole body impact task To evaluate the learned push recovery controller, we consider the situation where a humanoid robot punches another object. Robot gets various magnitude of perturbation according to the location and mass of the target object, which can make robot fall without proper stabilization. Punching motion is implemented by linearly interpolating three reference upper body keyframes. For stance parameters we use COM height = 0.295, ankle width d stance = 0.75 and body frontal tilt angle θ torso = 20. We use the ankle and the step strategy for push recovery, and decision boundaries of (19) with parameters (20) are used to select between the ankle and the step strategy based on current state estimated from inertial sensor readings. For push recovery controller parameters, we use values of K p = 0, K d = 0.15, xmax capture = 0.06 and step duration t ST EP = 0.35. To test different impact dynamics, we used a single target composed of a cardboard box with 2.8kg of weight attached at the bottom, and changed the distance between the robot and target. More details can be found in [14]. Fig. 7 summarizes the result of applying the impact motion controller for DARwIn-OP robot against targets with different distances. We can see that slight variance of the impact position is enough to make the robot fall down. On the other hand, the suggested push recovery controller can stabilize the robot against a wider range of perturbations from the impact. V. CONCLUSION In this paper, we investigate on the decision boundary of the high level push recovery controller that switches between low-level push recovery strategies based on current state information for three different simulated models of the robot and the physical DARwIn-OP small humanoid robot. We have found that the decision boundaries for more complex models are quite different from the one analytically derived from simplified model, which backs up our previous claim that learning approach is needed for physical robots, yet they are still low dimensional and can be effectively learned with limited training data in an online fashion. Using simulations and experiments on an actual DARwin-OP humanoid robot, we show how our method can effectively stabilize the robot against a wide range of perturbations during walking and whole body manipulation. Possible future work includes extending the push recovery controller to other humanoid robot tasks, and implementing these algorithms on full-size humanoid robots. 1646

(a) Without push recovery, target distance 25cm (d) With push recovery, target distance 25cm (b) Without push recovery, target distance 18cm (e) With push recovery, target distance 18cm (c) Without push recovery, no target Fig. 7. (f) With push recovery, no target Result of applying impact motion using a DARwIn-OP robot with and without push recovery control to targets with different distances. ACKNOWLEDGMENTS We acknowledge the support of the NSF PIRE program under contract OISE-0730206 and ONR SAFFIR program under contract N00014-11-1-0074. This work was also partially supported by the NRF grant of MEST (2012-0005643), as well as the BK21-IT program funded by MEST. R EFERENCES [1] Andreas G. Hofmann. Robust execution of bipedal walking tasks from biomechanical principles. PhD thesis, Cambridge, MA, USA, 2006. [2] Jerry Pratt, John Carff, and Sergey Drakunov. Capture point: A step toward humanoid push recovery. In in 6th IEEE-RAS International Conference on Humanoid Robots, pages 200 207, 2006. [3] B. Stephens. Humanoid push recovery. In IEEE-RAS International Conference on Humanoid Robots, 2007. [4] Bassam Jalgha, Daniel C. Asmar, and Imad Elhajj. A hybrid ankle/hip preemptive falling scheme for humanoid robots. In IEEE International Conference on Robotics and Automation, 2011. [5] Seung-Joon Yi, Byoung-Tak Zhang, Denis Hong, and Daniel D. Lee. Learning full body push recovery control for small humanoid robots. In IEEE International Conference on Robotics and Automation, 2011. [6] Baek-Kyu Cho, Sang-Sin Park, and Jun-Ho Oh. Stabilization of a hopping humanoid robot for a push. In IEEE-RAS International Conference on Humanoid Robots, pages 60 65, dec. 2010. [7] Mitsuharu Morisawa, Fumio Kanehiro, Kenji Kaneko, Nicolas Mansard, Joan Sol, Eiichi Yoshida, Kazuhito Yokoi, and Jean-Paul Laumond. Combining suppression of the disturbance and reactive stepping for recovering balance. In IEEE/RSJ International Conference on Intelligent Robots and Systems, pages 3150 3156. IEEE, 2010. [8] Seung-Joon Yi, Byoung-Tak Zhang, Dennis Hong, and Daniel D. Lee. Online learning of a full body push recovery controller for omnidirectional walking. In IEEE-RAS International Conference on Humanoid Robots, pages 1 6, 2011. [9] S. Kajita and K. Tani. Experimental study of biped dynamic walking in the linear inverted pendulum mode. In Robotics and Automation, 1995. Proceedings., 1995 IEEE International Conference on, volume 3, pages 2885 2891 vol.3, may 1995. [10] Inyong Ha, Yusuke Tamura, and Hajime Asama. Gait pattern generation and stabilization for humanoid robot based on coupled oscillators. In IEEE/RSJ International Conference on Intelligent Robots and Systems, pages 3207 3212, 2011. [11] Muhammad Abdallah and Ambarish Goswami. A biomechanically motivated two-phase strategy for biped upright balance control. In IEEE International Conference on Robotics and Automation, pages 2008 2013, 2005. [12] O. Michel. Webots: Professional mobile robot simulation. Journal of Advanced Robotics Systems, 1(1):39 42, 2004. [13] D. Gouaillier, C. Collette, and C. Kilner. Omni-directional closed-loop walk for nao. In IEEE-RAS International Conference on Humanoid Robots, pages 448 454, dec. 2010. [14] Seung-Joon Yi, Byoung-Tak Zhang, Denis Hong, and Daniel D. Lee. Active stabilization of a humanoid robot for impact motions with unknown reaction forces. In IROS, 2012. 1647