You are on page 1of 15

Compliant Leg Shape, Reduced-Order Models

and Dynamic Running

Jae Yun Jun, Duncan Haldane, and Jonathan E. Clark

Abstract. The groundbreaking running performances of RHex-like robots are ana-


lyzed from the perspective of their leg designs. In particular, two-segment-leg mod-
els are used both for studying the running with the legs currently employed and for
suggesting new leg designs that could improve the gait stability, running efficiency
and forward speed. New curved compliant monolithic legs are fabricated from these
models, and the running with these legs is tested by using a newly designed running
test robot. Both the simulations and the experimental trials seem to suggest that run-
ning with legs with unity-ratio of the leg segments is faster and more efficient than
running with the leg that is currently used on the RHex-like robots. The simulation
model predictions seem to match closely to experimental trials in some instances
but not always. In the future, a more sophisticated model is needed to capture the
actual running with curved legs more accurately.

1 Introduction
Animals have segmented legs with various geometric configurations depending on
their scale and the habitat in which they live. Large-sized mammalian animals have
upright posture to reduce the bending moments acting about their leg joints. Small
animals have more bent legs (crouched legs) because acceleration is crucial for their
survival even though this type of leg produces reduced effective mechanical advan-
tage [2]. Despite the rich biological precedent available, most analytical studies on
running have used the prismatic legs assumed in the Spring-Loaded Inverted Pen-
dulum (SLIP) model of running [1]. Only a few studies have been published that
represent running behavior by means of segmented-leg models [11]. Blickhan et al.
showed that the geometric configuration of a segmented leg influences the demand
for energy production, structural stability and velocity transmission from muscle
groups to the leg tip [2]. Rummel and Seyfarth showed that the two-segment-leg
model offers a much larger stability region (i.e., the range of the leg impact angles
that guarantee stable running) than the standard SLIP model could [11].
Jae Yun Jun · Duncan Haldane · Jonathan E. Clark
Dept. Mechanical Engineering, Florida State University, 2525 Pottsdamer Street Rm A229,
Tallahassee, Florida 32310
e-mail: {jaeyun,haldadu,clarkj}@eng.fsu.edu

O. Khatib, V. Kumar, and G. Sukhatme (eds.), Experimental Robotics, 759


Springer Tracts in Advanced Robotics 79,
DOI: 10.1007/978-3-642-28572-1_52,  c Springer-Verlag Berlin Heidelberg 2014
760 J.Y. Jun, D. Haldane, and J.E. Clark

Despite the apparent advantage of a more complex, articulated leg geometry,


most successful dynamic running robots have utilized simple, robust leg designs.
At the logical extreme of this approach is the hexapedal robot RHex which utilizes
compliant monolithic legs driven by a single actuator to run quickly over rough
terrains [12].
At the other extreme BigDog [10], Kenken [5] and Rush [14] are some of the
most successful running robots with articulated legs, however these robots have not
yet been able to best the speed over smooth terrain of iSprawl [8] or the ability to
quickly traverse rough terrain of RHex [12].
In this paper we investigate whether simple monolithic legs can be designed to
employ the dynamic advantages of more complex segmented legs for this type of
highly mobile platform. In particular, we consider the compliant curved legs utilized
by RHex, and argue that this type of leg can be viewed as a two-segment-leg with
its intersegmental joint placed at the flexion point on the leg. We investigate the
relationships between the parameters that define a leg segmentation and running
performance, so that these relationships can serve as a guide for designing new legs
or improving currently available leg designs.
First, we simulate the passive (conservative) dynamics of running with these legs
in a reduced-order model. These parameters are tuned to optimize the leg design
for running performances. Next, a more realistic reduced-order model is introduced
including energy dissipation and leg actuation. For both the original and the new
leg design parameters, the stability and the efficiency are optimized over the con-
troller parameter space by using a direct search method (Nelder-Mead algorithm)
to determine if the advantages found in the underlying (conservative) dynamics are
evidenced in the actuated system. Finally, a new set of compliant legs are designed
and manufactured based on the simulation results. A new bipedal test robot is built
and used to experimentally analyze the running dynamics with the original and the
new legs. The simulation-to-experiment and experiment-to-experiment results are
then compared.

2 Two-Segment-Leg Model in the Conservative System


2.1 Dynamical Model
Although the half circle legs utilized by RHex and Edubot were initially envisioned
and modeled as a prismatic spring as shown in Fig. 1(b), more recent studies [4]
have shown that a two-segment-leg model as shown in Figure 1(b) more accurately
models the compliant behavior of the leg. In this model, the body is represented by
a point mass, and the leg has two massless segments with length l1 and l2 for the
distal (shank) and proximal (thigh) segments, respectively. A pin joint contact model
is used to model the foot contact with the ground where no foot slip is allowed.
The motion of the body is constrained to be in the sagittal plane only. During its
locomotion, the energy is stored and returned to the system by using a torsional
spring (with constant stiffness K) at the joint of the two segments. When the leg is
not loaded, the effective leg length (l) and the intersegmental leg angle (β ) are at
rest with effective leg rest length (lo ) and leg rest angle (βo ), respectively.
Compliant Leg Shape, Reduced-Order Models and Dynamic Running 761

m
l2
K −θ
β
l
ψ γ
y l1 F
leg
x α

(a) (b)
Fig. 1 (a) Half-circle leg employed on EduBot[9], a RHex-like robot. (b) Two-segment-leg
model to approximate the dynamics of running with a half-circle leg. The parameters and the
variables of the model are explained in the text.

Each stride is comprised of one stance phase and one flight phase. The stance
phase is described by two dimensional generalized coordinates, q = [ψ , θ ]T , and
 T
generalized speeds, q̇ = ψ̇ , θ̇ (see Figure 1(b)). The equation of motion for stance
phase is
M(q)q̈ + N(q, q̇) + G(q) = 0 (1)
where M is the 2×2 inertia matrix, N is the 2×1 Coriolis force vector, and G is a 2×
1 gravitational force vector. All angles are defined to be positive counterclockwisely.
The motion of the Center of Mass(CoM) during flight phase is simulated by using
a simple ballistic model. The transition from stance phase to flight phase occurs
when the condition y ≤ lo sin α is triggered, where α is some predefined leg impact
angle with respect to the horizontal line, and y is the height of the CoM from the
ground. The transition from flight phase to stance phase occurs when the condition
l > lo is satisfied.
As shown by Rummel and Seyfarth in [11], we relate the leg torsional stiffness
to a reference stiffness defined as k10% = F10% /Δ l10% , where Δ l10% is a reference
leg compression at 10% of the effective leg rest length (i.e. 0.1lo), and F10% is the
leg force at the reference leg compression. Dimensionless reference stiffness (k̃10% )
are used in the rest of the presented work where k̃10% is defined as k10% lo /mg. The
relationship between the reference stiffness (k10% ) and the torsional stiffness (K) is
shown in [11].
The leg ground reaction force is directed from the leg contact point with the
ground to the center of mass (CoM), and its magnitude has the following expression

l K(βo − β )
Fleg = (2)
l1 l2 sin β

where l1 , l2 , l, K, βo , β are the length of the distal leg segment, the length of the
proximal leg segment, the effective leg length, the leg’s torsional stiffness, the leg’s
intersegmental rest angle, and the leg’s intersegmental angle, respectively. From
762 J.Y. Jun, D. Haldane, and J.E. Clark

the leg force calculation, the horizontal and vertical ground reaction forces can be
expressed as
Fx = −Fleg sin γ
(3)
Fy = Fleg cos γ

2.2 Analysis of Periodicity and Stability of Running


Let q be the state vector that describes the dynamical system approximated by the
two-segment-leg model. Taking the Poincaré section at apex, let qi be the state
vector when the center of mass (CoM) is at apex within the i-th stride. Let R
be the Poincaré map function defined as qi+1 = R(qi ). Then, when the condition
q∗ = qi+1 = qi is satisfied, q∗ is a fixed point (also defined as a periodic orbit in
the phase portrait). For the case of the two-segment-leg model in the conservative
system, a fixed point is searched by letting the body run up to N = 50 strides. If
within N strides, the norm-2 of the difference between the states corresponding to
the two consecutive apex events is less than a predefined tolerance value, then the
particular gait is considered as a periodic gait (i.e. a fixed point is found).
Once a fixed point is found, its stability is studied by analyzing the eigenvalues of
the Jacobian matrix of the linearized Poincaré map around the fixed point. A fixed
point (i.e. the associated periodic gait) is asymptotically stable or neutrally stable or
unstable, if all eigenvalues are less than unity or if one or more eigenvalues are unity
(and the rest being less than unity) or if there exists at least one eigenvalue larger
than unity, respectively.

2.3 Results and Discussion


The curved compliant leg currently used on the bipedal running robot is charac-
terized empirically by identifying its corresponding design parameters of the two-
segment-leg model as follows: βo = 90o , l2 /l1 = 0.35 and k̃10% = 12.46.
In order to search for better leg designs by using the two-segment-leg model,
various values of βo , l2 /l1 and k̃10% are considered:
• l2 /l1 ∈ {0.15, 0.35, 0.55, 0.75, 1, 1/0.75, 1/0.55, 1/0.35, 1/0.15}
• βo ∈ [90o , 170o ] with granularity of 10o
• k̃10% ∈ {12.46, 21.5, 28.5, 38.5}
Figure 2 shows four surfaces representing the size of the range of the leg impact
angles (α ) for which stable running gaits exist for the four leg reference stiffness
values considered. The 2D-domain on which each surface is depicted is formed by
the leg segments’ ratios (i.e. the ratio of the length of the leg proximal segment (l2 ) to
the length of the leg distal segment (l1 )) and the leg intersegmental rest angles (βo ).
For each triplet formed by leg segments’ ratio, leg intersegmental rest angle and
leg stiffness, leg impact angles (α ) that range from 30o to 90o are considered with
precision of 0.1o . For all running simulations, the same initial state values are em-
apex apex apex
ployed: (vox , yo ) = (2.5m/s, lo ), where vox is the horizontal forward speed of
apex
the CoM at apex, and yo is the height of the CoM from the ground at apex.
Compliant Leg Shape, Reduced-Order Models and Dynamic Running 763

Δα ΀ĚĞŐ΁ Δα ΀ĚĞŐ΁

βŽ ΀ĚĞŐ΁

βŽ ΀ĚĞŐ΁
;ĂͿ ;ďͿ
Δα ΀ĚĞŐ΁ Δα ΀ĚĞŐ΁
βŽ ΀ĚĞŐ΁

βŽ ΀ĚĞŐ΁

;ĐͿ ;ĚͿ
Fig. 2 Range of the leg’s impact angles of the stable running gaits (in degrees) for various
ratios of the length of the leg’s proximal segment (l2 ) to the length of the leg’s distal segment
(l1 ) and for various leg’s intersegmental rest angles (βo ). (a) for k̃10% = 12.46, (b) for k̃10% =
21.50, (c) for k̃10% = 28.50, and (d) for k̃10% = 38.50.

From Figure 2, one can observe that the size of the range of the stable leg impact
angles is largest when l2 /l1 = 1 for all considered leg stiffness values and leg inter-
segmental rest angles. Three aspects can be observed as the leg stiffness is increased.
First, the stability region expands in both directions of the 2D-domain formed by the
ratio of the lengths of the leg segments and the leg intersegmental rest angle with the
leg stiffness increase. Second, the size of the stability region increases from about
9o for k̃10% = 12.46 to about 29o for k̃10% = 38.50. Third, the leg intersegmental rest
angle that corresponds to the maximum size of the stability region shifts from about
90o to 170o as the leg stiffness value is increased from k̃10% = 12.46 to k̃10% = 38.50.
Some preliminary conclusions could be stated from the obtained results. First, it
is desirable to design legs with the same lengths of the proximal and distal segments.
This suggests that the current curved leg design is not optimum in terms of gait
stability, suggesting the first leg design change: altering the l2 /l1 ratio from about
0.35 in the current design to l2 /l1 = 1.
Second, for a given leg stiffness, there exists a leg intersegmental rest angle βo for
which the size of the range of the stable leg impact angles is maximized. Third, the
size of the stability increases with the leg intersegmental rest angle when l2 /l1 =
1 and when an appropriate leg stiffness value is chosen. This suggests the second
design change, increasing the leg stiffness k̃10% , and increasing βo =90o to βo =150o.
764 J.Y. Jun, D. Haldane, and J.E. Clark

3 Two-Segment-Leg Model in an Energy-Dissipative System


3.1 Dynamical Model
The running behavior of RHex, Edubot, and the new bipedal running robot (Figure
3(b)) is modeled by adding a torque actuator at the hip and a torsional damper at the
leg’s intersegmental joint to the two-segment-leg model of the energy-conservative
system, as shown in Figure 3(a). The torsional damper emulates the energetic losses
during running, and the torque actuator does work to compensate these losses. A
simple motor model is incorporated in the simulation with the purpose of captur-
ing basic aspects of the limitation of the motors employed on the bipedal running
robot. This model consists of a linear speed-torque curve with parameters specified
in Section 4.1.

τ Hip
m
l2
−θ γ l
K, B β F
leg
ψ −
τ Hip
LJ l1 l

dž
(a) (b)
Fig. 3 (a) Torque-driven and damped two-segment-leg model. A torque actuator (τHip ) is
added at the hip to model the motor that actuates half-circle leg, and a torsional damper (B) is
incorporated at the intersegmental joint to model the energetic losses of the leg. (b) Bipedal
test robot with curved legs is attached to a boom located at the center of the track.

The resulting equations of motion are similar to Eq. 1 with the difference of the
addition of the torque actuator (τHip ) and the torsional damper (B). The states in the
energy-dissipative system are the same as those of the conservative system.
The new expression for the leg force is given in Eq. (4).

l K(βo − β ) − Bβ̇
Fleg = (4)
l1 l2 sin β
And the expressions for the horizontal and vertical ground reaction forces are given
in Eq. (5)
τ
Fx = −Fleg sin γ − Hip cos γ
τHipl (5)
Fy = Fleg cos γ − l sin γ
where Fleg is shown in Eq. 4, γ is the angle between the effective leg and the vertical
line, and τHip is the amount of torque applied by the torsional actuator located at the
hip.
Compliant Leg Shape, Reduced-Order Models and Dynamic Running 765

3.2 Controller
A periodic function known as Buehler Clock [12] (used by RHex-like robots) is
considered as a reference signal for controlling the leg motion. This function is a
periodic two-piecewise linear function with one steeper slope for the fast swing
phase and one flatter slope for the slow stance phase. Four parameters characterize
this periodic function: desired leg touch-down angle (γA ), desired leg lift-off angle,
(γB ), duration of the stance phase (tB ), and the stride period (T ) (see [6] for details).
The control input signal is generated by using a simple proportional-derivative (PD)
controller:
τHip = kp (γdes − γ ) + kd(γ̇des − γ̇ ) (6)
where kp and kd are the proportional and the derivative gains, γdes and γ̇des are the
desired leg angular position and speed dictated by the reference signal, and γ and γ̇
are the actual angular position and speed of the effective leg, respectively.

3.3 Optimizer
Each of the three considered legs (legs with (βo = 90o , l2 /l1 = 0.35), (βo = 90o ,
l2 /l1 = 1) and (βo = 150o , l2 /l1 = 1)) requires distinct controller parameter settings
for stable running, and these parameters need to be optimized independently, done
here by using a direct search method (Nelder-Mead algorithm), in order to compare
their respective running performances. Two optimization cost functions have been
considered: stability and efficiency. For stability, the norm-2 of the maximum eigen-
value of the Jacobian matrix of the linearized Poincaré map is used as the stability
measure (see Section 2.2). For efficiency, specific resistance is used. Specific resis-
tance is defined as P/mgv, where P is the average mechanical power expenditure
within a stride, m is the body mass, g is the gravitational acceleration, and v is the
average forward running speed over a stride.
The results of this optimization process are described and compared to the exper-
imental results in Section 4.3.

4 Experiments of Running with Curved Legs


4.1 Description of Bipedal Running Robot and Test-Bed
A bipedal running robot (shown in Figure 3(b)) was built to test the running perfor-
mances with curved legs. The robot consists of an aluminum frame with two motors
(Maxon RE-Max 222049 11W brushed DC motors with Maxon 143978 24:1 plan-
etary gearhead) that actuate a pair of curved legs. The robot is attached to a boom
which in turn is anchored to the center of the track on which the robot runs. The
dimensions of the robot and the boom are specified in Table 2. Foot slippage during
running is minimized by padding the track with a carpet and also by attaching a
bicycle tread on the portion of the curved leg that contacts with the track.
766 J.Y. Jun, D. Haldane, and J.E. Clark

Three sets of legs are fabricated based on the simulation results and are listed in
Table 1.

Table 1 Types of curved legs employed for running tests

Leg βo [deg] l2 /l1 [−] k̃10% k[N/m]


A 90 0.35 12.46 1,100
B 90 1 12.46 1,100
C 150 1 28.50 2,516

A Gumstix Verdex-Pro with 400MHz Intel PXA270 microprocessor with an I/O


board(Breakout-vx) controls the running behavior of the biped. The angular posi-
tions of the motors are controlled via a 1kHz PD control loop. Two motor encoders
measure the angular positions of the legs, and another two encoders attached to the
boom measure the vertical and horizontal displacements of the biped. Ground reac-
tion forces are measured via a 6-axis force plate (AMTI Inc.) embedded in the track.
An electric current sensor measures the current drawn by motors during running tri-
als, and these values together with the commanded voltage values are employed to
estimate the electric power expenditure.

Table 2 Specifications for bipedal running robot and for test-bed

Parameters Symbols Values Units


Body mass mbody 1.026 kg
Body dimension W × L × H 220 × 240 × 55 mm × mm × mm
Leg mass mleg 0.045 kg
Leg length (hip to toe) lo 0.114 m
Motor stall torque τs 0.0569 Nm
Motor no load speed ωnls 789.6 rad/s
Gear ratio Gr 24 : 1 -
Boom length Lboom 1.5 m

4.2 Description of the Fabrication of Monolithic Compliant Legs


As shown in Figure 4, three monolithic compliant leg designs were fabricated using
the Vacuum Assisted Resin Transfer (VARTM) process. This method creates very
light and strong composites and allows extensive customization of the composite
layup. When using the VARTM process, dry reinforcing fibers are laid onto a mold
using a small amount of spray adhesive to maintain fiber alignment during process-
ing. After the entire dry fiber layup has been applied to the mold, a layer of nylon
peelply is applied, followed by a layer of infusion media. A vacuum is then drawn
on the composite layup and the fibers are then infused with resin.
Compliant Leg Shape, Reduced-Order Models and Dynamic Running 767

Vinyl Ester resin with 2% MEKP activator was used to create the composite
legs. This styrene based resin offers more control over composite compliance than
traditional epoxy systems. By post-curing the styrene based composite, stiffness can
be controllably increased.
Two types of fiberglass were used in the reinforcing matrix in the compliant legs.
A very light 6781 weave comprised of S2 glass was applied in alternating layers
with a heavier S-glass double bias weave. A portion of the composite needed to be
rigid to modify the location of the characteristic pivot. To create a light weight, rigid
composite section, a coring material called Soric was used. Soric is comprised of
small hexagonal foam sections interspersed with a light binding matrix.
The original legs (leg A, Figure 4(a)) are comprised of three layers of S2-6781
alternated with three layers of double bias for a total of six layers. The resulting
composite was post cured at 80◦ C for thirty minutes. The leg B (Figure 4(b)) is
comprised of three layers of S2-6781 and two layers of double bias. Additionally,
stiffening Soric was applied to the layup to shift the characteristic pivot to the mid-
dle of the leg. The resulting composite was post cured at 80◦C for fifteen minutes.
Finally, the leg C (Figure 4(c)) was fabricated in multiple sections: a rigid attach-
ment to the motor mount, a flexural segment and a rigid rolling contact adapter.
These sections were then bonded together using a methacrylate adhesive. The flex-
ural segment was fabricated separately to avoid the mechanical imperfections that
can result when pieces with complex geometry are VARTM processed. The flexural
section is comprised of four layers of 3K T-300 carbon fiber applied in alternating
45◦ layers. The rigid motor mount attachment section is made by reinforcing the
flexural layup with Soric. The rolling contact adapter is comprised of ten layers of
3K T-300 carbon fiber. After the legs have been fabricated, a section of rubber tread
is applied to increase traction.
Both the original leg (leg A) and leg B have a stiffness of 1,100 N/m (k̃10% =
12.46), leg C has a stiffness of 2,516 N/m (k̃10% = 28.50). The stiffness of the legs
was verified using a customized MTS Insight material testing machine. A leg mount
is attached to the load cell; the other end of the leg is placed on a linear slide rail
to allow rolling contact. The MTS system compresses the leg by 10% of the rest

(a) (b) (c)


Fig. 4 Curved compliant monolithic legs with: (a) βo = 90 , l2 /l1 = 0.35, k̃10% = 12.46 (leg
o

A) (b) βo = 90o , l2 /l1 = 1, k̃10% = 12.46 (leg B) (c) βo = 150o , l2 /l1 = 1, k̃10% = 28.50
(leg C).
768 J.Y. Jun, D. Haldane, and J.E. Clark

length, or 1.14 cm, and the (k̃10% )) value is calculated to verify that the leg is the
correct stiffness.

4.3 Experimental Results


The controller parameters optimized in simulation for all three types of legs are
tested on the robot. First, the controller parameters optimized for gait stability are
employed on the robot in order to validate the simulation model described in Sec-
tion 3. The running performances obtained by using these controller parameters in
simulation level are shown in Table 3.

Table 3 Simulation results by running with the controller parameters optimized for gait
stability
Study Leg |λmax | vx [m/s] Mechanical Power [W] SR
1 A 0.1762 1.1603 8.0460 0.6890
2 B 0.3949 1.7947 6.6977 0.3814
3 C 0.0723 0.9899 5.4196 0.5440

Table 3 shows that the running with leg C is the most stable (lowest |λmax |), and
running with leg B is fastest (highest vx ) and most efficient (lowest specific resis-
tance). Figure 5 and Figure 6 show the simulation-to-experiment comparison re-
sults for running with leg A and leg B, respectively. Repeatable running experiments
could not be performed with leg C because the leg repeatedly fractured at the flexion
point of the leg. To overcome this, new leg design is required in the future. In addi-
tion, the PD controller used on the robot operates in voltage mode, and, therefore,
the gains employed for experimental running trials needed to be modified manually
because the gains optimized in the simulation are for torque mode (see Eq. 6).
Figure 5 and Figure 6 show six graphs respectively: (a) height of CoM (ex-
periment and simulation); (b) electric current drawn by right motor (experiment);
(c) ground reaction forces (experiment); (d) forward speed of CoM (experiment
and simulation); (e) hip torque applied for right leg motion (simulation); and (f)
ground reaction forces (simulation). The horizontal axis of all six graphs repre-
sents time normalized by the respective stride period (T = 0.2840s for Study 1 and
T = 0.2780s for Study 2).
For Figures 5(a), 5(d), 6(a) and 6(d), the normalized time 0 (or 1) corresponds
to the instant of time when leg maximum compression occurs. On the other hand,
for Figures 5(b), 5(c), 5(e), 5(f), 6(b), 6(c), 6(e) and 6(f), the normalized time 0 (or
1) corresponds to the instant of time when leg touch-down occurs. The figures are
represented in this way because the data collected from the current sensor and the
force sensor are not synchronized with the data collected from the boom encoders.
The time normalization process for the data collected from the boom encoders is
possible by detecting the instant of time when the height of CoM is the lowest. For
the data collected from the current sensor and the force sensor, this process is easier
by detecting the instant of time when the leg touches the ground.
Compliant Leg Shape, Reduced-Order Models and Dynamic Running 769

^ƚƵĚLJϭ;>ĞŐͿ
^ŝŵƵůĂƚŝŽŶ
džƉĞƌŝŵĞŶƚ s'Z&
,'Z&

;ĂͿ ;ďͿ ;ĐͿ

^ŝŵƵůĂƚŝŽŶ
džƉĞƌŝŵĞŶƚ s'Z&
,'Z&

;ĚͿ ;ĞͿ ;ĨͿ

Fig. 5 Simulation-to-experiment comparison results for leg A (Study 1 shown in Table 3).
The horizontal axis of all the graphs is time normalized by the stride period (T = 0.2840s).

^ƚƵĚLJϮ;>ĞŐͿ
^ŝŵƵůĂƚŝŽŶ
džƉĞƌŝŵĞŶƚ s'Z&
,'Z&

;ĂͿ ;ďͿ ;ĐͿ

s'Z&
,'Z&

^ŝŵƵůĂƚŝŽŶ
džƉĞƌŝŵĞŶƚ

;ĚͿ ;ĞͿ ;ĨͿ

Fig. 6 Simulation-to-experiment comparison results for leg B (Study 2 shown in Table 3).
The horizontal axis of all the graphs is time normalized by the stride period (T = 0.2780s).
770 J.Y. Jun, D. Haldane, and J.E. Clark

The running with leg A for Study 1 is experimentally observed to be very sta-
ble and repeatable. The height fluctuations of the CoM observed during experimen-
tal running match closely to those predicted by the simulation model (Figure 5(a)).
However, the forward velocity of the robot deviates substantially from the simula-
tion prediction (Figure 5(d)). It appears that this discrepancy occurs mainly because
the simulation model does not capture the foot slippage that occurs in actual running
experiments. Figure 5(b) shows that during about 30% of the stride period the right
motor does negative work. Afterwards, the electric current draw grows positively in
order to correct the error produced between the desired and actual leg trajectories.
As the leg moves towards the instant of the lift-off events, the drawn current gets
smaller because the error signal is reduced and reaches its minimum value during
flight phase. Figure 5(e) shows the amount of applied torque for the right leg motion
in simulation level. Mechanical power is considered in simulation, but in experi-
ments electric power is used instead. The amount of applied torque can be related to
the electric current drawn by the motor using motor constant specified by the motor
manufacturer. Figure 5(b) and Figure 5(e) show similar trends. Note that in simula-
tion no torque is applied during flight phase because the leg is considered massless
and is assumed to track the desired leg trajectory perfectly. Figure 5(c) and Figure
5(f) show that the peak measured vertical ground reaction force is much lower than
predicted from the simulation. One source of this difference is the rolling contact
motion caused by running with curved legs which the two-segment-leg model does
not capture (because the model assumes pin joint contact model). Rolling contact
motion during stance phase has been shown to reduce the peak vertical ground re-
action forces, and the larger the radius of the rolling foot curvature, the smaller the
peak vertical ground reaction force is obtained for both walking [13] and running [7].

Table 4 Experimental running performances for Study 1 and Study 2

Study Leg v̄x [m/s] P̄[W] SR


1 A 0.8421 8.9768 1.0590
2 B 1.3353 7.6047 0.5658

Similar results are shown running with leg B for Study 2 (Figure 6). The average
forward speed for this study is larger than for Study 1 as predicted by the simula-
tion model (Table 3). The running motion with the optimal controller parameters
obtained for Study 2, however, was not as stable as running with Study 1 as one can
observe from Figure 5(a) and Figure 5(d).
Specific resistance values for Study 1 and 2 are estimated from the averaged
forward speeds and the estimated average electric powers shown in Table 4. These
results show that running with leg B is faster and more efficient than running with
leg A for these particular studies, and this result matches to the simulation results
shown in Table 3.
When the optimization cost function is changed from the gait stability measure
(|λmax |) to specific resistance (SR), neither leg A nor leg B could run stably with the
Compliant Leg Shape, Reduced-Order Models and Dynamic Running 771

optimized controller parameters in simulation. A primary reason why these gaits


were not realizable on the robotic running may be because the stride frequency is
very high reaching almost the no-load-speed during flight phase. In the simulation
model, no inertia is present in retracting legs during flight phase because the legs are
considered massless. However, the legs have non-negligible mass in actual running,
and, therefore, the angular speed that the legs could achieve during flight phase is
far below from the no-load-speed. Even augmenting the source voltage to motors
(as an attempt to increase the motor angular speed limit), the robot could not run
stably either with leg A or with leg B.
In order to seek for gaits that could be realizable in actual running experiments, a
second optimization was run with some modifications. A constant upper speed limit
(lower than no-load-speed) was added to the existing motor model for the flight
phase in simulation. Also, in order to avoid gaits which are optimal for specific
resistance but are unstable (i.e. |λmax | > 1), a penalty cost function is added in order
to avoid convergence to an unstable gait. Re-optimization with these constraints
yields gaits with running performances shown in Table 5.

Table 5 Simulation results by running with the controller parameters optimized for specific
resistance

Study Leg |λmax | vx [m/s] Mechanical Power [W] SR


4 A 0.8937 1.2757 6.7539 0.5260
5 B 0.9769 1.1447 5.0909 0.4419
6 C 0.9419 0.8412 4.3179 0.5100

Simulation-to-experiment comparison results for Study 4 and 5 are shown in Fig-


ure 7. Once again, running with leg C with controller parameters of Study 6 was not
realizable because of its fragility. The graphs shown in the first row of Figure 7
correspond to running with leg A for Study 4, and the graphs of the second row,
running with leg B for Study 5. The results obtained from running with both legs
differ significantly from the simulation predictions in the fluctuations of the height
and the forward speed of CoM.

Table 6 Experimental running performances for Study 4 and Study 5

Study Leg v̄x [m/s] P̄[W] SR


4 A 0.2910 7.9177 2.7029
5 B 0.5366 6.3050 1.1674

Table 6 shows that running with leg B is again faster and more efficient than
running with leg A, despite the fact that both experimental runnings diverge greatly
from the simulation results. All simulation and experimental results shown in this
772 J.Y. Jun, D. Haldane, and J.E. Clark

^ŝŵƵůĂƚŝŽŶ
džƉĞƌŝŵĞŶƚ

^ƚƵĚLJϰ;>ĞŐͿ
;ĂͿ ;ďͿ
^ŝŵƵůĂƚŝŽŶ
džƉĞƌŝŵĞŶƚ
^ƚƵĚLJϱ;>ĞŐͿ

;ĐͿ ;ĚͿ

Fig. 7 Simulation-to-experiment comparison results for Study 4 (leg A) and Study 5 (leg B).
The stride period for Study 4 is T = 0.3692s, and for Study 5, T = 0.3226s.

section seem to indicate that running with leg B results in faster and more stable
gait than running with leg A, although more experiments need to be performed to
confirm this.

5 Conclusion and Future Works


Motivated by the groundbreaking running performances of RHex-like robots using
curved legs [3], the present work first approximates running with curved legs in sim-
ulation by using two-segment-leg model both in energy-conservative and energy-
dissipative systems with the purpose of understanding aspects of the leg design that
contribute to stable, efficient and fast running. Using the two-segment-leg model for
the energy-conservative system, we have observed that the size of the stability re-
gion (Δ α ) is maximized when the ratio of the length of the proximal segment to the
length of the distal segment (l2 /l1 ) is unity for all leg intersegmental rest angles (βo )
and for all reference stiffness values considered. The size of the stability region also
increases with the leg intersegmental rest angle when an appropriate leg stiffness
value is chosen. These results led to the design of two novel compliant monolithic
legs.
A new bipedal robot was built and used to test the running behavior with these
legs. It was found that leg C (βo = 150o) could not be tested satisfactorily on the
experimental running trials because of its fragility, and, therefore, a new design is
needed in the future.
Although the articulated simulation model seems to predict the robotic running
behavior with curved legs in some instances, it does not appear to capture foot slip-
ping, rolling contact, or motor limitations particularly well. This suggests that bet-
ter contact and motor models need to be incorporated in the simulation model. In
Compliant Leg Shape, Reduced-Order Models and Dynamic Running 773

addition, curved legs have other aspects that the described model does not capture:
leg length change, rolling motion, and the stiffness directionality. The influence of
these aspects on the running performance (such as stability, efficiency and speed of
running) remains to be analyzed. Better results may be obtained if the controller
optimization process is performed on the actual robotic running. Another interest-
ing result that may bear further investigation is that the gaits optimized for stability
appear to map better to the robot than those optimized for specific resistance.
Despite the differences between the simulations and the experiments, the results
obtained from both systems seem to indicate that running with leg B (βo = 90o ,
l2 /l1 = 1, k̃10% = 12.46) is faster and more efficient than running with leg A (βo =
90o , l2 /l1 = 0.35, k̃10% = 12.46), and these results seem to be due to the ratio of the
lengths of the leg segments chosen for leg B.

References
1. Blickhan, R.: The spring mass model for running and hopping. Journal of Biomechan-
ics 22, 1217–1227 (1989)
2. Blickhan, R., Seyfarth, A., Geyer, H., Grimmer, S., Wagner, H., Gunther, M.: Intelli-
gence by mechanics. Philosophical Transactions of the Royal Society London, Series A
(Mathematical, Physical and Engineering Sciences) 365, 199–220 (2007)
3. Galloway, K.C.: Passive variable compliance for dynamic legged robots. Ph.D. thesis
(2010)
4. Galloway, K.C., Clark, J.E., Koditschek, D.E.: Design of a tunable stiffness composite
leg for dynamic locomotion. In: Proc. ASME IDETC/CIE (2009)
5. Hyon, S.H., Mita, T.: Development of a biologically inspired hopping robot - kenken. In:
Proceedings of IEEE International Conference on Robotics and Automation (2002)
6. Jun, J.Y., Clark, J.E.: Dynamic stability of variable stiffness running. In: Proceedings of
IEEE International Conference on Robotics and Automation (2009)
7. Jun, J.Y., Clark, J.E.: Effect of rolling on running performance. In: Proceedings of IEEE
International Conference on Robotics and Automation (2011) (submitted)
8. Kim, S., Clark, J.E., Cutkosky, M.R.: isprawl: design and tuning for high-speed au-
tonomous open-loop running. International Journal of Robotics Research 25, 903–912
(2006)
9. Komsuoglu, H.: Towards a comprehensive infrastructure for construction of modular
and extensible robotic systems. Tech. rep., Dept. of Computer and Information Science,
University of Pennsylvania (2007)
10. Raibert, M.: Bigdog, the rough-terrain quadruped robot. In: The International Federation
of Automatic Control (2008)
11. Rummel, J., Seyfarth, A.: Stable running with segmented legs. International Journal of
Robotics Research 27, 919–934 (2008)
12. Saranli, U., Buehler, M., Koditschek, D.E.: Rhex: A simple and highly mobile hexapod
robot. International Journal of Robotics Research 20, 616–631 (2001)
13. Whittington, B.R., Thelen, D.G.: A simple mass-spring model with roller feet can induce
the ground reactions observed in human walking. Journal of Biomechanical Engineering-
Transactions of the Asme 131 (2009)
14. Zhang, Z.G., Kimura, H.: Rush: a simple and autonomous quadruped running robot.
Proceedings of the Institution of Mechanical Engineers, Part I (Journal of Systems and
Control Engineering) 223, 323–336 (2009)

You might also like