Underactuated Robotics

Algorithms for Walking, Running, Swimming, Flying, and Manipulation

Russ Tedrake

© Russ Tedrake, 2024
Last modified .
How to cite these notes, use annotations, and give feedback.

Note: These are working notes used for a course being taught at MIT. They will be updated throughout the Spring 2024 semester. Lecture videos are available on YouTube.

Previous Chapter Table of contents Next Chapter

Simple Models of Walking and Running

Practical legged locomotion is one of the fundamental problems in robotics; we've seen amazing progress over the last few years, but there are still some major unsolved problems. Much of the recent progress is due to improvements in hardware -- a legged robot must carry all of its sensors, actuators and power and traditionally this meant underpowered motors that act as far-from-ideal force/torque sources -- Boston Dynamics and other companies have made incredible progress here. The control systems implemented on these systems, though, are still more heuristic than you might think. They also require dramatically higher bandwidth and lower latency than the human motor control system and still perform worse in challenging environments.

The control of walking robots is definitely underactuated: assuming we cannot pull on the ground (and barring any aerodynamic effects!), then no matter how powerful my actuators are, there is nothing that we can do to accelerate the center of mass of the robot towards the ground faster than gravity. But beyond the underactuated systems we have studied so far, the control of walking robots faces an additional complexity: controlling through contact. The fact that we can get non-zero contact forces if and only if two bodies are in contact introduces a fundamentally discrete/combinatorial element into the problemI've had some interesting arguments with folks on this point, because it's possible to write contact models that smooth the discontinuity, and/or to model systems that have neglible collision events. But replacing a discontinuity in the vector field with a stiff but smooth transition does not remove the need to decide whether or not your robot should make contact to accomplish a task..

Contact is essential for many aspects of robotics (manipulation is my other favorite example); it's sad to see so many robots going through life avoiding contact at any cost. Controlling contact means that your robot is capable of performing physical work on the environment; isn't that the point of robotics, afterall?

I like to start our discussion of contact with walking robots for a few important reasons. As you'll see in this chapter, there are a number of elegant "simple models" of walking that capture much of the essence of the problem with minimal complexity. But there is another, more subtle, reason: I want to continue to use the language of stability. At this point we've seen both finite-horizon and infinite-horizon problem formulations -- when we can say something about the behavior of a system in the limit as time goes to infinity, it's almost always cleaner than a finite-time analysis (because time drops away). Walking provides a very clean way to use the language of stability in studying behavior of systems that are sometimes in contact but sometimes out of contact in the limiting case; we just need to extend our notions from stability of a fixed point to the stability of a (periodic) cycle.

Limit Cycles

A limit cycle is an periodic orbit that is a limit set of the dynamical system; they can be stable (the limit set of neighboring trajectories as time goes to infinity) or unstable (the limit set of neighboring trajectories as time goes to negative infinity). One of the simplest models of limit cycle behavior is the Van der Pol oscillator. Let's examine that first...

Van der Pol Oscillator

The Van der Pol oscillator is described by a second-order differential equation in one variable: $$\ddot{q} + \mu (q^2 - 1)\dot{q} + q = 0, \quad \mu>0$$ One can think of this system as almost a simple spring-mass-damper system, except that it has nonlinear damping. In particular, the velocity term dissipates energy when $|q|>1$, and adds energy when $|q|<1$. Therefore, it is not terribly surprising to see that the system settles into a stable oscillation from almost any initial conditions (the exception is the state $q=0,\dot{q}=0$). This can be seen nicely in the phase portrait below.

System trajectories of the Van der Pol oscillator with $\mu =.2$. (Left) phase portrait. (Right) time domain.

However, if we examine system trajectories in the time domain (see the right panel of the figure above), then we can see that our traditional notion for stability of a trajectory, $\lim_{t\rightarrow \infty} |\bx(t) - \bx^*(t)| = 0$, is not satisfied here for any $\bx^*(t)$. Although the phase portrait clearly reveals that all trajectories converge to the same orbit, the time domain plot reveals that these trajectories do not necessarily synchronize in time.

This example shows that stability of a trajectory (in time) is not the definition we want. Instead we will define the stability of an orbit, $\bx^*(t)$, using \[ \min_\tau || \bx(t) - \bx^*(\tau) ||.\] Here the quantity of interest is the distance between our trajectory and the closest point on the orbit. Depending on exactly how this distance evolves with time, we can define orbital stability: for every small $\epsilon>0$, we can find a $\delta>0$ such that $\min_\tau||\bx(t_0) - \bx^*(\tau)|| < \delta$ implies $\forall t, \min_\tau ||\bx(t) - \bx^*(\tau)|| < \epsilon$). The definitions for asymptotic orbital stability, exponential orbital stability, finite-time orbital stability follow naturally, as does the definition of an unstable orbit. In the case of limit cycles, this $\bx^*(t)$ is a periodic solution with $\bx^*(t+t_{period}) = \bx^*(t)$.

As we begin to extend our notions of stability, allow me to make a few quick connections to the discussion of stability in the chapter on Lyapunov analysis. It's interesting to know that if we can find an invariant set in state space which does not contain any fixed points, then this set must contain a closed orbit (this is the Poincaré-Bendixson Theorem) Strogatz94. It's also interesting to note that gradient potential fields (and Lyapunov functions with $\dot{V}(\bx) \prec 0$) cannot have a closed orbit Strogatz94, and consequently we will also have to slightly modify the Lyapunov analysis we have introduced so far to handle limit cycles. We'll discuss this in some details in an upcoming chapter.

Poincaré Maps

The first tool one should know for analyzing the stability of a limit cycle, and one that will serve us quite well for the examples in this chapter, is the method of Poincaré. Let's consider a dynamical system, $\dot{\bx} = {\bf f}(\bx),$ with $\bx \in \Re^n.$ Define an $n-1$ dimensional surface of section, $S$. We will also require that $S$ is transverse to the flow (i.e., all trajectories starting on $S$ flow through $S$, not parallel to it). The Poincaré map (or return map) is a mapping from $S$ to itself: $$\bx_p[n+1] = {\bf P}(\bx_p[n]),$$ where $\bx_p[n]$ is the state of the system at the $n$th crossing of the surface of section. Note that we will use the notation $\bx_p$ to distinguish the state of the discrete-time system from the continuous-time system; they are related by $\bx_p[n] = \bx(t_c[n])$, where $t_c[n]$ is the time of the $n$th crossing of $S$.

For our purposes in this chapter, it will be useful for us to allow the Poincaré maps to also include fixed points of the original continuous-time system, and to define the return time to be zero. These points do not satisfy the transversality condition, and require some care numerically, but they do not compromise our analysis.

Return map for the Van der Pol Oscillator

Since the state space of this system is two-dimensional, a return map for the system must have dimension one. For instance, we can take $S$ to be the line segment where $q = 0$, $\dot{q} \ge 0$. It is easy to see that all trajectories that start on S are transverse to it.

One dimensional maps, like one dimensional flows, are amenable to graphical analysis.

(Left) Phase plot with the surface of section, $S$, drawn with a black dashed line. (Right) The resulting Poincaré first-return map (blue), and the line of slope one (red).

If we can demonstrate that ${\bf P}(\bx_p)$ exists for all $\bx_p \in S$, (all trajectories that leave $S$ eventually return to $S$), then something remarkable happens: we can reduce the stability analysis for a limit cycle in the original coordinates into the stability analysis of a fixed point on the discrete map. If $\bx^*_p$ is a stable (or unstable) fixed point of the map ${\bf P}$, then there exists a unique periodic orbit $\bx^*(t)$ which passes through $\bx^*_p$ that is a (stable, or unstable) limit cycle. If we are able to upper-bound the time it takes for trajectories leaving $S$ to return, then it is also possible to infer asympotitic orbital or even exponential orbital stability.

In practice it is often difficult or impossible to find ${\bf P}$ analytically, but it can be sampled quite reasonably numerically. Once a fixed point of ${\bf P}$ is obtained (by finding a solution to $\bx^*_p = {\bf P}(\bx^*_p)$), we can infer local limit cycle stability with an eigenvalue analysis. If this eigen-analysis is performed in the original coordinates (as opposed to the $n-1$ dimensional coordinates of $S$), then there will always be a at least one zero eigenvalue - corresponding to perturbations along the limit cycle which do not change the state of first return. The limit cycle is considered locally exponentially stable if all remaining eigenvalues, $\lambda_i$, have magnitude less than one, $|\lambda_i|<1$.

It is sometimes possible to infer more global stability properties of the return map by examining ${\bf P}$. We will study computational methods in the later chapter. Koditschek91 describes some less computational stability properties known for unimodal maps which are useful for the simple systems we study in this chapter.

the paragraph below treats x_p as living in the reduced space, whereas the text above kept it in the original space.

A particularly graphical method for understanding the dynamics of a one-dimensional iterated map is the "staircase method": Sketch the Poincaré map -- a plot of $x_p[n]$ vs $x_p[n+1]$ -- and also the line of slope one. In order to "iterate" the map, take an initial condition on the plot as a point $\left( x_p[k], x_p[k] \right)$, and move vertically on the plot to the point $\left( x_p[k], x_p[k+1] = f(x_p[k]) \right)$. Now move horizontally to the next point $\left( x_p[k+1], x_p[k+1] \right)$ and repeat.

The "staircase" on the return map of the Van der Pol Oscillator with $\mu = 0.2$, starting from two initial conditions ($\dot{q}_p[0] = .5$ and $\dot{q}_p[0] = 3.5$).

We can quickly recognize the fixed points as the points where the return map crosses this line. We can infer local exponential stability if the absolute value of the slope of return map at the fixed point is less than one ($|\lambda| < 1$). The figure above illustrates the stability of the Van der Pol limit cycle; it also illustrates the (one-sided) instability of the fixed point at the origin. Understanding the regions of attraction of fixed points on an iterated map (or any discrete-time system) graphically requires a little more care, though, than the continuous-time flows we examined before. The discrete maps can jump a finite distance on each step, and will in fact oscillate back and forth from either side of the fixed point when $\lambda < 0.$

Simple Models of Walking

Much of the fundamental work in the dynamics of legged robots was done originally in the context of studying passive-dynamic walkers (often just "passive walker"). We've already seen my favorite example of a passive walker, in the first chapter, but here it is again:

A 3D passive dynamic walker by Steve Collins and Andy RuinaCollins01.

This amazing robot has no motors and no controllers... it walks down a tiny ramp and is powered entirely by gravity. We can build up our understanding of this robot in a series of steps. As we will see, what is remarkable is that even though the system is passive, the periodic gait you see in the video is actually a stable limit cycle!

Well before the first passive walkers, one of the earliest models of walking was proposed by McMahonMcMahon80, who argued that humans use a mostly ballistic (passive) gait. He observed that muscle activity (recorded via EMG) in the "stance leg" is relatively high, but muscle activity in the "swing leg" is very low, except for at the very beginning and end of swing. He proposed a "ballistic walker" model -- a three-link pendulum in the sagittal plane, with one link from the ankle to the hip representing a straight "stance leg", the second link from the hip back down to the "swing leg" knee, and the final link from the knee down to the swing foot -- and argued that this model could largely reproduce the kinematics of gait. This model of "walking by vaulting" is considered overly simplistic today, as we now understand much better the role of compliance in the stance leg during walking. His model was also restricted to the continuous portion of gait, not the actual dynamics of taking a step. But it was the start. Nearly a decade later, McGeerMcGeer90 followed up with a series of similarly inspired walking machines, which he coined "passive-dynamic walkers".

The Rimless Wheel

The rimless wheel. The orientation of the stance leg, $\theta$, is measured clockwise from the vertical axis (the camera is looking in the positive $y$ direction in vehicle coordinates).

Perhaps the simplest possible model of a legged robot, introduced as the conceptual framework for passive-dynamic walking by McGeerMcGeer90, is the rimless wheel. This system has rigid legs and only a point mass at the hip as illustrated in the figure above. To further simplify the analysis, we make the following modeling assumptions:

  • Collisions with ground are inelastic and impulsive (only angular momentum is conserved around the point of collision).
  • The stance foot acts as a pin joint and does not slip.
  • The transfer of support at the time of contact is instantaneous (no double support phase).
  • $0 \le \gamma < \frac{\pi}{2}$, $0 < \alpha < \frac{\pi}{2}$, $l > 0$.

Note that the coordinate system used here is slightly different than for the simple pendulum ($\theta=0$ is at the top, and the sign of $\theta$ has changed).

update the coordinates (here and in drake)

The most comprehensive analysis of the rimless wheel was done by Coleman98a.

Stance Dynamics

The dynamics of the system when one leg is on the ground are given by $$\ddot\theta = \frac{g}{l}\sin(\theta).$$ If we assume that the system is started in a configuration directly after a transfer of support ($\theta(0^+) = \gamma-\alpha$), then forward walking occurs when the system has an initial velocity, $\dot\theta(0^+) > \omega_1,$ where $$\omega_1 = \sqrt{ 2 \frac{g}{l} \left[ 1 - \cos \left (\gamma-\alpha \right) \right]}.$$ $\omega_1$ is the threshold at which the system has enough kinetic energy to vault the mass over the stance leg and take a step. This threshold is zero for $\gamma = \alpha$ and does not exist for $\gamma > \alpha$ (because when $\gamma > \alpha$ the mass is always ahead of the stance foot, and the standing fixed point disappears). The next foot touches down when $\theta(t) = \gamma+\alpha$, at which point the conversion of potential energy into kinetic energy yields the velocity $$\dot\theta(t^-) = \sqrt{\dot\theta^2(0^+) + 4\frac{g}{l} \sin\alpha \sin\gamma }.$$ $t^-$ denotes the time immediately before the collision.

Foot Collision

Angular momentum is conserved around the point of impact

The angular momentum around the point of collision at time $t$ just before the next foot collides with the ground is $$L(t^-) = -ml^2\dot\theta(t^-) \cos(2\alpha).$$ The angular momentum at the same point immediately after the collision is $$L(t^+) = -ml^2\dot\theta(t^+).$$ Assuming angular momentum is conserved, this collision causes an instantaneous loss of velocity: $$\dot\theta(t^+) = \dot\theta(t^-) \cos(2\alpha).$$

Forward simulation

Given the stance dynamics, the collision detection logic ($\theta = \gamma \pm \alpha$), and the collision update, we can numerically simulate this simple model. Doing so reveals something remarkable... the wheel starts rolling, but then one of two things happens: it either runs out of energy and stands still, or it quickly falls into a stable periodic solution. Convergence to this stable periodic solution appears to happen from a huge range of initial conditions. Try it yourself.

Numerical simulation of the rimless wheel

I've setup the notebook to make it easy for you to try a handful of interesting initial conditions. And even made an interactive widget for you to watch the simulation phase portrait as you change those initial conditions. Give it a try!

The rimless wheel model actually uses a number of the more nuanced features of in order to simulate the hybrid system accurately (as do many of the examples in this chapter). It actually registers the collision events of the system -- the simulator uses zero-crossing detection to ensure that the time/state of collision is obtained accurately, and then applies the reset map.

Phase portrait of the rimless wheel ($m=1$, $l=1$, $g=9.8$, $\alpha=\pi/8$, $\gamma=0.08$).

One of the fantastic things about the rimless wheel is that the dynamics are exactly the undamped simple pendulum that we've thought so much about. So you will recognize the phase portraits of the system -- they are centered around the unstable fixed point of the simple pendulum.

Poincaré Map

We can now derive the angular velocity at the beginning of each stance phase as a function of the angular velocity of the previous stance phase. First, we will handle the case where $\gamma \le \alpha$ and $\dot\theta_n^+ > \omega_1$. The "step-to-step return map", factoring losses from a single collision, the resulting map is: $$\dot\theta^{+}_{n+1} = \cos(2\alpha) \sqrt{ ({\dot\theta_n}^{+})^{2} + 4\frac{g}{l}\sin\alpha \sin\gamma}.$$ where the $\dot{\theta}^{+}$ indicates the velocity just after the energy loss at impact has occurred.

Using the same analysis for the remaining cases, we can complete the return map. The threshold for taking a step in the opposite direction is $$\omega_2 = - \sqrt{2 \frac{g}{l} [1 - \cos(\alpha + \gamma)]}.$$ For $\omega_2 < \dot\theta_n^{+} < \omega_1,$ we have $$\dot\theta_{n+1}^{+} = -\dot\theta_n^{+} \cos(2\alpha).$$ Finally, for $\dot\theta_n^{+} < \omega_2$, we have $$\dot\theta_{n+1}^{+} = - \cos(2\alpha)\sqrt{(\dot\theta_n^{+})^2 - 4 \frac{g}{l} \sin\alpha \sin\gamma}.$$ Altogether, we have (for $\gamma \le \alpha$) $$\dot\theta_{n+1}^{+} = \begin{cases} \cos(2\alpha) \sqrt{(\dot\theta_n^{+})^2 + 4 \frac{g}{l} \sin\alpha \sin\gamma}, & \text{ } \omega_1 < \dot\theta_n^{+} \\ -\dot\theta_n^{+} \cos(2\alpha), & \text{ } \omega_2 < \dot\theta_n^{+} < \omega_1 \\ -\cos(2\alpha) \sqrt{(\dot\theta_n^{+})^2 - 4\frac{g}{l} \sin\alpha \sin\gamma}, & \dot\theta_n^{+} < w_2 \end{cases}.$$

Notice that the return map is undefined for $\dot\theta_n = \{ \omega_1, \omega_2 \}$, because from these configurations, the wheel will end up in the (unstable) equilibrium point where $\theta = 0$ and $\dot\theta = 0$, and will therefore never return to the map.

This return map blends smoothly into the case where $\gamma > \alpha$. In this regime, $$\dot\theta_{n+1}^{+} = \begin{cases} \cos(2\alpha) \sqrt{(\dot\theta_n^{+})^2 + 4 \frac{g}{l} \sin\alpha \sin\gamma}, & \text{ } 0 \le \dot\theta_n^{+} \\ -\dot\theta_n^{+} \cos(2\alpha), & \text{ } \omega_2 < \dot\theta_n^{+} < 0 \\ -\cos(2\alpha) \sqrt{(\dot\theta_n^{+})^2 - 4\frac{g}{l} \sin\alpha \sin\gamma}, & \dot\theta_n^{+} \le w_2 \end{cases}.$$ Notice that the formerly undefined points at $\{\omega_1,\omega_2\}$ are now well-defined transitions with $\omega_1 = 0$, because it is kinematically impossible to have the wheel statically balancing on a single leg.

Fixed Points and Stability

For a fixed point, we require that $\dot\theta_{n+1}^{+} = \dot\theta_n^{+} = \omega^*$. Our system has two possible fixed points, depending on the parameters: $$ \omega_{stand}^* = 0,~~~ \omega_{roll}^* = \cot(2 \alpha) \sqrt{4 \frac{g}{l} \sin\alpha\sin\gamma}.$$ The limit cycle plotted above illustrates a state-space trajectory in the vicinity of the rolling fixed point. $\omega_{stand}^*$ is a fixed point whenever $\gamma < \alpha$. $\omega_{roll}^*$ is a fixed point whenever $\omega_{roll}^* > \omega_1$. It is interesting to view these bifurcations in terms of $\gamma$. For small $\gamma$, $\omega_{stand}$ is the only fixed point, because energy lost from collisions with the ground is not compensated for by gravity. As we increase $\gamma$, we obtain a stable rolling solution, where the collisions with the ground exactly balance the conversion of gravitational potential to kinetic energy. As we increase $\gamma$ further to $\gamma > \alpha$, it becomes impossible for the center of mass of the wheel to be inside the support polygon, making standing an unstable configuration.

Limit cycle trajectory of the rimless wheel ($m=1,l=1,g=9.8,\alpha=\pi/8,\gamma=0.15$). All hatched regions converge to the rolling fixed point, $\omega_{roll}^*$; the white regions converge to zero velocity ($\omega_{stand}^*$).

Stability of standing still

I opened this chapter with the idea that the natural notion of stability for a walking system is periodic stability, and I'll stand by it. But we can also examine the stability of a fixed-point (of the original coordinates; no Poincaré this time) for a system that has contact mechanics. For a legged robot, a fixed point means standing still. This can actually be a little subtle in our hybrid models: in the rimless wheel, standing still is the limiting case where we have infinitely frequent collisions. [Details coming soon.]

call out to the fact that I left the origin OUT of the poincare map on the van der pol.

The Compass Gait

The rimless wheel models only the dynamics of the stance leg, and simply assumes that there will always be a swing leg in position at the time of collision. To remove this assumption, we take away all but two of the spokes, and place a pin joint at the hip. To model the dynamics of swing, we add point masses to each of the legs. I've derived the equations of motion for the system assuming that we have a torque actuator at the hip - resulting in swing dynamics equivalent to an Acrobot (although in a different coordinate frame) - but let's examine the passive dynamics of the system first ($\bu = 0$).

The compass gait

In addition to the modeling assumptions used for the rimless wheel, we also assume that the swing leg retracts in order to clear the ground without disturbing the position of the mass of that leg. This model, known as the compass gait, is well studied in the literature using numerical methods Goswami96a+Goswami99+Spong03, but relatively little is known about it analytically.

The state of this robot can be described by 4 variables: $\theta_{st},\theta_{sw},\dot\theta_{st}$, and $\dot\theta_{sw}$. The abbreviation $st$ is shorthand for the stance leg and $sw$ for the swing leg. Using $\bq = [ \theta_{st}, \theta_{sw} ]^T$ and $\bu = \tau$, we can write the dynamics as $$\bM(\bq)\ddot\bq + \bC(\bq,\dot\bq)\dot\bq = \btau_g(q) + \bB\bu,$$ with \begin{gather*} \bM = \begin{bmatrix} (m_h+m)l^2 + ma^2 & -mlb\cos(\theta_{st}-\theta_{sw}) \\ -mlb\cos(\theta_{st}-\theta_{sw}) & mb^2 \end{bmatrix}\\ \bC = \begin{bmatrix} 0 & -mlb\sin(\theta_{st}-\theta_{sw})\dot\theta_{sw} \\ mlb\sin(\theta_{st}-\theta_{sw})\dot\theta_{st} & 0 \end{bmatrix} \\ \btau_g(q) = \begin{bmatrix} (m_hl + ma + ml)g\sin(\theta_{st}) \\ -mbg\sin(\theta_{sw}) \end{bmatrix},\\ \bB = \begin{bmatrix} -1 \\ 1 \end{bmatrix} \end{gather*} and $l=a+b$.

The foot collision is an instantaneous change of velocity caused by a impulsive force at the foot that brings the foot to rest. The update to the velocities can be calculated using the derivation for impulsive collisions derived in the appendix. To use it, we proceed with the following steps:

  • Add a "floating base" to the system by adding a free (x,y) joint at the pre-impact stance toe, $\bq_{fb} = [x,y,\theta_{st},\theta_{sw}]^T.$
  • Calculate the mass matrix for this new system, call it $\bM_{fb}$.
  • Write the position of the (pre-impact) swing toe as a function $\bphi(\bq_{fb})$. Define the Jacobian of this function: $\bJ = \frac{\partial \bphi}{\partial \bq_{fb}}.$
  • The post-impact velocity that ensures that $\dot\bphi = 0$ after the collision is given by $$\dot\bq^+ = \left[ \bI - \bM_{fb}^{-1}\bJ^T[\bJ\bM_{fb}^{-1}\bJ^T]^{-1}\bJ\right] \dot\bq^-,$$ noting that $\dot{x}^- = \dot{y}^- = 0$, and that you need only read out the last two elements of $\dot{\bq}^+$. The velocity of the post-impact stance foot will be zero by definition, and the new velocity of the pre-impact stance foot can be completely determined from the minimal coordinates.
  • Switch the stance and swing leg positions and velocities.

Numerical simulation of the compass gait

You can simulate the passive compass gait using:

Try playing with the initial conditions. You'll find this system is much more sensitive to initial conditions than the rimless wheel. It actually took some work to find the initial conditions that make it walk stably.

Numerical integration of these equations reveals a stable limit cycle, plotted below. Notice that when the left leg is in stance (top part of the plot), the trajectory quite resembles the familiar pendulum dynamics of the rimless wheel. But this time, when the leg impacts, it takes a long arc around as the swing leg before returning to stance. The impacts are also clearly visible as discontinuous changes (decreases) in velocity. The dependence of this limit cycle on the system parameters has been studied extensively in Goswami96a.

Passive limit cycle trajectory of the compass gait. ($m=5$kg, $m_h=10$kg, $a=b=0.5$m, $\phi=0.0525$rad. $\bx(0)=[0,0,0.4,-2.0]^T$). Drawn is the phase portait of the left leg angle, which is recovered from $\theta_{st}$ and $\theta_{sw}$ in the simulation with some simple book-keeping.

The basin of attraction of the stable limit cycle is a narrow band of states surrounding the steady state trajectory. Although the simplicity of this model makes it attractive for conceptual explorations, this lack of stability makes it difficult to implement on a physical device.

The Kneed Walker

To achieve a more anthropomorphic gait, as well as to acquire better foot clearance and ability to walk on rough terrain, we want to model a walker that includes a kneeHsu07. The mathematical modeling would be simpler if we used a single link for the stance leg, but we'll go ahead and use two links for each leg (and each with a point mass) because this robot can actually be built!

The Kneed Walker

At the beginning of each step, the system is modeled as a three-link pendulum, like the ballistic walkerMcMahon80+Mochon80+Spong03. The stance leg is the one in front, and it is the first link of a pendulum, with two point masses. The swing leg has two links, with the joint between them unconstrained until knee-strike. Given appropriate mass distributions and initial conditions, the swing leg bends the knee and swings forward. When the swing leg straightens out (the lower and upper length are aligned), knee-strike occurs. The knee-strike is modeled as a discrete inelastic collision, conserving angular momentum and changing velocities instantaneously.

After this collision, the knee is locked and we switch to the compass gait model with a different mass distribution. In other words, the system becomes a two-link pendulum. Again, the heel-strike is modeled as an inelastic collision. After the collision, the legs switch instantaneously. After heel-strike then, we switch back to the ballistic walker's three-link pendulum dynamics. This describes a full step cycle of the kneed walker, which is shown above.

Limit cycle trajectory for kneed walker

By switching between the dynamics of the continuous three-link and two-link pendulums with the two discrete collision events, we characterize a full point-feed kneed walker walking cycle. After finding appropriate parameters for masses and link lengths, a stable gait is found. As with the compass gait's limit cycle, there is a swing phase (top) and a stance phase (bottom). In addition to the two heel-strikes, there are two more instantaneous velocity changes from the knee-strikes as marked in the figure. This limit cycle is traversed clockwise.

Limit cycle (phase portrait) of the kneed walker
reproduce this model in drake.

Curved feet

The region of attraction of the kneed-walking limit cycle can be improved considerably by the addition of curved feet...

also knee "de-bouncers" (suction cups). Simulation model?
Tad McGeer's kneed walker. Here is Matt Haberland's guide to launching it successfully. It's nontrivial!

In Collins05, we studied the benefits of adding a small amount of energy to these passive dynamic walkers in order to allow them to walk on flat terrain -- sometimes very efficiently! You can find the videos of the Cornell biped here, the TU Delft biped here, and the MIT learning biped (aka "Toddler") here. Toddler can sometimes still be found in the MIT Museum.

And beyond...

The world has seen all sorts of great variations on the passive-dynamic walker theme. Almost certainly the most impressive are the creations of Dutch artist Theo Jansen -- he likes to say that he is creating "new forms of life" which he calls the Strandbeest. There are many variations of these amazing machines -- including his beach walkers which are powered only by the wind (I've actually been able to visit Theo's workshop once; it is very windy there).

These results are very real. Robin Deits (an exceptionally talented student who felt particularly inspired once on a long weekend) once reproduced one of Theo's earliest creations in .

Drake simulation of the Strandbeest.

Simple Models of Running

Just like walking, our understanding of the dynamics and control of running has evolved by a nice interplay between the field of biomechanics and the field of robotics. But in running, I think it's fair to say that the roboticists got off the starting blocks first. And I think it's fair to say that it started with a series of hopping machines built by Marc Raibert and the Leg Laboratory (which was first at CMU, but moved to MIT) in the early 1980's. At a time where many roboticsts were building relatively clumsy walking robots that moved very slowly (when they managed to stay up!), the Leg Laboratory produced a series of hopping machines that threw themselves through the air with considerable kinetic energy and considerable flair.

To this day, I'm still a bit surprised that impressive running robots (assuming you accept hopping as the first form of running) appeared before the impressive walking robots. I would have thought that running would be a more difficult control and engineering task than walking. But these hopping machines turned out to be an incredibly clever way to build something simple and very dynamic.

Shortly after Marc Raibert built his machines, Dan Koditschek and Martin Buehler started analyzing them with simpler models Koditschek91. Years later, in collaboration with biologist Bob Full, they started to realize that the simple models used to describe the dynamics of Raibert's hoppers could also be used to describe experimental data obtained from running animals. In fact, they described an incredible range of experimental results, for animals ranging in size from a cockroach up to a horseFull99 (I think the best overall summary of this line of work is Holmes06). The most successful of the simple models was the so-called "spring-loaded inverted pendulum" (or "SLIP", for short).

The Spring-Loaded Inverted Pendulum (SLIP)

The model is a point mass, $m$, on top of a massless, springy leg with rest length of $l_0$, and spring constant $k$. The configuration of the system is given by the $x,z$ position of the center of mass, and the length, $l$, and angle $\theta$ of the leg. The dynamics are modeled piecewise - with one dynamics governing the flight phase, and another governing the stance phase.

The Spring-Loaded Inverted Pendulum (SLIP) model

In SLIP, we actually use different state variables for each phase of the dynamics. During the flight phase, we use the state variables: $\bx = [x,z,\dot{x},\dot{z}]^T.$ The flight dynamics are simply $$\dot{\bx} = \begin{bmatrix} \dot{x} \\ \dot{z} \\ 0 \\ - g \end{bmatrix}.$$ Since the leg is massless, we assume that the leg can be controlled to any position instantaneously, so it's reasonable to take the leg angle as the control input $\theta = u$.

During the "stance phase", we write the dynamics in polar coordinates, with the foot anchored at the origin. Using the state variables: $\bx = [r, \theta, \dot{r}, \dot\theta]^T,$ the location of the mass is given by $$\bx_m = \begin{bmatrix} - r \sin\theta \\ r \cos\theta \end{bmatrix}.$$ The total energy is given by the kinetic energy (of the mass) and the potential energy from gravity and from the leg spring: $$T = \frac{m}{2} (\dot{r}^2 + r^2 \dot\theta^2 ), \quad U = mgr\cos\theta + \frac{k}{2}(l_0 - r)^2.$$ Plugging these into Lagrange yields the stance dynamics: \begin{gather*} m \ddot{r} - m r \dot\theta^2 + m g \cos\theta - k (l_0 - r) = 0, \\ m r^2 \ddot{\theta} + 2mr\dot{r}\dot\theta - mgr \sin\theta = 0.\end{gather*} We assume that the stance phase is entirely ballistic; there are no control inputs during this phase.

Unlike the rimless wheel model, the idealization of a spring leg with a massless toe means that no energy is lost during the collision with the ground. The collision event causes a transition to a different state space, and to a different dynamics model, but no instantaneous change in velocity. The transition from flight to stance happens when $z - l_0\cos\theta \le 0.$ The transition from stance to flight happens when the stance leg reaches its rest length: $r \ge l_0$.

Simulation of the SLIP model

You can simulate the spring-loaded inverted pendulum using:

Take a moment to read the code that defines the SpringLoadedInvertedPendulum system. You can see that we carefully register the hybrid guards (witness functions) and reset maps, which tells the numerical integrator to use zero-crossing event detection to isolate precisely the time of the event and to apply the reset update at this event. It is quite elegant, and we've gone to some length to make sure that the numerical integration routines in can support it.

We used the same witness functions and resets to simulate the walking models, but those were implemented in C++. This is a purely Python implementation, making it a bit easier to see those details.

Analysis on the apex-to-apex map

The rimless wheel had only two state variables ($\theta$, and $\dot\theta$), and since the surface of section drops one variable, we ended up with a one-dimensional Poincaré map. We have a lot more state variables here. Is there any hope for graphical analysis?

For this system, one can still make progress if we define the Poincaré section to be at the apex of the hop (when $\dot{z}=0$), and analyze the dynamics from apex to apex. We will ignore the absolute horizontal position, $x$, and the leg angle, $\theta$, as they are irrelevant to the flight dynamics. This still leaves us with two state variables at the apex: the horizontal velocity, $\dot{x},$ and the hopping height, $z$.

Importantly, at the apex, these two variables are linked -- when you know one, you know the other. This is because no energy is ever added or removed from the system. If you know the (constant) total energy in the system, and you know the height at the apex, then you know the horizontal velocity (up to a sign): $$E_{apex}(z, \dot{x}) = \frac{1}{2}m \dot{x}^2 + mgz = E_0.$$ To make our analysis less dependent on our specific choice of parameters, we typically report any analysis using a dimensionless total energy, $\tilde{E} = \frac{E}{m g l_0},$ and a dimensionless spring constant, $\tilde{k} = \frac{k l_0}{mg}$ Geyer05.

Putting all of these together, we can numerically evaluate the apex-to-apex return map given a known total energy of the system as a one-dimensional Poincaré map.

Numerical SLIP apex-to-apex map

Numerically integrating the equations of motion from apex-to-apex, using zero-crossing event detection, can give us the apex-to-apex map.

Apex-to-apex return map of the SLIP model given $\tilde{k} = 10.7$, $\tilde{E} = 1.61$, and a fixed touchdown angle of $\theta = 30$ deg (the parameters used in Geyer05).

Using our graphical staircase analysis (for discrete-time systems), we can see that there are two fixed points of this map: one "stable" fixed point around 0.87 meters, and one unstable fixed point around 0.92 meters.

As always, you should run this code yourself to make sure you understand:

I'd also recommend running the simulation using initial conditions at various points on this plot. You should ask: why does the plot cut off so abruptly just below the stable fixed point? Well the apex-to-apex map becomes undefined if the apex is smaller that $l_0\cos\theta$ (the height at touchdown). Furthemore, when $z$ is too large (relative to $\tilde{E}$), then the system will not have enough horizontal energy to transition through the stance phase and continue forward motion: the center of mass will compress the spring and then fall backwards in the direction from which it came.

more videos. in particular a plot/video of the failure case of falling backward would help a lot here.

The system's conservation of total energy is quite convenient for our return map analysis, but it should give you pause. The rimless wheel actually relied on the dissipation due to collisions with the ground to acquire its stability. Is there any chance for a system that is energy conserving to be stable? The apex-to-apex return maps visualized above reveal "stable" fixed-points, but you must understand that the projection from $(z,\dot{x})$ to just $z$ is using the assumption that the energy is conserved. As a result, our analysis on the one-dimensional Poincaré map can only make a statement about partial stability (i.s.L., asymptotic, or exponential); our analysis does not say anything about stability against disturbances that change the total energy of the system.

Despite this limitation, it is quite interesting that such a simple system converges to particular hopping heights from various initial conditions. The observation that these "piecewise-Hamiltonian" systems (Hamiltonian because they conserve total energy) can exhibit this partial stability generated quite a bit of interest from the theoretical physics community. You can sense that enthusiasm when reading Holmes06.

Geyer05 also gives a more thorough analysis of this apex-to-apex map, including a closed-form approximation to the return map dynamics based on a linear (small angle) approximation of the stance dynamics.

include Harmut's return map analysis here?

SLIP Control

Our SLIP analysis was simplified considerably by the assumption that the leg is massless, and can be moved instantaneously during the flight phase into the desired touchdown position. This might sound to you like an offensive oversimplification, but in fact the strong trend in legged robots these days has been to build physical devices which make this sort of abstraction (approximately) valid. For quadrupeds especially, it's hard to fall over when you can move your foot instantaneously into the direction where your body is falling.

     
(Left) Spot-mini from Boston Dynamic, (right) Mini-Cheetah from MIT. These are two of my favorite representatives from a fleet of legged robots that have extremely light legs (and approximately point feet), which simplifies control and can improve stability.

Assuming a massless leg also gives us a nice conceptual advantage -- it becomes natural to think about control design as making a control decision once per step. This fits nicely with our Poincaré analysis -- we can think of the apex-to-apex map of the SLIP model as being parameterized by the (scalar) control decision $u[n]$ which sets the touchdown angle for the $n$th hop: $$x_p[n+1] = P(x_p[n], u[n]).$$ This system can be stabilized like any other discrete-time system. For example, it would be perfectly reasonable to linearize the apex-to-apex map around the hopping fixed point, and use the (discrete-time) linear quadratic regulator to design a controller.

It turns out that for the SLIP model, there is a more interesting approach available. If we set the cost on control effort to zero (consistent with our massless leg assumption), then the optimal action to pick is the one that drives the system exactly to the desired fixed point in a single hop. This is the so-called deadbeat controller. Not all systems can be stabilized in this way, but Geyer05 showed that the relationship between the touchdown angle and the hopping height in the apex-to-apex map was an invertible relationship (within the bounds from the total energy of the system): given a desired hopping height one can uniquely choose a touchdown angle that will achieve it at the next apex.

Take a moment to reflect on this. In continuous time, if I were to remove the cost on control input, then the optimal actions would be unbounded, or bang-bang if there are input limits. But in discrete-time systems, this is not necessarily the case -- one can often achieve perfect convergence in a finite number of steps with bounded control input -- this corresponds to a finite-time convergence in continuous time. For this simple running model, what is particularly elegant is that this convergence can even be achieved in just one step.

In fact, the story for deadbeat control of the SLIP model gets even better. Ernst09 observed that the desired touchdown angle to achieve a deadbeat apex-to-apex max is a smooth function of the desired hopping height (or equivalently forward speed). Now imagine that we are trying to run/hop at a constant speed over rough terrain. For bonus points, let's do it blind -- with no sensing of the terrain height. It turns out that, given this smooth relationship between touchdown angle and running speed, one can sweep the leg through an open-loop leg angle trajectory (based only on the time since the apex), such that whenever the foot happens to collide with the ground, it will always be in exactly the deadbeat touchdown angle!

One would not expect to use the terms "blind", "running", "rough-terrain", and "deadbeat control" all at once to describe a control strategy. But there you have it. A very nice result! The authors went on to study whether this observation could help to explain the known phenomenon in biomechanics of "ground-speed matching" -- a human/animals' tendency to actively swing the leg around to meet the ground at the moment of touchdown.

Add a simulation example, and a corresponding figure here.

SLIP extensions

The original SLIP model, perhaps inspired by the hopping robots turned out to be quite effective in describing the vertical center of mass dynamics of a wide range of animals Holmes06. Remarkably, experiments showed that the dimensionless individual leg stiffness was very close to 10 for a huge range of animals, ranging from a $0.001~kg$ cockroach to a $135~kg$ horse Koditschek04. But the SLIP models, and modest extensions to them, have also been used to understand lateral stability in cockroaches, resulting in some of my favorite experimental videos of all time Jindrich02.

insert cockroach gunpowder video?

Hopping robots from the MIT Leg Laboratory

The 2D Hopper

A three-part control strategy: decoupling the control of hopping height, body attitude, and forward velocity Raibert86a.

Running on four legs as though they were one

bipeds, quadrupeds, and backflips...Raibert86

Towards human-like running

Add videos of 3D running from Katja (Schultz10).

A simple model that can walk and run

Exercises

One-Dimensional Hopper

In this exercise we implement a one-dimensional version of the controller proposed in Raibert84. The goal is to control the vertical motion of a hopper by generating a stable resonant oscillation that causes the system to hop off the ground at a given height. We accomplish this via a careful energy analysis of the hopping cycle: all the details are in . Besides understanding in detail the dynamics of the hopping system, in the notebook, you are asked to write two pieces of code:

  1. A function that, given the state of the hopper, returns its total mechanical energy.
  2. The hopping controller class. This is a VectorSystem that, at each sampling time, reads the state of the hopper and returns the preload of the hopper spring necessary for the system to hop at the desired height. All the necessary information for the synthesis of this controller are given in the notebook.

References

  1. Steven H. Strogatz, "Nonlinear Dynamics and Chaos: With Applications to Physics, Biology, Chemistry, and Engineering", Perseus Books , 1994.

  2. Daniel E. Koditschek and Martin Buehler, "Analysis of a Simplified Hopping Robot", International Journal of Robotics Research, vol. 10, no. 6, pp. 587-605, Dec, 1991.

  3. Steven H. Collins and Martijn Wisse and Andy Ruina, "A Three-Dimensional Passive-Dynamic Walking Robot with Two Legs and Knees", International Journal of Robotics Research, vol. 20, no. 7, pp. 607-615, July, 2001.

  4. Simon Mochon and Thomas A. McMahon, "Ballistic walking: An improved model", Mathematical Biosciences, vol. 52, no. 3-4, pp. 241-260, December, 1980.

  5. Tad McGeer, "Passive Dynamic Walking", International Journal of Robotics Research, vol. 9, no. 2, pp. 62-82, April, 1990.

  6. Michael J. Coleman, "A Stability Study of a Three-Dimensional Passive-Dynamic Model of Human Gait", PhD thesis, Cornell University, 1998.

  7. Ambarish Goswami and Benoit Thuilot and Bernard Espiau, "Compass-Like Biped Robot Part {I} : Stability and Bifurcation of Passive Gaits", Tech. Report, RR-2996, October, 1996.

  8. A. Goswami, "Postural stability of biped robots and the foot rotation indicator ({FRI}) point", International Journal of Robotics Research, vol. 18, no. 6, 1999.

  9. Mark W. Spong and Gagandeep Bhatia, "Further Results on Control of the Compass Gait Biped", Proc. IEEE International Conference on Intelligent Robots and Systems (IROS) , pp. 1933-1938, 2003.

  10. Vanessa Hsu, "Passive dynamic walking with knees: A Point-foot model", , February, 2007. [ link ]

  11. Simon Mochon and Thomas A. McMahon, "Ballistic Walking", Journal of Biomechanics, vol. 13, pp. 49-57, 1980.

  12. Steven H. Collins and Andy Ruina and Russ Tedrake and Martijn Wisse, "Efficient bipedal robots based on passive-dynamic walkers", Science, vol. 307, pp. 1082-1085, February 18, 2005. [ link ]

  13. R.J. Full and D.E. Koditschek, "Templates and anchors: neuromechanical hypotheses of legged locomotion on land", Journal of Experimental Biology, vol. 202, no. 23, pp. 3325-3332, 1999.

  14. Philip Holmes and Robert J. Full and Dan Koditschek and John Guckenheimer, "The Dynamics of Legged Locomotion: Models, Analyses, and Challenges", Society for Industrial and Applied Mathematics (SIAM) Review, vol. 48, no. 2, pp. 207--304, 2006.

  15. Hartmut Geyer, "Simple models of legged locomotion based on compliant limb behavior", PhD thesis, University of Jena, 2005.

  16. Michael Ernst and Hartmut Geyer and Reinhard Blickhan, "Spring-legged locomotion on uneven ground: a control approach to keep the running speed constant", International Conference on Climbing and Walking Robots , pp. 639--644, 2009.

  17. Daniel E Koditschek and Robert J Full and Martin Buehler, "Mechanical aspects of legged locomotion control", Arthropod structure \& development, vol. 33, no. 3, pp. 251--272, 2004.

  18. DL Jindrich and RJ Full, "Dynamic stabilization of rapid hexapedal locomotion", J Exp Biol, vol. 205, no. Pt 18, pp. 2803-23, Sep, 2002.

  19. Marc H. Raibert, "Legged Robots That Balance", The MIT Press , 1986.

  20. Raibert and M. H. and Chepponis and M. and Brown and H. B., "Running on four legs as though they were one", IEEE Journal of Robotics and Automation, vol. 2, no. 2, pp. 70--82, 1986.

  21. Marc H. Raibert, "Hopping in legged systems: Modeling and simulation for the 2D one-legged case", IEEE Trans. Systems, Man, and Cybernetics, vol. 14, pp. 451-463, 1984.

Previous Chapter Table of contents Next Chapter