Skip to main content

Advertisement

Log in

Symmetry in legged locomotion: a new method for designing stable periodic gaits

  • Published:
Autonomous Robots Aims and scope Submit manuscript

Abstract

This paper introduces a method for achieving stable periodic walking for legged robots. This method is based on producing a type of odd-even symmetry in the system. A hybrid system with such symmetries is called a symmetric hybrid system (SHS). We discuss the properties of an SHS and, in particular, will show that an SHS can have an infinite number of synchronized periodic orbits. We describe how controllers can be obtained to make a legged robot an SHS. Then the stability of the synchronized periodic orbits of this SHS is studied, where the notion of self-synchronization is introduced. We show that such self-synchronized periodic orbits are neutrally stable in kinetic energy. As the final step in the process of achieving asymptotically stable periodic walking, we show how by introducing asymmetries (such as energy loss at impact) in the system, the synchronized periodic orbits of this SHS can be turned into asymptotically stable periodic orbits. Many numerical examples, including an 8-DOF 3D biped with 2 degrees of underactuation, are studied to demonstrate the effectiveness of the method.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13
Fig. 14
Fig. 15
Fig. 16
Fig. 17
Fig. 18
Fig. 19

Similar content being viewed by others

References

  • Altendorfer, R., Koditschek, D. E., & Holmes, P. (2004). Stability analysis of legged locomotion models by symmetry-factored return maps. The International Journal of Robotics Research, 23(10–11), 979–999.

    Article  Google Scholar 

  • Byrnes, C. I., & Isidori, A. (1985). Global feedback stabilization of nonlinear systems. In 24th IEEE conference on decision and control 1985 (pp. 1031–1037).

  • Chevallereau, C., Djoudi, D., & Grizzle, J. W. (2008). Stable bipedal walking with foot rotation through direct regulation of the zero moment point. IEEE Transactions on Robotics, 24(2), 390–401.

    Article  Google Scholar 

  • Chevallereau, C., Gabriel, A., Aoustin, Y., Plestan, F., Westervelt, E., de Wit, C. C., et al. (2003). Rabbit: A testbed for advanced control theory. IEEE Control Systems Magazine, 23(5), 57–79.

    Article  Google Scholar 

  • Chevallereau, C., Grizzle, J. W., & Shih, C.-L. (2009). Asymptotically stable walking of a five-link underactuated 3-d bipedal robot. IEEE Transactions on Robotics, 25(1), 37–50.

    Article  Google Scholar 

  • Collins, S., Ruina, A., Tedrake, R., & Wisse, M. (2005). Efficient bipedal robots based on passive-dynamic walkers. Science, 307(5712), 1082–1085.

    Article  Google Scholar 

  • Collins, S. H., Wisse, M., & Ruina, A. (2001). A three-dimensional passive-dynamic walking robot with two legs and knees. The International Journal of Robotics Research, 20(7), 607–615.

    Article  Google Scholar 

  • Dingwell, J. B., & Kang, H. G. (2007). Differences between local and orbital dynamic stability during human walking. Journal of Biomechanical Engineering, 129(4), 586–593.

    Article  Google Scholar 

  • Garcia, M., Chatterjee, A., Ruina, A., & Coleman, M. (1998). The simplest walking model: Stability, complexity, scaling. Journal of Biomechanical Engineering, 120(2), 281–288.

    Article  Google Scholar 

  • Geng, T., Porr, B., & Worgotter, F. (2006). Fast biped walking with a sensor-driven neuronal controller and real-time online learning. The International Journal of Robotics Research, 25(3), 243–259.

    Article  Google Scholar 

  • Goswami, A., Espiau, B., & Keramane, A. (1996). Limit cycles and their stability in a passive bipedal gait. In Proceedings of the 1996 IEEE international conference on robotics and automation (Vol. 1, pp. 246–251).

  • Goswami, A., Espiau, B., & Keramane, A. (1997). Limit cycles in a passive compass gait biped and passivity-mimicking control laws. Autonomous Robots, 4(3), 273–286.

    Article  Google Scholar 

  • Gregg, R. D., & Righetti, L. (2013). Controlled reduction with unactuated cyclic variables: Application to 3d bipedal walking with passive yaw rotation. IEEE Transactions on Automatic Control, 58(10), 2679–2685.

    Article  MathSciNet  Google Scholar 

  • Grizzle, J. W., Abba, G., & Plestan, F. (2001). Asymptotically stable walking for biped robots: Analysis via systems with impulse effects. IEEE Transactions on Automatic Control, 46(1), 51–64.

    Article  MathSciNet  MATH  Google Scholar 

  • Grizzle, J. W., Chevallereau, C., & Shih, C.-L. (2008). Hzd-based control of a five-link underactuated 3d bipedal robot. In 47th IEEE conference on decision and control, 2008. CDC 2008, (pp. 5206–5213).

  • Hamed, K. A., Buss, B. G., & Grizzle, J. W. (2014). Continuous-time controllers for stabilizing periodic orbits of hybrid systems: Application to an underactuated 3d bipedal robot. In 53rd IEEE conference of decision and control.

  • Holmes, P., Full, R. J., Koditschek, D., & Guckenheimer, J. (2006). The dynamics of legged locomotion: Models, analyses, challenges. Siam Review, 48(2), 207–304.

    Article  MathSciNet  MATH  Google Scholar 

  • Hurmuzlu, Y., & Marghitu, D. B. (1994). Rigid body collisions of planar kinematic chains with multiple contact points. The International Journal of Robotics Research, 13(1), 82–92.

    Article  Google Scholar 

  • Isidori, A. (1995). Nonlinear control systems. New York: Springer.

    Book  MATH  Google Scholar 

  • Kajita, S., Kanehiro, F., Kaneko, K., Yokoi, K., & Hirukawa, H. (2001). The 3d linear inverted pendulum mode: A simple modeling for a biped walking pattern generation. In Proceedings of 2001 IEEE/RSJ international conference on intelligent robots and systems, (Vol. 1, pp. 239–246).

  • Kuo, A. D. (1999). Stabilization of lateral motion in passive dynamic walking. The International Journal of Robotics Research, 18(9), 917–930.

    Article  Google Scholar 

  • Lee, J. M. (2003). Introduction to smooth manifolds. New York: Springer.

    Book  Google Scholar 

  • McGeer, T. (1990). Passive dynamic walking. The International Journal of Robotics Research, 9(2), 62–82.

    Article  Google Scholar 

  • Merker, A., Kaiser, D., Seyfarth, A., & Hermann, M. (2015). Stable running with asymmetric legs: A bifurcation approach. International Journal of Bifurcation and Chaos, 25(11), 1550152.

    Article  MathSciNet  MATH  Google Scholar 

  • Merker, A., Rummel, J., & Seyfarth, A. (2011). Stable walking with asymmetric legs. Bioinspiration & Biomimetics, 6(4), 045004.

    Article  Google Scholar 

  • Peuker, F., Maufroy, C., & Seyfarth, A. (2012). Leg-adjustment strategies for stable running in three dimensions. Bioinspiration & Biomimetics, 7(3), 036002.

    Article  Google Scholar 

  • Raibert, M. H. (1984). Hopping in legged systems, modeling and simulation for the two-dimensional one-legged case. IEEE Transactions on Systems, Man and Cybernetics, 3, 451–463.

    Article  Google Scholar 

  • Raibert, M. H. (1986a). Legged robots that balance (Vol. 3). Cambridge, MA: MIT Press.

    MATH  Google Scholar 

  • Raibert, M. H. (1986b). Symmetry in running. Science, 231(4743), 1292–1294.

    Article  Google Scholar 

  • Ramezani, A., Hurst, J. W., Hamed, K. A., & Grizzle, J. W. (2014). Performance analysis and feedback control of atrias, a three-dimensional bipedal robot. Journal of Dynamic Systems, Measurement, Control, 136(2), 021012.

    Article  Google Scholar 

  • Razavi, H., Bloch, A. M., Chevallereau, C., & Grizzle J. W. (2015). Restricted discrete invariance and self-synchronization for stable walking of bipedal robots. In American control conference (ACC), 2015, IEEE, (pp.4818–4824).

  • Rezazadeh, S., Hubicki, C., Jones, M., Peekema, A., Van Why, J., Abate, A., et al. (2015). Spring-mass walking with atrias in 3d: Robust gait control spanning zero to 4.3 kph on a heavily underactuated bipedal robot. In ASME 2015 dynamic systems and control conference (pp. V001T04A003–V001T04A003). American Society of Mechanical Engineers.

  • Seipel, J. E., & Holmes, P. (2005). Running in three dimensions: Analysis of a point-mass sprung-leg model. The International Journal of Robotics Research, 24(8), 657–674.

    Article  Google Scholar 

  • Seyfarth, A., Geyer, H., Gunther, M., & Blickhan, R. (2002). A movement criterion for running. Journal of Biomechanics, 35(5), 649–655.

    Article  Google Scholar 

  • Westervelt, E. R., Grizzle, J. W., hevallereau, C., Choi, J. H., & Morris, B. (2007). Feedback control of dynamic bipedal robot locomotion. London: Taylor & Francis/CRC.

    Book  Google Scholar 

  • Wisse, M., Schwab, A. L., Van der Linde, R. Q., & van der Helm, F. C. (2005). How to keep from falling forward: Elementary swing leg action for passive dynamic walkers. IEEE Transactions on Robotics, 21(3), 393–401.

    Article  Google Scholar 

Download references

Acknowledgments

We gratefully acknowledge partial support by NSF INSPIRE 1343720, NSF DMS 1613819, and Simons Foundations. Hamed Razavi was supported in part by Rackham Summer Fellowship.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Hamed Razavi.

Appendices

Appendix 1

Proposition 32

Suppose that the SHS has a synchronized solution \((\varPhi (t), \varPsi (t))\) with initial conditions \((\varPhi (0), \varPsi (0)) = (-\phi _0, \psi _0)\) and \(({\dot{\varPhi }}(0), {\dot{\varPsi }}(0)) = ({\dot{\varPhi }}_0, {\dot{\varPsi }}_0)\). Suppose that \(t = t_m\) is the time for which \(\varPhi (t_m) = 0\) and \({\dot{\varPsi }}(t_m) = 0\) such that \({\dot{\varPhi }}(t_m) \ne 0\) . Let \(\chi ({\dot{\phi }}_0,{\dot{\psi }}_0,t)\) be the flow map of the \((\phi _0,\psi _0)\)-invariant SHS. Therefore, \(\chi ({\dot{\phi }}_0,{\dot{\psi }}_0,0) = (-\phi _0,\psi _0,{\dot{\phi }}_0,{\dot{\psi }}_0)\) for every \(({\dot{\phi }}_0,{\dot{\psi }}_0) \in {\mathcal {T}}_{(-\phi _0,\psi _0)}{\mathcal {Q}}\). Let \(\chi = (\phi ,\psi ,{\dot{\phi }},{\dot{\psi }})\). Since, \((\varPhi (t), \varPsi (t))\) is synchronized,

$$\begin{aligned} \phi ({\dot{\varPhi }}_0, {\dot{\varPsi }}_0, t_m) = 0,~ {\dot{\psi }}({\dot{\varPhi }}_0, {\dot{\varPsi }}_0, t_m) = 0. \end{aligned}$$

If the Jacobian of \((\phi ,{\dot{\psi }})\) with respect to \(({\dot{\phi }}_0,{\dot{\psi }}_0)\) is invertible at \(({\dot{\varPhi }}_0, {\dot{\varPsi }}_0, t_m)\), then there exists a smooth function \(L:{\mathcal {T}}_{(-\phi _0,\psi _0)}{\mathcal {Q}}\rightarrow {\mathbb {R}}^{m+n-1}\) such that if \(L({\dot{\phi }}_0, {\dot{\psi }}_0) = 0\), the solution starting from \((-\phi _0,\psi _0)\) with initial velocity \(({\dot{\phi }}_0, {\dot{\psi }}_0)\) is synchronized.

Proof

Based on the definition of a synchronized solution, we are interested in the values of \(({\dot{\phi }}_0, {\dot{\psi }}_0, t)\) for which

$$\begin{aligned} \phi ({\dot{\phi }}_0,{\dot{\psi }}_0, t) = 0,~ {\dot{\psi }}({\dot{\phi }}_0,{\dot{\psi }}_0, t) = 0. \end{aligned}$$

Since \(({\dot{\varPhi }}_0, {\dot{\varPsi }}_0, t_m)\) is a solution to this system, we have

$$\begin{aligned} \phi ({\dot{\varPhi }}_0, {\dot{\varPsi }}_0, t_m) = 0, ~{\dot{\psi }}({\dot{\varPhi }}_0, {\dot{\varPsi }}_0, t_m) = 0. \end{aligned}$$

Because the Jacobian of \((\phi ,{\dot{\psi }})\) with respect to \(({\dot{\phi }}_0,{\dot{\psi }}_0)\) is invertible at \(({\dot{\varPhi }}_0, {\dot{\varPsi }}_0, t_m)\), by implicit function theorem there exists a smooth function F defined in a neighborhood of \(t_m\) such that in this neighborhood

$$\begin{aligned} \phi (F(t),t) = 0, ~{\dot{\psi }}(F(t),t) = 0. \end{aligned}$$
(60)

The function \(t\mapsto F(t)\) defines a smooth curve in the manifold \({\mathcal {T}}_{(-\phi _0,\psi _0)}{\mathcal {Q}}\), where \(F(t_m) = ({\dot{\varPhi }}_0, {\dot{\varPsi }}_0)\). Differentiating the two equations in (60) with respect to t at \(t = t_m\), since \({\dot{\varPhi }}(t_m) \ne 0\), we can show that \({\dot{F}}(t_m) \ne 0\). Therefore, the parametrization \(t\mapsto F(t)\) is a regular parametrization in a neighborhood of \(t_m\), and, hence, the image of \(t\mapsto F(t)\) defines an embedded 1-dimensional submanifold \({\mathcal {K}}\) of \({\mathcal {T}}_{(-\phi _0,\psi _0)}{\mathcal {Q}}\) (Lee 2003). Thus, there exists a smooth function \(L:{\mathcal {T}}_{(-\phi _0,\psi _0)}{\mathcal {Q}}\rightarrow {\mathbb {R}}^{m+n-1}\) with rank \(m+n-1\) such that

$$\begin{aligned} {\mathcal {K}} = \{({\dot{\phi }}_0,{\dot{\psi }}_0) \in {\mathcal {T}}_{(-\phi _0,\psi _0)}{\mathcal {Q}} | L({\dot{\phi }}_0,{\dot{\psi }}_0) = 0\}. \end{aligned}$$

Consequently, if \(L({\dot{\phi }}_0, {\dot{\psi }}_0) = 0\), the solution starting from \((-\phi _0,\psi _0)\) with initial velocity \(({\dot{\phi }}_0, {\dot{\psi }}_0)\) is synchronized. Function L is called the synchronization measure at \((-\phi _0, \psi _0)\) for the SHS.\(\square \)

Proposition 33

Let \({\mathcal {Q}}\) denote the configuration space of the \((x_0, y_0)\)-invariant 3D LIP biped. Assume that the switching surface is

$$\begin{aligned} {\mathcal {S}} = \{(x,y,{\dot{x}},{\dot{y}}) | h(x,y) = h(x_0, y_0) \}, \end{aligned}$$

where \(h:Q\rightarrow {\mathbb {R}}\) is a smooth function and \(\partial h/\partial y (x_0, y_0)\ne 0\). Let \(({\dot{x}}_*, {\dot{y}}_*)\) be a fixed point of the restricted Poincaré map \(P: {\mathcal {T}}_{(-x_0, y_0)}{\mathcal {Q}} \rightarrow {\mathcal {T}}_{(-x_0, y_0)}{\mathcal {Q}}\). The eigenvalues of P at \(({\dot{x}}_*, {\dot{y}}_*)\) are \(\{\lambda , 1\}\) with

$$\begin{aligned} \lambda = -1+\frac{2\omega ^2(y_0^2-Cx_0^2)}{CE_x^*-E_y^*}, \end{aligned}$$
(61)

where

$$\begin{aligned} C =\frac{y_0}{x_0} \left( \frac{\partial h}{\partial y}(x_0, y_0)\right) ^{-1} \frac{\partial h}{\partial x}(x_0, y_0), \end{aligned}$$

and

$$\begin{aligned} E_x^* = {\dot{x}}_*^2-\omega ^2 x_0^2, ~ E_y^* = {\dot{y}}_*^2 - \omega ^2 y_0^2 \end{aligned}$$

are the orbital energies in the x and y directions.

Proof

Let \(L:{\mathcal {T}}_{(-x_0, y_0)}{\mathcal {Q}} \rightarrow {\mathbb {R}}\) be the synchronization measure of the 3D LIP. By Proposition 15, \(\lambda = \partial L_1/\partial L_0(0,K^*)\), where \(K^* = (1/2) (({\dot{x}}_*)^2+({\dot{y}}_*)^2)\). Assume that \((-x_0, y_0, {\dot{x}}_*+\delta {\dot{x}}_0, {\dot{y}}_*+\delta {\dot{y}}_0)\) is the initial state of a solution of the system at the beginning of the step, and let \((-x_0, y_0, {\dot{x}}_*+\delta {\dot{x}}_1, {\dot{y}}_*+\delta {\dot{y}}_1)\) be its state at the beginning of the next step. Define \(L_0 = L({\dot{x}}_*+\delta {\dot{x}}_0, {\dot{y}}_*+\delta {\dot{y}}_0)\) and \(L_1 = L({\dot{x}}_*+\delta {\dot{x}}_1, {\dot{y}}_*+\delta {\dot{y}}_1)\). We have

$$\begin{aligned} \lambda = \lim _{L_0 \rightarrow 0} \frac{L_1}{L_0}. \end{aligned}$$

Denote the state of the system right before the transition by \((x_0 + \delta x_1 , y_0 + \delta y_1, {\dot{x}}_*+ \delta {\dot{x}}_1, -({\dot{y}}_*+\delta {\dot{y}}_1))\). Since \(L = {\dot{x}}{\dot{y}}-\omega ^2 xy\) is a conserved quantity for the 3D LIP in the continuous phase of motion, we have

$$\begin{aligned} L_0 = -({\dot{x}}_*+ \delta {\dot{x}}_1)({\dot{y}}_*+\delta {\dot{y}}_1) -\omega ^2 (x_0+\delta x_1)(y_0+\delta y_1). \end{aligned}$$

Moreover, since the system is \((x_0,y_0)\)-invariant,

$$\begin{aligned} L_1 = ({\dot{x}}_*+ \delta {\dot{x}}_1)({\dot{y}}_*+\delta {\dot{y}}_1)+\omega ^2 x_0y_0. \end{aligned}$$

Adding this equation to the previous one,

$$\begin{aligned} L_1+L_0 = -\omega ^2 (x_0\delta y_1 + y_0 \delta x_1). \end{aligned}$$
(62)

By definition of the switching surface, \(h(x_0+\delta x_1, y_0+\delta y_1) = h(x_0, y_0)\), from which,

$$\begin{aligned} \frac{\partial h}{\partial x}(x_0, y_0) \delta x_1 = - \frac{\partial h}{\partial y}(x_0, y_0) \delta y_1. \end{aligned}$$
(63)

By definition of C, \(\delta y_1 = -C\frac{x_0}{y_0} \delta x_1\). Substituting this into Eq. (62) results in

$$\begin{aligned} L_1+L_0 = -\omega ^2 \left( -C\frac{x_0^2}{y_0}+ y_0\right) \delta x_1. \end{aligned}$$

Therefore, from the equation above,

$$\begin{aligned} \lim _{L_0\rightarrow 0} \frac{L_1}{L_0} = -1-\omega ^2 \left( -C\frac{x_0^2}{y_0}+ y_0\right) \lim _{L_0 \rightarrow 0} \frac{\delta x_1}{L_0}. \end{aligned}$$
(64)

Thus, to find the limit on the left-hand side we need only find the limit on the right-hand side. Since in the continuous phase of motion the orbital energies, \({\dot{x}}^2-\omega ^2 x^2\) and \({y}^2-\omega ^2 y^2\), are conserved quantities, we have

$$\begin{aligned} \left( {\dot{x}}_*+\delta {\dot{x}}_0\right) ^2-\omega ^2 x_0= & {} \left( {\dot{x}}_*+\delta {\dot{x}}_1\right) ^2-\omega ^2 \left( x_0+\delta x_1\right) ,\\ \left( {\dot{y}}_*+\delta {\dot{y}}_0\right) ^2-\omega ^2 y_0= & {} \left( {\dot{y}}_*+\delta {\dot{y}}_1\right) ^2-\omega ^2 \left( y_0+\delta y_1\right) . \end{aligned}$$

From these two equations and definition of C,

$$\begin{aligned} {\dot{x}}_*\left( \delta {\dot{x}}_1-\delta {\dot{x}}_0\right) = \omega ^2 \delta x_1, ~{\dot{y}}_*\left( \delta {\dot{y}}_1-\delta {\dot{y}}_0\right) = -C\omega ^2 \frac{x_0}{y_0} \delta x_1.\nonumber \\ \end{aligned}$$
(65)

Since \(L = {\dot{x}}{\dot{y}}-\omega ^2 xy\) is a conserved quantity in the continuous phase of the motion, we can write \(L_0\) in terms of the states at the beginning of step, that is, \((-x_0, y_0, {\dot{x}}_*+\delta {\dot{x}}_0, {\dot{y}}_*+{\dot{y}}_0)\) or at end of the step, that is, \((x_0+\delta x_1, y_0+\delta y_1, {\dot{x}}_*+\delta {\dot{x}}_1, {\dot{y}}_*+{\dot{y}}_1)\). Hence,

$$\begin{aligned} L_0= & {} \left( {\dot{x}}_*+\delta {\dot{x}}_0\right) \left( {\dot{y}}_*+\delta {\dot{y}}_0\right) +\omega ^2 x_0 y_0,\\ L_0= & {} -\left( {\dot{x}}_*+ \delta {\dot{x}}_1\right) \left( {\dot{y}}_*+\delta {\dot{y}}_1\right) -\omega ^2 \left( x_0+\delta x_1\right) \left( y_0+\delta y_1\right) . \end{aligned}$$

From (65) and the two equations above,

$$\begin{aligned} \frac{2L_0}{\delta x_1} = -\omega ^2 \frac{x_0y_0\left( -C{\dot{x}}_*^2+{\dot{y}}_*^2\right) +\left( {\dot{x}}_*{\dot{y}}_*\right) \left( -Cx_0^2+y_0^2\right) }{y_0{\dot{x}}_*{\dot{y}}_*}. \end{aligned}$$

Substituting this into Eq. (64), we have

$$\begin{aligned} \lim _{L_0\rightarrow 0}\frac{L_1}{L_0}= & {} -1+ \frac{1}{\omega ^2}\\&\cdot&\frac{2\left( -Cx_0^2+ y_0^2\right) \left( {\dot{x}}_*{\dot{y}}_*\right) }{x_0y_0\left( -C{\dot{x}}_*^2+{\dot{y}}_*^2\right) +\left( {\dot{x}}_*{\dot{y}}_*\right) \left( -Cx_0^2+y_0^2\right) }. \end{aligned}$$

The limit on the left-hand side is \(\lambda \). Since \(L({\dot{x}}_*,{\dot{y}}_*) = 0\), we have \({\dot{x}}_*{\dot{y}}_* = -\omega ^2 x_0 y_0\). Therefore, if we replace \({\dot{x}}_*{\dot{y}}_* \) with \(-\omega ^2 x_0 y_0\) in the equation above, we obtain

$$\begin{aligned} \lambda= & {} -1+\frac{1}{\omega ^2}\\&\cdot \frac{2\left( -Cx_0^2+ y_0^2\right) \left( -\omega ^2 x_0y_0\right) }{x_0y_0\left( -C{\dot{x}}_*^2+{\dot{y}}_*^2\right) +\left( -\omega ^2 x_0 y_0\right) \left( -Cx_0^2+y_0^2\right) }. \end{aligned}$$

After simplification

$$\begin{aligned} \lambda =-1+\frac{2\omega ^2\left( y_0^2-Cx_0^2\right) }{C{\dot{x}}_*^2-{\dot{y}}_*^2+\omega ^2\left( y_0^2-Cx_0^2\right) }. \end{aligned}$$

By definition of \(E_x^*\) and \(E_y^*\) this equation is equivalent to Eq. (61).\(\square \)

Corollary 34

In Proposition 33, if \(h(x,y) = x^2+a^2y^2\) then

$$\begin{aligned} \lambda = -1+\frac{2\omega ^2\left( a^2y_0^2-x_0^2\right) }{E_x^*-a^2E_y^*}. \end{aligned}$$

In particular, if \(a = 1\),

$$\begin{aligned} \lambda = -1+\frac{2\omega ^2\left( y_0^2-x_0^2\right) }{E_x^*-E_y^*}. \end{aligned}$$

Appendix 2 Proof of Proposition 20

Lemma 35

Let \(\varSigma = (X, {\mathcal {Q}})\) be a second order hybrid system and \((\phi , \psi )\) be a local coordinate system defined on a symmetric neighborhood \({\mathcal {N}} \subset {\mathcal {Q}}\). Suppose that \((\phi , \psi , {v}_\phi , {v}_\psi )\) is a coordinate system defined on \({\mathcal {TN}}\), the tangent bundle of \({\mathcal {N}}\), and assume that \(G: {\mathcal {TN}} \rightarrow {\mathcal {TN}}\), which maps \((\phi , \psi , v_{\phi }, v_{\psi })\) to \((-\phi , \psi , v_{\phi }, -v_{\psi })\), is a well-defined function. If

$$\begin{aligned} {\dot{\phi }}(-\phi , \psi , v_{\phi }, -v_{\psi })= & {} {\dot{\phi }}(\phi , \psi , v_{\phi }, v_{\psi }),\nonumber \\ {\dot{\psi }}(-\phi , \psi , v_{\phi }, -v_{\psi })= & {} -{\dot{\psi }}(\phi , \psi , v_{\phi }, v_{\psi }), \end{aligned}$$
(66)

and the vector field X satisfies the equality

$$\begin{aligned} X \circ G = - \mathrm {d} G \cdot X, \end{aligned}$$
(67)

where \(\mathrm {d} G\) is the Jacobian of G, then \(\varSigma \) is a symmetric system in the coordinates \((\phi ,\psi )\).

Proof

From (66), in the coordinates \((\phi , \psi , {\dot{\phi }}, {\dot{\psi }})\), function G maps \((\phi , \psi , {\dot{\phi }}, {\dot{\psi }})\) to \((-\phi , \psi , {\dot{\phi }}, -{\dot{\psi }})\). Let \(X = (X_q, X_{{\dot{q}}})\), and suppose that in the coordinates \((\phi , \psi ,{\dot{\phi }}, {\dot{\psi }})\), \(X_{{\dot{q}}} = (f, g)\) for smooth functions f and g. We need to show that f and g satisfy Eqs. (8) and (9). However, these equations immediately follow by writing (67) in the coordinates \((\phi , \psi , {\dot{\phi }}, {\dot{\psi }})\).\(\square \)

The above lemma is a particular example of what is called time reversal symmetry in (Altendorfer et al. 2004).

Proof

(Proof of Proposition 20) Let \({\varvec{r}}_G\) denote the position vector of the COM of the 8-DOF 3D Biped, and let \({\varvec{H}}\) denote the total angular momentum of the system about the point of contact. We have

$$\begin{aligned} \frac{d {\varvec{H}}}{dt} = M{\varvec{r}}_G\times {\varvec{g}}, \end{aligned}$$
(68)

where M is the total mass of the biped, and \({\varvec{g}}\) is the vector of gravity. Let \((x_G, y_G, z_G)\) denote the coordinates of \({\varvec{r}}_G\) in a cartesian coordinate system attached to the inertial frame I, centered at the point of contact. Assume that \(H_x\) is the second component of \({\varvec{H}}\) and \(H_y\) is its first component (Normally, one might assign \(H_x\) to the first component and \(H_y\) to the second component; however, this choice may cause confusion with the notation of Lemma 35. To avoid this confusion, we assign \(H_x\) to the second component of \({\varvec{H}}\) and \(H_y\) to its first component).

One can check that \((x_G, y_G, H_x, H_y)\) is a coordinate system for the zero dynamics manifold and the coordinate systems \((x_G, y_G, H_x, H_y)\) and \((x_G, y_G, {\dot{x}}_G, {\dot{y}}_G)\) satisfy (66). As a result, there exist smooth one-to-one functions \(f_x\) and \(f_y\) such that on the zero dynamics manifold,

$$\begin{aligned} {\dot{x}}_G= & {} f_x(x_G, y_G, H_x, H_y),\nonumber \\ {\dot{y}}_G= & {} f_y(x_G, y_G, H_x, H_y), \end{aligned}$$
(69)

and

$$\begin{aligned} f_x(-x_G, y_G, H_x, -H_y)= & {} f_x(x_G, y_G, H_x, H_y),\nonumber \\ f_y(-x_G, y_G, H_x, -H_y)= & {} -f_y(x_G, y_G, H_x, H_y). \end{aligned}$$
(70)

Moreover, from Eq. (68),

$$\begin{aligned} {\dot{H}}_x= & {} Mg x_G,\nonumber \\ {\dot{H}}_y= & {} -Mg y_G. \end{aligned}$$
(71)

Equation (69) together with Eq. (71) define the equations of motion on the zero dynamics. Therefore, if \(X = (f_x, f_y, Mgx_G, -Mgy_G)\) and \(z = (x_G, y_G, H_x, H_y)\), then the system \({\dot{z}} = X(z)\) defines the equations of motion on the zero dynamics manifold. Hence, to show that this system is symmetric, according to Lemma 35, if G is the function that maps \((x_G, y_G, H_x, H_y)\) to \((-x_G, y_G, H_x, -H_y)\), it suffices to show that \(X \circ G = - \mathrm {d} G \cdot X\). It immediately follows that

$$\begin{aligned} X \circ G (x) = \left[ \begin{array}{c} f_x(-x_G, y_G, H_x, -H_y)\\ f_y(-x_G, y_G, H_x, -H_y)\\ -Mgx_G\\ -Mgy_G \end{array} \right] , \end{aligned}$$

and

$$\begin{aligned} \mathrm {d}G \cdot X(x)= \left[ \begin{array}{c} -f_x(x_G, y_G, H_x, H_y)\\ f_y(-x_G, y_G, H_x, H_y)\\ Mgx_G\\ Mgy_G \end{array} \right] . \end{aligned}$$

However, the above two equations together with Eq. (70) prove that \(X \circ G = - \mathrm {d} G \cdot X\). Hence, the zero dynamics is a symmetric system in the coordinates \((x_G, y_G)\).

Finally, from kinematics equations on the zero dynamics manifold, the coordinate systems \((q_1, q_2)\) and \((x_G, y_G)\) are equivalent in the sense of Proposition 6. As a result, the zero dynamics is a symmetric system in the coordinates \((q_1, q_2)\) as well.\(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Razavi, H., Bloch, A.M., Chevallereau, C. et al. Symmetry in legged locomotion: a new method for designing stable periodic gaits. Auton Robot 41, 1119–1142 (2017). https://doi.org/10.1007/s10514-016-9593-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10514-016-9593-x

Keywords

Navigation