Differential dynamic programming

Algorithm for trajectory optimization

Differential dynamic programming (DDP) is an optimal control algorithm of the trajectory optimization class. The algorithm was introduced in 1966 by Mayne[1] and subsequently analysed in Jacobson and Mayne's eponymous book.[2] The algorithm uses locally-quadratic models of the dynamics and cost functions, and displays quadratic convergence. It is closely related to Pantoja's step-wise Newton's method.[3][4]

Finite-horizon discrete-time problems

The dynamics

x i + 1 = f ( x i , u i ) {\displaystyle \mathbf {x} _{i+1}=\mathbf {f} (\mathbf {x} _{i},\mathbf {u} _{i})} (1)

describe the evolution of the state x {\displaystyle \textstyle \mathbf {x} } given the control u {\displaystyle \mathbf {u} } from time i {\displaystyle i} to time i + 1 {\displaystyle i+1} . The total cost J 0 {\displaystyle J_{0}} is the sum of running costs {\displaystyle \textstyle \ell } and final cost f {\displaystyle \ell _{f}} , incurred when starting from state x {\displaystyle \mathbf {x} } and applying the control sequence U { u 0 , u 1 , u N 1 } {\displaystyle \mathbf {U} \equiv \{\mathbf {u} _{0},\mathbf {u} _{1}\dots ,\mathbf {u} _{N-1}\}} until the horizon is reached:

J 0 ( x , U ) = i = 0 N 1 ( x i , u i ) + f ( x N ) , {\displaystyle J_{0}(\mathbf {x} ,\mathbf {U} )=\sum _{i=0}^{N-1}\ell (\mathbf {x} _{i},\mathbf {u} _{i})+\ell _{f}(\mathbf {x} _{N}),}

where x 0 x {\displaystyle \mathbf {x} _{0}\equiv \mathbf {x} } , and the x i {\displaystyle \mathbf {x} _{i}} for i > 0 {\displaystyle i>0} are given by Eq. 1. The solution of the optimal control problem is the minimizing control sequence U ( x ) argmin U J 0 ( x , U ) . {\displaystyle \mathbf {U} ^{*}(\mathbf {x} )\equiv \operatorname {argmin} _{\mathbf {U} }J_{0}(\mathbf {x} ,\mathbf {U} ).} Trajectory optimization means finding U ( x ) {\displaystyle \mathbf {U} ^{*}(\mathbf {x} )} for a particular x 0 {\displaystyle \mathbf {x} _{0}} , rather than for all possible initial states.

Dynamic programming

Let U i {\displaystyle \mathbf {U} _{i}} be the partial control sequence U i { u i , u i + 1 , u N 1 } {\displaystyle \mathbf {U} _{i}\equiv \{\mathbf {u} _{i},\mathbf {u} _{i+1}\dots ,\mathbf {u} _{N-1}\}} and define the cost-to-go J i {\displaystyle J_{i}} as the partial sum of costs from i {\displaystyle i} to N {\displaystyle N} :

J i ( x , U i ) = j = i N 1 ( x j , u j ) + f ( x N ) . {\displaystyle J_{i}(\mathbf {x} ,\mathbf {U} _{i})=\sum _{j=i}^{N-1}\ell (\mathbf {x} _{j},\mathbf {u} _{j})+\ell _{f}(\mathbf {x} _{N}).}

The optimal cost-to-go or value function at time i {\displaystyle i} is the cost-to-go given the minimizing control sequence:

V ( x , i ) min U i J i ( x , U i ) . {\displaystyle V(\mathbf {x} ,i)\equiv \min _{\mathbf {U} _{i}}J_{i}(\mathbf {x} ,\mathbf {U} _{i}).}

Setting V ( x , N ) f ( x N ) {\displaystyle V(\mathbf {x} ,N)\equiv \ell _{f}(\mathbf {x} _{N})} , the dynamic programming principle reduces the minimization over an entire sequence of controls to a sequence of minimizations over a single control, proceeding backwards in time:

V ( x , i ) = min u [ ( x , u ) + V ( f ( x , u ) , i + 1 ) ] . {\displaystyle V(\mathbf {x} ,i)=\min _{\mathbf {u} }[\ell (\mathbf {x} ,\mathbf {u} )+V(\mathbf {f} (\mathbf {x} ,\mathbf {u} ),i+1)].} (2)

This is the Bellman equation.

Differential dynamic programming

DDP proceeds by iteratively performing a backward pass on the nominal trajectory to generate a new control sequence, and then a forward-pass to compute and evaluate a new nominal trajectory. We begin with the backward pass. If

( x , u ) + V ( f ( x , u ) , i + 1 ) {\displaystyle \ell (\mathbf {x} ,\mathbf {u} )+V(\mathbf {f} (\mathbf {x} ,\mathbf {u} ),i+1)}

is the argument of the min [ ] {\displaystyle \min[\cdot ]} operator in Eq. 2, let Q {\displaystyle Q} be the variation of this quantity around the i {\displaystyle i} -th ( x , u ) {\displaystyle (\mathbf {x} ,\mathbf {u} )} pair:

Q ( δ x , δ u ) ( x + δ x , u + δ u ) + V ( f ( x + δ x , u + δ u ) , i + 1 ) ( x , u ) V ( f ( x , u ) , i + 1 ) {\displaystyle {\begin{aligned}Q(\delta \mathbf {x} ,\delta \mathbf {u} )\equiv &\ell (\mathbf {x} +\delta \mathbf {x} ,\mathbf {u} +\delta \mathbf {u} )&&{}+V(\mathbf {f} (\mathbf {x} +\delta \mathbf {x} ,\mathbf {u} +\delta \mathbf {u} ),i+1)\\-&\ell (\mathbf {x} ,\mathbf {u} )&&{}-V(\mathbf {f} (\mathbf {x} ,\mathbf {u} ),i+1)\end{aligned}}}

and expand to second order

1 2 [ 1 δ x δ u ] T [ 0 Q x T Q u T Q x Q x x Q x u Q u Q u x Q u u ] [ 1 δ x δ u ] {\displaystyle \approx {\frac {1}{2}}{\begin{bmatrix}1\\\delta \mathbf {x} \\\delta \mathbf {u} \end{bmatrix}}^{\mathsf {T}}{\begin{bmatrix}0&Q_{\mathbf {x} }^{\mathsf {T}}&Q_{\mathbf {u} }^{\mathsf {T}}\\Q_{\mathbf {x} }&Q_{\mathbf {x} \mathbf {x} }&Q_{\mathbf {x} \mathbf {u} }\\Q_{\mathbf {u} }&Q_{\mathbf {u} \mathbf {x} }&Q_{\mathbf {u} \mathbf {u} }\end{bmatrix}}{\begin{bmatrix}1\\\delta \mathbf {x} \\\delta \mathbf {u} \end{bmatrix}}} (3)

The Q {\displaystyle Q} notation used here is a variant of the notation of Morimoto where subscripts denote differentiation in denominator layout.[5] Dropping the index i {\displaystyle i} for readability, primes denoting the next time-step V V ( i + 1 ) {\displaystyle V'\equiv V(i+1)} , the expansion coefficients are

Q x = x + f x T V x Q u = u + f u T V x Q x x = x x + f x T V x x f x + V x f x x Q u u = u u + f u T V x x f u + V x f u u Q u x = u x + f u T V x x f x + V x f u x . {\displaystyle {\begin{alignedat}{2}Q_{\mathbf {x} }&=\ell _{\mathbf {x} }+\mathbf {f} _{\mathbf {x} }^{\mathsf {T}}V'_{\mathbf {x} }\\Q_{\mathbf {u} }&=\ell _{\mathbf {u} }+\mathbf {f} _{\mathbf {u} }^{\mathsf {T}}V'_{\mathbf {x} }\\Q_{\mathbf {x} \mathbf {x} }&=\ell _{\mathbf {x} \mathbf {x} }+\mathbf {f} _{\mathbf {x} }^{\mathsf {T}}V'_{\mathbf {x} \mathbf {x} }\mathbf {f} _{\mathbf {x} }+V_{\mathbf {x} }'\cdot \mathbf {f} _{\mathbf {x} \mathbf {x} }\\Q_{\mathbf {u} \mathbf {u} }&=\ell _{\mathbf {u} \mathbf {u} }+\mathbf {f} _{\mathbf {u} }^{\mathsf {T}}V'_{\mathbf {x} \mathbf {x} }\mathbf {f} _{\mathbf {u} }+{V'_{\mathbf {x} }}\cdot \mathbf {f} _{\mathbf {u} \mathbf {u} }\\Q_{\mathbf {u} \mathbf {x} }&=\ell _{\mathbf {u} \mathbf {x} }+\mathbf {f} _{\mathbf {u} }^{\mathsf {T}}V'_{\mathbf {x} \mathbf {x} }\mathbf {f} _{\mathbf {x} }+{V'_{\mathbf {x} }}\cdot \mathbf {f} _{\mathbf {u} \mathbf {x} }.\end{alignedat}}}

The last terms in the last three equations denote contraction of a vector with a tensor. Minimizing the quadratic approximation (3) with respect to δ u {\displaystyle \delta \mathbf {u} } we have

δ u = argmin δ u Q ( δ x , δ u ) = Q u u 1 ( Q u + Q u x δ x ) , {\displaystyle {\delta \mathbf {u} }^{*}=\operatorname {argmin} \limits _{\delta \mathbf {u} }Q(\delta \mathbf {x} ,\delta \mathbf {u} )=-Q_{\mathbf {u} \mathbf {u} }^{-1}(Q_{\mathbf {u} }+Q_{\mathbf {u} \mathbf {x} }\delta \mathbf {x} ),} (4)

giving an open-loop term k = Q u u 1 Q u {\displaystyle \mathbf {k} =-Q_{\mathbf {u} \mathbf {u} }^{-1}Q_{\mathbf {u} }} and a feedback gain term K = Q u u 1 Q u x {\displaystyle \mathbf {K} =-Q_{\mathbf {u} \mathbf {u} }^{-1}Q_{\mathbf {u} \mathbf {x} }} . Plugging the result back into (3), we now have a quadratic model of the value at time i {\displaystyle i} :

Δ V ( i ) = 1 2 Q u T Q u u 1 Q u V x ( i ) = Q x Q x u Q u u 1 Q u V x x ( i ) = Q x x Q x u Q u u 1 Q u x . {\displaystyle {\begin{alignedat}{2}\Delta V(i)&=&{}-{\tfrac {1}{2}}Q_{\mathbf {u} }^{T}Q_{\mathbf {u} \mathbf {u} }^{-1}Q_{\mathbf {u} }\\V_{\mathbf {x} }(i)&=Q_{\mathbf {x} }&{}-Q_{\mathbf {xu} }Q_{\mathbf {u} \mathbf {u} }^{-1}Q_{\mathbf {u} }\\V_{\mathbf {x} \mathbf {x} }(i)&=Q_{\mathbf {x} \mathbf {x} }&{}-Q_{\mathbf {x} \mathbf {u} }Q_{\mathbf {u} \mathbf {u} }^{-1}Q_{\mathbf {u} \mathbf {x} }.\end{alignedat}}}

Recursively computing the local quadratic models of V ( i ) {\displaystyle V(i)} and the control modifications { k ( i ) , K ( i ) } {\displaystyle \{\mathbf {k} (i),\mathbf {K} (i)\}} , from i = N 1 {\displaystyle i=N-1} down to i = 1 {\displaystyle i=1} , constitutes the backward pass. As above, the Value is initialized with V ( x , N ) f ( x N ) {\displaystyle V(\mathbf {x} ,N)\equiv \ell _{f}(\mathbf {x} _{N})} . Once the backward pass is completed, a forward pass computes a new trajectory:

x ^ ( 1 ) = x ( 1 ) u ^ ( i ) = u ( i ) + k ( i ) + K ( i ) ( x ^ ( i ) x ( i ) ) x ^ ( i + 1 ) = f ( x ^ ( i ) , u ^ ( i ) ) {\displaystyle {\begin{aligned}{\hat {\mathbf {x} }}(1)&=\mathbf {x} (1)\\{\hat {\mathbf {u} }}(i)&=\mathbf {u} (i)+\mathbf {k} (i)+\mathbf {K} (i)({\hat {\mathbf {x} }}(i)-\mathbf {x} (i))\\{\hat {\mathbf {x} }}(i+1)&=\mathbf {f} ({\hat {\mathbf {x} }}(i),{\hat {\mathbf {u} }}(i))\end{aligned}}}

The backward passes and forward passes are iterated until convergence.

Differential dynamic programming is a second-order algorithm like Newton's method. It therefore takes large steps toward the minimum and often requires regularization and/or line-search to achieve convergence.[6][7] Regularization in the DDP context means ensuring that the Q u u {\displaystyle Q_{\mathbf {u} \mathbf {u} }} matrix in Eq. 4 is positive definite. Line-search in DDP amounts to scaling the open-loop control modification k {\displaystyle \mathbf {k} } by some 0 < α < 1 {\displaystyle 0<\alpha <1} .

Monte Carlo version

Sampled differential dynamic programming (SaDDP) is a Monte Carlo variant of differential dynamic programming.[8][9][10] It is based on treating the quadratic cost of differential dynamic programming as the energy of a Boltzmann distribution. This way the quantities of DDP can be matched to the statistics of a multidimensional normal distribution. The statistics can be recomputed from sampled trajectories without differentiation.

Sampled differential dynamic programming has been extended to Path Integral Policy Improvement with Differential Dynamic Programming.[11] This creates a link between differential dynamic programming and path integral control,[12] which is a framework of stochastic optimal control.

Constrained problems

Interior Point Differential dynamic programming (IPDDP) is an interior-point method generalization of DDP that can address the optimal control problem with nonlinear state and input constraints.[13]

See also

  • Optimal control

References

  1. ^ Mayne, D. Q. (1966). "A second-order gradient method of optimizing non-linear discrete time systems". Int J Control. 3: 85–95. doi:10.1080/00207176608921369.
  2. ^ Mayne, David Q.; Jacobson, David H. (1970). Differential dynamic programming. New York: American Elsevier Pub. Co. ISBN 978-0-444-00070-5.
  3. ^ de O. Pantoja, J. F. A. (1988). "Differential dynamic programming and Newton's method". International Journal of Control. 47 (5): 1539–1553. doi:10.1080/00207178808906114. ISSN 0020-7179.
  4. ^ Liao, L. Z.; C. A Shoemaker (1992). "Advantages of differential dynamic programming over Newton's method for discrete-time optimal control problems". Cornell University. hdl:1813/5474.
  5. ^ Morimoto, J.; G. Zeglin; C.G. Atkeson (2003). "Minimax differential dynamic programming: Application to a biped walking robot". Intelligent Robots and Systems, 2003.(IROS 2003). Proceedings. 2003 IEEE/RSJ International Conference on. Vol. 2. pp. 1927–1932.
  6. ^ Liao, L. Z; C. A Shoemaker (1991). "Convergence in unconstrained discrete-time differential dynamic programming". IEEE Transactions on Automatic Control. 36 (6): 692. doi:10.1109/9.86943.
  7. ^ Tassa, Y. (2011). Theory and implementation of bio-mimetic motor controllers (PDF) (Thesis). Hebrew University. Archived from the original (PDF) on 2016-03-04. Retrieved 2012-02-27.
  8. ^ "Sampled differential dynamic programming". 2016 IEEE/RSJ International Conference on Intelligent Robots and Systems (IROS). doi:10.1109/IROS.2016.7759229. S2CID 1338737.
  9. ^ Rajamäki, Joose; Hämäläinen, Perttu (June 2018). Regularizing Sampled Differential Dynamic Programming - IEEE Conference Publication. 2018 Annual American Control Conference (ACC). pp. 2182–2189. doi:10.23919/ACC.2018.8430799. S2CID 243932441. Retrieved 2018-10-19.
  10. ^ Rajamäki, Joose (2018). Random Search Algorithms for Optimal Control. Aalto University. ISBN 978-952-60-8156-4. ISSN 1799-4942.
  11. ^ Lefebvre, Tom; Crevecoeur, Guillaume (July 2019). "Path Integral Policy Improvement with Differential Dynamic Programming". 2019 IEEE/ASME International Conference on Advanced Intelligent Mechatronics (AIM). pp. 739–745. doi:10.1109/AIM.2019.8868359. hdl:1854/LU-8623968. ISBN 978-1-7281-2493-3. S2CID 204816072.
  12. ^ Theodorou, Evangelos; Buchli, Jonas; Schaal, Stefan (May 2010). "Reinforcement learning of motor skills in high dimensions: A path integral approach". 2010 IEEE International Conference on Robotics and Automation. pp. 2397–2403. doi:10.1109/ROBOT.2010.5509336. ISBN 978-1-4244-5038-1. S2CID 15116370.
  13. ^ Pavlov, Andrei; Shames, Iman; Manzie, Chris (2020). "Interior Point Differential Dynamic Programming". arXiv:2004.12710 [math.OC].
  • A Python implementation of DDP
  • A MATLAB implementation of DDP