This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Controllability and Observability Imply Exponential Decay of Sensitivity in Dynamic Optimization

Sungho Shin and Victor M. Zavala Department of Chemical and Biological Engineering,
University of Wisconsin-Madison, Madison, WI 53706 USA
(e-mail: {sungho.shin,victor.zavala}@wisc.edu).
Abstract

We study a property of dynamic optimization (DO) problems (as those encountered in model predictive control and moving horizon estimation) that is known as exponential decay of sensitivity (EDS). This property indicates that the sensitivity of the solution at stage ii against a data perturbation at stage jj decays exponentially with |ij||i-j|. Building upon our previous results, we show that EDS holds under uniform boundedness of the Lagrangian Hessian, a uniform second order sufficiency condition (uSOSC), and a uniform linear independence constraint qualification (uLICQ). Furthermore, we prove that uSOSC and uLICQ can be obtained under uniform controllability and observability. Hence, we have that uniform controllability and observability imply EDS. These results provide insights into how perturbations propagate along the horizon and enable the development of approximation and solution schemes. We illustrate the developments with numerical examples.

keywords:
sensitivity analysis, nonlinear, model predictive control, moving horizon estimation
thanks: We acknowledge support from the Grainger Wisconsin Distinguished Graduate Fellowship.

1 Introduction

This work studies the discrete-time, dynamic optimization (DO) formulation:

minx0:Nu0:N1\displaystyle\min_{\begin{subarray}{c}x_{0:N}\\ u_{0:N-1}\end{subarray}}\; i=0N1i(xi,ui;di)+N(xN;dN)\displaystyle\sum_{i=0}^{N-1}\ell_{i}(x_{i},u_{i};d_{i})+\ell_{N}(x_{N};d_{N}) (1a)
s.t.\displaystyle\mathop{\text{s.t.}}\; Tx0=d1|λ1\displaystyle Tx_{0}=d_{-1}\quad|\quad\lambda_{-1} (1b)
xi+1=fi(xi,ui;di),i𝕀[0,N1]|λi.\displaystyle x_{i+1}=f_{i}(x_{i},u_{i};d_{i}),\;i\in\mathbb{I}_{[0,N-1]}\quad|\quad\lambda_{i}. (1c)

Here, N𝕀>0N\in\mathbb{I}_{>0} is the horizon length; for each stage (time) ii, xinxx_{i}\in\mathbb{R}^{n_{x}} are the states, uinuu_{i}\in\mathbb{R}^{n_{u}} are the controls, dindd_{i}\in\mathbb{R}^{n_{d}} are the data (parameters), λinx\lambda_{i}\in\mathbb{R}^{n_{x}} are the dual variables, i:nx×nu×nd\ell_{i}:\mathbb{R}^{n_{x}}\times\mathbb{R}^{n_{u}}\times\mathbb{R}^{n_{d}}\rightarrow\mathbb{R} are the stage cost functions, fi:nx×nu×ndnxf_{i}:\mathbb{R}^{n_{x}}\times\mathbb{R}^{n_{u}}\times\mathbb{R}^{n_{d}}\rightarrow\mathbb{R}^{n_{x}} are the dynamic mapping functions, N:nx×nd\ell_{N}:\mathbb{R}^{n_{x}}\times\mathbb{R}^{n_{d}}\rightarrow\mathbb{R} is the final cost function, and the symbol || is used to denote the associated dual variables. The initial state constraint is enforced with the initial state mapping Tn0×nxT\in\mathbb{R}^{n_{0}\times n_{x}} and parameter d1n0d_{-1}\in\mathbb{R}^{n_{0}}. We let x1,u1,uN,λNx_{-1},u_{-1},u_{N},\lambda_{N} be empty vectors (for convenience), and define zi:=[xi,ui]z_{i}:=[x_{i},u_{i}], wi:=[zi;λi]w_{i}:=[z_{i};\lambda_{i}], and ξi:=[wi;di]\xi_{i}:=[w_{i};d_{i}] for i𝕀[1,N]i\in\mathbb{I}_{[-1,N]}, and we use the syntax va:b:=[va;va+1;;vb]v_{a:b}:=[v_{a};v_{a+1};\cdots;v_{b}] for v=x,u,λ,d,z,w,ξv=x,u,\lambda,d,z,w,\xi. The DO problem (1) is a parametric nonlinear program that we denote as P0:N(d1:N)P_{0:N}(d_{-1:N}). We assume that all functions are twice continuously differentiable and potentially nonconvex. Typical MPC problems are formulated with T=IT=I and typical MHE problems are formulated with an empty matrix T0×nxT\in\mathbb{R}^{0\times n_{x}} (i.e., the initial constraint is not enforced). State-output mappings encountered in MHE problems are assumed to be embedded within the stage cost functions.

In this paper we study a property of DO problems that is known as exponential decay of sensitivity (EDS) (Na and Anitescu, 2020a; Shin et al., 2021). The property indicates that the sensitivity of the solution at stage ii against a data perturbation at stage jj decays exponentially with |ij||i-j|. This property helps understand how different data perturbations (e.g., disturbances or changes in set-points, initial conditions, and terminal penalties) propagate along with the time horizon. Moreover, EDS has been shown to be essential in constructing efficient discretization schemes for continuous-time DO formulations (Shin and Zavala, 2020; Grüne et al., 2020b) and in establishing convergence of algorithms (Na et al., 2020; Na and Anitescu, 2020b).

Building upon our previous results (Shin et al., 2021) we show that, under uniform boundedness of the Lagrangian Hessian (uBLH), a uniform second order sufficiency condition (uSOSC), and a uniform linear independence constraint qualification (uLICQ), the primal-dual solution of the DO problem at a given stage decays exponentially with the distance to the stage at which a data perturbation is introduced. In particular, given base data d1:Nd_{-1:N}^{\star} and associated primal-dual solution trajectory w1:Nw^{\star}_{-1:N} (at which uBLH, uSOSC, and uLICQ are satisfied), there exist uniform constants Υ>0\Upsilon>0 and ρ(0,1)\rho\in(0,1) and a neighborhood 𝔻1:N\mathbb{D}^{\star}_{-1:N} of d1:Nd_{-1:N}^{\star} such that the following holds for any d1:N,d1:N𝔻1:Nd_{-1:N},d^{\prime}_{-1:N}\in\mathbb{D}^{\star}_{-1:N}:

wi(d1:N)wi(d1:N)j=1NΥρ|ij|djdj,\displaystyle\|w^{\dagger}_{i}(d_{-1:N})-w^{\dagger}_{i}(d^{\prime}_{-1:N})\|\leq{\sum_{j=-1}^{N}}\Upsilon\rho^{|i-j|}\|d_{j}-d^{\prime}_{j}\|, (2)

where wi()w^{\dagger}_{i}(\cdot) is the primal-dual solution mapping at stage ii. That is, the sensitivity Υρ|ij|\Upsilon\rho^{|i-j|} of the solution at stage ii against a data perturbation at stage jj decays exponentially with respect to the distance |ij||i-j|. Here, it is important that (Υ,ρ)(\Upsilon,\rho) are uniform constants (independent of horizon length NN); this allows us to maintain (Υ,ρ)(\Upsilon,\rho) unchanged even if the horizon length becomes indefinitely long (e.g., when approaching an infinite horizon). Our key finding is that uBLH, uLICQ, and uSOSC can be obtained under uniform controllability/observability and under uniformly bounded system matrices (standard assumption). This result thus allows to establish EDS directly from fundamental system-theoretic properties.

In summary, our main contribution is showing that controllability and observability provide sufficient conditions for EDS. This result sheds light on how system-theoretic properties influence the propagation of perturbations along the solution trajectory. EDS for continuous-time, linear-quadratic MPC has been established under stablizability and detectability in Grüne et al. (2019, 2020b, 2020a). EDS has been established for discrete-time, nonlinear MPC under uniform SOSC and controllability in (Na and Anitescu, 2020a); the proof of this result uses a Riccati recursion representation of the optimality conditions of the DO problem. The proof that we present here is more compact and general; specifically, our approach is based on a graph-theoretic analysis of the optimality conditions. This general setting allows us to establish EDS for discrete-time, nonlinear MPC and MHE problems.

Basic Notation: The set of real numbers and the set of integers are denoted by \mathbb{R} and 𝕀\mathbb{I}, respectively, and we define 𝕀A:=𝕀A\mathbb{I}_{A}:=\mathbb{I}\cap A, where AA is a set; 𝕀>0:=𝕀(0,)\mathbb{I}_{>0}:=\mathbb{I}\cap(0,\infty); 𝕀0:=𝕀[0,)\mathbb{I}_{\geq 0}:=\mathbb{I}\cap[0,\infty). We consider vectors always as column vectors. We use the syntax: {Mi}i=1N:=[M1;;MN]:=[M1Mn]\{M_{i}\}_{i=1}^{N}:=[M_{1};\cdots;M_{N}]:=[M_{1}^{\top}\,\cdots\,M_{n}^{\top}]^{\top}. Furthermore, v[i]v[i] denotes the ii-th component of vv. For a function ϕ:n\phi:\mathbb{R}^{n}\rightarrow\mathbb{R} and variable vectors ypy\in\mathbb{R}^{p}, zqz\in\mathbb{R}^{q}, yz2ϕ(x):={({2y[i]z[j]ϕ(x)}j=1q)}i=1p\nabla^{2}_{yz}\phi(x):=\{(\{\frac{\partial^{2}}{\partial y[i]\partial z[j]}\phi(x)\}_{j=1}^{q})^{\top}\}_{i=1}^{p}. For a vector function φ:nm\varphi:\mathbb{R}^{n}\rightarrow\mathbb{R}^{m} and a variable vector wsw\in\mathbb{R}^{s}, wφ(x):={({w[j]φ(x)[i]}j=1s)}i=1m\nabla_{w}\varphi(x):=\{(\{\frac{\partial}{\partial w[j]}\varphi(x)[i]\}_{j=1}^{s})^{\top}\}_{i=1}^{m}. Vector 2-norms and induced 2-norms of matrices are denoted by \|\cdot\|. For matrices AA and BB with the same dimensions, A()BA\succ(\succeq)B indicates that ABA-B is positive (semi) definite. We use the convention: if mm or nn is zero, m×n\mathbb{R}^{m\times n} is a singleton only containing the m×nm\times n null matrix.

2 Main Results

2.1 Exponential Decay of Sensitivity

In this section, we establish EDS (2) for P0:N()P_{0:N}(\cdot). We first formally define sufficient conditions for EDS to hold (uBLH, uSOSC, and uLICQ). We begin by defining the notion of uniformly bounded quantities.

Definition 1 (Uniform Bounds).

A set {Ai}i𝒜\{A_{i}\}_{i\in\mathcal{A}} is called LL-uniformly bounded above (below) if there exists a uniform constant LL\in\mathbb{R} (independent of NN) such that Ai()L\|A_{i}\|\leq(\geq)L holds for any i𝒜i\in\mathcal{A}.

We say that a set of a quantity is uniformly bounded above (below) if there exists a uniform constant LL such that the set is LL-uniformly bounded above (below). Also, we will write that some quantity aa is uniformly bounded above (below) if {a}\{a\} is uniformly bounded above (below).

The Lagrangian function of P0:N(d1:N)P_{0:N}(d_{-1:N}) is defined as

0:N(w1:N;d1:N):=i=0Ni(zi,λi1:i;di),\displaystyle\mathcal{L}_{0:N}(w_{-1:N};d_{-1:N}):=\sum_{i=0}^{N}\mathcal{L}_{i}(z_{i},\lambda_{i-1:i};d_{i}),

where:

i(zi,λi1:i;di)\displaystyle\mathcal{L}_{i}(z_{i},\lambda_{i-1:i};d_{i}) :=i(zi;di)λi1xi+λifi(zi;di)\displaystyle:=\ell_{i}(z_{i};d_{i})-\lambda_{i-1}^{\top}x_{i}+\lambda_{i}^{\top}f_{i}(z_{i};d_{i})
N(xN,λN1;dN)\displaystyle\mathcal{L}_{N}(x_{N},\lambda_{N-1};d_{N}) :=N(xN;dN)λN1xN.\displaystyle:=\ell_{N}(x_{N};d_{N})-\lambda^{\top}_{N-1}x_{N}.
Definition 2 (uBLH).

Given d1:Nd_{-1:N}^{\star} and the solution w1:Nw_{-1:N}^{\star} of P0:N(d1:N)P_{0:N}(d_{-1:N}^{\star}), LL-uBLH holds if:

w1:Nξ1:N20:N(w1:N;d1:N)L,\displaystyle\|\nabla^{2}_{w_{-1:N}\xi_{-1:N}}\mathcal{L}_{0:N}(w_{-1:N}^{\star};d_{-1:N}^{\star})\|\leq L, (3)

with uniform constant L<L<\infty.

The primal Hessian 𝑯0:N\boldsymbol{H}_{0:N} of the Lagrangian and the constraint Jacobian 𝑱0:N\boldsymbol{J}_{0:N} are:

𝑯0:N\displaystyle\boldsymbol{H}_{0:N} :=z0:N,z0:N20:N(w1:N;d1:N)\displaystyle:=\nabla^{2}_{z_{0:N},z_{0:N}}\mathcal{L}_{0:N}(w_{-1:N}^{\star};d_{-1:N}^{\star}) (4a)
𝑱0:N\displaystyle\boldsymbol{J}_{0:N} :=z0:Nc1:N1(z0:N;d1:N),\displaystyle:=\nabla_{z_{0:N}}c_{-1:N-1}(z^{\star}_{0:N};d_{-1:N}^{\star}), (4b)

where c1:N1()c_{-1:N-1}(\cdot) is the constraint function for P0:N()P_{0:N}(\cdot); that is,

c1:N1(z0:N;d1:N):=[Tx0d1x1f1(z0;d0)xNfN1(zN1;dN1)].c_{-1:N-1}(z_{0:N};d_{-1:N}):=\begin{bmatrix}Tx_{0}-d_{-1}\\ x_{1}-f_{1}(z_{0};d_{0})\\ \cdots\\ x_{N}-f_{N-1}(z_{N-1};d_{N-1})\end{bmatrix}.
Definition 3 (uSOSC).

Given d1:Nd_{-1:N}^{\star} and the solution w1:Nw_{-1:N}^{\star} of P0:N(d1:N)P_{0:N}(d_{-1:N}^{\star}), γ\gamma-uSOSC holds if:

ReH(𝑯0:N,𝑱0:N)γI,\displaystyle ReH(\boldsymbol{H}_{0:N},\boldsymbol{J}_{0:N})\succeq\gamma I, (5)

with uniform constant γ>0\gamma>0.

Here, ReH(𝑯0:N,𝑱0:N):=Z𝑯0:NZReH(\boldsymbol{H}_{0:N},\boldsymbol{J}_{0:N}):=Z^{\top}\boldsymbol{H}_{0:N}Z is the reduced Hessian and ZZ is a null-space matrix of 𝑱0:N\boldsymbol{J}_{0:N}.

Definition 4 (uLICQ).

Given d1:Nd_{-1:N}^{\star} and the primal-dual solution w1:Nw_{-1:N}^{\star} of P0:N(d1:N)P_{0:N}(d_{-1:N}^{\star}), β\beta-uLICQ holds if:

𝑱0:N𝑱0:NβI\displaystyle\boldsymbol{J}_{0:N}\boldsymbol{J}_{0:N}^{\top}\succeq\beta I (6)

with uniform constant β>0\beta>0.

Note that uSOSC assumes that the smallest eigenvalue of the reduced Hessian is uniformly bounded below by γ\gamma, while uLICQ assumes that the smallest non-trivial singular value of the Jacobian is uniformly bounded below by β1/2{\beta}^{1/2}. Thus, these are strengthened versions of SOSC and LICQ. We require uSOSC and uLICQ because, under SOSC and LICQ, the smallest eigenvalue of reduced Hessian or the smallest non-trivial singular value of the Jacobian may become arbitrarily close to 0 as the horizon length NN is extended (e.g., see Shin et al. (2021, Example 4.18)). Under uSOSC and uLICQ, on the other hand, the lower bounds are independent of NN.

Assumption 5.

Given twice continuously differentiable functions {i()}i=0N\{\ell_{i}(\cdot)\}_{i=0}^{N}, {fi()}i=0N1\{f_{i}(\cdot)\}_{i=0}^{N-1} and base data d1:Nd^{\star}_{-1:N}, there exists a primal-dual solution w1:Nw_{-1:N}^{\star} of P0:N(d1:N)P_{0:N}(d_{-1:N}^{\star}) at which LL-uBLH, γ\gamma-uSOSC, and β\beta-uLICQ are satisfied.

The following lemma is a well-known characterization of solution mappings of parametric nonlinear programs (NLPs) (Robinson, 1980; Dontchev and Rockafellar, 2009).

Lemma 6.

Under Assumption 5, there exist neighborhoods 𝔻1:N\mathbb{D}^{\star}_{-1:N} of d1:Nd^{\star}_{-1:N} and 𝕎1:N\mathbb{W}^{\star}_{-1:N} of w1:Nw^{\star}_{-1:N} and continuous w1:N:𝔻1:N𝕎1:Nw^{\dagger}_{-1:N}:\mathbb{D}^{\star}_{-1:N}\rightarrow\mathbb{W}^{\star}_{-1:N} such that for any d1:N𝔻1:Nd_{-1:N}\in\mathbb{D}^{\star}_{-1:N}, w1:N(d1:N)w^{\dagger}_{-1:N}(d_{-1:N}) is a local solution of P0:N(d1:N)P_{0:N}(d_{-1:N}).

{pf}

From Shin et al. (2021, Lemma 3.3). ∎ We can thus see that there exists a well-defined solution mapping w1:N()w^{\dagger}_{-1:N}(\cdot) around the neighborhood of d1:Nd_{-1:N}^{\star}. We now study stage-wise solution sensitivity by characterizing the dependence of wi()w^{\dagger}_{i}(\cdot) on the data d1:Nd_{-1:N}.

Theorem 7.

Under Assumption 5, there exist uniform constants Υ>0\Upsilon>0 and ρ(0,1)\rho\in(0,1) (functions of L,γ,βL,\gamma,\beta) and neighborhoods 𝔻1:N\mathbb{D}^{\star}_{-1:N} of d1:Nd_{-1:N}^{\star} and 𝕎1:N\mathbb{W}^{\star}_{-1:N} of w1:Nw_{-1:N}^{\star} such that (2) holds for any d1:N,d1:N𝔻1:Nd_{-1:N},d^{\prime}_{-1:N}\in\mathbb{D}^{\star}_{-1:N} and i𝕀[1,N]i\in\mathbb{I}_{[-1,N]}.

{pf}

We observe that P0:N()P_{0:N}(\cdot) is graph-structured (induced by 𝒢N=(𝒱N,N)\mathcal{G}_{N}=(\mathcal{V}_{N},\mathcal{E}_{N}), where 𝒱N={1,0,,N}\mathcal{V}_{N}=\{-1,0,\cdots,N\} and N={{1,0},{0,1},,{N1,N}\mathcal{E}_{N}=\{\{-1,0\},\{0,1\},\cdots,\{N-1,N\}), and the maximum graph degree D=2D=2. From uBLH, uLICQ, and uSOSC, one can see that assumptions in Shin et al. (2021, Theorem 4.9) are satisfied. This implies that the singular values of w0:Nw0:N20:N(w1:N;d1:N)\nabla^{2}_{w_{0:N}w_{0:N}}\mathcal{L}_{0:N}(w^{\star}_{-1:N};d^{\star}_{-1:N}) are uniformly upper and lower bounded and those of w0:Nd1:N20:N(w1:N;d1:N)\nabla^{2}_{w_{0:N}d_{-1:N}}\mathcal{L}_{0:N}(w^{\star}_{-1:N};d^{\star}_{-1:N}) are uniformly upper bounded (uniform constants given by functions of L,β,γL,\beta,\gamma; see Shin et al. (2021, Equation (4.15))). We then apply Shin et al. (2021, Theorem 3.5) to obtain Υ>0\Upsilon>0 and ρ(0,1)\rho\in(0,1) as functions of the upper and lower bounds of the singular values (see Shin et al. (2021, Equation (3.17))). This allows expressing Υ,ρ\Upsilon,\rho as functions of L,β,γL,\beta,\gamma. ∎

Theorem 7 establishes EDS under the regularity conditions of Assumption 5. It is important that Υ,ρ\Upsilon,\rho can be determined solely in terms of L,γ,βL,\gamma,\beta (and do not depend on the horizon length NN). Practical DO problems typically have additional equality/inequality constraints that are not considered in (1). Thus, Theorem 7 may not be directly applicable to those problems. However, the results in Shin et al. (2021) are applicable to such problems as long as the DO problem is a graph-structured NLP. Specifically, under uniformly strong SOSC and uLICQ, we can establish EDS using Shin et al. (2021, Theorem 3.5, 4.9). The graph structure breaks when there exist globally coupled variables; typical MPC and MHE problems do not have such variables, but parameter estimation problems may have such variables. Specifically, in the presence of globally coupled variables, the graph distance between any pair of stages is not greater than two.

2.2 Regularity from System-Theoretic Properties

Although uSOSC and uLICQ are standard notions of NLP solution regularity, they are not intuitive notions from a system-theoretic perspective. However, we now show that uSOSC and uLICQ can be obtained from uniform controllability and observability. We begin by defining:

Qi\displaystyle Q_{i} :=xixi2i(zi,λi1:i;di)\displaystyle:=\nabla^{2}_{x_{i}x_{i}}\mathcal{L}_{i}(z^{\star}_{i},\lambda^{\star}_{i-1:i};d^{\star}_{i}) Ri:=uiui2i(zi,λi1:i;di)\displaystyle R_{i}:=\nabla^{2}_{u_{i}u_{i}}\mathcal{L}_{i}(z^{\star}_{i},\lambda^{\star}_{i-1:i};d^{\star}_{i})
Si\displaystyle S_{i} :=xiui2i(zi,λi1:i;di)\displaystyle:=\nabla^{2}_{x_{i}u_{i}}\mathcal{L}_{i}(z^{\star}_{i},\lambda^{\star}_{i-1:i};d^{\star}_{i}) Ei:=xidi2i(zi,λi1:i;di)\displaystyle E_{i}:=\nabla^{2}_{x_{i}d_{i}}\mathcal{L}_{i}(z^{\star}_{i},\lambda^{\star}_{i-1:i};d^{\star}_{i})
Fi\displaystyle F_{i} :=uidi2i(zi,λi1:i;di)\displaystyle:=\nabla^{2}_{u_{i}d_{i}}\mathcal{L}_{i}(z^{\star}_{i},\lambda^{\star}_{i-1:i};d^{\star}_{i}) Ai:=xifi(zi;di)\displaystyle A_{i}:=\nabla_{x_{i}}f_{i}(z^{\star}_{i};d^{\star}_{i})
Bi\displaystyle B_{i} :=uifi(zi;di)\displaystyle:=\nabla_{u_{i}}f_{i}(z^{\star}_{i};d^{\star}_{i}) Gi:=difi(zi;di).\displaystyle G_{i}:=\nabla_{d_{i}}f_{i}(z^{\star}_{i};d^{\star}_{i}).
Definition 8.

({Ai}i=1N1,{Bi}i=0N1)(\{A_{i}\}_{i=1}^{N-1},\{B_{i}\}_{i=0}^{N-1}) is (Nc,βc)(N_{c},\beta_{c})-uniformly controllable with Nc𝕀0N_{c}\in\mathbb{I}_{\geq 0} and βc>0\beta_{c}>0 (independent of NN) if, for any i,j𝕀[0,N1]i,j\in\mathbb{I}_{[0,N-1]} with |ij|Nc|i-j|\geq N_{c}, 𝒞i:j𝒞i:jβcI\mathcal{C}_{i:j}\mathcal{C}_{i:j}^{\top}\succeq\beta_{c}I holds, where

𝒞i:j:=[Ai+1:jBiAjBj1Bj].\displaystyle\mathcal{C}_{i:j}:=\begin{bmatrix}A_{i+1:j}B_{i}&\cdots&A_{j}B_{j-1}&B_{j}\end{bmatrix}.
Definition 9.

({Ai}i=0N1,{Qi}i=0N)(\{A_{i}\}_{i=0}^{N-1},\{Q_{i}\}_{i=0}^{N}) is (No,γo)(N_{o},\gamma_{o})-uniformly observable with No𝕀0N_{o}\in\mathbb{I}_{\geq 0} and γo>0\gamma_{o}>0 (independent of NN) if for any i,j𝕀[0,N1]i,j\in\mathbb{I}_{[0,N-1]} with |ij|No|i-j|\geq N_{o}, 𝒪i:j𝒪i:jγoI\mathcal{O}^{\top}_{i:j}\mathcal{O}_{i:j}\succeq\gamma_{o}I holds, where

𝒪i:j:=[QjAi:j1Qi+1AiQi].\displaystyle\mathcal{O}_{i:j}:=\begin{bmatrix}Q_{j}A_{i:j-1}\\ \ddots\\ Q_{i+1}A_{i}\\ Q_{i}\end{bmatrix}.

Here

Aa:b:={AbAb1Aa+1Aa, if abAbAb+1Aa1Aa, otherwise.A_{a:b}:=\begin{cases}A_{b}A_{b-1}\cdots A_{a+1}A_{a},\text{ if }a\leq b\\ A_{b}A_{b+1}\cdots A_{a-1}A_{a},\text{ otherwise.}\end{cases}

Note that uniform controllability and observability are stronger versions of their standard counterparts. One can establish the following duality between uniform controllability and observability.

Proposition 10.

({Ai}i=1N,{Bi}i=0N)(\{A_{i}\}_{i=1}^{N},\{B_{i}\}_{i=0}^{N}) is (N0,α0)(N_{0},\alpha_{0})-uniformly controllable if and only if ({Ai}i=N1,{Bi}i=N0)(\{A^{\top}_{i}\}_{i=N}^{1},\{B^{\top}_{i}\}_{i=N}^{0}) is (N0,α0)(N_{0},\alpha_{0})-uniformly observable (here, note that the orders of sequences {Ai}i=N1,{Bi}i=N0\{A^{\top}_{i}\}_{i=N}^{1},\{B^{\top}_{i}\}_{i=N}^{0} are inverted).

{pf}

The proof is straightforward and thus omitted. ∎

The following technical lemma is needed to show that uniform controllability implies uLICQ.

Lemma 11.

Consider a block row/column operator UU with LL-uniformly bounded above block VV of the form:

U:=[IVII],[IVII].\displaystyle U:=\begin{bmatrix}I\\ V&I\\ &&\ddots\\ &&&I\end{bmatrix},\begin{bmatrix}I&&&V\\ &I\\ &&\ddots\\ &&&I\end{bmatrix}.

We have that U,U1U,U^{-1} are (L+1)(L+1)-uniformly bounded above.

{pf}

The proof is straightforward and thus omitted. ∎ We now show that uniform controllability implies uLICQ.

Lemma 12.

KK-uniform upper boundedness of {Ai}i=0N1\{A_{i}\}_{i=0}^{N-1} and {Bi}i=0N1\{B_{i}\}_{i=0}^{N-1}, TTδITT^{\top}\succeq\delta I for uniformly lower bounded δ>0\delta>0, and (Nc,βc)(N_{c},\beta_{c})-uniform controllability of ({Ai}i=1N1,{Bi}i=0N1)(\{A_{i}\}_{i=1}^{N-1},\{B_{i}\}_{i=0}^{N-1}) implies (6), where β>0\beta>0 is a function of K,δ,Nc,βcK,\delta,N_{c},\beta_{c}.

{pf}

The Jacobian 𝑱0:N\boldsymbol{J}_{0:N} has the following form:

𝑱0:N=[TA0B0IAN2BN2IAN1BN1I].\displaystyle\small\boldsymbol{J}_{0:N}=\begin{bmatrix}T\\ -A_{0}&-B_{0}&I\\ &&\ddots\\ &&-A_{N-2}&-B_{N-2}&I\\ &&&&-A_{N-1}&-B_{N-1}&I\end{bmatrix}.

By inspecting the block structure of 𝑱0:N\boldsymbol{J}_{0:N} and Shin et al. (2021, Lemma 4.15), one can see that it suffices to show that the smallest non-trivial singular value of

[SAiBiIAj1Bj1IAjBj]\displaystyle\begin{bmatrix}S\\ -A_{i}&-B_{i}&I\\ &&\ddots\\ &&-A_{j-1}&-B_{j-1}&I\\ &&&&-A_{j}&-B_{j}\end{bmatrix} (7)

is β1/2{\beta}^{1/2}-uniformly bounded below for S=TS=T or II and for any i,j𝕀[0,N1]i,j\in\mathbb{I}_{[0,N-1]} with Nc|ij|2NcN_{c}\leq|i-j|\leq 2N_{c}, where 0<β10<\beta\leq 1 is a function of K,δ,Nc,βcK,\delta,N_{c},\beta_{c}. This follows from the observation that one can always partition 𝕀[0,N1]\mathbb{I}_{[0,N-1]} into a family of blocks with size between NcN_{c} and 2Nc2N_{c}. For now, we assume S=IS=I. By applying a set of suitable block row and column operations (in particular, first apply block row operations to eliminate Ai,,AjA_{i},\cdots,A_{j}, and then apply block column operations to eliminate Bi,,Ai:j1Bj2-B_{i},\cdots,-A_{i:j-1}B_{j-2}) and permutations, one can obtain the following:

[IAi+1:jBiAjBj1Bj].\displaystyle\begin{bmatrix}I\\ &-A_{i+1:j}B_{i}&\cdots&-A_{j}B_{j-1}&-B_{j}\end{bmatrix}. (8)

The lower-right blocks constitute the controllability matrix 𝒞i:j\mathcal{C}_{i:j}; from uniform controllability, the smallest non-trivial singular value of the matrix in (8) is uniformly lower bounded by min(1,βc1/2)\min(1,\beta_{c}^{1/2}). Here, we have applied block-row and block-column operations as the ones that appear in Lemma 11 (each multiplied block is uniformly bounded above due to KK-uniform boundedness of {Ai}i=0N1\{A_{i}\}_{i=0}^{N-1} and {Bi}i=0N1\{B_{i}\}_{i=0}^{N-1}). Also, we have applied such operations only uniformly bounded many times (the number of operations is independent of NN since the number of blocks in the matrix in (7) is bounded by 4(2Nc+1)(Nc+1)4(2N_{c}+1)(N_{c}+1), which is uniformly bounded above). We thus have that the smallest non-trivial singular value of the matrix in (7) is uniformly lower bounded with uniform constant β01/2{\beta_{0}}^{1/2}, and β0>0\beta_{0}>0 is given by a function of K,Nc,βcK,N_{c},\beta_{c}. Now we consider the S=TS=T case. One can observe that the smallest non-trivial singular value of the matrix in (7) with S=TS=T is lower bounded by that with S=[T~;T]S=[\widetilde{T};T] (here, T~\widetilde{T}^{\top} is a null space matrix of TT); and again, it is lower bounded by δ1/2\delta^{1/2} times that with S=IS=I. We thus have that the smallest non-trivial singular value of the matrix in (7) with S=TS=T is uniformly lower bounded by β01/2δ1/2\beta_{0}^{1/2}\delta^{1/2}. Therefore, the smallest non-trivial singular values of the matrices in (7) with S=IS=I or TT are β1/2\beta^{1/2}-uniformly lower bounded for any i,j𝕀[0,N1]i,j\in\mathbb{I}_{[0,N-1]} with Nc|ij|2NcN_{c}\leq|i-j|\leq 2N_{c}, where β:=min(β0,δβ0,1)\beta:=\min(\beta_{0},\delta\beta_{0},1). Thus, by Shin et al. (2021, Lemma 4.15) we have (6). ∎

If T0×nxT\in\mathbb{R}^{0\times n_{x}}, the assumption TTδITT^{\top}\succeq\delta I for uniformly lower bounded δ>0\delta>0 holds for an arbitrary δ>0\delta>0 due to the convention introduced in the Notation in Section 1.

We now show that uniform observability implies uSOSC.

Lemma 13.

KK-uniform boundedness of {Ai}i=0N1\{A_{i}\}_{i=0}^{N-1}, {Bi}i=0N1\{B_{i}\}_{i=0}^{N-1}, and {Qi}i=0N\{Q_{i}\}_{i=0}^{N}, Qi0Q_{i}\succeq 0, Si=0S_{i}=0, RirIR_{i}\succeq rI (r>0r>0 is independent of NN), and (No,γo)(N_{o},\gamma_{o})-uniform observability of ({Ai}i=0N1,{Qi}i=0N)(\{A_{i}\}_{i=0}^{N-1},\{Q_{i}\}_{i=0}^{N}) implies (5), where γ>0\gamma>0 is a function of K,No,γo,rK,N_{o},\gamma_{o},r.

{pf}

The primal Hessian 𝑯0:N\boldsymbol{H}_{0:N} has the following form:

𝑯0:N[Q0R0QN1RN1QN].\displaystyle\boldsymbol{H}_{0:N}\begin{bmatrix}Q_{0}\\ &R_{0}\\ &&\ddots\\ &&&Q_{N-1}\\ &&&&R_{N-1}\\ &&&&&Q_{N}\end{bmatrix}.

By inspecting the block structure of 𝑯0:N\boldsymbol{H}_{0:N} and 𝑱0:N\boldsymbol{J}_{0:N} and Shin et al. (2021, Lemma 4.14), one can observe that it suffices to show that: first,

ReH([𝑸i:j𝑹i:j1],[𝑨i:j𝑩i:j1])\displaystyle ReH\left(\begin{bmatrix}\boldsymbol{Q}_{i:j}\\ &\boldsymbol{R}_{i:j-1}\end{bmatrix},\begin{bmatrix}\boldsymbol{A}_{i:j}&\boldsymbol{B}_{i:j-1}\end{bmatrix}\right) (9)

has γ\gamma-uniformly lower bounded smallest eigenvalue with γ>0\gamma>0 for any i,j𝕀[0,N1]i,j\in\mathbb{I}_{[0,N-1]} with No|ij|2NoN_{o}\leq|i-j|\leq 2N_{o}, where:

𝑨i:j:=[AiIAj1I],𝑩i:j1:=[BiBj1]\displaystyle\boldsymbol{A}_{i:j}:=\begin{bmatrix}-A_{i}&I\\ &\ddots&\ddots\\ &&-A_{j-1}&I\end{bmatrix},\boldsymbol{B}_{i:j-1}:=\begin{bmatrix}-B_{i}\\ &&\ddots\\ &&&-B_{j-1}\end{bmatrix}
𝑸i:j:=[QiQi+1Qj],𝑹i:j1:=[RiRi+1Rj1],\displaystyle\boldsymbol{Q}_{i:j}:=\begin{bmatrix}Q_{i}\\ &Q_{i+1}\\ &&\ddots\\ &&&Q_{j}\end{bmatrix},\boldsymbol{R}_{i:j-1}:=\begin{bmatrix}R_{i}\\ &R_{i+1}\\ &&\ddots\\ &&&R_{j-1}\end{bmatrix},

and second, RiγIR_{i}\succeq\gamma I for any i𝕀[0,N1]i\in\mathbb{I}_{[0,N-1]}. This follows from the observation that one can always partition 𝕀[0,N1]\mathbb{I}_{[0,N-1]} into a family of blocks with size between NcN_{c} and 2Nc2N_{c}. We consider 𝒙i:j,𝒖i:j1\boldsymbol{x}_{i:j},\boldsymbol{u}_{i:j-1} such that 𝑨i:j𝒙i:j+𝑩i:j1𝒖i:j1=0\boldsymbol{A}_{i:j}\boldsymbol{x}_{i:j}+\boldsymbol{B}_{i:j-1}\boldsymbol{u}_{i:j-1}=0 holds. By uniform positive definiteness of {Ri}i=0N1\{R_{i}\}_{i=0}^{N-1} and uniform boundedness of {Bi}i=0N1\{B_{i}\}_{i=0}^{N-1}, for κ:=r/2K2\kappa:=r/2K^{2}, we have

12𝒖i:j1𝑹i:j1𝒖i:j1\displaystyle\frac{1}{2}\boldsymbol{u}_{i:j-1}^{\top}\boldsymbol{R}_{i:j-1}\boldsymbol{u}_{i:j-1} κ𝒖i:j1𝑩i:j1𝑩i:j1𝒖i:j1\displaystyle\geq\kappa\boldsymbol{u}_{i:j-1}^{\top}\boldsymbol{B}_{i:j-1}^{\top}\boldsymbol{B}_{i:j-1}\boldsymbol{u}_{i:j-1}
=κ𝒙i:j𝑨i:j𝑨i:j𝒙i:j,\displaystyle=\kappa\boldsymbol{x}_{i:j}^{\top}\boldsymbol{A}_{i:j}^{\top}\boldsymbol{A}_{i:j}\boldsymbol{x}_{i:j},

where the equality follows from 𝑨i:j𝒙i:j+𝑩i:j1𝒖i:j1=0\boldsymbol{A}_{i:j}\boldsymbol{x}_{i:j}+\boldsymbol{B}_{i:j-1}\boldsymbol{u}_{i:j-1}=0. Furthermore, from 𝑸i:j0\boldsymbol{Q}_{i:j}\succeq 0, we have that:

𝒙i:j𝑸i:j2𝒙i:j=(𝑸i:j1/2𝒙i:j)𝑸i:j(𝑸i:j1/2𝒙i:j)K𝒙i:j𝑸i:j𝒙i:j,\displaystyle\boldsymbol{x}_{i:j}^{\top}\boldsymbol{Q}_{i:j}^{2}\boldsymbol{x}_{i:j}=(\boldsymbol{Q}_{i:j}^{1/2}\boldsymbol{x}_{i:j})^{\top}\boldsymbol{Q}_{i:j}(\boldsymbol{Q}_{i:j}^{1/2}\boldsymbol{x}_{i:j})\leq K\boldsymbol{x}_{i:j}^{\top}\boldsymbol{Q}_{i:j}\boldsymbol{x}_{i:j},

where the inequality follows from that the largest eigenvalue of 𝑸i:j\boldsymbol{Q}_{i:j} is bounded by KK. Thus, 𝒙i:j𝑸i:j𝒙i:j+𝒖i:j1𝑹i:j1𝒖i:j1\boldsymbol{x}_{i:j}^{\top}\boldsymbol{Q}_{i:j}\boldsymbol{x}_{i:j}+\boldsymbol{u}_{i:j-1}^{\top}\boldsymbol{R}_{i:j-1}\boldsymbol{u}_{i:j-1} is not less than:

min(1/K,κ)𝒙i:j[𝑸i:j𝑨i:j][𝑸i:j𝑨i:j]𝒙i:j+r2𝒖i:j12.\displaystyle\min(1/K,\kappa)\boldsymbol{x}_{i:j}^{\top}\begin{bmatrix}\boldsymbol{Q}_{i:j}&\boldsymbol{A}_{i:j}^{\top}\end{bmatrix}\begin{bmatrix}\boldsymbol{Q}_{i:j}\\ \boldsymbol{A}_{i:j}\end{bmatrix}\boldsymbol{x}_{i:j}+\frac{r}{2}\|\boldsymbol{u}_{i:j-1}\|^{2}.

Observe that [𝑸i:j𝑨i:j]\begin{bmatrix}\boldsymbol{Q}_{i:j}&\boldsymbol{A}_{i:j}^{\top}\end{bmatrix} can be permuted to:

[QjIAj1Qj1IAi+1Qi+1IAiQi].\displaystyle\begin{bmatrix}Q_{j}&I\\ &-A_{j-1}&Q_{j-1}&I\\ &&&\ddots\\ &&&-A^{\top}_{i+1}&Q_{i+1}&I\\ &&&&&-A^{\top}_{i}&Q_{i}\end{bmatrix}. (10)

We apply block row and column operations (as those of Lemma 6) uniformly bounded many times to obtain:

[IAj1:iQjAiQi+1Qi].\displaystyle\begin{bmatrix}I\\ &A^{\top}_{j-1:i}Q_{j}&\cdots&A^{\top}_{i}Q_{i+1}&Q_{i}\\ \end{bmatrix}. (11)

From Proposition 10 and (No,γo)(N_{o},\gamma_{o})-uniform observability of ({Ai}i=0N1,{Qi}i=0N)(\{A_{i}\}_{i=0}^{N-1},\{Q_{i}\}_{i=0}^{N}), we have that

({Ai}i=N10,{Qi}i=N0)(\{A^{\top}_{i}\}_{i=N-1}^{0},\{Q_{i}\}_{i=N}^{0}) (12)

is (No,γo)(N_{o},\gamma_{o})-uniformly controllable. We thus have that the matrix in (11) has min(1,γo)\min(1,\gamma_{o})-uniformly lower bounded smallest non-trivial singular value. This implies that the smallest non-trivial singular value of the matrix in (10) is uniformly lower bounded by γ\gamma^{\prime}, where γ\gamma^{\prime} is given by a function of KK, NoN_{o}, γo\gamma_{o}. Therefore, we have that: 𝒙i:j𝑸i:j𝒙i:j+𝒖i:j1𝑹i:j1𝒖i:j1γ(𝒙i:j2+𝒖i:j12)\boldsymbol{x}_{i:j}^{\top}\boldsymbol{Q}_{i:j}\boldsymbol{x}_{i:j}+\boldsymbol{u}_{i:j-1}^{\top}\boldsymbol{R}_{i:j-1}\boldsymbol{u}_{i:j-1}\geq\gamma(\|\boldsymbol{x}_{i:j}\|^{2}+\|\boldsymbol{u}_{i:j-1}\|^{2}), where γ:=min(γ/K,κγ,r/2)\gamma:=\min(\gamma^{\prime}/K,\kappa\gamma^{\prime},r/2). One can observe that RiγIR_{i}\succeq\gamma I for any i𝕀[0,N1]i\in\mathbb{I}_{[0,N-1]}. Consequently, the smallest eigenvalues of the matrix in (9) for any i,j𝕀[0,N1]i,j\in\mathbb{I}_{[0,N-1]} with No|ij|2NoN_{o}\leq|i-j|\leq 2N_{o} and RiR_{i} are γ\gamma-uniformly lower bounded. Thus, by Shin et al. (2021, Lemma 4.14), (5) holds. One can confirm that γ\gamma is a function of K,No,γo,rK,N_{o},\gamma_{o},r. ∎

Finally, we show that uniform boundedness of system matrices implies uBLH.

Lemma 14.

If {Qi}i=0N\{Q_{i}\}_{i=0}^{N}, {Ri}i=0N1\{R_{i}\}_{i=0}^{N-1}, {Si}i=0N1\{S_{i}\}_{i=0}^{N-1}, {Ai}i=0N1\{A_{i}\}_{i=0}^{N-1}, {Bi}i=0N1\{B_{i}\}_{i=0}^{N-1}, {Ei}i=0N1\{E_{i}\}_{i=0}^{N-1}, {Fi}i=0N1\{F_{i}\}_{i=0}^{N-1}, {Gi}i=0N1\{G_{i}\}_{i=0}^{N-1}, and TT are KK-uniformly bounded above, (3) holds, where L<L<\infty is a function of KK.

{pf}

Uniform boundedness of the system matrices implies that for any i,j𝕀[1,N]i,j\in\mathbb{I}_{[-1,N]}, wiξi20:N(w1:N;d1:N)\nabla^{2}_{w_{i}\xi_{i}}\mathcal{L}_{0:N}(w^{\star}_{-1:N};d^{\star}_{-1:N}) are uniformly bounded above by 4K4K (Shin et al. (2021, Lemma 4.6)). Furthermore, by inspecting the problem structure, we can see that wiξj20:N(w1:N;d1:N)1\|\nabla^{2}_{w_{i}\xi_{j}}\mathcal{L}_{0:N}(w^{\star}_{-1:N};d^{\star}_{-1:N})\|\leq 1 for iji\neq j (there is only one identity block). Thus, wiξj20:N(w1:N;d1:N)max(4K,1)\|\nabla^{2}_{w_{i}\xi_{j}}\mathcal{L}_{0:N}(w^{\star}_{-1:N};d^{\star}_{-1:N})\|\leq\max(4K,1). By noting that the maximum graph degree D=2D=2 and applying Shin et al. (2021, Lemma 4.5), we have that w1:Nξ1:N20:N(w1:N;d1:N)\nabla^{2}_{w_{-1:N}\xi_{-1:N}}\mathcal{L}_{0:N}(w^{\star}_{-1:N};d^{\star}_{-1:N}) is 4max(4K,1)4\max(4K,1)-uniformly bounded above. We set L:=4max(4K,1)L:=4\max(4K,1).∎ We now state EDS in terms of uniformly bounded system matrices and uniform controllability/observability.

Assumption 15.

Given twice continuously differentiable functions {i()}i=0N\{\ell_{i}(\cdot)\}_{i=0}^{N}, {fi()}i=0N1\{f_{i}(\cdot)\}_{i=0}^{N-1}, and data d1:Nd^{\star}_{-1:N}, there exists a primal-dual solution w1:Nw_{-1:N}^{\star} of P0:N(d1:N)P_{0:N}(d_{-1:N}^{\star}) at which the assumptions in Lemma 12, 13, 14 hold.

Corollary 16.

Under Assumption 15, there exist uniform constants Υ>0\Upsilon>0 and ρ(0,1)\rho\in(0,1) (functions of KK, rr, NcN_{c}, βc\beta_{c}, NoN_{o}, γo\gamma_{o}, δ\delta) and neighborhoods 𝔻1:N\mathbb{D}^{\star}_{-1:N} of d1:Nd_{-1:N}^{\star} and 𝕎1:N\mathbb{W}^{\star}_{-1:N} of w1:Nw_{-1:N}^{\star} such that (2) holds for any d1:N,d1:N𝔻1:Nd_{-1:N},d^{\prime}_{-1:N}\in\mathbb{D}^{\star}_{-1:N} and i𝕀[1,N]i\in\mathbb{I}_{[-1,N]}.

{pf}

From Theorem 7 and Lemma 12, 13, 14. ∎

2.3 Time-Invariant Setting

Assume now that the system is time-invariant and focus on a region around a steady-state. A corollary of Theorem 7 for such a setting is derived. We present this result since this setting has been of particular interest in the MPC literature. Consider a time-invariant system with a stage-cost function ()\ell(\cdot), initial regularization function b()\ell_{b}(\cdot), terminal cost function f()\ell_{f}(\cdot), and dynamic mapping f()f(\cdot). The DO problem is given by (1) with fi()=f()f_{i}(\cdot)=f(\cdot) for i𝕀[0,N1]i\in\mathbb{I}_{[0,N-1]}, i()=()\ell_{i}(\cdot)=\ell(\cdot) for i𝕀[1,N1]i\in\mathbb{I}_{[1,N-1]}, 0(x,u)=(x,u;d)+b(x;d)\ell_{0}(x,u)=\ell(x,u;d)+\ell_{b}(x;d), and N(x;d)=f(x;d)\ell_{N}(x;d)=\ell_{f}(x;d). The steady-state optimization problem is:

minx,u\displaystyle\min_{x,u}\; (x,u;d)s.t.x=f(x,u;d)|λ.\displaystyle\ell(x,u;d)\;\mathop{\text{s.t.}}\;x=f(x,u;d)\quad|\quad\lambda. (13)

For given dsd^{s} and an associated primal-dual solution ws:=[xs;us;λs]w^{s}:=[x^{s};u^{s};\lambda^{s}] of (13), we define:

Q\displaystyle Q :=xx2(ws;ds),S:=xu2(ws;ds),\displaystyle:=\nabla^{2}_{xx}\mathcal{L}(w^{s};d^{s}),\,S:=\nabla^{2}_{xu}\mathcal{L}(w^{s};d^{s}),
R\displaystyle R :=uu2(ws;ds),A:=xf(zs;ds),B:=uf(zs;ds),\displaystyle:=\nabla^{2}_{uu}\mathcal{L}(w^{s};d^{s}),\,A:=\nabla_{x}f(z^{s};d^{s}),\,B:=\nabla_{u}f(z^{s};d^{s}),

where (w;d):=f(z;d)λx+λf(z;d)\mathcal{L}(w;d):=f(z;d)-\lambda^{\top}x+\lambda^{\top}f(z;d); for the initial and terminal cost functions b()\ell_{b}(\cdot) and f()\ell_{f}(\cdot), we define:

λb\displaystyle\lambda_{b} :=xb(xs;ds),Qb:=xx2b(xs;ds)\displaystyle:=\nabla_{x}\ell_{b}(x^{s};d^{s}),\;Q_{b}:=\nabla^{2}_{xx}\ell_{b}(x^{s};d^{s})
λf\displaystyle\lambda_{f} :=xf(xs;ds),Qf:=xx2f(xs;ds).\displaystyle:=\nabla_{x}\ell_{f}(x^{s};d^{s}),\;Q_{f}:=\nabla^{2}_{xx}\ell_{f}(x^{s};d^{s}).

The quantities defined above (QQ, RR, etc.) are independent of NN since wsw^{s} can be determined independently of NN.

Assumption 17.

Given twice continuously differentiable ()\ell(\cdot), b()\ell_{b}(\cdot), f()\ell_{f}(\cdot), f()f(\cdot), and data dsd^{s}, there exists a steady-state solution wsw^{s}, at which QfQ0Q_{f}\succeq Q\succeq 0, Qb0Q_{b}\succeq 0, S=0S=0, R0R\succ 0, (A,B)(A,B) controllable, (A,Q)(A,Q) observable, TT0TT^{\top}\succ 0, λb+λsRange(T)\lambda_{b}+\lambda^{s}\in\text{Range}(T^{\top}) and λf=λs\lambda_{f}=\lambda^{s} hold.

Corollary 18.

Under the time invariance setting and Assumption 17, there exist uniform constants Υ>0\Upsilon>0 and ρ(0,1)\rho\in(0,1) such that the following holds: for any N𝕀0N\in\mathbb{I}_{\geq 0}, there exist neighborhoods 𝔻1:Ns\mathbb{D}^{s}_{-1:N} of d1:Ns:=[Txs;ds;;ds]d^{s}_{-1:N}:=[Tx^{s};d^{s};\cdots;d^{s}] and 𝕎1:Ns\mathbb{W}^{s}_{-1:N} of w1:Ns:=[λ1s;ws;;ws;xs]w_{-1:N}^{s}:=[\lambda^{s}_{-1};w^{s};\cdots;w^{s};x^{s}] such that (2) holds for any d1:N,d1:N𝔻1:Nsd_{-1:N},d^{\prime}_{-1:N}\in\mathbb{D}^{s}_{-1:N}, where λ1s\lambda^{s}_{-1} is the solution of Tλ1s=λb+λsT^{\top}\lambda^{s}_{-1}=\lambda_{b}+\lambda^{s}.

{pf}

From the existence (follows from b+sRange(T)\ell_{b}+\ell^{s}\in\text{Range}(T^{\top})) and uniqueness (follows from TT0TT^{\top}\succ 0) of the solution of Tλ1s=λb+λsT^{\top}\lambda^{s}_{-1}=\lambda_{b}+\lambda^{s}, we have well-defined λ1s\lambda^{s}_{-1}. From Tλ1s=λb+λsT^{\top}\lambda^{s}_{-1}=\lambda_{b}+\lambda^{s} and the optimality of wsw^{s} for (13), w1:Nsw^{s}_{-1:N} satisfies the first-order optimality conditions for P0:N(d1:Ns)P_{0:N}(d_{-1:N}^{s}). Furthermore, all the assumptions in Lemma 14 are satisfied with some uniform constant KK because ()\ell(\cdot), b()\ell_{b}(\cdot), f()\ell_{f}(\cdot), f()f(\cdot), TT, wsw^{s}, and dsd^{s} are independent of NN; thus, by Lemma 14, we have (3) for a uniform constant L<L<\infty. Moreover, TTδITT^{\top}\succeq\delta I holds for some uniform constant δ>0\delta>0, and RirIR_{i}\succeq rI for i𝕀[0,N1]i\in\mathbb{I}_{[0,N-1]} with some uniform constant r>0r>0, since ()\ell(\cdot), wsw^{s}, ds,Td^{s},T are independent of NN. Similarly, (A,B)(A,B) controllability implies (Nc,βc)(N_{c},\beta_{c})-uniform controllability of ({Ai}i=1N1,{Bi}i=0N1)(\{A_{i}\}_{i=1}^{N-1},\{B_{i}\}_{i=0}^{N-1}) with some uniform constant Nc,βcN_{c},\beta_{c}, and (A,Q)(A,Q) observability implies (No,γo)(N_{o},\gamma_{o})-uniform observability of ({Ai}i=0N1,{Qi}i=0N)(\{A_{i}\}_{i=0}^{N-1},\{Q_{i}\}_{i=0}^{N}) for some uniform constants No,γoN_{o},\gamma_{o} (for now, we assume that Qb=0Q_{b}=0 and Qf=QQ_{f}=Q). From Lemma 12, 13, we have (5) and (6) for uniform β,γ>0\beta,\gamma>0. Now, observe that (5) for Qb=0Q_{b}=0 and Qf=QQ_{f}=Q implies (5) for any Qb0Q_{b}\succeq 0 and QfQQ_{f}\succeq Q; thus, we have (5) with uniform γ>0\gamma>0 for any Qb,QfQ_{b},Q_{f}. Since the first and second order conditions of optimality and constraint qualifications are satisfied, w1:Nsw^{s}_{-1:N} is a strict minimizer for P0:N(d1:N)P_{0:N}(d_{-1:N}). Since we have (3), (5), and (6) with uniform L,γ,βL,\gamma,\beta, we have uBLH, uLICQ, and uSOSC at (w1:Ns,d1:Ns)(w^{s}_{-1:N},d^{s}_{-1:N}). By applying Theorem 7, we can obtain (2). Lastly, since the parameters K,r,Nc,βc,No,γoK,r,N_{c},\beta_{c},N_{o},\gamma_{o} are independent of NN, so do Υ\Upsilon and ρ\rho.

Initial and terminal cost functions that satisfy Assumption 17 can be constructed as:

b(x;d)\displaystyle\ell_{b}(x;d) :=((IT+T)λs)x\displaystyle:=-((I-T^{+}T)\lambda^{s})^{\top}x
f(x;d)\displaystyle\ell_{f}(x;d) :=(xxs)Q(xxs)+(λs)x,\displaystyle:=(x-x^{s})^{\top}Q(x-x^{s})+(\lambda^{s})^{\top}x,

where ()+(\cdot)^{+} is the pseudoinverse of the argument. One can observe that b()\ell_{b}(\cdot) can be set to constantly zero if T=IT=I.

3 Numerical Results

Refer to caption Refer to caption
Refer to caption Refer to caption
Refer to caption Refer to caption
Figure 1: Base and perturbed solutions. Left: Case 1 (q=1,b=1q=1,b=1). Right: Case 2 (q=0,b=0q=0,b=0).

We illustrate the results of Theorem 7 and of Corollaries 16, 18. In this study, we solve the problem with base data d1:Nd^{\star}_{-1:N} to obtain the base solution w1:Nw^{\star}_{-1:N}. We then solve a set of problems with perturbed data; in each of these problems, a random perturbation Δdj\Delta d_{j} is introduced at a selected time stage jj, while the rest of the data do not have perturbation (i.e., Δdi=0\Delta d_{i}=0 for iji\neq j). The obtained solutions w1:N(d1:N+Δd1:N)w^{\dagger}_{-1:N}(d^{\star}_{-1:N}+\Delta d_{-1:N}) for the perturbed problems are visualized along with the base solution w1:Nw^{\star}_{-1:N}. The scripts can be found here https://github.com/zavalab/JuliaBox/tree/master/SensitivityNMPC. We consider a quadrotor motion planning problem (Hehn and D’Andrea, 2011) with the time-invariant setting; the cost functions are given by:

(z;d):=\displaystyle\ell(z;d):= (xd)Q(xd)+uRu\displaystyle(x-d)^{\top}Q(x-d)+u^{\top}Ru
f(x;d):=\displaystyle\ell_{f}(x;d):= (xd)Qf(xd),b(x;d)=0,\displaystyle(x-d)^{\top}Q_{f}(x-d),\quad\ell_{b}(x;d)=0,

where Q:=diag(1,1,1,q,q,q,1,1,1)Q:=\mathop{\text{diag}}(1,1,1,q,q,q,1,1,1), R:=IR:=I, Qf:=IQ_{f}:=I, and T:=IT:=I; and the dynamic mapping is obtained from:

d2Xdt2\displaystyle\frac{d^{2}X}{dt^{2}} =a(cosγsinβcosα+sinγsinα)\displaystyle=a(\cos\gamma\sin\beta\cos\alpha+\sin\gamma\sin\alpha) (14a)
d2Ydt2\displaystyle\frac{d^{2}Y}{dt^{2}} =a(cosγsinβsinαsinγcosα)\displaystyle=a(\cos\gamma\sin\beta\sin\alpha-\sin\gamma\cos\alpha) (14b)
d2Zdt2\displaystyle\frac{d^{2}Z}{dt^{2}} =acosγcosβg\displaystyle=a\cos\gamma\cos\beta-g (14c)
dγdt\displaystyle\frac{d\gamma}{dt} =(bωXcosγ+ωYsinγ)/cosβ\displaystyle=(b\omega_{X}\cos\gamma+\omega_{Y}\sin\gamma)/\cos\beta (14d)
dβdt\displaystyle\frac{d\beta}{dt} =bωXsinγ+ωYcosγ\displaystyle=-b\omega_{X}\sin\gamma+\omega_{Y}\cos\gamma (14e)
dαdt\displaystyle\frac{d\alpha}{dt} =bωXcosγtanβ+ωYsinγtanβ+ωZ,\displaystyle=b\omega_{X}\cos\gamma\tan\beta+\omega_{Y}\sin\gamma\tan\beta\ +\omega_{Z}, (14f)

where the state and control variables are defined as: x:=(X,X˙,Y,Y˙,Z,Z˙,γ,β,α)x:=(X,\dot{X},Y,\dot{Y},Z,\dot{Z},\gamma,\beta,\alpha) and u:=(a,ωX,ωY,ωZ)u:=(a,\omega_{X},\omega_{Y},\omega_{Z}). We use qq and bb as parameters that influence controllability and observability. In particular, the system becomes less observable if qq becomes small and the system loses controllability as bb becomes small (the effect of manipulation on ωX\omega_{X} becomes weak). We have empirically tested the sensitivity behavior for q=b=1q=b=1 (Case 1) and q=b=0q=b=0 (Case 2). One can see that some of the assumptions (e.g., Si=0S_{i}=0 in Corollary 16) may be violated, but one can also see that, qualitatively, the system is more observable and controllable in Case 1 than in Case 2. The results are presented in Figure 1. The base trajectories are shown as dashed lines, the perturbed trajectories are shown as solid gray lines, and the perturbed stages are highlighted using vertical lines. We can see that, for Case 1 (q=1,b=1q=1,b=1), the differences between the base and perturbed solutions become small as moving away from the perturbation point (EDS holds). On the other hand, for Case 2 (q=0,b=0q=0,b=0) one cannot observe EDS; this confirms that observability and controllability induce EDS.

4 Conclusions

We have shown that uniform controllability and observability provide sufficient conditions for exponential decay of sensitivity in dynamic optimization. As part of future work, we will aim to establish exponential decay of sensitivity under mesh refinement settings and will aim to establish formal connections with continuous-time results.

References

  • Dontchev and Rockafellar (2009) Dontchev, A.L. and Rockafellar, R.T. (2009). Implicit functions and solution mappings, volume 543. Springer.
  • Grüne et al. (2019) Grüne, L., Schaller, M., and Schiela, A. (2019). Sensitivity analysis of optimal control for a class of parabolic PDEs motivated by model predictive control. SIAM Journal on Control and Optimization, 57(4), 2753–2774.
  • Grüne et al. (2020a) Grüne, L., Schaller, M., and Schiela, A. (2020a). Abstract nonlinear sensitivity and turnpike analysis and an application to semilinear parabolic PDEs. arXiv preprint arXiv:2008.13001.
  • Grüne et al. (2020b) Grüne, L., Schaller, M., and Schiela, A. (2020b). Exponential sensitivity and turnpike analysis for linear quadratic optimal control of general evolution equations. Journal of Differential Equations, 268(12), 7311–7341.
  • Hehn and D’Andrea (2011) Hehn, M. and D’Andrea, R. (2011). A flying inverted pendulum. In 2011 IEEE International Conference on Robotics and Automation, 763–770. IEEE.
  • Na and Anitescu (2020a) Na, S. and Anitescu, M. (2020a). Exponential decay in the sensitivity analysis of nonlinear dynamic programming. SIAM Journal on Optimization, 30(2), 1527–1554.
  • Na and Anitescu (2020b) Na, S. and Anitescu, M. (2020b). Superconvergence of online optimization for model predictive control. arXiv preprint arXiv:2001.03707.
  • Na et al. (2020) Na, S., Shin, S., Anitescu, M., and Zavala, V.M. (2020). Overlapping schwarz decomposition for nonlinear optimal control. arXiv preprint arXiv:2005.06674.
  • Robinson (1980) Robinson, S.M. (1980). Strongly regular generalized equations. Mathematics of Operations Research, 5(1), 43–62.
  • Shin et al. (2021) Shin, S., Anitescu, M., and Zavala, V.M. (2021). Exponential decay of sensitivity in graph-structured nonlinear programs. arXiv preprint arXiv:2101.03067v1.
  • Shin and Zavala (2020) Shin, S. and Zavala, V.M. (2020). Diffusing-horizon model predictive control. arXiv preprint arXiv:2002.08556.