You are on page 1of 6

This article has been accepted for publication in a future issue of this journal, but has not been

fully edited. Content may change prior to final publication. Citation information: DOI 10.1109/LCSYS.2020.3001514, IEEE Control
Systems Letters

Recursive feasibility of continuous-time model


predictive control without stabilising constraints
Willem Esterhuizen1 , Karl Worthmann2 and Stefan Streif1

Abstract—We consider sampled-data Model Predictive Con- so-called cost-controllability [14], i.e., the same property as
trol (MPC) of nonlinear continuous-time control systems. We typically invoked for showing asymptotic stability. To this end,
derive sufficient conditions to guarantee recursive feasibility and a sufficiently long prediction horizon is required, which is
asymptotic stability without stabilising costs and/or constraints.
Moreover, we present formulas to explicitly estimate the required quantified in dependence of the problem data. In this paper,
length of the prediction horizon based on the concept of (local) we derive a continuous-time analogon of these results: We
cost controllability. For the linear-quadratic case, cost controlla- establish a condition that allows one to determine the length
bility can be inferred from standard assumptions. In addition, we of the prediction horizon such that recursive feasibility and
extend results on the relationship between the horizon length and asymptotic stability are guaranteed. To this end, a thorough
the distance of the initial state to the boundary of the viability
kernel from the discrete-time to the continuous-time setting. investigation of the interplay between open- and closed-loop
control is of key importance. In addition, we consider the
Index Terms—model predictive control, state constraints, re- Linear-Quadratic (LQ) case in detail, showing that the viability
cursive feasibility, asymptotic stability, sampled-data systems
kernel [15], [16] is closely related to the sublevel sets of the
finite-horizon value function.
I. I NTRODUCTION The outline of the paper is as follows: In Section II, we
present the setting, i.e. nonlinear continuous-time systems with
T WO key aspects in Model Predictive Control (MPC) are
asymptotic stability and recursive feasibility, see, e.g.
[1]. The latter property refers to existence of a solution to
state and input constraints, and the MPC algorithm. Section III
presents our main result, i.e. a sufficient condition for recursive
the optimisation problem recursively invoked in the MPC feasibility and asymptotic stability. Then, we focus on the LQ
algorithm. Both properties are often ensured by constructing case in Section IV before our findings are illustrated by an
suitable terminal costs and constraints, which are then added to example in Section V and conclusions are drawn in Section VI.
the optimisation problem. However, recently, researchers have Notation: R≥0 and R>0 to the set of nonnegative and
thoroughly analysed MPC schemes without stabilising con- positive real numbers resp. For x ∈ Rn , |x| denotes its
straints, i.e. so-called unconstrained MPC, w.r.t. their stability Euclidean norm while kF k := max|x|=1 |F x| denotes the
behaviour, see, e.g. [2], [3], [4], [5] for references dealing with induced matrix norm of F ∈ Rm×n . For a set S ⊆ Rq with
discrete-time systems and [6], [7] for extensions to continuous- q ≥ n, ∂S and projRn (S) refer to its boundary and projection
time systems and the relation to the former. See [8] for a study onto Rn resp. A continuous function η : R≥0 → R≥0
of how these two approaches are linked. The main motivation belongs to class K∞ if it is strictly increasing, unbounded,
behind this conceptual shift is simplicity of the approach and η(0) = 0. The space L∞ m
loc (R≥0 , R ) denotes the set
and its reduced numerical complexity, which explains its pre- of Lebesgue-measurable functions u : R≥0 → Rm that
dominant use in industry, cp. the discussion in [9, Section 7.4]. are locally essentially bounded. Given two sets S1 ⊂ Rn
However, a rigorous treatment of recursive feasibility has and S2 ⊂ Rn , dist(S1 , S2 ) := inf x∈S1 dist(x, S2 ) where
only been done for discrete-time systems, see [9], [10], [11], dist(x, S) is the Euclidean distance from x ∈ Rn to S ⊂ Rn .
while in the continuous-time case only particular cases were
studied, see [12], [13]. On the one hand, only qualitative II. P ROBLEM SETTING
statements (sufficiently large prediction horizon) were given, We consider the continuous-time nonlinear system:
but without any estimates on the required length. On the other ẋ(t) = f (x(t), u(t)) (1)
hand, more restrictive assumptions were made, i.e. a combina-
tion of local controllability (linearized system is controllable) with initial condition x(0) = x0 , where x(t) ∈ Rn and u(t) ∈
and reachability, see [13, assumption (ii) of Proposition 1]. Rm are the state and the control at time t ∈ R≥0 resp. Let the
Indeed, we implicitly show the second assumption en passant map f : Rn × Rm → Rn be continuous and locally Lipschitz
in our line of reasoning. In the references [9], [10], [11], continuous with respect to its first argument x on Rn \ {0}.
the authors show that sublevel sets of the finite-horizon Then, for a given control function u ∈ L∞ m
loc (R≥0 , R ), there
value function can be made recursively feasible assuming exists a unique solution of the initial value problem, which is
denoted by x(t) = x(t; x0 , u), t ∈ Ix0 ,u , where Ix0 ,u denotes
1 Willem Esterhuizen & Stefan Streif, Technische Universität Chemnitz,
its maximal interval of existence, see for example [17, App.
Automatic Control and System Dynamics Laboratory, Chemnitz, Germany;
[willem.esterhuizen, stefan.streif]@etit.tu-chemnitz.de C]. Using the set E ⊆ Rn × Rm , we impose the constraints
2 Karl Worthmann, Institute for Mathematics, Technische Universität Ilme-
nau, Ilmenau, Germany; karl.worthmann@tu-ilmenau.de (x(t), u(t)) ∈ E ∀ t ∈ [0, ∞). (2)

2475-1456 (c) 2020 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission. See http://www.ieee.org/publications_standards/publications/rights/index.html for more information.
Authorized licensed use limited to: UNIVERSITY OF BIRMINGHAM. Downloaded on June 15,2020 at 05:08:40 UTC from IEEE Xplore. Restrictions apply.
This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI 10.1109/LCSYS.2020.3001514, IEEE Control
Systems Letters

Moreover, the sets U (x) := {u ∈ Rm : (x, u) ∈ E}, x ∈ X, Algorithmically, ensuring admissibility of a control function,
and X := projRn (E) = {x : U (x) 6= ∅} are introduced. With i.e. u ∈ UT (x̂), is a non-trivial task. Hence, the derivation of
x0 ∈ X and T ∈ R>0 , we let UT (x0 ) (resp. U∞ (x0 )) denote sufficient conditions such that only finitely many inequality
the set of all u ∈ L∞ m
loc (R≥0 , R ) such that the solution exists constraints have to be checked to ensure admissibility is of
and satisfies the constraint (2) on [0, T ] (resp. for all t ≥ 0). particular interest, see, e.g. [18, Lemma 1] for an example
We will refer to the set UT (x0 ) (resp. U∞ (x0 )) as the set of tailored to the non-holonomic robot. One of our goals for
admissible control functions w.r.t. the initial value x0 and the future research is to alleviate this burden by using the char-
horizon T (resp. the infinite horizon). acterization of the viability kernel’s boundary based on the
Next, we define the viability kernel [16] (also called ad- so-called theory of barriers [15]. There it is shown that parts
missible set in, e.g., [15]) as the set of states, for which an of the boundary are made up of integral curves of the system
admissible control exists on [0, ∞). that satisfy a minimum-like principle, which yields conditions
on the control.
Definition 1 (Viability Kernel). The viability kernel is defined
The following property is essential in order to ensure well-
by A := {x0 ∈ Rn : U∞ (x0 ) 6= ∅}.
posedness of the MPC algorithm.
Let x̄ ∈ X be a controlled equilibrium, that is, there exists
a ū ∈ U (x̄) such that f (x̄, ū) = 0 holds. Our goal is to use Definition 2. A set S ⊆ X is said to be forward invariant
MPC to asymptotically steer the state to x̄ while satisfying w.r.t. the sampled feedback law µT,δ if and only if, for each
the state and control constraints (2). To that end, we define a x̂ ∈ S, the conditions
finite-horizon cost functional • (x(t; x̂, u? ), u? (t)) ∈ E for all t ∈ [0, δ] and
Z T • x(δ; x̂, u? ) ∈ S
JT (x0 , u) := `(x(s; x0 , u), u(s)) ds, hold with u? (t) := µT,δ (t; x̂) depending on x̂ and T .
0
with continuous stage cost ` : Rn × Rm → R≥0 , and consider If the initial state is located in a forward-invariant set, then
the finite-horizon Optimal Control Problem (OCP) the state and input remain feasible on [0, δ] under the MPC
feedback, and the state is contained in this set at time δ. Then,
inf JT (x0 , u). (3)
u∈UT (x0 ) the OCP to be solved in the third step of the proposed MPC
algorithm is feasible in each iteration as can be shown by
The value function VT : X → R≥0 ∪ {+∞} is defined as
induction. In conclusion, one gets recursive feasibility of the
VT (x0 ) := inf u∈UT (x0 ) JT (x0 , u). By convention, we will say
MPC closed loop. Note that a forward-invariant set w.r.t. some
that VT (x0 ) = ∞ when UT (x0 ) = ∅. Moreover, we use the
feedback is, of course, a subset of the viability kernel but,
abbreviation VT−1 [0, C] := {x ∈ X|VT (x) ≤ C}.
in general, not vice versa. However, the viability kernel may
We impose the following assumption on `, which is, e.g.,
contain states that cannot be steered to the origin as shown in
satisfied for quadratic stage cost `(x, u) := xT Qx + uT Ru
the following example.
with positive definite weighting Q ∈ Rn×n if x̄ = 0, ū = 0.
(A1) Let there be K∞ -functions η, η satisfying Example 1. Consider the scalar system ẋ(t) = x(t) + u(t)
with the constraints |x(t)| ≤ 2 and |u(t)| ≤ 1. Then, ∂A is
η(|x − x̄|) ≤ `? (x) ≤ η(|x − x̄|) ∀x ∈ X given by the set {−1, 1}. For each x0 ∈ ∂A, u(t) ≡ −x0
using the abbreviation `? (x) := inf u∈U (x) `(x, u). with state trajectory x(t; x0 , u) ≡ x0 is the only element of
Given a time shift δ ∈ R>0 , a number N ∈ N that specifies the U∞ (x0 ). Hence, we get V∞ (x0 ) = ∞ for all x0 ∈ ∂A if the
length of the prediction (optimization) horizon, and an initial stage cost satisfies Assumption A1.
value x0 ∈ X, the MPC algorithm is as follows: Hence, our goal is to show both forward invariance (and,
1. Set the prediction horizon T ← N δ, let p ← 0 thus, recursive feasiblity of the MPC closed loop) and asymp-
2. Measure the current state x̂ = x(pδ; x0 , uMPC ) totic stability for a (hopefully) large set S ⊆ A.
3. Find a minimiser u? ∈ arg inf u∈UT (x̂) JT (x̂, u)
4. Implement uMPC (t) = u? (t), t ∈ [pδ, (p + 1)δ)
5. Set p ← p + 1 and go to step 2 III. R ECURSIVE FEASIBILITY AND ASYMPTOTIC
We assume that a minimiser of problem (3) in step 3 exists if STABILITY
UT (x0 ) 6= ∅ holds to facilitate the upcoming analysis, see for To address stability and recursive feasibility of the MPC
e.g. [9, p.56] for a detailed discussion of that assumption. closed loop, we impose cost controllability [14].
The MPC algorithm implicitly defines a sampled-data feed-
back law µT,δ : [0, δ) × X → Rm , µT,δ (t, x̂) = u? (t). (A2) There exists γ ∈ R>0 and a neighbourhood N of x̄
Here, u? ∈ UT (x̂) denotes the solution of the finite-horizon such that V∞ (x) ≤ γ`? (x) holds for all x ∈ N ∩ X.
optimal control problem depending on the initial state x̂. This Note that we employ the local version following [10], which
(open-loop) control is applied on the time interval [0, δ). The assumes cost controllability only in a neighbourhood of the
resulting closed-loop solution at time t emanating from x0 controlled equilibrium x̄. It is possible to further weaken
is denoted by xµT ,δ (t; x0 ). For practical purposes, the set of Assumption A2 by using a growth bound γ depending on
admissible control functions is often limited to sampled-data the length of the prediction horizon, see, e.g. [19] and the
with zero-order hold, e.g., u? (t) ≡ û ∈ Rm , t ∈ [0, δ). references therein for a detailed discussion. Assumption A2,

2475-1456 (c) 2020 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission. See http://www.ieee.org/publications_standards/publications/rights/index.html for more information.
Authorized licensed use limited to: UNIVERSITY OF BIRMINGHAM. Downloaded on June 15,2020 at 05:08:40 UTC from IEEE Xplore. Restrictions apply.
This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI 10.1109/LCSYS.2020.3001514, IEEE Control
Systems Letters

RT
˜ ds ≤ t `(s) ds + VT −t+δ (x(t; x̂, u? ))
R
in fact, holds on arbitrary sublevel sets of the finite-horizon and VT (x̃) = 0 `(s) δ
value function. To show this, note that the constant with ˜ := `(x(s; x̃, ũ? ), ũ? (s)). Hence, using (8) to replace
`(s)
Rt
M := inf `? (x) > 0 (4) δ
`(s) ds in the last inequality yields
x∈X\N
Z δ Z T
is well-defined in view of Assumption A1. VT (x̃) ≤ VT (x̂) − `(s) ds − `(s) ds + VT −t+δ (x(t))
0 t
Proposition 1. Let Assumptions A1 and A2 hold. Then, for (9)
each C ∈ R>0 such that N ∩ X ⊆ VT−1 [0, C], the following
inequality holds with β := max{C/M, γ} with x(t) := x(t; x̂, u? ).
Before we proceed, we make the following preliminary
VT (x) ≤ β`? (x) ∀ x ∈ VT−1 [0, C]. (5) considerations: Since the assumption VT (x̂) ≤ C implies
Proof. Using Inequality (4), we have VT (x) ≤ for all C ?
M ` (x)
VT −t (x(t; x̂, u? )) ≤ C for all t ∈ [0, T ], using Inequal-
x ∈ VT−1 [0, C] \ N . Moreover, from Assumption A2 we have ity (5) yields VT −t (x(t; x̂, u? ))R≤ β`? (x(t; x̂, u? )) ≤ β`(t).
t
VT (x) ≤ γ`? (x) for all x ∈ N ∩ X. Combining these two Therefore, we have VT (x̂) ≤ 0 `(s) ds + β`(t) and, thus,
RT
inequalities yields the assertion. `(s) ds ≤ β`(t) for all t ∈ [0, T ]. For any t̄ ∈ [0, T − δ],
t
We impose the following assumption, which can be consid- this implies the inequality
ered as a consistency condition ensuring that low-order terms Z t̄+δ Z T ! Z t̄+δ
of the value function can be estimated from below by using the `(s) ds dt ≤ β `(t) dt,
stage cost, cp. the framework on stage cost design presented t̄ t t̄

in [14] for further insight, see also Section IV, in which we RT R t̄+δ RT
verify Assumption A3 in the LQ case. which is, using t `(s) ds = t `(s) ds + t̄+δ `(s) ds,
(A3) For given prediction horizon T > 0 and C > 0, there equivalent to:
exists a constant C̄ ∈ R>0 such that the following Z t̄+δ Z t̄+δ ! Z T Z t̄+δ
inequality holds for all δ ∈ (0, T ] `(s) ds dt + δ `(s) ds ≤ β `(t) dt.
t̄ t t̄+δ t̄
δ`? (x̂) ≤ C̄Vδ (x̂) ∀ x̂ ∈ VT−1 [0, C].
Since the first term on the left-hand side is nonnegative, we
We now present the main contribution of the paper: the RT R t̄+δ
get δ t̄+δ `(s) ds ≤ β t̄ `(s) ds for all t̄ ∈ [0, T −δ]. Thus,
extension of [10, Theorem 4] from discrete to continuous time.
setting t̄ = pδ for p ∈ {0, 1, . . . , N − 1}, leads to
Theorem 1 shows that the (controlled) equilibrium x̄ is asymp-
totically stable w.r.t. the sampled-data feedback law µT,δ δ Nδ
Z Z (p+1)δ
with domain of attraction containing the set VT−1 [0, C]. The `(s) ds ≤ `(s) ds.
β (p+1)δ pδ
latter implies recursive feasibility of the MPC closed loop for
arbitrary initial values x0 ∈ VT−1 [0, C]. Adding the term
R Nδ
`(s) ds on both sides of this inequality
(p+1)δ
Theorem 1. Consider the system (1) and constraint (2). For reads as
given C > 0, let Assumptions A1, A2 with γ, and A3 with C̄ β+δ
Z Nδ Z Nδ
hold. Moreover, let M be defined by (4) and β := max{ MC
, γ}. `(s) ds ≤ `(s) ds.
β (p+1)δ pδ
Then, for δ ∈ (0, β), and prediction horizon T = N δ, N ∈ N,
satisfying the condition Iteratively applying this inequality until p = N − 2 to further
(  2 )  N −1 estimate the left hand side yields
C β β
max , C̄ · < 1, (6) Z Nδ N −p−1 Z N δ
Mδ δ β+δ

β+δ
`(s) ds ≥ `(s) ds
the following relaxed Lyapunov inequality holds for all x̂ ∈ pδ β (N −1)δ
VT−1 [0, C] with α = αN,δ := C̄(β/δ)2 (β/(β + δ)N −1 ):
for p ∈ {0, 1, . . . , N − 1}. Then, first taking p = 0 and
invoking x̂ ∈ VT−1 [0, C] Assumption A2, and the definition
Z δ
VT (x(δ; x̂, µT,δ )) ≤ VT (x̂) − (1 − α) `(s) ds, (7)
0
of β afterwards, yields
where `(s) := `(x(s; x̂, u? ), u? (s)). 
β+δ
N −1 Z T
?
min{β` (x̂), C} ≥ VT (x̂) ≥ `(s) ds.
Proof. Consider x̂ ∈ VT−1 [0, C], and let u? ∈ UT (x̂) be a β T −δ
solution of the OCP (3), which implies JT (x̂, u? ) ≤ JT (x̂, u) RT (10)
for all u ∈ UT (x̂). Define x̃ := x(δ; x̂, u? ), and let ũ? be the Using (the first term of) Condition (6), we get T −δ`(s) ds <
RT
respective solution of (3), i.e. JT (x̃, ũ? ) = VT (x̃).1 For any M δ, which implies T −δ `? (x(s; x̂, u? )) ds < M δ. Since
t ∈ [0, T ], we have   Z T
Z δ Z t Z T ? ?
δ inf ` (x(t; x̂, u )) ≤ `? (x(s; x̂, u? )) ds
VT (x̂) = `(s) ds + `(s) ds + `(s) ds (8) t∈[T −δ,T ] T −δ
0 δ t (11)
1 Note
that this equality also holds (VT (x̃) = ∞) if there does not exist an holds, there exists a time instant t̂ ∈ [T − δ, T ] such that the
admissible solution of the OCP (3). inequality `? (x(t̂; x̂, u? )) < M holds. As a consequence, the

2475-1456 (c) 2020 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission. See http://www.ieee.org/publications_standards/publications/rights/index.html for more information.
Authorized licensed use limited to: UNIVERSITY OF BIRMINGHAM. Downloaded on June 15,2020 at 05:08:40 UTC from IEEE Xplore. Restrictions apply.
This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI 10.1109/LCSYS.2020.3001514, IEEE Control
Systems Letters

−1
definition of M ensures x(t̂; x̂, u? ) ∈ N , cp. (4). Therefore, the importance of considering compact sets K ⊂ V∞ [0, ∞) \
we have V∞ (x(t̂; x̂, u? )) ≤ β`? (x(t̂; x̂, u? )), which implies2 O in the next section.
In the following section, we show that some of the constants
VT −t̂+δ (x(t̂; x̂, u? )) ≤ β`? (x(t̂; x̂, u? )). used in Theorem 1 can be determined in the LQ case using
Note that the definition of t̂ and the line of reasoning used to the spectrum of the weighting matrices in the stage cost and
derive Inequality (11) imply the solution to the algebraic Riccati equation. In the nonlinear
Z T Z T settting, the system at hand has to be considered in detail, see,
? ?
δ` (x(t̂; x̂, u )) ≤ ? ?
` (x(s; x̂, u ) ds ≤ `(s) ds. e.g. [20], but is, in general, a non-trivial task.
T −δ T −δ

Thus, using (9) with t = t̄ and applying the last two IV. C ONSTRAINED L INEAR -Q UADRATIC C ASE
inequalities, we get In Section III we presented conditions rendering sublevel sets
Z δ Z T of the finite-horizon value function forward invariant w.r.t.
β T
Z
VT (x̃) ≤ VT (x̂) − `(s) ds − `(s) ds + `(s) ds. µT,δ and ensuring asymptotic stability of the equilibrium. We
0 t̂ δ T −δ
now consider the LQ case and establish a relationship between
RT the viability kernel, see Definition 1, and the level sets of the
Then, dropping the term t̂ `(s) ds, t̂ ∈ [T − δ, T ], and using
Inequality (10), we get value function. Then, we specialise Corollary 1 to the LQ case;
Z δ N −1 an important result because there exist algorithms capable of
β2

β computing compact inner-approximations of viability kernels,
VT (x̃) ≤ VT (x̂) − `(s) ds + `? (x̂).
0 δ β + δ see [21], [22]. Most of the results derived for discrete-time
systems in [10] carry over to our continuous-time setting in
Invoking Assumption A3, we get the desired relaxed Lyapunov
an analogous way. Hence, we only present novel aspects in the
inequality (7) where αN,δ is defined as in (the second term
current work due to space limitations and refer to an extended
of) Condition (6), i.e. the assertion.
ArXiv-version [23] containing the details in its appendix.
Theorem 1 relates the easily checkable Condition (6) with We focus on linear systems, i.e. system (1) given by
the Lyapunov-like decrease (7) of the finite-horizon value
function, which allows one to conclude asymptotic stability f (x(t), u(t)) := Ax(t) + Bu(t) (13)
of the equilibrium and can be, thus, considered as a sufficient with matrices A ∈ Rn×n and B ∈ Rn×m . Moreover, we
stability condition, cp. [5]. impose pure control constraints, i.e. u(t) ∈ U , U ⊂ Rm , for
For the following statement, we require the set of excep- all t ≥ 0 and the following assumptions:
−1
tional points defined by O := limn→∞ V∞ [n, ∞), cp. [10]. (A4) The pair (A, B) is stabilisable
The set O consists of all elements of the state space at which (A5) The sets U , E := {(x, u) ∈ Rn+m : g(x, u) ≤ 0}
the value function may blow up. Hence, entering the set O with g ∈ C 2 (Rn × Rm ) → Rp are convex, compact,
should be avoided. Using this definition, we can derive the and contain the origin in their interior. The mapping
following corollary of Theorem 1, which is the continuous- u 7→ gi (x, u), i ∈ {1, . . . , p}, is convex on Rn .
time analogon of [10, Thm. 6], which involves compact subsets
of the infinite-horizon value function’s sublevel sets. Proposition 2. Let the dynamics be given by (13) and Assump-
tion A5 hold. Then, the viability kernel is compact, convex, and
Corollary 1. Let Assumptions A1-A3 hold and let K ⊂ contains the origin in its interior.
−1
V∞ [0, ∞) \ O be a compact set. Then, for a chosen time shift
δ > 0 there exists a horizon length T̄ = T̄ (δ, K) ∈ (δ, ∞) Proof. The proof that the viability kernel is bounded, convex,
such that the origin is asymptotically stable w.r.t. the MPC and contains the origin in its interior easily adapts from the
closed loop with domain of attraction containing the set K. discrete-time case [10], see [23, Prop. 5 and 6]. As detailed
in [24, Prop. 3.1] and [25], to have closedness of the viability
Proof. By the same arguments as in [10], for any compact kernel one needs to impose assumptions on the dynamics f ,
−1
set K ⊂ V∞ [0, ∞) \ O, there exists a C < ∞ such that in addition to Assumption A5. These are that f is at least
−1
K ⊆ V∞ [0, C]. Therefore, from Condition (6) and with a C 2 from Rn × U1 to Rn , U1 an open subset containing
chosen time shift δ, for any initial condition x0 ∈ K the MPC U ; that there exists a constant 0 < C < ∞ such that
closed loop is stable and recursively feasible if N ≥ N̄ , i.e. supu∈U |xT f (x, u)| ≤ C(1 + kxk2 ), for all x ∈ Rn ; and that
T̄ := δ N̄ , where N̄ = N̄ (δ, K) ∈ N satisfies the set f (x, U ) is convex for all x ∈ Rn . We note that all three
max{ln(δ) + ln(M ) − ln(C), 2 ln(δ/β) − ln(C̄)} of these assumptions are satisfied by the system (13).
N̄ > + 1.
ln(β) − ln(β + δ) Remark 1. If the functions gi do not depend on the input,
(12)
that is gi : Rn → R, the set A is closed without assuming
convexity of gi (x, ·), see [15, Prop. 4.1] for details.
We stress that Corollary 1 implies recursive feasibility for all
Remark 2. Note that in the discrete-time setting with dynam-
initial values x0 contained in the compact set K. We clarify the
ics xk+1 = f (xk , uk ), a set S is said to be control invariant if,
2 We stress that the following inequality implies finiteness of V (x̃) and,
T
for each x ∈ S, there exists u ∈ U (x) such that f (x, u) ∈ S
thus, the existence of an admissible control function ũ ∈ UT (x̃). holds. Hence, establishing closedness of A requires fewer

2475-1456 (c) 2020 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission. See http://www.ieee.org/publications_standards/publications/rights/index.html for more information.
Authorized licensed use limited to: UNIVERSITY OF BIRMINGHAM. Downloaded on June 15,2020 at 05:08:40 UTC from IEEE Xplore. Restrictions apply.
This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI 10.1109/LCSYS.2020.3001514, IEEE Control
Systems Letters

assumptions: Briefly, if g is continuous, closedness can be di- minimal and maximal eigenvalue of the matrix Q, respectively.
rectly shown by considering the limit of converging sequences Invoking Assumptions A4 and A5 imply that the unique,
{xk }k∈N ⊆ A. positive definite solution P of the algebraic Riccati equation
satisfies V∞ (x) = xT P x on a neighbourhood N of the origin.
Assumption A4 implies existence of a matrix F such that
Hence, we have
the state feedback µF (x) = F x renders the matrix (A + BF )
Hurwitz. Thus, see, e.g. [26], there exist constants Γ > 0 and σmax (P ) ?
V∞ (x) ≤ σmax (P )|x|2 ≤ ` (x)
η > 0 such that for all x0 ∈ Rn : σmin (Q)
|x(t; x0 , uF )| ≤ Γe−η(t−t0 ) |x0 | ∀t ∈ [t0 , ∞). (14) for all x ∈ N , i.e. Assumption A2.
Next, we prove Assumption A3, which trivially holds in
These facts are used to arrive at the following proposition, the discrete-time setting with C̄ = 1. To this end, we extend
which uses ideas originally proposed in [27]. the previously presented argumentation based on the algebraic
Proposition 3. Consider the system (13) under Assump- Riccati equation. Taking the constraints into account yields
tions A4 and A5. For all x ∈ λA, with λ ∈ [0, 1) there exists Vδ (x) ≥ x> P x ≥ σmin (P )|x|2 . Hence, Assumption A3 holds
a constant M (λ) such that V∞ (x) ≤ M (λ). with C̄ := T σmax (Q)/σmin (P ).

Proof. Since the proof essentially uses the same ideas as its Remark 3. The proof of Theorem 2 identifies some of the con-
discrete-time analogon, we only provide a sketch and refer to stants appearing in Theorem 1 for the LQ case: γ = σσmax (P )
min (Q)
,
[23, Proposition 7] for the details. The idea of the proof is to C̄ = δ σσmax (Q)
min (P )
≤ T σσmax (Q)
min (P )
.
consider a convex combination of two control functions: one
Finally, we state the continuous-time analogon of [10,
that keeps the state-control pair admissible for all time, and the
Cor. 15], which relates the sufficient horizon length to the
other, given by the linear-quadratic-regulator (LQR) feedback,
distance of the state to A’s boundary, see the appendix of
driving the state to the origin asymptotically. Then, upon
[23] for a detailed proof.
reaching a neighbourhood of the origin (which is guaranteed
due to the choice of the convex combination) switching to Corollary 2. Consider the system (13) under Assumptions A4
the LQR feedback, which drives the state to the origin while and A5, with the quadratic stage cost (15), and a compact set
satisfying the constraints. K ⊂ int(A). The upper bound of the infinite-horizon value
function on K is inversely proportional to the distance of the
Proposition 3 states that, for linear systems satisfying As-
state from the boundary of the viability kernel. That is, there
sumptions A4 and A5, the infinite-horizon value function is
exists a constant D, such that:
uniformly bounded on λA, λ ∈ [0, 1). On the one hand,
for given δ, this ensures asymptotic stability of the origin D
sup V∞ (x) ≤ .
w.r.t. the MPC closed loop with the interior of A contained x∈K dist(K; ∂A)
in the basin of attraction (for a sufficiently large prediction C
Moreover, for β set to max{ M dist(K;∂A) , γ}, if the time shift δ
horizon T = N δ). On the other hand, the set O (if present, √
see Example 1) is restricted to the boundary of A. is smaller than min{C/M, γ C̄} and N̄ (K, δ) satisfies (12),
Next, we state a theorem which combines Corollary 1 with the origin is asymptotically stable w.r.t. the MPC closed loop
Proposition 3 to provide a result for the LQ case. with basin of attraction S containing the set K.

Theorem 2. Consider the system (13) and suppose that In the future we intend to further explore conditions under
Assumptions A4 and A5 hold with the symmetric, quadratic, which the horizon length does not blow up as the state
and positive definite stage cost approaches the boundary of A by using the theory of barriers
   as mentioned in Section II.
T T Q N x
`(x, u) := (x , u ) . (15)
NT R u
V. E XAMPLE
Let K ⊂ int(A) be compact and δ > 0 be given. Then, the We demonstrate the growth of the sufficient horizon length N
origin is asymptotically stable w.r.t. the MPC closed loop with for a chosen time shift δ as the initial condition approaches the
basin of attraction S containing K for each prediction horizon viability kernel’s boundary. We consider the double integrator:
T = N δ such that N satisfies Condition (6).
ẋ1 (t) = x2 (t), ẋ2 (t) = u(t),
Proof. Clearly, there exists λ ∈ [0, 1) such that K ⊂ λA.
Hence, Proposition 3 implies boundedness of VT on K. Con- with |u| ≤ 1 and |xi | ≤ 1, i ∈ {1, 2}. As shown in [15], parts
−1
sequently, K ⊆ V∞ [0, ∞) \ O holds. Hence, Theorem 1 and of the boundary of the viability kernel (called the barrier)
Corollary 1 imply the assertion supposing that Assumptions consist of integral curves that, together with a particular
A1-A3 hold. control function, satisfy a minimum-like principle and intersect
Validity of Assumptions A1 and A2 can be shown analo- the boundary of the constrained state space tangentially. We
gously to the discrete-time case, cp. [10, Thm. 13]: Since the use this fact to construct the two solid curves that form the
cost is quadratic and positive definite, Assumption A1 holds. barrier, labelled [∂A]− , see Figure 1.
In particular, we have `? (x) = xT Qx, and σmin (Q)|x|2 ≤ We run the continuous-time MPC algorithm (solving the
`? (x) ≤ σmax (Q)|x|2 where σmin (Q) and σmax (Q) denote the optimisation problem via a direct method) with the running

2475-1456 (c) 2020 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission. See http://www.ieee.org/publications_standards/publications/rights/index.html for more information.
Authorized licensed use limited to: UNIVERSITY OF BIRMINGHAM. Downloaded on June 15,2020 at 05:08:40 UTC from IEEE Xplore. Restrictions apply.
This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI 10.1109/LCSYS.2020.3001514, IEEE Control
Systems Letters

R EFERENCES
[1] J. B. Rawlings, D. Q. Mayne, M. M. Diehl, Model Predictive Control:
Theory, Computation, and Design, 2nd ed. Nob Hill Publishing, 2018.
[2] V. Nevistić and J. A. Primbs, “Receding horizon quadratic optimal
control: performance bounds for a finite horizon strategy,” in Proc.
European Control Conf. (ECC), 1997, pp. 3584–3589.
[3] A. Jadbabaie, J. Yu, and J. Hauser, “Unconstrained receding-horizon
control of nonlinear systems,” IEEE Trans. Automat. Control, vol. 46,
no. 5, pp. 776–783, 2001.
[4] G. Grimm, M. J. Messina, S. E. Tuna, A. R. Teel, “Model predictive
control: for want of a local control Lyapunov function, all is not lost,”
IEEE Trans. Automat. Control, vol. 50, no. 5, pp. 546–558, 2005.
[5] L. Grüne, J. Pannek, M. Seehafer, and K. Worthmann, “Analysis of
unconstrained nonlinear MPC schemes with varying control horizon,”
SIAM J. Control Optim., vol. 48, no. 8, pp. 4938–4962, 2010.
[6] M. Reble and F. Allgöwer, “Unconstrained model predictive control
and suboptimality estimates for nonlinear continuous-time systems,”
Fig. 1. Simulation of the MPC closed loop from various initial conditions Automatica, vol. 48, no. 8, pp. 1812 – 1817, 2012.
with δ = 0.1. Red curve: x0 = (0.5, 0.5)T , N = 4; blue curve: x0 = [7] K. Worthmann, M. Reble, L. Grüne, and F. Allgöwer, “The role of
(0.6, 0.6)T , N = 4; green curve: x0 = (0.7, 0.7)T , N = 5; magenta curve: sampling for stability and performance in unconstrained nonlinear model
x0 = (0.733, 0.73)T , N = 7. predictive control,” SIAM J. Control Optim., vol. 52, no. 1, pp. 581–605,
2014.
[8] M. S. Darup and M. Cannon, “A missing link between nonlinear
MPC schemes with guaranteed stability,” in Proc. 54th IEEE Conf. on
cost `(x, u) = x21 + x22 + u2 from various initial conditions Decision and Control (CDC), 2015, pp. 4977–4983.
[9] L. Grüne and J. Pannek, Nonlinear Model Predictive Control. Theory
that approach the boundary of A, with various time shifts δ and Algorithms, 2nd ed. London: Springer, 2017.
and horizons N . Table I displays the smallest N = N (x0 , δ), [10] A. Boccia, L. Grüne, and K. Worthmann, “Stability and feasibility of
for which MPC steers the particular initial state to the origin state constrained MPC without stabilizing terminal constraints,” Systems
& Control Letters, vol. 72, pp. 14 – 21, 2014.
while maintaining constraint satisfaction. It is interesting to [11] ——, “Stability and feasibility of state-constrained linear MPC without
note that the magenta curve, initiating from x0 ∈ [∂A]− , stabilizing terminal constraints ,” in Proc. 21st Int. Symp. Math. Theory
results from a finite horizon length, N = 7. This emphasises Networks Syst., 2014, pp. 453–460.
[12] T. Faulwasser and D. Bonvin, “On the design of economic NMPC based
that the infinite-horizon value function may be bounded on on an exact turnpike property,” IFAC-PapersOnLine, vol. 48, no. 8, pp.
the viability kernel’s boundary, and that the statement in 525–530, 2015.
Corollary 2 is merely sufficient. [13] ——, “On the design of economic NMPC based on approximate turnpike
properties,” in Proc. 54th IEEE Conf. on Decision and Control (CDC),
2015, pp. 4964–4970.
TABLE I [14] J.-M. Coron, L. Grüne, and K. Worthmann, “Model predictive control,
S MALLEST HORIZON LENGTH N (δ) FOR TIME SHIFT δ, FROM x0 . cost controllability, and homogeneity,” arXiv:1906.05112, 2019.
[15] J. De Dona and J. Lévine, “On barriers in state and input constrained
x0 δ = 0.1 δ = 0.05 δ = 0.03 nonlinear systems,” SIAM J. Control Optim., vol. 51, no. 4, pp. 3208–
3234, 2013.
(0.5, 0.5)T 4 7 10 [16] J.-P. Aubin, Viability theory. Springer Science & Business Media, 2009.
[17] E. D. Sontag, Mathematical control theory: deterministic finite dimen-
(0.6, 0.6)T 4 7 11 sional systems. Springer Science & Business Media, 2013, vol. 6.
(0.7, 0.7)T 5 10 14 [18] M. S. Darup and K. Worthmann, “Tailored MPC for mobile robots with
very short prediction horizons,” in Proc. European Control Conf. (ECC),
2018, pp. 1361–1366.
[19] K. Worthmann, “Stability analysis of unconstrained receding horizon
VI. C ONCLUSION control schemes,” Ph.D. dissertation, University of Bayreuth, 2011.
We considered MPC without (stabilising) terminal costs and/or [Online]. Available: https://epub.uni-bayreuth.de/273/
[20] K. Worthmann, M. W. Mehrez, M. Zanon, G. K. I. Mann, R. G. Gosine,
constraints for continuous-time systems. We proposed suffi- and M. Diehl, “Regulation of differential drive robots using continuous
cient conditions to ensure asymptotic stability of an equilib- time mpc without stabilizing constraints or costs,” IFAC-PapersOnLine,
rium. Simultaneously, we showed recursive feasibility for all vol. 48, no. 23, pp. 129–135, 2015.
[21] P. Saint-Pierre, “Approximation of the viability kernel,” Applied Math-
initial values contained in some compact set. In particular, we ematics and Optimization, vol. 29, no. 2, pp. 187–209, 1994.
provided a formula that allows one to determine a prediction [22] D. Monnet, J. Ninin, and L. Jaulin, “Computing an inner and an outer
horizon such that the finite-horizon value function is guaran- approximation of the viability kernel,” Reliable Computing, vol. 22,
2016.
teed to enjoy a Lyapunov-like decrease along trajectories of [23] W. Esterhuizen, K. Worthmann, and S. Streif, “Recursive feasibility
the MPC closed loop. For the linear-quadratic case, the interior of continuous-time model predictive control without stabilising con-
of the viability kernel can (essentially) be covered using this straints,” arXiv:2003.07598, 2020.
[24] W. Esterhuizen and J. Lévine, “Barriers in nonlinear control systems
technique using standard assumptions. with mixed constraints,” arXiv:1508.01708, 2015.
[25] ——, “Barriers and potentially safe sets in hybrid systems: Pendulum
ACKNOWLEDGEMENT with non-rigid cable,” Automatica, vol. 73, pp. 248 – 255, 2016.
[26] R. W. Brockett, Finite Dimensional Linear Systems. John Wiley &
This work was funded by BMBF-Projekts 05M18OCA / Sons, 1970.
05M18SIA within KONSENS: “Konsistente Optimierung und [27] R. Gondhalekar, J. Imura, and K. Kashima, “Controlled invariant fea-
sibility — a general approach to enforcing strong feasibility in MPC
Stabilisierung elektrischer Netzwerksysteme”. Karl Worth- applied to move-blocking,” Automatica, vol. 45, no. 12, pp. 2869 –
mann thanks the German Research Foundation (DFG) for their 2875, 2009.
support (Grant No. WO 2056/6-1).

2475-1456 (c) 2020 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission. See http://www.ieee.org/publications_standards/publications/rights/index.html for more information.
Authorized licensed use limited to: UNIVERSITY OF BIRMINGHAM. Downloaded on June 15,2020 at 05:08:40 UTC from IEEE Xplore. Restrictions apply.

You might also like