Generic Stability Implication from Full Information Estimation to Moving-Horizon Estimation
Abstract
Optimization-based state estimation is useful for handling of constrained linear or nonlinear dynamical systems. It has an ideal form, known as full information estimation (FIE) which uses all past measurements to perform state estimation, and also a practical counterpart, known as moving-horizon estimation (MHE) which uses most recent measurements of a limited length to perform the estimation. Due to the theoretical ideal, conditions for robust stability of FIE are relatively easier to establish than those for MHE, and various sufficient conditions have been developed in literature. This work reveals a generic link from robust stability of FIE to that of MHE, showing that the former implies at least a weaker robust stability of MHE which implements a long enough horizon. The implication strengthens to strict robust stability of MHE if the corresponding FIE satisfies a mild Lipschitz continuity condition. The revealed implications are then applied to derive new sufficient conditions for robust stability of MHE, which further reveal an intrinsic relation between the existence of a robustly stable FIE/MHE and the system being incrementally input/output-to-state stable.
Index Terms:
Nonlinear systems; moving-horizon estimation; full information estimation; state estimation; disturbances; robust stability; incremental input/output-to-state stabilityI Introduction
Optimization-based state estimation is an estimation approach which performs state estimation by solving an optimization problem. Compared to conventional approaches like Kalman filtering (KF) which deals with linear dynamical systems, and its extensions like extended KF and unscented KF which can deal with nonlinear dynamical systems based on linearization techniques, optimization-based approach has the advantage of handling linear and nonlinear dynamical systems directly and also including various physical or operational constraints [1]. The optimization formulation also admits flexible definition of the objective function to be optimized, which in some cases is necessary to accurately recover the state [2].
Optimization-based state estimation generally takes one of the two forms: full information estimation (FIE) which uses all past measurements to perform the estimation, and moving-horizon estimation (MHE) which uses most recent measurements of a limited length to perform the estimation. MHE is a practical approximate of FIE which is ideal and computationally intractable. The interest in FIE lies in two aspects: serving as a benchmark to MHE and providing useful insights for the stability analysis of MHE. Recent studies on FIE and MHE are concentrated on analyses of their stability and robustness, as are critical to guide the applications of MHE. While earlier literature assumes restrictive and idealistic conditions such as observability and/or zero or a priori known convergent disturbances [3, 4, 5, 6, 7, 8, 9], recent literature considers more practical conditions like detectability and/or in the presence of bounded disturbances [10, 11, 1].
An important progress was made in [11], which introduced the incremental input/output-to-state stability (i-IOSS) concept on detectability of nonlinear systems (as developed in [12]) to study robust stability of FIE and MHE. Given an i-IOSS system, it was shown that under mild conditions FIE is robustly stable and convergent for convergent disturbances. Whereas, conditions for FIE and MHE to be robustly stable under bounded disturbances were posted as an open research challenge. Reference [13] provided a prompt response to this challenge,111The main results were obtained in the late of 2012 though the paper was not able to be published until 2015. identifying a general set of conditions for FIE to be robustly stable for i-IOSS systems. Onwards, a series of researches have been inspired to close the challenge.
A particular type of cost functions with a max-term was investigated for a class of i-IOSS systems in [14], establishing robust stability of the FIE. The conditions were enhanced in [15], enabling robust stability also of MHE if a sufficiently long horizon is applied. The conclusion was extended to MHE without a max-term in the cost function [16]. Meanwhile, it was shown that MHE is convergent for convergent disturbances, with or without a max-term. On the other hand, reference [2] revealed an implication link from robust stability of FIE to that of MHE, and consequently identified rather general conditions for MHE to be robustly stable by inheriting conditions which ensure robust stability of the corresponding FIE. By making use of a Lipschitz continuity condition introduced in [2], reference [17] streamlined and generalized the analysis and results of [16], showing that the key is essentially to assume that the system satisfies a form of exponential detectability. When global exponential detectability is assumed, the MHE can further be shown to be robustly globally exponentially stable by implementing properly time-discounted stage costs [18].
The reviewed robust stabilities of FIE and MHE were concluded based on the concept defined in [11], which is however found to be flawed in that such defined robustly stable estimator does not imply convergence for convergent disturbances [19, 20]. This motivates a necessary modification of the stability concept in [19], which redefines the estimate error bound with a worst-case time-discounted instead of uniformly-weighted impact of the disturbances. The new concept is an enhancement of the old one, and was shown to be compatible with the original i-IOSS detectability of a system, which is necessary for establishing robust stability of both FIE and MHE. With this new concept, it becomes straightforward to understand the earlier robust local/global exponential stability results reported of FIE and MHE [16, 18]. It also motivates some new results as developed in [1, Ch. 4], which introduced a kind of stabilizability condition to establish robust stability of FIE and MHE. Despite conceptual elegance, the new condition involves an inequality which is uneasy to verify in general.
Motivated by the new and stronger concept of robust stability, this work aims to motivate general sufficient conditions for robust stability of MHE by firstly establishing a generic implication link from robust stability of FIE to that of MHE, and then transforming the challenge into identifying sufficient conditions for ensuring robust stability of FIE. While the reasoning approach is inspired by the ideas introduced in [2], the contributions of this work are three-fold:
-
•
A generic implication is established from robust stability of FIE to practical robust stability of the corresponding MHE, and the implication becomes stronger to robust stability of the MHE if the FIE admits a certain Lipschitz continuity property;
-
•
Explicit computations of sufficient MHE horizons are provided for the aforementioned implications to be effective in a local or global sense of robust stability. Analyses on two exemplary cases further illustrate that the upper bound of the MHE estimate error can eventually decrease with respect to (w.r.t.) the horizon size once it is large enough and satisfies certain conditions.
-
•
Given the revealed implications, new sufficient conditions are derived for robust stability of MHE by firstly developing those for the corresponding FIE. An interesting finding is that a system being i-IOSS is necessary but also sufficient for existence of a robustly stable FIE. Consequently is a similar but weaker conclusion applicable to MHE.
The remaining of this work is organized as follows. Sec. II introduces notation, setup and necessary preliminaries. Sec. III defines general forms of FIE and MHE, and introduces robust stability concepts. Sec. IV reveals the implication from global (or local) robust stability of FIE to that of the corresponding MHE. Sec. V applies the implication to establish robust stability of MHE, by firstly developing conditions for ensuring robust stability of the corresponding FIE. Finally, Sec. VI concludes the work.
II Notation, Setup and Preliminaries
The notation mostly follows the convention in [2, 1]. The symbols , and denote the sets of real numbers, nonnegative real numbers and nonnegative integers, respectively, and denotes the set of integers from to . The constraints and are used interchangeably to refer to the set of discrete times. The symbol denotes the Euclidean norm of a vector. The bold symbol , denotes a sequence of vector-valued variables , and with a function acting on a vector , stands for the sequence of function values . Given a scalar function , any and any scalar function , the notation refers to the fold composition of function by function subject to the given second argument , and equals the identity function.
Throughout the paper, refers to a discrete time, and as a subscript it indicates dependence on time . Whereas, the subscripts or superscripts , and are used exclusively to indicate a function or variable that is associated with the state (), process disturbance () or measurement noise (). The symbols and refer to the integers that are closest to from below and above, respectively. Given two scalars and , let . The operator is both associative and commutative, i.e., and , and furthermore is distributive with respect to (w.r.t.) increasing functions. That is, if the function is increasing in the argument. The frequently used , and functions are defined as follows.
Definition 1.
(, and functions) A function is a function if it is continuous, zero at zero, and strictly increasing. A function is a function if it is continuous, nonincreasing and satisfies as . A function is a function if, for each , is a function and for each , is a function.
Consider a discrete-time system described by
(1) |
where is the system state, the process disturbance, the measurement, the measurement disturbance, all at time . Here we study state estimation as an independent subject, and so control inputs (if there were any) known up to the estimation time are treated as given constants, which do not cause difficulty to later defined optimization and related analyses and hence are neglected in the problem formulation for brevity [11, 1]. The functions and are assumed to be continuous and known. The initial state and the disturbances are assumed to be unknown but bounded.
Consider two state trajectories. Let
(4) |
for all , where corresponds to if additive measurement noises are present while identical measurements are assumed. Hence collects a sequence of deviation vectors, and its domain is denoted as . Let extract the time index (i.e., the original index above) of . Specifically, we have
(5) |
Note that the time index of is defined as which refers to the time associated with a priori information.
Definition 2.
(Concise definition of i-IOSS) The system , is i-IOSS if there exists such that for every two initial states , and two sequences of disturbances , the following inequality holds for all :
(6) |
where is a shorthand of for . Furthermore, the system is exp-i-IOSS if in the above inequality admits an exponential form as with certain , and for all .
As proved in [19], this definition of i-IOSS is equivalent to the original one introduced in [12] which applies instead of function to each uncertainty for all .
The following defines Lipschitz continuity of a function at a given point, and presents a related proposition.
Definition 3.
(Lipschitz continuity at a point) Given a subset , a function is said to be Lipschitz continuous at a point over the subset if there is a constant such that for all .
Proposition 1.
(Properties related to Lipschitz continuity) Given any , , and , define . The following conclusions hold: a) the minimum is well defined, i.e., it exists; b) the inequality holds for all ; and c) if is Lipschitz continuous at the origin over the range , then a) and b) hold also with .
Proof:
a) With , if we take , which certainly exists and has a finite value by the definition of a function, then we must have for all . That is, is a feasible solution of and hence also an upper bound for the considered minimization problem. Consequently, it is sufficient to consider the decision variable in the finite set . By enumerating in the set , the inequality can be checked for all subject to each given . All valid values of can then be collected in a set . Since is nonempty and finite, there must exist a minimum value, that is, is well defined. This completes the proof of a).
b) Since the inequality holds for all by the definition of , it is remaining to show that it holds also for all . In the latter case, we have , where the first inequality owes to the property of a function and the second results from a). This completes the proof of b).
c) Since is a function, being Lipschitz continuous at the origin over implies that for all and any , where is the Lipschitz constant which is nonincreasing in and furthermore goes to zero as , i.e., . Consequently, is well defined, which implies that for all . Therefore, in this case both a) and b) hold true for all , which includes the origin. ∎
III Optimization-based State Estimation
Consider the system described in (1). Given a present time , the state estimation problem is to find an optimal estimate of state based on historical measurements for in a time set. Ideally, all measurements up to time are used, leading to the so-called FIE; and practically, only measurements are used within a limited distance backwards from time , yielding the so-called MHE. Both FIE and MHE can be cast as optimization problems.222Readers are referred to [11] for a brief introduction of their connection to control problems and difference from probabilistic formulations. To be concise, we will first define MHE and then treat FIE as a variant.
Let MHE implement a moving horizon of size . The decision variables are denoted as , which correspond to the system variables .333As in [1], the last measurement is not considered for ease of presentation, though the inclusion does not change the conclusions. And let the optimal decision variables be . Since is uniquely determined from and , the decision variables essentially reduce to .
In addition, let be a priori estimate of , and in particular be bounded. Without loss of generality, the prior estimates of the disturbances are assumed to be zero. Denote the cost function as , which penalizes uncertainties in the initial state, the process and the measurements. Then, the MHE instance at time is defined by the following optimization problem:
(7) | ||||
s.t., | ||||
As is uniquely determined by and , it is kept mainly for the convenience of expressing the disturbance set and the objective function. Since the global optimal solution , for any , is dependent on time when the MHE instance is defined, to be unambiguous we use to represent that is solved from the instance defined at time . This keeps unchanged, while the realization varies as MHE renews itself in time.
To define FIE, it suffices to adapt the horizon size in the MHE formulation to taking the time-varying value . The yielded FIE has the form of an MHE but with complete data originating from the zero initial time. To link them easily, an FIE is called the corresponding FIE of an MHE based on which the FIE is derived, and conversely is the MHE called a corresponding MHE of the FIE. Since it becomes computationally intractable as time elapses, FIE is studied mainly for its theoretical interest: its performance is viewed as a limit or benchmark that MHE attempts to approach, and its stability can be a good start point for analysis of MHE.
An important issue in designing FIE or MHE is to identify conditions under which the associated optimization admits optimal estimates that satisfy a robust stability property defined below. Let denote a state sequence generated from an initial state , and a disturbance sequence . In addition, define a bounded set for any given :
(8) |
Definition 4.
(Robust stable estimation) The estimate of state is based on partial or full sequence of the noisy measurements, . The estimate is robustly asymptotically stable (RAS) if given any , there exist functions such that the following inequality holds for all , , and :
(9) |
The estimate is further said to be robustly globally asymptotically stable (RGAS) if the inequality is satisfied for all . Moreover, if the estimate is RAS (or RGAS) and the function admits an exponential form as with certain , and for all , then the estimate is said to be robustly exponentially stable (RES) (or robustly globally exponentially stable (RGES)).
The last measurement and hence the corresponding noise is not considered in the above inequality, to keep the definition consistent with the formulations of FIE and MHE. Here the definition of RGAS strengthens the one introduced in [11] which applies instead of functions to the disturbances. This change is necessary to enable a desirable feature that a state estimator which is RGAS must be convergent under convergent disturbances [1, 19]. The next definition presents a weaker alternative of the above robust stability, which will also be needed in later analysis.
Definition 5.
(Practical robust stable estimations) The estimate defined in Definition 4 is practically RAS (pRAS) if given any , there exist functions such that the following inequality holds for all , , and :
(10) |
The estimate is further said to be practically RGAS (pRGAS) if the inequality is satisfied for all .
The adverb “practically” before RGAS is employed to keep it in line with the practical stability concept developed in control literature (e.g., [21]). Compared to an RGAS estimate, the pRGAS estimate admits a looser bound with a non-vanishing constant term . As will be shown later, this term can be made arbitrarily small if MHE implements a long enough horizon. In addition, it is worthwhile to remark that there is no need to define such stability variant for the case of RES or RGES because in either case the exponential property ensures that the term will be fully tempered and removed.
By applying the notation used in the concise definition of i-IOSS in (6), conciser forms of the above stability definitions can also be obtained, which will be useful to convey stability conditions for both FIE and MHE later on. Towards that, let , , , , , and for all . Consequently, the notation defined in (4) has a new realization, and its corresponding domain is denoted as if or if .
With the new notation, it is straightforward to prove the following equivalent definition of robust stabilities.
Lemma 1.
(Concise equivalent definition of robust stable estimations) The estimate defined in Definition 4 is RAS if and only if given any , there exists a function such that the following inequality holds for all and :
(11) |
And it is RGAS if and only if the above inequality holds for all . The estimate being RAS (or RGAS) is further said to be RES (or RGES), if and only if the function admits the form as with certain , and for all and in the corresponding domain.
Similar concise definitions of pRAS and pRGAS can both be obtained by adding a constant to the right hand side of (11), in which the function will then be dependent on . For brevity, we do not state them with another lemma.
IV Stability Implication from FIE to MHE
At any discrete time, an MHE instance can be interpreted as the corresponding FIE initiating from the start of the horizon over which the MHE instance is defined. Thus, the corresponding FIE being robustly stable implies that each MHE instance is robustly stable within the time horizon over which the instance is defined. If we interpret this as MHE being instance-wise robustly stable, then the challenge reduces to identifying conditions under which instance-wise robust stability implies robust stability of MHE. This observation was made in [2], and the challenge was solved there for a weaker definition of RGAS. This section resolves the challenge subject to the stronger stability concept given by Definition 4.
To that end, as in [2], we apply an ordinary assumption on the prior estimate of the initial state of an MHE instance.
Assumption 1.
Given any time , the prior estimate of is given such that
The assumption is obviously true if is set to , which is the MHE estimate obtained at time . Alternatively, a better might be obtained with smoothing techniques which use measurements both before and after time [22, 1].
Next, we present an important lemma which links global robust stability of MHE with that of its corresponding FIE.
Lemma 2.
(Global stability implication from FIE to MHE) Consider MHE under Assumption 1. The following two conclusions hold:
a) (RGAS–>pRGAS/RGAS/RGES) If FIE is RGAS as per (9), then there exists such that the corresponding MHE under Assumption 1 is pRGAS for all . If further the function in (9) is globally Lipschitz continuous at the origin, then the implication strengthens to the existence of such that the MHE is RGAS for all ; and if furthermore the functions and are also globally Lipschitz continuous at the origin, then the MHE is RGES for all .
Proof:
a) RGAS–>pRGAS. Let , and so . For all , the MHE and the corresponding FIE estimates are the same, both denoted as . So, given any , the absolute estimation error satisfies the RGAS inequality given by (9). That is, we have
(12) |
for all . Next, we proceed to show that the RGAS property is maintained for all .
Let and for all . The MHE instance at time can be viewed as the corresponding FIE confined to time interval . Thus, the MHE satisfies the RGAS property within this interval. That is, by (9) we have:
where Assumption 1 has been applied to produce the first term of the right hand side of the inequality. Repeat this reasoning for the MHE instance defined at time and then apply the above inequality, yielding
By induction, we obtain
Since satisfies inequality (12), this implies that
(13) | ||||
where the relation that for all and has been used to establish the first equality, and the three functions , and are defined by one-to-one correspondence to the three preceding terms, each of which is dependent on the horizon size . By definition, it is obvious that , and are functions for any given . However, it is not necessary that , and are functions for all , and so , and are not necessarily functions. The remaining proof is to show that there exists such that , and are bounded by functions plus a positive constant for all .
As increases from zero to infinity, the function instances form a sequence , for each . By the definition of , given any and , we have for each . That is, point-wisely converges to as . Next, we prove that the convergence is uniform w.r.t. the arguments and . Since is assumed to be bounded, there exists such that . Given any , there must exist satisfying , such that for all we will have for any ,
and for any ,
That is, holds true for all . This implies that uniformly converges to . So can and be shown in the same way to uniformly converge to and , respectively.444If any of the uncertainties does not have a certain bound, then the valid will depend on the actual uncertainty magnitude and consequently only semi-uniformly convergence will be proved.
Consequently, given any , there exists such that and for all , , , , , and . Given each , let
for all . Given any , it is easy to verify that is a function as is, and so for any we will have for all in aforementioned domains. By applying these inequalities to the last inequality of above, we conclude that given any , there exists such that
for all , where for all . This implies that the MHE is pRGAS by definition, and hence completes proof of the first conclusion of part a).
RGAS–>RGAS. If further from (9) is globally Lipschitz continuous at the origin, then for any there must exist a function such that for all because . By applying this property to (13), for all satisfying with certain , the inequality there proceeds as
(14) |
Consequently, the MHE is RGAS by definition, which completes the proof of the second conclusion of part a).
RGAS–>RGES. If in addition to , the functions and from (9) are globally Lipschitz continuous at the origin, then inequality (14) proceeds as
where is the Lipschitz constant of at the origin for all . Consequently, the MHE is RGES by definition, which completes the third conclusion of part a).
b) RGES–>RGES. In this case, we have with certain and for each . Since RGES implies RGAS, the deduction in the proof of part a) is applicable. In particular, the three functions , and are derived explicitly as:
(15) | ||||
for all , , , and . If , it is obvious that , and are all upper bounded by functions in exponential forms, which implies RGES of the MHE for all . And if , by defining () we have for all , where the penultimate equality holds only for . Consequently , and are again all upper bounded by functions in exponential forms which imply RGES of the MHE. Combining the two cases, we conclude that the MHE is RGES for all , which completes the proof of b). ∎
Lemma 2.a) indicates that RGAS of FIE implies pRGAS of the corresponding MHE which implements a long enough horizon, and that the implication strengthens to RGAS or RGES of the MHE if the FIE additionally satisfies certain global Lipschitz continuity conditions. In the latter case, it is not certain that the MHE will be RGES unless the three bound functions {, , } of the FIE all satisfy the global Lipschitz conditions. Also note that the main conclusion of Lemma 2.b) is implied by the third conclusion of Lemma 2.a) and that its proof above is presented mainly for derivation of an explicit bound of the sufficient horizon size.
Remark 1.
In the second case of Lemma 2, an FIE being RGAS and its bound function being globally Lipschitz at the origin imply that the system is exp-i-IOSS. This can be proved by extending the proof for the local case in [17, Proposition 2] to the global case. However, the two conditions do not necessarily imply that the FIE will be RGES (or at least the proof is unclear yet), though practical RGES is guaranteed by Lemma 2 if and are also globally Lipschitz at the origin. On the other hand, if an FIE is RGES, then the Lipschitz conditions in the third case are valid and hence the corresponding MHE will be RGES by Lemma 2, as is in line with the existing results reported in [18, 23].
Remark 2.
Reference [24] proved existence of a finite horizon for MHE to be RGAS if the corresponding FIE induces a global contraction for the estimation error which essentially requires the dynamical system to be exp-i-IOSS. As referred to Remark 2, this conveys the same necessary condition as in the second case of Lemma 2. Overall, Lemma 2 covers more general conclusions, which links RGAS of FIE to pRGAS of the corresponding MHE but also RGAS and RGES of MHE as two naturally induced cases subject to extra mild conditions.
Remark 3.
Reference [25] proved existence of a finite horizon for a model predictive control scheme to exponentially stabilize an unperturbed nonlinear system via state feedback. Though the result could be relevant to the strongest conclusion of Lemma 2 when the FIE is RGES, a strict and meaningful connection is yet to be explored for establishing robustly stable estimators here when both disturbances and i-IOSS stability are in place which are yet absent in the system setup of [25].
Next, we present additional results when the conditions of Lemma 2 are relaxed and confined to local regions of the estimation problem. In this case, as a byproduct, an explicit bound for the sufficient MHE horizon size is derived when the corresponding FIE is known to be RAS or RGAS.
Lemma 3.
(Local/global stability implication from FIE to MHE) Consider MHE under Assumption 1. Let the disturbances be bounded as and for all . Given any , the following four conclusions hold:
a) (RAS–>pRAS/RAS) Given any and , where , if FIE is RAS as per (9), then the corresponding MHE is pRAS for all and . If in addition is Lipschitz continuous at the origin over for all , then the corresponding MHE is RAS for all and .
b) (RGAS–>pRGAS/RGAS) If the conditions in a) are satisfied for all (which replaces ), then the two conclusions in a) hold globally.
c) (RES–>RES) Given defined in a) and any , if FIE is RES as per (9), in which for each the function has an exponential form as for certain , and all and , then the corresponding MHE is RES for all and .
d) (RGES–>RGES) If the conditions in c) are satisfied for all , then the conclusion in c) holds globally.
Proof:
a) RAS–>pRAS. Given any and , define for all . The proof follows the same procedure of the proof for Lemma 2.a), but further bounds the term after the function composition with by defined here in each induction step. That the value of lies in makes the same compose-and-then-bound approach applicable throughout the induction steps.
Let , which is well defined by Proposition 1.a). For any , Proposition 1.b) is applicable, and so the aforementioned induction leads to the following bound of the MHE estimate error:
(16) |
where the three functions , and are defined by one-to-one correspondence to the three preceding terms, each of which is dependent on the horizon size . With , it is obvious that , and are all functions. Consequently, the above estimate error bound implies that the MHE is pRAS by definition.
RAS–>RAS. When is Lipschitz continuous at the origin over the domain for all , by Proposition 1.c) the value is well defined also at . As a consequence, the same proof above will show that the implied pRAS collapses to RAS, as the ambiguity caused by disappears.
b) RGAS–>pRGAS/RGAS. The proof is straightforward by extending the reasoning in the proof of a) to the entire applicable domain with .
c) RES–>RES. In this case, we have with certain , and for all and . If , let and hence for all and . And if , let and hence for all and . In either case, the induction steps in the proof of Lemma 2.b) are applicable as the domain of the function composition via is kept within due to the contraction induced by the inequality . As a consequence, the state estimate error will again be bounded as per (15). By applying the same argument right after (15) subject to (which consolidates the two cases), we conclude that the MHE corresponding to the FIE is RES. This completes the proof of c).
d) RGES–>RGES. The proof is straightforward by extending the reasoning in the proof of c) to the entire problem domain with . ∎
Remark 4.
Two remarks follow on the proof of Lemma 3.a). a) The proof shows that the upper bound of the estimate error is scaled by the factor , where is a value that bonds with the horizon bound . Given any , the error bound can in fact be tightened by replacing the factor with , where is a certain function (and is short for ). As shown in Appendix A, as time is large enough, the derivative of this tighter factor will be negative and hence the error bound will be decreasing w.r.t. if the horizon size and the associated bound satisfy certain conditions that depend on the function . b) The proof indicates that it is valid for MHE to change its horizon size over time according to the amplitude of uncertainty in an intermediate initial state, as long as sufficient contraction of the uncertainty is maintained. This leaves space for adapting the horizon size of MHE such that a flexible and time-varying tradeoff can be made between estimation quality and computational complexity.
Lemma 3 indicates that there always exists a large enough horizon size such that RAS/RGAS of its corresponding FIE implies pRAS/pRGAS of MHE, and further RAS/RGAS of the MHE if the FIE admits a certain Lipschitz continuity property. When the corresponding FIE is RES/RGES, the implication directly strengthens to RES/RGES of the MHE.
Note that the implication for pRGAS of MHE in Lemma 3.b) is equivalent to the one concluded in Lemma 2.a), but here it is derived from a different proof. The advantage is to have an explicit way of computing a sufficient horizon size. The implication for RGES of MHE in Lemma 3.d) is again equivalent to that presented in Lemma 2.b), while here it is extended naturally from a local implication.
The bound for a sufficient horizon size as given in Lemma 3.a) can be derived explicitly if the function has particular forms. Given and , two expressions of are derived as follows:
The given constants for are such that the two ’s above are Lipschitz continuous at the origin over for all .
V Robust Stability of FIE and MHE
This section presents new sufficient conditions for robust stability of FIE, and also for that of the corresponding MHE by applying the stability implication revealed in last section.
To make the conditions easy to interpret and the proof concise to present, the following notations are introduced in the spirit of (4):
(17) | ||||
for all , where and refer to the optimal estimates of and , respectively. Given any , notation collects the sequence of vector variables and the corresponding domain is denoted by if and as if , where is defined in (8). Function extracts the time index (i.e., the original index ) of as per (5). Given any , it is easy to verify that and that .
Given any , the next two assumptions are introduced to establish robust stability of FIE.
Assumption 2.
There exists such that the cost function of FIE , which is continuous, satisfies the following inequality for all and :
(18) | |||
Assumption 3.
Overall, Assumption 2 requires that the FIE has a property that mimics the i-IOSS property of the system, while Assumption 3 ensures that the FIE is more sensitive than the system to the uncertainties so that accurate inference of the state is possible. The interpretation of the relatively more obscure Assumption 3 becomes clear with a concrete realization below.
Lemma 4.
Proof:
Theorem 1.
The following two conclusions hold:
a) (RAS/RGAS/RES/RGES of FIE) The FIE is RAS if the system is i-IOSS and the cost function of FIE satisfies Assumptions 2 and 3, and is RGAS if further the two assumptions are valid for all . And if, furthermore, the functions in Assumption 3 admit specific forms as and with certain , and for all and in the applicable domain, then the FIE which is RAS or RGAS will be RES or RGES, respectively.
b) (pRAS/pRGAS/RES/RGES/RGAS of MHE) In the four cases of a), the corresponding MHE under additional Assumption 1 with a sufficiently long horizon is pRAS, pRGAS, RES and RGES, respectively. And the pRGAS of MHE strengthens to RGAS if the functions and in Assumption 3 are globally Lipschitz continuous at the origin.
Proof:
a) RAS/RGAS of FIE. The global optimal solution of for the FIE is denoted as (cf. Eq. (17)), yielding a minimum cost . It follows that for all ,
Consequently, by Assumption 2 we have , and further .
Since for each , by applying the triangle inequality this implies that
(21) |
for all . Substitute (21) into the i-IOSS property of (6), yielding
where the equality and Assumption 3 have been used to derive the last inequality, and for all and . Since is a function, the FIE is RAS by definition. With the same reasoning, the conclusion immediately extends to that the FIE is RGAS when Assumptions 2 and 3 are applicable to the entire problem domain as specified by .
RES/RGES of FIE. If in addition, and with certain and and for all and in an applicable domain, then it is immediate that , with and . Here is a function in an exponential form, so the FIE which is RAS (or RGAS) in this case will be RES (or RGES) by Lemma 1.
Next, we present a lemma indicating that Assumptions 2 and 3 do not impose special difficulty as there always exists a cost function satisfying both of them if the system is i-IOSS.
Lemma 5.
(Satisfaction of Assumptions 2 and 3) If the system is i-IOSS as per (6), then, given any and satisfying for all and , it is feasible to specify the cost function of FIE as
(22) |
such that Assumptions 2 and 3 hold true. The same form of cost function remains valid for Assumptions 2 and 3 to hold globally with . In either the local or the global case, if the system is exp-i-IOSS, then the function admits a form as with certain and and for and in the applicable domain.
Proof:
Given the cost function specified as per (22), Assumption 2 is automatically met. With for all and , Assumption 3 is also met by Lemma 4. While, existence of such function is guaranteed as it is always feasible to let . Since the reasoning approach does not rely on the domain of , the conclusion remains valid if Assumptions 2 and 3 are extended to the entire problem domain with .
If the system is exp-i-IOSS with for certain and , then it is valid to let with for all and in the applicable local/global domain. ∎
Lemma 5 indicates that a valid cost function can always be designed from the i-IOSS bound function , and hence implies an important result below.
Corollary 1.
There exists a cost function for FIE to be RGAS (or RGES) if and only if the system is i-IOSS (or exp-i-IOSS).
Proof:
Sufficiency. By Lemma 5, the system being i-IOSS implies that the FIE admits a cost function such that Assumptions 2 and 3 hold true, which consequently implies that the FIE is RGAS by Theorem 1. When the system is exp-i-IOSS, the conclusion trivially strengthens to that the FIE is RGES by following the same approach of reasoning.
By Lemma 2, it follows immediately from Corollary 1 that there exists a cost function for the MHE to be pRGAS (or RGES) if the system is i-IOSS (or exp-i-IOSS).
Remark 5.
The conclusion of Corollary 1 coincides with a key finding reported in a latest paper [24], which had been submitted for review. The derivation approaches are, however, quite different. Here, we apply the reasoning approach of [2], focusing on developing most general conditions for robust stability of FIE, and the aforementioned conclusion appears as a corollary for an endeavour to understand the developed conditions. In contrast, reference [24] reaches the conclusion by starting with a particular cost function which is constructed directly from the i-IOSS property of the system and is not necessarily the only form admitted by our derived conditions.
VI Conclusion
This work proved that robust global (or local) asymptotic stability of full information estimation (FIE) implies practical robust global (or local) asymptotic stability of moving horizon estimation (MHE) which implements a sufficiently long horizon. The “practical” becomes exact if the FIE admits a certain Lipschitz continuity. In both exact and inexact cases, explicit ways were also provided of computing a sufficient horizon size for a robustly stable MHE. With the revealed implication, sufficient conditions for the MHE to be robustly stable were derived by firstly developing those for ensuring robust stability of the corresponding FIE. A particular realization of these conditions indicates that the system being i-IOSS is not only necessary but also sufficient to ensure the existence of a robustly globally asymptotically stable of FIE. With the revealed implication, the sufficiency remains valid for existence of a practically robustly stable MHE.
Since it is generic and relies only on robust stability of the corresponding FIE, the revealed stability link implies that existing conditions which ensure robust stability of FIE can all be inherited to establish that of the corresponding MHE, but also paves the way for derivation of new sufficient conditions via deeper analysis of FIE. This may also contribute to developing robustly stable MHE which exploits sub-optimal solutions for resource-constrained or faster estimations. Readers are referred to Sec. VII of [2] for related discussions, and [26, 27, 28, 29] for some recent developments in this line of research.
Appendix A. Derivative of the Error Bound Factor w.r.t. the Moving Horizon Size
The symbols mostly come from Lemma 3.a) and its proof. Given , it is easy to show that the estimate error bound in (16) can be tightened by replacing with . This appendix analyzes the derivative of this new bound factor w.r.t. , in order to understand monotonicity of the bound w.r.t. the horizon size .
Firstly, we show that there exists such that for all . By Lemma 8 in [30], given , there exists and such that , for all . This implies that the RGAS property of FIE can equivalently be expressed with the product of and functions. Therefore, it does not lose generality by assuming . Consequently, . Since is a function, given , it is feasible to define and will be a function for all .
Next, we derive the derivative of w.r.t. . By definition we have for all . Let . Hence . The chain rule of derivatives implies
The first factor of the last expression is positive, so the sign of the derivative is uniquely determined by the second factor. When is much larger than , we have . Then, the second factor approximates to
As is positive, it is sufficient to analyze the sign of
As examples, two particular forms of are considered. Firstly, let us consider the exponential form: with . Then, . So, and hence will be strictly decreasing w.r.t. if . Secondly, let us consider the fractional form: with . Then, So, and hence will again be strictly decreasing w.r.t. if . Since the condition is more restrictive in this case, the monotonic decrement property seems harder to establish than that in the previous case.
To conclude, the analysis shows that if the moving horizon size satisfies certain conditions (being large enough in general), the bound of the estimate error as controlled by the factor will eventually decrease with .
References
- [1] J. B. Rawlings, D. Q. Mayne, and M. Diehl, Model Predictive Control: Theory, Computation, and Design, 2nd ed. Nob Hill Publishing Madison, WI, 2020.
- [2] W. Hu, “Robust stability of optimization-based state estimation,” https://arxiv.org/abs/1702.01903, 2017.
- [3] H. Michalska and D. Mayne, “Moving horizon observers,” in Nonlinear Control Systems Design 1992. Elsevier, 1993, pp. 185–190.
- [4] K. R. Muske, J. B. Rawlings, and J. H. Lee, “Receding horizon recursive state estimation,” in American Control Conference (ACC), San Francisco, CA, USA, Jun 1993, pp. 900–904.
- [5] H. Michalska and D. Q. Mayne, “Moving horizon observers and observer-based control,” IEEE Trans. on Autom. Control, vol. 40, no. 6, pp. 995–1006, 1995.
- [6] C. Rao, J. Rawlings, and D. Mayne, “Constrained state estimation for nonlinear discrete-time systems: Stability and moving horizon approximations,” IEEE Trans. on Autom. Control, vol. 48, no. 2, pp. 246–258, 2003.
- [7] A. Alessandri, M. Baglietto, and G. Battistelli, “Moving-horizon state estimation for nonlinear discrete-time systems: New stability results and approximation schemes,” Automatica, vol. 44, no. 7, pp. 1753–65, 2008.
- [8] J. B. Rawlings and D. Q. Mayne, Model Predictive Control: Theory and Design. Madison, WI: Nob Hill Pub., 2009.
- [9] A. Alessandri, M. Baglietto, G. Battistelli, and V. Zavala, “Advances in moving horizon estimation for nonlinear systems,” in The 49th IEEE Conference on Decision and Control (CDC), Atlanta, Georgia, USA, Dec. 2010, pp. 5681–5688.
- [10] J. Liu, “Moving horizon state estimation for nonlinear systems with bounded uncertainties,” Chem. Eng. Sci., vol. 93, pp. 376–386, 2013.
- [11] J. Rawlings and L. Ji, “Optimization-based state estimation: Current status and some new results,” Journal of Process Control, vol. 22, pp. 1439–1444, 2012.
- [12] E. Sontag and Y. Wang, “Output-to-state stability and detectability of nonlinear systems,” Systems & Control Letters, vol. 29, no. 5, pp. 279–290, 1997.
- [13] W. Hu, L. Xie, and K. You, “Optimization-based state estimation under bounded disturbances,” in The 54th IEEE Conference on Decision and Control (CDC), Osaka, Japan, Dec 2015.
- [14] L. Ji, J. B. Rawlings, W. Hu, A. Wynn, and M. Diehl, “Robust stability of moving horizon estimation under bounded disturbances,” IEEE Trans. on Autom. Control, vol. 61, no. 11, pp. 3509 – 3514, 2016.
- [15] M. A. Müller, “Nonlinear moving horizon estimation for systems with bounded disturbances,” in American Control Conference (ACC), Boston, USA, Jul 2016, pp. 883–888.
- [16] ——, “Nonlinear moving horizon estimation in the presence of bounded disturbances,” Automatica, vol. 79, pp. 306–314, 2017.
- [17] D. A. Allan and J. B. Rawlings, Handbook of Model Predictive Conttrol. Springer Nature Switzerland AG, 2019, ch. Moving Horizon Estimation, pp. 99–124.
- [18] S. Knüfer and M. A. Müller, “Robust global exponential stability for moving horizon estimation,” in The 57th IEEE Conference on Decision and Control, Miami Beach, FL, USA, Dec. 2018, pp. 3477–3482.
- [19] D. A. Allan, J. Rawlings, and A. R. Teel, “Nonlinear detectability and incremental input/output-to-state stability,” SIAM Journal on Control and Optimization, vol. 59, no. 4, pp. 3017–3039, 2021.
- [20] S. Knüfer and M. A. Müller, “Time-discounted incremental input/output-to-state stability,” in The 59th IEEE Conference on Decision and Control (CDC), Jeju Island, Korea, Dec. 2020, pp. 5394–5400.
- [21] L. Grüne and J. Pannek, “Nonlinear model predictive control: Theory and algorithms,” in Nonlinear Model Predictive Control, 2nd ed. Springer, 2017, pp. 45–69.
- [22] A. Aravkin, J. V. Burke, L. Ljung, A. Lozano, and G. Pillonetto, “Generalized kalman smoothing: Modeling and algorithms,” arXiv preprint arXiv:1609.06369, 2016.
- [23] D. A. Allan, “A lyapunov-like function for analysis of model predictive control and moving horizon estimation,” Ph.D. dissertation, 2020.
- [24] S. Knuefer and M. A. Mueller, “Nonlinear full information and moving horizon estimation: Robust global asymptotic stability,” arXiv preprint arXiv:2105.02764, 2021.
- [25] A. Jadbabaie and J. Hauser, “On the stability of receding horizon control with a general terminal cost,” IEEE Trans. on Autom. Control, vol. 50, no. 5, pp. 674–678, 2005.
- [26] A. Wynn, M. Vukov, and M. Diehl, “Convergence guarantees for moving horizon estimation based on the real-time iteration scheme,” IEEE Trans. on Autom. Control, vol. 59, no. 8, pp. 2215–2221, 2014.
- [27] A. Alessandri and M. Gaggero, “Fast moving horizon state estimation for discrete-time systems using single and multi iteration descent methods,” IEEE Trans. on Autom. Control, vol. 62, no. 9, pp. 4499–4511, 2017.
- [28] M. Gharbi, B. Gharesifard, and C. Ebenbauer, “Anytime proximity moving horizon estimation: Stability and regret,” arXiv preprint arXiv:2006.14303, 2020.
- [29] J. D. Schiller, S. Knüfer, and M. A. Müller, “Robust stability of suboptimal moving horizon estimation using an observer-based candidate solution,” arXiv preprint arXiv:2011.08723, 2020.
- [30] E. D. Sontag, “Comments on integral variants of ISS,” Systems & Control Letters, vol. 34, no. 1, pp. 93–100, 1998.