• Tidak ada hasil yang ditemukan

Contraction Theory for Discrete-time Systems

Chapter 2: Contraction Theory

2.4 Contraction Theory for Discrete-time Systems

The results presented so far can be readily extended to those for discrete-time nonlinear systems.

2.4.I Deterministic Perturbation

Let us consider the following nonlinear system with bounded deterministic pertur- bation๐‘‘๐‘˜ :R๐‘›ร—Nโ†ฆโ†’R๐‘›with ยฏ๐‘‘ โˆˆRโ‰ฅ0s.t. ยฏ๐‘‘ =sup๐‘ฅ , ๐‘˜ โˆฅ๐‘‘๐‘˜(๐‘ฅ , ๐‘˜) โˆฅ:

๐‘ฅ(๐‘˜ +1)= ๐‘“๐‘˜(๐‘ฅ(๐‘˜), ๐‘˜) +๐‘‘๐‘˜(๐‘ฅ(๐‘˜), ๐‘˜) (2.50) where ๐‘˜ โˆˆ N, ๐‘ฅ : N โ†ฆโ†’ R๐‘› is the discrete system state, and ๐‘“๐‘˜ : R๐‘› ร—N โ†ฆโ†’ R๐‘› is a smooth function. Although this thesis focuses mainly on continuous-time nonlinear systems, let us briefly discuss contraction theory for (2.50) to imply that the techniques in the subsequent chapters are applicable also to discrete-time nonlinear systems.

Let ๐œ‰0(๐‘˜) and ๐œ‰1(๐‘˜) be solution trajectories of (2.50) with ๐‘‘๐‘˜ = 0 and ๐‘‘๐‘˜ โ‰  0, respectively. Then a virtual system of ๐‘ž(๐œ‡, ๐‘˜) parameterized by๐œ‡ โˆˆ [0,1], which has๐‘ž(๐œ‡=0, ๐‘˜) =๐œ‰0(๐‘˜)and๐‘ž(๐œ‡=1, ๐‘˜) =๐œ‰1(๐‘˜)as its particular solutions, can be expressed as follows:

๐‘ž(๐œ‡, ๐‘˜ +1) = ๐‘“๐‘˜(๐‘ž(๐œ‡, ๐‘˜), ๐‘˜) +๐œ‡ ๐‘‘๐‘˜(๐œ‰1(๐‘˜), ๐‘˜). (2.51) The discrete version of robust contraction in Theorem 2.4 is given in the following theorem.

Theorem 2.8. Let๐‘ฅ๐‘˜ =๐‘ฅ(๐‘˜) and๐‘ž๐‘˜ =๐‘ž(๐œ‡, ๐‘˜)for any๐‘˜ โˆˆN. If there exists a uni- formly positive definite matrix๐‘€๐‘˜(๐‘ฅ๐‘˜, ๐‘˜) = ฮ˜๐‘˜(๐‘ฅ๐‘˜, ๐‘˜)โŠคฮ˜๐‘˜(๐‘ฅ๐‘˜, ๐‘˜) โ‰ป0, โˆ€๐‘ฅ๐‘˜, ๐‘˜, where ฮ˜๐‘˜ defines a smooth coordinate transformation of๐›ฟ๐‘ฅ๐‘˜, i.e.,๐›ฟ ๐‘ง๐‘˜ = ฮ˜๐‘˜(๐‘ฅ๐‘˜, ๐‘˜)๐›ฟ๐‘ฅ๐‘˜, s.t.

either of the following equivalent conditions holds forโˆƒ๐›ผโˆˆ (0,1),โˆ€๐‘ฅ๐‘˜, ๐‘˜:

ฮ˜๐‘˜+1(๐‘ฅ๐‘˜+1, ๐‘˜ +1)๐œ• ๐‘“๐‘˜

๐œ• ๐‘ฅ๐‘˜

ฮ˜๐‘˜(๐‘ฅ๐‘˜, ๐‘˜)โˆ’1

โชฏ ๐›ผ (2.52)

๐œ• ๐‘“๐‘˜

๐œ• ๐‘ฅ๐‘˜

โŠค

๐‘€๐‘˜+1(๐‘ฅ๐‘˜+1, ๐‘˜ +1)๐œ• ๐‘“๐‘˜

๐œ• ๐‘ฅ๐‘˜

โชฏ ๐›ผ2๐‘€๐‘˜(๐‘ฅ๐‘˜, ๐‘˜), (2.53)

then we have the following bound as long as we have๐‘šI โชฏ ๐‘€๐‘ฅ(๐‘ฅ๐‘˜, ๐‘˜) โชฏ ๐‘šI, โˆ€๐‘ฅ๐‘˜, ๐‘˜, as in(2.26):

โˆฅ๐œ‰1(๐‘˜) โˆ’๐œ‰0(๐‘˜) โˆฅ โ‰ค ๐‘‰โ„“(0)

โˆš ๐‘š

๐›ผ๐‘˜ + ๐‘‘ยฏ(1โˆ’๐›ผ๐‘˜) 1โˆ’๐›ผ

โˆš๏ธ„

๐‘š ๐‘š

(2.54) where๐‘‰โ„“(๐‘˜) = โˆซ๐œ‰1

๐œ‰0 โˆฅฮ˜๐‘˜(๐‘ž๐‘˜, ๐‘˜)๐›ฟ๐‘ž๐‘˜โˆฅ as in(2.22) for the unperturbed trajectory๐œ‰0, perturbed trajectory๐œ‰1, and virtual state๐‘ž๐‘˜ =๐‘ž(๐‘˜)given in(2.51).

Proof. If (2.52) or (2.53) holds, we have that ๐‘‰โ„“(๐‘˜+1) โ‰ค

โˆซ 1 0

โˆฅฮ˜๐‘˜+1(๐œ•๐‘ž

๐‘˜๐‘“๐‘˜(๐‘ž๐‘˜, ๐‘˜)๐œ•๐œ‡๐‘ž๐‘˜ +๐‘‘๐‘˜(๐‘ฅ๐‘˜, ๐‘˜)) โˆฅ๐‘‘๐œ‡

โ‰ค ๐›ผ

โˆซ 1 0

โˆฅฮ˜๐‘˜(๐‘ž๐‘˜, ๐‘˜)๐œ•๐œ‡๐‘ž๐‘˜โˆฅ๐‘‘๐œ‡+๐‘‘ยฏ

โˆš

๐‘š =๐›ผ๐‘‰โ„“(๐‘˜) +๐‘‘ยฏ

โˆš ๐‘š

where ฮ˜๐‘˜+1 = ฮ˜๐‘˜+1(๐‘ž๐‘˜+1, ๐‘˜ +1), ๐œ•๐‘ž

๐‘˜๐‘“๐‘˜(๐‘ž๐‘˜, ๐‘˜) = ๐œ• ๐‘“๐‘˜/๐œ• ๐‘ž๐‘˜, and ๐œ•๐œ‡๐‘ž๐‘˜ = ๐œ• ๐‘ž๐‘˜/๐œ• ๐œ‡. Applying this inequality iteratively results in (2.54).

Theorem 2.8 can be used with Theorem 2.4 for stability analysis of hybrid nonlin- ear systems [33]โ€“[35], or with Theorem 2.5 for stability analysis of discrete-time stochastic nonlinear systems [6], [8], [35]. For example, it is shown in [6] that if the time interval in discretizing (2.1) as (2.50) is sufficiently small, contraction of discrete-time systems with stochastic perturbation reduces to that of continuous-time systems as follows.

2.4.II Stochastic Perturbation

Let us also present a discrete-time version of Theorem 2.5, which can be extensively used for proving the stability of discrete-time and hybrid stochastic nonlinear sys- tems, along with known results for deterministic systems [33], [34]. Consider the discrete-time nonlinear system with stochastic perturbation modeled by the stochas- tic difference equation

๐‘ฅ(๐‘˜ +1)= ๐‘“๐‘˜(๐‘ฅ(๐‘˜), ๐‘˜) +๐บ๐‘˜(๐‘ฅ(๐‘˜), ๐‘˜)๐‘ค(๐‘˜) (2.55) where๐บ๐‘˜ :R๐‘›ร—Nโ†’R๐‘›ร—๐‘‘is a matrix-valued function and๐‘ค(๐‘˜)is a๐‘‘-dimensional sequence of zero mean uncorrelated normalized Gaussian random variables. Con- sider the following two systems with trajectories ๐œ‰0(๐‘˜) and ๐œ‰1(๐‘˜) driven by two independent stochastic perturbation๐‘ค0(๐‘˜) and๐‘ค1(๐‘˜):

๐œ‰๐‘–(๐‘˜ +1) = ๐‘“๐‘˜(๐œ‰๐‘–(๐‘˜), ๐‘˜) +๐บ๐‘–, ๐‘˜(๐œ‰๐‘–(๐‘˜), ๐‘˜)๐‘ค๐‘–(๐‘˜), ๐‘– =0,1, (2.56)

Similar to (2.36), a virtual system of ๐‘ž(๐œ‡, ๐‘˜) parameterized by ๐œ‡ โˆˆ [0,1], which has๐‘ž(๐œ‡=0, ๐‘˜) =๐œ‰0(๐‘˜)and๐‘ž(๐œ‡=1, ๐‘˜) =๐œ‰1(๐‘˜)as its particular solutions, can be given as follows:

๐‘ž(๐œ‡, ๐‘˜ +1) = ๐‘“๐‘˜(๐‘ž(๐œ‡, ๐‘˜), ๐‘˜) +๐บ๐‘˜(๐œ‡, ๐œ‰0(๐‘˜), ๐œ‰1(๐‘˜), ๐‘˜)๐‘ค(๐‘˜) (2.57) where๐บ๐‘˜(๐œ‡, ๐œ‰0(๐‘˜), ๐œ‰1(๐‘˜), ๐‘˜) =[(1โˆ’๐œ‡)๐บ0, ๐‘˜(๐œ‰0(๐‘˜), ๐‘˜), ๐œ‡๐บ1, ๐‘˜(๐œ‰1(๐‘˜), ๐‘˜)]and๐‘ค(๐‘˜) = [๐‘ค0(๐‘˜)โŠค, ๐‘ค1(๐‘˜)โŠค]โŠค. The following theorem analyzes stochastic incremental stabil- ity for discrete-time nonlinear systems (2.56), which is different from [26], [35] in that the stability is studied in a differential sense and its Riemannian metric is state- and time-dependent.

Theorem 2.9. Suppose that(2.53)holds for the discrete-time deterministic system (2.56) with ๐›ผ2 = 1โˆ’ ๐›พ๐‘‘ and thatโˆƒ๐‘š, ๐‘š โˆˆ R>0 and ๐‘”ยฏ0๐‘‘,๐‘”ยฏ1๐‘‘ โˆˆ Rโ‰ฅ0 s.t. ๐‘š ๐ผ โชฏ ๐‘€๐‘˜(๐‘ฅ , ๐‘˜) โชฏ ๐‘š ๐ผ , โˆ€๐‘ฅ , ๐‘˜, sup๐‘ฅ , ๐‘˜ โˆฅ๐บ1, ๐‘˜(๐‘ฅ , ๐‘˜) โˆฅ๐น = ๐‘”ยฏ0๐‘‘, and sup๐‘ฅ , ๐‘˜ โˆฅ๐บ2, ๐‘˜(๐‘ฅ , ๐‘˜) โˆฅ๐น =

ยฏ

๐‘”1๐‘‘. Suppose also that โˆƒ๐›พ2 โˆˆ (0,1) s.t. ๐›พ2 โ‰ค 1โˆ’ (๐‘š/๐‘š) (1โˆ’๐›พ๐‘‘), where ๐›พ๐‘‘ is the contraction rate. Consider the generalized squared length with respect to a Riemannian metric๐‘€๐‘˜(๐‘ž(๐œ‡, ๐‘˜), ๐‘˜)defined as

๐‘‰๐‘ โ„“(๐‘ž, ๐›ฟ๐‘ž, ๐‘˜) =

โˆซ ๐œ‰1 ๐œ‰0

๐›ฟ๐‘žโŠค๐‘€๐‘˜(๐‘ž(๐œ‡, ๐‘˜), ๐‘˜)๐›ฟ๐‘ž =

โˆซ 1 0

๐œ• ๐‘ž

๐œ• ๐œ‡

โŠค

๐‘€๐‘˜(๐‘ž(๐œ‡, ๐‘˜), ๐‘˜)๐œ• ๐‘ž

๐œ• ๐œ‡

๐‘‘๐œ‡(2.58) s.t. ๐‘‰๐‘˜(๐‘ž, ๐›ฟ๐‘ž, ๐‘˜) โ‰ฅ ๐‘šโˆฅ๐œ‰1(๐‘˜) โˆ’๐œ‰0(๐‘˜) โˆฅ2. Then the mean squared distance between the two trajectories of the system(2.56)is bounded as follows:

E

โˆฅ๐œ‰1(๐‘˜) โˆ’๐œ‰0(๐‘˜) โˆฅ2

โ‰ค 1โˆ’๐›พหœ๐‘˜

๐‘‘

1โˆ’๐›พหœ๐‘‘

๐ถ๐‘‘+ ๐›พหœ๐‘˜

๐‘‘

๐‘š

๐ธ[๐‘‰๐‘ โ„“(0)]. (2.59)

where๐‘‰๐‘ โ„“(0) =๐‘‰๐‘ โ„“(๐‘ž(0), ๐›ฟ๐‘ž(0),0), ๐ถ๐‘‘ = (๐‘š/๐‘š) (๐‘”ยฏ2

0๐‘‘ +๐‘”ยฏ2

1๐‘‘), and ๐›พหœ๐‘‘ = 1โˆ’๐›พ2 โˆˆ (0,1).

Proof. Let ๐‘ž๐‘˜ = ๐‘ž(๐œ‡, ๐‘˜), ๐‘ค๐‘˜ = ๐‘ค(๐‘˜), ๐‘‰๐‘˜ = ๐‘‰๐‘ โ„“(๐‘ž(๐œ‡, ๐‘˜), ๐›ฟ๐‘ž(๐œ‡, ๐‘˜), ๐‘˜), and ๐‘€๐‘˜ = ๐‘€๐‘˜(๐‘ž(๐œ‡, ๐‘˜), ๐‘˜) for any ๐‘˜ โˆˆNfor notational simplicity. Using the assumed bounds along with (2.53) (๐›ผ2=1โˆ’๐›พ๐‘‘) and (2.57), we have, forโ„“ โˆˆN, that

๐‘‰โ„“+1 โ‰ค ๐‘š

โˆซ 1 0

๐œ• ๐‘“โ„“

๐œ• ๐‘žโ„“

๐œ• ๐‘žโ„“

๐œ• ๐œ‡

+ ๐œ• ๐บโ„“

๐œ• ๐œ‡ ๐‘คโ„“

2

๐‘‘๐œ‡ (2.60)

โ‰ค ๐‘š ๐‘š

(1โˆ’๐›พ๐‘‘)

โˆซ 1 0

๐œ• ๐‘žโ„“

๐œ• ๐œ‡

โŠค

๐‘€โ„“

๐œ• ๐‘žโ„“

๐œ• ๐œ‡ ๐‘‘๐œ‡

+๐‘š

โˆซ 1 0

2๐œ• ๐‘žโ„“

๐œ• ๐œ‡

โŠค๐œ• ๐‘“โ„“

๐œ• ๐‘žโ„“

โŠค๐œ• ๐บโ„“

๐œ• ๐œ‡

๐‘คโ„“ +๐‘คโŠค

โ„“

๐œ• ๐บโ„“

๐œ• ๐œ‡

โŠค๐œ• ๐บโ„“

๐œ• ๐œ‡ ๐‘คโ„“

๐‘‘๐œ‡.

Taking the conditional expected value of (2.60) when๐‘žโ„“, ๐›ฟ๐‘žโ„“, andโ„“ are given, we have that (see also: Theorem 2 of [26])

E๐œโ„“[๐‘‰โ„“+1] โ‰ค๐›พ๐‘š๐‘‰โ„“ +๐‘šE๐œโ„“

โˆซ 1 0

๐‘คโŠค

โ„“

๐œ• ๐บโ„“

๐œ• ๐œ‡

โŠค๐œ• ๐บโ„“

๐œ• ๐œ‡ ๐‘คโ„“๐‘‘๐œ‡

โ‰ค ๐›พ๐‘š๐‘‰โ„“ + โˆ‘๏ธ

๐‘–=1,2

๐‘šE๐œโ„“ h

Tr

๐‘ค๐‘–,โ„“๐‘คโŠค

๐‘–,โ„“๐บโŠค

๐‘–,โ„“๐บ๐‘–,โ„“ i

โ‰ค ๐›พ๐‘š๐‘‰โ„“ +๐‘š

โˆ‘๏ธ

๐‘–=1,2

Tr ๐บโŠค

๐‘–,โ„“๐บ๐‘–,โ„“

โ‰ค ๐›พหœ๐‘‘๐‘‰โ„“ +๐‘š๐ถ๐‘‘, (2.61) where๐›พ๐‘š =๐‘š/๐‘š(1โˆ’๐›พ๐‘‘), and๐‘žโ„“,๐›ฟ๐‘žโ„“, andโ„“ are denoted as๐œโ„“. Here, we used the condition: โˆƒ๐›พ2 โˆˆ (0,1) s.t. ๐›พ๐‘š โ‰ค 1โˆ’๐›พ2 = ๐›พหœ๐‘‘. Taking expectation over ๐œโ„“โˆ’1 in (2.61) with the tower ruleE๐œโ„“โˆ’1[๐‘‰โ„“+1] =E๐œโ„“โˆ’1[E๐œโ„“[๐‘‰โ„“+1]]gives us that

E๐œโ„“โˆ’1[๐‘‰โ„“+1] โ‰ค๐›พหœ2

๐‘‘๐‘‰โ„“โˆ’1+๐‘š๐ถ๐‘‘+๐‘š๐ถ๐‘‘๐›พหœ๐‘‘

where หœ๐›พ๐‘‘ is defined as หœ๐›พ๐‘‘ = 1โˆ’ ๐›พ2. Continuing this operation with the relation ๐‘šE๐œ0

โˆฅ๐œ‰1,โ„“+1โˆ’๐œ‰2,โ„“+1โˆฅ2

โ‰ค E๐œ0 [๐‘‰โ„“+1]yields E๐œ0

โˆฅ๐œ‰1, ๐‘˜ โˆ’๐œ‰2, ๐‘˜โˆฅ2

โˆ’ ๐›พหœ๐‘˜

๐‘‘

๐‘š

๐‘‰0 โ‰ค ๐ถ๐‘‘

๐‘˜โˆ’1

โˆ‘๏ธ

๐‘–=0

หœ ๐›พ๐‘–

๐‘‘ = 1โˆ’๐›พหœ๐‘˜

๐‘‘

1โˆ’๐›พหœ๐‘‘ ๐ถ๐‘‘

where ๐‘˜ = โ„“ + 1. Taking expectation over ๐œ0 and rearranging terms result in (2.59).

2.4.III Connection between Continuous and Discrete Stochastic Contraction Theory

Let us now consider the case where the time intervalฮ”๐‘ก =๐‘ก๐‘˜+1โˆ’๐‘ก๐‘˜for discretization is sufficiently small, i.e.,ฮ”๐‘ก โ‰ซ (ฮ”๐‘ก)2. Then the continuous-time stochastic system (2.29) can be discretized as

๐‘ฅ(๐‘˜ +1)=๐‘ฅ(๐‘˜) +

โˆซ ๐‘ก๐‘˜+1

๐‘ก๐‘˜

๐‘“(๐‘ฅ(๐‘ก), ๐‘ก)๐‘‘ ๐‘ก+๐บ(๐‘ฅ(๐‘ก), ๐‘ก)๐‘‘๐’ฒ(๐‘ก)

=๐‘ฅ(๐‘˜) + ๐‘“(๐‘ฅ(๐‘˜), ๐‘ก๐‘˜)ฮ”๐‘ก+๐บ(๐‘ฅ(๐‘˜), ๐‘ก๐‘˜)ฮ”๐’ฒ(๐‘˜) + O ฮ”๐‘ก2

where๐‘ฅ(๐‘˜)=๐‘ฅ(๐‘ก๐‘˜),ฮ”๐’ฒ(๐‘˜) =โˆš

ฮ”๐‘ก ๐‘ค(๐‘˜), and๐‘ค(๐‘˜)is a๐‘‘-dimensional sequence of zero mean uncorrelated normalized Gaussian random variables. Whenฮ”๐‘ก โ‰ซ (ฮ”๐‘ก)2, ๐‘“๐‘˜(๐‘ฅ(๐‘˜), ๐‘˜)and๐บ๐‘˜(๐‘ฅ(๐‘˜), ๐‘˜)in (2.55) can be approximated as ๐‘“๐‘˜(๐‘ฅ(๐‘˜), ๐‘˜) โ‰ƒ ๐‘ฅ(๐‘˜) + ๐‘“(๐‘ฅ(๐‘˜), ๐‘ก๐‘˜)ฮ”๐‘ก and ๐บ๐‘˜(๐‘ฅ(๐‘˜), ๐‘˜) โ‰ƒ

โˆš

ฮ”๐‘ก ๐บ(๐‘ฅ(๐‘˜), ๐‘ก๐‘˜). In this situation, we have the following theorem that connects the stochastic incremental stability of discrete-time systems with that of continuous-time systems.

Theorem 2.10. Suppose that(2.61)in Theorem 2.9 holds with๐›พหœ๐‘‘ =1โˆ’๐›พ2 โˆˆ (0,1). Then the expected value of ๐‘‰๐‘˜+1 up to first order in ฮ”๐‘ก is given as E๐œ๐‘˜[๐‘‰๐‘˜+1] = ๐‘‰๐‘˜ +ฮ”๐‘กโ„’๐‘‰๐‘˜, where๐‘‰๐‘˜ = ๐‘‰๐‘ โ„“(๐‘ž(๐œ‡, ๐‘˜), ๐›ฟ๐‘ž(๐œ‡, ๐‘˜), ๐‘˜) for ๐‘‰๐‘ โ„“ of (2.58) and โ„’ is the infinitesimal differential generator. Furthermore, the following inequality holds:

โ„’๐‘‰๐‘ โ„“(๐‘ž๐‘˜, ๐›ฟ๐‘ž๐‘˜, ๐‘ก๐‘˜) โ‰ค โˆ’๐›พ2 ฮ”๐‘ก

๐‘‰๐‘ โ„“(๐‘ž,๐›ฟ๐‘ž๐‘˜, ๐‘ก๐‘˜) +๐‘š๐ถหœ๐‘ (2.62) where๐‘ž๐‘˜ =๐‘ž(๐œ‡, ๐‘˜)๐ถหœ๐‘is a positive constant given as

หœ ๐ถ๐‘ = ๐ถ๐‘‘

ฮ”๐‘ก

= ๐‘š ๐‘šฮ”๐‘ก

(๐‘”ยฏ2

0๐‘‘+๐‘”ยฏ2

1๐‘‘) = ๐‘š ๐‘š

(๐‘”ยฏ2

0+๐‘”ยฏ2

1) (2.63)

with๐‘”ยฏ0and๐‘”ยฏ1defined in Theorem 2.5.

Proof. Let๐‘€๐‘˜ =๐‘€๐‘˜(๐‘ž(๐œ‡, ๐‘˜), ๐‘˜). ๐‘€๐‘˜+1up to first order inฮ”๐‘ก is written as ๐‘€๐‘˜+1= ๐œ• ๐‘€๐‘˜

๐œ• ๐‘ก๐‘˜ ฮ”๐‘ก+

๐‘›

โˆ‘๏ธ

๐‘–=1

๐œ• ๐‘€๐‘˜

๐œ•(๐‘ž๐‘˜)๐‘–

(๐‘“๐‘, ๐‘˜ฮ”๐‘ก+๐บ๐‘, ๐‘˜ฮ”๐’ฒ๐‘˜)๐‘– (2.64) + 1

2

๐‘›

โˆ‘๏ธ

๐‘–=1 ๐‘›

โˆ‘๏ธ

๐‘—=1

๐œ•2๐‘€๐‘˜

๐œ•(๐‘ž๐‘˜)๐‘–๐œ•(๐‘ž๐‘˜)๐‘—

(๐บ๐‘, ๐‘˜ฮ”๐’ฒ๐‘˜)๐‘–(๐บ๐‘, ๐‘˜ฮ”๐’ฒ๐‘˜)๐‘— +๐‘€๐‘˜ + O ฮ”๐‘ก2

where ๐‘“๐‘, ๐‘˜ and ๐บ๐‘, ๐‘˜ are defined as ๐‘“๐‘, ๐‘˜ = ๐‘“(๐‘ž๐‘˜, ๐‘ก๐‘˜) and ๐บ๐‘, ๐‘˜ = ๐บ(๐‘ž๐‘˜, ๐‘ก๐‘˜) for notational simplicity. The subscripts๐‘–and ๐‘— denote the corresponding vectorsโ€™๐‘–th and ๐‘—th elements. Similarly, ๐œ• ๐‘ž๐‘˜+1/๐œ• ๐œ‡up to first order inฮ”๐‘ก can be computed as

๐œ• ๐‘ž๐‘˜+1

๐œ• ๐œ‡

= ๐œ• ๐‘ž๐‘˜

๐œ• ๐œ‡

+ ๐œ• ๐‘“๐‘, ๐‘˜

๐œ• ๐‘ž๐‘˜

๐œ• ๐‘ž๐‘˜

๐œ• ๐œ‡

ฮ”๐‘ก+ ๐œ• ๐บ๐‘, ๐‘˜

๐œ• ๐œ‡

ฮ”๐’ฒ๐‘˜+ O ฮ”๐‘ก2

. (2.65)

Substituting (2.64) and (2.65) intoE๐œ๐‘˜[๐‘‰๐‘˜+1] yields E๐œ๐‘˜[๐‘‰๐‘˜+1] =E๐œ๐‘˜

โˆซ 1 0

๐œ• ๐‘ž๐‘˜+1

๐œ• ๐œ‡

โŠค

๐‘€๐‘˜+1

๐œ• ๐‘ž๐‘˜+1

๐œ• ๐œ‡ ๐‘‘๐œ‡

=๐‘‰๐‘˜+ (๐‘‘๐‘‰๐‘‘ , ๐‘˜ +๐‘‘๐‘‰๐‘ , ๐‘˜)ฮ”๐‘ก+ O (ฮ”๐‘ก3/2) where๐‘‘๐‘‰๐‘‘ , ๐‘˜ and๐‘‘๐‘‰๐‘ , ๐‘˜ are given by

๐‘‘๐‘‰๐‘‘ , ๐‘˜ =

โˆซ 1 0

๐œ• ๐‘ž๐‘˜

๐œ• ๐œ‡

โŠค

๐œ• ๐‘“๐‘, ๐‘˜

๐œ• ๐‘ž๐‘˜

โŠค

๐‘€๐‘˜+ ยค๐‘€๐‘˜+๐‘€๐‘˜

๐œ• ๐‘“๐‘, ๐‘˜

๐œ• ๐‘ž๐‘˜ ๐œ• ๐‘ž๐‘˜

๐œ• ๐œ‡ ๐‘‘๐œ‡

with๐‘€ยค๐‘˜ =๐œ• ๐‘€๐‘˜/๐œ• ๐‘ก๐‘˜ +ร๐‘›

๐‘–=1(๐œ• ๐‘€๐‘˜/๐œ•(๐‘ž๐‘˜)๐‘–)๐‘“๐‘, ๐‘˜ and ๐‘‘๐‘‰๐‘ , ๐‘˜ =

โˆซ 1 0

๏ฃฎ

๏ฃฏ

๏ฃฏ

๏ฃฏ

๏ฃฏ

๏ฃฐ

๐‘›

โˆ‘๏ธ

๐‘–=1 ๐‘›

โˆ‘๏ธ

๐‘—=1

(๐‘€๐‘˜)๐‘– ๐‘—

๐œ• ๐บ๐‘, ๐‘˜

๐œ• ๐œ‡

๐œ• ๐บ๐‘, ๐‘˜

๐œ• ๐œ‡

โŠค

๐‘– ๐‘—

+2๐œ•(๐‘€๐‘˜)๐‘–

๐œ•(๐‘ž๐‘˜)๐‘—

๐œ• ๐‘ž๐‘˜

๐œ• ๐œ‡

๐บ๐‘, ๐‘˜

๐œ• ๐บ๐‘, ๐‘˜

๐œ• ๐œ‡

โŠค

๐‘– ๐‘—

+1 2

๐œ• ๐‘ž๐‘˜

๐œ• ๐œ‡

โŠค ๐œ•2๐‘€๐‘˜

๐œ•(๐‘ž๐‘˜)๐‘–๐œ•(๐‘ž๐‘˜)๐‘—

๐œ• ๐‘ž๐‘˜

๐œ• ๐œ‡

(๐บ๐‘, ๐‘˜๐บโŠค

๐‘, ๐‘˜)๐‘– ๐‘—

๐‘‘๐œ‡.

We note that the properties of๐‘ค(๐‘˜) as a๐‘‘-dimensional sequence of zero mean un- correlated normalized Gaussian random variables are used to derive these relations.

Since๐‘‘๐‘‰๐‘‘ , ๐‘˜ +๐‘‘๐‘‰๐‘ , ๐‘˜ = โ„’๐‘‰๐‘˜ whereโ„’ is the infinitesimal differential generator, we haveE๐œ๐‘˜[๐‘‰๐‘˜+1] =๐‘‰๐‘˜+ฮ”๐‘กโ„’๐‘‰๐‘˜. Thus, the conditionE๐œ๐‘˜[๐‘‰๐‘˜+1] โ‰ค (1โˆ’๐›พ2)๐‘‰๐‘˜+๐‘š๐ถ๐‘‘ given by (2.61) in Theorem 2.9 reduces to the following inequality:

โ„’๐‘‰๐‘˜(๐‘ž๐‘˜, ๐œ•๐œ‡๐‘ž๐‘˜, ๐‘ก๐‘˜) โ‰ค โˆ’๐›พ2 ฮ”๐‘ก

๐‘‰๐‘˜(๐‘ž๐‘˜, ๐œ•๐œ‡๐‘ž๐‘˜, ๐‘ก๐‘˜) +๐‘š ๐ถ๐‘‘

ฮ”๐‘ก

. (2.66)

Finally, (2.66) with the relations หœ๐ถ๐‘=๐ถ๐‘‘/ฮ”๐‘กand๐บ๐‘˜(๐‘ž๐‘˜, ๐‘˜)=

โˆš

ฮ”๐‘ก ๐บ(๐‘ž๐‘˜, ๐‘ก๐‘˜)results in (2.62) and (2.63).

For example, in practical control applications, we use the same control input at ๐‘ก = ๐‘ก๐‘˜ for a finite time interval๐‘ก โˆˆ [๐‘ก๐‘˜, ๐‘ก๐‘ก+1). Theorems 2.5 and 2.10 indicate that if ฮ”๐‘ก is sufficiently small, a discrete-time stochastic controller can be viewed as a continuous-time counterpart with contraction rate 2๐›พ1 = ๐›พ2/ฮ”๐‘ก. We will illustrate how to select the sampling periodฮ”๐‘ก large enough without deteriorating the control performance as demonstrated in [6].

We finally remark that the steady-state upper bounds of (2.27) in Theorem 2.4, (2.39) in Theorem 2.5, and (2.54) in Theorem 2.8 are all functions of๐‘š/๐‘š. This property is to be used extensively in Chapter 4 for designing a convex optimization-based control and estimation synthesis algorithm via contraction theory.

References

[1] W. Lohmiller and J.-J. E. Slotine, โ€œOn contraction analysis for nonlinear systems,โ€Automatica, vol. 34, no. 6, pp. 683โ€“696, 1998, issn: 0005-1098.

[2] J.-J. E. Slotine, โ€œModular stability tools for distributed computation and control,โ€Int. J. Adapt. Control Signal Process., vol. 17, no. 6, pp. 397โ€“416, 2003.

[3] W. Wang and J.-J. E. Slotine, โ€œOn partial contraction analysis for coupled nonlinear oscillators,โ€ Biol. Cybern., vol. 92, no. 1, pp. 38โ€“53, Jan. 2005, issn: 0340-1200.

[4] Q. Pham, N. Tabareau, and J.-J. E. Slotine, โ€œA contraction theory approach to stochastic incremental stability,โ€ IEEE Trans. Autom. Control, vol. 54, no. 4, pp. 816โ€“820, Apr. 2009.

[5] S.-J. Chung, S. Bandyopadhyay, I. Chang, and F. Y. Hadaegh, โ€œPhase syn- chronization control of complex networks of Lagrangian systems on adaptive digraphs,โ€Automatica, vol. 49, no. 5, pp. 1148โ€“1161, 2013.

[6] H. Tsukamoto and S.-J. Chung, โ€œRobust controller design for stochastic nonlinear systems via convex optimization,โ€ IEEE Trans. Autom. Control, vol. 66, no. 10, pp. 4731โ€“4746, 2021.

[7] A. P. Dani, S.-J. Chung, and S. Hutchinson, โ€œObserver design for stochastic nonlinear systems via contraction-based incremental stability,โ€IEEE Trans.

Autom. Control, vol. 60, no. 3, pp. 700โ€“714, Mar. 2015.

[8] H. Tsukamoto and S.-J. Chung, โ€œConvex optimization-based controller de- sign for stochastic nonlinear systems using contraction analysis,โ€ in IEEE Conf. Decis. Control, Dec. 2019, pp. 8196โ€“8203.

[9] H. K. Khalil,Nonlinear Systems, 3rd. Upper Saddle River, NJ: Prentice-Hall, 2002.

[10] D. E. Kirk,Optimal Control Theory: An Introduction. Dover Publications, Apr. 2004, isbn: 0486434842.

[11] D. Angeli, โ€œA Lyapunov approach to incremental stability properties,โ€IEEE Trans. Autom. Control, vol. 47, no. 3, pp. 410โ€“421, Mar. 2002.

[12] J. Jouffroy and J.-J. E. Slotine, โ€œMethodological remarks on contraction theory,โ€ inIEEE Conf. Decis. Control, vol. 3, Dec. 2004, pp. 2537โ€“2543.

[13] W. J. Rugh, Linear Systems Theory. USA: Prentice-Hall, Inc., 1996, isbn:

0134412052.

[14] J.-J. E. Slotine and W. Li,Applied Nonlinear Control. Upper Saddle River, NJ: Pearson, 1991.

[15] H. Robbins and S. Monro, โ€œA stochastic approximation method,โ€Ann. Math.

Statist., vol. 22, no. 3, pp. 400โ€“407, 1951.

[16] P. M. Wensing and J.-J. E. Slotine, โ€œBeyond convexity โ€“ Contraction and global convergence of gradient descent,โ€ PLOS ONE, vol. 15, pp. 1โ€“29, Aug. 2020.

[17] S.-I. Amari, โ€œNatural gradient works efficiently in learning,โ€ Neural Com- put., vol. 10, no. 2, pp. 251โ€“276, 1998.

[18] C. Udriste, Convex functions and optimization methods on Riemannian manifolds. Springer Science & Business Media, 1994, vol. 297.

[19] H. Tsukamoto and S.-J. Chung, โ€œNeural contraction metrics for robust esti- mation and control: A convex optimization approach,โ€ IEEE Control Syst.

Lett., vol. 5, no. 1, pp. 211โ€“216, 2021.

[20] H. Tsukamoto, S.-J. Chung, and J.-J. E. Slotine, โ€œNeural stochastic contrac- tion metrics for learning-based control and estimation,โ€IEEE Control Syst.

Lett., vol. 5, no. 5, pp. 1825โ€“1830, 2021.

[21] I. R. Manchester and J.-J. E. Slotine, โ€œControl contraction metrics: Convex and intrinsic criteria for nonlinear feedback design,โ€ IEEE Trans. Autom.

Control, vol. 62, no. 6, pp. 3046โ€“3053, Jun. 2017.

[22] R. A. Horn and C. R. Johnson,Matrix Analysis, 2nd. Cambridge University Press, 2012, isbn: 0521548233.

[23] S. Singh, A. Majumdar, J.-J. E. Slotine, and M. Pavone, โ€œRobust online motion planning via contraction theory and convex optimization,โ€ inIEEE Int. Conf. Robot. Automat., May 2017, pp. 5883โ€“5890.

[24] L. Arnold,Stochastic Differential Equations: Theory and Applications. Wi- ley, 1974.

[25] H. J. Kushner, Stochastic Stability and Control, English. Academic Press New York, 1967, xiv, 161 p.

[26] T.-J. Tarn and Y. Rasis, โ€œObservers for nonlinear stochastic systems,โ€IEEE Trans. Autom. Control, vol. 21, no. 4, pp. 441โ€“448, Aug. 1976.

[27] M. Zakai, โ€œOn the ultimate boundedness of moments associated with so- lutions of stochastic differential equations,โ€SIAM J. Control, vol. 5, no. 4, pp. 588โ€“593, 1967.

[28] D. S. Geoffrey R. Grimmett,Probability and Random Processes, 3rd. United Kingdom: Oxford University Press, 2001.

[29] E. Mazumdar, T. Westenbroek, M. I. Jordan, and S. Shankar Sastry, โ€œHigh confidence sets for trajectories of stochastic time-varying nonlinear sys- tems,โ€ inIEEE Conf. Decis. Control, 2020, pp. 4275โ€“4280.

[30] S. Han and S.-J. Chung,Incremental nonlinear stability analysis for stochas- tic systems perturbed by Lรฉvy noise, arXiv:2103.13338, Mar. 2021.

[31] J.-J. E. Slotine and W. Lohmiller, โ€œModularity, evolution, and the binding problem: A view from stability theory,โ€Neural Netw., vol. 14, no. 2, pp. 137โ€“

145, 2001.

[32] S.-J. Chung and J.-J. E. Slotine, โ€œCooperative robot control and concurrent synchronization of Lagrangian systems,โ€IEEE Trans. Robot., vol. 25, no. 3, pp. 686โ€“700, Jun. 2009.

[33] J.-J. E. Slotine, W. Wang, and K. El Rifai, โ€œContraction analysis of synchro- nization in networks of nonlinearly coupled oscillators,โ€ inInt. Symp. Math.

Theory Netw. Syst., Jul. 2004.

[34] W. Lohmiller and J.-J. E. Slotine, โ€œNonlinear process control using contrac- tion theory,โ€AIChE Journal, vol. 46, pp. 588โ€“596, Mar. 2000.

[35] Q. Pham, โ€œAnalysis of discrete and hybrid stochastic systems by nonlinear contraction theory,โ€ inInt. Conf. Control Automat. Robot. Vision, Dec. 2008, pp. 1054โ€“1059.

C h a p t e r 3

ROBUST NONLINEAR CONTROL AND ESTIMATION VIA CONTRACTION THEORY

[1] H. Tsukamoto and S.-J. Chung, โ€œRobust controller design for stochastic nonlinear systems via convex optimization,โ€ IEEE Trans. Autom. Control, vol. 66, no. 10, pp. 4731โ€“4746, 2021.

[2] H. Tsukamoto and S.-J. Chung, โ€œConvex optimization-based controller de- sign for stochastic nonlinear systems using contraction analysis,โ€ in IEEE Conf. Decis. Control, Dec. 2019, pp. 8196โ€“8203.

As shown in Theorem 2.4 for deterministic disturbance and in Theorem 2.5 for stochastic disturbance, contraction theory provides explicit bounds on the distance of any couple of perturbed system trajectories. This property is useful in design- ing robust and optimal feedback controllers for a nonlinear system such as Hโˆž control [1]โ€“[11], which attempts to minimize the systemL2gain for optimal distur- bance attenuation.

Most of such feedback control and estimation schemes are, however, based on the assumption that we know a Lyapunov function candidate. This chapter thus delineates one approach to solve a nonlinear optimal feedback control problem via contraction theory [12], [13], thereby proposing one explicit way to construct a Lyapunov function and contraction metric for general nonlinear systems for the sake of robustness. This approach is also utilizable for optimal state estimation problems as shall be seen in Chapter 4.

We consider the following smooth nonlinear system, perturbed by bounded deter- ministic disturbances ๐‘‘๐‘(๐‘ฅ , ๐‘ก) with sup๐‘ฅ ,๐‘กโˆฅ๐‘‘๐‘(๐‘ฅ , ๐‘ก) โˆฅ = ๐‘‘ยฏ๐‘ โˆˆ Rโ‰ฅ0 or by Gaussian white noise, driven by a Wiener process๐’ฒ(๐‘ก) with sup๐‘ฅ ,๐‘กโˆฅ๐บ๐‘(๐‘ฅ , ๐‘ก) โˆฅ๐น =๐‘”ยฏ๐‘ โˆˆRโ‰ฅ0:

ยค

๐‘ฅ = ๐‘“(๐‘ฅ , ๐‘ก) +๐ต(๐‘ฅ , ๐‘ก)๐‘ข+๐‘‘๐‘(๐‘ฅ , ๐‘ก) (3.1)

๐‘‘๐‘ฅ = (๐‘“(๐‘ฅ , ๐‘ก) +๐ต(๐‘ฅ , ๐‘ก)๐‘ข)๐‘‘ ๐‘ก+๐บ๐‘(๐‘ฅ , ๐‘ก)๐‘‘๐’ฒ(๐‘ก) (3.2)

ยค

๐‘ฅ๐‘‘ = ๐‘“(๐‘ฅ๐‘‘, ๐‘ก) +๐ต(๐‘ฅ๐‘‘, ๐‘ก)๐‘ข๐‘‘ (3.3)

where ๐‘ฅ : Rโ‰ฅ0 โ†ฆโ†’ R๐‘› is the system state, ๐‘ข โˆˆ R๐‘š is the system control input, ๐‘“ : R๐‘›ร— Rโ‰ฅ0 โ†ฆโ†’ R๐‘› and ๐ต : R๐‘› ร—Rโ‰ฅ0 โ†ฆโ†’ R๐‘›ร—๐‘š are known smooth functions,

๐‘‘๐‘ :R๐‘›ร—Rโ‰ฅ0 โ†ฆโ†’R๐‘› and๐บ๐‘ : R๐‘›ร—Rโ‰ฅ0โ†ฆโ†’ R๐‘›ร—๐‘ค are unknown bounded functions for external disturbances, and๐’ฒ : Rโ‰ฅ0 โ†ฆโ†’R๐‘คis a๐‘ค-dimensional Wiener process.

Also, for (3.3),๐‘ฅ๐‘‘ : Rโ‰ฅ0 โ†ฆโ†’ R๐‘› and๐‘ข๐‘‘ : Rโ‰ฅ0 โ†ฆโ†’ R๐‘š denote the desired target state and control input trajectories, respectively.

Remark 3.1. We consider control-affine nonlinear systems(3.1)โ€“(3.3)in Chapter 3, 4, and 6 โ€“ 8.1. This is primarily because the controller design techniques for control-affine nonlinear systems are less complicated than those for control non- affine systems (which often result in๐‘ขgiven implicitly by๐‘ข =๐‘˜(๐‘ฅ , ๐‘ข, ๐‘ก) [14], [15]), but still utilizable even for the latter, e.g., by treating๐‘ขยคas another control input (see Example 3.1), or by solving the implicit equation ๐‘ข = ๐‘˜(๐‘ฅ , ๐‘ข, ๐‘ก) iteratively with a discrete-time controller (see Example 3.2 and Remark 3.3).

Example 3.1. By using๐‘ขยคinstead of๐‘ขin(3.1)and(3.2), a control non-affine system,

ยค

๐‘ฅ = ๐‘“(๐‘ฅ , ๐‘ข, ๐‘ก), can be rewritten as ๐‘‘

๐‘‘ ๐‘ก

"

๐‘ฅ ๐‘ข

#

=

"

๐‘“(๐‘ฅ , ๐‘ข, ๐‘ก) 0

# +

"

0 I

#

ยค ๐‘ข

which can be viewed as a control-affine nonlinear system with the state [๐‘ฅโŠค, ๐‘ขโŠค]โŠค and control๐‘ขยค.

Example 3.2. One drawback of the technique in Example 3.1 is that we have to control ๐‘ขยค instead of ๐‘ข, which could be difficult in practice. In this case, we can utilize the following control non-affine nonlinear system decomposed into control- affine and non-affine parts:

ยค

๐‘ฅ = ๐‘“(๐‘ฅ , ๐‘ข, ๐‘ก) = ๐‘“๐‘Ž(๐‘ฅ , ๐‘ก) +๐ต๐‘Ž(๐‘ฅ , ๐‘ก)๐‘ข+๐‘Ÿ(๐‘ฅ , ๐‘ข, ๐‘ก)

where ๐‘Ÿ(๐‘ฅ , ๐‘ข, ๐‘ก) = ๐‘“(๐‘ฅ , ๐‘ข, ๐‘ก) โˆ’ ๐‘“๐‘Ž(๐‘ฅ , ๐‘ก) โˆ’ ๐ต๐‘Ž(๐‘ฅ , ๐‘ก)๐‘ข. The controller ๐‘ข can now be designed implicitly as

๐ต๐‘Ž(๐‘ฅ , ๐‘ก)๐‘ข= ๐ต๐‘Ž(๐‘ฅ , ๐‘ก)๐‘ขโˆ—โˆ’๐‘Ÿ(๐‘ฅ , ๐‘ข, ๐‘ก) (3.4) where ๐‘ขโˆ— is a stabilizing controller for the control-affine system ๐‘ฅยค = ๐‘“๐‘Ž(๐‘ฅ , ๐‘ก) + ๐ต๐‘Ž(๐‘ฅ , ๐‘ก)๐‘ขโˆ—. Since solving such an implicit equation in (3.4) in real-time could be unrealistic in practice, we will derive a learning-based approach to solve it iteratively for unknown ๐‘Ÿ(๐‘ฅ , ๐‘ข, ๐‘ก), without deteriorating its stability performance (see Lemma 8.2 and Theorem 8.4 of Chapter 8).

3.1 Overview of Nonlinear Control and Estimation

We briefly summarize the advantages and disadvantages of existing nonlinear feed- back control and state estimation schemes, so that one can identify which strategy is appropriate for their study and refer to the relevant parts of this thesis.

Table 3.1: Comparison between the SDC and CCM formulation (note that๐›พ(๐œ‡ = 0, ๐‘ก)=๐‘ฅ๐‘‘and๐›พ(๐œ‡=1, ๐‘ก) =๐‘ฅ).

SDC (Theorem 4.2) [12], [13],

[16]โ€“[18] CCM (Theorem 4.6) [19], [20]

Control law ๐‘ข = ๐‘ข๐‘‘ โˆ’๐พ(๐‘ฅ , ๐‘ฅ๐‘‘, ๐‘ข๐‘‘, ๐‘ก) (๐‘ฅโˆ’๐‘ฅ๐‘‘)

or๐‘ข๐‘‘โˆ’๐พ(๐‘ฅ , ๐‘ก) (๐‘ฅโˆ’๐‘ฅ๐‘‘) ๐‘ข=๐‘ข๐‘‘+โˆซ1

0 ๐‘˜(๐›พ(๐œ‡, ๐‘ก), ๐œ•๐œ‡๐›พ(๐œ‡, ๐‘ก), ๐‘ข, ๐‘ก)๐‘‘๐œ‡ Computation Evaluates๐พ(๐‘ฅ , ๐‘ฅ๐‘‘, ๐‘ข๐‘‘, ๐‘ก)for given

(๐‘ฅ , ๐‘ฅ๐‘‘, ๐‘ข๐‘‘, ๐‘ก)as in LTV systems

Computes geodesics๐›พfor given(๐‘ฅ , ๐‘ฅ๐‘‘, ๐‘ก) and integrates๐‘˜

Generality Captures nonlinearity by (multi-

ple) SDC matrices Handles general differential dynamics Contraction Depends on(๐‘ฅ , ๐‘ฅ๐‘‘, ๐‘ข๐‘‘, ๐‘ก)or(๐‘ฅ , ๐‘ก)

(partial contraction) Depends on(๐‘ฅ , ๐‘ก)(contraction)

3.1.I Systems with Known Lyapunov Functions

As discussed in Sec. 1.2, there are several nonlinear systems equipped with a known contraction metric/Lyapunov function, such as Lagrangian systems [21, p. 392], whose inertia matrixH (q) defines its contraction metric (see Example 2.6), or the nonlinear SLAM problem [18], [22] with virtual synthetic measurements, which can be reduced to an LTV estimation problem [22]. Once we have a contraction metric/Lyapunov function, stabilizing control and estimation laws can be easily derived by using, e.g., [23]โ€“[25]. Thus, those dealing primarily with such nonlinear systems should skip this chapter and proceed to Part II of this thesis (Chapter 5 โ€“ 8) on learning-based and data-driven control using contraction theory. Note that these known contraction metrics are not necessarily optimal, and the techniques to be derived in Chapter 3 and Chapter 4 are for obtaining contraction metrics with an optimal disturbance attenuation property [12], [13].

3.1.II Linearization of Nonlinear Systems

If a contraction metric of a given nonlinear system is unknown, we could linearize it to apply methodologies inspired by LTV systems theory such asHโˆžcontrol [6]โ€“[11], iterative Linear Quadratic Regulator (iLQR) [26], [27], or Extended Kalman Filter (EKF). Their stability is typically analyzed by decomposing ๐‘“(๐‘ฅ , ๐‘ก) as ๐‘“(๐‘ฅ , ๐‘ก) =

๐ด๐‘ฅ+ (๐‘“(๐‘ฅ , ๐‘ก) โˆ’๐ด๐‘ฅ)assuming that the nonlinear part ๐‘“(๐‘ฅ , ๐‘ก) โˆ’๐ด๐‘ฅ is bounded, or by finding a local contraction region for the sake of local exponential stability as in [16], [28]. Since the decomposition ๐‘“(๐‘ฅ , ๐‘ก) = ๐ด๐‘ฅ+ (๐‘“(๐‘ฅ , ๐‘ก) โˆ’ ๐ด๐‘ฅ) allows applying the result of Theorem 2.4, we could exploit the techniques in Chapter 3 and Chapter 4 for providing formal robustness and optimality guarantees for the LTV systems- type approaches. For systems whose nonlinear part ๐‘“(๐‘ฅ , ๐‘ก) โˆ’ ๐ด๐‘ฅ is not necessarily bounded, Sec. 8.2.II elucidates how contraction theory can be used to stabilize them with the learned dynamics for control synthesis.

3.1.III State-Dependent Coefficient (SDC) Formulation

It is shown in [12], [13], [16]โ€“[18] that the SDC-based control and estimation [29]โ€“

[32], which capture nonlinearity using a state-dependent matrix๐ด(๐‘ฅ , ๐‘ก)s.t. ๐‘“(๐‘ฅ , ๐‘ก)= ๐ด(๐‘ฅ , ๐‘ก)๐‘ฅ (e.g., we have ๐ด(๐‘ฅ , ๐‘ก) = cos๐‘ฅ for ๐‘“(๐‘ฅ , ๐‘ก) = ๐‘ฅcos๐‘ฅ), result in exponential boundedness of system trajectories both for deterministic and stochastic systems due to Theorems 2.4 and 2.5 [16]. Because of the extended linear form of SDC (see Table 3.1), the results to be presented in Chapter 3 โ€“ 4 based on the SDC formulation are applicable to linearized dynamics that can be viewed as an LTV system with some modifications (see Remark 3.2).

This idea is slightly generalized in [17] to explicitly consider incremental stability with respect to a target trajectory (e.g.,๐‘ฅ๐‘‘ for control and๐‘ฅ for estimation) instead of using๐ด(๐‘ฅ , ๐‘ก)๐‘ฅ = ๐‘“(๐‘ฅ , ๐‘ก). Let us derive the following lemma for this purpose [12], [13], [17], [18], [32]. Let us derive the following lemma for this purpose [12], [13], [17], [18], [32].

Lemma 3.1. Let ๐‘“ : R๐‘› ร—Rโ‰ฅ0 โ†ฆโ†’ R๐‘› and ๐ต : R๐‘› ร—Rโ‰ฅ0 โ†ฆโ†’ R๐‘›ร—๐‘š be piecewise continuously differentiable functions. Then there exists a matrix-valued function ๐ด :R๐‘›ร—R๐‘›ร—R๐‘š ร—Rโ‰ฅ0โ†ฆโ†’R๐‘›ร—๐‘›s.t.,โˆ€๐‘ โˆˆR๐‘›,๐‘ ยฏโˆˆR๐‘›,๐‘ขยฏ โˆˆR๐‘š, and๐‘ก โˆˆRโ‰ฅ0, ๐ด(๐‘ ,๐‘ ,ยฏ ๐‘ข, ๐‘กยฏ )e = ๐‘“(๐‘ , ๐‘ก) +๐ต(๐‘ , ๐‘ก)๐‘ขยฏโˆ’ ๐‘“(๐‘ , ๐‘กยฏ ) โˆ’๐ต(๐‘ , ๐‘กยฏ )๐‘ขยฏ

wheree=๐‘ โˆ’๐‘ ยฏ, and one such ๐ดis given as follows:

๐ด(๐‘ ,๐‘ ,ยฏ ๐‘ข, ๐‘กยฏ ) =

โˆซ 1 0

๐œ•๐‘“ยฏ

๐œ• ๐‘ 

(๐‘ ๐‘ + (1โˆ’๐‘)๐‘ ,ยฏ ๐‘ข, ๐‘กยฏ )๐‘‘๐‘ (3.5) where ๐‘“ยฏ(๐‘ ,๐‘ข, ๐‘กยฏ ) = ๐‘“(๐‘ , ๐‘ก) +๐ต(๐‘ , ๐‘ก)๐‘ขยฏ. We call๐ดan SDC matrix if it is constructed to satisfy the controllability (or observability for estimation) condition. Furthermore, the choice of ๐ด is not unique for ๐‘› โ‰ฅ 2, where ๐‘› is the number of states, and

the convex combination of such non-unique SDC matrices also verifies extended linearization as follows:

๐‘“(๐‘ , ๐‘ก) +๐ต(๐‘ , ๐‘ก)๐‘ขยฏโˆ’ ๐‘“(๐‘ , ๐‘กยฏ ) โˆ’๐ต(๐‘ , ๐‘กยฏ )๐‘ขยฏ

= ๐ด(๐œš, ๐‘ ,๐‘ ,ยฏ ๐‘ข, ๐‘กยฏ ) (๐‘ โˆ’๐‘ ยฏ) =

๐‘ ๐ด

โˆ‘๏ธ

๐‘–=1

๐œš๐‘–๐ด๐‘–(๐‘ ,๐‘ ,ยฏ ๐‘ข, ๐‘กยฏ ) (๐‘ โˆ’๐‘ ยฏ) (3.6) where ๐œš = (๐œš1,ยท ยท ยท , ๐œš๐‘ 

๐ด), ร๐‘ ๐ด

๐‘–=1๐œš๐‘– = 1, ๐œš๐‘– โ‰ฅ 0, and each ๐ด๐‘– satisfies the relation

ยฏ

๐‘“(๐‘ ,๐‘ข, ๐‘กยฏ ) โˆ’ ๐‘“ยฏ(๐‘ ,ยฏ ๐‘ข, ๐‘กยฏ ) = ๐ด๐‘–(๐‘ ,๐‘ ,ยฏ ๐‘ข, ๐‘กยฏ ) (๐‘ โˆ’๐‘ ยฏ).

Proof. The first statement on (3.5) follows from the integral relation given as

โˆซ 1 0

๐‘‘๐‘“ยฏ ๐‘‘๐‘

(๐‘ ๐‘ + (1โˆ’๐‘)๐‘ ,ยฏ ๐‘ข, ๐‘กยฏ )๐‘‘๐‘= ๐‘“ยฏ(๐‘ ,๐‘ข, ๐‘กยฏ ) โˆ’ ๐‘“ยฏ(๐‘ ,ยฏ ๐‘ข, ๐‘กยฏ ).

If there are multiple SDC matrices ๐ด๐‘–, we clearly have ๐œš๐‘–๐ด๐‘–(๐‘ ,๐‘ ,ยฏ ๐‘ข, ๐‘กยฏ ) (๐‘  โˆ’ ๐‘ ยฏ) = ๐œš๐‘–(๐‘“ยฏ(๐‘ ,๐‘ข, ๐‘กยฏ ) โˆ’ ๐‘“ยฏ(๐‘ ,ยฏ ๐‘ข, ๐‘กยฏ )), โˆ€๐‘–, and therefore, the relationร๐‘ ๐ด

๐‘–=1๐œš๐‘– = 1, ๐œš๐‘– โ‰ฅ 0 gives (3.6).

Example 3.3. Let us illustrate how Lemma 3.1 can be used in practice taking the following nonlinear system as an example:

ยค

๐‘ฅ =[๐‘ฅ2,โˆ’๐‘ฅ1๐‘ฅ2]โŠค+ [0,cos๐‘ฅ1]โŠค๐‘ข (3.7)

where๐‘ฅ = [๐‘ฅ1, ๐‘ฅ2]โŠค. If we use(๐‘ ,๐‘ ,ยฏ ๐‘ขยฏ) = (๐‘ฅ , ๐‘ฅ๐‘‘, ๐‘ข๐‘‘)in Lemma 3.1 for a given target trajectory (๐‘ฅ๐‘‘, ๐‘ข๐‘‘)that satisfies(3.7), evaluating the integral of(3.5)gives

๐ด1(๐‘ฅ , ๐‘ฅ๐‘‘, ๐‘ข๐‘‘, ๐‘ก) =โˆ’

"

0 1

๐‘ฅ2+๐‘ฅ2๐‘‘

2 โˆ’ ๐‘ข๐‘‘(cos๐‘ฅ1โˆ’cos๐‘ฅ๐‘‘1)

๐‘ฅ1โˆ’๐‘ฅ๐‘‘1

๐‘ฅ1+๐‘ฅ1๐‘‘ 2

#

(3.8) due to the relation ๐œ•๐‘“ยฏ/๐œ• ๐‘  = 0 1

โˆ’๐‘ 2 โˆ’๐‘ 1

+ 0 0

โˆ’๐‘ข๐‘‘sin๐‘ 1 0

for ๐‘“ยฏ(๐‘ , ๐‘ข๐‘‘, ๐‘ก) = ๐‘“(๐‘ , ๐‘ก) + ๐ต(๐‘ , ๐‘ก)๐‘ข๐‘‘, where๐‘ฅ๐‘‘ = [๐‘ฅ1๐‘‘, ๐‘ฅ2๐‘‘]โŠค. Note that we have

(cos๐‘ฅ1โˆ’cos๐‘ฅ๐‘‘1) ๐‘ฅ1โˆ’๐‘ฅ๐‘‘1

=โˆ’sin๐‘ฅ1+๐‘ฅ1๐‘‘ 2

sinc๐‘ฅ1โˆ’๐‘ฅ1๐‘‘ 2

and thus ๐ด(๐‘ฅ , ๐‘ฅ๐‘‘, ๐‘ข๐‘‘, ๐‘ก) is defined for all ๐‘ฅ, ๐‘ฅ๐‘‘, ๐‘ข๐‘‘, and ๐‘ก. The SDC matrix (3.8) indeed verifies๐ด1(๐‘ฅ , ๐‘ฅ๐‘‘, ๐‘ข๐‘‘, ๐‘ก) (๐‘ฅโˆ’๐‘ฅ๐‘‘) = ๐‘“ยฏ(๐‘ฅ , ๐‘ก) โˆ’ ๐‘“ยฏ(๐‘ฅ๐‘‘, ๐‘ก).

We can see that the following is also an SDC matrix of the nonlinear system(3.7):

๐ด2(๐‘ฅ , ๐‘ฅ๐‘‘, ๐‘ข๐‘‘, ๐‘ก) =โˆ’

"

0 1

๐‘ฅ2โˆ’ ๐‘ข๐‘‘(cos๐‘ฅ1โˆ’cos๐‘ฅ๐‘‘1)

๐‘ฅ1โˆ’๐‘ฅ๐‘‘1

๐‘ฅ1๐‘‘

#

. (3.9)

Therefore, the convex combination of๐ด1in(3.8)and๐ด2in(3.9), ๐ด= ๐œš1๐ด1+๐œš2๐ด2 with ๐œš1+ ๐œš2=1, ๐œš1, ๐œš2 โ‰ฅ 0, is also an SDC matrix due to Lemma 3.1.

The major advantage of the formalism in Lemma 3.1 lies in its systematic connection to LTV systems based on uniform controllability and observability, adequately accounting for the nonlinear nature of underlying dynamics through๐ด(๐œš, ๐‘ฅ , ๐‘ฅ๐‘‘, ๐‘ข๐‘‘, ๐‘ก) for global stability, as shall be seen in Chapter 3 and Chapter 4. Since ๐ดdepends also on (๐‘ฅ๐‘‘, ๐‘ข๐‘‘) in this case unlike the original SDC matrix, we could consider contraction metrics using a positive definite matrix๐‘€(๐‘ฅ , ๐‘ฅ๐‘‘, ๐‘ข๐‘‘, ๐‘ก)instead of๐‘€(๐‘ฅ , ๐‘ก) in Definition 2.3, to improve the representation power of ๐‘€ at the expense of computational efficiency. Another interesting point is that the non-uniqueness of ๐ด in Lemma 3.1 for ๐‘› โ‰ฅ 2 creates additional degrees of freedom for selecting the coefficients ๐œš, which can also be treated as decision variables in constructing optimal contraction metrics as proposed in [12], [13], [18].

We focus mostly on the generalized SDC formulation in Chapter 3 and Chapter 4, as it yields optimal control and estimation laws with global stability [17] while keeping the analysis simple enough to be understood as in LTV systems theory.

Remark 3.2. This does not mean that contraction theory works only for the SDC parameterized nonlinear systems but implies that it can be used with the other techniques discussed in Sec. 3.1. For example, due to the extended linear form given in Table 3.1, the results to be presented in Chapter 3 and in Chapter 4 based on the SDC formulation are applicable to linearized dynamics that can be viewed as an LTV system with some modifications, regarding the dynamics modeling error term as an external disturbance as in Sec. 3.1.II. Also, the original SDC formulation with respect to a fixed point (e.g., (๐‘ ,๐‘ ,ยฏ ๐‘ขยฏ) = (๐‘ฅ ,0,0) in Lemma 3.1) can still be used to obtain contraction conditions independent of a target trajectory (๐‘ฅ๐‘‘, ๐‘ข๐‘‘) (see Theorem 3.2 for details).

3.1.IV Control Contraction Metric (CCM) Formulation

We could also consider using the partial derivative of ๐‘“ of the dynamical system directly for control synthesis through differential state feedback๐›ฟ๐‘ข =๐‘˜(๐‘ฅ , ๐›ฟ๐‘ฅ , ๐‘ข, ๐‘ก).

This idea, formulated as the concept of a CCM [3], [14], [15], [19], [20], [33], constructs contraction metrics with global stability guarantees independently of target trajectories, achieving greater generality while requiring added computation in evaluating integrals involving minimizing geodesics. Similar to the CCM, we could design a state estimator using a general formulation based on geodesics distances between trajectories [34], [35]. These approaches are well compatible with the

convex optimization-based schemes in Chapter 4, and hence will be discussed in Sec. 4.3.

The differences between the SDC and CCM formulation are summarized in Ta- ble 3.1. Considering such trade-offs would help determine which form of the control law is the best fit when using contraction theory for nonlinear stabilization.

Remark 3.3. For control non-affine nonlinear systems, we could find ๐‘“(๐‘ฅ , ๐‘ข, ๐‘ก) โˆ’ ๐‘“(๐‘ฅ๐‘‘, ๐‘ข๐‘‘, ๐‘ก) = ๐ด(๐‘ฅ , ๐‘ฅ๐‘‘, ๐‘ข, ๐‘ข๐‘‘, ๐‘ก) (๐‘ฅ โˆ’๐‘ฅ๐‘‘) +๐ต(๐‘ฅ , ๐‘ฅ๐‘‘, ๐‘ข, ๐‘ข๐‘‘, ๐‘ก) (๐‘ข โˆ’๐‘ข๐‘‘) by Lemma 3.1 on the SDC formulation and use it in Theorem 4.2, although (3.10) has to be solved implicitly as ๐ตdepends on๐‘ข in this case. A similar approach for the CCM formulation can be found in [14], [15]. As discussed in Example 3.2, designing such implicit control laws will be discussed in Lemma 8.2 and Theorem 8.4 of Sec. 8.2.II.

3.2 LMI Conditions for Contraction Metrics

We design a nonlinear feedback tracking control law parameterized by a matrix- valued function๐‘€(๐‘ฅ , ๐‘ฅ๐‘‘, ๐‘ข๐‘‘, ๐‘ก) (or๐‘€(๐‘ฅ , ๐‘ก), see Theorem 3.2) as follows:

๐‘ข =๐‘ข๐‘‘โˆ’๐พ(๐‘ฅ , ๐‘ฅ๐‘‘, ๐‘ข๐‘‘, ๐‘ก) (๐‘ฅโˆ’๐‘ฅ๐‘‘) (3.10)

=๐‘ข๐‘‘โˆ’๐‘…(๐‘ฅ , ๐‘ฅ๐‘‘, ๐‘ข๐‘‘, ๐‘ก)โˆ’1๐ต(๐‘ฅ , ๐‘ก)โŠค๐‘€(๐‘ฅ , ๐‘ฅ๐‘‘, ๐‘ข๐‘‘, ๐‘ก) (๐‘ฅโˆ’๐‘ฅ๐‘‘)

where ๐‘…(๐‘ฅ , ๐‘ฅ๐‘‘, ๐‘ข๐‘‘, ๐‘ก) โ‰ป 0 is a weight matrix on the input๐‘ข and๐‘€(๐‘ฅ , ๐‘ฅ๐‘‘, ๐‘ข๐‘‘, ๐‘ก) โ‰ป 0 is a positive definite matrix (which satisfies the matrix inequality constraints for a contraction metric, to be given in Theorem 3.1). As discussed in Sec. 3.1.III, the extended linear form of the tracking control (3.10) enables LTV systems-type ap- proaches to Lyapunov function construction, while being general enough to capture the nonlinearity of the underlying dynamics due to Lemma 3.2 [36].

Lemma 3.2. Consider a general feedback controller๐‘ขdefined as๐‘ข= ๐‘˜(๐‘ฅ , ๐‘ฅ๐‘‘, ๐‘ข๐‘‘, ๐‘ก) with๐‘˜(๐‘ฅ๐‘‘, ๐‘ฅ๐‘‘, ๐‘ข๐‘‘, ๐‘ก) =๐‘ข๐‘‘, where๐‘˜ :R๐‘›ร—R๐‘›ร—R๐‘š ร—Rโ‰ฅ0โ†ฆโ†’ R๐‘š. If๐‘˜ is piecewise continuously differentiable, then โˆƒ๐พ : R๐‘› ร— R๐‘› ร— R๐‘š ร— Rโ‰ฅ0 โ†ฆโ†’ R๐‘šร—๐‘› s.t. ๐‘ข = ๐‘˜(๐‘ฅ , ๐‘ฅ๐‘‘, ๐‘ข๐‘‘, ๐‘ก) =๐‘ข๐‘‘โˆ’๐พ(๐‘ฅ , ๐‘ฅ๐‘‘, ๐‘ข๐‘‘, ๐‘ก) (๐‘ฅโˆ’๐‘ฅ๐‘‘).

Proof. Using๐‘˜(๐‘ฅ๐‘‘, ๐‘ฅ๐‘‘, ๐‘ข๐‘‘, ๐‘ก) =๐‘ข๐‘‘,๐‘ขcan be decomposed as๐‘ข=๐‘ข๐‘‘+(๐‘˜(๐‘ฅ , ๐‘ฅ๐‘‘, ๐‘ข๐‘‘, ๐‘ก)โˆ’

๐‘˜(๐‘ฅ๐‘‘, ๐‘ฅ๐‘‘, ๐‘ข๐‘‘, ๐‘ก)). Since we have๐‘˜(๐‘ฅ , ๐‘ฅ๐‘‘, ๐‘ข๐‘‘, ๐‘ก) โˆ’๐‘˜(๐‘ฅ๐‘‘, ๐‘ฅ๐‘‘, ๐‘ข๐‘‘, ๐‘ก) =โˆซ1

0 (๐‘‘ ๐‘˜(๐‘๐‘ฅ+ (1โˆ’ ๐‘)๐‘ฅ๐‘‘, ๐‘ฅ๐‘‘, ๐‘ข๐‘‘, ๐‘ก)/๐‘‘๐‘)๐‘‘๐‘, selecting๐พas

๐พ =โˆ’

โˆซ 1 0

๐œ• ๐‘˜

๐œ• ๐‘ฅ

(๐‘๐‘ฅ+ (1โˆ’๐‘)๐‘ฅ๐‘‘, ๐‘ฅ๐‘‘, ๐‘ข๐‘‘, ๐‘ก)๐‘‘๐‘ gives the desired relation [36].