When a control system has all its vector fields tangent to the level set of a given smooth function u at a point ˆx, under appropriate assumptions that function can still have a negative rate of decrease with respect to the trajectories of the control system in an appropriate sense. In the case when the system is symmetric and u has a decrease rate of the second order, we characterise this fact and investigate the existence of a best possible rate in the class of piecewise constant controls. The problem turns out to be purely algebraic and depends on the eigenvalues of matrices constructed from a basis matrix whose elements are the second order Lie derivatives of u at ˆx with respect to the vector fields of the system.
Citation: Mauro Costantini, Pierpaolo Soravia. On the optimal second order decrease rate for nonlinear and symmetric control systems[J]. AIMS Mathematics, 2024, 9(10): 28232-28255. doi: 10.3934/math.20241369
[1] | Wajdi Kallel, Noura Allugmani . Finite-time stabilization of stochastic systems with varying parameters. AIMS Mathematics, 2023, 8(8): 17687-17701. doi: 10.3934/math.2023903 |
[2] | Hongyan Wang, Shaoping Jiang, Yudie Hu, Supaporn Lonapalawong . Analysis of drug-resistant tuberculosis in a two-patch environment using Caputo fractional-order modeling. AIMS Mathematics, 2024, 9(11): 32696-32733. doi: 10.3934/math.20241565 |
[3] | Jinji Du, Chuangliang Qin, Yuanxian Hui . Optimal control and analysis of a stochastic SEIR epidemic model with nonlinear incidence and treatment. AIMS Mathematics, 2024, 9(12): 33532-33550. doi: 10.3934/math.20241600 |
[4] | Aziz Belmiloudi . Time-varying delays in electrophysiological wave propagation along cardiac tissue and minimax control problems associated with uncertain bidomain type models. AIMS Mathematics, 2019, 4(3): 928-983. doi: 10.3934/math.2019.3.928 |
[5] | Saim Ahmed, Ahmad Taher Azar, Ibraheem Kasim Ibraheem . Model-free scheme using time delay estimation with fixed-time FSMC for the nonlinear robot dynamics. AIMS Mathematics, 2024, 9(4): 9989-10009. doi: 10.3934/math.2024489 |
[6] | Zhiqiang Chen, Alexander Yurievich Krasnov . Disturbance observer based fixed time sliding mode control for a class of uncertain second-order nonlinear systems. AIMS Mathematics, 2025, 10(3): 6745-6763. doi: 10.3934/math.2025309 |
[7] | Yuanyuan Cheng, Yuan Li . A novel event-triggered constrained control for nonlinear discrete-time systems. AIMS Mathematics, 2023, 8(9): 20530-20545. doi: 10.3934/math.20231046 |
[8] | Miao Xiao, Zhe Lin, Qian Jiang, Dingcheng Yang, Xiongfeng Deng . Neural network-based adaptive finite-time tracking control for multiple inputs uncertain nonlinear systems with positive odd integer powers and unknown multiple faults. AIMS Mathematics, 2025, 10(3): 4819-4841. doi: 10.3934/math.2025221 |
[9] | Liandi Fang, Li Ma, Shihong Ding . Finite-time fuzzy output-feedback control for $ p $-norm stochastic nonlinear systems with output constraints. AIMS Mathematics, 2021, 6(3): 2244-2267. doi: 10.3934/math.2021136 |
[10] | Asma Hanif, Azhar Iqbal Kashif Butt . Atangana-Baleanu fractional dynamics of dengue fever with optimal control strategies. AIMS Mathematics, 2023, 8(7): 15499-15535. doi: 10.3934/math.2023791 |
When a control system has all its vector fields tangent to the level set of a given smooth function u at a point ˆx, under appropriate assumptions that function can still have a negative rate of decrease with respect to the trajectories of the control system in an appropriate sense. In the case when the system is symmetric and u has a decrease rate of the second order, we characterise this fact and investigate the existence of a best possible rate in the class of piecewise constant controls. The problem turns out to be purely algebraic and depends on the eigenvalues of matrices constructed from a basis matrix whose elements are the second order Lie derivatives of u at ˆx with respect to the vector fields of the system.
In this paper, we consider a nonlinear and symmetric control system,
{˙xt=σ(xt)at=∑mi=1(ai)tσi(xt),x0=x∈Rn, | (1.1) |
where σ:Rn→Mn,m(R) is continuous and matrix valued, the control function a⋅:R+→B1 is measurable, a=(ai)i=1,…,m, and B1={a∈Rm:|a|≤1} is the control set. The columns σi of σ=(σi)i=1,…,m are a family of smooth vector fields in Rn. We call xt a trajectory of the control system (depending on the control and the initial condition). Given a smooth function u:Rn→R and a point ˆx∈Rn such that ∇u(ˆx)≠0 but ∇uσ(ˆx)=0, we are interested in the behaviour of the trajectories of (1.1) in the neighborhood of ˆx satisfying at their initial point u(x0)≥u(ˆx), in particular the fact that trajectories can nonetheless enter in small time the interior of the sublevel set
U={x:u(x)≤u(ˆx)}. | (1.2) |
This property is named "small time local attainability" (STLA) of U at ˆx by the control system, and it is more precisely stated by saying that the minimum time function to reach U is continuous (and null) at ˆx.
When we deal with a simple dynamical system
˙xt=f(xt),x0=x, |
the classical Lyapunov method, see e.g., Lyapunov [16], aims at finding a function u having at the origin a strict minimum, with the additional property that u is strictly decreasing along the trajectory of the system, in order to show that the trajectory reaches the origin (either in finite time or asymptotically). The classical Lyapunov method requires that
Hfu(x):=⟨f(x),∇u(x)⟩<0,x≠0 |
in a neighborhood of the origin. The operator Hf:C1(Rn)→C(R) is called the Hamiltonian of the system and Hfu is also called the Lie derivative of u with respect to the vector field f. However, even when Hfu(x0)=0, at some x0≠0, one can still prove that after a short time t>0, one has u(xt)<u(x0) by using the second order Taylor expansion
u(xt)=u(x0)+12d2dt2u(xt)|t=0t2+t2o(1) |
provided
d2dt2u(xt)|t=0=Hf∘Hfu(x0)=:H(2)fu(x0)<0. |
The previous equation defines the second order Lie derivative of u at x0 and we name the operator H(2)f:C2(Rn)→C(R) a second order Hamiltonian, see e.g., [25,27]. We may view the quantity (1/2)H(2)fu(x0) as a second order decrease rate of u at x0 along the trajectory of the system.
For control systems the situation is more varied. One can extend the Lyapunov idea to control systems like (1.1), see e.g., Artstein [1] and Clarke et al. [8], even in much more general cases, by imposing, in a suitable sense if u is nonsmooth,
H(x,∇u(x))=|∇uσ(x)|=maxa∈B1{−Hσ(x)au(x)}>0,x≠0 | (1.3) |
in a neighborhood of the origin (u is a control Lyapunov function). If u is smooth at x, the negative quantity
−H(x,∇u(x)) |
can still be seen as the optimal decrease rate of the trajectories of the control system at x, and the control
ˉa=t(∇uσ(x))H(x,∇u(x))∈B1, |
where the maximum is attained in (1.3) as an optimal control in that sense. Control Lyapunov functions will not be smooth in general but allowing the inequality (1.3) to be large, we may sometimes be able to find better behaved functions, see e.g., Motta and Rampazzo [20], when we look for conditions also of a different nature. When at some point ˆx∈Rn we have ∇uσ(ˆx)=0, i.e., all the vector fields of the control system are tangent to the level set of u at ˆx, it is well known that if there exists a Lie bracket of the vector fields σi pointing inward the sublevel set in (1.2) at ˆx, we can still reach the interior of U in small time with the trajectories of the control system starting at ˆx or in its neighborhood, although u is no longer strictly decreasing along the trajectories in general. The goal of this paper is to determine, in this case, an optimal second order decrease rate for the trajectories of the system (1.1) in some appropriate sense. We have not seen this optimality question treated before in the literature, but we believe it is interesting in order to investigate the differences between local optimization and global optimization when we want, for instance, to reach a target. We will restrict our investigation to trajectories determined by piecewise constant control functions. Even if the system is nonlinear, the problem will turn out to be purely algebraic for a symmetric system, and we will end up computing the eigenvalues of a sequence of matrices and their asymptotic behaviour. Our method will mix some classical tools of nonlinear control systems as the use of Lie algebra in the problem of small time attainability, some newer Hamilton–Taylor expansions developed by one of the authors [24,25,26,27], and linear algebra of matrices.
It is known that in order to have some negative second order decrease rate, it is necessary and sufficient, see e.g., [3], that the system satisfies a second order attainability condition, namely in the neighborhood of ˆx the minimum time function T to reach U satisfies an estimate of the form
T(x)≤C|x−ˆx|1/2, | (1.4) |
where T(x)=inf{t≥0:xt∈U,a(⋅)∈L∞(0,+∞;B1)}. The function u we use in stating our problem is generic, but it can well be, for instance, the smooth distance function d(x)=dist(x,T) from a closed set T⊂Rn either convex or with at least a C2 boundary, to give the problem a more metric significance.
In mathematical control theory, controllability to a point is well studied, has a long story, and has a huge literature. Symmetric systems are studied for instance in Chow and Rashevski [7], where the famous sufficient condition using the Lie algebra is derived. For general nonlinear systems Petrov [21,22] introduced the positive basis condition, and Liverowskii [17] extended that result to second order sufficient and necessary conditions. For affine systems, some classical results are, for instance, due to Sussmann [28,29]. Frankowska [10] and Kawski [11] discussed higher order conditions for affine systems at an equilibrium point. Other results can be found in Bianchini and Stefani [4,5,6] and Krastanov [13] for dynamics on manifolds. A summary of the main classical results for the point is contained in the chapter on controllability of control systems in the book by Coron [9], where many additional references can be found. For attainability of sets, we recall Bardi and Falcone [2] for necessary and sufficient first order conditions, and our paper with Bardi and Feleqi [3] for necessary and sufficient second order conditions, see also [23]. For affine systems with drift vanishing on the target, sufficient conditions with different generality can be found in Krastanov and Quincampoix [12,13,14], Marigonda and coauthors [15,18,19].
As notations used in this paper, given a square matrix A, we denote its transposed as tA, and respectively its symmetric part A∗ and its antisymmetric part Ae as
A∗=(A+tA)/2,Ae=(A−tA)/2. |
We also indicate by Mn,m(R) the space of n×m matrices with elements in R, and also Mn(R)≡Mn,n(R). We denote by I the identity matrix (of the appropriate dimension).
In the following, we always assume that u:Rn→R is a C2 function, σ:Rn→Mn,m(RN) is C1, ˆx∈Rn is given, and that ∇uσ(ˆx)=0. In this case a first order decay rate at ˆx for u relative to the trajectories of the system (1.1) is not feasible. One therefore needs to look for higher order decay rates that exploit the nonlinearity of the controlled vector field or that of the level set U. A very general definition to express this is in a quantitative way is the following:
Definition 2.1. We say that a function u∈C2(Rn) has a second order decrease rate v<0 for the system (1.1) at ˆx if there are sequences (a[n](⋅))n≥1 of control functions and tn→0+ such that if x[n](⋅) are the corresponding trajectories in (1.1) with initial point ˆx then
u(x[n]tn)=u(ˆx)+vt2n+t2no(1),as n→+∞. |
Sometimes in the literature one may find the terminology second order variation of u(xt) for v in the definition.
Remark 2.2. It is important to stress the fact that in the previous definition we have a sequence of trajectories and that we are checking each one at a specific time that will change with the parameter. We are therefore building a fictitious discrete trajectory (x[n]tn)n along which we expect the function u to decrease at a determined rate. Our following construction is more specific as the trajectory family parameter is continuous and all trajectories in the family are constructed similarly and consistently.
There is a standard way in the literature where the idea of Definition 2.1 is implemented and applies in the following way. Suppose that we fix a measurable control ˆa(⋅)∈L∞([0,1],B1). We define the following family of control functions parametrized by t>0 (as t→0+)
a[t]s=as/t,s∈[0,t]. |
Computing the corresponding family of trajectories (x[t](⋅))t of (1.1) at the end time t, in order to check Definition 2.1, we hope to find some second order rate v<0 such that
u(xt)≡u(x[t]t)=u(ˆx)+vt2+t2o(1),as t→0+. |
We will usually avoid showing explicitly the parameter of the family. In order to exploit the best performance of the system relative to the function u, in this paper we will adopt the previous construction, although we further limit ourselves to piecewise constant control functions in the interval [0,1] and compute the lowest possible decrease rate among them.
Example 2.3. In order to bridge our approach with the literature, the best known example of what we just discussed is when for controls a1,a2∈B1, we take a control function
as={a1,s∈[0,1/4[,a2,s∈[1/4,1/2[,−a1,s∈[1/2,3/4[,−a2,s∈[3/4,1], |
and consider the family of trajectories (x[t](⋅))t of (1.1) corresponding to the control functions a[t](⋅)≡a⋅/t with the same initial point ˆx. It is then well known that
u(x[t]t)=u(ˆx)+116⟨[σa1,σa2](ˆx),∇u(ˆx)⟩t2+t2o(1),as t→0+, | (2.1) |
where [σa1,σa2]=D(σa2)σa1−D(σa1)σa2 is the Lie bracket of the two vector fields. Therefore u has a second order decrease rate 116⟨[σa1,σa2](ˆx),∇u(ˆx)⟩ when negative.
In order to make things more general, following one of the authors [24,25,27], we introduce the matrix valued function
S:Rn→Mm(R),tS(x)=D(∇uσ)σ(x)=(Hσj∘Hσiu(x))i,j=1,…,m, |
which is the matrix of all second order Lie derivatives of u with respect to the vector fields (σi)i=1,…,m.
Remark 2.4. Notice that S(x) is not a symmetric matrix in general, as one easily gets that
2Se(x)=S(x)−tS(x)=(⟨[σi,σj],∇u(x)⟩)i,j=1,…,m. |
Example 2.5. Suppose that in R3,
σ=(100100) |
and u(x,y,z)=z−x2−y2. At the origin we have ∇uσ(0,0,0)=0 and we compute
tS(0,0,0)=D(−2x,−2y)σ=(−200−2), |
which is symmetric. Notice that in this case there is no sensible Lie bracket of the vector fields since they are constant.
Example 2.6. Again, in R3, suppose that
σ(x,y,z)=(1001y−x) |
and u(x,y,z)=z, then at any point of the z-axis
tS(0,0,z)=D((0,0,1)σ)σ=(01−10), |
which is antisymmetric.
We will return to these examples to comment on the results in the next section. To see the role of the matrix valued function S(x), notice that given two controls a1,a2∈B1 and corresponding vector fields f(x)=σ(x)a1, g(x)=σ(x)a2, we can compute the second order Lie derivative of u with respect to the vector fields f,g as
Hf∘Hgu=⟨f,∇⟨g,∇u⟩⟩=⟨σa1,∇⟨σa2,∇u⟩⟩=⟨D(∇uσ)σa1,a2⟩=⟨Sa2,a1⟩. |
That is indeed a bilinear operator on the controls (a1,a2). Consider, for instance, a constant control function at≡a,|a|≤1 and ⟨σ(ˆx)a,∇u(ˆx)⟩=0, then the standard Taylor estimate gives
u(xt)=u(ˆx)+t22!H(2)σau(ˆx)+t2o(1),as t→0+, |
where we used the notation H(2)σa=Hσa∘Hσa. In particular, H(2)σau(ˆx)=⟨S(ˆx)a,a⟩=⟨S∗(ˆx)a,a⟩ and, if negative, it is twice the second order decrease rate of u for the trajectory of the control system. Thus, the minimum decrease rate among all constant controls |a|≤1 is 1/2 the minimum eigenvalue of the symmetric matrix S∗(ˆx).
As an example of a more general trajectory, consider the following family of control functions parametrized by t>0
a[t]s={a1,if s∈[0,t/2[,a2,if s∈[t/2,t], | (2.2) |
then the trajectories corresponding to the control functions a[t](⋅), all starting at ˆx satisfy (see [25])
u(xt)≡u(x[t]t)=u(ˆx)+12!(t2)2(Hσa1⊞Hσa2)2u(ˆx)+t2o(1),as t→0+ |
where
(Hσa1⊞Hσa2)2u(ˆx):=⟨S(ˆx)a1,a1⟩+⟨S(ˆx)a2,a2⟩+2⟨S(ˆx)a2,a1⟩ |
defines the square of the sum of two Hamiltonians, which is a helpful operator. Indeed, it has been previously proved in [24] that (1.4) for the system (1.1) is equivalent to the following algebraic property of S(ˆx), see e.g., [3,25], we can find a1,a2∈B1 such that
(Hσa1⊞Hσa2)2u(ˆx)<0. | (2.3) |
Notice that the left hand side in (2.3) identifies a quadratic form in R2m as
(Hσa1⊞Hσa2)2u(ˆx)=⟨K2(S(ˆx))(a1a2),(a1a2)⟩,K2(S(ˆx))=(S∗(ˆx)S(ˆx)tS(ˆx)S∗(ˆx)). |
Therefore, if (2.3) holds, u has a second order decrease rate at ˆx given by v=18(Hσa1⊞Hσa2)2u(ˆx).
Remark 2.7. Note that, if f,g:Rn→Rn are vector fields, then (see [25])
(Hf⊞Hg)2u=Tr(t(f+g)D2u(f+g))+⟨[f,g],∇u⟩. |
Therefore (Hf⊞Hg)2 is, as a differential operator, second order degenerate elliptic and has two parts: the second order measures the curvature of u in the direction of the sum of the vector fields, and the first order that compares the direction of the Lie bracket of the vector fields and the normal to the level set of u.
We can more generally consider piecewise constant control functions with an arbitrary number of switches. Namely for any given t>0, a family of k controls ai∈B1,i=1,…,k and real numbers αi∈]0,1], ∑ki=1αi=1, we indicate ti=tαi and define the piecewise constant control function
a[t]s={a1,if s∈[0,t1[,a2,if s∈[t1,t1+t2[,…ak,if s∈[t−tk,t]. | (2.4) |
From the system (1.1) with initial point ˆx, we therefore obtain the corresponding family of trajectories indexed by t, that we indicate as (x[t]s)s∈[0,t]. By the results of one of the authors, see [27] and also [17], such a family of trajectories, satisfies the following Hamilton–Taylor expansion
u(xt)≡u(x[t]t)=u(ˆx)+t22!(Hα1σa1⊞⋯⊞Hαkσak)2u(ˆx)+t2o(1)as t→0+, | (2.5) |
where for k≥2 the coefficient v of the second order term above contains the square of the sum of the k Hamiltonians which is defined as follows:
2!v=(Hα1σa1⊞⋯⊞Hαkσak)2u(ˆx):=(∑ki=1H(2)αiσaiu(ˆx)+2∑ki<j=1Hαiσai∘Hαjσaju(ˆx))=(∑ki=1⟨S(ˆx)(αiai),(αiai)⟩+2∑ki<j=1⟨S(ˆx)(αjaj),(αiai)⟩)=(∑ki=1α2i⟨S(ˆx)ai,ai⟩+2∑ki<j;i,j=1αiαj⟨S(ˆx)aj,ai⟩). |
Also, in this case, v is given by a quadratic form since
v=12Kk(S(ˆx))(α1a1⋮αkak)⋅(α1a1⋮αkak),|a1|,…,|ak|≤1,k∑i=1αi=1, | (2.6) |
where
Kk(S(ˆx))=(S∗(ˆx)S(ˆx)⋯S(ˆx)tS(ˆx)S∗(ˆx)S(ˆx)⋮tS(ˆx)tS(ˆx)S∗(ˆx)⋮tS(ˆx)⋯tS(ˆx)S∗(ˆx))∈Mkm(R), |
as it is easily seen by an induction argument. When negative, v is the decrease rate of u with respect to the family of control functions in (2.4). We have therefore defined a sequence of matrices (Kk(S(ˆx)))k where the k−th element appears in the Hamilton–Taylor expansion of trajectories of the system constructed as above and having k−1 switches.
The main goal of the paper is now to compute by induction on k the (nonpositive, according to the properties of the matrix S(ˆx)) minimum of each quadratic form (2.6) in order to obtain the lowest possible decrease rate of the trajectories of the system (1.1) among all piecewise constant controls and the indicated construction. We also want to understand if the minimum is reached for a specific number of switches among the vector fields or in the limit. We could not find this type of analysis in the literature, but we think it could also help identify time optimal trajectories for the system for appropriate choices of the function u.
We now anticipate the main result of the paper. Its proof can be found in the next section at the end of Subsection 3.2.
Theorem 2.8. Let u∈C2(Rn) and σ∈C1(Rn;Mn,m(R)). Assume that at ˆx∈Rn we have ∇uσ(ˆx)=0 and let S(ˆx)=(Hσi∘Hσju(ˆx))i,j=1,…,r. Then u has at ˆx a (negative) second order decrease rate for the control system (1.1) if and only if S(ˆx) is not symmetric and positive semidefinite. In this case, the highest rate of decrease among all piecewise constant controls is
12infk≥1λ(k)1k(=infk,ai,αi{12(Hα1σa1⊞⋯⊞Hαkσak)2u(ˆx)}), | (2.7) |
where λ(k)1 is the minimal eigenvalue of Kk(S(ˆx)).
The right hand side of Eq (2.7) shows how the highest decrease rate is computed from a fully nonlinear elliptic operator that plays the role of the Hamilton-Jacobi-Bellman one in the case of higher order conditions.
Notice that (2.6) also contains information about the controls and times that we can use in the system to achieve the decay rates appearing in Theorem 2.8. In particular, if w is an eigenvector of Kk(tS(ˆx)) with eigenvalue λ(k)1 and |w|=1/√k, then
v=12λ(k)1k. |
This is the case when w=(αiai)i=1,…,k with |ai|=1, αi=1/k for all i=1,…,k; see Propositions 3.1 and 3.3 on how controls ai are choosen and the proof of the theorem.
As main examples of the applicability of Theorem 2.8, where calculations can be performed explicitly, we can consider in the next section, three cases when S(ˆx) is antisymmetric, symmetric or the sum of an antisymmetric and a scalar matrix. This is done below in Theorems 3.14, 3.17, and 3.20, also with examples of control systems where this happens. Furthermore, in addition to showing the calculation of the optimal decreasing rate in practice, we also show how to easily construct the controls that give the optimal rate λk/(2k) above for a given integer k. This is done in the Propositions 3.3 and 3.1.
Let A be in Mr(R). Let us pose K1(A)=A∗, and
Kn(A)=(A∗A⋯AtAA∗A⋮tAtAA∗⋮tA⋯tAA∗)=(AKn−1(A)⋮AtA⋯tAA∗)∈Mnr(R) |
if n≥2. Let v=t(v1,…,vn)∈Rnr.
We now study the properties of the eigenvalues and the eigenvectors of Kn(A). We notice that when we will discuss Theorem 2.8 we will choose A=tS(ˆx).
Let λ be nonzero real. Since Ae is alternating, ±λ is not eigenvalue of Ae hence, both λI+Ae and λI−Ae are invertible. We define Γλ=(λI+Ae)−1(λI−Ae). Note that in case Ae is non singular, then we may drop the assumption λ≠0: in this case Γ0=(Ae)−1(−Ae)=−I. We always have detΓλ=1, if λ≠0, see also Remark 3.5.
Proposition 3.1. Let v=t(v1,…,vn)∈Rnr be an eigenvector of Kn(A) with vi∈Rr for i=1,…,n so that
Kn(A)v=λv |
for some λ∈R. Then the following holds true:
(ⅰ) Ae(vi+vi+1)=λ(vi−vi+1), i=1,…,n−1. Assume λ≠0 if Ae is singular.
(ⅱ) Γλ is an isometry of Rr.
(ⅲ) vi+1=Γλvi, i=1,…,n−1.
(ⅳ) |vi|=|vj| for each i, j.
Proof. Let us prove (ⅰ). Consider the i-th and i+1-th equations of the linear system Kn(A)v=λv in block form:
tAv1+tAv2+⋯+A∗vi+Avi+1+⋯+Avn=λvi,tAv1+tAv2+⋯+tAvi+A∗vi+1+⋯+Avn=λvi+1. |
Subtracting gives
Ae(vi+vi+1)=(A∗−tA)vi+(A−A∗)vi+1=λ(vi−vi+1). | (3.1) |
(ⅱ) Since Γλ=(λI+Ae)−1(λI−Ae) and Ae is alternating, we get ⟨Γλv,w⟩ = ⟨v,Γ−1λw⟩ for each v, w∈Rr, so Γλ is an isometry. Unless Ae is non singular and λ=0 (in which case Γ0=−I), −1 is not an eigenvalue of Γλ.
(ⅲ) From Ae(vi+vi+1)=λ(vi−vi+1) it follows (λI+Ae)vi+1=(λI−Ae)vi, hence
vi+1=(λI+Ae)−1(λI−Ae)vi=Γλvi. |
Finally (ⅳ) follows from (ⅱ) and (ⅲ).
Remark 3.2. We note that from vi+1=Γλvi it follows that vk=Γk−1λv1 and then
v=(v1v2⋮vn)=(v1Γλv1⋮Γn−1λv1) |
is the form on an eigenvector of Kn(A). Moreover
12⟨Kn(A)v,v⟩=n2|v1|2λ. |
The case A symmetric will be specifically treated in the next section. Otherwise the following necessary condition for eigenvalues holds.
Proposition 3.3. Suppose that the antisymmetric part Ae of A is not singular. Then, λ is an eigenvalue of Kn(A) if and only if
det(AΓλn−tA)=0. |
Moreover
v1∈ker (AΓλn−tA)(I−Γλ)−1, |
if v=t(v1,…,vn)∈Rnr is an eigenvector of Kn(A).
Proof. We begin by observing that since Ae is not singular, 1 is not an eigenvalue of Γλ, and so I−Γλ is invertible.
Let v=t(v1,…,vn)∈Rnr, λ∈R. By the previous proposition, the linear system Kn(A)v=λv is equivalent to
A∗v1+Av2+⋯+Avn=λv1,vi+1=Γλvi,i=1,…,n−1. |
Therefore v≠0 is an eigenvector of Kn(A) relative to λ if and only if
A∗v1+AΓλv1+⋯+AΓn−1λv1=λv1,A(v1+Γλv1+⋯+Γλn−1v1)=(λI+Ae)v1,A(I−Γλn)(I−Γλ)−1v1=(λI+Ae)v1=(λI+Ae)(I−Γλ)(I−Γλ)−1v1. |
Setting w=(I−Γλ)−1v1, we get
A(I−Γλn)w=(λI+Ae)w−(λI−Ae)w=2Aew,tAw−AΓλnw=0. | (3.2) |
Since w≠0, λ is an eigenvalue of Kn(A) if and only if det(AΓλn−tA)=0.
Remark 3.4. Observe that if Ae is non singular, the determinant condition of Proposition 3.3 allows us to detect when λ=0 is an eigenvalue of Kn(A). This happens if and only if det(A(−1)n−tA)=0. If n is even, λ=0 is not eigenvalue of Kn(A), since we are assuming detAe≠0. If instead n is odd, then λ=0 is an eigenvalue of Kn(A) if and only if detA∗=0. In this case
kerKn(A)={v=(v1,−v1,…,v1,−v1,v1)∣A∗v1=0}≅kerA∗≠{0}. |
In general, one can easily show that for n even, λ=0 is an eigenvalue of Kn(A) if and only if Ae is singular, while for n odd, λ=0 is an eigenvalue of Kn(A) either if Ae is non singular and A∗ is singular, or if Ae is singular.
Remark 3.5. Let a≠0 and
B=(0a−a0)∈M2(R). |
Then
Γλ=(λI+B)−1(λI−B)=(λ2−a2λ2+a2−2aλλ2+a22aλλ2+a2λ2−a2λ2+a2)=(cosϑ−sinϑsinϑcosϑ) |
is the rotation of angle ϑ, where λa=cot(ϑ/2), 0<ϑ<2π. Observe that posing (e1,e2) the canonical basis of R2, we have B(e1+ie2)=ia(e1+ie2), B(e1−ie2)=−ia(e1−ie2). In addition, if λ≠0,
(I+Γλ)−1(I−Γλ)=1λB. |
Let A∈Mr(R) be antisymmetric. Then there exists P orthogonal such that P−1AP=R,
R=(B10⋯00B20⋮00Bs⋮0⋯00k)∈Mr(R) | (3.3) |
with
Bj=(0aj−aj0)∈M2(R) | (3.4) |
aj>0 for j=1,…,s, and we can assume a1≥a2≥⋯≥as>0. Then the eigenvalues of A are 0 with multiplicity k, k≥0 and ia1,−ia1,…,ias,−ias. Let (v1,w1,v2,w2,…,vs,ws,z1,…,zk) be the orthonormal basis of Rr given by the columns of P. If λ∈R, λ≠0, then
(λI+Bj)−1(λI−Bj)=(cosϑj−sinϑjsinϑjcosϑj) |
is the rotation of angle ϑj in the plane ⟨vi,wi⟩, where aj=λtan(ϑj/2), −π<ϑj<π, ϑj≠0. Then
(λI+R)−1(λI−R)=(cosϑ1−sinϑ1sinϑ1cosϑ10⋯00cosϑ2−sinϑ2sinϑ2cosϑ20⋮00cosϑs−sinϑssinϑscosϑs⋮0⋯01k) |
the product of s rotations (this orthogonal matrix does not have -1 as its eigenvalue), in particular
det(λI+R)−1(λI−R)=1. |
The vectors vj, wj are also constructed from eigenvectors relative to iaj and −iaj, while the zk constitutes an orthonormal basis of the core of R. Note that every isometry X of Rr of this form can be achieved from an appropriate antisymmetrix matrix R: R=(I+X)−1(I−X).
We can characterize the cases when the matrix K2(A) has a negative eigenvalue.
Lemma 3.6. Let A be in Mr(R). Then K2(A) is positive semidefinite if and only if A is symmetric and positive semidefinite.
Proof. If A is symmetric and positive semidefinite, then
⟨K2(A)(v1v2),(v1v2)⟩=⟨A(v1+v2),(v1+v2)⟩ | (3.5) |
and therefore K2(A) is positive semidefinite as well. We next prove the converse. Notice that choosing v1=−v2∈Rr also shows that K2(A) always has 0 as an eigenvalue.
Suppose first that A is symmetric. By (3.5), if K2(A) is positive semidefinite and we choose v1=v2∈Rr, then we get 0≤4⟨Av1,v1⟩, for all v1∈Rr and A is also positive semidefinite.
Suppose now that A is not symmetric and we show that K2(A) must have a negative eigenvalue. In particular, A is not the null matrix. Consider thus the positive semidefinite matrix tAA; it will have at least one positive eigenvalue λ2, with λ>0, with corresponding unit eigenvector v2. Thus,
tAAv2=λ2v2 |
and then
|Av2|2=⟨Av2,Av2⟩=⟨tAAv2,v2⟩=λ2⟨v2,v2⟩=λ2, |
so that λ=|Av2|>0. Just notice that if ˉv is an eigenvector of tAA with null eigenvalue, then the same argument shows that Aˉv=0. Let
v1=−Av2λ, |
so that |v1|=1. We can now obtain
tAv1=−λv2,⟨Av2,v1⟩=−λ,⟨Av1,v1⟩=−λ⟨v1,v2⟩=⟨Av2,v2⟩. |
Thus we conclude that
⟨K2(A)(v1v2),(v1v2)⟩=−2λ(1+⟨v1,v2⟩) |
and we reach our conclusion provided v1≠−v2. If instead v1=−v2, it then follows
Av1=λv1,tAv1=λv1. |
If this happens for all eigenvectors of tAA with positive eigenvalues, and we consider an orthonormal basis of eigenvectors of tAA, then this is also a family of eigenvectors for A which can then be diagonalised by an orthogonal matrix, and it is thus symmetric, which was supposed not to be the case.
Corollary 3.7. Let A be in Mr(R). Then for all n≥2, Kn(A) is positive semidefinite if and only if A is symmetric and positive semidefinite.
Proof. If n=2, then the result is the content of Lemma 3.6. Let us assume n≥3. If Kn(A) is positive semidefinite, then K2(A) is positive semidefinite, hence A is symmetric positive semidefinite. Assume A symmetric positive semidefinite. Then
⟨Kn(A)(v1v2⋮vn),(v1v2⋮vn)⟩=⟨A(v1+⋯+vn),(v1+⋯+vn)⟩ | (3.6) |
hence Kn(A) is positive semidefinite.
Corollary 3.8. Let A be in Mr(R). Then, for all n≥2, Kn(A) is never positive definite.
Proof. If Kn(A) is positive semidefinite, then A is symmetric. Hence Ae=0 is singular, and therefore kerKn(A) is not zero.
Given A∈Mr(R) and n≥1, we consider the quadratic form Qn(A):Rnr→R,
Qn(A): v↦ ⟨Kn(A)v,v⟩ =tvKn(A)v,v∈Rnr. | (3.7) |
We want to compute the minimum of Qn(A) on different compact subsets D⊂Rnr.
Remark 3.9. It is well known that on Dn,2={v∈Rnr:|v|=1}, then minDQn(A) is the minimal eigenvalue λ(n)1 of Kn(A). If instead D=B1, then minB1Qn(A)=0 if Kn(A) is positive semidefinite, while again minB1Qn(A)=λ(n)1 if Kn(A) is not positive semidefinite.
Remark 3.10. Suppose that we have a sequence of domains Dn⊂Rnr, n≥2, with the property that if v=(v1,…,vn−1)∈Dn−1 then ˆv=(v1,…,vn−1,0)∈Dn so that we can identify Dn−1 with a subset of Dn. Since ⟨Kn−1(A)v,v⟩=⟨Kn(A)ˆv,ˆv⟩, it is then clear that minDnQn(A)≤minDn−1Qn−1(A). Therefore the sequence (minDQn(A))n is nonincreasing and infnminDnQn(A) is either attained at some ˉn if the sequence is constant for n≥ˉn, or it is attained asymptotically as n→+∞.
From now on we will suppose that A is not symmetric and positive semidefinite so that λ(n)1, the minimal eigenvalue of Kn(A), is negative for n≥2, as we proved in the previous section. We turn to a more interesting case for us, which is, n≥1,
Dn,∞={v=(v1,…,vn)∈Rnr:maxi|vi|=1}. |
The property of Remark 3.10 holds true for the sequence (Dn,∞)n. Notice that if v∈Dn,∞ then |v|≤√nmaxi|vi|=√n and that the equality is reached only if |vi|=1, for all i=1,…,n.
Lemma 3.11. If A is not symmetric and positive semidefinite, then
minDn,∞Qn(A)=nλ(n)1. |
Proof. If n=1, the thesis is obvious since D1,2=D1,∞. For n≥2 and v∈Dn,∞, we also have that v/|v|∈Dn,2. Therefore, as λ(n)1<0,
Qn(A)(v)=|v|2Qn(A)(v|v|)≥|v|2λ(n)1≥nλ(n)1 |
and by Proposition 3.1(i), the equalities are actually reached for v=(vi)i being an eigenvector for λ(n)1 with |vi|=1 for all i=1,…,n. Thus minDn,∞Qn(A)=nλ(n)1.
We now consider
Dn,1={v=t(v1,…,vn)∈Rnr:n∑i=1|vi|=1}. |
Remark 3.10 applies to this case as well, and D1,1=D1,2.
Proposition 3.12. Suppose that A is not symmetric and positive semidefinite, so that Kn(A) has a negative eigenvalue, for n≥2. Then
minDn,1Qn(A)=min{λ(1)1,λ(2)12…,λ(n)1n}, |
where λ(i)1 is the minimal eigenvalue of Ki(A).
Proof. We proceed by an induction argument on n. If n=1 then K1(A)=A∗ and D1,1=∂B1=D1,2. Therefore, minD1,1Q1(A)=λ(1)1 is the minimal eigenvalue of A∗.
For n≥2, we use the Lagrange multipliers necessary condition. Let v=(v1,…,vn)∈Dn,1 be a minimum point of Qn(A). If some vi=0, then we can reduce the problem to that of the minimum of Qm(A) with m<n (and v is at some corner of the domain), which we can assume is solved by the inductive hypothesis. So we assume vi≠0 for all i in order to get necessary conditions genuinely for the given n. The Lagrangian is
L(v,λ)=12⟨Kn(A)v,v⟩−λ(n∑i=1|vi|−1), |
so that the Lagrange necessary system is as follows:
Kn(A)(v1⋮vn)=λ(v1/|v1|⋮vn/|vn|),v∈Dn,1. | (3.8) |
Notice that then
Qn(A)(v)=⟨Kn(A)v,v⟩=λ(n∑i=1|vi|)=λ, |
therefore, the Lagrange multiplier gives the possible minimum. If λ=0, then
Qn(A)(v) = ⟨Kn(A)v,v⟩ =0 |
which is the minimum only if Kn(A) is positive semidefinite, i.e., if A is symmetric and positive semidefinite by Corollary 3.7, which is not the case by assumption. Hence λ≠0. Let us define ˆvi=vi|vi|, then vi=γiˆvi, with γi=|vi|>0, γ1+⋯+γn=1. We will prove that all γi are equal (to 1/n), or that we can find anyway
v′=(v′1⋮v′n) |
with the v′i all having the same norm equal to 1/n, such that (3.8) is also satisfied by v′ with the same λ. Consider the i-th and (i+1)-th equations in (3.8), i=1,…,n−1,
tAv1+tAv2+⋯+A∗vi+Avi+1+⋯+Avn=λˆvi,tAv1+tAv2+⋯+tAvi+A∗vi+1+⋯+Avn=λˆvi+1. |
Subtracting this gives
Ae(vi+vi+1)=λ(ˆvi−ˆvi+1). |
But now Ae is alternating, so
⟨vi+vi+1,λ(ˆvi−ˆvi+1)⟩ = ⟨vi+vi+1,Ae(vi+vi+1)⟩ =0 |
from which it follows
⟨vi+vi+1,ˆvi−ˆvi+1⟩=0 |
being λ≠0. Recall that vi=γiˆvi, vi+1=γi+1ˆvi+1, so
(γi−γi+1)(1−⟨ˆvi,ˆvi+1⟩)=0. |
We have two possible cases. Suppose first ˆvi≠ˆvi+1. Then γi=γi+1 (and thus |vi|=|vi+1|=γi).
Suppose next ˆvi=ˆvi+1. The equation
Ae(vi+vi+1)=λ(ˆvi−ˆvi+1) |
becomes (γi+γi+1)Aeˆvi=0 and therefore Aeˆvi=0. Rewrite for convenience
Kn(A)=Kn(A∗+Ae)=(A∗A∗+Ae⋯A∗+AeA∗−AeA∗A∗+Ae⋮A∗−AeA∗−AeA∗⋮A∗−Ae⋯A∗−AeA∗)∈Mnr(R). |
It is possible that some other ˆvi+2,ˆvi+3,…,ˆvj is also equal to ˆvi. We then consider the string
ˆvi=ˆvi+1=ˆvi+2=ˆvi+3=…=ˆvr |
where 1≤i<r≤n. Let us start supposing either i≥2 or r≤n−1, and therefore either ˆvi−1≠ˆvi or ˆvr≠ˆvr+1. Thus either γi−1=γi or γr=γr+1 by the discussion above. Recall that vs=γsˆvs=γsˆvi and hence Aevs=0 for i≤s≤r. The j-th equation in (3.8) results in
A∗(v1+⋯+vn)+Ae(vj+1+⋯+vn)−Ae(v1+⋯+vj−1)=λˆvj. |
Let δ1,…,δn in R, positive, with γi+⋯+γr=δi+⋯+δr and δj=γj if j<i or j>r. We consider the vector
v′=(v′1⋮v′n)∈Mn(Rr) |
where we put δsˆvs=v′s for 1≤s≤n, respectively. Then
v1+⋯+vn=v′1+⋯+v′n |
and
A∗(v1+⋯+vn)=A∗(v′1+⋯+v′n). |
But also if j<n
Ae(vj+1+⋯+vn)=Ae(v′j+1+⋯+v′n) |
and if j>1
Ae(v1+⋯+vj−1)=Ae(v′1+⋯+v′j−1), |
since at places s where we varied the coefficient, we have Aevs=0. We therefore have
Kn(A∗+Ae)(v′1⋮v′n)=λ(v1/|v1|⋮vn/|vn|)=λ(v′1/|v′1|⋮v′n/|v′n|),⟨Kn(A∗+Ae)v′,v′⟩ = ⟨λ(ˆv′1⋮ˆv′n),(δ1ˆv′1⋮δnˆv′n)⟩ =λ(δ1+⋯+δn)=λ. |
Since we can modify at will the δs, s=i or s=r, this contradicts one of the following two
γi−1=γi,γr=γr+1, |
unless we are really in the extreme case, i.e., i=1, r=n, that is, all ˆvi are equal.
Let us then put ourselves in this case. We then have
vi=γiˆv1,Aevi=0 |
for every i=1,…,n, and again, as above, we can modify all vi by choosing δs=1/n for every 1≤s≤n. Then
v′=1n(ˆv1⋮ˆv1)∈Rrn, |
and this choice satisfied
Kn(A)v′=nλv′,Qn(A)(v′)=⟨Kn(A)v′,v′⟩ =λ |
therefore, v′ is an eigenvector of Kn(A) with nλ as an eigenvalue and Qn(A)(v′)=λ. Thus the best choice to reach a minimum of Qn(A) is for the v′ eigenvector of Kn(A) with λ(n)1=nλ as eigenvalue, and finally Qn(A)(v′)=λ(n)1n as we intended.
Then, by the induction assumption, the minimum is
minDn,1Qn(A)=min{minDn,1Qn−1(A),λ(n)1n}=min{λ(1)11,…,λ(n)1n} | (3.9) |
as we wanted.
We are now in the position to prove Theorem 2.8.
Proof of Theorem 2.8. As we saw in Section 2, (2.6) is the expression of the second order coefficient for the family of trajectories corresponding to the family of controls defined in (2.4), where n is the number of controls being used and α1,…,αn are the percentage of the time being spent on each control respectively. We will use the above with A=tS(ˆx). We need to compute for each n the minimum of the quadratic form (1/2)Qn(A) on the following domain
ˆDn={v=t(α1v1,…,αnvn)∈Rnr:|vi|≤1,αi≥0,∑iαi=1}. |
We will show that minˆDnQn(A)=minDn,1Qn(A) reaching our thesis. This fact will be a consequence of Proposition 3.12 and the proof that
ˆDn={v=t(v1,…,vn)∈Rnr:n∑i=1|vi|≤1}. |
Indeed, on one hand, for v=t(α1v1,…,αnvn)∈ˆDn we have that
n∑i=1|αivi|≤n∑i=1αi=1. | (3.10) |
On the other hand let v=t(v1,…,vn) and ∑ni=1|vi|≤1. We may clearly suppose v≠0 and thus S=∑ni=1|vi|>0. Then we put
αi=|vi|S,ˆvi={viαi,if vi≠0,vi,if vi=0. |
Therefore ∑ni=1αi=1, αiˆvi=vi, and either ˆvi=0 or |ˆvi|=S≤1 so that v∈ˆDn.
Also notice that by choosing t(v1,…,vn) an eigenvector of Kn(A) with eigenvalue λ(n)1, |vi|=1 and αi=1/n for all i, so that t(α1v1,…,αnvn)∈ˆDn, we obtain
Qn(A)(v)=1n2Kn(A)(v1⋮vn)⋅(v1⋮vn)=1n2λ(n)1n∑i=1|vi|2=λ(n)1n. |
The first example deals with the case when A is antisymmetric.
Proposition 3.13. Let A∈Mr(R) be antisymmetric, A≠0. Then the extremal eigenvalues of Kn(A), n≥2, are
±α1cot(π2n) |
where α1 is the biggest modulus of the eigenvalues of A.
Proof. Let A be antisymmetric, n≥2. Then
Kn(A)=(0A⋯A−A0A⋮−A−A0⋮−A⋯−A0)∈Mnr(R). |
We will compute all its eigenvalues. A convenient way to denote Kn(A) is obtained by introducing the antisymmetric matrix
Tn=(01⋯1−101⋮−1−10⋮−1⋯−10)∈Mn(R). |
Then
Kn(A)=Tn⊗A. |
If λ1,…,λn are the eigenvalues of Tn and μ1,…μr are those of A, then the eigenvalues of Tn⊗A are all the products λiμj, i=1,…,n, j=1,…,r.
Since A is antisymmetric, there exists P orthogonal such that P−1AP=R, as in (3.3) and (3.4). Then the eigenvalues of A are 0 with multiplicity k, k≥0 and iα1,−iα1,…,iαs,−iαs, with αi>0 for i=1,…,s, and we can assume α1≥α2≥⋯≥αs>0.
We come to the matrix Tn. Let v=t(x1,…,xn)∈Rn be an eigenvector of Tn, so that Tnv=λv for some λ∈iR. Consider the i-th and (i+1)-th equations of Tnv=λv:
−x1−x2+⋯−xi−1+xi+1+⋯+xn=λxi,−x1−x2+⋯−xi+xi+2+⋯+xn=λxi+1. |
Subtracting gives
xi+xi+1=λ(xi−xi+1). | (3.11) |
Hence xi+1=γλxi, where γλ=λ−1λ+1. It follows that xk=γk−1λx1 and then
v=(x1x2⋮xn)=(x1γλx1⋮γn−1λx1) |
is the form on an eigenvector of Tn. Therefore, v≠0 is an eigenvector of Tn relative to λ if and only if
x2+⋯+xn=λx1,γλx1+⋯+γn−1λx1=λx1,x1+γλx1+⋯+γn−1λx1=x1+λx1,(1+γλ+⋯+γn−1λ)x1=(1+λ)x1. |
Note that γλ≠1, hence v≠0 is an eigenvector of Tn relative to λ if and only if
(γλ−1)(1+γλ+⋯+γn−1λ)x1=(γλ−1)(1+λ)x1,(γnλ−1)x1=−2x1,(γnλ+1)x1=0. |
It follows that λ is an eigenvalue of Tn if and only if γnλ=−1, i.e., γλ=eiϑ, ϑ=πn+k2πn, k=0,…,n−1. From γλ=λ−1λ+1, we get
λ=1+γλ1−γλ=1+eiϑ1−eiϑ=icotϑ2. |
The eigenvalues of Tn are thus
λ=icot(π2n+kπn),k=0,…,n−1. |
The maximum is cot(π2n), the minimum is
cot(π2n+(n−1)πn)=−cot(π2n). |
Recalling that the biggest modulus of the eigenvalues of A is α1, α1>0, the maximum eigenvalue of Kn(A) is then
α1cot(π2n)(∼α12πn,as n→+∞) |
and the minimum is
−α1cot(π2n). |
The consequence for the control problem is as follows: This case happens when the level set of the function u is flat.
Theorem 3.14. Let A=tS(ˆx)∈Mr(R) be antisymmetric, A≠0. Then the highest decrease rate of u at ˆx for the control system is −α1/π, where α1 is the biggest modulus of the eigenvalues of A. Moreover as n→+∞ the error for using the best rate with n controls is of the order |λ(n)1/n+α1/π|∼π/(12n2).
Proof. Since the sequence
an=−12ncot(π2n) |
is strictly decreasing; by Theorem 2.8, we just need to compute limn→+∞anα1=−α1/π. The highest rate is attained in the limit as n→+∞.
Example 3.15. An instance of a control system where an antisymmetric matrix appears is the well-known system of the generators of the Heisenberg group. In this case (1.1) is in dimension 3 and has the following data
σ(x,y,z)=(1001y−x). |
We choose u(x,y,z)=z and want to find a decrease rate of u with respect to the system at a point (0,0,z). Notice that ∇uσ(0,0,z)=(0,0) so we can expect a second order decrease rate. As in Section 2, we define
A≡tS(0,0,z)=D(∇uσ)σ(0,0,z)=(01−10), |
whose eigenvalues are ±i. Of course, K1(A)=A∗=0, so we look at
K2(A)=(000100−100−1001000). |
This matrix has λ(2)1=−1 and as an eigenvector v=(0,1,1,0) (the space of eigenvectors has dimension 2). Therefore if in an interval [0,t] with t small we use controls (0,1) and (1,0) in consecutive subintervals of equal length we can expect a second order rate
12λ(2)22=−14. |
This is already significant since from the classical formula (2.1), the best rate from a Lie bracket is easily computed to be −1/8. Therefore a Lie bracket never gives the best rate of decrease.
We could do better by exploiting Kn(A) with higher n and a control function with n−1 switches instead, because the expected optimal second order rate is −1/π as we proved above. Indeed computing K3(A) we find the minimal eigenvalue −√3, which corresponds to a rate −1/(2√3), with an eigenvector for instance (−1/2,√3/2,1/2,√3/2,1,0), which gives us the three controls to use in order to achieve that rate for sufficiently small time.
Next we consider the case when A∈Mr(R) is symmetric.
Proposition 3.16. Let A∈Mr(R) be symmetric, A≠0. Then 0 is always an eigenvalue of Kn(A) for n≥2, the maximum eigenvalue of Kn(A) is max{0,nα1} and the minimum is min{0,nαr}, where α1,αr are the maximal and minimal eigenvalues of A, respectively.
Proof. Since A is symmetric,
Kn(A)=(AA⋯AAAA⋮AAA⋮A⋯AA)∈Mnr(R). |
In this case, to denote Kn(A), we consider the symmetric matrix
Rn=(11⋯1111⋮111⋮1⋯11)∈Mn(R). |
Then
Kn(A)=Rn⊗A. |
If λ1,…,λn are the eigenvalues of Rn and μ1,…μr are those of A, then the eigenvalues of Rn⊗A are all the products λiμj, i=1,…,n, j=1,…,r.
Let α1,…,αr be the eigenvalues of A, with α1≥α2≥⋯≥αr. The eigenvalues of Rn are 0, with multiplicity n−1, and n. Therefore the maximum eigenvalue of Kn(A) is max{0,nα1} and the minimum is min{0,nαr}.
The consequence of the previous proposition on the control system (1.1) is the following: This case happens when the optimal decrease rate is due to the curvature of the level set of u.
Theorem 3.17. Let A=tS(ˆx)∈Mr(R) be symmetric. If A is not positive semidefinite, then u at ˆx has a (negative) decrease rate for the control system; the highest decrease rate is (1/2)αr, where αr is the minimal eigenvalue of A.
Proof. The sequence min{0,nαr}/2n=αr/2 is constant since αr<0 by the assumption.
Example 3.18. The case of a symmetric matrix for the control system comes from Example 2.5, where we found
A=tS(0,0,0)=(−200−2). |
Then the optimal second order rate among piecewise constant controls is −1, and it is constant on the number of switches. Any control will lead to the optimal rate in this case.
Another special case that we can explicitly deal with, is when A=sI+Ae, Ae is antisymmetric. We omit the details.
Proposition 3.19. Let A be such that A∗=sI is scalar, s≠0, and Ae≠0. Let α1 be the biggest modulus of the eigenvalues of Ae, θ∈(0,π/2) be such that cotθ=|s|α1. Then the extremal eigenvalues of Kn(A), n≥2 are
λmax=α1cot(θn)>ns,λmin=α1cot(θn−πn)<0,if s>0, |
λmax=−α1cot(θn−πn)>0,λmin=−α1cot(θn)<ns,if s<0. |
When we apply the previous proposition to control systems, if s<0 then the curvature of the level set helps u to decrease, while the opposite happens if s>0.
Theorem 3.20. Let A=tS(ˆx)∈Mr(R) be such that A∗=sI is scalar, s≠0, and Ae≠0. Let α1 be the biggest modulus of the eigenvalues of Ae, θ∈(0,π/2) be such that cotθ=|s|α1. Then the highest decrease rate of u at ˆx for the control system is
α12(θ−π)<0,if s>0, |
−α12θ<s/2,if s<0. |
Example 3.21. In R3 we consider the control system where
σ(x,y,z)=(1001y−x) |
and u(x,y,z)=2z−x2−y2. Here we can compute at the points of the z−axis ∇uσ(0,0,z)=(0,0),
A=tS(0,0,z)=(−22−2−2)=−2I+(02−20). |
Then α1=2,s=−2,θ=π4. We can also compute 12λ(1)1=−1 and
λ(n)12n=−2cot(π4n)2n=−cot(π4n)n, |
which is negative and decreasing and whose limit is −4/π<−1. For n=2 we have λ(2)1/4=−cot(π/8)/2 which is smaller in modulus but close to −4/π.
We considered the problem of optimising the second-order decay rate of a function at a point with respect to the trajectories of a symmetric control system. We defined the second-order decrease rate by fixing any piecewise constant control function in a reference interval. We have demonstrated a formula for the minimum rate of decrease as the infimum of the rates when we allow at most k−1 switches in the control function. The k controls used are given by eigenvectors (of appropriate norm) of block matrices constructed recursively from the matrix of the second order Lie derivatives of the function. We have provided a way to explicitly calculate the corresponding optimal control functions through linear algebra methods. We performed explicit full calculations in three cases.
Mauro Costantini: Conceptualization, Formal Analysis, Investigation, Methodology, Validation, Writing – original draft, review & editing; Pierpaolo Soravia: Conceptualization, Formal Analysis, Investigation, Methodology, Validation, Writing – original draft, review & editing. Both authors have read and approved the final version of the manuscript for publication.
The second author was supported in part by University of Padova research grant DOR2192733.
Prof. Pierpaolo Soravia is a Guest Editor of special issue "Mathematical Control of Nonlinear Systems and its Applications" for AIMS Mathematics. Prof. Soravia was not involved in the editorial review and the decision to publish this article. The authors declare no other conflicts of interest regarding this paper.
[1] |
Z. Artstein, Stabilization with relaxed controls, Nonlinear Anal., 7 (1983), 1163–1173. https://doi.org/10.1016/0362-546X(83)90049-4 doi: 10.1016/0362-546X(83)90049-4
![]() |
[2] |
M. Bardi, M. Falcone, An approximation scheme for the minimum time function, SIAM J. Control Optim., 28 (1990), 950–965. https://doi.org/10.1137/0328053 doi: 10.1137/0328053
![]() |
[3] |
M. Bardi, E. Feleqi, P. Soravia, Regularity of the minimum time and of viscosity solutions of degenerate eikonal equations via generalized Lie brackets, Set-Valued Var. Anal., 29 (2021), 83–108. https://doi.org/10.1007/s11228-020-00539-z doi: 10.1007/s11228-020-00539-z
![]() |
[4] | R. M. Bianchini, G. Stefani, Local controllability for analytic families of vector fields, Rapporto No. 19, Istituto Matematico U. Dini, Firenze, Italy, 1981/82. |
[5] | R. M. Bianchini, G. Stefani, Stati localmente controllabili, Rend. Sem. Mat. Univ. Pol. Torino, 42 (1984), 15–23. |
[6] | R. M. Bianchini, G. Stefani, Time-optimal problem and time-optimal map, Rend. Sem. Mat. Univ. Pol. Torino, 48 (1990), 401–429. |
[7] |
W. L. Chow, Über systeme von linearen partiellen differentialgleichungen erster ordnung, Math. Ann., 117 (1940), 98–105. https://doi.org/10.1007/BF01450011 doi: 10.1007/BF01450011
![]() |
[8] |
F. H. Clarke, Y. S. Ledyaev, L. Rifford, R. J. Stern, Feedback stabilization and Lyapunov functions, SIAM J. Control Optim., 39 (2000), 25–48. https://doi.org/10.1137/S0363012999352297 doi: 10.1137/S0363012999352297
![]() |
[9] | J. M. Coron, Control and nonlinearity, Mathematical Surveys and Monographs, Vol. 136, American Mathematical Society, 2007. https://doi.org/10.1090/surv/136 |
[10] |
H. Frankowska, An open mapping principle for set-valued maps, J. Math. Anal. Appl., 127 (1987), 172–180. https://doi.org/10.1016/0022-247X(87)90149-1 doi: 10.1016/0022-247X(87)90149-1
![]() |
[11] | M. Kawski, High-order small-time local controllability, In: H. J. Sussmann, Nonlinear controllability and optimal control, New York: Routledge, 133 (1990), 431–467. |
[12] |
M. Krastanov, M. Quincampoix, Local small time controllability and attainability of a set for nonlinear control system, ESAIM: Control Optim. Calc. Var., 6 (2001), 499–516. https://doi.org/10.1051/cocv:2001120 doi: 10.1051/cocv:2001120
![]() |
[13] |
M. Krastanov, A sufficient condition for small-time local controllability, SIAM J. Control Optim., 48 (2009), 2296–2322. https://doi.org/10.1137/070707117 doi: 10.1137/070707117
![]() |
[14] | M. Krastanov, High-order variations and small-time local attainability, Control Cybern., 38 (2009), 1411–1427. |
[15] |
T. T. Le Thuy, A. Marigonda, Small-time local attainability for a class of control systems with state constraints, ESAIM: Control Optim. Calc. Var., 23 (2017), 1003–1021. https://doi.org/10.1051/cocv/2016022 doi: 10.1051/cocv/2016022
![]() |
[16] | A. Lyapunov, Problème général de la stabilité du mouvement, Ann. Fac. Sci. Univ. Toulouse, 9 (1907), 203–474. |
[17] | A. A. Liverovskii, A Hölder condition for Bellman's functions, Differ. Uravn., 13 (1977), 2180–2187. |
[18] |
A. Marigonda, Second order conditions for the controllability of nonlinear systems with drift, Commun. Pure Appl. Anal., 5 (2006), 861–885. https://doi.org/10.3934/cpaa.2006.5.861 doi: 10.3934/cpaa.2006.5.861
![]() |
[19] |
A. Marigonda, S. Rigo, Controllability of some nonlinear systems with drift via generalized curvature properties, SIAM J. Control Optim., 53 (2015), 434–474. https://doi.org/10.1137/130920691 doi: 10.1137/130920691
![]() |
[20] |
M. Motta, F. Rampazzo, Asymptotic controllability and Lyapunov-like functions determined by Lie brackets, SIAM J. Control Optim., 56 (2018), 1508–1534. https://doi.org/10.1137/16M1086947 doi: 10.1137/16M1086947
![]() |
[21] | N. N. Petrov, Controllability of autonomous systems, Differ. Uravn., 4 (1968), 606–617. |
[22] |
N. N. Petrov, On the bellman function for the time-optimal process problem: PMM vol. 34, n∘=5, 1970, pp. 820–826, J. Appl. Math. Mech., 34 (1970), 785–791. https://doi.org/10.1016/0021-8928(70)90060-2 doi: 10.1016/0021-8928(70)90060-2
![]() |
[23] |
P. Soravia, Hölder continuity of the minimum-time function for C1-manifold targets, J. Optim. Theory Appl., 75 (1992), 401–421. https://doi.org/10.1007/BF00941476 doi: 10.1007/BF00941476
![]() |
[24] |
P. Soravia, Some results on second order controllability conditions, 2019 IEEE 58th Conference on Decision and Control (CDC), 2019, 1468–1473. https://doi.org/10.1109/CDC40024.2019.9029843 doi: 10.1109/CDC40024.2019.9029843
![]() |
[25] | P. Soravia, A degenerate elliptic equation for second order controllability of nonlinear systems, Minimax Theory Appl., 5 (2020), 413–437. |
[26] |
P. Soravia, The Aronsson equation, Lyapunov functions, and local Lipschitz regularity of the minimum time function, Abstr. Appl. Anal., 2019 (2019), 6417074. https://doi.org/10.1155/2019/6417074 doi: 10.1155/2019/6417074
![]() |
[27] |
P. Soravia, A Hamiltonian approach to small time local attainability of manifolds for nonlinear control systems, Appl. Math. Optim., 88 (2023), 1. https://doi.org/10.1007/s00245-023-09973-5 doi: 10.1007/s00245-023-09973-5
![]() |
[28] |
H. J. Sussmann, Lie brackets and local controllability: a sufficient condition for scalar-input systems, SIAM J. Control Optim., 21 (1983), 686–713. https://doi.org/10.1137/0321042 doi: 10.1137/0321042
![]() |
[29] |
H. J. Sussmann, A general theorem on local controllability, SIAM J. Control Optim., 25 (1987), 158–194. https://doi.org/10.1137/0325011 doi: 10.1137/0325011
![]() |