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

Asymptotically stable control problems by infinite horizon optimal control with negative discounting

Fumihiko Nakamura Kitami Institute of Technology, Kitami, 090-8507, Japan nfumihiko@mail.kitami-it.ac.jp
Abstract.

In the paper, for the system which possesses both an attractor and a stable fixed point, we first formulate new stable control problems to find the asymptotically stable control function which realizes to transit a state moving around the attractor to the stable fixed point. Then by using the ordinary differential equation based on the infinite horizon optimal control model with negative discounts, we give one of answers for the stable control problem in a two-dimensional case. Furthermore, under some conditions, we verify that the phase space can be separated to some open connected components depending on the asymptotic behavior of the orbit starting from the initial point in their components. This classification of initial points suggests that it is enable to robustly achieve a stable control. Moreover, we illustrate some numerical results for the stable control obtained by applying our focused system for the Bonhoeffer-van der Pol model.

Key words and phrases:
asymptotic stability, Infinite horizon, negative discount
2010 Mathematics Subject Classification:
49J15, 34D05

1. Problems and main theorems

The asymptotic stability in infinite horizon optimal control problems are well discussed in previous study. For instance, the paper [1] raised a global stable problem for modified Hamiltonian dynamical systems and gave a sufficient condition for the existence of asymptotically stable solutions. In [2], an overtaking and GG-supported properties which guarantee the stable control for a class of non-convex systems are introduced. Also [3] provided some conditions to obtain stable control for nonlinear systems and illustrated numerical examples of its control.

The Bonhoeffer-van der Pol (BvP) model or FitzHugh-Nagumo model are well known example on the neuroscience, which have an unique stable fixed point and stable limit cycle for appropriate parameters [4]. In [5, 6], the method how the state can transit between the stable fixed point and the limit cycle is discussed and some numerical results are displayed. Their idea is based on variational principles and they simulate by using the method called first-order gradient algorithm [7]. Since the transit problems in [5, 6] are beyond the problem supposed in [1], their study inspires a generalization of the problem, that is, the problem is whether it is possible to transit the stable state such as a periodic orbit to another stable state by the additive type of perturbations. From such background, in this paper, we first formulate the generalized problem as follows.

Let F:nnF:\mathbb{R}^{n}\to\mathbb{R}^{n} be a smooth function and consider an autonomous ordinary differential equation z˙=F(z)\dot{z}=F(z) for znz\in\mathbb{R}^{n}. Let the origin 0n0\in\mathbb{R}^{n} be a stable fixed point of the system, and assume that there is an attractor AA whose basin B(A)B(A) does not contain the origin.

Problem 1.

Find a control function χ(t)n\chi(t)\in\mathbb{R}^{n} such that

(i)limt|χ(t)|=0and(ii)U(χ)A.{\rm(i)\ }\displaystyle\lim_{t\to\infty}|\chi(t)|=0\quad\quad{\rm and}\quad\quad\ {\rm(ii)\ }U(\chi)\cap A\neq\emptyset.

where U(χ):={z0n|limtφt(z0)=0}\displaystyle U(\chi):=\{z_{0}\in\mathbb{R}^{n}\ |\lim_{t\to\infty}\varphi^{t}(z_{0})=0\} and φt(z0)\varphi^{t}(z_{0}) denotes the solution of z˙=F(z)+χ\dot{z}=F(z)+\chi with z(0)=z0z(0)=z_{0}. Furthermore, find properties of U(χ)U(\chi).

The properties of U(χ)U(\chi) are, for instance, whether the Lebesgue measure of U(χ)U(\chi) is large enough, small or zero. If the Lebesgue measure of U(χ)U(\chi) is small, then the orbit may not be able to reach to zero by small perturbations.

Now, if the control function χ(t)\chi(t) is determined by some differential equation χ˙(t)=G(z,χ)\dot{\chi}(t)=G(z,\chi), the problem 1 can be reformulated as the following initial value problem.

Problem 2.

Find a smooth function G:2nnG:\mathbb{R}^{2n}\to\mathbb{R}^{n} and a set U2nU\subset\mathbb{R}^{2n} such that

(i)limtΦt(z0,χ0)=0for(z0,χ0)U,and(ii)π1(U)A.{\rm(i)\ }\lim_{t\to\infty}\Phi^{t}(z_{0},\chi_{0})=0\quad{\rm for}\quad(z_{0},\chi_{0})\in U,\quad\quad{\rm and}\quad\quad{\rm(ii)\ }\pi_{1}(U)\cap A\neq\emptyset.

where Φt(z0,χ0)\Phi^{t}(z_{0},\chi_{0}) denotes the solution of z˙=F(z)+χ~\dot{z}=F(z)+\tilde{\chi} and χ˙=G(z,χ)\dot{\chi}=G(z,\chi) with z(0)=z0z(0)=z_{0}, χ(0)=χ0\chi(0)=\chi_{0}, χ~=(χ1,,χk,0,,0)\tilde{\chi}=(\chi_{1},\cdots,\chi_{k},0,\cdots,0) for some knk\leq n, and π1(z,χ)=z\pi_{1}(z,\chi)=z is a projection for zz. Furthermore, find properties of UU.

Here, the reason why we use χ~\tilde{\chi} is because it is desired that we use less control factors from control point of view.

In this paper, we succeeded to give the partial answer to the problems in the case n=2n=2. More precisely, consider the two-dimensional system z˙=F(z)\dot{z}=F(z) which has only two attractors, stable fixed point 0 and stable limit cycle γs\gamma^{s} such that 0I(γs)0\in I(\gamma^{s}) and their basins satisfy cl(B(γs)B(0))=2{\rm cl}(B(\gamma^{s})\cup B(0))=\mathbb{R}^{2}, where I(γ)I(\gamma) is a set of inside points of the closed curve γ\gamma and cl(X){\rm cl}(X) implies a closure of the set XX. In this case, there is also a unique unstable limit cycle γu\gamma^{u} between 0 and γs\gamma^{s}. Then, we focus on the next autonomous ordinary differential equation:

(){z˙=F(z)+q~q˙=[ρI+DFT(z)]qwithq=(q1,q2),q~=(q1,0).\displaystyle(\ast)\ \ \begin{cases}\dot{z}=F(z)+\tilde{q}\\ \dot{q}=-[\rho I+D_{F}^{T}(z)]q\end{cases}\ \ \ \text{with}\ \ \ q=(q_{1},q_{2}),\ \tilde{q}=(q_{1},0).

where ρ\rho is a positive constant, DF(z)D_{F}(z) is the Jacobian matrix of FF at zz, II is the n×nn\times n identity matrix and XTX^{T} is transpose of XX. Assume that the solution Φt(z,q)\Phi^{t}(z,q) of the system (\ast) exists for any initial point (z,q)4(z,q)\in\mathbb{R}^{4} and tt\in\mathbb{R}. Denoting z=(x,y)z=(x,y) and F=(f,g)F=(f,g), we can write it as the following differential equation:

{x˙=f(x,y)+q1y˙=g(x,y)q1˙=(ρ+fx(x,y))q1gx(x,y)q2q2˙=fy(x,y)q1(ρ+gy(x,y))q2\displaystyle\begin{cases}\dot{x}=f(x,y)+q_{1}\\ \dot{y}=g(x,y)\\ \dot{q_{1}}=-(\rho+f_{x}(x,y))q_{1}-g_{x}(x,y)q_{2}\\ \dot{q_{2}}=-f_{y}(x,y)q_{1}-(\rho+g_{y}(x,y))q_{2}\end{cases}

The next theorem shows that, by using the system (\ast), we can obtain the control function χ=(q1,0)\chi=(q_{1},0) which satisfies (i) and (ii) in Problem 1 under the assumption (A1) introduced later.

Theorem A.

For the system (\ast) with the assumptions (A1), there exists a set U4U\subset\mathbb{R}^{4} such that

(i)limtΦt(z0,q0)=0for(z0,q0)U,and(ii)π1(U)γs.{\rm(i)\ }\lim_{t\to\infty}\Phi^{t}(z_{0},q_{0})=0\quad{\rm for}\quad(z_{0},q_{0})\in U,\quad\quad{\rm and}\quad\quad{\rm(ii)\ }\pi_{1}(U)\cap\gamma^{s}\neq\emptyset.

Note that we can immediately find Φt(z0,q0)\Phi^{t}(z_{0},q_{0}) decays exponentially by adjusting ρ\rho satisfying (A1). In addition to (A1), if the conditions (A2)-(A5) and (AA) are assumed, we can show the next theorem which tells us that the set UU of initial points exists as an open connected subset in 4\mathbb{R}^{4}.

Theorem B.

For the system (\ast) with the assumptions (A1)-(A5) and (AA), there exist four disjoint open connected subset A1,A2,A3,A44A_{1},A_{2},A_{3},A_{4}\subset\mathbb{R}^{4} such that cl(A1A2A3A4)=4{\rm cl}(A_{1}\cup A_{2}\cup A_{3}\cup A_{4})=\mathbb{R}^{4} and

limtΦt(z0,q0)={(0,0)for (z0,q0)A2,γs×{0}for (z0,q0)A3,for (z0,q0)A1A4.\displaystyle\lim_{t\to\infty}\Phi^{t}(z_{0},q_{0})=\begin{cases}(0,0)&\text{for }(z_{0},q_{0})\in A_{2},\\ \gamma^{s}\times\{0\}&\text{for }(z_{0},q_{0})\in A_{3},\\ \infty&\text{for }(z_{0},q_{0})\in A_{1}\cup A_{4}.\end{cases}

The optimal solution for the control is generally calculated by using the variational principle which is a general method to find functions which extremize the value of some giving functional. One of the most famous theorem in such optimal problems is Pontryagin’s Maximum Principle [8, 9] which tell us necessary or sufficient conditions for optimal solutions. Moreover, the principle for infinite horizon are well-discussed in [2, 10].

It is known that generally the equations derived from variational principles are act on a space with double dimensions, because it consists of original state variables and control variables. However, the original stable fixed point becomes saddle in the space for the additive type of perturbations. This implies that, in the case that the target state is the stable fixed point, the control variables may diverge even if the state can reach to a target point. Then, our problem is to find the optimal solution which realizes that not only the state can move to the target state but also the control variables converges to zero.

To solve this issue, we consider the equations (\ast) from the variational principle for the functional having discount effects. There are many previous study for the model with positive discounting, for example [12, 11, 13]. In the papers [1, 2, 3], they also consider the positive discounting. On the other hand, there is less study focused on a negative discount. The book [12] says that the fixed point can be stable if the discount rate is negative for linear differential equations. Although the book [11] does not treat the negative discount case, they say that the negative discount rate gives the most importance to what happens in the distant future and this should amplified the stabilizing force. Thus, the reason why we focus on the system (\ast) is the fact that we expect one of answer for the Problem 1 and 2 can be achieved by using the system associated with the infinite horizon optimal control problem with “negative” discounting. We summarize the process to obtain the system (\ast) in Appendix A.

We consider the two-dimensional systems in this paper in order to apply the concrete example (BvP model). Indeed, we apply our theorem and display our numerical results in section 4. Moreover, for one-dimensional case, we can derive the same results under less conditions. Although the one-dimensional case is relatively simple, we summarize it in Appendix B since the argument helps us to understand the two-dimensional model.

1.1. Notations and assumptions

Before we mention our assumptions, we give some notations. First the Jacobian matrix of the system (\ast) can be calculate by

J(z,q)=(fx(z)fy(z)10gx(z)gy(z)00h1(z,q)h2(z,q)(ρ+fx(z))gx(z)h3(z,q)h4(z,q)fy(z)(ρ+gy(z)))J_{*}(z,q)=\begin{pmatrix}f_{x}(z)&f_{y}(z)&1&0\\ g_{x}(z)&g_{y}(z)&0&0\\ -h_{1}(z,q)&-h_{2}(z,q)&-(\rho+f_{x}(z))&-g_{x}(z)\\ -h_{3}(z,q)&-h_{4}(z,q)&-f_{y}(z)&-(\rho+g_{y}(z))\end{pmatrix}

where

(h1(z,q)h2(z,q)h3(z,q)h4(z,q))=(fxx(x,y)q1+gxx(x,y)q2fxy(x,y)q1+gxy(x,y)q2fyx(x,y)q1+gyx(x,y)q2fyy(x,y)q1+gyy(x,y)q2).\displaystyle\begin{pmatrix}h_{1}(z,q)&h_{2}(z,q)\\ h_{3}(z,q)&h_{4}(z,q)\end{pmatrix}=\begin{pmatrix}f_{xx}(x,y)q_{1}+g_{xx}(x,y)q_{2}&f_{xy}(x,y)q_{1}+g_{xy}(x,y)q_{2}\\ f_{yx}(x,y)q_{1}+g_{yx}(x,y)q_{2}&f_{yy}(x,y)q_{1}+g_{yy}(x,y)q_{2}\end{pmatrix}. (1)

Here fxf_{x} or fxyf_{xy} denote the partial derivative. It is obvious that the divergence of the system (\ast) satisfies div=2ρ{\rm div}_{*}=-2\rho, which implies volume contracting as tt\to\infty.

Moreover, we prepare the following stable and unstable sets for the fixed points (z,q)(z_{*},q_{*}) and periodic orbit γ\gamma of the system (\ast) as follows:

Ws(z,q)\displaystyle W^{s}(z_{*},q_{*}) :=\displaystyle:= {(z,q)4|Φt(z,q)(z,q) as t}\displaystyle\{(z,q)\in\mathbb{R}^{4}\ |\ \Phi^{t}(z,q)\to(z_{*},q_{*})\text{ as }t\to\infty\}
Wu(z,q)\displaystyle W^{u}(z_{*},q_{*}) :=\displaystyle:= {(z,q)4|Φt(z,q)(z,q) as t}\displaystyle\{(z,q)\in\mathbb{R}^{4}\ |\ \Phi^{t}(z,q)\to(z_{*},q_{*})\text{ as }t\to-\infty\}
Ws(γ)\displaystyle W^{s}(\gamma) :=\displaystyle:= {(z,q)4|d(Φt(z,q),γ)0 as t}\displaystyle\{(z,q)\in\mathbb{R}^{4}\ |\ d(\Phi^{t}(z,q),\gamma)\to 0\text{ as }t\to\infty\}
Wu(γ)\displaystyle W^{u}(\gamma) :=\displaystyle:= {(z,q)4|d(Φt(z,q),γ)0 as t}\displaystyle\{(z,q)\in\mathbb{R}^{4}\ |\ d(\Phi^{t}(z,q),\gamma)\to 0\text{ as }t\to-\infty\}

where

d((z,q),γ):=infwγ|(z,q)(w,0)|.d((z,q),\gamma):=\inf_{w\in\gamma}|(z,q)-(w,0)|.

and |||\cdot| denotes the usual Euclidean norm.

Next, we define the set Dρ2D_{\rho}\in\mathbb{R}^{2} as

Dρ:={z2|the matrix [ρI+DF(z)] is negative definite},D_{\rho}:=\{z\in\mathbb{R}^{2}\ |\ \text{the matrix }-[\rho I+D_{F}(z)]\text{ is negative definite}\},

where a n×nn\times n matrix AA is positive (negative) definite if the symmetric matrix (A+AT)/2(A+A^{T})/2 is positive (negative) definite, that is, xTA+AT2x>0x^{T}\frac{A+A^{T}}{2}x>0 (<0<0) holds for any vector xn\{0}x\in\mathbb{R}^{n}\backslash\{0\}. It is well-known that a symmetric matrix AA is positive (negative) definite if and only if all eigenvalues for AA are positive (negative).

Now, the assumptions (A1)-(A5) are stated as follows:

  • (A1)

    I(γs)DρI(\gamma^{s})\subset D_{\rho}.

  • (A2)

    #{z2|det[ρI+DF(z)]=0,g(z)=0}=2\#\{z\in\mathbb{R}^{2}\ |\ det[\rho I+D_{F}(z)]=0,g(z)=0\}=2.

  • (A3)

    For any nontrivial fixed point (z~,q~)(\tilde{z},\tilde{q}) for system (\ast) satisfying the equations

    det[ρI+DF(z)]=0,[ρI+DF(z)]q=0(q0),F(z)+q~=0,\displaystyle det[\rho I+D_{F}(z)]=0,\ \ [\rho I+D_{F}(z)]q=0\ (q\neq 0),\ \ F(z)+\tilde{q}=0, (2)

    the following inequality holds:

    (gx(z~)gy(z~))(h4(z~,q~)h2(z~,q~)h3(z~,q~)h1(z~,q~))(gx(z~)gy(z~))\displaystyle\begin{pmatrix}g_{x}(\tilde{z})&g_{y}(\tilde{z})\end{pmatrix}\begin{pmatrix}h_{4}(\tilde{z},\tilde{q})&-h_{2}(\tilde{z},\tilde{q})\\ -h_{3}(\tilde{z},\tilde{q})&h_{1}(\tilde{z},\tilde{q})\end{pmatrix}\begin{pmatrix}g_{x}(\tilde{z})\\ g_{y}(\tilde{z})\end{pmatrix}
    +ρ(h1(z~,q~)gy(z~)h2(z~,q~)gx(z~))>0\displaystyle\hskip 142.26378pt+\rho(h_{1}(\tilde{z},\tilde{q})g_{y}(\tilde{z})-h_{2}(\tilde{z},\tilde{q})g_{x}(\tilde{z}))>0
  • (A4)

    There exist a set K2K\subset\mathbb{R}^{2} and a function V:2V:\mathbb{R}^{2}\to\mathbb{R} such that

    (i) V(z)>0V(z)>0 in 2\K\mathbb{R}^{2}\backslash K,  (ii) V˙(z(t))<0\dot{V}(z(t))<0 in 2\K\mathbb{R}^{2}\backslash K, (iii) V(z)V(z)\to\infty if |z||z|\to\infty.
  • (A5)

    The matrix [ρI+DF(z)]-[\rho I+D_{F}(z)] is positive definite for z2\Kz\in\mathbb{R}^{2}\backslash K where the set KK is of (A4).

Fortunately, Theorem A can be hold under only the assumption (A1) for our system (\ast). (A2) means that there are two nontrivial fixed point of the system (\ast) which satisfy the equation (2). Although this condition seems to be a strong condition, we can calculate only two fixed points for BvP model in the section 4 by choosing appropriate ρ\rho. (A3) plays an important role in order that the stable set Ws(z~,q~)W^{s}(\tilde{z},\tilde{q}) becomes a three-dimensional stable manifold in 4\mathbb{R}^{4}. The Lyapunov function-like assumption (A4) might be a natural condition since we consider that the original system z˙=F(z)\dot{z}=F(z) has only one fixed point and stable limit cycles as its attractors. (A5) is also important for the separation of 4\mathbb{R}^{4} in Theorem B.

Finally, one of difficulties of the arguments is that we do not know the existence of non-trivial closed orbit for our four-dimensional differential equation. The famous Poincare-Bendixson Theorem cannot be applied to more than three-dimensional system generally. Although some of previous study said about the non-existence of closed orbits, for instance [14] showed it by the condition of sum of the first and second eigenvalues, it is difficult to apply their results to our system. Thus, to prove Theorem B, assume the following condition:

  • (AA)

    The system (\ast) has no non-wandering point in K\Dρ×2K\backslash D_{\rho}\times\mathbb{R}^{2}, where (z,q)4(z,q)\in\mathbb{R}^{4} is said to be non-wandering point if, for any open set U4U\subset\mathbb{R}^{4} containing (z,q)(z,q) and any T>0T>0, there exists t>Tt>T such that Φt(U)U\Phi^{t}(U)\cap U\neq\emptyset.

2. Proof of Theorem A

To prove the Theorem A, we prepare the next proposition and the lemma.

Proposition 2.1.

For a linear system q˙(t)=A(x)q(t)\dot{q}(t)=A(x)q(t) with q(0)=q0q(0)=q_{0} where A(x)A(x) is a n×nn\times n matrix for x(t)x(t) which is a continuous map in some compact subset KnK\subset\mathbb{R}^{n} and q(t)nq(t)\in\mathbb{R}^{n} is a vector for tt\in\mathbb{R}. If A(x)A(x) is negative definite for any xKx\in K, then |q(t)||q(t)| converges to 0 as tt\to\infty.

Proof.

Since A(x)A(x) is negative definite, we can calculate as follows:

ddt|q(t)|2\displaystyle\frac{d}{dt}\lvert q(t)\rvert^{2} =\displaystyle= qT(t)q˙(t)+q˙T(t)q(t)\displaystyle q^{T}(t)\dot{q}(t)+\dot{q}^{T}(t)q(t)
=\displaystyle= qT(t)A(x)q(t)+qT(t)AT(x)q(t)\displaystyle q^{T}(t)A(x)q(t)+q^{T}(t)A^{T}(x)q(t)
=\displaystyle= qT(t)(A(x)+AT(x))q(t)<0for any t>0.\displaystyle q^{T}(t)(A(x)+A^{T}(x))q(t)<0\ \text{for any $t>0$}.

Therefore, |q(t)|2\lvert q(t)\rvert^{2} is monotonically decreasing as tt\to\infty, and |q(t)|\lvert q(t)\rvert must converge to 0 as tt\to\infty since ddt|q(t)|2<0\frac{d}{dt}\lvert q(t)\rvert^{2}<0 for any q(t)0q(t)\neq 0. ∎

Lemma 2.2.

The point (0,0)4(0,0)\in\mathbb{R}^{4} becomes a stable fixed point of the system (\ast). Moreover, the set γs×{0}\gamma^{s}\times\{0\} and γu×{0}\gamma^{u}\times\{0\} become a stable and unstable limit cycle in 4\mathbb{R}^{4}, respectively.

Proof.

By calculating Jacobian matrix for the fixed point (0,0)4(0,0)\in\mathbb{R}^{4}, we immediately find the eigenvalues of the matrix are λ1(0)\lambda_{1}(0), λ2(0)\lambda_{2}(0), λ1(0)ρ-\lambda_{1}(0)-\rho and λ2(0)ρ-\lambda_{2}(0)-\rho. Then (0,0)(0,0) becomes a stable fixed point since the assumption (A1) since a negative definite matrix [ρI+DF(0)]-[\rho I+D_{F}(0)] has two eigenvalues with a negative real part.

Next, from the assumption (A1), when z(t)z(t) moves around the neighborhood of the periodic orbit in DρD_{\rho}, then [ρI+DF(z)]-[\rho I+D_{F}(z)] is always negative definite. Thus, by Proposition 2.1, q1(t)q_{1}(t) and q2(t)q_{2}(t) converge to 0 Therefore, γs×{0,0}\gamma^{s}\times\{0,0\} and γu×{0,0}\gamma^{u}\times\{0,0\} become a stable and unstable limit cycle on 4\mathbb{R}^{4}. ∎

Proof of Theorem A

By the assumption (A1), there is a small ε>0\varepsilon>0 such that limtΦt(Bε(0,0))=0\displaystyle\lim_{t\to\infty}\Phi^{t}(B_{\varepsilon}(0,0))=0 where Bε(0,0)B_{\varepsilon}(0,0) is a open ball in 4\mathbb{R}^{4} with center (0,0)(0,0) and radius ε>0\varepsilon>0. We will show that π1(Φt(Bε(0,0)))γs\pi_{1}(\Phi^{-t}(B_{\varepsilon}(0,0)))\cap\gamma^{s}\neq\emptyset. Assume that the set is empty set. Since 0I(γs)20\in I(\gamma^{s})\subset\mathbb{R}^{2}, it must be hold that π1(Φt(Bε(0,0)))I(γs)\pi_{1}(\Phi^{-t}(B_{\varepsilon}(0,0)))\subset I(\gamma^{s}) which implies π1(Φt(Bε(0,0)))Dρ\pi_{1}(\Phi^{-t}(B_{\varepsilon}(0,0)))\subset D_{\rho}. Then, any orbit z(t)z(-t) is included in the set DρD_{\rho} for any t0t\geq 0 so that the matrix [ρI+DF(z)][\rho I+D_{F}(z)] is always positive definite. From the similar argument of Proposition 2.1, we have |q||q| monotonically increases to infinity as tt\to-\infty for any (z0,q0)Bε(0,0)(z_{0},q_{0})\in B_{\varepsilon}(0,0).

Assume that q1(t)q_{1}(-t) is bounded for any t>0t>0, then |q2||q_{2}| must increase since |q||q| is increasing. Moreover, there is some TT such that q2(t)q_{2}(-t) is always positive (or negative) for any t>Tt>T and goes to \infty (or -\infty). For the case q2(t)q_{2}(-t)\to\infty (the case q2(t)q_{2}(-t)\to-\infty is same), consider the equation q˙=(ρ+fx)q1+gxq2\dot{q}=(\rho+f_{x})q_{1}+g_{x}q_{2}. Note that the minus sings of the original equation are replaced by plus sings because we now consider the inverse time direction, tt\to-\infty. If gx=0g_{x}=0, since ρ+fx\rho+f_{x} is positive due to positive definite matrix [ρI+DF(z)][\rho I+D_{F}(z)], we have q1(t)e(ρ+minfx)tq_{1}(t)\geq e^{(\rho+\min f_{x})t}\to\infty as tt\to\infty, which contradicts to the boundedness of q1q_{1}. If gxg_{x} is always positive (or negative), since z(t)z(t) and q1(t)q_{1}(t) are bounded but q2(t)q_{2}(t) increase, we have q˙1(t)>C\dot{q}_{1}(t)>C (or q˙1(t)<C\dot{q}_{1}(t)<-C) for some C>0C>0 and sufficient large tt. This means that q1q_{1} cannot be bounded which is contradiction. If the case that gx(z(t))g_{x}(z(t)) has oscillated and repeated positive and negative values, the function G(t):=gx(t)q2(t)G(t):=g_{x}(t)q_{2}(t) oscillates and diverges. In this case, let {si}\{s_{i}\} and {ti}\{t_{i}\} be times satisfying G(si)=G(ti)=0G(s_{i})=G(t_{i})=0, G˙(si)>0\dot{G}(s_{i})>0 and G˙(ti)<0\dot{G}(t_{i})<0. Then, since z(t)z(t) and q1(t)q_{1}(t) are bounded but G(t)G(t) is unbounded, the time interval ti+1tit_{i+1}-t_{i} must converge to zero as ii\to\infty. Indeed, letting

ai=sitiG(t)𝑑tandbi=tisi+1G(t)𝑑t,a_{i}=\int_{s_{i}}^{t_{i}}G(t)dt\quad{\rm and}\quad b_{i}=\int_{t_{i}}^{s_{i+1}}G(t)dt,

we have bi+1bi<0<aiai+1b_{i+1}\leq b_{i}<0<a_{i}\leq a_{i+1} and s1ti+1G(t)𝑑t=ai+bi\int_{s_{1}}^{t_{i+1}}G(t)dt=\sum a_{i}+\sum b_{i} must be bounded since q1q_{1} is bounded. This implies ti+1ti0t_{i+1}-t_{i}\to 0 as ii\to\infty due to the increasing q2q_{2}.

However, if ti+1tit_{i+1}-t_{i}, either z˙(t)\dot{z}(t) is unbounded or gx(t)0g_{x}(t)\to 0, and both lead a contradiction. Therefore, we found q1(t)q_{1}(t) is unbounded for tt\to-\infty.

Finally, considering the equation x˙=f(z)q1\dot{x}=-f(z)-q_{1}, the function q1(t)q_{1}(t) must oscillate and diverge. Indeed if q1q_{1} does not oscillate and does diverge, there are C>0C>0 and T>0T>0 such that x˙(t)>C\dot{x}(t)>C (or x˙(t)<C\dot{x}(t)<-C) for any t>Tt>T, which implies x(t)x(t) is unbounded. Hence, let {si}\{s_{i}\} and {ti}\{t_{i}\} be times satisfying q1(si)=q1(ti)=0q_{1}(s_{i})=q_{1}(t_{i})=0, q˙1(si)>0\dot{q}_{1}(s_{i})>0 and q˙1(ti)<0\dot{q}_{1}(t_{i})<0. Then, since x(t)x(t) and f(z)f(z) are bounded but q1(t)q_{1}(t) is unbounded, the time interval ti+1tit_{i+1}-t_{i} must converge to zero as ii\to\infty. However, this is impossible since the imaginary part of eigenvalues of the matrix [ρI+DF(z)][\rho I+D_{F}(z)] (if they are complex numbers) is bounded so that the frequency of the oscillation of q1q_{1} cannot be infinity which contradicts to ti+1ti0t_{i+1}-t_{i}\to 0. (When the eigenvalues are real number, ti+1ti0t_{i+1}-t_{i}\to 0 cannot occur since q1q_{1} must get away from 0.)

From these arguments, we can show that z(t)z(t) must go out in DρD_{\rho} as tt\to-\infty for any (z0,q0)Bε(0,0)(z_{0},q_{0})\in B_{\varepsilon}(0,0), which implies π1(Φt(Bε(0,0)))γs\pi_{1}(\Phi^{-t}(B_{\varepsilon}(0,0)))\cap\gamma^{s}\neq\emptyset. This completes the proof.

3. Proof of Theorem B

The ideas of the proof of Theorem B are to show that firstly the stable sets Ws(z~,q~)W^{s}(\tilde{z},\tilde{q}) and Ws(γu)W^{s}(\gamma^{u}) become three-dimensional smooth manifolds by the Lemma 3.1 and 3.2. Secondary, by knowing behaviors of almost all orbits Φt(z0,q0)\Phi^{t}(z_{0},q_{0}) as tt\to-\infty from Lemma 3.4, we can derive these stable manifolds separate 4\mathbb{R}^{4} to four-disjoint open connected sets. To prove this, the separation theorem in [17] which is well-known in differential topology are used. Denote the two non-trivial fixed point by (z~1,q~1)(\tilde{z}_{1},\tilde{q}_{1}) and (z~2,q~2)(\tilde{z}_{2},\tilde{q}_{2}) satisfying the equation (2).

Lemma 3.1.

Under the assumptions (A2) and (A3), the Jacobian D(z~i,q~i)D_{*}(\tilde{z}_{i},\tilde{q}_{i}) has three eigenvalues with negative real parts for the non-trivial fixed points (z~i,q~i)(\tilde{z}_{i},\tilde{q}_{i}), i=1,2i=1,2, so that Ws(z~i,q~i)W^{s}(\tilde{z}_{i},\tilde{q}_{i}) becomes a three-dimensional smooth manifold.

Proof.

For the fixed point (z~i,q~i)(\tilde{z}_{i},\tilde{q}_{i}), we can calculate the characteristic polynomial as follows:

Λ(λ)\displaystyle\Lambda(\lambda) :=\displaystyle:= det[λIDF(z~i,q~i)]\displaystyle det[\lambda I-D_{F}(\tilde{z}_{i},\tilde{q}_{i})] (3)
=\displaystyle= λ4+2ρλ3{ρ2+3(fx+gy)ρ+(fx+gy)2h1}λ2\displaystyle\lambda^{4}+2\rho\lambda^{3}-\{\rho^{2}+3(f_{x}+g_{y})\rho+(f_{x}+g_{y})^{2}-h_{1}\}\lambda^{2}
{2ρ2+3ρ(fx+gy)+(fx+gy)2h1}ρλB(z~i)\displaystyle-\{2\rho^{2}+3\rho(f_{x}+g_{y})+(f_{x}+g_{y})^{2}-h_{1}\}\rho\lambda-B(\tilde{z}_{i})

where

B(z~i)=B(x~i,y~i)=(gxgy)(h4h2h3h1)(gxgy)ρ(h1gyh2gx).B(\tilde{z}_{i})=B(\tilde{x}_{i},\tilde{y}_{i})=-\begin{pmatrix}g_{x}&g_{y}\end{pmatrix}\begin{pmatrix}h_{4}&-h_{2}\\ -h_{3}&h_{1}\end{pmatrix}\begin{pmatrix}g_{x}\\ g_{y}\end{pmatrix}-\rho(h_{1}g_{y}-h_{2}g_{x}).

Then we have

ddλΛ(λ)\displaystyle\frac{d}{d\lambda}\Lambda(\lambda) =\displaystyle= 4λ3+6ρλ22{ρ2+3(fx+gy)ρ+(fx+gy)2h1}λ\displaystyle 4\lambda^{3}+6\rho\lambda^{2}-2\{\rho^{2}+3(f_{x}+g_{y})\rho+(f_{x}+g_{y})^{2}-h_{1}\}\lambda
2{2ρ2+3(fx+gy)ρ+(fx+gy)2h1}\displaystyle-2\{2\rho^{2}+3(f_{x}+g_{y})\rho+(f_{x}+g_{y})^{2}-h_{1}\}
=\displaystyle= 2(λ+ρ/2)(2λ2+3ρλ{2ρ2+3(fx+gy)ρ+(fx+gy)2h1})\displaystyle 2(\lambda+\rho/2)(2\lambda^{2}+3\rho\lambda-\{2\rho^{2}+3(f_{x}+g_{y})\rho+(f_{x}+g_{y})^{2}-h_{1}\})

Thus, one of the extremal values of Λ(λ)\Lambda(\lambda) takes at λ=ρ2\lambda=-\frac{\rho}{2}, and we find that if Λ(0)<0\Lambda(0)<0, then the equation Λ(λ)=0\Lambda(\lambda)=0 has three negative solutions and a positive solution when all eigenvalues are real numbers. The case Λ(λ)\Lambda(\lambda) has two complex eigenvalues, we can calculate its real parts are both negative if Λ(0)<0\Lambda(0)<0. Therefore, the stable manifold theorem [16] and the assumption (A3) which implies the inequality Λ(0)<0\Lambda(0)<0 guarantees Ws(z~i,q~i)W^{s}(\tilde{z}_{i},\tilde{q}_{i}) is a three-dimensional smooth manifold.

Lemma 3.2.

Ws(γu)S1×2W^{s}(\gamma^{u})\simeq S^{1}\times\mathbb{R}^{2} is a three-dimensional smooth manifold.

Proof.

Due to γuDρ\gamma^{u}\subset D_{\rho}, |q(t)||q(t)| decreases monotonically and converges to 0 as tt\to\infty if z(t)z(t) moves around the neighbor of γu\gamma^{u}, which implies both q1(t)q_{1}(t) and q2(t)q_{2}(t) are stable. Thus, from the stable manifold theorem for a periodic orbit [16], Ws(γu)S1×2W^{s}(\gamma^{u})\simeq S^{1}\times\mathbb{R}^{2} becomes a three-dimensional smooth manifold. ∎

Lemma 3.3.

If (A4) is assumed for the original two-dimensional system z˙=F(z)\dot{z}=F(z), then, for any z0z_{0}, there exists T>0T>0 such that the solution z(t)z(t) with z(0)=z0z(0)=z_{0} belongs to DρD_{\rho} for t>Tt>T.

Proof.

Assume that the solution z(t)z(t) does not belong to DρD_{\rho} for any tt. Since V(z(t))V(z(t)) is monotonic decreasing and positive, V(z(t))V(z(t)) converges to some a>0a>0, that is, aV(z)V(z(0))a\leq V(z)\leq V(z(0)) holds. Letting AA be a set {z2|aV(z)V(z(0))}\{z\in\mathbb{R}^{2}\ |\ a\leq V(z)\leq V(z(0))\}, the set AA becomes clearly closed and bounded set. Thus AA is compact. Then, since V˙\dot{V} has a supremum in AA, that is, supzAV˙(z)=b<0\displaystyle\sup_{z\in A}\dot{V}(z)=-b<0, we have

V(z(t))=V(z(0))+0tV˙(z(t))𝑑tV(x(0))bt.V(z(t))=V(z(0))+\int_{0}^{t}\dot{V}(z(t))dt\leq V(x(0))-bt.

Therefore, V(x(t))<0V(x(t))<0 holds for t>V(x(0))/bt>V(x(0))/b which contradicts to the assumption (i) in (A4).

Lemma 3.4.

Assume (A1)-(A5). Then |z(t)||z(t)|\to\infty and |q(t)|0|q(t)|\to 0 as tt\to-\infty for any initial points (z0,q0)(z_{0},q_{0}) in 4\mathbb{R}^{4} except with Wu(z~1,q~1)Wu(z~2,q~2)Wu(γu){0,0}γs×{0}W^{u}(\tilde{z}_{1},\tilde{q}_{1})\cup W^{u}(\tilde{z}_{2},\tilde{q}_{2})\cup W^{u}(\gamma^{u})\cup\{0,0\}\cup\gamma^{s}\times\{0\}.

Proof.

First we have already know the convergent of orbits from initial points in Wu(z~1,q~1)Wu(z~2,q~2)Wu(γu){0,0}γs×{0}W^{u}(\tilde{z}_{1},\tilde{q}_{1})\cup W^{u}(\tilde{z}_{2},\tilde{q}_{2})\cup W^{u}(\gamma^{u})\cup\{0,0\}\cup\gamma^{s}\times\{0\}. Thus, we chose the initial point (z0,q0)(z_{0},q_{0}) in 4\mathbb{R}^{4} except with these sets.

From Theorem A, there is T>0T>0 such that π1(Φt(z0,q0))2\Dρ\pi_{1}(\Phi^{-t}(z_{0},q_{0}))\in\mathbb{R}^{2}\backslash D_{\rho} for t>Tt>T. Assume that π1(Φt(z0,q0))K\pi_{1}(\Phi^{-t}(z_{0},q_{0}))\in K for any t>Tt>T. In the case that |q||q| is increasing and diverge, we can lead a contradiction from the similar arguments with Theorem A. In the case |q||q| is bounded, π1(Φt(z0,q0))\pi_{1}(\Phi^{-t}(z_{0},q_{0})) moves in some bounded region for any t>Tt>T, which implies the existence of non-wandering point (z¯,q¯)(\overline{z},\overline{q}) in KK since we can take subsequence {ti}\{t_{i}\} such that Φti(z0,q0)(z¯,q¯)\Phi^{-t_{i}}(z_{0},q_{0})\to(\overline{z},\overline{q}) as ii\to\infty (by Bolzano-Weierstrass theorem). This contradicts to (AA). Hence, although the orbit may go back to KK again after it goes out from KK, by (AA), there is T>TT^{\prime}>T such that π1(Φt(z0,q0))2\K\pi_{1}(\Phi^{-t}(z_{0},q_{0}))\in\mathbb{R}^{2}\backslash K for any t>Tt>T^{\prime}.

Then, since [ρI+DF(z)][\rho I+D_{F}(z)] is negative definite if z2\Kz\in\mathbb{R}^{2}\backslash K by (A5), we have |q(t)|0|q(t)|\to 0 as tt\to-\infty. This implies q1(t)0q_{1}(t)\to 0 as tt\to-\infty and the orbit follows the original system z˙=F(z)\dot{z}=F(z). By substituting w(t)=z(t)w(t)=z(-t), the system becomes w˙(t)=F(w)\dot{w}(t)=-F(w) and V˙(w(t))=V˙(z(t))>0\dot{V}(w(t))=-\dot{V}(-z(t))>0 in 2\Dρ\mathbb{R}^{2}\backslash D_{\rho}. Thus, V(w(t))V(w(t)) monotonically increases and goes to infinity as tt\to\infty. If w(t)w(t) is bounded, V(w)V(w) is also bounded, which is contradiction. Therefore, |z(t)||z(t)|\to\infty as tt\to-\infty. This completes the proof. ∎

Proof of Theorem B

First prove that Ws(z~,q~)W^{s}(\tilde{z},\tilde{q}) separates 4\mathbb{R}^{4} to two connected sets where (z~,q~)(\tilde{z},\tilde{q}) is the fixed point satisfying Eq.(2). Consider Φt(Ws(z~,q~)Bε(z~,q~))\Phi^{-t}(W^{s}(\tilde{z},\tilde{q})\cap B_{\varepsilon}(\tilde{z},\tilde{q})). From Lemma 3.4, any orbits Φt(z0,q0)\Phi^{-t}(z_{0},q_{0})\to\infty as tt\to\infty for (z0,q0)Ws(z~,q~)Bε(z~,q~)(z_{0},q_{0})\in W^{s}(\tilde{z},\tilde{q})\cap B_{\varepsilon}(\tilde{z},\tilde{q}).

Now we show that the closure of Ws(z~,q~)W^{s}(\tilde{z},\tilde{q}) has no boundary. Assume that the closure of the boundary is not empty, namely, we have AWs(z~,q~)A\subset\partial W^{s}(\tilde{z},\tilde{q}). Since Ws(z~,q~)W^{s}(\tilde{z},\tilde{q}) is a smooth manifold, AWs(z~,q~)A\nsubseteq W^{s}(\tilde{z},\tilde{q}). Thus, for any yAy\in A, there exist {ti}i=1\{t_{i}\}_{i=1}^{\infty} and (zi,qi)Bε(z~,q~)(z_{i},q_{i})\in\partial B_{\varepsilon}(\tilde{z},\tilde{q}) such that Φti(zi,qi)y\Phi^{-t_{i}}(z_{i},q_{i})\to y for some small ε>0\varepsilon>0. Since Bε(z~,q~)\partial B_{\varepsilon}(\tilde{z},\tilde{q}) is compact set, there is subsequence {ik}k\{i_{k}\}_{k} such that (zik,qik)(z¯,q¯)Bε(z~,q~)(z_{i_{k}},q_{i_{k}})\to(\overline{z},\overline{q})\in\partial B_{\varepsilon}(\tilde{z},\tilde{q}). From Lemma 3.4, for any L>0L>0, there is kk with tik>Tt_{i_{k}}>T such that |π1(Φtik(zik,qik))|L|\pi_{1}(\Phi^{-t_{i_{k}}}(z_{i_{k}},q_{i_{k}}))|\geq L for any LL, which contradicts to Φtik(zik,qik)yA\Phi^{-t_{i_{k}}}(z_{i_{k}},q_{i_{k}})\to y\in A as kk\to\infty.

Then, Theorem 4.6 in [17] tells us that if NN is a simply connected manifold and MNM\subset N is a connected closed submanifold of codimension 1 with M=N=\partial M=\partial N=\emptyset, then MM separates NN. In our case, although Ws(z~,q~)W^{s}(\tilde{z},\tilde{q}) is not closed, we can prove that Ws(z~,q~)W^{s}(\tilde{z},\tilde{q}) separates 4\mathbb{R}^{4} by modifying the proof of the theorem since Ws(z~,q~)W^{s}(\tilde{z},\tilde{q}) spreads to infinity. (Indeed, we can check similar proof does work even our situation since, for any r>0r>0, the closed disc D4(r)4D^{4}(r)\subset\mathbb{R}^{4} satisfies (Ws(z~,q~)D4(r))D4(r)\partial(W^{s}(\tilde{z},\tilde{q})\cap D^{4}(r))\subset\partial D^{4}(r). See [17] for detail.)

Similarly, we have that Ws(γu)W^{s}(\gamma^{u}) also separates 4\mathbb{R}^{4}. Hence, there exist three separatrixes Ws(z~1,q~1)W^{s}(\tilde{z}_{1},\tilde{q}_{1}), Ws(z~2,q~2)W^{s}(\tilde{z}_{2},\tilde{q}_{2}) and Ws(γu)W^{s}(\gamma^{u}) which separate 4\mathbb{R}^{4} to four connected components since each stable set has no intersection due to a uniqueness of the solution. The four sets denote by A1,A2,A3A_{1},A_{2},A_{3} and A4A_{4}. Ws(γu)W^{s}(\gamma^{u}) separates 0 and γs\gamma^{s} since Wu(γu)=I(γs)×{0}\{0,0}W^{u}(\gamma^{u})=I(\gamma^{s})\times\{0\}\backslash\{0,0\} and Wu(γu)W^{u}(\gamma^{u}) and Ws(γu)W^{s}(\gamma^{u}) do not intersect transversally. Hence we can take the sets satisfying 0A20\in A_{2} and γsA3\gamma^{s}\in A_{3}. Moreover, the set A2A_{2} and A3A_{3} are bounded for qq, that is, there is a constant L>0L>0 such that |q|<L|q|<L for (z,q)A2A3(z,q)\in A_{2}\cup A_{3} since A2A3A_{2}\cup A_{3} are between Ws(z~1,q~1)W^{s}(\tilde{z}_{1},\tilde{q}_{1}) and Ws(z~2,q~2)W^{s}(\tilde{z}_{2},\tilde{q}_{2}).

Finally, there is no non-trivial attractor for our system (\ast) since if zDρz\in D_{\rho} (or zKcz\in K^{c}), then |q||q| has decreased (or increased), and there is no non-wandering point in K\DρK\backslash D_{\rho}. Since there is no attractor except with the stable fixed point 0 in A2A_{2}, the orbit Φt(z0,q0)\Phi^{t}(z_{0},q_{0}) goes to either 0 or \infty for (z0,q0)A2(z_{0},q_{0})\in A_{2}. Assuming that Φt(z0,q0)\Phi^{t}(z_{0},q_{0}) diverges, (x(t),q(t))(x(t),q(t)) must go to (,0)(\infty,0) since |q||q| is bounded, which contradicts to (A4). Thus, Φt(z0,q0)\Phi^{t}(z_{0},q_{0}) converges to 0 for any (z0,q0)A2(z_{0},q_{0})\in A_{2}. Similarly, Φt(z0,q0)\Phi^{t}(z_{0},q_{0}) converges to γs\gamma^{s} for (z0,q0)A3(z_{0},q_{0})\in A_{3}. Since there is no attractor in A1A_{1} and A4A_{4}, the orbit Φt(z0,q0)\Phi^{t}(z_{0},q_{0}) must diverge for (z0,q0)A1A4(z_{0},q_{0})\in A_{1}\cup A_{4}. This completes the proof.∎

Remark 3.5.

Even if the assumption (A4) and (A5) are not assumed, the stable set Ws(z~,q~)W^{s}(\tilde{z},\tilde{q}) can separate 4\mathbb{R}^{4} if any orbits Φt(z0,q0)\Phi^{-t}(z_{0},q_{0})\to\infty as tt\to\infty for (z0,q0)Ws(z~,q~)(z_{0},q_{0})\in W^{s}(\tilde{z},\tilde{q}). However, in this case, we cannot conclude that Φt(z0,q0)0\Phi^{t}(z_{0},q_{0})\to 0 (or γs\gamma^{s}) for any (z0,q0)A2(z_{0},q_{0})\in A_{2} (or A3A_{3}) since the sets A2A_{2} and A3A_{3} may be unbounded for qq and some orbits in the sets may diverge.

Furthermore, if not only (A4) and (A5) but also (AA) are not imposed, some non-trivial invariant set may exist. The system has no fixed point except with 0 and (z~i,q~i)(\tilde{z}_{i},\tilde{q}_{i}), i=1,2i=1,2, and non-trivial attractor for tt\to-\infty cannot exists because of div=2ρ{}_{*}=2\rho. However, we cannot verify the existence of invariant set like a periodic orbit or chaos orbit. If it has two or three dimension as its unstable directions, a heteroclinic orbit between it and (z~,q~)(\tilde{z},\tilde{q}) or γu\gamma^{u} may exist and this makes our discussions more complicated.

4. Application for Bonhoeffer-van der Pol model

In this section, we apply our results to the concrete example called Bonhoeffer-van der Pol (BvP) model described by

{x˙=c(xx3/3+yr)y˙=(x+a+by)/c.\displaystyle\begin{cases}\dot{x}=c(x-x^{3}/3+y-r)\\ \dot{y}=-(x+a+by)/c\end{cases}. (4)

It is well known that this system has a stable limit cycle, a unstable limit cycle and a stable fixed point when a=0.7,b=0.8,c=3.0,r=0.342a=0.7,b=0.8,c=3.0,r=0.342, and the fixed point (x,y)(0.958366,0.322957)(x_{*},y_{*})\fallingdotseq(0.958366,0.322957). We can calculate the Jacobian matrix as

DF(x,y)=(c(1x2)c1/cb/c).D_{F}(x,y)=\begin{pmatrix}c(1-x^{2})&c\\ -1/c&-b/c\end{pmatrix}.

The eigenvalues λ±\lambda_{\pm} of DF(x,y)D_{F}(x_{*},y_{*}) are 0.011031±0.966773i-0.011031\pm 0.966773i. Next, by applying the system (\ast) for BvP model, we obtain the following autonomous four-dimensional ordinary differential equation.

{x˙=c(x+x3/3+yr)+q1y˙=(x+a+by)/cq1˙=(ρ+c(1x2))q1+1cq2q2˙=cq1(ρbc)q2\displaystyle\begin{cases}\dot{x}=c(x+x^{3}/3+y-r)+q_{1}\\ \dot{y}=-(x+a+by)/c\\ \dot{q_{1}}=-(\rho+c(1-x^{2}))q_{1}+\frac{1}{c}q_{2}\\ \dot{q_{2}}=-cq_{1}-(\rho-\frac{b}{c})q_{2}\end{cases} (5)

For example, if we give the initial point (x0,y0)=(1.35,0.26)(x_{0},y_{0})=(1.35,-0.26), (q1(0),q2(0))=(0.0,5.0)(q_{1}(0),q_{2}(0))=(0.0,-5.0) and ρ=2.5\rho=2.5, we can observe numerically in the figure 1 that the point near the stable limit cycle goes to the stable fixed point.

Refer to caption
Figure 1. The orbit of (x(t),y(t))(x(t),y(t)) for the initial point (x0,y0)=(1.35,0.26)(x_{0},y_{0})=(1.35,-0.26) and (q1(0),q2(0))=(0.0,5.0)(q_{1}(0),q_{2}(0))=(0.0,-5.0) by our discount model with the rate ρ=2.5\rho=2.5 is illustrated. The big (small) gray closed curve in the left is the stable (unstable) limit cycle for the original BvP model.

Now, we can define DρD_{\rho} for the model as

Dρ=[1+ρc(c21)24c2(cρb),1+ρc(c21)24c2(cρb)]×D_{\rho}=\left[-\sqrt{1+\frac{\rho}{c}-\frac{(c^{2}-1)^{2}}{4c^{2}(c\rho-b)}},\ \sqrt{1+\frac{\rho}{c}-\frac{(c^{2}-1)^{2}}{4c^{2}(c\rho-b)}}\right]\times\mathbb{R}

Next we estimate ρ\rho satisfying our assumption (A1)-(A5):

  • (A1)

    Since I(γs)[2,2]I(\gamma^{s})\subset[-2,2], we must choose ρ>9.2\rho>9.2. However, such ρ\rho is so large that the set of initial points becomes small and it may be difficult to capture the appropriate initial points. Moreover, Figure 1 and 2 suggest the assumption (A1) might be weakened.

  • (A2)

    Considering the equations det[ρI+DF(z)]=(ρ+c(1x2))(ρb/c)+1=0det[\rho I+D_{F}(z)]=(\rho+c(1-x^{2}))(\rho-b/c)+1=0 and g(x,y)=0g(x,y)=0, we find that the number of solution is always just two.

  • (A3)

    From Λ(0)=f(z)fxx(z)(gy2(z)+ρgy(z))<0\Lambda(0)=f(z)f_{xx}(z)(g_{y}^{2}(z)+\rho g_{y}(z))<0 and f(z)fxx(z)>0f(z)f_{xx}(z)>0, then (A3) is satisfied if ρ>gy=1/c0.333\rho>-g_{y}=1/c\fallingdotseq 0.333\cdots.

  • (A4)

    The assumption is independent of ρ\rho.

  • (A5)

    Unfortunately, for any z2\Dρz\in\mathbb{R}^{2}\backslash D_{\rho}, the matrix [ρI+DF(z)]-[\rho I+D_{F}(z)] cannot be positive definite so that this assumption cannot be satisfied. However, Remark 3.5 suggest that even if (A5) is not satisfied, the space 4\mathbb{R}^{4} can be separated if almost orbits diverge as tt\to-\infty. We can check it numerically, the assumption (A5) also might be weakened.

  • (AA)

    It is difficult to show that the assumption is satisfied. However, we expect to notice the non-trivial closed orbit in the numerical experiments if it exists.

The figure 2 describes classifications of initial points of control variables (q1,q2)(q_{1},q_{2}) [5,5]2\in[-5,5]^{2} for the several initial points (x0,y0)(x_{0},y_{0}) in the case ρ=2.5\rho=2.5 which satisfied all assumptions. The orbit starting from the black, white and gray region goes to the stable fixed point, goes to the stable limit cycle and diverge respectively.

Refer to caption
Figure 2. The classifications of (q1(q_{1},q2)q_{2}) planes are illustrated for some (x0,y0)(x_{0},y_{0}). If (q1(0),q2(0))(q_{1}(0),q_{2}(0)) is chosen in the black, white and gray region, the orbit (x(t),y(t))(x(t),y(t)) goes to the stable fixed point, goes to the stable limit cycle and diverges respectively. The initial point (x0,y0)(x_{0},y_{0}) is (a) (-1.66,0.42), (b) (-1.0,1.23), (c) (0.5,1.27), (d) (-0.62,0.04), (e) (1.98,0.93), (f) (0.5,-0.22), (g) (1.35,-0.26), (h) (1.73,0.25).
Remark 4.1.

In the equation (\ast), although we give the control variables q1q_{1} to the only xx direction, we naturally can consider the system which has a second control variable q2q_{2} into the yy direction as follows:

{z˙=F(z)+qq˙=[ρI+DF(z)]qwithq=(q1,q2).\displaystyle\begin{cases}\dot{z}=F(z)+q\\ \dot{q}=-[\rho I+D_{F}(z)]q\end{cases}\ \ \ \text{with}\ \ \ q=(q_{1},q_{2}). (6)

that is,

{x˙=f(x,y)+q1y˙=g(x,y)+q2q1˙=(ρ+fx(x,y))q1gx(x,y)q2q2˙=fy(x,y)q1(ρ+gy(x,y))q2\displaystyle\begin{cases}\dot{x}=f(x,y)+q_{1}\\ \dot{y}=g(x,y)+q_{2}\\ \dot{q_{1}}=-(\rho+f_{x}(x,y))q_{1}-g_{x}(x,y)q_{2}\\ \dot{q_{2}}=-f_{y}(x,y)q_{1}-(\rho+g_{y}(x,y))q_{2}\end{cases}

We can also prove Theorem A and Theorem B for this system if all assumptions are hold. However, it might be more difficult to satisfy the assumptions. For example, the characteristic polynomial becomes

Λ(λ)\displaystyle\Lambda(\lambda) :=\displaystyle:= det[λIDF(z,q)]\displaystyle det[\lambda I-D_{F}(z,q)] (7)
=\displaystyle= λ4+2ρλ3{ρ2+3(fx+gy)ρ+(fx+gy)2(a+d)}λ2\displaystyle\lambda^{4}+2\rho\lambda^{3}-\{\rho^{2}+3(f_{x}+g_{y})\rho+(f_{x}+g_{y})^{2}-(a+d)\}\lambda^{2}
{2ρ2+3ρ(fx+gy)+(fx+gy)2(a+d)}ρλB~(z,q),\displaystyle-\{2\rho^{2}+3\rho(f_{x}+g_{y})+(f_{x}+g_{y})^{2}-(a+d)\}\rho\lambda-\tilde{B}(z,q),

where

B~(z,q)\displaystyle\tilde{B}(z,q) =\displaystyle= (fxfy)(h4h2h3h1)(fxfy)(gxgy)(h4h2h3h1)(gxgy)\displaystyle-\begin{pmatrix}f_{x}&f_{y}\end{pmatrix}\begin{pmatrix}h_{4}&-h_{2}\\ -h_{3}&h_{1}\end{pmatrix}\begin{pmatrix}f_{x}\\ f_{y}\end{pmatrix}-\begin{pmatrix}g_{x}&g_{y}\end{pmatrix}\begin{pmatrix}h_{4}&-h_{2}\\ -h_{3}&h_{1}\end{pmatrix}\begin{pmatrix}g_{x}\\ g_{y}\end{pmatrix}
ρ(h1gyh2gxh3fy+h4fx)(h1h4h2h3),\displaystyle-\rho(h_{1}g_{y}-h_{2}g_{x}-h_{3}f_{y}+h_{4}f_{x})-(h_{1}h_{4}-h_{2}h_{3}),

then the relation B~(z~,q~)<0\tilde{B}(\tilde{z},\tilde{q})<0 must be hold instead of assumption (A3). For the BvP model, by using the equation (7), ρ\rho must be satisfied ρ>c3+1/c27.33\rho>c^{3}+1/c\fallingdotseq 27.33\cdots. Therefore we find that using the system (\ast) is better than the system (6). Note that if we consider the control variables to the only yy direction, we can show that Λ(0)\Lambda(0) is always positive so that the assumption (A3) cannot be hold. This implies that the initial points cannot be classified by the separatrix Ws(z~,q~)W^{s}(\tilde{z},\tilde{q}).

Appendix A Infinite horizon with discounting

Consider the infinite horizon optimal control problem described by

{Minimize0|p(t)|2/2𝑑tsubject toz˙(t)=F(z(t))+p(t),z(0)=z0,limtz(t)=z\displaystyle\begin{cases}\text{Minimize}&\int_{0}^{\infty}|p(t)|^{2}/2dt\\ \text{subject to}&\dot{z}(t)=F(z(t))+p(t),\ \ \ \displaystyle z(0)=z_{0},\lim_{t\to\infty}z(t)=z_{\infty}\end{cases} (8)

where F:nnF:\mathbb{R}^{n}\to\mathbb{R}^{n} is a smooth function, z(t),p(t)nz(t),p(t)\in\mathbb{R}^{n} are vector valued functions and z0,znz_{0},z_{\infty}\in\mathbb{R}^{n}. The problem is to find the optimal control function p(t)p(t) to make the state z(t)z(t) to move from the initial state z0z_{0} to the terminal state zz_{\infty} under the differential equation z˙(t)=F(z(t))+p(t)\dot{z}(t)=F(z(t))+p(t). Define the functional called performance function or Lagrangian by

(z,p,μ)=|p|2/2+μT(z˙F(z)p)\mathcal{L}(z,p,\mu)=|p|^{2}/2+\mu^{T}(\dot{z}-F(z)-p)

where the vector valued function μ(t)n\mu(t)\in\mathbb{R}^{n} is as a Lagrange multiplier and xTx^{T} denotes the transpose of vector xx. By the usual variational principle, we can obtain the following system on 2n\mathbb{R}^{2n},

{z˙=F(z)+pp˙=DF(z)Tp\displaystyle\begin{cases}\dot{z}=F(z)+p\\ \dot{p}=-D_{F}(z)^{T}p\end{cases} (9)

where DF(x)D_{F}(x) is the Jacobian matrix of FF for zz and ATA^{T} denotes the transpose of the matrix AA.

Now we will focus on the stability problem for a fixed point of the system (9). Assume that the system z˙=f(z)\dot{z}=f(z) has a fixed point zz_{*} which have non-zero eigenvalues {λi}i=1n\{\lambda_{i}\}_{i=1}^{n} of the Jacobian matrix DF(z)D_{F}(z_{*}). Then we immediately find that (z,0)2n(z_{*},0)\in\mathbb{R}^{2n} becomes the fixed point of the system (9) on 2n\mathbb{R}^{2n} whose eigenvalues are {±λi}i=1n\{\pm\lambda_{i}\}_{i=1}^{n} which implies the point (z,0)(z_{*},0) becomes always a saddle. This suggests that the control variable p(t)p(t) may extremely increase or diverge even if z(t)z(t) goes to zz_{*} for sufficiently large tt.

To avoid the trouble, we propose the optimal control problem with a discount parameter ρ~\tilde{\rho} as follows:

{Minimize0eρ~t|p(t)|2/2𝑑tsubject toz˙(t)=F(z(t))+p(t),z(0)=z0,limtz(t)=z.\displaystyle\begin{cases}\text{Minimize}&\int_{0}^{\infty}e^{-\tilde{\rho}t}|p(t)|^{2}/2dt\\ \text{subject to}&\dot{z}(t)=F(z(t))+p(t),z(0)=z_{0},\displaystyle\lim_{t\to\infty}z(t)=z_{\infty}.\end{cases} (10)

Give the Lagrangian with a discount term eρ~te^{-\tilde{\rho}t} by

(z,p,μ)=eρ~t|p|2/2+μT(z˙f(z)p),\mathcal{L}(z,p,\mu)=e^{-\tilde{\rho}t}|p|^{2}/2+\mu^{T}(\dot{z}-f(z)-p),

and by the variational principle, we have

{z˙=F(z)+peρ~tp˙=DF(z)Tp.\displaystyle\begin{cases}\dot{z}=F(z)+pe^{\tilde{\rho}t}\\ \dot{p}=-D_{F}(z)^{T}p.\end{cases}

Changing the variables q(t)=p(t)eρ~tq(t)=p(t)e^{\tilde{\rho}t} leads the autonomous system on 2n\mathbb{R}^{2n},

{z˙=F(z)+qq˙=ρ~qDf(z)Tq.\displaystyle\begin{cases}\dot{z}=F(z)+q\\ \dot{q}=\tilde{\rho}q-D_{f}(z)^{T}q.\end{cases} (11)

The eigenvalues of the Jacobian matrix for the fixed point (z,0)(z_{*},0) for the system (11) become λi\lambda_{i} and λi+ρ~-\lambda_{i}+\tilde{\rho}. Thus, when λi<0\lambda_{i}<0 for all ii, the point (x,0)(x_{*},0) can become stable fixed point by taking ρ<0\rho<0 satisfying λiρ~<0-\lambda_{i}-\tilde{\rho}<0 for all ii. By substituting ρ=ρ~\rho=-\tilde{\rho}, we obtain the system (\ast). Therefore, we can expect that it is possible to realize the stable control by considering infinite horizon optimal control problem with “negative” discount.

Note that, by Legendre transformation, the system (11) can be rewritten as

{z˙=Hqq˙=Hz+ρ~q\displaystyle\begin{cases}\dot{z}=\frac{\partial H}{\partial q}\\ \dot{q}=-\frac{\partial H}{\partial z}+\tilde{\rho}q\end{cases} (12)

where the Hamiltonian HH is given by

H(z,q,μ):=eρ~t|q|2/2+μT(F(z)+q)\displaystyle H(z,q,\mu):=e^{-\tilde{\rho}t}|q|^{2}/2+\mu^{T}(F(z)+q)

Remark that we do not know generally whether the system is a Hamilton system if ρ0\rho\neq 0. In the case ρ\rho is positive, it is known that the integral in (10) converges as TT\to\infty for a bounded p(t)p(t) (See [11]). On the other hand, for the negative rate ρ\rho, the integral in (10) does not converge generally. Although the infinite horizon problem with positive discount rate is well discussed in many previous works, for example [12, 11, 13], for the negative discount rate, it has not developed enough due to the difficulty of the nonlinearity.

Appendix B For one dimensional control problems

Our discount model applied to one-dimensional autonomous system x˙=f(x)\dot{x}=f(x) can be calculated by elementary methods since we know a non-existence of closed orbit by Poincare-Bendixson theorem in two-dimensional systems. Because these calculus help us to understand arguments for the higher dimensional model, we summarise in the appendix section.

Let f:f:\mathbb{R}\to\mathbb{R} be a smooth map and ff has more than two stable fixed points. We denote the sets of stable and unstable fixed points by FPs:={x1s,x2s,,xN+1s}FP^{s}:=\{x_{1}^{s},x_{2}^{s},\cdots,x_{N+1}^{s}\} and FPu:={x1u,x2u,,xNu}FP^{u}:=\{x_{1}^{u},x_{2}^{u},\cdots,x_{N}^{u}\}. Consider the two-dimensional system with discount as follows:

{x˙=f(x)+qq˙=(ρ+f(x))q\displaystyle\begin{cases}\dot{x}=f(x)+q\\ \dot{q}=-(\rho+f^{\prime}(x))q\end{cases} (13)

Let Dρ:={x|ρ+f(x)>0}D_{\rho}:=\{x\in\mathbb{R}\ |\ \rho+f^{\prime}(x)>0\}. Non-trivial fixed points of (13) can be calculated by

ρ+f(x)=0,f(x)+q=0.\displaystyle\rho+f^{\prime}(x)=0,\ \ \ f(x)+q=0. (14)

We assume the followings:

  • (A1)’

    FPs,FPuDρFP^{s},FP^{u}\in D_{\rho}.

  • (A2)’

    #{x|ρ+f(x)=0}=2\#\{x\in\mathbb{R}\ |\ \rho+f^{\prime}(x)=0\}=2.

  • (A3)’

    f(x)f′′(x)>0f(x)f^{\prime\prime}(x)>0 in \Dρ\mathbb{R}\backslash D_{\rho},

  • (A4)’

    f(x)f^{\prime}(x)\to-\infty if |x||x|\to\infty

For one-dimensional model, we usually give a Lyapunov function as V(x)=f(x)V(x)=-f^{\prime}(x). Then, V(x)>0V(x)>0 and V˙(x(t))<0\dot{V}(x(t))<0 in \\mathbb{R}\backslash by (A2)’ and (A3)’. Therefore the assumption (A4) is satisfied by (A2)’ and (A3)’ together with (A4)’. Finally, the assumption (A5) is automatically satisfied by taking K=DρK=D_{\rho} in the case of one-dimensional model. Then, we can show Theorem B.3 under the assumption (A1)’-(A4)’.

The Jacobian matrix for the system FF is given by

DF(x,q)=(f(x)1f(x)q(ρ+f(x)))D_{F}(x,q)=\begin{pmatrix}f^{\prime}(x)&1\\ -f^{\prime}(x)q&-(\rho+f^{\prime}(x))\end{pmatrix}
Lemma B.1.

If xFPsx_{*}\in FP^{s}, then (x,0)(x_{*},0) becomes a stable fixed point of FF . If xFPux_{*}\in FP^{u}, then (x,0)(x_{*},0) becomes a saddle.

Proof.

For the fixed point (x,0)(x_{*},0), we have the Jacobian matrix is

DF(x,0)=(f(x)10(ρ+f(x))),D_{F}(x_{*},0)=\begin{pmatrix}f^{\prime}(x_{*})&1\\ 0&-(\rho+f^{\prime}(x_{*}))\end{pmatrix},

and the eigenvalues of the matrix are f(x)f^{\prime}(x_{*}) and f(x)ρ-f^{\prime}(x_{*})-\rho. If xFPux_{*}\in FP^{u}, then (x,0)(x_{*},0) becomes a saddle since f(x)ρ<0-f^{\prime}(x_{*})-\rho<0. If xFPsx_{*}\in FP^{s}, then (x,0)(x_{*},0) becomes stable from the assumption (A1).

Lemma B.2.

There are two non-trivial fixed points of FF satisfying (14), which are saddles.

Proof.

From the assumption (A2)’, we have only two non-trivial fixed points of FF. For the fixed point (x~,q~)(\tilde{x},\tilde{q}), the Jacobian matrix is

DF(x^,q^)=(ρ1f′′(x^)f(x^)0),D_{F}(\hat{x},\hat{q})=\begin{pmatrix}-\rho&1\\ f^{\prime\prime}(\hat{x})f(\hat{x})&0\end{pmatrix},

and the eigenvalues of the matrix are (ρ±ρ2+4f′′(x^)f(x^))/2(-\rho\pm\sqrt{\rho^{2}+4f^{\prime\prime}(\hat{x})f(\hat{x})})/2. Therefore (x^,q^)(\hat{x},\hat{q}) must be saddle from (A4). ∎

Theorem B.3.

Assume that (A1)’-(A4)’ hold. Then, there exist (N+2)(N+2) disjoint simply connected subset A0,A1,,AN+12A_{0},A_{1},\cdots,A_{N+1}\subset\mathbb{R}^{2} such that cl(A0AN+1)=2{\rm cl}(A_{0}\cup\cdots\cup A_{N+1})=\mathbb{R}^{2} and

limtΦt(x0,q0)={(xis,0) for (x0,q0)Ai(i=1,,N) for (x0,q0)A0AN+1.\lim_{t\to\infty}\Phi^{t}(x_{0},q_{0})=\begin{cases}(x_{i}^{s},0)&\text{ for }(x_{0},q_{0})\in A_{i}\ \ (i=1,\cdots,N)\\ \infty&\text{ for }(x_{0},q_{0})\in A_{0}\cup A_{N+1}.\end{cases}
Refer to caption
Figure 3. The illustration for one-dimensional control problems. The four stable manifolds of each saddle point, (x1u,0),(x2u,0),(x~L,q~L)(x_{1}^{u},0),(x_{2}^{u},0),(\tilde{x}_{L},\tilde{q}_{L}) and (x~R,q~R)(\tilde{x}_{R},\tilde{q}_{R}), separate 2\mathbb{R}^{2}.
Proof.

Since DρD_{\rho} is an open interval set, denote Dρ=(xL,xR)D_{\rho}=(x_{L},x_{R}). First we show that Ws(x~,q~)W^{s}(\tilde{x},\tilde{q}) with x~=xR\tilde{x}=x_{R} separates 2\mathbb{R}^{2} to two connected sets and is their boundary. Since (x~,q~)(\tilde{x},\tilde{q}) is a saddle, the neighbor of the fixed point contains sets of initial points from which the orbit converges to (x~,q~)(\tilde{x},\tilde{q}) starting. If the initial point (x0,q0)(x_{0},q_{0}) in the neighbor satisfies q0+f(x0)>0q_{0}+f(x_{0})>0, then (x0,q0)Dρ(x_{0},q_{0})\in D_{\rho} and q˙>0\dot{q}>0 which mean x(t)x(t) decreases and q(t)q(t) increases monotonically as tt\to-\infty until x(t)x(t) reaches xLx_{L}. After that, q˙\dot{q} becomes negative, so that, q(t)0q(t)\to 0 and x(t)x(t)\to-\infty as tt\to-\infty since the line {(x,q):x<xL,q=0}\{(x,q):\ x<x_{L},q=0\} becomes a part of separatrixes. If the initial point (x0,q0)(x_{0},q_{0}) in the neighbor satisfies q0+f(x0)<0q_{0}+f(x_{0})<0, then q(t)0q(t)\to 0 and x(t)x(t)\to\infty as tt\to-\infty by similar arguments. Therefore, the set Ws(x~,q~)W^{s}(\tilde{x},\tilde{q}) separates 2\mathbb{R}^{2} to two connected sets. (More rigorously, we may need the theorem 4.6 in [17] or discussions of one-point compactification and Jordan-Brouwer separation theorem.)

By the same arguments, we can see that the all stable manifolds Ws(x~,q~)W^{s}(\tilde{x},\tilde{q}) for x~=xL\tilde{x}=x_{L} and Ws(xis,0)W^{s}(x_{i}^{s},0) for i=1,,N+1i=1,\cdots,N+1 also separate 2\mathbb{R}^{2}. Moreover, since they have no intersection each other, the stable manifolds Ws(x~1,q~1)W^{s}(\tilde{x}_{1},\tilde{q}_{1}), Ws(x~2,q~2)W^{s}(\tilde{x}_{2},\tilde{q}_{2}) and Ws(xis,0)W^{s}(x_{i}^{s},0) separate 2\mathbb{R}^{2} to N+2N+2 regions {Ai}i=0N+1\{A_{i}\}_{i=0}^{N+1}. We can obviously name {Ai}i=0N+1\{A_{i}\}_{i=0}^{N+1} satisfy (i)-(iii) of the theorem.

Finally, each set AiA_{i}, i=1,,Ni=1,\cdots,N, has only unique stable fixed points xisx_{i}^{s} in the inside, and the orbit cannot move to other AjA_{j} since all stable manifolds are separatrixes. The orbit clearly does not diverge and there is no non-trivial attractor in each AiA_{i} since there is no fixed points except with (x~1,q~1)(\tilde{x}_{1},\tilde{q}_{1}), (x~2,q~2)(\tilde{x}_{2},\tilde{q}_{2}) and (xis,0)(x_{i}^{s},0). Therefore, the orbit Φy(x0,q0)\Phi^{y}(x_{0},q_{0}) with (x0,q0)Ai(x_{0},q_{0})\in A_{i} must be converge to (xis,0)(x_{i}^{s},0) as tt\to\infty. For (x0,q0)A0AN+1(x_{0},q_{0})\in A_{0}\cup A_{N+1}, the orbit must diverge since A0A_{0} and AN+1A_{N+1} have no attractor in its inside. The proof is completed.

Acknowledgement

I am deeply grateful to Prof. Ichiro Tsuda (Chubu university) and Prof. Hideo Kubo (Hokkaido university) for constructive comments and warm encouragement. Moreover, I would like to thank Prof. Atsuro Sannami (Kitami institute of Technology), Prof. Okihiro Sawada (Kitami institute of Technology) and Prof. Tomoo Yokoyama (Kyoto university of Education) to give me insightful comments and suggestion. Finally, this work is supported by JST CREST Grant Number JPMJCR17A4, Japan.

References

  • [1] Brock, William A., and Jose A. Scheinkman. ”Global asymptotic stability of optimal control systems with applications to the theory of economic growth.” The Hamiltonian approach to dynamic economics. Academic Press, 1976. 164-190.
  • [2] Haurie, Alain. “Existence and global asymptotic stability of optimal trajectories for a class of infinite-horizon, nonconvex systems.” Journal of Optimization Theory and Applications 31.4 (1980): 515-533.
  • [3] Gaitsgory, Vladimir, Lars Grüne, and Neil Thatcher. “Stabilization with discounted optimal control.” Systems & Control Letters 82 (2015): 91-98.
  • [4] Barnes, Belinda, and Roger Grimshaw. “Analytical and numerical studies of the Bonhoeffer van der Pol system.” The ANZIAM Journal 38.4 (1997): 427-453.
  • [5] Forger, Daniel B., and David Paydarfar. “Starting, stopping, and resetting biological oscillators: in search of optimum perturbations.” Journal of theoretical biology 230.4 (2004): 521-532.
  • [6] Chang, Joshua, and David Paydarfar. “Switching neuronal state: optimal stimuli revealed using a stochastically-seeded gradient algorithm.” Journal of computational neuroscience 37.3 (2014): 569-582.
  • [7] Bryson, Arthur Earl. Applied optimal control: optimization, estimation and control. Routledge, 2018.
  • [8] Aseev, Sergey M., and Vladimir M. Veliov. Maximum principle for infinite-horizon optimal control problems with dominating discount. na, 2012.
  • [9] Tauchnitz, Nico. “The Pontryagin maximum principle for nonlinear optimal control problems with infinite horizon.” Journal of Optimization Theory and Applications 167.1 (2015): 27-48.
  • [10] de Oliveira, Valeriano Antunes, and Geraldo Nunes Silva. “A note on the sufficiency of the maximum principle for infinite horizon optimal control problems.” Optimal Control Applications and Methods 39.4 (2018): 1573-1580.
  • [11] Carlson, Dean A., Alain B. Haurie, and Arie Leizarowitz. Infinite horizon optimal control: deterministic and stochastic systems. Springer Science & Business Media, 2012.
  • [12] Kamien, Morton I., and Nancy Lou Schwartz. Dynamic optimization: the calculus of variations and optimal control in economics and management. Courier Corporation, 2012.
  • [13] Aseev, Sergei M., and Arkady V. Kryazhimskiy. “The Pontryagin maximum principle and transversality conditions for a class of optimal control problems with infinite time horizons.” SIAM Journal on Control and Optimization 43.3 (2004): 1094-1119.
  • [14] Smith, Russell A. “Some applications of Hausdorff dimension inequalities for ordinary differential equations.” Proceedings of the Royal Society of Edinburgh Section A: Mathematics 104.3-4 (1986): 235-259.
  • [15] Robinson, Clark. Dynamical systems: stability, symbolic dynamics, and chaos. CRC press, 1998.
  • [16] Teschl, Gerald. Ordinary differential equations and dynamical systems. Vol. 140. American Mathematical Soc., 2012.
  • [17] Hirsch, Morris W. Differential topology. Vol. 33. Springer Science & Business Media, 2012.