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

Morse index and determinant of block Jacobi matrices via optimal control

Stefano Baranzini 111Università degli studi di Torino, Via Carlo Alberto 10, Torino, Italy    Ivan Beschastnyi 222Univesidade de Aveiro, Campus Santiago, Aveiro, Portugal
Abstract

We describe the relation between block Jacobi matrices and minimization problems for discrete time optimal control problems. Using techniques developed for the continuous case, we provide new algorithms to compute spectral invariants of block Jacobi matrices. Some examples and applications are presented.

1 Introduction

1.1 Motivation

The goal of this paper is to explore an interesting connection between block Jacobi matrices and a class of discrete optimal control problems. This gives effective algorithms for computing the determinant, the negative inertia index and more generally the number of eigenvalues smaller than some threshold λ\lambda^{*}\in\mathbb{R} of large block Jacobi matrices.

Recall that a block Jacobi matrix \mathcal{I} is a matrix of the form

=(S1R1000R1tS2R2000R2tS300000SN1RN1000RN1tSN),{\mathcal{I}}=\begin{pmatrix}S_{1}&R_{1}&0&\dots&0&0\\ R_{1}^{t}&S_{2}&R_{2}&\dots&0&0\\ 0&R_{2}^{t}&S_{3}&\dots&0&0\\ \vdots&\vdots&\vdots&\ddots&\vdots&\vdots\\ 0&0&0&\dots&S_{N-1}&R_{N-1}\\ 0&0&0&\dots&R^{t}_{N-1}&S_{N}\end{pmatrix}, (1)

where NN\in\mathbb{N}, SkS_{k} are symmetric matrices of order nn. Usually it is also assumed that RkR_{k} are positive symmetric matrices of the same order. Here, we only require them to be non-degenerate, i.e. RkGL(n,)R_{k}\in GL(n,\mathbb{R}).

Jacobi matrices find applications in numerical analysis [11], statistical physics [12], knot theory [8] and many other areas. Moreover, any symmetric matrix can be put in a tridiagonal form using the Householder method [7]. Therefore, understanding their spectral properties is an extremely important topic.

In this article we establish a correspondence between Jacobi matrices and discrete linear quadratic regulator (LQR) problem. Then we show how the general machinery of optimal control theory allows us to compute the index and the determinant of such matrices. Similar formulas for determinants also appeared in [15], see also [16, 17] for a geometric approach to the problem.

The LQR approach nevertheless has its benefits. In particular, as a small application of our method we give a novel proof of the generalized Euler identity

asinbbsinha=j=1(1a2+b2a2+(πj)2)\frac{a\sin b}{b\sinh a}=\prod_{j=1}^{\infty}\left(1-\frac{a^{2}+b^{2}}{a^{2}+(\pi j)^{2}}\right) (2)

using only methods of linear algebra and basic analysis.

In order to state the main results, let us introduce the necessary notations. Consider the following discrete differential equation:

xk+1xk=Akxk+Bkuk,k=0,,N+1,x_{k+1}-x_{k}=A_{k}x_{k}+B_{k}u_{k},\qquad k=0,\dots,N+1, (3)

Where AkA_{k} and BkB_{k} are n×nn\times n matrix, BkB_{k} are invertible and xk,uknx_{k},u_{k}\in\mathbb{R}^{n}. Variables xkx_{k} are called state variables while uku_{k} are the controls. We denote by x=(xk)x=(x_{k}) and u=(uk)u=(u_{k}) the vectors having xkx_{k} and uku_{k} respectively as kk-th component. Note that we will often consider them as elements of the space of functions on a finite set of points with values in n\mathbb{R}^{n}, clearly isomorphic to n(N+2)\mathbb{R}^{n(N+2)} and n(N+1)\mathbb{R}^{n(N+1)} correspondingly. On the space of solutions of eq. 3 with fixed initial condition x0x_{0} we consider the following quadratic functional given by:

𝒥(u)=12i=0N(|ui|2Qixi,xi),\mathcal{J}(u)=\frac{1}{2}\sum_{i=0}^{N}\left(|u_{i}|^{2}-\langle Q_{i}x_{i},x_{i}\rangle\right), (4)

where QiQ_{i} are some symmetric n×nn\times n matrices.

The problem of minimizing 𝒥\mathcal{J} over all possible discrete trajectories of (3) is the classical linear quadratic regulator problem, which is one of the central problems in optimal control theory. Since we assume that BkB_{k} are invertible, we can resolve the constraints (3) with respect to uku_{k} and plug them into (4). This yields

𝒥(x)=12(i=0N|Bi1(xi+1(1+Ai)xi)|2i=0NQixi,xi).\mathcal{J}(x)=\frac{1}{2}\left(\sum_{i=0}^{N}|B_{i}^{-1}(x_{i+1}-(1+A_{i})x_{i})|^{2}-\sum_{i=0}^{N}\langle Q_{i}x_{i},x_{i}\rangle\right).

If we assume additionally that x0=0=xN+1x_{0}=0=x_{N+1} and write down the matrix associated to the quadratic form above, we find that it is given by a block Jacobi matrix (1). Indeed, set

Γk1:=(Bk1)tBk1,\Gamma_{k}^{-1}:=(B_{k}^{-1})^{t}B_{k}^{-1},

which is a symmetric non degenerate matrix. Then the functional 2𝒥2\mathcal{J} is the represented by the symmetric matrix (1) with:

Sk\displaystyle S_{k} =Γk11+(1+Ak)tΓk1(1+Ak)Qk,\displaystyle=\Gamma_{k-1}^{-1}+(1+A_{k})^{t}\Gamma_{k}^{-1}(1+A_{k})-Q_{k},
Rk\displaystyle R_{k} =(1+Ak)tΓk1,\displaystyle=-(1+A_{k})^{t}\Gamma_{k}^{-1}, (5)
1kN1.\displaystyle 1\leq k\leq N-1.

We now see that to each LQR problem (3)-(4) we can associate a Jacobi matrix. The converse is also true. However, there are several LQR problems which correspond to the same Jacobi matrix. We can get rid of this ambiguity by choosing fixed invertible BkB_{k}, 0kn0\leq k\leq n and an arbitrary matrix A0A_{0}. Thus to every LQR problem (3)-(4) one can associate a Jacobi matrix and vice versa. Fixing BkB_{k}, 0kn0\leq k\leq n and A0A_{0} beforehand makes this correspondence an affine bijection.

1.2 Main results

Through out the paper we will make the following assumption. It is equivalent to requiring RkGL(n,)R_{k}\in GL(n,\mathbb{R}) in (5).

Assumption 1.

The matrices BkB_{k} and 1+Ak1+A_{k} in (3) are invertible for every kk.

Similarly to the case of continuous optimal control problems, one can characterize the extremal points of the functional (4) under the constraints (3) as solutions to a system of difference equations. To be more precise define the following matrices for 0ijN+10\leq i\leq j\leq N+1

Pij:=r=ij1(1+Ar),Pjj=1.P_{i}^{j}:=\prod_{r=i}^{j-1}(1+A_{r}),\quad P_{j}^{j}=1.

and for 0i<jN+10\leq i<j\leq N+1 we take:

Pji=(Pij)1.P_{j}^{i}=(P_{i}^{j})^{-1}.

Next, we define 2n×2n2n\times 2n matrices:

Mk=((Pk+1k)t0Γk(Pk+1k)tPkk+1)(1Qk01)\displaystyle M_{k}=\begin{pmatrix}(P^{k}_{k+1})^{t}&0\\ \Gamma_{k}(P^{k}_{k+1})^{t}&P^{k+1}_{k}\end{pmatrix}\begin{pmatrix}1&-Q_{k}\\ 0&1\end{pmatrix}

and

Φk=(Φk1Φk2Φk3Φk4):=i=0k1Mi.\Phi_{k}=\begin{pmatrix}\Phi^{1}_{k}&\Phi^{2}_{k}\\ \Phi^{3}_{k}&\Phi^{4}_{k}\end{pmatrix}:=\prod_{i=0}^{k-1}M_{i}.

Matrices Φk\Phi_{k} correspond to the flow of the extremal equations that we will derive in Section 2 using the Lagrange multiplier rule. Now we can formulate the two main results of this paper.

Theorem 1.

Let 𝒥\mathcal{J} be as in (4) and consider constraints as in (3). The negative inertia index of 𝒥\mathcal{J} restricted to solutions of (3) with x0=xN+1=0x_{0}=x_{N+1}=0 is given by the following formula:

ind𝒥=k=1Nind𝒬k+dimkerΦk3,ind^{-}\mathcal{J}=\sum_{k=1}^{N}ind^{-}\mathcal{Q}_{k}^{\perp}+\dim\ker\Phi^{3}_{k}, (6)

where

𝒬k=(Φk3)t((Pkk+1)tΓk1(Pkk+1)Qk)Φk3+(Φk3)tΦk1\mathcal{Q}_{k}^{\perp}=(\Phi_{k}^{3})^{t}\left((P^{k+1}_{k})^{t}\Gamma_{k}^{-1}(P^{k+1}_{k})-Q_{k}\right)\Phi_{k}^{3}+(\Phi_{k}^{3})^{t}\Phi_{k}^{1}

Notice that we have reduced the problem of computing the negative inertia index of a nN×nNnN\times nN matrix to computing the inertia of NN matrices of dimension n×nn\times n, which greatly reduces the dimensionality of the problem. In the same spirit, we also prove a formula for the determinant.

Theorem 2.

Let 𝒥\mathcal{J} be as in (4) and consider constraints as in (3). The determinant of 𝒥\mathcal{J} restricted to solutions of (3) with x0=xN+1=0x_{0}=x_{N+1}=0 is given by the following formula:

det(𝒥)=2nNdet(Φ3)det(P0N+1)k=0NdetΓk1.\det(\mathcal{J})=2^{-nN}\det(\Phi_{3})\det(P_{0}^{N+1})\prod_{k=0}^{N}\det\Gamma_{k}^{-1}. (7)

Let us rephrase the two statements in terms of Jacobi matrices of the form (1). We can choose A0=0A_{0}=0 and Γk=1\Gamma_{k}=1 for all 0kN0\leq k\leq N and obtain the following relations from eq. 5 for 1kN11\leq k\leq N-1:

{Qk=1Sk+RkRkt,Pkk+1=Rkt,\begin{cases}Q_{k}=1-S_{k}+R_{k}R^{t}_{k},\\ P_{k}^{k+1}=-R_{k}^{t},\end{cases}
Mk=(Rk1Rk1(1Sk+RkRkt)Rk1Rk1(1Sk)).M_{k}=\begin{pmatrix}-R^{-1}_{k}&R_{k}^{-1}(1-S_{k}+R_{k}R_{k}^{t})\\ -R_{k}^{-1}&R_{k}^{-1}(1-S_{k})\end{pmatrix}.
Corollary 1.

Index and determinant of the matrix \mathcal{I} defined in eq. 1 can be computed as:

ind\displaystyle ind^{-}\mathcal{I} =k=1Nind𝒬k+dimkerΦk3,\displaystyle=\sum_{k=1}^{N}ind^{-}\mathcal{Q}_{k}^{\perp}+\dim\ker\Phi^{3}_{k},
det()\displaystyle\det(\mathcal{I}) =(1)nNdet(Φ3)k=1NdetRk.\displaystyle=(-1)^{nN}\det(\Phi_{3})\prod_{k=1}^{N}\det R_{k}.

where 𝒬k=(Φk3)t(Sk1)Φk3+(Φk3)tΦk1\mathcal{Q}_{k}^{\perp}=(\Phi_{k}^{3})^{t}\left(S_{k}-1\right)\Phi_{k}^{3}+(\Phi_{k}^{3})^{t}\Phi_{k}^{1}.

Remark 1.

Clearly, applying formula (6) with Q~k=Qk+λ\tilde{Q}_{k}=Q_{k}+\lambda^{*} instead of QkQ_{k} (or equivalently S~k=Skλ\tilde{S}_{k}=S_{k}-\lambda^{*} instead of SkS_{k}), gives a formula to compute the number of eigenvalues of 𝒥\mathcal{J} (respectively of \mathcal{I}) smaller than λ\lambda^{*}\in\mathbb{R}. For some oscillation results in this setting, see [5].

The techniques used in our proofs are deeply connected to symplectic geometry. One can extend them to much more general settings [3, 4, 6]. We have chosen to keep the exposition as simple as possible and minimize the use of symplectic language in this work. In this way, both formulas (6) and (7) can be applied directly.

The article has the following structure. In Section 2 we derive the Hamiltonian system for the extremal curves using the end-point map and Lagrange multiplier rule. In Section 3 we prove Theorem 1 and in Section 4 Theorem 2. Finally, in Section 5, we consider some examples and prove the generalized Euler identity (2).

2 Eigenvalues of Jacobi matrices and the Lagrange multiplier rule

In order to compute the index and the determinant of a Jacobi matrix {\mathcal{I}} in (1) we will use the flow of a certain system of difference equations. First, though, we need some notions from optimal control theory.

The first object we introduce is the endpoint map:

Ek+1:uxk+1(u),0kN,E^{k+1}:u\mapsto x_{k+1}(u),\qquad 0\leq k\leq N,

This map takes a control uu and gives the solution of the differential equation  (3) with x0=0x_{0}=0 at step k+1{k+1}. We can write this map explicitly by iterating (3).

Ek+1(u)=xk+1(u0,,uk)=j=1k+1Pjk+1Bj1uj1.E^{k+1}(u)=x_{k+1}(u_{0},\dots,u_{k})=\sum_{j=1}^{k+1}P_{j}^{k+1}B_{j-1}u_{j-1}. (8)

In particular, from this formula, it is clear that the value of Ek+1(u)E^{k+1}(u) depends only on the first kk components of the control uu. It makes sense, thus, to introduce the following filtration (i.e. flag of subspaces) in the space of controls. Set U:=nNU:=\mathbb{R}^{nN} the space of controls and define:

Uk:={(u0,u1,,uk,0,,0):ujn}.U_{k}:=\{(u_{0},u_{1},\dots,u_{k},0,\dots,0):u_{j}\in\mathbb{R}^{n}\}. (9)

Denote by prlpr_{l} the orthogonal projection on UlU_{l} with respect to the standard Euclidean scalar product. We have:

Ex0k+1(u)=Ex0k+1(prlu)lk.E_{x_{0}}^{k+1}(u)=E_{x_{0}}^{k+1}(pr_{l}u)\quad\forall\,l\geq k.

Under 1, we can resolve the constraints ukerEk+1Uku\in\ker E^{k+1}\cap U_{k} quite easily. This will play an important role in all of the proofs. Indeed, consider the map

Fk:Uk1n,F_{k}:U_{k-1}\to\mathbb{R}^{n},

defined as

Fk(u):=Bk1j=1kPjk+1Bj1uj1.F_{k}(u):=-B_{k}^{-1}\sum_{j=1}^{k}P^{k+1}_{j}B_{j-1}u_{j-1}. (10)

It is straightforward to check, using eq. 8, that:

(u0,,uk)kerEk+1Ukuk=Fk(u).(u_{0},\dots,u_{k})\in\ker E^{k+1}\cap U_{k}\iff u_{k}=F_{k}(u). (11)

We can identify FkF_{k} with a row block-matrix of size n×nkn\times nk.


As we have seen in the previous section, a Jacobi matrix can be interpreted as the functional 𝒥\mathcal{J} in eq. 4 restricted to kerEN+1\ker E^{N+1}. We would like to find the eigenvalues of this restriction, i.e. of the corresponding Jacobi matrix {\mathcal{I}}.

By definition τ\tau\in\mathbb{R} is an eigenvalue of 𝒥\mathcal{J} if and only if the bilinear form 𝒥τ|u|2\mathcal{J}-\tau|u|^{2} is degenerate on kerEN+1\ker E^{N+1}. Since the form 𝒥(u)τ|u|2\mathcal{J}(u)-\tau|u|^{2} is quadratic and kerEN+1\ker E^{N+1} is a linear space, this is equivalent to finding critical points of 𝒥(u)τ|u|2\mathcal{J}(u)-\tau|u|^{2} on kerEN+1\ker E^{N+1}. This is a typical task in constrained optimization and can be done via Lagrange multiplier rule. Notice that τ=12\tau=\frac{1}{2} is not an eigenvalue of 𝒥\mathcal{J} if and only if all QiQ_{i}, 1iN1\leq i\leq N are non degenerate.

Let us multiply 𝒥τ\mathcal{J}-\tau by (12τ)1(1-2\tau)^{-1}, operation that clearly preserves the kernel. After we introduce a new parameter s=1/(12τ)s=1/(1-2\tau) we obtain a family of functionals

𝒥s(u):\displaystyle\mathcal{J}_{s}(u): =s𝒥(u)s12|u|2\displaystyle=s\mathcal{J}(u)-\frac{s-1}{2}|u|^{2}
=12(i=0N|ui|2si=1Nxi,Qixi).\displaystyle=\frac{1}{2}\left(\sum_{i=0}^{N}|u_{i}|^{2}-s\sum_{i=1}^{N}\langle x_{i},Q_{i}x_{i}\rangle\right).

The following result is a discrete version of the Pontryagin maximum principle for LQR problems, which characterizes critical points of the constrained variational problem as solutions to a Hamiltonian system.

Lemma 1.

The quadratic form 𝒥s\mathcal{J}_{s} is degenerate on kerEN+1\ker E^{N+1} if and only if the boundary value problem

{xk+1=Pkk+1xk+BkBktλk+1,λk=(Pkk+1)tλk+1+sQkxk,\begin{cases}x_{k+1}=P^{k+1}_{k}x_{k}+B_{k}B_{k}^{t}\lambda_{k+1},\\ \lambda_{k}=(P^{k+1}_{k})^{t}\lambda_{k+1}+sQ_{k}x_{k},\end{cases} (12)

with x0=0x_{0}=0, xN+1=0x_{N+1}=0 has a non-trivial solution. Moreover the dimension of the space of solutions of eq. 12 and of ker𝒥s|kerEN+1\ker\mathcal{J}_{s}|_{\ker E^{N+1}} coincide.

The former quadratic form is degenerate if and only if (1s)/s(1-s)/s is an eigenvalue of the restriction to kerdEN+1\ker dE^{N+1} of:

i=0N|vi|2j=1N+1(Ei(v))tQiEi(v).\sum_{i=0}^{N}|v_{i}|^{2}-\sum_{j=1}^{N+1}(E^{i}(v))^{t}Q_{i}E^{i}(v).
Proof.

If uu is critical point of 𝒥s\mathcal{J}_{s} restricted to kerEN+1\ker E^{N+1}, there exists λN+1n\lambda_{N+1}\in\mathbb{R}^{n} such that

du𝒥N=λN+1tduEN+1.d_{u}\mathcal{J}^{N}=\lambda_{N+1}^{t}d_{u}E^{N+1}. (13)

Since EN+1E^{N+1} is linear, we have duEN+1=EN+1d_{u}E^{N+1}=E^{N+1}. We will often use the latter notation in the formulas that follow.

In order to obtain a discrete analogue of a Hamiltonian system of PMP, we need to introduce for 1kN1\leq k\leq N a family of functionals

𝒥sk(u):=12(i=0k|ui|2si=1kxi,Qixi)\mathcal{J}_{s}^{k}(u):=\frac{1}{2}\left(\sum_{i=0}^{k}|u_{i}|^{2}-s\sum_{i=1}^{k}\langle x_{i},Q_{i}x_{i}\rangle\right)

and their differentials:

du𝒥sk(v)=i=0kui,visi=1kxitQiEi(v).d_{u}\mathcal{J}^{k}_{s}(v)=\sum_{i=0}^{k}\langle u_{i},v_{i}\rangle-s\sum_{i=1}^{k}x_{i}^{t}Q_{i}E^{i}(v). (14)

Functionals 𝒥sk\mathcal{J}^{k}_{s} differ from 𝒥s\mathcal{J}_{s} only by the range of summation and clearly 𝒥sN=𝒥s\mathcal{J}^{N}_{s}=\mathcal{J}_{s}. We look for λkn\lambda_{k}\in\mathbb{R}^{n}, 1kN+11\leq k\leq N+1 which satisfy:

du𝒥sk=λk+1tduEk+1,d_{u}\mathcal{J}^{k}_{s}=\lambda_{k+1}^{t}d_{u}E^{k+1}, (15)

Using the explicit formula (8) for the end-point map we find the recurrence relation

Ek+1(v)=Pkk+1Ek(prk1v)+BkvkE^{k+1}(v)=P^{k+1}_{k}E^{k}(pr_{k-1}v)+B_{k}v_{k} (16)

for all vUkv\in U_{k}. From here, we can already derive the expression for the kthk-th component of the control. Indeed, assume that vi=0v_{i}=0 for all i<ki<k. In this case

Ei(pri1v)=0,ik.E^{i}(pr_{i-1}v)=0,\quad\forall i\leq k.

Hence, if we substite such a control vv in (15), then using (14) and (16) we obtain the control law

uk=Bktλk+1,u_{k}=B_{k}^{t}\lambda_{k+1}, (17)

which we can plug in (3).

The next step is to obtain a discrete equation for λk\lambda_{k}. To do this we compare the multipliers λk\lambda_{k} and λk+1\lambda_{k+1} by subtracting (15) from the same formula with kk replaced by k1k-1. Using vUk1Ukv\in U_{k-1}\subset U_{k} as a test variation and formulas (14) and (16) we find that:

λk+1tEk+1(v)λktEk(v)=du𝒥sk(v)du𝒥sk1(v)\displaystyle\lambda_{k+1}^{t}E^{k+1}(v)-\lambda_{k}^{t}E^{k}(v)=d_{u}\mathcal{J}^{k}_{s}(v)-d_{u}\mathcal{J}^{k-1}_{s}(v)
(λk+1tPkk+1λkt+sxktQk)Ek(v)=0.\displaystyle\iff\left(\lambda^{t}_{k+1}P^{k+1}_{k}-\lambda^{t}_{k}+sx_{k}^{t}Q_{k}\right)E^{k}(v)=0.

Since by our assumption BkB_{k} are invertible for all 0kN0\leq k\leq N, the end-point map EkE^{k} is surjective, as can be seen from the explicit expression (8). Thus the term in the bracket must vanish.

Collecting everything gives eq. 12. The boundary conditions for this system come form the fact that we have to add the equation EN+1(u)=0=xN+1E^{N+1}(u)=0=x_{N+1} and that we are assuming x0=0x_{0}=0.

Now, notice that the initial covector λ0\lambda_{0} of the lift uniquely determines the whole trajectory and the control. This means that the correspondence λ0u\lambda_{0}\mapsto u is a bijection between the space of solutions of the boundary value problem and the kernel of 𝒥s\mathcal{J}_{s}. ∎

Under 1, we can rewrite system (12) as a forward equation. To do so, recall that Γk=BkBkt\Gamma_{k}=B_{k}B_{k}^{t}, multiply the second equation by (Pk+1k)t(P^{k}_{k+1})^{t} and plug in the new expression for λk+1\lambda_{k+1} into the first equation. This gives

{xk+1=(Pkk+1sΓk(Pk+1k)tQk)xk+Γk(Pk+1k)tλk,λk+1=(Pk+1k)tλks(Pk+1k)tQkxk.\begin{cases}x_{k+1}=(P^{k+1}_{k}-s\Gamma_{k}(P^{k}_{k+1})^{t}Q_{k})x_{k}+\Gamma_{k}(P^{k}_{k+1})^{t}\lambda_{k},\\ \lambda_{k+1}=(P^{k}_{k+1})^{t}\lambda_{k}-s(P^{k}_{k+1})^{t}Q_{k}x_{k}.\end{cases} (18)

We thus have

(λk+1xk+1)=Mk(s)(λkxk),\begin{pmatrix}\lambda_{k+1}\\ x_{k+1}\end{pmatrix}=M_{k}(s)\begin{pmatrix}\lambda_{k}\\ x_{k}\end{pmatrix},

where

Mk(s)=((Pk+1k)t0Γk(Pk+1k)tPkk+1)(1sQk01).\displaystyle M_{k}(s)=\begin{pmatrix}(P^{k}_{k+1})^{t}&0\\ \Gamma_{k}(P^{k}_{k+1})^{t}&P^{k+1}_{k}\end{pmatrix}\begin{pmatrix}1&-sQ_{k}\\ 0&1\end{pmatrix}.

Both matrices in the product are symplectic, which makes Mk(s)M_{k}(s) symplectic as well.

Define the flow up to the point kk as

Φk(s)=(Φk1(s)Φk2(s)Φk3(s)Φk4(s)):=i=0k1Mi(s).\Phi_{k}(s)=\begin{pmatrix}\Phi^{1}_{k}(s)&\Phi^{2}_{k}(s)\\ \Phi^{3}_{k}(s)&\Phi^{4}_{k}(s)\end{pmatrix}:=\prod_{i=0}^{k-1}M_{i}(s).

When k=N+1k=N+1, we just write Φ(s)=ΦN+1(s)\Phi(s)=\Phi_{N+1}(s). Also in accordance with the notations given in the introduction, we write for s=1s=1:

𝒥k:=𝒥1k,Φk:=Φk(1),Mk:=Mk(1).\mathcal{J}^{k}:=\mathcal{J}^{k}_{1},\qquad\Phi_{k}:=\Phi_{k}(1),\qquad M_{k}:=M_{k}(1).
Remark 2.

We can reformulate Lemma 1 and the boundary value problem (12) in terms of Φk(s)\Phi_{k}(s). We are looking for solutions of (12) with boundary values x0=xk+1=0x_{0}=x_{k+1}=0. Writing as above Φk+1(s)\Phi_{k+1}(s) as a block matrix shows that the existence of a solution is equivalent to the vanishing of the determinant of the block Φk+13\Phi^{3}_{k+1}:

ker𝒥sk|kerEk+10detΦk+13(s)=0.\ker\mathcal{J}^{k}_{s}|_{\ker E^{k+1}}\neq 0\iff\det\Phi^{3}_{k+1}(s)=0.

3 A recursive formula for the index

In this section we prove Theorem 1. The tools we will employ are essentially taken from linear algebra. The same kind of ideas has been applied in [4] to reformulate problems of intersection theory for paths in the Lagrange Grassmannian in terms of finite dimensional linear algebra. The classical approach to second order minimality conditions rephrases the problem of counting the number of negative eigenvalues of the Second Variation (the counter part of our quadratic form 𝒥\mathcal{J}) as counting the intersection number of a curve in a finite-dimensional manifolds, known as the Lagrangian Grassmannian, with an appropriate submanifold. For further references on the continuum approach see for instance [9, 14, 1] and reference therein.

In contrast to the continuous case, here we work only with an ordered sets of points constructed from the solutions of eq. 12. Still, all the information about the index can be recovered from them.

We will use the following lemma:

Lemma 2.

Suppose that a quadratic form 𝒬\mathcal{Q} is defined on finite dimensional vector space XX and a subspace VXV\subseteq X is given. Denote by V𝒬V^{\perp_{\mathcal{Q}}} the 𝒬\mathcal{Q}-orthogonal subspace to VV i.e.:

V𝒬:={uX:𝒬(u,v)=0,vV}.V^{\perp_{\mathcal{Q}}}:=\{u\in X:\mathcal{Q}(u,v)=0,\forall v\in V\}.

Denote by indQind^{-}Q the number of negative eigenvalues of QQ, then:

ind𝒬\displaystyle ind^{-}\mathcal{Q} =ind𝒬|V+ind𝒬|V𝒬\displaystyle=ind^{-}\mathcal{Q}|_{V}+ind^{-}\mathcal{Q}|_{V^{\perp_{\mathcal{Q}}}}
+dim((VV𝒬)/(VV𝒬ker𝒬)).\displaystyle+\dim\big{(}(V\cap V^{\perp_{\mathcal{Q}}})/(V\cap V^{\perp_{\mathcal{Q}}}\cap\ker\mathcal{Q})\big{)}.

We will apply iteratively this formula to the subspaces Vk:=UkkerEk+1V_{k}:=U_{k}\cap\ker E^{k+1}. The subspaces UkU_{k} were defined in (9) and provide a filtration of our space of variations. One has a natural filtration Vk1VkV_{k-1}\subset V_{k} as well. Indeed, if uVk1u\in V_{k-1}, then xk=Ek(u)=0x_{k}=E^{k}(u)=0. Hence, if we extend the control uu to VkV_{k} by taking uk=0u_{k}=0, we get from (8)

xk+1=Ek+1(u)=Pkk+1Ek(u)=0.x_{k+1}=E^{k+1}(u)=P^{k+1}_{k}E^{k}(u)=0.

Notice also that for any kk, the subspace Vk1V_{k-1} has codimension nn in VkV_{k}.

We denote for brevity

𝒬k:=𝒥k|kerEk+1,Vk:=Vk𝒬k+1,𝒬k:=𝒬k|Vk1.\mathcal{Q}_{k}:=\mathcal{J}^{k}|_{\ker E^{k+1}},\quad V_{k}^{\perp}:=V_{k}^{\perp_{\mathcal{Q}_{k+1}}},\quad\mathcal{Q}_{k}^{\perp}:=\mathcal{Q}_{k}|_{V_{k-1}^{\perp}}.

Let us apply Lemma 2 with 𝒬=𝒬k+1\mathcal{Q}=\mathcal{Q}_{k+1} and V=VkVk+1=XV=V_{k}\subseteq V_{k+1}=X. This gives:

ind𝒬k+1\displaystyle ind^{-}\mathcal{Q}_{k+1} =ind𝒬k+ind𝒬k+1\displaystyle=ind^{-}\mathcal{Q}_{k}+ind^{-}\mathcal{Q}_{k+1}^{\perp}
+dim((VkVk)/(VkVkker𝒬k+1)).\displaystyle+\dim\big{(}(V_{k}\cap V_{k}^{\perp})/(V_{k}\cap V_{k}^{\perp}\cap\ker\mathcal{Q}_{k+1})\big{)}.

Thus, iteration of this formula gives:

ind𝒥|kerEN+1\displaystyle ind^{-}\mathcal{J}|_{\ker E^{N+1}} =k=0N1ind𝒬k+1\displaystyle=\sum_{k=0}^{N-1}ind^{-}\mathcal{Q}_{k+1}^{\perp}
+dim((VkVk)/(VkVkker𝒬k+1)).\displaystyle+\dim\big{(}(V_{k}\cap V_{k}^{\perp})/(V_{k}\cap V_{k}^{\perp}\cap\ker\mathcal{Q}_{k+1})\big{)}.

What is left to do is to express every term using the discrete Hamiltonian system given in eq. 12.

First of all, let us describe the subspaces VkVkV_{k}^{\perp}\cap V_{k} and VkVkker𝒬k+1V_{k}^{\perp}\cap V_{k}\cap\ker\mathcal{Q}_{k+1}.

Lemma 3.

Let VkV_{k} be defined as above and let Π\Pi be the vertical subspace

Π={(λ,x)2n:x=0}.\Pi=\{(\lambda,x)\in\mathbb{R}^{2n}\,:\,x=0\}.

Then we have the following identifications:

VkVk=ker(𝒬k)=Φk+1(Π)Π;VkVkker(𝒬k+1)=Φk+1(Π)Mk+11(Π)Π.\begin{split}V_{k}^{\perp}\cap V_{k}&=\ker(\mathcal{Q}_{k})=\Phi_{k+1}(\Pi)\cap\Pi;\\ V_{k}^{\perp}\cap V_{k}\cap\ker(\mathcal{Q}_{k+1})&=\Phi_{k+1}(\Pi)\cap M_{k+1}^{-1}(\Pi)\cap\Pi.\end{split}
Proof.

A direct consequence of the definitions of VkV_{k} and VkV_{k}^{\perp} is that VkVk=ker(𝒬k)V_{k}^{\perp}\cap V_{k}=\ker(\mathcal{Q}_{k}). Lemma 1 identifies the elements of the kernel with solutions of the equation (12) with x0=xk+1=0x_{0}=x_{k+1}=0. In terms of the corresponding flow Φk+1\Phi_{k+1} we get:

ker(𝒬k)=Φk+1(Π)Π.\ker(\mathcal{Q}_{k})=\Phi_{k+1}(\Pi)\cap\Pi.

Similarly we have ker(𝒬k+1)=Φk+2(Π)Π\ker(\mathcal{Q}_{k+1})=\Phi_{k+2}(\Pi)\cap\Pi. Applying Mk+11M_{k+1}^{-1} gives

ker(𝒬k+1)=Φk+1(Π)Mk+11(Π).\ker(\mathcal{Q}_{k+1})=\Phi_{k+1}(\Pi)\cap M_{k+1}^{-1}(\Pi).

Combining this with the previous formula proves the second isomorphism in the statement. ∎

Notice that the previous lemma holds for general LQR. In this particular case, Mk+11(Π)M^{-1}_{k+1}(\Pi) is transversal to the fibre. Indeed, from the explicit expression of Mk+1M_{k+1} we have that Mk+1(Π)Π0M_{k+1}(\Pi)\cap\Pi\neq 0 if and only if kerΓk+1(Pk+2k+1)t0\ker\Gamma_{k+1}(P^{k+1}_{k+2})^{t}\neq 0, which is never the case under 1. Thus

VkVk=ker(𝒬k)=Φk+1(Π)Π;VkVkker(𝒬k+1)=0.\begin{split}V_{k}^{\perp}\cap V_{k}=\ker(\mathcal{Q}_{k})&=\Phi_{k+1}(\Pi)\cap\Pi;\\ V_{k}^{\perp}\cap V_{k}\cap\ker(\mathcal{Q}_{k+1})&=0.\end{split}

The next step is to compute ind𝒬kind^{-}\mathcal{Q}_{k}^{\perp}. To do so consider the standard symplectic form on 2n\mathbb{R}^{2n}:

σ((λ1,x1),(λ2,x2))=λ1tx2λ2tx1=(λ1tx1t)(0II0)(λ2x2).\sigma((\lambda_{1},x_{1}),(\lambda_{2},x_{2}))=\lambda_{1}^{t}x_{2}-\lambda_{2}^{t}x_{1}=\begin{pmatrix}\lambda_{1}^{t}&x_{1}^{t}\end{pmatrix}\begin{pmatrix}0&I\\ -I&0\end{pmatrix}\begin{pmatrix}\lambda_{2}\\ x_{2}\end{pmatrix}.

A plane Λ2n\Lambda\subset\mathbb{R}^{2n} is called Lagrangian if σ|Λ=0\sigma|_{\Lambda}=0 and dimΛ=n\dim\Lambda=n, i.e., it is a maximally isotropic space.

The Maslov form m(Λ1,Λ2,Λ3)m(\Lambda_{1},\Lambda_{2},\Lambda_{3}) of a triple of Lagrangian spaces Λ1,Λ2,Λ3\Lambda_{1},\Lambda_{2},\Lambda_{3} is a symmetric quadratic form on (Λ1+Λ3)Λ2(\Lambda_{1}+\Lambda_{3})\cap\Lambda_{2} defined as follows. Let p2=p1+p3p_{2}=p_{1}+p_{3} where piΛip_{i}\in\Lambda_{i}. Then

m(Λ1,Λ2,Λ3)(p2)=σ(p1,p3).m(\Lambda_{1},\Lambda_{2},\Lambda_{3})(p_{2})=\sigma(p_{1},p_{3}).
Lemma 4.

The negative index of 𝒬k\mathcal{Q}_{k} restricted to Vk1V_{k-1}^{\perp} coincides with the negative index of m(Φk(Π),Mk1(Π),Π)m(\Phi_{k}(\Pi),M_{k}^{-1}(\Pi),\Pi).

Proof.

An element uu belongs to VkV_{k} if and only if it is of the form:

u=(u0,,uk1,Fk(u)),u=(u_{0},\dots,u_{k-1},F_{k}(u)),

where FkF_{k} is defined in (10). Hence we can identify VkV_{k} with nk\mathbb{R}^{nk}.

It follows that an element uu belongs to Vk1V_{k-1}^{\perp} if and only if for all vVk1n(k1)v\in V_{k-1}\simeq\mathbb{R}^{n(k-1)}:

0=𝒬k(u,v)=12(i=0kui,vii=1kQiEiu,Eiv)\displaystyle 0=\mathcal{Q}_{k}(u,v)=\frac{1}{2}\left(\sum_{i=0}^{k}\langle u_{i},v_{i}\rangle-\sum_{i=1}^{k}\langle Q_{i}E^{i}u,E^{i}v\rangle\right)
=12(i=0k2ui,vi(Ei+1)tQi+1Ei+1u,vL2+uk1,Fk1(v))\displaystyle=\frac{1}{2}\Bigg{(}\sum_{i=0}^{k-2}\langle u_{i},v_{i}\rangle-\langle(E^{i+1})^{t}Q_{i+1}E^{i+1}u,v\rangle_{L^{2}}+\langle u_{k-1},F_{k-1}(v)\rangle\Bigg{)}
=12ui=0k2(Ei+1)tQi+1Ei+1u+Fk1t(uk1),vL2\displaystyle=\frac{1}{2}\langle u-\sum_{i=0}^{k-2}(E^{i+1})^{t}Q_{i+1}E^{i+1}u+F_{k-1}^{t}(u_{k-1}),v\rangle_{L^{2}}

where the L2L^{2}-scalar product is defined on functions on a finite set with values in n\mathbb{R}^{n}. Here we used that Ek(v)=0E^{k}(v)=0 since vVk1v\in V_{k-1}.

We have thus shown that if uVk1u\in V_{k-1}^{\perp} then

ui=0k2(Ei+1)tQi+1Ei+1u,vL2=Fk1t(uk1),vL2,\langle u-\sum_{i=0}^{k-2}(E^{i+1})^{t}Q_{i+1}E^{i+1}u,v\rangle_{L^{2}}=-\langle F_{k-1}^{t}(u_{k-1}),v\rangle_{L^{2}},

for all functions v:{0,,k1}nv:\{0,\dots,k-1\}\to\mathbb{R}^{n}.

Let us now simplify the form 𝒬k\mathcal{Q}_{k}^{\perp} using this expression. Take an element uVk1u\in V_{k-1}^{\perp}. Then

𝒬k(u)=12((i=0k2ui,uiu,(Ei+1)tQi+1Ei+1(u))+uk1,uk1+Fk(u),Fk(u)Ek(u),QkEk(u))=12(Fk1(u),uk1+uk1,uk1+Fk(u),Fk(u)xk,Qkxk).\begin{split}&\mathcal{Q}_{k}(u)=\frac{1}{2}\Bigg{(}\left(\sum_{i=0}^{k-2}\langle u_{i},u_{i}\rangle-\langle u,(E^{i+1})^{t}Q_{i+1}E^{i+1}(u)\rangle\right)\\ &+\langle u_{k-1},u_{k-1}\rangle+\langle F_{k}(u),F_{k}(u)\rangle-\langle E^{k}(u),Q_{k}E^{k}(u)\rangle\Bigg{)}\\ &=\frac{1}{2}(-\langle F_{k-1}(u),u_{k-1}\rangle+\langle u_{k-1},u_{k-1}\rangle+\langle F_{k}(u),F_{k}(u)\rangle\\ &-\langle x_{k},Q_{k}x_{k}\rangle).\end{split}

Where we used the characterization of Vk1V_{k-1}^{\perp} given in the previous lemma and substituted Ek(u)E^{k}(u) with xkx_{k}. We now rewrite each term in symplectic form. In particular recall the following relations, which are consequences of (17) and (3)

uk\displaystyle u_{k} =Bktλk+1,\displaystyle=B_{k}^{t}\lambda_{k+1},
xk\displaystyle x_{k} =Pk1kxk1+Γk1λk,\displaystyle=P_{k-1}^{k}x_{k-1}+\Gamma_{k-1}\lambda_{k},
Bk1tλk\displaystyle B_{k-1}^{t}\lambda_{k} =Fk1(u)=Bk11Pk1kxk1.\displaystyle=F_{k-1}(u)=-B_{k-1}^{-1}P_{k-1}^{k}x_{k-1}.

This implies that:

uk1,uk1\displaystyle\langle u_{k-1},u_{k-1}\rangle =λk,Γk1λk,\displaystyle=\langle\lambda_{k},\Gamma_{k-1}\lambda_{k}\rangle,
Fk1(u),uk1\displaystyle\langle F_{k-1}(u),u_{k-1}\rangle =λk,Pk1kxk1.\displaystyle=-\langle\lambda_{k},P_{k-1}^{k}x_{k-1}\rangle.

Similarly, Fk(u)F_{k}(u) and xkx_{k} can be expressed in terms of λk+1\lambda_{k+1}.

Pkk+1xk=BkFk(u)=Γkλk+1.-P_{k}^{k+1}x_{k}=B_{k}F_{k}(u)=\Gamma_{k}\lambda_{k+1}.

We thus get the following expression for uVk1u\in V_{k-1}^{\perp} after substituting the previous relations and grouping separately terms with λk\lambda_{k} and λk+1\lambda_{k+1}:

𝒬k(u)=12(λk,Pk1kxk1+Γk1λk+λk+1,Γk(λk+1+(Pk+1k)tQkxk))=12(λk,xk+λk+1,Γk(λk+1+(Pk+1k)tQkxk)).\begin{split}\mathcal{Q}_{k}(u)&=\frac{1}{2}(\langle\lambda_{k},P_{k-1}^{k}x_{k-1}+\Gamma_{k-1}\lambda_{k}\rangle\\ &\quad+\langle\lambda_{k+1},\Gamma_{k}\left(\lambda_{k+1}+(P_{k+1}^{k})^{t}Q_{k}x_{k}\right)\rangle)\\ &=\frac{1}{2}\left(\langle\lambda_{k},x_{k}\rangle+\langle\lambda_{k+1},\Gamma_{k}\left(\lambda_{k+1}+(P_{k+1}^{k})^{t}Q_{k}x_{k}\right)\rangle\right).\end{split}

On the other hand let us consider the Maslov form m(Φk(Π),Mk1(Π),Π)m(\Phi_{k}(\Pi),M^{-1}_{k}(\Pi),\Pi) and write down its explicit expression. We have on Mk1(Π)(Φk(Π)+Π)M^{-1}_{k}(\Pi)\cap(\Phi_{k}(\Pi)+\Pi).

Mk1(λk+10)\displaystyle M_{k}^{-1}\begin{pmatrix}\lambda_{k+1}\\ 0\end{pmatrix} =(((Pkk+1)tQkPk+1kΓk)λk+1Pk+1kΓkλk+1)\displaystyle=\begin{pmatrix}\left((P_{k}^{k+1})^{t}-Q_{k}P_{k+1}^{k}\Gamma_{k}\right)\lambda_{k+1}\\ -P_{k+1}^{k}\Gamma_{k}\lambda_{k+1}\end{pmatrix}
=(λk+νxk)Π+Φk(Π).\displaystyle=\begin{pmatrix}\lambda_{k}+\nu\\ x_{k}\end{pmatrix}\in\Pi+\Phi_{k}(\Pi). (19)

Thus the value of m(λk+1)=σ((λk,xk),(ν,0))=ν,xkm(\lambda_{k+1})=\sigma((\lambda_{k},x_{k}),(\nu,0))=-\langle\nu,x_{k}\rangle is determined by:

m(λk+1)=((Pkk+1)tQkPk+1kΓk)λk+1λk,xk=λk+1,Γk(λk+1+(Pk+1k)tQkxk)+λk,xk=2𝒬k(u).\begin{split}m(\lambda_{k+1})&=-\langle\left((P_{k}^{k+1})^{t}-Q_{k}P_{k+1}^{k}\Gamma_{k}\right)\lambda_{k+1}-\lambda_{k},x_{k}\rangle\\ &=\langle\lambda_{k+1},\Gamma_{k}(\lambda_{k+1}+(P_{k+1}^{k})^{t}Q_{k}x_{k})\rangle+\langle\lambda_{k},x_{k}\rangle\\ &=2\mathcal{Q}_{k}(u).\end{split} (20)

Here, in the second equality we substituted ν\nu from (3) and in the third equality xkx_{k} from the same equation. It follows that negative inertia index of 𝒬k\mathcal{Q}_{k}^{\perp} is the same as the negative inertia index of m(Φk(Π),Mk1(Π),Π)m(\Phi_{k}(\Pi),M^{-1}_{k}(\Pi),\Pi). ∎

To arrive at formula (6) we need to write down explicitly the kernel of the Maslov form in terms of the initial covector λ0\lambda_{0}. We have

(λkxk)=(Φk1λ0Φk3λ0),Φk=(Φk1Φk2Φk3Φk4).\begin{pmatrix}\lambda_{k}\\ x_{k}\end{pmatrix}=\begin{pmatrix}\Phi_{k}^{1}\lambda_{0}\\ \Phi_{k}^{3}\lambda_{0}\end{pmatrix},\quad\Phi_{k}=\begin{pmatrix}\Phi_{k}^{1}&\Phi_{k}^{2}\\ \Phi_{k}^{3}&\Phi_{k}^{4}\end{pmatrix}.

Moreover, λk+1\lambda_{k+1} is directly determined by xkx_{k}, see section 3, and thus by λ0\lambda_{0} as:

λk+1=Γk1Pkk+1Φk3λ0.\lambda_{k+1}=-\Gamma_{k}^{-1}P^{k+1}_{k}\Phi_{k}^{3}\lambda_{0}.

After inserting inside the explicit expression of the Maslov form given in (20) we find that the following quadratic form is the one to consider:

λ0,((Φk3)t((Pkk+1)tΓk1(Pkk+1)Qk)Φk3+(Φk3)tΦk1)λ0.\langle\lambda_{0},\Big{(}(\Phi_{k}^{3})^{t}\big{(}(P^{k+1}_{k})^{t}\Gamma_{k}^{-1}(P^{k+1}_{k})-Q_{k}\big{)}\Phi_{k}^{3}+(\Phi_{k}^{3})^{t}\Phi_{k}^{1}\Big{)}\lambda_{0}\rangle.

Now each term in the formula for ind𝒥|kerEN+1ind^{-}\mathcal{J}|_{\ker E^{N+1}} is identified and collecting all the pieces together finishes the proof of Theorem 1.

4 Proof of the determinant formula

In this section we prove Theorem 2. Consider the quadratic form 2𝒥2\mathcal{J}. As discussed in the introduction, we can pass from control variables uu to state variables xx. This gives the Jacobi matrix {\mathcal{I}} once we impose the boundary conditions x0=xN+1=0x_{0}=x_{N+1}=0. This matrix decomposes as the sum of two terms. A diagonal matrix QQ and a tridiagonal one T=(Ti,j)i,jT=(T_{i,j})_{i,j} given by:

s=TsQ,Q=diag(Qi),i=1,,N,\displaystyle{\mathcal{I}}_{s}=T-sQ,\quad Q=\mathrm{diag}(Q_{i}),\,i=1,\dots,N,
Ti=Γi11+(Pii+1)tΓi1Pii+1,Ti,i+1=(Pii+1)tΓi1.\displaystyle T_{i}=\Gamma_{i-1}^{-1}+(P_{i}^{i+1})^{t}\Gamma^{-1}_{i}P_{i}^{i+1},\quad T_{i,i+1}=-(P_{i}^{i+1})^{t}\Gamma^{-1}_{i}.

Notice, moreover, that the matrix TT is positive definite, since it is a multiple of the quadratic form i=0N|ui|2\sum_{i=0}^{N}|u_{i}|^{2}. For this reason we can define its square root T1/2T^{1/2} which is again positive definite and symmetric. We will use this observation to relate the determinant of Φ3(s)\Phi_{3}(s) to the characteristic polynomial of a suitable nNnN matrix.

Lemma 5.

The function det(Φ3(s))\det(\Phi_{3}(s)) is a polynomial of degree nNk=1Ndim(ker(Qk))nN-\sum_{k=1}^{N}\dim(\ker(Q_{k})) and in particular it is a multiple of det(TsQ)\det(T-sQ).

Proof.

The first thing we have to notice is that the entries of Φ(s)\Phi(s) are polynomials of degree at most NN since Φ(s)\Phi(s) is the product of NN affine in ss matrices. This implies that det(Φ3(s))\det(\Phi_{3}(s)) has at most degree nNnN.

This is exactly the dimension of the space of controls. We know by Lemma 1 that this polynomial has at least d:=nNidim(kerQi)d:=nN-\sum_{i}\dim(\ker Q_{i}) roots. This is because our perturbation 𝒥s\mathcal{J}_{s} sees every eigenvalue of 𝒥\mathcal{J} but 11, whose eigenspace corresponds to the kernel of QQ. In particular when the QiQ_{i} are all invertible, det(Φ3(s))\det(\Phi_{3}(s)) has degree nNnN, vanishes on the same set as det(TsQ)\det(T-sQ) and thus they are multiples. In general we have:

det(TsQ)=det(T)det(1sT1/2QT1/2)=snNdet(T)det(s1T1/2QT1/2)=λnNdet(T)det(λT1/2QT1/2)|λ=s1.\begin{split}\det(T-sQ)&=\det(T)\det(1-sT^{-1/2}QT^{-1/2})\\ &=s^{nN}\det(T)\det(s^{-1}-T^{-1/2}QT^{-1/2})\\ &=\lambda^{-nN}\det(T)\det(\lambda-T^{-1/2}QT^{-1/2})|_{\lambda=s^{-1}}.\end{split}

Order the non zero eigenvalues of T1/2QT1/2T^{-1/2}QT^{-1/2} as {λi}i=1d\{\lambda_{i}\}_{i=1}^{d}. Since det(λT1/2QT1/2)\det(\lambda-T^{-1/2}QT^{-1/2}) has dd roots different from zero and a root of order idim(ker(Qi))\sum_{i}\dim(\ker(Q_{i})) at λ=0\lambda=0 we can rewrite the formula as:

det(TsQ)=i=1d(1sλi)det(T).\det(T-sQ)=\prod_{i=1}^{d}(1-s\lambda_{i})\det(T).

Thus we have that det(Φ3(s))\det(\Phi_{3}(s)) and det(TsQ)\det(T-sQ) are both polynomials of degree dd vanishing on the same set and thus they must be multiples. ∎

By Lemma 5 there exists a constant C0C\neq 0 such that

det(TsQ)=CdetΦ3(s).\det(T-sQ)=C\det\Phi_{3}(s).

The constant can be computed by evaluating both sides at s=0s=0. Using induction we can prove that

Φ(0)=((PN+10)T0i=0NPi+1N+1Γi(Pi+10)TP0N+1).\Phi(0)=\begin{pmatrix}(P^{0}_{N+1})^{T}&0\\ \sum_{i=0}^{N}P^{N+1}_{i+1}\Gamma_{i}(P^{0}_{i+1})^{T}&P^{N+1}_{0}\end{pmatrix}. (21)

Hence

det(Φ3(s))|s=0=det(i=0NPi+1N+1Γi(Pi+10)T)=det(i=0NPi+1N+1Γi(Pi+1N+1)T)det(PN+10).\begin{split}\det(\Phi_{3}(s))|_{s=0}&=\det\left(\sum_{i=0}^{N}P^{N+1}_{i+1}\Gamma_{i}(P^{0}_{i+1})^{T}\right)\\ &=\det\left(\sum_{i=0}^{N}P^{N+1}_{i+1}\Gamma_{i}(P_{i+1}^{N+1})^{T}\right)\det(P^{0}_{N+1}).\end{split}

Now it only remains to find detT\det T. Recall that TT corresponds to the quadratic form 12k=0N|uk|2\frac{1}{2}\sum_{k=0}^{N}|u_{k}|^{2}.

Given a trajectory of (3) with x0=xN+1=0x_{0}=x_{N+1}=0 we can always reconstruct the control. This gives a bijection between (u0,,uN1)(u_{0},\dots,u_{N-1}) and (x1,xN)(x_{1},\dots x_{N}) which can be written using a block-triangular matrix

L=(B01000B11P12B110000BN11PN1NBN11)L=\begin{pmatrix}B_{0}^{-1}&0&\dots&0&0\\ -B_{1}^{-1}P^{2}_{1}&B_{1}^{-1}&\dots&0&0\\ \vdots&\vdots&\ddots&\vdots&\vdots\\ 0&0&\dots&-B_{N-1}^{-1}P^{N}_{N-1}&B_{N-1}^{-1}\end{pmatrix}

Now we wish to use FNF_{N} (see eq. 10) to pull back to UN1U_{N-1} the quadratic form:

i=0N|ui|2\sum_{i=0}^{N}|u_{i}|^{2}

seen as a quadratic form on kerEN+1\ker E^{N+1}. To do so we simply need to resolve for uNu_{N} using (11) and plug it inside the form above. This gives the form q(v)=(1+FNFN)v,vq(v)=\langle(1+F_{N}^{*}F_{N})v,v\rangle on UN1U_{N-1}.

We can now change coordinates and work with the xx variable. One can see that TT can be written as a symmetric matrix on nN×nN\mathbb{R}^{nN}\times\mathbb{R}^{nN} of the form:

T=LT(1+FNFN)L.T=L^{T}(1+F_{N}^{*}F_{N})L.

So if we want to compute its determinant, we need to find determinant of LL and of 1+FNFN1+F_{N}^{*}F_{N}. We clearly have

detL=i=0N1Bi1.\det L=\prod_{i=0}^{N-1}B_{i}^{-1}.

So it only remains to find det(1+FNFN)\det(1+F_{N}^{*}F_{N}).

Notice that, since FNF_{N} is a surjection, it has a large kernel of dimension N(n1)N(n-1). If vkerFNv\in\ker F_{N}, then vv is an eigenvector of qq with eigenvalue 11. So the determinant can be computed multiplying the remaining nn eigenvalues. We claim that they are all equal to the eigenvalues of 1+FNFN1+F_{N}F_{N}^{*}. Indeed, if vkerFNv\notin\ker F_{N} is an eigenvalue of qq, then

(1+FNFN)v=λv.(1+F_{N}^{*}F_{N})v=\lambda v.

Applying FF to both sides gives

(1+FNFN)FNv=λFNv.(1+F_{N}F_{N}^{*})F_{N}v=\lambda F_{N}v.

Hence we obtain that:

detT=i=0N1det(Γi)1×det(1+j=0N1BN1Pj+1N+1Γj1(Pj+1N+1)t(BN1)t)=i=0Ndet(Γi)1det(ΓN+j=0N1Pj+1N+1Γj1(Pj+1N+1)t).\begin{split}\det T&=\prod_{i=0}^{N-1}\det(\Gamma_{i})^{-1}\\ &\times\det\left(1+\sum_{j=0}^{N-1}B_{N}^{-1}P^{N+1}_{j+1}\Gamma_{j}^{-1}(P^{N+1}_{j+1})^{t}(B_{N}^{-1})^{t}\right)\\ &=\prod_{i=0}^{N}\det(\Gamma_{i})^{-1}\det\left(\Gamma_{N}+\sum_{j=0}^{N-1}P^{N+1}_{j+1}\Gamma_{j}^{-1}(P^{N+1}_{j+1})^{t}\right).\end{split}

This implies that C=det(P0N+1)k=0NΓk1C=\det(P_{0}^{N+1})\prod_{k=0}^{N}\Gamma_{k}^{-1} and thus:

det(Φ3(s))det(P0N+1)k=0NdetΓk1=det(TsQ),\det(\Phi_{3}(s))\det(P_{0}^{N+1})\prod_{k=0}^{N}\det\Gamma_{k}^{-1}=\det(T-sQ), (22)

which gives formula (7) when s=1s=1.

5 Some examples and applications

5.1 Tridiagonal Toeplitz Matrices

Let us compute the determinant of a symmetric tridiagonal Toeplitz matrix

TN(s,r)=(sr000rsr000rs00000sr000rs),T_{N}(s,r)=\begin{pmatrix}s&r&0&\dots&0&0\\ r&s&r&\dots&0&0\\ 0&r&s&\dots&0&0\\ \vdots&\vdots&\vdots&\ddots&\vdots&\vdots\\ 0&0&0&\dots&s&r\\ 0&0&0&\dots&r&s\end{pmatrix}, (23)

where s,rs,r\in\mathbb{R} and r0r\neq 0.

Recall that eigenvalues of TN(s,r)T_{N}(s,r) are explicitly known [13] and are of the form

λj=s+2|r|cosπjN+1,1jN.\lambda_{j}=s+2|r|\cos\frac{\pi j}{N+1},\qquad 1\leq j\leq N.

So we already a have a way of computing the determinant of the matrix as the product of all eigenvalues:

detTN(s,r)=j=1N(s+2|r|cosπjN+1).\det T_{N}(s,r)=\prod_{j=1}^{N}\left(s+2|r|\cos\frac{\pi j}{N+1}\right). (24)

Let us compute the same determinant using the control theoretic approach.

Using formulas (5) we reformulate the problem of computing detTN(r,s)\det T_{N}(r,s) as a LQR problem. We choose for 1iN1\leq i\leq N:

B0=Bi=1,A0=0,Ai=(1+r),Qi=1+r2s.B_{0}=B_{i}=1,\quad A_{0}=0,\quad A_{i}=-(1+r),\quad Q_{i}=1+r^{2}-s.

In particular, this immediately implies that

Γi=1,P0N+1=(r)N.\Gamma_{i}=1,\qquad P^{N+1}_{0}=(-r)^{N}.

After comparing with (7), we see that only Φ3\Phi_{3} needs to be computed. In this particular case

Φ=(r1r1(1+r2s)r1r1(1s))N(1011).\Phi=\begin{pmatrix}-r^{-1}&r^{-1}(1+r^{2}-s)\\ -r^{-1}&r^{-1}(1-s)\end{pmatrix}^{N}\begin{pmatrix}1&0\\ 1&1\end{pmatrix}.

Denote

M=(r1r1(1+r2s)r1r1(1s)).M=\begin{pmatrix}-r^{-1}&r^{-1}(1+r^{2}-s)\\ -r^{-1}&r^{-1}(1-s)\end{pmatrix}.

The eigenvalues of this matrix are given by

μ±=s±s24r22r.\mu_{\pm}=\frac{-s\pm\sqrt{s^{2}-4r^{2}}}{2r}. (25)

Assume that μ+μ\mu_{+}\neq\mu_{-}. Then we can diagonalize MM using the matrix

S=(1+rμ1+rμ+11),S=\begin{pmatrix}1+r\mu_{-}&1+r\mu_{+}\\ 1&1\end{pmatrix},

as can be verified by a direct computation. Thus

Φ=S(μ+N00μN)S1(1011).\Phi=S\begin{pmatrix}\mu_{+}^{N}&0\\ 0&\mu_{-}^{N}\end{pmatrix}S^{-1}\begin{pmatrix}1&0\\ 1&1\end{pmatrix}.

After all of the matrices are multiplied, one can take the element which lies under the diagonal and use formula (7). This way, after some simplification, we obtain an expression which can be compactly written in terms of the two eigenvalues

detTN(s,r)=(r)N(μ+1+Nμ1+N)(μ+μ)=(r)Nj=0Nμ+jμNj.\det T_{N}(s,r)=(-r)^{N}\frac{(\mu_{+}^{1+N}-\mu_{-}^{1+N})}{(\mu_{+}-\mu_{-})}=(-r)^{N}\sum_{j=0}^{N}\mu_{+}^{j}\mu_{-}^{N-j}. (26)

If μ+=μ=μ\mu_{+}=\mu_{-}=\mu, then using the continuity of the determinant this formula reduces to

detTN(s,r)=N(rμ)N.\det T_{N}(s,r)=N(-r\mu)^{N}.

5.2 Generalized Euler’s identity

The formula for the determinant of a tridiagonal matrix has been known for a long time. However, we would like to underline that control theoretic approach can be conceptually advantageous. To illustrate this claim, we now prove the generalized Euler’s identity (2).

Notice that we get the classical Euler’s identity if a=0a=0:

sinbb=j=1(1b2(πj)2).\frac{\sin b}{b}=\prod_{j=1}^{\infty}\left(1-\frac{b^{2}}{(\pi j)^{2}}\right).

The identity (2) appeared in [2]. Here we give a proof which uses only elementary tools from linear algebra and analysis. The idea is to take a certain sequence of tridiagonal Toeplitz matrices of increasing rank and to compute the asymptotic expansion of the determinants in two different ways: using formula (24) and (26).

We start with the following family of one-dimensional LQR problems

x˙=ax+u,1201(u(t)2ε(a2+b2)x2(t))𝑑tmin.\dot{x}=ax+u,\quad\frac{1}{2}\int_{0}^{1}\left(u(t)^{2}-\varepsilon(a^{2}+b^{2})x^{2}(t)\right)dt\to\min.

and x(0)=x(1)=0x(0)=x(1)=0. Here ε{0,1}\varepsilon\in\{0,1\}. Morally we want to compute the determinant of the quadratic functional restricted to the curves which satisfy the constraints. The difficulty here is that the functional in question is a quadratic form on an infinite-dimensional space. We will compute the determinant using finite-dimensional approximations coming from discretizations of this continuous LQR problem with time step equal to 1/N1/N. We consider discrete LQR problems of the form

xk+1xk=1N(axk+uk),12Nk=0Nuk2ε(a2+b2)xk2min.x_{k+1}-x_{k}=\frac{1}{N}(ax_{k}+u_{k}),\,\,\frac{1}{2N}\sum_{k=0}^{N}u_{k}^{2}-\varepsilon(a^{2}+b^{2})x_{k}^{2}\to\min.

In order to apply our formula we introduce a new control function v=Nuv=\sqrt{N}u, which gives an equivalent optimal control problem:

xk+1xk=aNxk+vkN,12k=0Nvk2εa2+b2Nx2min.x_{k+1}-x_{k}=\frac{a}{N}x_{k}+\frac{v_{k}}{\sqrt{N}},\quad\frac{1}{2}\sum_{k=0}^{N}v_{k}^{2}-\varepsilon\frac{a^{2}+b^{2}}{N}x^{2}\to\min.

Hence we get as parameters of the LQR problem

Ak=aN1,Bk=N1/2,\displaystyle A_{k}=aN^{-1},\quad B_{k}=N^{-1/2},
Qk=ε(a2+b2)N1,Γk=N1.\displaystyle Q_{k}=\varepsilon(a^{2}+b^{2})N^{-1},\quad\Gamma_{k}=N^{-1}.

We can now use formulas (5) to show that we are actually computing the determinants of a tridiagonal Toeplitz matrices TN(r,s)T_{N}(r,s) from the previous subsection with parameters:

r=(a+N),s=N+N(1+aN)2εa2+b2N.r=-(a+N),\qquad s=N+N\left(1+\frac{a}{N}\right)^{2}-\varepsilon\frac{a^{2}+b^{2}}{N}.

Let us denote those matrices as T^ε(a,b,N)\hat{T}_{\varepsilon}(a,b,N).

As discussed previously we want to compute detT^ε(a,b,N)\det\hat{T}_{\varepsilon}(a,b,N) in two different ways and then take the limit NN\to\infty. Foreshadowing a bit the computation, the sequence detT^ε(a,b,N)\det\hat{T}_{\varepsilon}(a,b,N) does not converge. So instead we compute the limit

limNdetT^1(a,b,N)detT^0(a,b,N).\lim_{N\to\infty}\frac{\det\hat{T}_{1}(a,b,N)}{\det\hat{T}_{0}(a,b,N)}.

There are two motivations for using exactly this ratio. First of all we can notice from formula (26) that if two systems have the same matrices AkA_{k} and BkB_{k}, then the ratio of the corresponding determinants depends only on the associated discrete flows Φ\Phi. Since we have started with an approximation to a continuous system, it is natural to expect that the flows will converge and this way we obtain a finite quantity. Another reason is the celebrated Gelfand-Yaglom formula in quantum mechanics [10], which computes the ratio of determinants of a 1D quantum particle with a potential and a 1D free quantum particle as a ratio of determinants of two finite-rank matrices.

Let us first use formula (26) to compute the ratio. The two eigenvalues of the flow matrix now depend on a,b,N,εa,b,N,\varepsilon. To keep formulas clean we omit the dependence on a,b,Na,b,N in our notations and indicate explicit dependence on ε\varepsilon through a sub-index or super-index. We also assume that a>0a>0, b>0b>0. Other sign choices will be a direct consequence of (2) itself.

First note that rr is independent of ε\varepsilon. If we write Δε=sε2r2\Delta_{\varepsilon}=s_{\varepsilon}^{2}-r^{2}, then the formula (25) for eigenvalues reads as

μ±ε=sε±Δε2r.\mu^{\varepsilon}_{\pm}=\frac{-s_{\varepsilon}\pm\sqrt{\Delta_{\varepsilon}}}{2r}.

Formula (26) in this notations reads as

detT^1(a,b,N)detT^0(a,b,N)=((μ+1)1+N(μ1)1+N)((μ+0)1+N(μ0)1+N)(μ+0μ0)(μ+1μ1)\displaystyle\frac{\det\hat{T}_{1}(a,b,N)}{\det\hat{T}_{0}(a,b,N)}=\frac{\left((\mu_{+}^{1})^{1+N}-(\mu_{-}^{1})^{1+N}\right)}{\left((\mu_{+}^{0})^{1+N}-(\mu_{-}^{0})^{1+N}\right)}\frac{(\mu_{+}^{0}-\mu_{-}^{0})}{(\mu_{+}^{1}-\mu_{-}^{1})}
=Δ0Δ1((s1+Δ1)1+N(s1Δ1)1+N)((s0+Δ0)1+N(s0Δ0)1+N).\displaystyle=\frac{\sqrt{\Delta_{0}}}{\sqrt{\Delta_{1}}}\frac{\left((-s_{1}+\sqrt{\Delta_{1}})^{1+N}-(-s_{1}-\sqrt{\Delta_{1}})^{1+N}\right)}{\left((-s_{0}+\sqrt{\Delta_{0}})^{1+N}-(-s_{0}-\sqrt{\Delta_{0}})^{1+N}\right)}.

Plugging formulas for rr and sεs_{\varepsilon} we find that

Δε=4(1ε)a24εb2+O(N1),N.\Delta_{\varepsilon}=4(1-\varepsilon)a^{2}-4\varepsilon b^{2}+O(N^{-1}),\qquad N\to\infty.

Hence

limNΔ0Δ1=2a2bi.\lim_{N\to\infty}\frac{\sqrt{\Delta_{0}}}{\sqrt{\Delta_{1}}}=\frac{2a}{2bi}.

Similarly we find that

sε=2N+O(1),N.s_{\varepsilon}=2N+O(1),\qquad N\to\infty.

Hence

limN(s1±Δ1)1+N(2N)1+N=e±ib\lim_{N\to\infty}\frac{(-s_{1}\pm\sqrt{\Delta_{1}})^{1+N}}{(2N)^{1+N}}=e^{\pm ib}

and

limN(s0±Δ0)1+N(2N)1+N=e±a.\lim_{N\to\infty}\frac{(-s_{0}\pm\sqrt{\Delta_{0}})^{1+N}}{(2N)^{1+N}}=e^{\pm a}.

Thus after collecting everything together, we find

limNdetT^1(a,b,N)detT^0(a,b,N)=2a2bieibeibeaea=asinbbsinha\lim_{N\to\infty}\frac{\det\hat{T}_{1}(a,b,N)}{\det\hat{T}_{0}(a,b,N)}=\frac{2a}{2bi}\frac{e^{ib}-e^{-ib}}{e^{a}-e^{-a}}=\frac{a\sin b}{b\sinh a}

which is exactly the left side of (2).

Now we use formula (24) to compute the same ratio. We get

j=1N(N+N(1+aN)2+2(a+N)cosπjN+1a2+b2N)j=1N(N+N(1+aN)2+2(a+N)cosπjN+1)\displaystyle\frac{\prod_{j=1}^{N}\left(N+N(1+\frac{a}{N})^{2}+2(a+N)\cos\frac{\pi j}{N+1}-\frac{a^{2}+b^{2}}{N}\right)}{\prod_{j=1}^{N}\left(N+N(1+\frac{a}{N})^{2}+2(a+N)\cos\frac{\pi j}{N+1}\right)}
=detT^1(a,b,N)detT^0(a,b,N).\displaystyle=\frac{\det\hat{T}_{1}(a,b,N)}{\det\hat{T}_{0}(a,b,N)}.

Expanding and simplifying the denominator gives us

N+N(1+aN)2+2(a+N)cosπjN+1\displaystyle N+N\left(1+\frac{a}{N}\right)^{2}+2(a+N)\cos\frac{\pi j}{N+1}
=1N(a2+4(N2+aN)cos2πj2(N+1)).\displaystyle=\frac{1}{N}\left(a^{2}+4(N^{2}+aN)\cos^{2}\frac{\pi j}{2(N+1)}\right).

Hence

detT^1(a,b,N)detT^0(a,b,N)=j=1N(1a2+b2a2+4(N2+aN)cos2πj2(N+1)).\frac{\det\hat{T}_{1}(a,b,N)}{\det\hat{T}_{0}(a,b,N)}=\prod_{j=1}^{N}\left(1-\frac{a^{2}+b^{2}}{a^{2}+4(N^{2}+aN)\cos^{2}\frac{\pi j}{2(N+1)}}\right).

Changing the summation index jN+1jj\mapsto N+1-j gives

detT^1(a,b,N)detT^0(a,b,N)=j=1N(1a2+b2a2+4(N2+aN)sin2πj2(N+1)).\frac{\det\hat{T}_{1}(a,b,N)}{\det\hat{T}_{0}(a,b,N)}=\prod_{j=1}^{N}\left(1-\frac{a^{2}+b^{2}}{a^{2}+4(N^{2}+aN)\sin^{2}\frac{\pi j}{2(N+1)}}\right).

Since sin(y)|y|\sin(y)\leq|y|, we have the following inequality:

4(N2+aN)(πj)2sin2πj2(N+1)1+(a2)N14(N+1)2\frac{4(N^{2}+aN)}{(\pi j)^{2}}\sin^{2}\frac{\pi j}{2(N+1)}\leq 1+\frac{(a-2)N-1}{4(N+1)^{2}}

Thus for any ϵ>0\epsilon>0 there exists N0N_{0}\in\mathbb{N} such that NN0\forall N\geq N_{0}:

a2+b2a2+4(N2+aN)sin2πj2(N+1)a2+b2a2+(1+ϵ)(πj)2\frac{a^{2}+b^{2}}{a^{2}+4(N^{2}+aN)\sin^{2}\frac{\pi j}{2(N+1)}}\geq\frac{a^{2}+b^{2}}{a^{2}+(1+\epsilon)(\pi j)^{2}} (27)

Define the following partial product:

Πkm(N)=j=km(1a2+b2a2+4(N2+aN)sin2πj2(N+1)).\Pi_{k}^{m}(N)=\prod_{j=k}^{m}\left(1-\frac{a^{2}+b^{2}}{a^{2}+4(N^{2}+aN)\sin^{2}\frac{\pi j}{2(N+1)}}\right).

By continuity we have that:

limNΠ0N(N)=limNΠ0N0(N)limNΠN0N(N)\displaystyle\lim_{N\mapsto\infty}\Pi_{0}^{N}(N)=\lim_{N\mapsto\infty}\Pi_{0}^{N_{0}}(N)\lim_{N\to\infty}\Pi_{N_{0}}^{N}(N)
=j=0N0(1a2+b2a2+(πj)2)limNΠN0N(N).\displaystyle=\prod_{j=0}^{N_{0}}\left(1-\frac{a^{2}+b^{2}}{a^{2}+(\pi j)^{2}}\right)\lim_{N\to\infty}\Pi_{N_{0}}^{N}(N).

Since the second factor is the tail of a convergent product, thanks to the estimates in eq. 27, it converges to 11 when we take the limit for N0N_{0}\to\infty and get:

limN0j=0N0(1a2+b2a2+(πj)2)=limNΠ0N(N).\lim_{N_{0}\to\infty}\prod_{j=0}^{N_{0}}\left(1-\frac{a^{2}+b^{2}}{a^{2}+(\pi j)^{2}}\right)=\lim_{N\mapsto\infty}\Pi_{0}^{N}(N).

Thus we have proven (2).

References

  • [1] A. A. Agrachëv. Quadratic mappings in geometric control theory. In Problems in geometry, Vol. 20 (Russian), Itogi Nauki i Tekhniki, pages 111–205. Akad. Nauk SSSR, Vsesoyuz. Inst. Nauchn. i Tekhn. Inform., Moscow, 1988. Translated in J. Soviet Math. 51 (1990), no. 6, 2667–2734.
  • [2] A. A. Agrachev. Spectrum of the second variation. Tr. Mat. Inst. Steklova, 304(Optimal noe Upravlenie i Differentsial nye Uravneniya):32–48, 2019.
  • [3] Andrei Agrachev and Ivan Beschastnyi. Jacobi fields in optimal control: Morse and Maslov indices. Nonlinear Anal., 214:Paper No. 112608, 47, 2022.
  • [4] Baranzini S. Agrachev A. and I. Beschastnyi. Index theorems for graph-parametrized optimal control problems. Nonlinearity, Submitted.
  • [5] Kerstin Ammann and Gerald Teschl. Relative oscillation theory for jacobi matrices. arXiv:0810.5648, 2008.
  • [6] Stefano Baranzini. Functional determinants for the second variation. submitted to Journal of Differential Equations, 2022.
  • [7] Richard L. Burden, J. Douglas Faires, and Albert C. Reynolds. Numerical analysis. Prindle, Weber & Schmidt, Boston, Mass., 1978.
  • [8] Stephan D. Burton. The determinant and volume of 2-bridge links and alternating 3-braids. New York J. Math., 24:293–316, 2018.
  • [9] J. J. Duistermaat. On the Morse index in variational calculus. Advances in Math., 21(2):173–195, 1976.
  • [10] Gerald Dunne. Functional determinants in quantum field theory. Journal of Physics A: Mathematical and Theoretical, 41, 12 2007.
  • [11] Sondre Tesdal Galtung and Katrin Grunert. A numerical study of variational discretizations of the Camassa-Holm equation. BIT, 61(4):1271–1309, 2021.
  • [12] Alfred Hucht. The square lattice Ising model on the rectangle III: Hankel and Toeplitz determinants. J. Phys. A, 54(37):Paper No. 375201, 29, 2021.
  • [13] Said Kouachi. Eigenvalues and eigenvectors of tridiagonal matrices. Electron. J. Linear Algebra, 15:115–133, 2006.
  • [14] Yiming Long. Index Theory for Symplectic Paths with Applications. 01 2002.
  • [15] Luca Guido Molinari. Determinants of block tridiagonal matrices. Linear Algebra Appl., 429(8-9):2221–2226, 2008.
  • [16] Hermann Schulz-Baldes. Sturm intersection theory for periodic jacobi matrices and linear hamiltonian systems. Linear Algebra and its Applications, 436(3):498–515, 2012.
  • [17] Hermann Schulz-Baldes and Liam Urban. Space versus energy oscillations of prüfer phases for matrix sturm-liouville and jacobi operators. Electronic Journal of Differential Equations, 2020(76):1–23, 2020.