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

Finsler geometries on strictly accretive matrices

( Axel Ringh111Department of Electronic and Computer Engineering, The Hong Kong University of Science and Technology, Clear Water Bay, Kowloon, Hong Kong, China. Email: [email protected] (A. Ringh), [email protected] (L. Qiu)   and  Li Qiu111Department of Electronic and Computer Engineering, The Hong Kong University of Science and Technology, Clear Water Bay, Kowloon, Hong Kong, China. Email: [email protected] (A. Ringh), [email protected] (L. Qiu) )
Abstract

In this work we study the set of strictly accretive matrices, that is, the set of matrices with positive definite Hermitian part, and show that the set can be interpreted as a smooth manifold. Using the recently proposed symmetric polar decomposition for sectorial matrices, we show that this manifold is diffeomorphic to a direct product of the manifold of (Hermitian) positive definite matrices and the manifold of strictly accretive unitary matrices. Utilizing this decomposition, we introduce a family of Finsler metrics on the manifold and charaterize their geodesics and geodesic distance. Finally, we apply the geodesic distance to a matrix approximation problem, and also give some comments on the relation between the introduced geometry and the geometric mean for strictly accretive matrices as defined by S. Drury in [S. Drury, Linear Multilinear Algebra. 2015 63(2):296–301].

Key words: Accretive matrices; matrix manifolds; Finsler geometry; numerical range; geometric mean

22footnotetext: This work was supported by the Knut and Alice Wallenberg foundation, Stockholm, Sweden, under grant KAW 2018.0349, and the Hong Kong Research Grants Council, Hong Kong, China, under project GRF 16200619.

1 Introduction

Given a complex number zz\in\mathbb{C} we can write it in its Cartesian form z=a+ibz=a+ib, where a=(z)a=\Re(z) is the real part and b=(z)b=\Im(z) is the imaginary part, or we can write it in its polar form as z=reiθz=re^{i\theta}, where r=|z|r=|z| is the magnitude and θ=z\theta=\angle\,z is the phase. The standard metric on \mathbb{C} defines the (absolute) distance between z1z_{1} and z2z_{2} as |z1z2|=(z1z2)2+(z1z2)2|z_{1}-z_{2}|=\sqrt{\Re(z_{1}-z_{2})^{2}+\Im(z_{1}-z_{2})^{2}} which is efficiently computed using the Cartesian form as (a1a2)2+(b1b2)2\sqrt{(a_{1}-a_{2})^{2}+(b_{1}-b_{2})^{2}}. However, sometimes a logarithmic (relative) distance between the numbers contains information that is more relevant for the problem at hand. One such distance is given by log(r1/r2)2+[(θ1θ2)mod 2π]2\sqrt{\log(r_{1}/r_{2})^{2}+[(\theta_{1}-\theta_{2})\mid\text{mod }2\pi]^{2}}, and in this distance measure the point 11 is as close to 10eiθ10e^{i\theta} as it is to 0.1eiθ0.1e^{-i\theta}. This type of distances have wide application in engineering problems, e.g., as demonstrated in the use of Bode plots and Nichols charts in control theory [41]. Moreover, this type of logarithmic metric has been generalized to (Hermitian) positive definite matrices, with plenty of applications, for example in computing geometric means between such matrices [37], [11, Chp. 6], [29, Chp. XII]. This generalization can done by identifying the set of positive matrices as a smooth manifold and introducing a Riemannian or Finsler metric on it. Here, we follow a similar path and extend this type of logarithmic metrics to so called strictly accretive matrices. More specifically, the outline of the paper is as follows: in Section 2 we review relevant background material and set up the notation used in the paper. Section 3 is devoted to showing that the set of strictly accretive matrices can be interpreted as a smooth manifold, and that this manifold is diffeomorphic to a direct product of the smooth manifold of positive definite matrices and the smooth manifold of strictly accretive unitary matrices. The latter is done using the newly introduced symmetric polar decomposition for sectorial matrices [44]. In Section 4 we introduce a family of Finsler metrics on the manifold, by means of the decomposition from the previous section and so called (Minkowskian) product functions [39]. In particular, this allows us to characterize the corresponding geodesics and compute the geodesic distance. Finally, in Section 5 we given an application of the metric to a matrix approximation problem and also give some comments on the relation between the geodesic midpoint and the geometric mean between strictly accretive matrices as introduced in [16].

2 Background and notation

In the following section we introduce some background material needed for the rest of the paper. At the same time, this section is also be used to set up the notation used throughout. To this end, let 𝕄n{\mathbb{M}}_{n} denote the set of n×nn\times n matrices over the filed \mathbb{C} of complex numbers. For A𝕄nA\in{\mathbb{M}}_{n}, let AA^{*} denotes its complex conjugate transpose, let H(A):=12(A+A)H(A):=\tfrac{1}{2}(A+A^{*}) denote its Hermitian part, and let S(A):=12(AA)S(A):=\tfrac{1}{2}(A-A^{*}) denote its skew-Hermitian part.111Note that this is equivalent to the Toeplitz decomposition since if A=(A)+i(A)A=\Re(A)+i\Im(A), then (A)=H(A)\Re(A)=H(A) and (A)=1iS(A)\Im(A)=\tfrac{1}{i}S(A), see [23, p. 7]. Moreover, by II we denote the identity matrix, and for A𝕄nA\in{\mathbb{M}}_{n} by λ(A)\lambda(A) we denote its spectrum, i.e., λ(A):={λdet(λIA)=0}\lambda(A):=\{\lambda\in\mathbb{C}\mid\det(\lambda I-A)=0\}, and by σ(A)\sigma(A) we denote it singular values, i.e., σ(A)=λ(AA)\sigma(A)=\sqrt{\lambda(A^{*}A)}.

The following is a number of different sets of matrices that will be used throughout: 𝔾𝕃n{\mathbb{G}\mathbb{L}_{n}} denotes the set of invertible matrices, 𝕌n{\mathbb{U}}_{n} denotes the set of unitary matrices, n{\mathbb{H}}_{n} denotes the set of Hermitian matrices, n{\mathbb{P}}_{n} denotes the set of positive definite matrices, i.e., AnA\in{\mathbb{H}}_{n} s.t. λ(A)+{0}\lambda(A)\subset\mathbb{R}_{+}\setminus\{0\}, 𝕊n\mathbb{S}_{n} denotes the set of skew-Hermitian matrices, and 𝔸n{\mathbb{A}}_{n} denotes the set of strictly accretive matrices, i.e., A𝔸nA\in{\mathbb{A}}_{n} if and only if H(A)nH(A)\in{\mathbb{P}}_{n}.222The naming used here is the same as in [28, p. 281], in contrast to [8].

Two matrices A,B𝕄nA,B\in{\mathbb{M}}_{n} are said to be congruent if there exists a matrix C𝔾𝕃nC\in{\mathbb{G}\mathbb{L}_{n}} such that A=CBCA=C^{*}BC. For matrices A,B𝕄nA,B\in{\mathbb{M}}_{n} we define the inner product A,B:=tr(AB)\langle A,B\rangle:={\text{\rm tr}}(A^{*}B), which gives the Frobenius norm A:=A,A=j=1nσj(A)2\|A\|:=\sqrt{\langle A,A\rangle}=\sqrt{\sum_{j=1}^{n}\sigma_{j}(A)^{2}}. By sp\|\cdot\|_{\text{sp}} we denote the spectral norm, i.e., Asp=supxn{0}Ax2/x2=σmax(A)\|A\|_{\text{sp}}=\sup_{x\in\mathbb{C}^{n}\setminus\{0\}}\|Ax\|_{2}/\|x\|_{2}=\sigma_{\max}(A), the larges singular value of AA. Next, a function Φ:n\Phi:\mathbb{R}^{n}\to\mathbb{R} is called a symmetric gauge function if for all x,ynx,y\in\mathbb{R}^{n} and all β\beta\in\mathbb{R} i) Φ(x)>0\Phi(x)>0 if x0x\neq 0, ii) Φ(βx)=|β|Φ(x)\Phi(\beta x)=|\beta|\Phi(x), iii) Φ(x+y)Φ(x)+Φ(y)\Phi(x+y)\leq\Phi(x)+\Phi(y), and iv) Φ(x)=Φ(x~)\Phi(x)=\Phi(\tilde{x}) for all x~=[±xα(i)]i=1n\tilde{x}=[\pm x_{\alpha(i)}]_{i=1}^{n} where α\alpha is any permutation of {1,,n}\{1,\ldots,n\} [36], [34, Sec. 3.I.1]. For any unitary invariant norm, i.e., norms |||\|\cdot\|| such that |UAV|=|A||\|UAV\||=|\|A\|| for all A𝕄nA\in{\mathbb{M}}_{n} and all U,V𝕌nU,V\in{\mathbb{U}}_{n}, there exists a symmetric gauge function Φ\Phi such that |A|=Φ(σ(A))|\|A\||=\Phi(\sigma(A)) [10, Thm IV.2.1 ], [34, Thm. 10.A.1]. For this reason we will henceforth denote such norms Φ\|\cdot\|_{\Phi}. Moreover, we will call a symmetric gauge function, and the corresponding norm, smooth if it is smooth outside of the origin, cf. [32, Thm. 8.5].

For a vector xnx\in\mathbb{R}^{n}, by xx^{\downarrow} we denote the vector obtained by sorting the elements in xx in a nonincreasing order. More precisely, xx^{\downarrow} is obtained by permuting the elements of xx such that x=[xk]k=1nx^{\downarrow}=[x_{k}^{\downarrow}]_{k=1}^{n} where x1x2xnx_{1}^{\downarrow}\geq x_{2}^{\downarrow}\geq\ldots\geq x_{n}^{\downarrow}. For two vectors x,ynx,y\in\mathbb{R}^{n}, we say that xx is submajorized (weakly submajorized) by yy if k=1xkk=1yk\sum_{k=1}^{\ell}x_{k}^{\downarrow}\leq\sum_{k=1}^{\ell}y_{k}^{\downarrow} for =1,,n1\ell=1,\ldots,n-1 and k=1nxk=()k=1nyk\sum_{k=1}^{n}x_{k}^{\downarrow}=(\leq)\sum_{k=1}^{n}y_{k}^{\downarrow} [34, p. 12]. Submajorization (weak submajorization) is a preorder on n\mathbb{R}^{n}, and we write x(w)yx\prec(\prec_{w})y. On the equivalence classes of vectors sorted in nonincreasing order it is a partial ordering [34, p. 19].

2.1 Sectorial matrices and the phases of a matrix

Given a matrix A𝕄nA\in{\mathbb{M}}_{n}, we define the numerical range (field of values) as

W(A):={zz=xAx,xn and x2=xx=1}.W(A):=\big{\{}z\in\mathbb{C}\mid z=x^{*}Ax,\;x\in\mathbb{C}^{n}\text{ and }\|x\|^{2}=x^{*}x=1\big{\}}.

Using the numerical range, we can define the set of so called sectorial matrices as

𝕎n:={A𝕄n0W(A)}.\mathbb{W}_{n}:=\{A\in{\mathbb{M}}_{n}\mid 0\not\in W(A)\}.

The name comes from the fact that the numerical range of a matrix A𝕎nA\in\mathbb{W}_{n} is contained in a sector of opening angle less than π\pi. The latter can be seen from the Toeplitz-Hausdorff theorem, which states that for any matrix A𝕄nA\in{\mathbb{M}}_{n}, W(A)W(A) is a convex set [45, Thm. 4.1], [20, Thm. 1.1-2]. Recently, sectorial matrices have received considerable attention in the literature, see, e.g., [8, 35, 17, 33, 46, 44, 12].

Sectorial matrices have several interesting properties. In particular, if AA is sectorial it is congruent to a unitary diagonal matrix DD, i.e., A=TDTA=T^{*}DT for some T𝔾𝕃nT\in{\mathbb{G}\mathbb{L}_{n}} [21, 15, 19, 26, 24]. Although the decomposition is not unique, the elements in DD are unique up to permutation, and any such decomposition is called a sectorial decoposition [46]. Using this decomposition, we define the phases of AA, denoted ϕ1(A),ϕ2(A),,ϕn(A)\phi_{1}(A),\phi_{2}(A),\ldots,\phi_{n}(A), as the phases of the eigenvalues of DD [19, 44, 12];333In [19], these were called canonical angles. by convention we defined them to belong the an interval of length strictly less than π\pi. With this definition we have, e.g., that A𝔸nA\in{\mathbb{A}}_{n} if and only if A𝕎nA\in\mathbb{W}_{n} and H(A)nH(A)\in{\mathbb{P}}_{n}, which is true if and only if (ϕ1(A),,ϕn(A))(π/2,π/2)(\phi_{1}(A),\ldots,\phi_{n}(A))\subset(-\pi/2,\pi/2). Note that the phases of a sectorial matrix AA is different from the angles of the eigenvalues, i.e., in general ϕ(A)φ(A)\phi(A)\neq\varphi(A) where φ(A):=λ(A)\varphi(A):=\angle\,\lambda(A). More precisely, equality holds for normal matrices. The phases have a number of desirable properties that the angles of the eigenvalues do not, see [44].

Another decomposition of sectorial matrices, which will in fact be central to this work, is the so called symmetric polar decomposition [44, Thm. 3.1]: for A𝕎nA\in\mathbb{W}_{n} there is a unique decomposition given by

A=PUP,A=PUP,

where PnP\in{\mathbb{P}}_{n} and U𝕎𝕌n:=𝕌n𝕎nU\in\mathbb{W}{\mathbb{U}}_{n}:={\mathbb{U}}_{n}\cap\mathbb{W}_{n}. The latter is the set of sectorial unitary matrices. The phases of AA are given by the phases of UU, which is in fact the phases of the eigenvalues of UU. Therefore, we have that A𝔸nA\in{\mathbb{A}}_{n} if and only if it has a symmetric polar decomposition such that U𝔸𝕌n:={U𝕎𝕌nH(U)n}U\in{\mathbb{A}}{\mathbb{U}}_{n}:=\{U\in\mathbb{W}{\mathbb{U}}_{n}\mid H(U)\in{\mathbb{P}}_{n}\}, i.e., the set of strictly accretive unitary matrices.

2.2 Riemannian and Finsler manifolds

Smooth manifolds are important mathematical objects which show up in such diverse fields as theoretical physics [40], robotics [38], and statistics and information theory [2]. Intuitively, they can be thought of as spaces that locally look like the Euclidean space, and on these spaces one can introduce geometric concepts such as curves and metrics. In particular, all smooth manifolds admit a so called Riemannian metric [27, Thm. 1.4.1], [30, Prop. 13.3], and Riemannian geometry is a well-studied subjects, see, e.g., one of the monographs [40, 29, 27, 30, 31]. An relaxation of Riemannian geometry leads to so called Finsler geometry [14]; loosely expressed it can be interpreted as chaining the tangent space from a Hilbert space to a Banach space. For an introduction to Finsler geometry, see, e.g., [9, 42, 13].

More specifically, given a smooth manifold \mathcal{M}, for xx\in\mathcal{M} we denote the tangent space by TxT_{x}\mathcal{M} and the tangent bundle by T:=x{x}×TxT\mathcal{M}:=\cup_{x\in\mathcal{M}}\{x\}\times T_{x}\mathcal{M}. A Riemannian metric is induced by an inner product on the tangent space, ,x:Tx×Tx\langle\cdot,\cdot\rangle_{x}:T_{x}\mathcal{M}\times T_{x}\mathcal{M}\to\mathbb{R}, that varies smoothly with the base point xx. Using this inner product, one defines the norm ,x\sqrt{\langle\cdot,\cdot\rangle_{x}}, which in fact defines a smooth function on the slit tangent bundle Tx(x,0)T\mathcal{M}\setminus\cup_{x\in\mathcal{M}}(x,0). In this work we consider Finsler structures on smooth (matrix) manifolds, but we will limit the scope to Finsler structures F:T+F:T\mathcal{M}\to\mathbb{R}_{+}, (x,X)Xx(x,X)\mapsto\|X\|_{x}, where x\|\cdot\|_{x} is a norm on Tx𝒳T_{x}\mathcal{X} which is not necessarily induced by an inner product, and such that FF is smooth on the slit tangent bundle Tx(x,0)T\mathcal{M}\setminus\cup_{x\in\mathcal{M}}(x,0). Given a piece-wise smooth curve γ:[0,1]\gamma:[0,1]\to\mathcal{M}, the arc length on the manifold is defined using this Finsler structure. More precisely, it is defined as

(γ):=01F(γ(t),γ˙(t))𝑑t,\mathcal{L}(\gamma):=\int_{0}^{1}F(\gamma(t),\dot{\gamma}(t))dt,

where γ˙(t)\dot{\gamma}(t) is the derivative of γ\gamma with respect to tt. Using arc length, the geodesic distance between two points x,yx,y\in\mathcal{M} is defined as

δ(x,y):=infγ(γ):γ is a piece-wise smooth curve such that γ(0)=x,γ(1)=y,\delta(x,y):=\inf_{\gamma}\mathcal{L}(\gamma)\;:\;\gamma\text{ is a piece-wise smooth curve such that }\gamma(0)=x,\gamma(1)=y,

and a minimizing curve (if one exists) is called a geodesic. A final notion we need is that of diffeomorphic manifolds. More precisely, two smooth manifolds \mathcal{M} and 𝒩\mathcal{N} are said to be diffeomorphic if there exists a diffeomorphism f:𝒩f:\mathcal{M}\to\mathcal{N}, i.e., a function ff which is a smooth bijection with a smooth inverse. In this case we write 𝒩\mathcal{M}\cong\mathcal{N}.

Next, we summarize some results regarding two matrix manifolds, namely n{\mathbb{P}}_{n} and 𝕌n{\mathbb{U}}_{n}, together with specific Finsler structures. These will be needed later.

2.2.1 A family of Finsler metrics on n{\mathbb{P}}_{n} and their geodesics

Riemannian and Finsler geometry on n{\mathbb{P}}_{n} is a well-studied subject, and we refer the reader to, e.g., [37], [11, Chp. 6] or [29, Chp. XII]. Here, we summarize some of the results we will need for later. To this end, note that the tangent space at PnP\in{\mathbb{P}}_{n} is n{\mathbb{H}}_{n}, and given PnP\in{\mathbb{P}}_{n}, X,YnX,Y\in{\mathbb{H}}_{n} we can introduce the inner product on the tangent space as

X,YP=tr((P1/2XP1/2)(P1/2YP1/2)),\langle X,Y\rangle_{P}={\text{\rm tr}}((P^{-1/2}XP^{-1/2})^{*}(P^{-1/2}YP^{-1/2})), (2.1)

with corresponding norm Fn(P,X)=XP=j=1nσj(P1/2XP1/2)2F_{{\mathbb{P}}_{n}}(P,X)=\|X\|_{P}=\sqrt{\sum_{j=1}^{n}\sigma_{j}(P^{-1/2}XP^{-1/2})^{2}}. The geodesic between P,QnP,Q\in{\mathbb{P}}_{n} in the induced Riemannian metric is given by

γn(t)=P1/2(P1/2QP1/2)tP1/2=Petlog(P1Q)\gamma_{{\mathbb{P}}_{n}}(t)=P^{1/2}(P^{-1/2}QP^{-1/2})^{t}P^{1/2}=Pe^{t\log(P^{-1}Q)} (2.2)

and the length of the curve, i.e., the Riemannian distance between PP and QQ, is given by

δn(P,Q)=log(P1/2QP1/2).\delta_{{\mathbb{P}}_{n}}(P,Q)=\|\log(P^{-1/2}QP^{-1/2})\|.

Interestingly, if the norm P\|\cdot\|_{P} on TPnT_{P}{\mathbb{P}}_{n} is changed to any other unitary invariant matrix norm

Φ,P=Φ(σ(P1/2P1/2))\|\cdot\|_{\Phi,P}=\Phi(\sigma(P^{-1/2}\cdot P^{-1/2}))

the expressions for a geodesic between two matrices remains unchanged and the corresponding distance is given by δnΦ(P,Q)=log(P1/2QP1/2)Φ\delta_{{\mathbb{P}}_{n}}^{\Phi}(P,Q)=\|\log(P^{-1/2}QP^{-1/2})\|_{\Phi} [11, Sec. 6.4]. However, the geodesic (2.2) might no longer be the unique shortest curve [11, p. 223].

An alternative expression for the geodesic (2.2) is given by the following proposition.

Proposition 2.1.

Let Φ\|\cdot\|_{\Phi} be any smooth unitarily invariant norm and consider the Finsler structure given by FnΦ:Tn+F_{{\mathbb{P}}_{n}}^{\Phi}:T{\mathbb{P}}_{n}\to\mathbb{R}_{+}, FnΦ:(P,X)P1/2XP1/2ΦF_{{\mathbb{P}}_{n}}^{\Phi}:(P,X)\mapsto\|P^{-1/2}XP^{-1/2}\|_{\Phi}. For P,QnP,Q\in{\mathbb{P}}_{n}, a geodesic between them can be written as γ(t)=SΛtS\gamma(t)=S\Lambda^{t}S^{*} where P=SSP=SS^{*}, Q=SΛSQ=S\Lambda S^{*} is a simultaneous diagonalization by congruence of PP and QQ, i.e., S𝔾𝕃nS\in{\mathbb{G}\mathbb{L}_{n}} and Λ\Lambda is diagonal with positive elements on the diagonal. Moreover, the geodesic distance is

δnΦ(P,Q)=log(P1/2QP1/2)Φ=Φ(log(λ(P1Q)))=log(Λ)Φ.\delta_{{\mathbb{P}}_{n}}^{\Phi}(P,Q)=\|\log(P^{-1/2}QP^{-1/2})\|_{\Phi}=\Phi\Big{(}\log\big{(}\lambda(P^{-1}Q)\big{)}\Big{)}=\|\log(\Lambda)\|_{\Phi}. (2.3)
Proof.

To show the second equality in (2.3), note that

log(P1/2QP1/2)Φ\displaystyle\|\log(P^{-1/2}QP^{-1/2})\|_{\Phi} =Φ(σ(log(P1/2QP1/2)))=Φ(|λ(log(P1/2QP1/2))|)\displaystyle=\Phi\Big{(}\sigma\big{(}\log(P^{-1/2}QP^{-1/2})\big{)}\Big{)}=\Phi\Big{(}|\lambda\big{(}\log(P^{-1/2}QP^{-1/2})\big{)}|\Big{)}
=Φ(λ(log(P1/2QP1/2)))=Φ(log(λ(P1/2QP1/2)))\displaystyle=\Phi\Big{(}\lambda\big{(}\log(P^{-1/2}QP^{-1/2})\big{)}\Big{)}=\Phi\Big{(}\log\big{(}\lambda(P^{-1/2}QP^{-1/2})\big{)}\Big{)}
=Φ(log(λ(P1Q)))\displaystyle=\Phi\Big{(}\log\big{(}\lambda(P^{-1}Q)\big{)}\Big{)}

where the second equality comes from that the singular values of a Hermitian matrix, i.e., the matrix log(P1/2QP1/2)\log(P^{-1/2}QP^{-1/2}), are the absolute values of the eigenvalues, the third equality follows since the symmetric gauge function is invariant under sign changes, the forth can be seen by using a unitary diagonalization of P1/2QP1/2P^{-1/2}QP^{-1/2}, and the fifth equality comes from that the spectrum is invariant under similarity. Next, since both P,QnP,Q\in{\mathbb{P}}_{n}, by [23, Thm. 7.6.4] we can simultaneously diagonalize PP and QQ by congruence, i.e., there exists an S𝔾𝕃nS\in{\mathbb{G}\mathbb{L}_{n}} such that P=SSP=SS^{*} and Q=SΛSQ=S\Lambda S^{*}, where Λ=diag(λ1,,λn)\Lambda={\text{\rm diag}}(\lambda_{1},\ldots,\lambda_{n}) and where λ1,,λn\lambda_{1},\ldots,\lambda_{n} are all strictly larger than 0. In fact, λ1,,λn\lambda_{1},\ldots,\lambda_{n} are the eigenvalues of P1QP^{-1}Q, which means that log(λ(P1Q))=log(λ(Λ))=log(Λ),\log\big{(}\lambda(P^{-1}Q)\big{)}=\log(\lambda(\Lambda))=\log(\Lambda), which in turn gives the last equality in (2.3). Finally, this also gives that γ(t)=Petlog(P1Q)=SSSetlog(Λ)S=SΛtS.\gamma(t)=Pe^{t\log(P^{-1}Q)}=SS^{*}S^{-*}e^{t\log(\Lambda)}S^{*}=S\Lambda^{t}S^{*}.

2.2.2 A family of Finsler metrics on 𝕌n{\mathbb{U}}_{n} and their geodesics

The set of unitary matrices is a Lie group, and results related to Riemannian and Finsler geometry on 𝕌n{\mathbb{U}}_{n} can be found in, e.g., [3, 5, 4, 7, 6]. Again, we here summarize some of the results that we will need for later. To this end, note that the tangent space at U𝕌nU\in{\mathbb{U}}_{n} is 𝕊n\mathbb{S}_{n} and given U𝕌nU\in{\mathbb{U}}_{n}, X,Y𝕊nX,Y\in\mathbb{S}_{n} we can introduce the inner product on the tangent space as

X,YU=tr(XY),\langle X,Y\rangle_{U}={\text{\rm tr}}(X^{*}Y), (2.4)

with corresponding induced norm F𝕌n(U,X)=XU=X,XUF_{{\mathbb{U}}_{n}}(U,X)=\|X\|_{U}=\sqrt{\langle X,X\rangle_{U}}. The induced Riemannian metric have shortest curves between U,V𝕌nU,V\in{\mathbb{U}}_{n} given by

γ𝕌n(t)=UeitZ,\gamma_{{\mathbb{U}}_{n}}(t)=Ue^{itZ}, (2.5)

where V=UeiZV=Ue^{iZ}, and where ZnZ\in{\mathbb{H}}_{n} is such that Zspπ\|Z\|_{\text{sp}}\leq\pi. Moreover, the geodesic distance is

δ𝕌n(U,V)=Z,\delta_{{\mathbb{U}}_{n}}(U,V)=\|Z\|, (2.6)

and the geodesic is unique if Zsp<π\|Z\|_{\text{sp}}<\pi. Similarly to the results for the smooth manifold n{\mathbb{P}}_{n}, if the norm U\|\cdot\|_{U} on TU𝕌nT_{U}{\mathbb{U}}_{n} is changed to any other unitary invariant matrix norm Φ,U\|\cdot\|_{\Phi,U} the expressions for a geodesic in (2.5) is unchanged and the expression for the geodesic distance (2.6) is only changed to using the corresponding norm Φ\|\cdot\|_{\Phi} [4, 7]. However, even if Zsp<π\|Z\|_{\text{sp}}<\pi the geodesic (2.5) might not be unique in this case [7, Sec. 3.2].

3 The smooth manifold of strictly accretive matrices

In this section we prove that 𝔸n{\mathbb{A}}_{n} is a smooth manifold diffemorphic to n×𝔸𝕌n{\mathbb{P}}_{n}\times{\mathbb{A}}{\mathbb{U}}_{n}. The latter is fundamental for the introduction of the Finsler structures in the following section. For improved readability, some of the technical results of this section are deferred to Appendix A. To this end, we start by proving the following.

Theorem 3.1.

𝔸n{\mathbb{A}}_{n} is a connected smooth manifold, and at a point A𝔸nA\in{\mathbb{A}}_{n} the tangent space is TA𝔸n=𝕄nT_{A}{\mathbb{A}}_{n}={\mathbb{M}}_{n}.

Proof.

This follows by Lemma A.1, A.2, and A.3, and applying [30, Ex. 1.26]. ∎

Next, we prove the following characterization of the manifold.

Theorem 3.2.

𝔸nn×𝔸𝕌n{\mathbb{A}}_{n}\cong{\mathbb{P}}_{n}\times{\mathbb{A}}{\mathbb{U}}_{n}, where n{\mathbb{P}}_{n} and 𝔸𝕌n{\mathbb{A}}{\mathbb{U}}_{n} are embedded submanifolds.

This theorem follows as a corollary to the following proposition.

Proposition 3.3.

For A𝔸nA\in{\mathbb{A}}_{n}, let A=PUPA=PUP be the symmetric polar decomposition. Then the mapping A(P2,U)A\mapsto(P^{2},U) is a diffeomorphism between the smooth manifolds 𝔸n{\mathbb{A}}_{n} and n×𝔸𝕌n{\mathbb{P}}_{n}\times{\mathbb{A}}{\mathbb{U}}_{n}.

Proof.

Since n{\mathbb{P}}_{n} and 𝔸𝕌n{\mathbb{A}}{\mathbb{U}}_{n} are smooth manifolds (see [11, Chp. 6] and Lemma A.4, respectively), n×𝔸𝕌n{\mathbb{P}}_{n}\times{\mathbb{A}}{\mathbb{U}}_{n} is also a smooth manifold [30, Ex. 1.34]. Next, note that the matrix square root is a diffeomorphism of n{\mathbb{P}}_{n} to itself, with the matrix square as the inverse, cf. [23, Thm. 7.2.6]. Therefore, it suffices to show that the mapping A(P,U)A\mapsto(P,U) is a diffeomorphism between 𝔸n{\mathbb{A}}_{n} and n×𝔸𝕌n{\mathbb{P}}_{n}\times{\mathbb{A}}{\mathbb{U}}_{n}. To this end, first observe that the latter is a bijection due to the existence and uniqueness of a symmetric polar decomposition [44, Thm. 3.1]. Moreover, that the inverse is smooth follows since the components in AA are polynomial in the components in PP and UU.

To show that PP and UU are smooth in AA, we note that since AA is strictly accretive, H(A)0H(A)\succ 0. This means that we can write

A\displaystyle A =H(A)+i(1iS(A))=H(A)1/2(I+iH(A)1/21iS(A)H(A)1/2)H(A)1/2\displaystyle=H(A)+i(\tfrac{1}{i}S(A))=H(A)^{1/2}\big{(}I+iH(A)^{-1/2}\tfrac{1}{i}S(A)H(A)^{-1/2}\big{)}H(A)^{1/2}
=H(A)1/2KH(A)1/2,\displaystyle=H(A)^{1/2}KH(A)^{1/2},

where K:=I+iH(A)1/21iS(A)H(A)1/2K:=I+iH(A)^{-1/2}\tfrac{1}{i}S(A)H(A)^{-1/2} is a normal matrix444To see this, note that a matrix AA is normal if and only if H(A)H(A) and 1iS(A)\tfrac{1}{i}S(A) commute, see, e.g., [45, Thm. 9.1]. (cf. [46, Proof of Cor. 2.5]) which by construction depends smoothly on AA. Now, let K=VKQKK=V_{K}Q_{K} be the polar decomposition of KK. Since KK depends smoothly on AA, and since the polar decomposition is smooth in the matrix (Lemma A.5), VKV_{K} and QKQ_{K} are smooth in AA. Moreover, since KK is normal, VKV_{K} and QKQ_{K} commute [45, Thm. 9.1], and thus VKV_{K} and QK1/2Q_{K}^{1/2} commute. Therefore, A=H(A)1/2QK1/2UKQK1/2H(A)1/2A=H(A)^{1/2}Q_{K}^{1/2}U_{K}Q_{K}^{1/2}H(A)^{1/2}, where all components depend smoothly on AA. Now, let L:=QK1/2H(A)1/2L:=Q_{K}^{1/2}H(A)^{1/2} and let L=VLQLL=V_{L}Q_{L} be the polar decomposition of LL. Similar to before, VLV_{L} and QLQ_{L} are thus both smooth in AA. Finally, we thus have that A=QLVLUKVLQL,A=Q_{L}V_{L}^{*}U_{K}V_{L}Q_{L}, and since VLV_{L} and UKU_{K} are unitary so is VLUKVLV_{L}^{*}U_{K}V_{L}. By the uniqueness of the symmetric polar decomposition [44, Thm. 3.1] it follows that P=QLP=Q_{L} and U=VLUKVLU=V_{L}^{*}U_{K}V_{L}, which are both smooth in AA. ∎

Proof of Theorem 3.2.

By Proposition 3.3, 𝔸nn×𝔸𝕌n{\mathbb{A}}_{n}\cong{\mathbb{P}}_{n}\times{\mathbb{A}}{\mathbb{U}}_{n}, and by [30, Prop. 5.3], both n×{I}n{\mathbb{P}}_{n}\times\{I\}\cong{\mathbb{P}}_{n} and {I}×𝔸𝕌n𝔸𝕌n\{I\}\times{\mathbb{A}}{\mathbb{U}}_{n}\cong{\mathbb{A}}{\mathbb{U}}_{n} are embedded submanifolds of n×𝔸𝕌n{\mathbb{P}}_{n}\times{\mathbb{A}}{\mathbb{U}}_{n}. ∎

Remark 3.4.

The results in Theorem 3.1 and 3.2 can easily be generalized to other subsets of sectorial matrices, namely any subset 𝔸~n𝕎n\tilde{{\mathbb{A}}}_{n}\subset\mathbb{W}_{n} of all matrices AA such that there exists α,β\alpha,\beta\in\mathbb{R}, α<β\alpha<\beta, and βα=π\beta-\alpha=\pi, for which mink=1,nϕk(A)>α\min_{k=1,\ldots n}\phi_{k}(A)>\alpha and maxk=1,nϕk(A)<β\max_{k=1,\ldots n}\phi_{k}(A)<\beta. To see this, note that for A~=P~U~P~𝔸~n\tilde{A}=\tilde{P}\tilde{U}\tilde{P}\in\tilde{{\mathbb{A}}}_{n} we have that A=P~(e(β+α)/2U~)P~𝔸nA=\tilde{P}(e^{-(\beta+\alpha)/2}\tilde{U})\tilde{P}\in{\mathbb{A}}_{n}, i.e., that e(β+α)/2𝔸~n=𝔸ne^{-(\beta+\alpha)/2}\tilde{{\mathbb{A}}}_{n}={\mathbb{A}}_{n} with a diffeomorphism between the components in the symmetric polar decomposition. Examples of such sets of matrices are and the set of strictly dissipative matrices, i.e., matrices such that (A)0\Re(A)\prec 0 [28, p. 279], and matrices such that (A)0\Im(A)\succ 0.555Note that the latter has unfortunately also been termed dissipative in the literature, see [28, p. 279].

4 A family of Finsler metrics on 𝔸n{\mathbb{A}}_{n} and their geodesics

In this section we introduce a family of Finsler structures on 𝔸n{\mathbb{A}}_{n} and in particular we will characterize the geodesics and geodesic distances corresponding to these structures. To this end, by Theorem 3.2 we have that 𝔸nn×𝔸𝕌n{\mathbb{A}}_{n}\cong{\mathbb{P}}_{n}\times{\mathbb{A}}{\mathbb{U}}_{n}. Moreover, n{\mathbb{P}}_{n} and 𝕌n{\mathbb{U}}_{n} are smooth manifolds that are well-studied in the literature, and since n{\mathbb{P}}_{n} and 𝔸𝕌n{\mathbb{A}}{\mathbb{U}}_{n} are embedded submanifolds of 𝔸n{\mathbb{A}}_{n} a desired property would be that when restricted to any of the two embedded submanifold the introduced Finsler structure would yield the corresponding known Finsler structure. To this end, we first characterize the geodesics and the geodesic distance on 𝔸𝕌n{\mathbb{A}}{\mathbb{U}}_{n}.

Proposition 4.1.

Let Φ\|\cdot\|_{\Phi} be any smooth unitarily invariant norm and consider the Finsler structure given by F𝔸𝕌nΦ:T𝔸𝕌n+F_{{\mathbb{A}}{\mathbb{U}}_{n}}^{\Phi}:T{\mathbb{A}}{\mathbb{U}}_{n}\to\mathbb{R}_{+}, F𝔸𝕌nΦ:(U,X)XΦF_{{\mathbb{A}}{\mathbb{U}}_{n}}^{\Phi}:(U,X)\mapsto\|X\|_{\Phi}. A geodesic between U𝔸𝕌nU\in{\mathbb{A}}{\mathbb{U}}_{n} and V𝔸𝕌nV\in{\mathbb{A}}{\mathbb{U}}_{n} is given by

γ𝔸𝕌n(t)=Uetlog(U1V)=U1/2(U1/2VU1/2)tU1/2.\gamma_{{\mathbb{A}}{\mathbb{U}}_{n}}(t)=Ue^{t\log(U^{-1}V)}=U^{1/2}(U^{-1/2}VU^{-1/2})^{t}U^{1/2}.

Moreover, the geodesic distance is given by

δ𝔸𝕌nΦ(U,V)=log(U1V)Φ=log(U1/2VU1/2)Φ.\delta_{{\mathbb{A}}{\mathbb{U}}_{n}}^{\Phi}(U,V)=\|\log(U^{-1}V)\|_{\Phi}=\|\log(U^{-1/2}VU^{-1/2})\|_{\Phi}. (4.1)
Proof.

Let U,V𝔸𝕌nU,V\in{\mathbb{A}}{\mathbb{U}}_{n}. The proposition follows if we can show that a geodesic on 𝕌n{\mathbb{U}}_{n} between UU and VV, given by (2.5), remains in 𝔸𝕌n{\mathbb{A}}{\mathbb{U}}_{n} for t[0,1]t\in[0,1], and that ZZ in (2.5) and (2.6) takes the form Z=ilog(UV)Z=-i\log(U^{*}V). To show the latter, note that eiZ=U1V=UVe^{iZ}=U^{-1}V=U^{*}V, where λ(UV)=\lambda(U^{*}V)\cap\mathbb{R}_{-}=\emptyset since both UU^{*} and VV are strictly accretive, see [43], [44, Thm. 6.2]. Therefore we can use the principle branch of the logarithm, which gives Z=ilog(UV)Z=-i\log(U^{*}V). Next, to show that γ𝕌n(t)𝔸𝕌n\gamma_{{\mathbb{U}}_{n}}(t)\in{\mathbb{A}}{\mathbb{U}}_{n} for t[0,1]t\in[0,1], note that γ𝕌n(1/2)=U1/2(U1/2VU1/2)1/2U1/2\gamma_{{\mathbb{U}}_{n}}(1/2)=U^{1/2}(U^{-1/2}VU^{-1/2})^{1/2}U^{1/2}. By [16, Prop. 3.1 and Thm. 3.4], γ𝕌n(1/2)\gamma_{{\mathbb{U}}_{n}}(1/2) is strictly accretive, and a repeated argument now gives that γ𝕌n(t)\gamma_{{\mathbb{U}}_{n}}(t) is strictly accretive for a dense set of t[0,1]t\in[0,1]. By continuity of the map tγ𝕌n(t)t\mapsto\gamma_{{\mathbb{U}}_{n}}(t), the result follows. ∎

Next, let Φ1\Phi_{1} and Φ2\Phi_{2} be two smooth symmetric gauge functions and consider the Finsler manifolds (n,FnΦ1)({\mathbb{P}}_{n},F_{{\mathbb{P}}_{n}}^{\Phi_{1}}) and (𝔸𝕌n,F𝕌nΦ2)({\mathbb{A}}{\mathbb{U}}_{n},F_{{\mathbb{U}}_{n}}^{\Phi_{2}}) as defined in Proposition 2.1 and Proposition 4.1, respectively. In the Riemannian case, there is a canonical way to introduce a metric on n×𝔸𝕌n{\mathbb{P}}_{n}\times{\mathbb{A}}{\mathbb{U}}_{n}, namely the product metric [30, Ex. 13.2]. However, in the case of products of Finsler manifolds there is no canonical way to introduce a Finsler structure on a product space, cf. [9, Ex. 11.1.6], [39]. Here, we consider so called (Minkowskian) product manifolds [39] and to this end we next define so called (Minkowskian) product functions.

Definition 4.2 ([39]).

A function Ψ:+×++\Psi:\mathbb{R}_{+}\times\mathbb{R}_{+}\to\mathbb{R}_{+} is called a product function if it satisfies the following conditions:

  1. i)

    Ψ(x1,x2)=0\Psi(x_{1},x_{2})=0 if and only if (x1,x2)=(0,0)(x_{1},x_{2})=(0,0),

  2. ii)

    Ψ(αx1,αx2)=αΨ(x1,x2)\Psi(\alpha x_{1},\alpha x_{2})=\alpha\Psi(x_{1},x_{2}) for all (x1,x2)+×+(x_{1},x_{2})\in\mathbb{R}_{+}\times\mathbb{R}_{+} and all α+\alpha\in\mathbb{R}_{+},

  3. iii)

    Ψ\Psi is smooth on +×+{(0,0)}\mathbb{R}_{+}\times\mathbb{R}_{+}\setminus\{(0,0)\},

  4. iv)

    xΨ0\partial_{x_{\ell}}\Psi\neq 0 on +×+{(0,0)}\mathbb{R}_{+}\times\mathbb{R}_{+}\setminus\{(0,0)\}, for =1,2\ell=1,2,

  5. v)

    x1Ψx2Ψ2Ψx1x2Ψ0\partial_{x_{1}}\Psi\,\partial_{x_{2}}\Psi-2\Psi\partial_{x_{1}}\partial_{x_{2}}\Psi\neq 0 on +×+{(0,0)}\mathbb{R}_{+}\times\mathbb{R}_{+}\setminus\{(0,0)\}.

For any product function Ψ\Psi, (n×𝔸𝕌n,Ψ((FnΦ1)2,(F𝔸𝕌nΦ2)2))({\mathbb{P}}_{n}\times{\mathbb{A}}{\mathbb{U}}_{n},\sqrt{\Psi((F_{{\mathbb{P}}_{n}}^{\Phi_{1}})^{2},(F_{{\mathbb{A}}{\mathbb{U}}_{n}}^{\Phi_{2}})^{2})}) is a Finsler manifold [39], and we therefore define the Finsler manifold (𝔸n,F𝔸nΦ1,Φ2,Ψ)({\mathbb{A}}_{n},F_{{\mathbb{A}}_{n}}^{\Phi_{1},\Phi_{2},\Psi}) as follows.666In [39], the convention is that the Finsler structure is squared compared to the one in [9]. We follow the convention of the latter.777Note that functions Ψ\Psi fulfilling i)-v) are not necessarily symmetric gauge functions. As an example, consider Ψ(x,y)=(x2+3xy+y2)2\Psi(x,y)=(x^{2}+3xy+y^{2})^{2} [39, Rem. 6]; this is not a symmetric gauge function since in general Ψ(x,y)Ψ(x,y)\Psi(x,y)\neq\Psi(x,-y). Conversely, symmetric gauge functions do not in general fulfill i)-v). As an example, consider Φ(x,y)=max{|x|,|y|}\Phi(x,y)=\max\{|x|,|y|\} [34, p. 138]; at any point (x,y)+×+(x,y)\in\mathbb{R}_{+}\times\mathbb{R}_{+} such that x>yx>y, yΦ=0\partial_{y}\Phi=0 and hence condition iv) is not fulfilled for this function.

Definition 4.3.

Let Φ1\Phi_{1} and Φ2\Phi_{2} be two smooth symmetric gauge functions, let (n,FnΦ1)({\mathbb{P}}_{n},F_{{\mathbb{P}}_{n}}^{\Phi_{1}}) and (𝔸𝕌n,F𝕌nΦ2)({\mathbb{A}}{\mathbb{U}}_{n},F_{{\mathbb{U}}_{n}}^{\Phi_{2}}) be the Finsler manifolds as defined in Proposition 2.1 and Proposition 4.1, respectively, and let Ψ\Psi be a product function. The Finsler manifold (𝔸n,F𝔸nΦ1,Φ2,Ψ)({\mathbb{A}}_{n},F_{{\mathbb{A}}_{n}}^{\Phi_{1},\Phi_{2},\Psi}) is defined via the diffeomorphis in Proposition 3.3 as

(𝔸n,F𝔸nΦ1,Φ2,Ψ):=(n×𝔸𝕌n,Ψ((FnΦ1)2,(F𝔸𝕌nΦ2)2)).({\mathbb{A}}_{n},F_{{\mathbb{A}}_{n}}^{\Phi_{1},\Phi_{2},\Psi}):=\left({\mathbb{P}}_{n}\times{\mathbb{A}}{\mathbb{U}}_{n},\sqrt{\Psi((F_{{\mathbb{P}}_{n}}^{\Phi_{1}})^{2},(F_{{\mathbb{A}}{\mathbb{U}}_{n}}^{\Phi_{2}})^{2}})\right).

One particular example of a product function is Ψ:(x1,x2)x1+x2\Psi:(x_{1},x_{2})\mapsto x_{1}+x_{2}, which in [39] this was called “the Euclidean product”, and in the Riemannian case this leads to the canonical product manifold. Moreover, the geodesics and geodesic distance can be characterized for general product functions Ψ\Psi. This leads to the following result.

Theorem 4.4.

Let A,B𝔸nA,B\in{\mathbb{A}}_{n}, and let A=PAUAPAA=P_{A}U_{A}P_{A} and B=PBUBPBB=P_{B}U_{B}P_{B} be the corresponding symmetric polar decompositions. On the Finsler manifold (𝔸n,F𝔸nΦ1,Φ2,Ψ)({\mathbb{A}}_{n},F_{{\mathbb{A}}_{n}}^{\Phi_{1},\Phi_{2},\Psi}), a geodesic from AA to BB is given by

γ𝔸n(t)=γn(t)1/2γ𝔸𝕌n(t)γn(t)1/2,\gamma_{{\mathbb{A}}_{n}}(t)=\gamma_{{\mathbb{P}}_{n}}(t)^{1/2}\cdot\gamma_{{\mathbb{A}}{\mathbb{U}}_{n}}(t)\cdot\gamma_{{\mathbb{P}}_{n}}(t)^{1/2}, (4.2a)
where
γn(t):=PA(PA1PBPBPA1)tPA=PA2etlog(PA2PB2),\displaystyle\gamma_{{\mathbb{P}}_{n}}(t):=P_{A}(P_{A}^{-1}P_{B}P_{B}P_{A}^{-1})^{t}P_{A}=P_{A}^{2}e^{t\log(P_{A}^{-2}P_{B}^{2})}, (4.2b)
γ𝔸𝕌n(t):=UA1/2(UA1/2UBUA1/2)tUA1/2=UAetlog(UAUB).\displaystyle\gamma_{{\mathbb{A}}{\mathbb{U}}_{n}}(t):=U_{A}^{1/2}(U_{A}^{-1/2}U_{B}U_{A}^{-1/2})^{t}U_{A}^{1/2}=U_{A}e^{t\log(U_{A}^{*}U_{B})}. (4.2c)

Moreover, the geodesic distance from AA to BB is given by

δ𝔸nΦ1,Φ2,Ψ(A,B)\displaystyle\delta_{{\mathbb{A}}_{n}}^{\Phi_{1},\Phi_{2},\Psi}(A,B) =Ψ(δnΦ1(PAPA,PBPB)2,δ𝔸𝕌nΦ2(UA,UB)2)\displaystyle=\sqrt{\Psi\Big{(}\delta_{{\mathbb{P}}_{n}}^{\Phi_{1}}(P_{A}P_{A},P_{B}P_{B})^{2},\delta_{{\mathbb{A}}{\mathbb{U}}_{n}}^{\Phi_{2}}(U_{A},U_{B})^{2}\Big{)}}
=Ψ(log(PA1PBPBPA1)Φ12,log(UAUB)Φ22)\displaystyle=\sqrt{\Psi\Big{(}\|\log(P_{A}^{-1}P_{B}P_{B}P_{A}^{-1})\|_{\Phi_{1}}^{2},\|\log(U_{A}^{*}U_{B})\|_{\Phi_{2}}^{2}\Big{)}} (4.3)
=Ψ(Φ1(λ(log(PA1PBPBPA1)))2,Φ2(φ(UA1UB))2),\displaystyle=\sqrt{\Psi\Big{(}\Phi_{1}(\lambda(\log(P_{A}^{-1}P_{B}P_{B}P_{A}^{-1})))^{2},\Phi_{2}(\varphi(U^{-1}_{A}U_{B}))^{2}\Big{)}},

where φ()\varphi(\cdot) denotes the angles of the eigenvalues.

Proof.

That (𝔸n,F𝔸nΦ1,Φ2,Ψ)({\mathbb{A}}_{n},F_{{\mathbb{A}}_{n}}^{\Phi_{1},\Phi_{2},\Psi}) is a Finsler manifold follows from the discussion leading up to the theorem; see [39]. Moreover, that (4.2) is a geodesic follows (by construction) by using [39, Thm. 3] together with Proposition 2.1 and Proposition 4.1; this also gives the first two equalities in (4.3). To prove the last equality, first observe that P1PBPBP1P^{-1}P_{B}P_{B}P^{-1} is the geometric mean of PA2P_{A}^{2} and PB2P_{B}^{2} and hence positive definite [11, Thm. 4.1.3], [16, Sec. 3]. Therefore, log(PA1PBPBPA1)\log(P_{A}^{-1}P_{B}P_{B}P_{A}^{-1}) is Hermitian and thus

σ(log(PA1PBPBPA1)=|λ(log(PA1PBPBPA1))|.\sigma(\log(P_{A}^{-1}P_{B}P_{B}P_{A}^{-1})=|\lambda(\log(P_{A}^{-1}P_{B}P_{B}P_{A}^{-1}))|.

Similarly, UA1UBU^{-1}_{A}U_{B} is unitary and thus log(UA1UB)\log(U_{A}^{-1}U_{B}) is skew-Hermitian. Therefore, λ(log(UA1UB))=iφ(UA1UB)\lambda(\log(U_{A}^{-1}U_{B}))=-i\varphi(U_{A}^{-1}U_{B}) and hence

σ(log(UA1UB))=|λ(log(UA1UB))|=|φ(UA1UB)|.\sigma(\log(U_{A}^{-1}U_{B}))=|\lambda(\log(U_{A}^{-1}U_{B}))|=|\varphi(U_{A}^{-1}U_{B})|.

Finally, for unitary invariant norms we have that Φ=Φ(σ())\|\cdot\|_{\Phi}=\Phi(\sigma(\cdot)), and since for any symmetric gauge function Φ(|x|)=Φ(x)\Phi(|x|)=\Phi(x), the last equality follows. ∎

Next, we derive some properties of the geodesic distance in Theorem 4.4.

Proposition 4.5.

For matrices A,B(𝔸n,F𝔸nΦ1,Φ2,Ψ)A,B\in({\mathbb{A}}_{n},F_{{\mathbb{A}}_{n}}^{\Phi_{1},\Phi_{2},\Psi}) we have that

  1. 1)

    δ𝔸nΦ1,Φ2,Ψ(A1,B1)=δ𝔸nΦ1,Φ2,Ψ(A,B)\displaystyle\delta_{{\mathbb{A}}_{n}}^{\Phi_{1},\Phi_{2},\Psi}(A^{-1},B^{-1})=\delta_{{\mathbb{A}}_{n}}^{\Phi_{1},\Phi_{2},\Psi}(A,B)

  2. 2)

    δ𝔸nΦ1,Φ2,Ψ(A,B)=δ𝔸nΦ1,Φ2,Ψ(A,B)\displaystyle\delta_{{\mathbb{A}}_{n}}^{\Phi_{1},\Phi_{2},\Psi}(A^{*},B^{*})=\delta_{{\mathbb{A}}_{n}}^{\Phi_{1},\Phi_{2},\Psi}(A,B)

  3. 3)

    δ𝔸nΦ1,Φ2,Ψ(A1,A)=2δ𝔸nΦ1,Φ2,Ψ(I,A)=2δ𝔸nΦ1,Φ2,Ψ(I,A1)\displaystyle\delta_{{\mathbb{A}}_{n}}^{\Phi_{1},\Phi_{2},\Psi}(A^{-1},A)=2\delta_{{\mathbb{A}}_{n}}^{\Phi_{1},\Phi_{2},\Psi}(I,A)=2\delta_{{\mathbb{A}}_{n}}^{\Phi_{1},\Phi_{2},\Psi}(I,A^{-1}), and the geodesic midpoint between A1A^{-1} and AA is γ𝔸n(1/2)=I\gamma_{{\mathbb{A}}_{n}}(1/2)=I

  4. 4)

    for any U𝕌nU\in{\mathbb{U}}_{n} we have that δ𝔸nΦ1,Φ2,Ψ(UAU,UBU)=δ𝔸nΦ1,Φ2,Ψ(A,B).\displaystyle\delta_{{\mathbb{A}}_{n}}^{\Phi_{1},\Phi_{2},\Psi}(U^{*}AU,U^{*}BU)=\delta_{{\mathbb{A}}_{n}}^{\Phi_{1},\Phi_{2},\Psi}(A,B).

Proof.

To prove the statements, first note that A1=PA1UA1PA1A^{-1}=P_{A}^{-1}U_{A}^{-1}P_{A}^{-1}, and that A=PAUA1PAA^{*}=P_{A}U_{A}^{-1}P_{A}. To prove 1), we observe that

δ𝔸nΦ1,Φ2,Ψ(A1,B1)=Ψ(Φ1(λ(log(PAPB1PB1PA)))2,Φ2(φ(UAUB1))2).\delta_{{\mathbb{A}}_{n}}^{\Phi_{1},\Phi_{2},\Psi}(A^{-1},B^{-1})=\sqrt{\Psi\Big{(}\Phi_{1}(\lambda(\log(P_{A}P_{B}^{-1}P_{B}^{-1}P_{A})))^{2},\Phi_{2}(\varphi(U_{A}U_{B}^{-1}))^{2}\Big{)}}.

For the positive definite part, we have that

λ(log(PAPB1PB1PA))=λ(log((PAPB1PB1PA)1))=λ(log(PA1PBPBPA1)),\lambda(\log(P_{A}P_{B}^{-1}P_{B}^{-1}P_{A}))=\lambda(-\log((P_{A}P_{B}^{-1}P_{B}^{-1}P_{A})^{-1}))=-\lambda(\log(P_{A}^{-1}P_{B}P_{B}P_{A}^{-1})),

and since the symmetric gauge function Φ1\Phi_{1} is invariant under sign changes the distance corresponding to the positive definite part is equal. Similarly, for the strictly accretive unitary part we have that |φ(UAUB1)|=|φ((UAUB1))|=|φ(UBUA1)|=|φ(UA1UB)||\varphi(U_{A}U_{B}^{-1})|=|\varphi((U_{A}U_{B}^{-1})^{*})|=|\varphi(U_{B}U_{A}^{-1})|=|\varphi(U_{A}^{-1}U_{B})|, where the first equality follows from that the absolute value of the angles of the eigenvalues of a unitary matrix are invariant under the operation of taking conjugate transpose, and the last equality follows since the angles of the eigenvalues are invariant under unitary congruence. Statement 2) follows by a similar argument. To prove statement 3),

δ𝔸nΦ1,Φ2,Ψ(A1,A)\displaystyle\delta_{{\mathbb{A}}_{n}}^{\Phi_{1},\Phi_{2},\Psi}(A^{-1},A) =Ψ(Φ1(λ(log(PA4)))2,Φ2(φ(UA2))2)\displaystyle=\sqrt{\Psi\Big{(}\Phi_{1}(\lambda(\log(P_{A}^{-4})))^{2},\Phi_{2}(\varphi(U_{A}^{-2}))^{2}\Big{)}}
=Ψ(Φ1(2λ(log(PA2)))2,Φ2(2φ(UA))2)\displaystyle=\sqrt{\Psi\Big{(}\Phi_{1}(-2\lambda(\log(P_{A}^{2})))^{2},\Phi_{2}(-2\varphi(U_{A}))^{2}\Big{)}}
=2Ψ(Φ1(λ(log(PA2)))2,Φ2(φ(UA))2)=2δ𝔸nΦ1,Φ2,Ψ(I,A)\displaystyle=2\sqrt{\Psi\Big{(}\Phi_{1}(\lambda(\log(P_{A}^{2})))^{2},\Phi_{2}(\varphi(U_{A}))^{2}\Big{)}}=2\delta_{{\mathbb{A}}_{n}}^{\Phi_{1},\Phi_{2},\Psi}(I,A)

where the second equity follows by an argument similar to previous ones, and the third equality follows from property ii) for symmetric gauge functions and property ii) for product functions. The second equality in 3) now follows from 1), and that γ𝔸n(1/2)=I\gamma_{{\mathbb{A}}_{n}}(1/2)=I follows by a direct calculation using (4.2). Finally, to prove 4), simply note that UAU=UPAUAPAU=UPAUUUAUUPAUU^{*}AU=U^{*}P_{A}U_{A}P_{A}U=U^{*}P_{A}UU^{*}U_{A}UU^{*}P_{A}U, i.e., the same unitary congruence transformation applied to PAP_{A} and UAU_{A} individually. A direct calculation, using the unitary invariance of eigenvalues, the matrix logarithm, and the norms, gives the result. ∎

Remark 4.6.

Note that (𝔸n,F𝔸nΦ1,Φ2,Ψ)({\mathbb{A}}_{n},F_{{\mathbb{A}}_{n}}^{\Phi_{1},\Phi_{2},\Psi}) is in general not a complete metric space. In particular, in the Riemannian case, i.e., with symmetric gauge functions Φ:xnk=1nxk2\Phi_{\ell}:x\in\mathbb{R}^{n}\mapsto\sqrt{\sum_{k=1}^{n}x_{k}^{2}}, =1,2\ell=1,2, and product function Ψ:(x1,x2)x1+x2\Psi:(x_{1},x_{2})\mapsto x_{1}+x_{2}, (𝔸n,F𝔸nΦ1,Φ2,Ψ)({\mathbb{A}}_{n},F_{{\mathbb{A}}_{n}}^{\Phi_{1},\Phi_{2},\Psi}) is not a complete metric space since it is not geodesically complete [31, Thm. 6.19]. The latter is due to the fact that (𝔸𝕌n,F𝔸𝕌nΦ2)({\mathbb{A}}{\mathbb{U}}_{n},F_{{\mathbb{A}}{\mathbb{U}}_{n}}^{\Phi_{2}}) is not geodesically complete; for (𝔸𝕌n,F𝔸𝕌nΦ2)({\mathbb{A}}{\mathbb{U}}_{n},F_{{\mathbb{A}}{\mathbb{U}}_{n}}^{\Phi_{2}}) geodesics are not defined for all t+t\in\mathbb{R}_{+} since they will reach the boundary. For an example of a Cauchy sequence that does not converge to an element in (𝔸n,F𝔸nΦ1,Φ2,Ψ)({\mathbb{A}}_{n},F_{{\mathbb{A}}_{n}}^{\Phi_{1},\Phi_{2},\Psi}), consider the sequence (A)=1(A_{\ell})_{\ell=1}^{\infty}, where A=ei(π/2π/(2))I𝔸nA_{\ell}=e^{i(\pi/2-\pi/(2\ell))}I\in{\mathbb{A}}_{n} for all \ell. In this case, the geodesic distance between AA_{\ell} and AkA_{k} is given by

δ𝔸n(A,Ak)=Z(,k)=nπ2|11k|,\delta_{{\mathbb{A}}_{n}}(A_{\ell},A_{k})=\|Z^{(\ell,k)}\|=\sqrt{n}\frac{\pi}{2}\left|\frac{1}{\ell}-\frac{1}{k}\right|,

since Z(,k):=ilog(AAk)=ilog(ei(π/(2)π/(2k))I)=π/2(1/1/k)IZ^{(\ell,k)}:=-i\log(A_{\ell}^{*}A_{k})=-i\log(e^{i(\pi/(2\ell)-\pi/(2k))}I)=\pi/2(1/\ell-1/k)I. Thus (A)=1(A_{\ell})_{\ell=1}^{\infty} is a Cauchy sequence, however limA=eiπ/2I𝔸n\lim_{\ell\to\infty}A_{\ell}=e^{i\pi/2}I\not\in{\mathbb{A}}_{n}. Since the Hopf-Rinow theorem [31, Thm. 6.19] also carries over to Finsler geometry [9, Thm. 6.6.1], similar statements are true also in the general case.

Remark 4.7.

Using Remark 3.4, the above results can easily be generalized to the same subsets 𝔸~n\tilde{{\mathbb{A}}}_{n} of sectorial matrices. In fact, a direct calculation shows that all the algebraic expressions in Theorem 4.4 still hold in this case. However, statements 1)-3) of Proposition 4.5 use the fact that if A𝔸nA\in{\mathbb{A}}_{n} then A1,A𝔸nA^{-1},A^{*}\in{\mathbb{A}}_{n}. This is in general not true for other sets 𝔸~n\tilde{{\mathbb{A}}}_{n}.

5 An application and some related results

By construction, on the Finsler manifold (𝔸n,F𝔸nΦ1,Φ2,Ψ)({\mathbb{A}}_{n},F_{{\mathbb{A}}_{n}}^{\Phi_{1},\Phi_{2},\Psi}) the question “given A𝔸nA\in{\mathbb{A}}_{n}, which matrix BnB\in{\mathbb{P}}_{n} is closest to AA” have the answer “B=P2B=P^{2}, where A=PUPA=PUP is the symmetric polar decomposition.” Similarly, the corresponding question “which matrix B𝔸𝕌nB\in{\mathbb{A}}{\mathbb{U}}_{n} is closest to AA” have the answer “B=UB=U”. In this section we consider an application of the distance on the Finsler manifold (𝔸n,F𝔸nΦ1,Φ2,Ψ)({\mathbb{A}}_{n},F_{{\mathbb{A}}_{n}}^{\Phi_{1},\Phi_{2},\Psi}) to another matrix approximation problem, namely finding the closest matrix of bounded log-rank, the definition of which is given in Section 5.1. Moreover, in Section 5.2 we consider the relation between the midpoint of geodesics on (𝔸n,F𝔸nΦ1,Φ2,Ψ)({\mathbb{A}}_{n},F_{{\mathbb{A}}_{n}}^{\Phi_{1},\Phi_{2},\Psi}) and the geometric mean of strictly accretive matrices as introduced in [16].

5.1 Closest matrix of bounded log-rank

For a positive definite matrix PP, we defined the log-rank as the rank of the matrix logarithm of PP. This is equivalent to the number of eigenvalues of PP that are different from 11. We denote this log-rankn(){\text{\rm log-rank}}_{{\mathbb{P}}_{n}}(\cdot). Analogously, for a unitary matrix UU the log-rank can be defined as the rank of the matrix logarithm of UU, which is equivalent to the number of eigenvalues of UU with phase different from 0. We denote this log-rank𝕌n(){\text{\rm log-rank}}_{{\mathbb{U}}_{n}}(\cdot). For strictly accretive matrices we define the log-rank as follows.

Definition 5.1.

For A𝔸nA\in{\mathbb{A}}_{n} with symmetric polar decomposition A=PUPA=PUP, we define the log-rank of AA as

log-rank𝔸n(A):=max{log-rankn(P2),log-rank𝕌n(U)}.{\text{\rm log-rank}}_{{\mathbb{A}}_{n}}(A):=\max\{{\text{\rm log-rank}}_{{\mathbb{P}}_{n}}(P^{2}),\;{\text{\rm log-rank}}_{{\mathbb{U}}_{n}}(U)\}.

We now consider the log-rank approximation problem: given A𝔸nA\in{\mathbb{A}}_{n} find Ar𝔸nA_{r}\in{\mathbb{A}}_{n}, the latter with log-rank bounded by rr, that is closest to AA in the geodesic distance δ𝔸nΦ1,Φ2,Ψ\delta_{{\mathbb{A}}_{n}}^{\Phi_{1},\Phi_{2},\Psi}. This can be formulated as the optimization problem

infAr𝔸n\displaystyle\inf_{A_{r}\in{\mathbb{A}}_{n}} δ𝔸nΦ1,Φ2,Ψ(Ar,A)\displaystyle\quad\delta_{{\mathbb{A}}_{n}}^{\Phi_{1},\Phi_{2},\Psi}(A_{r},A) (5.1a)
subject to log-rank𝔸n(Ar)r.\displaystyle\quad{\text{\rm log-rank}}_{{\mathbb{A}}_{n}}(A_{r})\leq r. (5.1b)

Let Ar=PrUrPrA_{r}=P_{r}U_{r}P_{r} be the symmetric polar decomposition. By properties i) - iv) in the Definition 4.2 of product functions Ψ\Psi, for each such function the distance is nondecreasing in each argument separately. Therefore, by the form of the geodesic distance (4.3) and the definition of log-rank on 𝔸n{\mathbb{A}}_{n}, it follows that (5.1) splits into two separate problems over n{\mathbb{P}}_{n} and 𝔸𝕌n{\mathbb{A}}{\mathbb{U}}_{n}, namely

infPrn\displaystyle\inf_{P_{r}\in{\mathbb{P}}_{n}} log(Pr1P2Pr1)Φ1\displaystyle\quad\|\log(P_{r}^{-1}P^{2}P_{r}^{-1})\|_{\Phi_{1}} (5.2a)
subject to log-rankn(Pr2)r,\displaystyle\quad{\text{\rm log-rank}}_{{\mathbb{P}}_{n}}(P_{r}^{2})\leq r, (5.2b)

and

infUr𝔸𝕌n\displaystyle\inf_{U_{r}\in{\mathbb{A}}{\mathbb{U}}_{n}} log(UrU)Φ2\displaystyle\quad\|\log(U_{r}^{*}U)\|_{\Phi_{2}} (5.3a)
subject to log-rank𝕌n(Ur)r.\displaystyle\quad{\text{\rm log-rank}}_{{\mathbb{U}}_{n}}(U_{r})\leq r. (5.3b)

In fact, this gives the following theorem.

Theorem 5.2.

Assume that P^r\hat{P}_{r} and U^r\hat{U}_{r} are optimal solutions to (5.2) and (5.3), respectively. Then an optimal solution to (5.1) is given by Ar^=P^rU^rP^r\hat{A_{r}}=\hat{P}_{r}\hat{U}_{r}\hat{P}_{r}. Conversely, if (5.1) has an optimal solution Ar^=P^rU^rP^r\hat{A_{r}}=\hat{P}_{r}\hat{U}_{r}\hat{P}_{r}, then P^r\hat{P}_{r} and U^r\hat{U}_{r} are optimal solutions to (5.2) and (5.3), respectively.

In [47, Thm. 3] it was shown that (5.3) always has an optimal solution, and that it is the same for all symmetric gauge functions Φ2\Phi_{2}. More precisely, the optimal solution U^r\hat{U}_{r} is obtained from UU by setting the nrn-r phases of UU with smallest absolute value equal to 0. That is, let U=VDVU=V^{*}DV be a diagonalization of UU where D=diag([eiϕ~k(U)]k=1n)D={\text{\rm diag}}([e^{i\tilde{\phi}_{k}(U)}]_{k=1}^{n}) and where [ϕ~k(U)]k=1n[\tilde{\phi}_{k}(U)]_{k=1}^{n} are the phases of UU ordered so that |ϕ~1(U)||ϕ~2(U)||ϕ~n(U)||\tilde{\phi}_{1}(U)|\geq|\tilde{\phi}_{2}(U)|\geq\ldots\geq|\tilde{\phi}_{n}(U)|. Then

U^r=Vdiag(eiϕ~1(U),eiϕ~r(U),1,,1)V\hat{U}_{r}=V^{*}{\text{\rm diag}}(e^{i\tilde{\phi}_{1}(U)},\ldots e^{i\tilde{\phi}_{r}(U)},1,\ldots,1)V

is the optimal solution to (5.3). In the same spirit, the optimal solution to (5.2) can be charaterized as follows.

Proposition 5.3.

Let PnP\in{\mathbb{P}}_{n}, and let P2=Vdiag(λ1,,λn)VP^{2}=V^{*}{\text{\rm diag}}(\lambda_{1},\ldots,\lambda_{n})V be a diagonlization of P2P^{2} where the eigenvalues are ordered so that |log(λ1)||log(λ2)||log(λn)||\log(\lambda_{1})|\geq|\log(\lambda_{2})|\geq\ldots\geq|\log(\lambda_{n})|. Then P^r2=Vdiag(λ1,,λr,1,,1)V\hat{P}_{r}^{2}=V^{*}{\text{\rm diag}}(\lambda_{1},\ldots,\lambda_{r},1,\ldots,1)V is a minimizer to (5.2) for all symmetric gauge functions Φ1\Phi_{1}.

Proof.

Clearly, log-rankn(P^r)=r{\text{\rm log-rank}}_{{\mathbb{P}}_{n}}(\hat{P}_{r})=r and hence P^r\hat{P}_{r} is feasible to (5.2). Next, log(Pr1P2Pr1)Φ1=Φ1(λ(log(Pr1P2Pr1)))\|\log(P_{r}^{-1}P^{2}P_{r}^{-1})\|_{\Phi_{1}}=\Phi_{1}(\lambda(\log(P_{r}^{-1}P^{2}P_{r}^{-1}))). Moreover, since Pr1P2Pr1nP_{r}^{-1}P^{2}P_{r}^{-1}\in{\mathbb{P}}_{n} and hence is unitary diagonalizable and has positive eigenvalues, we have that λ(log(Pr1P2Pr1))=log(λ(Pr1P2Pr1))\lambda(\log(P_{r}^{-1}P^{2}P_{r}^{-1}))=\log(\lambda(P_{r}^{-1}P^{2}P_{r}^{-1})). Now, to show that P^r\hat{P}_{r} is the minimizer to (5.2) for all symmetric gauge functions Φ1\Phi_{1}, it is equivalent to show that |log(λ(P^r1P2P^r1))|w|log(λ(Pr1P2Pr1))||\log(\lambda(\hat{P}_{r}^{-1}P^{2}\hat{P}_{r}^{-1}))|\prec_{w}|\log(\lambda(P_{r}^{-1}P^{2}P_{r}^{-1}))| for all PrP_{r} such that log-rankn(Pr2)r{\text{\rm log-rank}}_{{\mathbb{P}}_{n}}(P_{r}^{2})\leq r; see, e.g., [18, Thm. 4], [36, Thm. 1], [22, Sec. 3.5], [34, Prop. 4.B.6], [45, Thm. 10.35 ].

To this end, using [34, Thm. 9.H.1.f] (or [10, Cor III.4.6 ], [45, Thm. 10.30 ]) we have that log(λ(P2))log(λ(Pr2))log(λ(Pr1P2Pr1))\log(\lambda(P^{2}))^{\downarrow}-\log(\lambda(P_{r}^{2}))^{\downarrow}\prec\log(\lambda(P_{r}^{-1}P^{2}P_{r}^{-1})),888To see this, take V=Pr1P2Pr1nV=P_{r}^{-1}P^{2}P_{r}^{-1}\in{\mathbb{P}}_{n} and U=Pr2U=P_{r}^{2} in [34, Thm. 9.H.1.f], and use the fact that the eigenvalues of UV=PrP2Pr1UV=P_{r}P^{2}P_{r}^{-1} are invariant under the similarity transform Pr1PrP_{r}^{-1}\cdot P_{r}. and by [10, Ex 11.3.5] we therefore have that |log(λ(P2))log(λ(Pr2))|w|log(λ(Pr1P2Pr1))||\log(\lambda(P^{2}))^{\downarrow}-\log(\lambda(P_{r}^{2}))^{\downarrow}|\prec_{w}|\log(\lambda(P_{r}^{-1}P^{2}P_{r}^{-1}))|. Moreover, by a direct calculation it can be verified that log(λ(P2))log(λ(P^r2))=log(λ(P^r1P2P^r1))\log(\lambda(P^{2}))^{\downarrow}-\log(\lambda(\hat{P}_{r}^{2}))^{\downarrow}=\log(\lambda(\hat{P}_{r}^{-1}P^{2}\hat{P}_{r}^{-1}))^{\downarrow} holds. Therefore, if we can show that

|log(λ(P2))log(λ(P^r2))|w|log(λ(P2))log(λ(Pr2))|\displaystyle\text{$|\log(\lambda(P^{2}))^{\downarrow}-\log(\lambda(\hat{P}_{r}^{2}))^{\downarrow}|\prec_{w}|\log(\lambda(P^{2}))^{\downarrow}-\log(\lambda(P_{r}^{2}))^{\downarrow}|$} (5.4)
for all for all PrP_{r} such that log-rankn(Pr2)r{\text{\rm log-rank}}_{{\mathbb{P}}_{n}}(P_{r}^{2})\leq r ,

we would have that for all such PrP_{r},

|log(λ(P^r1P2P^r1))|\displaystyle|\log(\lambda(\hat{P}_{r}^{-1}P^{2}\hat{P}_{r}^{-1}))^{\downarrow}| =|log(λ(P2))log(λ(P^r2))|w|log(λ(P2))log(λ(Pr2))|\displaystyle=|\log(\lambda(P^{2}))^{\downarrow}-\log(\lambda(\hat{P}_{r}^{2}))^{\downarrow}|\prec_{w}|\log(\lambda(P^{2}))^{\downarrow}-\log(\lambda(P_{r}^{2}))^{\downarrow}|
w|log(λ(Pr1P2Pr1))|\displaystyle\prec_{w}|\log(\lambda(P_{r}^{-1}P^{2}P_{r}^{-1}))|

and by transitivity of preorders the result follows. To show (5.4), we formulate the following equivalent optimization problem: let a=log(λ(P2))a=\log(\lambda(P^{2}))^{\downarrow} and consider

minw\displaystyle\min_{\prec_{w}} |ax|\displaystyle\quad|a-x|
subject to xn,x1x2xn\displaystyle\quad x\in\mathbb{R}^{n},\quad x_{1}\geq x_{2}\geq\ldots\geq x_{n}
at most r elements of x are nonzero,\displaystyle\quad\text{at most }r\text{ elements of }x\text{ are nonzero},

where minw\min_{\prec_{w}} is minimizing with respect to the preordering w\prec_{w}. The solution to the latter is to take xi=aix_{i}=a_{i} for the rr elements of aa with largest absolute value. ∎

5.2 On the geometric mean for strictly accretive matrices

The geometric mean of strictly accretive matrices, denoted by A#BA\#B, was introduced in [16] as a generalization of the geometric mean for positive definite matrices [11, Chp. 4 and 6], [37]. In particular, in [16] it was shown that for A,B𝔸nA,B\in{\mathbb{A}}_{n} there is a unique solution G𝔸nG\in{\mathbb{A}}_{n} to the equation GA1G=BGA^{-1}G=B. This solution is given explicitly as

A#B:=G=A1/2(A1/2BA1/2)1/2A1/2,A\#B:=G=A^{1/2}(A^{-1/2}BA^{-1/2})^{1/2}A^{1/2},

which is also the same algebraic expression as for the geometric mean of positive definite matrices.

The geometric mean for positive definite matrices can also be interpreted as the midpoint on the geodesic connecting the matrices [11, Sec. 6.1.7]. With the Finsler geometry (𝔸n,F𝔸nΦ1,Φ2,Ψ)({\mathbb{A}}_{n},F_{{\mathbb{A}}_{n}}^{\Phi_{1},\Phi_{2},\Psi}), we can therefore get an alternative definition of the geometric mean between strictly accretive matrices as the geodesic midpoint. However, for A,B(𝔸n,F𝔸nΦ1,Φ2,Ψ)A,B\in({\mathbb{A}}_{n},F_{{\mathbb{A}}_{n}}^{\Phi_{1},\Phi_{2},\Psi}) we in general have that γ𝔸n(1/2)A#B\gamma_{{\mathbb{A}}_{n}}(1/2)\neq A\#B. This can be seen by the following simple example.

Example 5.4.

Let A=IA=I and let B=PUP(𝔸n,F𝔸nΦ1,Φ2,Ψ)B=PUP\in({\mathbb{A}}_{n},F_{{\mathbb{A}}_{n}}^{\Phi_{1},\Phi_{2},\Psi}). Then we have that A#B=B1/2=(PUP)1/2A\#B=B^{1/2}=(PUP)^{1/2} and γ𝔸n(1/2)=P1/2U1/2P1/2\gamma_{{\mathbb{A}}_{n}}(1/2)=P^{1/2}U^{1/2}P^{1/2}. Thus, in general A#Bγ𝔸n(1/2)A\#B\neq\gamma_{{\mathbb{A}}_{n}}(1/2). In fact, equality holds in this case if and only if PP and UU commute.

Instead, the midpoint of the geodesic between two matrices A,B(𝔸n,F𝔸nΦ1,Φ2,Ψ)A,B\in({\mathbb{A}}_{n},F_{{\mathbb{A}}_{n}}^{\Phi_{1},\Phi_{2},\Psi}) can be expressed using the geometric mean #\# as

γ𝔸n(1/2)=(PA2#PB2)1/2(UA#UB)(PA2#PB2)1/2,\gamma_{{\mathbb{A}}_{n}}(1/2)=(P_{A}^{2}\#P_{B}^{2})^{1/2}\,(U_{A}\#U_{B})\,(P_{A}^{2}\#P_{B}^{2})^{1/2}, (5.5)

which follows directly from (4.2). Using this representation, we can characterize when A#B=γ(1/2)A\#B=\gamma(1/2). In order to do so, we first need the following two auxiliary result.

Lemma 5.5.

Let A𝕎nA\in\mathbb{W}_{n}, and let A=VQA=VQ be is polar decomposition, where V𝕌nV\in{\mathbb{U}}_{n} and QnQ\in{\mathbb{P}}_{n}. AA is normal if and only if A=Q1/2VQ1/2A=Q^{1/2}VQ^{1/2} is the symmetric polar decomposition of AA.

Proof.

First, using [21, Lem. 9] we conclude that since AA is sectorial, VV is also sectorial. Now, AA is normal if and only if VV and QQ commute [45, Thm. 9.1], which is true if and only if VV and Q1/2Q^{1/2} commute. Hence AA is normal if and only if A=VQ=Q1/2VQ1/2A=VQ=Q^{1/2}VQ^{1/2}, and by the existence and uniqueness of the symmetric polar decomposition the result follows. ∎

Lemma 5.6.

Let A,B𝔸nA,B\in{\mathbb{A}}_{n} and G=A#BG=A\#B. For all X𝔾𝕃nX\in{\mathbb{G}\mathbb{L}_{n}}, the unique strictly accretive solution to

H(XAX)1H=XBXH(X^{*}AX)^{-1}H=X^{*}BX

is H=XGXH=X^{*}GX.

Proof.

That H=XGXH=X^{*}GX solves the equations is easily verified by simply plugging it in. Moreover, that HH is unique follows from the uniqueness of the geometric mean for strictly accretive matrices [16, Sec. 3] and the fact that for any X𝔾𝕃nX\in{\mathbb{G}\mathbb{L}_{n}} we have that XAX,XBX𝔸nX^{*}AX,X^{*}BX\in{\mathbb{A}}_{n}. ∎

Proposition 5.7.

For A,B(𝔸n,F𝔸nΦ1,Φ2,Ψ)A,B\in({\mathbb{A}}_{n},F_{{\mathbb{A}}_{n}}^{\Phi_{1},\Phi_{2},\Psi}), let A=PAUAPAA=P_{A}U_{A}P_{A} and B=PBUBPBB=P_{B}U_{B}P_{B} be the corresponding symmetric polar decompositions. We have that A#B=γ𝔸n(1/2)A\#B=\gamma_{{\mathbb{A}}_{n}}(1/2) if one of the following holds:

  1. i)

    UA=UB=IU_{A}=U_{B}=I,

  2. ii)

    PA=PBP_{A}=P_{B},

  3. iii)

    AA and BB are commuting normal matrices.

Proof.

Using (5.5), the first statement follows immediately. To prove the second statement, let A=PUAPA=PU_{A}P and B=PUBPB=PU_{B}P. From (5.5) it therefor follows that γ𝔸n(1/2)=P(UA#UB)P\gamma_{{\mathbb{A}}_{n}}(1/2)=P(U_{A}\#U_{B})P. Using Lemma 5.6 with G=UA#UBG=U_{A}\#U_{B} and X=PX=P, where therefore have that

A#B=(PUAP)#(PUBP)=P(UA#UB)P=γ𝔸n(1/2).A\#B=(PU_{A}P)\#(PU_{B}P)=P(U_{A}\#U_{B})P=\gamma_{{\mathbb{A}}_{n}}(1/2).

To prove the third statement, by Lemma 5.5 we have that PAP_{A}, UAU_{A}, and PBP_{B}, UBU_{B} commute. Moreover, since commuting normal matrices are simultaneously unitarilty diagonalizable [23, Thm. 2.5.5], and since a unitary diagonlization is unique up to permutation of the eigenvalues and eigenvectors, it follows that PA,UA,PBP_{A},U_{A},P_{B} and UBU_{B} all commute. Using this together with (5.5), a direct calculation gives the result. ∎

As noted in the above proof, if AA and BB are normal and commute they are also simultaneously unitarily diagonalizable [23, Thm. 2.5.5], i.e., A=VΛAVA=V^{*}\Lambda_{A}V and B=VΛBVB=V^{*}\Lambda_{B}V for some V𝕌nV\in{\mathbb{U}}_{n}. In this case, using Lemma 5.6 we have that A#B=V(ΛA#ΛB)VA\#B=V^{*}(\Lambda_{A}\#\Lambda_{B})V, and the geometric mean between AA and BB can thus be interpreted as the (independent) geometric mean between the corresponding pairs of eigenvalues. In fact, the latter observation can be generalized to all pairs of matrices that can be simultaneously diagonalized by congruence, albeit that the elements of the diagonal matrices are not necessarily eigenvalues in this case (cf. Proposition 2.1).

Proposition 5.8.

Let A,B𝔸nA,B\in{\mathbb{A}}_{n} and assume that A=TDATA=T^{*}D_{A}T and B=TDBTB=T^{*}D_{B}T, where T𝔾𝕃nT\in{\mathbb{G}\mathbb{L}_{n}} and where DAD_{A} and DBD_{B} are diagonal matrices. Then A#B=T(DA#DB)TA\#B=T^{*}(D_{A}\#D_{B})T.

Proof.

Let A,B𝔸nA,B\in{\mathbb{A}}_{n} and assume that A=TDATA=T^{*}D_{A}T and B=TDBTB=T^{*}D_{B}T, where T𝔾𝕃nT\in{\mathbb{G}\mathbb{L}_{n}} and where DAD_{A} and DBD_{B} are diagonal matrices. A direct application of Lemma 5.6, with G=DA#DBG=D_{A}\#D_{B} and X=TX=T, gives the result. ∎

As a final remark, note that if DAD_{A} and DBD_{B} in Proposition 5.8 are unitary, then A=TDATA=T^{*}D_{A}T and B=TDBTB=T^{*}D_{B}T are sectorial decompositions of AA and BB, respectively. Now, let T=VPT=VP be the polar decomposition of TT, with V𝕌nV\in{\mathbb{U}}_{n} and PnP\in{\mathbb{P}}_{n}. Hence we have that A=PVDAVP=PUAPA=PV^{*}D_{A}VP=PU_{A}P and B=PVDBVP=PUBPB=PV^{*}D_{B}VP=PU_{B}P, i.e., PP is the positive definite part and VDAVV^{*}D_{A}V and VDBVV^{*}D_{B}V are the strictly accretive unitary part in the symmetric polar decomposition of AA and BB, respectively. From Proposition 5.7.ii) we therefore have that A#B=γ𝔸n(1/2)A\#B=\gamma_{{\mathbb{A}}_{n}}(1/2) in this case.

6 Conclusions

In this work we show that the set of strictly accretive matrices is a smooth manifold that is diffeomorphic to a direct product of the smooth manifold of positive definite matrices and the smooth manifold of strictly accretive unitary matrices. Using this decomposition, we introduced a family of Finsler metrics and studied their geodesics and geodesic distances. Finally, we consider the matrix approximation problem of finding the closest strictly accretive matrix of bounded log-rank, and also discuss the relation between the geodesic midpoint and the previously introduced geometric mean between accretive matrices.

There are several interesting ways in which these results can be extended. For example, in the case of positive definite matrices the geometric framework offered by the Riemannian manifold construction gives yet another interpretation of the geometric mean. In fact, the geometric mean between two positive definite matrices AA and BB is also the (unique) solution to the variational problem minGnδn(A,G)2+δn(B,G)2\min_{G\in{\mathbb{P}}_{n}}\delta_{{\mathbb{P}}_{n}}(A,G)^{2}+\delta_{{\mathbb{P}}_{n}}(B,G)^{2} [11, Sec. 6.2.8], [37, Prop. 3.5], and this interpretation can be used to extend the geometric mean to a mean between several matrices [37]. In a similar way, a geometric mean between the strictly accretive matrices A1,,ANA_{1},\ldots,A_{N} can be defined as the solution to

minG𝔸ni=1Nδ𝔸nΦ1,Φ2,Ψ(Ai,G)2,\min_{G\in{\mathbb{A}}_{n}}\;\sum_{i=1}^{N}\delta_{{\mathbb{A}}_{n}}^{\Phi_{1},\Phi_{2},\Psi}(A_{i},G)^{2},

however such a generalization would need more investigation. For example, even in the case of the Riemannian metric on the manifold of positive definite matrices, analytically computing the geometric mean between several matrices is nontrivial [37, Prop. 3.4]. Nevertheless, there are efficient numerical algorithms for solving the latter problem, see, e.g., the survey [25] or the monograph [1] and references therein.

The idea of this work was to introduce a metric that separates the “magnitudes” and the “phases” of strictly accretive matrices. However, the similarities between the manifold of positive definite matrices and the manifold of unitary matrices raises a question about another potential geometry on 𝔸n{\mathbb{A}}_{n} that does not explicitly use the product structure. More precisely, note that since all strictly accretive matrices have a unique, strictly accretive square root, the inner product on the tangent space TU𝔸𝕌nT_{U}{\mathbb{A}}{\mathbb{U}}_{n}, given by (2.4), can be defined analogously to the one on n{\mathbb{P}}_{n}, given by (2.1), namely as

X,YU=tr((U1/2XU1/2)(U1/2YU1/2))=tr(XY),\langle X,Y\rangle_{U}={\text{\rm tr}}((U^{-1/2}XU^{-1/2})^{*}(U^{-1/2}YU^{-1/2}))={\text{\rm tr}}(X^{*}Y),

for U𝔸𝕌nU\in{\mathbb{A}}{\mathbb{U}}_{n} and X,YTU𝔸𝕌nX,Y\in T_{U}{\mathbb{A}}{\mathbb{U}}_{n}. Based on the similarities between the inner products, and the corresponding geodesics and geodesic distances, we ask the following question: for A,B𝔸nA,B\in{\mathbb{A}}_{n} and X,YTA𝔸nX,Y\in T_{A}{\mathbb{A}}_{n}, if we define the inner product on TA𝔸nT_{A}{\mathbb{A}}_{n} as

X,YA=tr((A1/2XA1/2)(A1/2YA1/2)),\langle X,Y\rangle_{A}={\text{\rm tr}}((A^{-1/2}XA^{-1/2})^{*}(A^{-1/2}YA^{-1/2})),

what is the form of the geodesics and the geodesic distance?

Acknowledgments

The authors would like to thank Wei Chen, Dan Wang, Xin Mao, Di Zhao, and Chao Chen for valuable discussions.

Appendix A Technical results from Section 3

The following is a number of lemmata use in the proofs of Theorem 3.1 and 3.2.

Lemma A.1.

𝔸n{\mathbb{A}}_{n} is an open sets in 𝔾𝕃n{\mathbb{G}\mathbb{L}_{n}}.

Proof.

To show that 𝔸n{\mathbb{A}}_{n} is open in 𝔾𝕃n{\mathbb{G}\mathbb{L}_{n}}, note that since n𝕊n{\mathbb{H}}_{n}\perp\mathbb{S}_{n}, cf. [34, Thm. 10.B.1 and 10.B.2], [45, Prob. 10.7.20], we have that A+B=H(A+B)+S(A+B)A+B=H(A+B)+S(A+B). Moreover, H(A+B)=H(A)+H(B)H(A+B)=H(A)+H(B), and since the set n{\mathbb{P}}_{n} is open in n{\mathbb{H}}_{n}, n𝕊n{\mathbb{P}}_{n}\oplus\mathbb{S}_{n} is open in 𝔾𝕃n{\mathbb{G}\mathbb{L}_{n}}. ∎

Lemma A.2.

𝔸n{\mathbb{A}}_{n}is connected.

Proof.

By [30, Prop. 1.11], 𝔸n{\mathbb{A}}_{n} is connected if and only if it is path-connected. To show the latter, it suffices to show that any A𝔸nA\in{\mathbb{A}}_{n} is path-connected to II. To this end, let A=TDTA=T^{*}DT be the sectorial decomposition of AA. A piece-wise smooth path connecting AA and II is given by

γ(t):={TD12tT,for t[0,1/2)(TT)22t,for t[1/2,1)I,for t=1,\gamma(t):=\begin{cases}T^{*}D^{1-2t}T,&\text{for }t\in[0,1/2)\\ (T^{*}T)^{2-2t},&\text{for }t\in[1/2,1)\\ I,&\text{for }t=1,\end{cases}

and hence 𝔸n{\mathbb{A}}_{n} is connected. ∎

Lemma A.3.

The tangent space at an A𝔸nA\in{\mathbb{A}}_{n} is given by TA𝔸n=𝕄nT_{A}{\mathbb{A}}_{n}={\mathbb{M}}_{n}.

Proof.

This follows since 𝔸n{\mathbb{A}}_{n} is an open subset of 𝕄n{\mathbb{M}}_{n}. ∎

Lemma A.4.

𝔸𝕌n{\mathbb{A}}{\mathbb{U}}_{n} is a connected smooth manifold and at a point U𝔸𝕌nU\in{\mathbb{A}}{\mathbb{U}}_{n} the tangent space is TU𝔸𝕌n=𝕊nT_{U}{\mathbb{A}}{\mathbb{U}}_{n}=\mathbb{S}_{n}.

Proof.

Since 𝔸n{\mathbb{A}}_{n} is open in 𝔾𝕃n{\mathbb{G}\mathbb{L}_{n}} (Lemma A.1), 𝔸𝕌n=𝕌n𝔸n{\mathbb{A}}{\mathbb{U}}_{n}={\mathbb{U}}_{n}\cap{\mathbb{A}}_{n} is open in 𝕌n{\mathbb{U}}_{n} in the relative topology with respect to 𝔾𝕃n{\mathbb{G}\mathbb{L}_{n}}. Thus it is a smooth manifold [30, Ex. 1.26]. Moreover, the proof of Lemma A.2 holds, mutatis mutandis, showing that it is connected. Finally, since it is open in 𝕌n{\mathbb{U}}_{n}, the tangent space at U𝔸𝕌nU\in{\mathbb{A}}{\mathbb{U}}_{n} is TU𝔸𝕌n=𝕊nT_{U}{\mathbb{A}}{\mathbb{U}}_{n}=\mathbb{S}_{n}, cf. [30, Prob. 8.29]. ∎

Lemma A.5 (Cf. [29, Prop. VII.2.5]).

For A𝔾𝕃nA\in{\mathbb{G}\mathbb{L}_{n}}, let A=VQA=VQ where V𝕌nV\in{\mathbb{U}}_{n} and QnQ\in{\mathbb{P}}_{n} be the polar decomposition of AA. The mapping A(V,Q)A\mapsto(V,Q) is a diffeomorphis between the manifolds 𝔾𝕃n{\mathbb{G}\mathbb{L}_{n}} and 𝕌n×n{\mathbb{U}}_{n}\times{\mathbb{P}}_{n}.

Proof.

First, since 𝕌n{\mathbb{U}}_{n} and n{\mathbb{P}}_{n} are smooth manifolds so is 𝕌n×n{\mathbb{U}}_{n}\times{\mathbb{P}}_{n} [30, Ex. 1.34]. Next, for each A𝔾𝕃nA\in{\mathbb{G}\mathbb{L}_{n}} the polar decomposition is unique [23, Thm. 7.3.1], [45, Prob. 3.2.20], and for each pair of matrices (V,Q)𝕌n×n(V,Q)\in{\mathbb{U}}_{n}\times{\mathbb{P}}_{n} we have that VQ𝔾𝕃nVQ\in{\mathbb{G}\mathbb{L}_{n}}; thus the mapping is bijective. Now, AA is smooth in VV and QQ since it is polynomial in the coefficients, i.e., the inverse mapping is smooth. To prove the converse, note that the components in the polar decomposition A=VQA=VQ are given by Q=(AA)1/2Q=(A^{*}A)^{1/2} and V=A(AA)1/2V=A(A^{*}A)^{-1/2}, cf. [23, p. 449], [45, p. 288]. Since A𝔾𝕃nA\in{\mathbb{G}\mathbb{L}_{n}}, AAnA^{*}A\in{\mathbb{P}}_{n}, and the matrix square root is a smooth function on n{\mathbb{P}}_{n}, cf. [23, Thm. 7.2.6]. Therefore, since both QQ and VV are compositions of smooth functions of AA, they both depend smoothly on the components of AA. ∎

References

  • [1] Pierre-Antoine Absil, Robert Mahony, and Rodolphe Sepulchre. Optimization algorithms on matrix manifolds. Princeton University Press, Princeton, NJ, 2008.
  • [2] Shun-ichi Amari and Hiroshi Nagaoka. Methods of information geometry. American Mathematical Society, Providence, RI, 2000.
  • [3] Esteban Andruchow. Short geodesics of unitaries in the L2L_{2} metric. Canadian Mathematical Bulletin, 48(3):340–354, 2005.
  • [4] Esteban Andruchow, Gabriel Larotonda, and Lázaro A. Recht. Finsler geometry and actions of the pp-Schatten unitary groups. Transactions of the American Mathematical Society, 362(1):319–344, 2010.
  • [5] Esteban Andruchow and Lázaro A. Recht. Geometry of unitaries in a finite algebra: variation formulas and convexity. International Journal of Mathematics, 19(10):1223–1246, 2008.
  • [6] Jorge Antezana, Eduardo Ghiglioni, and Demetrio Stojanoff. Minimal curves in 𝒰(n)\mathcal{U}(n) and 𝒢l(n)+\mathcal{G}l(n)^{+} with respect to the spectral and the trace norms. Journal of Mathematical Analysis and Applications, 483(2):123632, 2020.
  • [7] Jorge Antezana, Gabriel Larotonda, and Alejandro Varela. Optimal paths for symmetric actions in the unitary group. Communications in Mathematical Physics, 328(2):481–497, 2014.
  • [8] Charles S. Ballantine and Charles R. Johnson. Accretive matrix products. Linear and Multilinear Algebra, 3(3):169–185, 1975.
  • [9] David Bao, Shiing-Shen Chern, and Zhongmin Shen. An introduction to Riemann-Finsler geometry. Springer, New York, NY, 2000.
  • [10] Rajendra Bhatia. Matrix analysis. Springer, New York, NY, 1997.
  • [11] Rajendra Bhatia. Positive definite matrices. Princeton university press, Princton, NJ, 2007.
  • [12] Wei Chen, Dan Wang, Sei Zhen Khong, and Li Qiu. Phase analysis of MIMO LTI systems. In 2019 IEEE 58th Conference on Decision and Control (CDC), pages 6062–6067. IEEE, 2019.
  • [13] Xinyue Cheng and Zhongmin Shen. Finsler geometry. Springer, Berlin, Heidelberg, 2012.
  • [14] Shiing-Shen Chern. Finsler geometry is just Riemannian geometry without the quadratic restriction. Notices of the American Mathematical Society, 43(9):959–963, 1996.
  • [15] Charles R. DePrima and Charles R. Johnson. The range of A1AA^{-1}A^{*} in GL(n, C). Linear Algebra and its Applications, 9:209–222, 1974.
  • [16] Stephen Drury. Principal powers of matrices with positive definite real part. Linear and Multilinear Algebra, 63(2):296–301, 2015.
  • [17] Stephen Drury and Minghua Lin. Singular value inequalities for matrices with numerical ranges in a sector. Operators and Matrices, 8(4):1143–1148, 2014.
  • [18] Ky Fan. Maximum properties and inequalities for the eigenvalues of completely continuous operators. Proceedings of the National Academy of Sciences of the United States of America, 37(11):760, 1951.
  • [19] Susana Furtado and Charles R. Johnson. Spectral variation under congruence. Linear and Multilinear Algebra, 49(3):243–259, 2001.
  • [20] Karl E. Gustafson and Duggirala K.M. Rao. Numerical range. Springer, New York, NY, 1997.
  • [21] Alfred Horn and Robert Steinberg. Eigenvalues of the unitary part of a matrix. Pacific Journal of Mathematics, 9(2):541–550, 1959.
  • [22] Roger A. Horn and Charles R. Johnson. Topics in matrix analysis. Cambridge University Press, New York, NY, 1994.
  • [23] Roger A. Horn and Charles R. Johnson. Matrix analysis. Cambridge university press, New York, NY, 2013.
  • [24] Roger A. Horn and Vladimir V. Sergeichuk. Canonical forms for complex matrix congruence and congruence. Linear algebra and its applications, 416(2-3):1010–1032, 2006.
  • [25] Ben Jeuris, Raf Vandebril, and Bart Vandereycken. A survey and comparison of contemporary algorithms for computing the matrix geometric mean. Electronic Transactions on Numerical Analysis, 39:379–402, 2012.
  • [26] Charles R. Johnson and Susana Furtado. A generalization of Sylvester’s law of inertia. Linear Algebra and its Applications, 338(1-3):287–290, 2001.
  • [27] Jürgen Jost. Riemannian geometry and geometric analysis. Springer, Berlin, Heidelberg, 2008.
  • [28] Tosio Kato. Perturbation theory for linear operators. Springer, Berlin, Heidelberg, 1995.
  • [29] Serge Lang. Fundamentals of differential geometry. Springer, New York, NY, 1999.
  • [30] John M. Lee. Introduction to Smooth Manifolds. Springer, New York, NY, 2013.
  • [31] John M. Lee. Introduction to Riemannian Manifolds. Springer, Cham, 2018.
  • [32] Adrian S. Lewis. Group invariance and convex matrix analysis. SIAM Journal on Matrix Analysis and Applications, 17(4):927–949, 1996.
  • [33] Chi-Kwong Li and Nung-Sing Sze. Determinantal and eigenvalue inequalities for matrices with numerical ranges in a sector. Journal of Mathematical Analysis and Applications, 410(1):487–491, 2014.
  • [34] Albert W Marshall, Ingram Olkin, and Barry C Arnold. Inequalities: theory of majorization and its applications. Springer, New York, NY, 2nd edition, 2011.
  • [35] Roy Mathias. Matrices with positive definite Hermitian part: Inequalities and linear systems. SIAM journal on matrix analysis and applications, 13(2):640–654, 1992.
  • [36] Leon Mirsky. Symmetric gauge functions and unitarily invariant norms. The quarterly journal of mathematics, 11(1):50–59, 1960.
  • [37] Maher Moakher. A differential geometric approach to the geometric mean of symmetric positive-definite matrices. SIAM Journal on Matrix Analysis and Applications, 26(3):735–747, 2005.
  • [38] Richard M. Murray, Zexiang Li, and S. Shankar Sastry. A mathematical introduction to robotic manipulation. CRC press, Boca Raton, FL, 1994.
  • [39] Tsutomu Okada. Minkowskian product of Finsler spaces and Berwald connection. Journal of Mathematics of Kyoto University, 22(2):323–332, 1982.
  • [40] Barrett O’Neill. Semi-Riemannian geometry with applications to relativity. Academic press, San Diego, CA, 1983.
  • [41] Karl Johan Åström and Richard M. Murray. Feedback systems. Princeton university press, Princeton, NJ, 2008.
  • [42] Hideo Shimada and Vasile Soriin Sabău. Finsler geometry. In P.L. Antonelli, editor, Finslerian Geometries: A Meeting of Minds, pages 15–24. Springer, Dordrecht, 2000.
  • [43] Robert C. Thompson. On the eigenvalues of a product of unitary matrices I. Linear and Multilinear Algebra, 2(1):13–24, 1974.
  • [44] Dan Wang, Wei Chen, Sei Zhen Khong, and Li Qiu. On the phases of a complex matrix. Linear Algebra and its Applications, 593:152–179, 2020.
  • [45] Fuzhen Zhang. Matrix theory: basic results and techniques. Springer, New York, NY, 2011.
  • [46] Fuzhen Zhang. A matrix decomposition and its applications. Linear and Multilinear Algebra, 63(10):2033–2042, 2015.
  • [47] Di Zhao, Axel Ringh, Li Qiu, and Sei Zhen Khong. Low phase-rank approximation. Submitted, 2020.