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

The Brown measure of non-Hermitian sums of projections

Max Sun Zhou
Abstract.

We compute the Brown measure of the non-normal operators X=p+iqX=p+iq, where pp and qq are Hermitian, freely independent, and have spectra consisting of 22 atoms. The computation relies on the model of the non-trivial part of the von Neumann algebra generated by 2 projections as 2×22\times 2 random matrices. We observe that these measures are supported on hyperbolas and note some other properties related to their atoms and symmetries.

1. Introduction

Let MM be a von Neumann algebra with a faithful, normal, tracial state τ\tau. Let X(M,τ)X\in(M,\tau). The Brown measure of XX, introduced in [3], is a complex Borel probability measure supported on the spectrum of XX. When XX is normal, the Brown measure of XX is the spectral measure. When XX is a random matrix, the Brown measure of XX is the empirical spectral distribution of XX.

The first interesting explicit computations for the Brown measure of non-trivial operators were provided in [6], nearly 20 years after the Brown measure was first introduced. This class of operators are the “RR-diagonal operators” which include Voiculescu’s circular operator. In recent years, there has been much research in computing the Brown measure for a wide variety of families of operators (see [8], [9], and [2] for some examples).

In this paper, we will explicitly compute the Brown measure of operators of the form X=p+iqX=p+iq, where p,q(M,τ)p,q\in(M,\tau) are Hermitian, freely independent, and have spectra consisting of 22 atoms. Note that when pp or qq only has 11 atom in their spectra (i.e. is a real number), then XX is normal, so the Brown measure can be trivially computed. In the course of studying these operators, we will deduce that when pp and qq have 22 atoms, then XX is non-normal.

The key tool is the model of the von Neumann algebra generated by 22 projections from [14] to study the non-trivial part of this space. This reduces the computation in (M,τ)(M,\tau) to (M2(L((0,1),ν)),𝔼ν[12tr])\left(M_{2}\left(L^{\infty}((0,1),\nu^{*})\right),\mathbb{E}_{\nu^{*}}[\frac{1}{2}\text{{tr}}]\right) for some Borel probability measure ν\nu^{*} on (0,1)(0,1). To completely determine the Brown measure, we use freeness of pp and qq to explicitly compute ν\nu^{*} and some other relevant parameters.

We will show these measures are supported on hyperbolas in the complex plane and observe some other interesting properties about their atoms and symmetries.

One important aspect of the Brown measure is that it is the candidate for the limit of the empirical spectral distribution for random matrices: Given a random matrix model XnMn()X_{n}\in M_{n}(\mathbb{C}), suppose that XnX_{n} converges in *-moments to X(M,τ)X\in(M,\tau). The Brown measure of XX is typically the limit of the empirical spectral distributions of the XnX_{n} (see [11] for a precise statement of this). However, this does not always hold, as [[]Chapter 11, Exercise 15]SpeicherBook provides a simple counterexample. On the other hand, two notable results that demonstrate this principle are the Circular Law in [13] and the Single Ring Theorem in [5].

In a following paper, we will consider a natural random matrix model XnX_{n} corresponding to X=p+iqX=p+iq and prove that the empirical spectral distributions of the XnX_{n} converge to the Brown measure of XX.

The rest of the paper is organized as follows. In Section 2, we recall the definition of the Brown measure, the model of the algebra generated by 22 projections, and some free probability transforms. In Section 3, we fix some notation and give an outline of the steps of the computation. In Section 4, we compute the Brown measure of XX up to some parameters that depend on the joint law of pp and qq (Proposition 4.3). In Section 5, we use the freeness of pp and qq to compute these parameters (Propositions 5.3 and 5.6) and also deduce that XX is non-normal (Corollary 5.5). In Section 6, we combine these results to state the Brown measure of XX (Theorem 6.1). We also deduce some properties and provide accompanying figures to illustrate these properties.

2. Preliminaries

2.1. Brown measure

In this subsection, we recall the definition of the Brown measure of an operator X(M,τ)X\in(M,\tau) and describe how to compute it.

First, we recall the definition of the Fuglede-Kadison determinant, first introduced in [4]:

Definition 2.1.

Let x(M,τ)x\in(M,\tau). Let μ|x|\mu_{\left\lvert x\right\rvert} be the spectral measure of |x|=(xx)1/2\left\lvert x\right\rvert=(x^{*}x)^{1/2}. Then, the Fuglede-Kadison determinant of xx, Δ(x)\Delta(x), is given by:

(1) Δ(x)=exp[0logtdμ|x|(t)].\Delta(x)=\exp\left[\int_{0}^{\infty}\log t\,d\mu_{\left\lvert x\right\rvert}(t)\right]\,.

When x=Xn(Mn(),1ntr)x=X_{n}\in(M_{n}(\mathbb{C}),\frac{1}{n}\text{{tr}}), then Δ(Xn)=|det(Xn)|1/n\Delta(X_{n})=\left\lvert\det(X_{n})\right\rvert^{1/n}. Thus, the Fuglede-Kadison determinant is a generalization of the normalized, positive determinant of a complex matrix.

By applying the functional calculus to a decreasing sequence of continuous functions on [0,)[0,\infty) converging to logt\log t, we have:

(2) logΔ(x)=0logtdμ|x|(t)=120logtdμ|x|2(t).\log\Delta(x)=\int_{0}^{\infty}\log t\,d\mu_{\left\lvert x\right\rvert}(t)=\frac{1}{2}\int_{0}^{\infty}\log t\,d\mu_{\left\lvert x\right\rvert^{2}}(t)\,.

In practice, it is more convenient to compute the right-hand side to compute Δ(X)\Delta(X).

Then, the Brown measure is the following distribution:

Definition 2.2.

Let x(M,τ)x\in(M,\tau). The Brown measure of xx is defined as:

(3) μx=12π2logΔ(zx).\mu_{x}=\frac{1}{2\pi}\nabla^{2}\log\Delta(z-x)\,.

In [3], some properties of the Brown measure are proven (see also [10], [12], [7] for more exposition on some of the basic results). In particular, the Brown measure of xx is a probability measure supported on the spectrum of xx. When xx is normal, then the Brown measure is the spectral measure. When xx is a random matrix, then the Brown measure is the empirical spectral distribution.

2.2. The von Neumann algebra generated by 22 projections

In this subsection, we recall some results from [[]Section 12]VoiculescuPaper characterizing the tracial von Neumann algebra generated by 22 projections. Let (M,τ)(M,\tau) be the von Neumann algebra generated by two projections pp and qq.

Let (M,τ)(M,\tau) be the von Neumann algebra generated by two projections pp and qq.

Consider the following elements of MM:

(4) e00\displaystyle e_{00} =(1p)(1q)\displaystyle=(1-p)\wedge(1-q)
e01\displaystyle e_{01} =(1p)q\displaystyle=(1-p)\wedge q
e10\displaystyle e_{10} =p(1q)\displaystyle=p\wedge(1-q)
e11\displaystyle e_{11} =pq\displaystyle=p\wedge q
e\displaystyle e =1(e00+e01+e10+e11).\displaystyle=1-(e_{00}+e_{01}+e_{10}+e_{11})\,.

These elements are central, mutually orthogonal projections and

(5) e00+e01+e10+e11+e=1.e_{00}+e_{01}+e_{10}+e_{11}+e=1\,.

If we consider the matrix of mMm\in M with respect to the eij,ee_{ij},e (i.e. the 5×55\times 5 matrix with entries of the form fmffmf^{\prime} where f,f{ei,j,e}f,f^{\prime}\in\{e_{i,j},e\}), then this matrix is diagonal. Further, the 4 diagonal terms eijmeije_{ij}me_{ij} are constants. Hence, the only interesting part of MM is the subalgebra eMeeMe. In particular,

(6) p\displaystyle p =p(1q)+pq+epe=e10+e11+epe\displaystyle=p\wedge(1-q)+p\wedge q+epe=e_{10}+e_{11}+epe
q\displaystyle q =(1p)q+pq+eqe=e01+e11+eqe.\displaystyle=(1-p)\wedge q+p\wedge q+eqe=e_{01}+e_{11}+eqe\,.

When e0e\neq 0, consider (eMe,τ|eMe)(eMe,{\left.\kern-1.2pt\tau\vphantom{\big{|}}\right|_{eMe}}) as a von Neumann subalgebra with identity ee and τ|eMe(eme)=τ(eme)τ(e){\left.\kern-1.2pt\tau\vphantom{\big{|}}\right|_{eMe}}(eme)=\frac{\tau(eme)}{\tau(e)}. Then,

(7) (eMe,τ|eMe)(M2(L((0,1),ν)),𝔼ν[12tr]).(eMe,{\left.\kern-1.2pt\tau\vphantom{\big{|}}\right|_{eMe}})\cong\left(M_{2}\left(L^{\infty}((0,1),\nu^{*})\right),\mathbb{E}_{\nu^{*}}[\frac{1}{2}\text{{tr}}]\right)\,.

where ν\nu^{*} is a Borel measure on (0,1)(0,1).

For any mMm\in M, let m~=emeeMe\tilde{m}=eme\in eMe. The isomorphism has the following correspondence between m~eMe\tilde{m}\in eMe and matrix-valued functions of t(0,1)t\in(0,1):

(8) p~\displaystyle\tilde{p} (t(tt2)1/2(tt2)1/21t)\displaystyle\leftrightarrow\begin{pmatrix}t&(t-t^{2})^{1/2}\\ (t-t^{2})^{1/2}&1-t\end{pmatrix}
q~\displaystyle\tilde{q} (1000).\displaystyle\leftrightarrow\begin{pmatrix}1&0\\ 0&0\end{pmatrix}.

The measure ν\nu^{*} is recovered by noting that for the central element x=pqp+(1p)(1q)(1p)=1pq+pq+qpx=pqp+(1-p)(1-q)(1-p)=1-p-q+pq+qp,

(9) x~(t00t).\tilde{x}\leftrightarrow\begin{pmatrix}t&0\\ 0&t\end{pmatrix}.

Hence, ν\nu^{*} is the spectral measure of the element x~\tilde{x} in (eMe,τ|eMe)(eMe,{\left.\kern-1.2pt\tau\vphantom{\big{|}}\right|_{eMe}}).

Observe that with the change of variables t=cos2θt=\cos^{2}\theta, θ(0,π/2)\theta\in(0,\pi/2),

(10) p~=Rθq~Rθ1,\tilde{p}=R_{\theta}\tilde{q}R_{\theta}^{-1}\,,

where RθM2()R_{\theta}\in M_{2}(\mathbb{C}) is the rotation matrix with angle θ\theta.

In practice, we will work with θ\theta instead of tt. Let (cos2)1:[0,1][0,π/2](\cos^{2})^{-1}:[0,1]\to[0,\pi/2] be the inverse of cos2:[0,π/2][0,1]\cos^{2}:[0,\pi/2]\to[0,1] and consider the measure ν\nu on (0,π/2)(0,\pi/2) given by:

(11) ν=((cos2)1)(ν).\nu=\left((\cos^{2})^{-1}\right)_{*}(\nu^{*})\,.

Then, we will use the isomorphism:

(12) (eMe,τ|eMe)(M2(L((0,π/2),ν)),𝔼ν[12tr]).(eMe,{\left.\kern-1.2pt\tau\vphantom{\big{|}}\right|_{eMe}})\cong\left(M_{2}\left(L^{\infty}((0,\pi/2),\nu)\right),\mathbb{E}_{\nu}[\frac{1}{2}\text{{tr}}]\right)\,.

Let us highlight a few facts about this algebra. First, we explain why in the matrix algebras the domains are open instead of closed. This is because of the following fact:

Lemma 2.3.

The measure ν\nu^{*} does not have atoms at 0 or 11, i.e. ν({0,1})=0\nu^{*}(\{0,1\})=0. This is equivalent to ν\nu not having atoms at 0 or π/2\pi/2.

Proof.

Let p~=epe\tilde{p}=epe and q~=eqe\tilde{q}=eqe. Since ee commutes with pp and qq, p~=pe\tilde{p}=p\wedge e and q~=qe\tilde{q}=q\wedge e.

Consider τ((q~p~q~)n)τ(p~q~)=τ(e(pq))\tau((\tilde{q}\tilde{p}\tilde{q})^{n})\to\tau(\tilde{p}\wedge\tilde{q})=\tau(e\wedge(p\wedge q)) as nn\to\infty. Recall that ee and pqp\wedge q are mutually orthogonal, so τ(e(pq))=0\tau(e\wedge(p\wedge q))=0. Under the isomorphism, computation shows that

(13) τ((q~p~q~)n)=01tn𝑑ν(t).\tau((\tilde{q}\tilde{p}\tilde{q})^{n})=\int_{0}^{1}t^{n}\,d\nu^{*}(t)\,.

As nn\to\infty the integral on the right-hand side decreases to ν({1})\nu^{*}(\{1\}). Hence, ν({1})=0\nu^{*}(\{1\})=0.

A similar argument considering τ(((1q~)(p~)(1q~))n)τ((1q~)p~)\tau(((1-\tilde{q})(\tilde{p})(1-\tilde{q}))^{n})\to\tau((1-\tilde{q})\wedge\tilde{p}) using that ee and (1q)p(1-q)\wedge p are mutually orthogonal shows that ν({0})=0\nu^{*}(\{0\})=0.

Finally, we note that from the definition of pushforward measure, ν({0,1})=ν({0,π/2})\nu^{*}(\{0,1\})=\nu(\{0,\pi/2\}). ∎

Next, we note that on eMeeMe, epeepe and eqeeqe have trace 1/21/2. This follows from ee being mutually orthogonal to all of the eije_{ij}:

Lemma 2.4.

Let τ\tau be the trace on eMeeMe. Then, τ(epe)=τ(eqe)=1/2\tau(epe)=\tau(eqe)=1/2.

Proof.

If this is false, then we may choose one of epeepe or 1epe1-epe and one of eqeeqe or 1eqe1-eqe so that the sum of the traces is greater than 11. Without loss of generality assume that τ(epe)+τ(eqe)>1\tau(epe)+\tau(eqe)>1. Since ee commutes with pp and qq, epe=peepe=p\wedge e and eqe=qeeqe=q\wedge e. Then, from the parallelogram law, we have the following contradiction:

(14) 0\displaystyle 0 =τ(e(pq))=τ(epeeqe)=τ(epe)+τ(eqe)τ(epeeqe)\displaystyle=\tau(e\wedge(p\wedge q))=\tau(epe\wedge eqe)=\tau(epe)+\tau(eqe)-\tau(epe\vee eqe)
τ(epe)τ(eqe)1>0.\displaystyle\geq\tau(epe)-\tau(eqe)-1>0\,.

2.3. Free probability transforms

In this final preliminary subsection, we review some definitions and properties of free probability transforms. First, we recall that the definition of freeness in (M,τ)(M,\tau):

Definition 2.5.

Let (M,τ)(M,\tau) be a tracial von Neumann algebra, and let {Ai}iI\{A_{i}\}_{i\in I} be a family of unital subalgebras of MM. {Ai}iI\{A_{i}\}_{i\in I} are freely independent if for any ajAk(j)a_{j}\in A_{k(j)} with k(j)k(j+1)k(j)\neq k(j+1), j=1,,n=1j=1,\ldots,n=1 and τ(ai)=0\tau(a_{i})=0, then

(15) τ(a1,an)=0.\tau(a_{1}\ldots,a_{n})=0\,.

Let r,(mk)1krr,(m_{k})_{1\leq k\leq r} be positive integers. The sets {X1,p,,Xmp,p}1pr\{X_{1,p},\ldots,X_{m_{p},p}\}_{1\leq p\leq r} of non-commutative random variables are free if the algebras they generate are free.

If AiA_{i} generate (M,τ)(M,\tau) and are freely independent, then τ|Ai{\left.\kern-1.2pt\tau\vphantom{\big{|}}\right|_{A_{i}}} determines τ\tau.

Now, we recall the definition of some free probability transforms of real measures and state some of their basic properties. We refer to [1], [10], and [12] for proofs of these facts.

For a Hermitian x(M,τ)x\in(M,\tau), we will abuse notation by using the subscript xx to denote the integral transform with respect to the spectral measure μx\mu_{x}.

First, we define the Stieltjes transform of a real probability measure:

Definition 2.6.

Let μ\mu be a probability measure on \mathbb{R}. Then, the Stieltjes transform of μ\mu is the function Gμ:supp(μ)G_{\mu}:\mathbb{C}\setminus\text{supp}(\mu)\to\mathbb{C} given by:

(16) Gμ(z)=1zt𝑑μ(t).G_{\mu}(z)=\int_{\mathbb{R}}\frac{1}{z-t}\,d\mu(t)\,.

We list some of the well-known facts about GμG_{\mu} in the following Proposition:

Proposition 2.7.

Let GμG_{\mu} be the Stieltjes transform of μ\mu. Then,

  • GμG_{\mu} is analytic on supp(μ)\mathbb{C}\setminus\text{supp}(\mu).

  • Let ±()\mathbb{H}^{\pm}(\mathbb{C}) be the upper/lower half-planes of \mathbb{C}. Then, Gμ:±G_{\mu}:\mathbb{H}^{\pm}\to\mathbb{H}^{\mp}. In particular, Gμ(z)G_{\mu}(z)\in\mathbb{R} if and only if zsupp(μ)z\in\mathbb{R}\setminus\text{supp}(\mu).

  • Gμ(z)¯=Gμ(z¯)\overline{G_{\mu}(z)}=G_{\mu}(\overline{z}) for zsupp(μ)z\in\mathbb{C}\setminus\text{supp}(\mu).

  • |Gμ(z)|1|Im(z)|\left\lvert G_{\mu}(z)\right\rvert\leq\frac{1}{\left\lvert\text{{Im}}(z)\right\rvert}.

  • GμG_{\mu} is analytic on supp(μ)\mathbb{C}\setminus\text{supp}(\mu).

  • μ\mu is compactly supported, Gμ(z)G_{\mu}(z) has the following Laurent series expansion for |z|>supλsupp(μ)|λ|\left\lvert z\right\rvert>\sup_{\lambda\in\text{supp}(\mu)}\left\lvert\lambda\right\rvert:

    (17) Gμ(z)=n=0(tn𝑑μ(t))zn1.G_{\mu}(z)=\sum_{n=0}^{\infty}\left(\int_{\mathbb{R}}t^{n}\,d\mu(t)\right)z^{-n-1}\,.

Thus, for compact measures, the Stieltjes transform of a measure μ\mu contains the same information as the moments of μ\mu.

We highlight one special fact about the Stieltjes transform. For z=a+ibz=a+ib, note that

(18) Im1zt=b(ta)2+b2.\text{{Im}}\,\frac{1}{z-t}=-\frac{b}{(t-a)^{2}+b^{2}}\,.

Recall that the Poisson kernel on the upper half-plane is given by:

(19) Pb(a)=1πba2+b2,b>0.P_{b}(a)=\frac{1}{\pi}\frac{b}{a^{2}+b^{2}}\,,\quad b>0\,.

Combining these two facts shows that

(20) 1πImGμ(a+ib)=(μPb)(a).-\frac{1}{\pi}\text{{Im}}\,G_{\mu}(a+ib)=(\mu*P_{b})(a)\,.

Recall that PbP_{b} are approximations to the identity as b0+b\to 0^{+}, so then

(21) limb0+1πImGμ(+ib)=μ\lim\limits_{b\to 0^{+}}-\frac{1}{\pi}\text{{Im}}\,G_{\mu}(\cdot+ib)=\mu

in the vague topology on \mathbb{R}.

By exploiting the conjugate symmetry of GμG_{\mu}, we can also write this in terms of the discontinuity of GμG_{\mu} across \mathbb{R}:

(22) limb0+Gμ(+ib)Gμ(ib)2πi=μ\lim\limits_{b\to 0^{+}}-\frac{G_{\mu}(\cdot+ib)-G_{\mu}(\cdot-ib)}{2\pi i}=\mu

in the vague topology on \mathbb{R}.

Additionally, there is an explicit formula for intervals ([10], Theorem 6):

Proposition 2.8.

For a,ba,b\in\mathbb{R} and a<ba<b,

(23) limy0+ab1πImGμ(x+iy)dx=μ((a,b))+12μ({a})+12μ({b}).\lim\limits_{y\to 0^{+}}\int_{a}^{b}-\frac{1}{\pi}\text{{Im}}G_{\mu}(x+iy)\,dx=\mu((a,b))+\frac{1}{2}\mu(\{a\})+\frac{1}{2}\mu(\{b\})\,.

When μ\mu is a compactly supported real measure, then

(24) lim|z|zGμ(z)=1.\lim\limits_{\left\lvert z\right\rvert\to\infty}zG_{\mu}(z)=1\,.

Computation using FμF_{\mu} shows that RμR_{\mu} is analytic in a neighborhood of 0, as the 1/w1/w is exactly the pole of Gμ1G_{\mu}^{\langle-1\rangle} at w=0w=0.

For non-compactly supported measures μ\mu, then Gμ1G_{\mu}^{\langle-1\rangle} can still be defined on wedge-shaped domains containing 0 [[, see]Theorem 33]SpeicherBook, but we will not need this fact.

Similar to the Stieltjes transform, we can define ψμ\psi_{\mu}:

Definition 2.9.

Let μ\mu be a probability measure on [0,)[0,\infty). Let supp(μ)1={1/x:xsupp(μ)}\text{supp}(\mu)^{-1}=\{1/x:x\in\text{supp}(\mu)\}. Define ψμ:supp(μ)1\psi_{\mu}:\mathbb{C}\setminus\text{supp}(\mu)^{-1}\to\mathbb{C} by:

(25) ψμ(z)=tz1tz𝑑μ(t).\psi_{\mu}(z)=\int_{\mathbb{R}}\frac{tz}{1-tz}\,d\mu(t)\,.

We list some well-known properties of ψμ\psi_{\mu} in the following Proposition:

Proposition 2.10.

Let ψμ\psi_{\mu} be as above. Then,

  • ψμ(0)=0\psi_{\mu}(0)=0.

  • ψμ\psi_{\mu} is analytic on supp(μ)1\mathbb{C}\setminus\text{supp}(\mu)^{-1}.

  • ψμ(z)¯=ψμ(z¯)\overline{\psi_{\mu}(z)}=\psi_{\mu}(\overline{z}) for all zsupp(μ)1z\in\mathbb{C}\setminus\text{supp}(\mu)^{-1}.

  • Let ±()\mathbb{H}^{\pm}(\mathbb{C}) be the upper/lower half-planes of \mathbb{C}. If μ({0})<1\mu(\{0\})<1, then ψμ:±()±()\psi_{\mu}:\mathbb{H}^{\pm}(\mathbb{C})\to\mathbb{H}^{\pm}(\mathbb{C}). In particular, ψμ(z)\psi_{\mu}(z)\in\mathbb{R} if and only if zsupp(μ)1z\in\mathbb{R}\setminus\text{supp}(\mu)^{-1}.

  • If μ\mu is compactly supported, then Gμ(z)G_{\mu}(z) has the following Taylor series expansion for |z|<infλsupp(μ)1|λ|\left\lvert z\right\rvert<\inf_{\lambda\in\text{supp}(\mu)^{-1}}\left\lvert\lambda\right\rvert:

    (26) ψμ(z)=n=1(tn𝑑μ(t))zn.\psi_{\mu}(z)=\sum_{n=1}^{\infty}\left(\int_{\mathbb{R}}t^{n}\,d\mu(t)\right)z^{n}\,.

We highlight the following equation, valid for zsupp(μ)z\in\mathbb{C}\setminus\text{supp}(\mu):

(27) Gμ(z)=1z(ψμ(1z)+1).G_{\mu}(z)=\frac{1}{z}\left(\psi_{\mu}\left(\frac{1}{z}\right)+1\right)\,.

To consider the inverse of ψμ\psi_{\mu} at z=0z=0, note that

(28) ψ(0)=t𝑑μ(t).\psi^{\prime}(0)=\int_{\mathbb{R}}t\,d\mu(t)\,.

As long as μ({0})<1\mu(\{0\})<1, then ψμ\psi_{\mu} is invertible in a neighborhood of 0. In this situation, define the following functions:

Definition 2.11.

Let μ\mu be a compactly supported probability measure on [0,)[0,\infty) such that μ({0})<1\mu(\{0\})<1. Then, ψμ\psi_{\mu} is invertible in a neighborhood of 0 in \mathbb{C}. Define

(29) χμ(w)\displaystyle\chi_{\mu}(w) =ψμ1(w)\displaystyle=\psi_{\mu}^{\langle-1\rangle}(w)
Sμ(w)\displaystyle S_{\mu}(w) =χμ(w)w+1w.\displaystyle=\chi_{\mu}(w)\frac{w+1}{w}\,.

SμS_{\mu} is called the 𝐒\bm{S}-transform of μ\mu.

In particular, since χμ(0)=0\chi_{\mu}(0)=0 then SμS_{\mu} is well-defined in a neighborhood of 0.

Finally, we recall the addition (resp. multiplication) laws for free additive (resp. free multiplicative) convolution:

Theorem 2.12.

Let x,y(M,τ)x,y\in(M,\tau) be Hermitian and freely independent. Let xyx\boxplus y be the spectral measure of x+yx+y. Then, where the functions are defined,

(30) Rxy(z)=Rμx(z)+Rμy(z).R_{x\boxplus y}(z)=R_{\mu_{x}}(z)+R_{\mu_{y}}(z)\,.

If x,yx,y are positive, let xyx\boxtimes y be the spectral measure of x1/2yx1/2x^{1/2}yx^{1/2}. Then, where the functions are defined,

(31) Sxy(z)=Sμx(z)Sμy(z).S_{x\boxtimes y}(z)=S_{\mu_{x}}(z)S_{\mu_{y}}(z)\,.

3. Notation and Outline of Computation

Recall that we consider operators of the form X=p+iqX=p+iq, where pp and qq are Hermitian, freely independent, and have 22 atoms in their spectra.

We will use the following notation for the atoms of pp and qq and their weights:

(32) μp\displaystyle\mu_{p} =aδα+(1a)δα\displaystyle=a\delta_{\alpha}+(1-a)\delta_{\alpha^{\prime}}
μq\displaystyle\mu_{q} =bδβ+(1b)δβ,\displaystyle=b\delta_{\beta}+(1-b)\delta_{\beta^{\prime}}\,,

where a,b(0,1)a,b\in(0,1), α,α,β,β\alpha,\alpha^{\prime},\beta,\beta^{\prime}\in\mathbb{R}, αα\alpha\neq\alpha^{\prime}, and ββ\beta\neq\beta^{\prime}.

We fix pp^{\prime} (resp. qq^{\prime}) to be the following spectral projection of pp (resp. qq):

Definition 3.1.

Let p,q(M,τ)p,q\in(M,\tau) be Hermitian with laws:

(33) μp\displaystyle\mu_{p} =aδα+(1a)δα\displaystyle=a\delta_{\alpha}+(1-a)\delta_{\alpha^{\prime}}
μq\displaystyle\mu_{q} =bδβ+(1b)δβ.\displaystyle=b\delta_{\beta}+(1-b)\delta_{\beta^{\prime}}\,.

Let p,q(M,τ)p^{\prime},q^{\prime}\in(M,\tau) be the following projections:

(34) p\displaystyle p^{\prime} =χ{α}(p)\displaystyle=\chi_{\{\alpha^{\prime}\}}(p)
q\displaystyle q^{\prime} =χ{β}(q).\displaystyle=\chi_{\{\beta^{\prime}\}}(q)\,.

As a consequence,

(35) 1p=χ{α}(p)\displaystyle 1-p^{\prime}=\chi_{\{\alpha\}}(p)
1q=χ{β}(q).\displaystyle 1-q^{\prime}=\chi_{\{\beta\}}(q)\,.

Hence,

(36) p\displaystyle p =α(1p)+αp=(αα)p+α\displaystyle=\alpha(1-p^{\prime})+\alpha^{\prime}p^{\prime}=(\alpha^{\prime}-\alpha)p^{\prime}+\alpha
q\displaystyle q =β(1q)+βq=(ββ)q+β\displaystyle=\beta(1-q^{\prime})+\beta^{\prime}q^{\prime}=(\beta^{\prime}-\beta)q^{\prime}+\beta\,

or equivalently

(37) p\displaystyle p^{\prime} =pααα\displaystyle=\frac{p-\alpha}{\alpha^{\prime}-\alpha}
q\displaystyle q^{\prime} =qβββ.\displaystyle=\frac{q-\beta}{\beta^{\prime}-\beta}\,.

Recall the Brown measure of XX is defined as:

(38) μ=12π2logΔ(zX)=12π2120log(x)𝑑νz(x),\mu=\frac{1}{2\pi}\nabla^{2}\log\Delta(z-X)=\frac{1}{2\pi}\nabla^{2}\frac{1}{2}\int_{0}^{\infty}\log(x)\,d\nu_{z}(x)\,,

where νz\nu_{z} is the spectral measure of Hz(X)=(zX)(zX)H_{z}(X)=(z-X)^{*}(z-X).

To compute the Brown measure, we need to complete the following steps:

  1. (1)

    Compute νz\nu_{z}.

  2. (2)

    Compute logΔ(zX)=120log(x)𝑑νz(x)\log\Delta(z-X)=\frac{1}{2}\int_{0}^{\infty}\log(x)\,d\nu_{z}(x).

  3. (3)

    Compute μ=12π2logΔ(zX)\mu=\frac{1}{2\pi}\nabla^{2}\log\Delta(z-X).

In Section 4, we will compute the Brown measure of XX up to some parameters that come from the von Neumann algebra generated by 22 projections in Subsection 2.2. In Section 5, we use freeness of pp and qq to compute these parameters and also deduce that XX is non-normal. In Section 6, we combine these results to state the Brown measure of XX and deduce some properties. In Section LABEL:sec:further_work, we discuss some further work that has been completed on related families of operators.

4. Brown Measure up to Weights and Measures

In this section, we will describe the Brown measure of X=p+iqX=p+iq as a convex combination of 4 atoms and another probability measure, μ\mu^{\prime}. Applying the model of the von Neumann algebra generated by 22 projections with the projections pp^{\prime} and qq^{\prime} defined in the previous section, observe that:

(39) e00\displaystyle e_{00} =χ{α}(p)χ{β}(q)\displaystyle=\chi_{\{\alpha\}}(p)\wedge\chi_{\{\beta\}}(q)
e01\displaystyle e_{01} =χ{α}(p)χ{β}(q)\displaystyle=\chi_{\{\alpha\}}(p)\wedge\chi_{\{\beta^{\prime}\}}(q)
e10\displaystyle e_{10} =χ{α}(p)χ{β}(q)\displaystyle=\chi_{\{\alpha^{\prime}\}}(p)\wedge\chi_{\{\beta\}}(q)
e11\displaystyle e_{11} =χ{α}(p)χ{β}(q)\displaystyle=\chi_{\{\alpha^{\prime}\}}(p)\wedge\chi_{\{\beta^{\prime}\}}(q)
e\displaystyle e =1(e00+e01+e10+e11).\displaystyle=1-(e_{00}+e_{01}+e_{10}+e_{11})\,.

The measure μ\mu^{\prime} depends on the measure ν\nu in the isomorphism from the von Neumann algebra generated by 22 projections: (eMe,τ|eMe)(M2(L((0,π/2),ν)),𝔼ν[1ntr])(eMe,{\left.\kern-1.2pt\tau\vphantom{\big{|}}\right|_{eMe}})\cong\left(M_{2}\left(L^{\infty}((0,\pi/2),\nu)\right),\mathbb{E}_{\nu}[\frac{1}{n}\text{{tr}}]\right). The weights in the convex combination are the weights τ(eij),τ(e)\tau(e_{ij}),\tau(e) from Subsection 2.2. We will determine the Brown measure up to determining the weights τ(eij),τ(e)\tau(e_{ij}),\tau(e) and the measure ν\nu. These parameters depend on the joint law of pp and qq. In particular, the results in this section are valid for arbitrary projections p,q(M,τ)p,q\in(M,\tau). In Section 5, we will compute these parameters when pp and qq are freely independent.

To compute the Brown measure of XX, recall we need to first compute νz\nu_{z}, the spectral measure of Hz(X)=(zX)(zX)H_{z}(X)=(z-X)^{*}(z-X).

The result of the computation is the following:

Proposition 4.1.

If e=0e=0,

(40) νz=τ(e00)δ|z(α+iβ)|2+τ(e01)δ|z(α+iβ)|2+τ(e10)δ|z(α+iβ)|2+τ(e11)δ|z(α+iβ)|2.\nu_{z}=\tau(e_{00})\delta_{\left\lvert z-(\alpha+i\beta)\right\rvert^{2}}+\tau(e_{01})\delta_{\left\lvert z-(\alpha+i\beta^{\prime})\right\rvert^{2}}+\tau(e_{10})\delta_{\left\lvert z-(\alpha^{\prime}+i\beta)\right\rvert^{2}}+\tau(e_{11})\delta_{\left\lvert z-(\alpha^{\prime}+i\beta^{\prime})\right\rvert^{2}}\,.

If e0e\neq 0, there exists continuous functions σz,1,σz,2:(0,π/2)[0,)\sigma_{z,1},\sigma_{z,2}:(0,\pi/2)\to[0,\infty) such that σz,1(θ),σz,2(θ)\sigma_{z,1}(\theta),\sigma_{z,2}(\theta) are the singular values of the element of (M2(L((0,π/2),ν)),𝔼ν[1ntr])\left(M_{2}\left(L^{\infty}((0,\pi/2),\nu)\right),\mathbb{E}_{\nu}[\frac{1}{n}\text{{tr}}]\right) corresponding to e(z(p+iq))e(eMe,τ|eMe)e(z-(p+iq))e\in(eMe,{\left.\kern-1.2pt\tau\vphantom{\big{|}}\right|_{eMe}}).

Then,

(41) νz\displaystyle\nu_{z} =τ(e00)δ|z(α+iβ)|2+τ(e01)δ|z(α+iβ)|2+τ(e10)δ|z(α+iβ)|2+\displaystyle=\tau(e_{00})\delta_{\left\lvert z-(\alpha+i\beta)\right\rvert^{2}}+\tau(e_{01})\delta_{\left\lvert z-(\alpha+i\beta^{\prime})\right\rvert^{2}}+\tau(e_{10})\delta_{\left\lvert z-(\alpha^{\prime}+i\beta)\right\rvert^{2}}+
τ(e11)δ|z(α+iβ)|2+τ(e)(σz,12)(ν)+(σz,22)(ν)2.\displaystyle\qquad\qquad\tau(e_{11})\delta_{\left\lvert z-(\alpha^{\prime}+i\beta^{\prime})\right\rvert^{2}}+\tau(e)\frac{(\sigma_{z,1}^{2})_{*}(\nu)+(\sigma_{z,2}^{2})_{*}(\nu)}{2}\,.
Proof.

Let Hz=Hz(X)=(zX)(zX)H_{z}=H_{z}(X)=(z-X)^{*}(z-X). Since the eij,ee_{ij},e are central, mutually orthogonal projections that sum to 11, then

(42) (Hz)n\displaystyle(H_{z})^{n} =(e00+e01+e10+e11+e)(Hz)n(e00+e01+e10+e11+e)\displaystyle=(e_{00}+e_{01}+e_{10}+e_{11}+e)(H_{z})^{n}(e_{00}+e_{01}+e_{10}+e_{11}+e)
=e00(Hz)ne00+e01(Hz)ne01+e10(Hz)ne10+e11(Hz)ne11+e(Hz)ne\displaystyle=e_{00}(H_{z})^{n}e_{00}+e_{01}(H_{z})^{n}e_{01}+e_{10}(H_{z})^{n}e_{10}+e_{11}(H_{z})^{n}e_{11}+e(H_{z})^{n}e
=(e00Hze00)n+(e01Hze01)n+(e10Hze10)n+(e11Hze11)n+(eHze)n.\displaystyle=(e_{00}H_{z}e_{00})^{n}+(e_{01}H_{z}e_{01})^{n}+(e_{10}H_{z}e_{10})^{n}+(e_{11}H_{z}e_{11})^{n}+(eH_{z}e)^{n}\,.

Taking traces of both sides,

(43) τ((Hz)n)\displaystyle\tau((H_{z})^{n})
=τ((e00Hze00)n)+τ((e01Hze01)n)+\displaystyle=\tau((e_{00}H_{z}e_{00})^{n})+\tau((e_{01}H_{z}e_{01})^{n})+
τ((e10Hze10)n)+τ((e11Hze11)n)+τ((eHze)n).\displaystyle\qquad\qquad\tau((e_{10}H_{z}e_{10})^{n})+\tau((e_{11}H_{z}e_{11})^{n})+\tau((eH_{z}e)^{n})\,.

From the definitions of the eije_{ij},

(44) (e00Hze00)n\displaystyle(e_{00}H_{z}e_{00})^{n} =(|z(α+iβ)|2)ne00\displaystyle=(\left\lvert z-(\alpha+i\beta)\right\rvert^{2})^{n}e_{00}
(e01Hze01)n\displaystyle(e_{01}H_{z}e_{01})^{n} =(|z(α+iβ)|2)ne01\displaystyle=(\left\lvert z-(\alpha+i\beta^{\prime})\right\rvert^{2})^{n}e_{01}
(e10Hze10)n\displaystyle(e_{10}H_{z}e_{10})^{n} =(|z(α+iβ)|2)ne10\displaystyle=(\left\lvert z-(\alpha^{\prime}+i\beta)\right\rvert^{2})^{n}e_{10}
(e11Hze11)n\displaystyle(e_{11}H_{z}e_{11})^{n} =(|z(α+iβ)|2)ne11.\displaystyle=(\left\lvert z-(\alpha^{\prime}+i\beta^{\prime})\right\rvert^{2})^{n}e_{11}\,.

Combining the previous two equations,

(45) τ((Hz)n)\displaystyle\tau((H_{z})^{n}) =τ(e00)(|z(α+iβ)|2)n\displaystyle=\tau(e_{00})(\left\lvert z-(\alpha+i\beta)\right\rvert^{2})^{n}
+τ(e01)(|z(α+iβ)|2)n\displaystyle\quad+\tau(e_{01})(\left\lvert z-(\alpha+i\beta^{\prime})\right\rvert^{2})^{n}
+τ(e10)(|z(α+iβ)|2)n\displaystyle\quad+\tau(e_{10})(\left\lvert z-(\alpha^{\prime}+i\beta)\right\rvert^{2})^{n}
+τ(e11)(|z(α+iβ)|2)n\displaystyle\quad+\tau(e_{11})(\left\lvert z-(\alpha^{\prime}+i\beta^{\prime})\right\rvert^{2})^{n}
+τ(e)τ|eMe((eHze)n).\displaystyle\quad+\tau(e){\left.\kern-1.2pt\tau\vphantom{\big{|}}\right|_{eMe}}((eH_{z}e)^{n})\,.

The right-hand side is the nn-th moment of the following convex combination of probability measures:

(46) τ(e00)δ|z(α+iβ)|2+τ(e01)δ|z(α+iβ)|2+\displaystyle\tau(e_{00})\delta_{\left\lvert z-(\alpha+i\beta)\right\rvert^{2}}+\tau(e_{01})\delta_{\left\lvert z-(\alpha+i\beta^{\prime})\right\rvert^{2}}+
+τ(e10)δ|z(α+iβ)|2+τ(e11)δ|z(α+iβ)|2+τ(e)μeHze,\displaystyle\qquad+\tau(e_{10})\delta_{\left\lvert z-(\alpha^{\prime}+i\beta)\right\rvert^{2}}+\tau(e_{11})\delta_{\left\lvert z-(\alpha^{\prime}+i\beta^{\prime})\right\rvert^{2}}+\tau(e)\mu_{eH_{z}e}\,,

where μeHze\mu_{eH_{z}e} is the spectral measure of eHzeeH_{z}e in (eMe,τ|eMe)\left(eMe,{\left.\kern-1.2pt\tau\vphantom{\big{|}}\right|_{eMe}}\right).

As moments determine compact measures on \mathbb{R}, then this is νz\nu_{z}.

If e=0e=0, then νz\nu_{z} is the desired convex combination of atoms.

If e0e\neq 0, consider μeHze\mu_{eH_{z}e} using the isomorphism described in Section LABEL:sec:two_projections, i.e. we proceed by assuming that (eMe,τ|eMe)=(M2(L((0,π/2),ν)),𝔼ν[1ntr])(eMe,{\left.\kern-1.2pt\tau\vphantom{\big{|}}\right|_{eMe}})=\left(M_{2}\left(L^{\infty}((0,\pi/2),\nu)\right),\mathbb{E}_{\nu}[\frac{1}{n}\text{{tr}}]\right).

For any mMm\in M, let m~=emeeMe\tilde{m}=eme\in eMe. Then, Hz~=(z(p~+iq~))(z(p~+iq~))\tilde{H_{z}}=(z-(\tilde{p}+i\tilde{q}))^{*}(z-(\tilde{p}+i\tilde{q})).

Now, we claim that it is possible to choose continuous functions σz,1(θ),σz,2(θ)\sigma_{z,1}(\theta),\sigma_{z,2}(\theta) that are the singular values of (z(p~+iq~))=e(z(p+iq))e(z-(\tilde{p}+i\tilde{q}))=e(z-(p+iq))e for θ(0,π/2)\theta\in(0,\pi/2). First, note that from the expressions for p~,q~\tilde{p^{\prime}},\tilde{q^{\prime}} in Subsection 2.2 that Hz~\tilde{H_{z}} is continuous in θ\theta. Hence, the characteristic polynomial of Hz~\tilde{H_{z}} is a monic polynomial with coefficients that are continuous in θ\theta. From the quadratic formula, the eigenvalues of Hz~\tilde{H_{z}} have expressions in terms of the coefficients of the characteristic polynomial of Hz~\tilde{H_{z}}. As Hz~\tilde{H_{z}} are positive operators, then the eigenvalues are real, i.e. the discriminant of the characteristic polynomial is non-negative. Hence, the two branches of the square root of the discriminant are continuous in θ\theta. Thus, we may choose continuous expressions for the two eigenvalues of Hz~\tilde{H_{z}}. Taking square roots of these two non-negative continuous functions produces σz,1(θ),σz,2(θ)\sigma_{z,1}(\theta),\sigma_{z,2}(\theta).

Thus,

(47) τ|eMe((Hz~)n)\displaystyle{\left.\kern-1.2pt\tau\vphantom{\big{|}}\right|_{eMe}}((\tilde{H_{z}})^{n}) =0π/212tr((Hz~)n)𝑑ν(θ)\displaystyle=\int_{0}^{\pi/2}\frac{1}{2}\text{{tr}}((\tilde{H_{z}})^{n})\,d\nu(\theta)
=0π/2(σz,12(θ))n+(σz,22(θ))n2𝑑ν(θ).\displaystyle=\int_{0}^{\pi/2}\frac{(\sigma_{z,1}^{2}(\theta))^{n}+(\sigma_{z,2}^{2}(\theta))^{n}}{2}\,d\nu(\theta)\,.

The right-hand side is the nn-th moment of the following measure:

(48) (σz,12)(ν)+(σz,22)(ν)2.\frac{(\sigma_{z,1}^{2})_{*}(\nu)+(\sigma_{z,2}^{2})_{*}(\nu)}{2}\,.

Since moments determine compact measures on \mathbb{R}, then

(49) μeHze=(σz,12)(ν)+(σz,22)(ν)2.\mu_{eH_{z}e}=\frac{(\sigma_{z,1}^{2})_{*}(\nu)+(\sigma_{z,2}^{2})_{*}(\nu)}{2}\,.

By combining this with (46), we get the desired result:

(50) νz\displaystyle\nu_{z} =τ(e00)δ|z(α+iβ)|2+τ(e01)δ|z(α+iβ)|2+τ(e10)δ|z(α+iβ)|2+\displaystyle=\tau(e_{00})\delta_{\left\lvert z-(\alpha+i\beta)\right\rvert^{2}}+\tau(e_{01})\delta_{\left\lvert z-(\alpha+i\beta^{\prime})\right\rvert^{2}}+\tau(e_{10})\delta_{\left\lvert z-(\alpha^{\prime}+i\beta)\right\rvert^{2}}+
τ(e11)δ|z(α+iβ)|2+τ(e)(σz,12)(ν)+(σz,22)(ν)2.\displaystyle\qquad\qquad\tau(e_{11})\delta_{\left\lvert z-(\alpha^{\prime}+i\beta^{\prime})\right\rvert^{2}}+\tau(e)\frac{(\sigma_{z,1}^{2})_{*}(\nu)+(\sigma_{z,2}^{2})_{*}(\nu)}{2}\,.

As an intermediate step in computing the Brown measure of XX, we compute logΔ(zX)\log\Delta(z-X):

Proposition 4.2.

If e=0e=0,

(51) logΔ(zX)\displaystyle\log\Delta(z-X) =120log(x)𝑑νz(x)\displaystyle=\frac{1}{2}\int_{0}^{\infty}\log(x)\,d\nu_{z}(x)
=τ(e00)log|z(α+iβ)|\displaystyle=\tau(e_{00})\log\left\lvert z-(\alpha+i\beta)\right\rvert
+τ(e01)log|z(α+iβ)|\displaystyle\quad+\tau(e_{01})\log\left\lvert z-(\alpha+i\beta^{\prime})\right\rvert
+τ(e10)log|z(α+iβ)|\displaystyle\quad+\tau(e_{10})\log\left\lvert z-(\alpha^{\prime}+i\beta)\right\rvert
+τ(e11)log|z(α+iβ)|.\displaystyle\quad+\tau(e_{11})\log\left\lvert z-(\alpha^{\prime}+i\beta^{\prime})\right\rvert.

If e0e\neq 0, there exists continuous functions λ1,λ2:(0,π/2)\lambda_{1},\lambda_{2}:(0,\pi/2)\to\mathbb{C} such that λ1(θ),λ2(θ)\lambda_{1}(\theta),\lambda_{2}(\theta) are the eigenvalues of the element of (M2(L((0,π/2),ν)),𝔼ν[1ntr])\left(M_{2}\left(L^{\infty}((0,\pi/2),\nu)\right),\mathbb{E}_{\nu}[\frac{1}{n}\text{{tr}}]\right) corresponding to e(p+iq)e(eMe,τ|eMe)e(p+iq)e\in(eMe,{\left.\kern-1.2pt\tau\vphantom{\big{|}}\right|_{eMe}}).

Let 𝒜=αα\mathscr{A}=\alpha^{\prime}-\alpha and =ββ\mathscr{B}=\beta^{\prime}-\beta and z\sqrt{z} denote the principal branch of the square root defined on (,0)\mathbb{C}\setminus(-\infty,0). Then,

(52) λ1(θ)={α+α2+iβ+β212𝒜22+2i𝒜cos(2θ) when |𝒜|||α+α2+iβ+β2i22𝒜22i𝒜cos(2θ) when |𝒜|<||\lambda_{1}(\theta)=\begin{dcases}\frac{\alpha+\alpha^{\prime}}{2}+i\frac{\beta+\beta^{\prime}}{2}-\frac{1}{2}\sqrt{\mathscr{A}^{2}-\mathscr{B}^{2}+2i\mathscr{A}\mathscr{B}\cos(2\theta)}&\text{ when }\left\lvert\mathscr{A}\right\rvert\geq\left\lvert\mathscr{B}\right\rvert\\ \frac{\alpha+\alpha^{\prime}}{2}+i\frac{\beta+\beta^{\prime}}{2}-\frac{i}{2}\sqrt{\mathscr{B}^{2}-\mathscr{A}^{2}-2i\mathscr{A}\mathscr{B}\cos(2\theta)}&\text{ when }\left\lvert\mathscr{A}\right\rvert<\left\lvert\mathscr{B}\right\rvert\end{dcases}
(53) λ2(θ)={α+α2+iβ+β2+12𝒜22+2i𝒜cos(2θ) when |𝒜|||α+α2+iβ+β2+i22𝒜22i𝒜cos(2θ) when |𝒜|<||.\lambda_{2}(\theta)=\begin{dcases}\frac{\alpha+\alpha^{\prime}}{2}+i\frac{\beta+\beta^{\prime}}{2}+\frac{1}{2}\sqrt{\mathscr{A}^{2}-\mathscr{B}^{2}+2i\mathscr{A}\mathscr{B}\cos(2\theta)}&\text{ when }\left\lvert\mathscr{A}\right\rvert\geq\left\lvert\mathscr{B}\right\rvert\\ \frac{\alpha+\alpha^{\prime}}{2}+i\frac{\beta+\beta^{\prime}}{2}+\frac{i}{2}\sqrt{\mathscr{B}^{2}-\mathscr{A}^{2}-2i\mathscr{A}\mathscr{B}\cos(2\theta)}&\text{ when }\left\lvert\mathscr{A}\right\rvert<\left\lvert\mathscr{B}\right\rvert.\end{dcases}

Then,

(54) logΔ(zX)\displaystyle\log\Delta(z-X) =120log(x)𝑑νz(x)\displaystyle=\frac{1}{2}\int_{0}^{\infty}\log(x)\,d\nu_{z}(x)
=τ(e00)log|z(α+iβ)|\displaystyle=\tau(e_{00})\log\left\lvert z-(\alpha+i\beta)\right\rvert
+τ(e01)log|z(α+iβ)|\displaystyle\quad+\tau(e_{01})\log\left\lvert z-(\alpha+i\beta^{\prime})\right\rvert
+τ(e10)log|z(α+iβ)|\displaystyle\quad+\tau(e_{10})\log\left\lvert z-(\alpha^{\prime}+i\beta)\right\rvert
+τ(e11)log|z(α+iβ)|\displaystyle\quad+\tau(e_{11})\log\left\lvert z-(\alpha^{\prime}+i\beta^{\prime})\right\rvert
+τ(e)0π/2log|zλ1(θ)|+log|zλ2(θ)|2𝑑ν(θ).\displaystyle\quad+\tau(e)\int_{0}^{\pi/2}\frac{\log\left\lvert z-\lambda_{1}(\theta)\right\rvert+\log\left\lvert z-\lambda_{2}(\theta)\right\rvert}{2}\,d\nu(\theta)\,.
Proof.

When e=0e=0, the formula for logΔ(zX)\log\Delta(z-X) follows easily from Proposition 4.1.

For e0e\neq 0, from Proposition 4.1, for any continuous f:[0,)f:[0,\infty)\to\mathbb{C},

(55) 0f(x)𝑑νz(x)\displaystyle\int_{0}^{\infty}f(x)\,d\nu_{z}(x) =τ(e00)f(|z(α+iβ)|2)\displaystyle=\tau(e_{00})f(\left\lvert z-(\alpha+i\beta)\right\rvert^{2})
+τ(e01)f(|z(α+iβ)|2)\displaystyle\quad+\tau(e_{01})f(\left\lvert z-(\alpha+i\beta^{\prime})\right\rvert^{2})
+τ(e10)f(|z(α+iβ)|2)\displaystyle\quad+\tau(e_{10})f(\left\lvert z-(\alpha^{\prime}+i\beta)\right\rvert^{2})
+τ(e11)f(|z(α+iβ)|2)\displaystyle\quad+\tau(e_{11})f(\left\lvert z-(\alpha^{\prime}+i\beta^{\prime})\right\rvert^{2})
+τ(e)0f(x)d((σz,12)(ν)+(σz,22)(ν)2).\displaystyle\quad+\tau(e)\int_{0}^{\infty}f(x)\,d\left(\frac{(\sigma_{z,1}^{2})_{*}(\nu)+(\sigma_{z,2}^{2})_{*}(\nu)}{2}\right)\,.

Rewriting the final integral using the change of variables formula,

(56) 0f(x)d((σz,12)(ν)+(σz,22)(ν)2)\displaystyle\int_{0}^{\infty}f(x)\,d\left(\frac{(\sigma_{z,1}^{2})_{*}(\nu)+(\sigma_{z,2}^{2})_{*}(\nu)}{2}\right)
=12(0f(x)d(σz,12)(ν)+0f(x)(σz,22)(ν))\displaystyle=\frac{1}{2}\left(\int_{0}^{\infty}f(x)\,d(\sigma_{z,1}^{2})_{*}(\nu)+\int_{0}^{\infty}f(x)(\sigma_{z,2}^{2})_{*}(\nu)\right)
=12(0π/2f(σz,12(θ))𝑑ν(θ)+0π/2f(σz,22(θ))𝑑ν(θ))\displaystyle=\frac{1}{2}\left(\int_{0}^{\pi/2}f(\sigma_{z,1}^{2}(\theta))\,d\nu(\theta)+\int_{0}^{\pi/2}f(\sigma_{z,2}^{2}(\theta))\,d\nu(\theta)\right)
=120π/2f(σz,12(θ))+f(σz,22(θ))dν(θ).\displaystyle=\frac{1}{2}\int_{0}^{\pi/2}f(\sigma_{z,1}^{2}(\theta))+f(\sigma_{z,2}^{2}(\theta))\,d\nu(\theta)\,.

Let fn(x)=log(x+1/n)/2f_{n}(x)=\log(x+1/n)/2 for n=1,2,n=1,2,\ldots. Then, fn:[0,)f_{n}:[0,\infty)\to\mathbb{R} are continuous and decrease to log(x)/2\log(x)/2. Applying two previous equations for f=fnf=f_{n} and using the monotone convergence theorem to take the limit as nn\to\infty,

(57) logΔ(zX)\displaystyle\log\Delta(z-X) =120log(x)𝑑νz(x)\displaystyle=\frac{1}{2}\int_{0}^{\infty}\log(x)\,d\nu_{z}(x)
=τ(e00)log|z(α+iβ)|\displaystyle=\tau(e_{00})\log\left\lvert z-(\alpha+i\beta)\right\rvert
+τ(e01)log|z(α+iβ)|\displaystyle\quad+\tau(e_{01})\log\left\lvert z-(\alpha+i\beta^{\prime})\right\rvert
+τ(e10)log|z(α+iβ)|\displaystyle\quad+\tau(e_{10})\log\left\lvert z-(\alpha^{\prime}+i\beta)\right\rvert
+τ(e11)log|z(α+iβ)|\displaystyle\quad+\tau(e_{11})\log\left\lvert z-(\alpha^{\prime}+i\beta^{\prime})\right\rvert
+τ(e)20π/2log(σz,12(θ))+log(σz,22(θ))2𝑑ν(θ).\displaystyle\quad+\frac{\tau(e)}{2}\int_{0}^{\pi/2}\frac{\log(\sigma_{z,1}^{2}(\theta))+\log(\sigma_{z,2}^{2}(\theta))}{2}\,d\nu(\theta)\,.

For the rest of the proof, assume that (eMe,τ|eMe)=(M2(L((0,π/2),ν)),𝔼ν[1ntr])(eMe,{\left.\kern-1.2pt\tau\vphantom{\big{|}}\right|_{eMe}})=\left(M_{2}\left(L^{\infty}((0,\pi/2),\nu)\right),\mathbb{E}_{\nu}[\frac{1}{n}\text{{tr}}]\right).

Recall that σz,1(θ),σz,2(θ)\sigma_{z,1}(\theta),\sigma_{z,2}(\theta) are the singular values of e(zX)ee(z-X)e. Then, the integrand of the final integral can be simplified as:

(58) log(σz,12(θ))+log(σz,22(θ))2\displaystyle\frac{\log(\sigma_{z,1}^{2}(\theta))+\log(\sigma_{z,2}^{2}(\theta))}{2} =log(σz,12(θ)σz,22(θ))2\displaystyle=\frac{\log(\sigma_{z,1}^{2}(\theta)\sigma_{z,2}^{2}(\theta))}{2}
=log(det((e(zX)e)e(zX)e))2\displaystyle=\frac{\log(\det((e(z-X)e)^{*}e(z-X)e))}{2}
=log(|det((e(zX)e))|2)2\displaystyle=\frac{\log(\left\lvert\det((e(z-X)e))\right\rvert^{2})}{2}
=log|det((e(zX)e))|\displaystyle=\log\left\lvert\det((e(z-X)e))\right\rvert
=log|det(zeXe)|.\displaystyle=\log\left\lvert\det(z-eXe)\right\rvert.

Given that we can verify the formulas for λ1(θ)\lambda_{1}(\theta), λ2(θ)\lambda_{2}(\theta), then zλ1(θ),zλ2(θ)z-\lambda_{1}(\theta),z-\lambda_{2}(\theta) are the eigenvalues for zeXez-eXe. Thus, combining the previous two equations,

(59) 0π/2log(σz,12(θ))+log(σz,22(θ))2𝑑ν(θ)\displaystyle\int_{0}^{\pi/2}\frac{\log(\sigma_{z,1}^{2}(\theta))+\log(\sigma_{z,2}^{2}(\theta))}{2}\,d\nu(\theta)
=0π/2log|det(zeXe)|𝑑ν(θ)\displaystyle=\int_{0}^{\pi/2}\log\left\lvert\det(z-eXe)\right\rvert\,d\nu(\theta)
=0π/2log|zλ1(θ)|+log|zλ2(θ)|dν(θ).\displaystyle=\int_{0}^{\pi/2}\log\left\lvert z-\lambda_{1}(\theta)\right\rvert+\log\left\lvert z-\lambda_{2}(\theta)\right\rvert\,d\nu(\theta)\,.

Substituting this expression into (57) produces the desired formula for logΔ(zX)\log\Delta(z-X).

We return to verifying the formulas for λ1(θ)\lambda_{1}(\theta), λ2(θ)\lambda_{2}(\theta). Straightforward computation shows that the characteristic polynomial of eXe(M2(L((0,π/2),ν)),𝔼ν[1ntr])eXe\in\left(M_{2}\left(L^{\infty}((0,\pi/2),\nu)\right),\mathbb{E}_{\nu}[\frac{1}{n}\text{{tr}}]\right) is:

(60) p(λ)=λ2((α+α)+i(β+β))λ+(ααββ+i2((α+α)(β+β)𝒜cos(2θ))).p(\lambda)=\lambda^{2}-((\alpha+\alpha^{\prime})+i(\beta+\beta^{\prime}))\lambda+\left(\alpha\alpha^{\prime}-\beta\beta^{\prime}+\frac{i}{2}((\alpha+\alpha^{\prime})(\beta+\beta^{\prime})-\mathscr{A}\mathscr{B}\cos(2\theta))\right).

The eigenvalues of eXeeXe are:

(61) α+α2+iβ+β2±12𝒜22+2i𝒜cos(2θ).\frac{\alpha+\alpha^{\prime}}{2}+i\frac{\beta+\beta^{\prime}}{2}\pm\frac{1}{2}\sqrt{\mathscr{A}^{2}-\mathscr{B}^{2}+2i\mathscr{A}\mathscr{B}\cos(2\theta)}\,.

where the square root is any branch of the square root.

When |𝒜|||\left\lvert\mathscr{A}\right\rvert\geq\left\lvert\mathscr{B}\right\rvert, then Re(𝒜22+2i𝒜cos(2θ))0\text{{Re}}(\mathscr{A}^{2}-\mathscr{B}^{2}+2i\mathscr{A}\mathscr{B}\cos(2\theta))\geq 0 and the expression for the eigenvalues of e(p+iq)ee(p+iq)e is continuous and well-defined if we take the square root to be the principal branch of the square root. This gives the formulas for λ1(θ),λ2(θ)\lambda_{1}(\theta),\lambda_{2}(\theta) for |𝒜|||\left\lvert\mathscr{A}\right\rvert\geq\left\lvert\mathscr{B}\right\rvert.

When |𝒜|<||\left\lvert\mathscr{A}\right\rvert<\left\lvert\mathscr{B}\right\rvert, then ±i2𝒜22i𝒜cos(2θ)\pm i\sqrt{\mathscr{B}^{2}-\mathscr{A}^{2}-2i\mathscr{A}\mathscr{B}\cos(2\theta)} are also expressions for the square roots of 𝒜22+2i𝒜cos(2θ)\mathscr{A}^{2}-\mathscr{B}^{2}+2i\mathscr{A}\mathscr{B}\cos(2\theta), where the square root is the principal branch of the square root. Since |𝒜|<||\left\lvert\mathscr{A}\right\rvert<\left\lvert\mathscr{B}\right\rvert, then Re(2𝒜22i𝒜cos(2θ))>0\text{{Re}}(\mathscr{B}^{2}-\mathscr{A}^{2}-2i\mathscr{A}\mathscr{B}\cos(2\theta))>0 and this expression is continuous and well-defined.. This gives the formulas for λ1(θ),λ2(θ)\lambda_{1}(\theta),\lambda_{2}(\theta) for |𝒜|<||\left\lvert\mathscr{A}\right\rvert<\left\lvert\mathscr{B}\right\rvert. ∎

Finally, we compute the Brown measure of XX, μ=12π2logΔ(zX)\mu=\frac{1}{2\pi}\nabla^{2}\log\Delta(z-X) in the following Proposition:

Proposition 4.3.

Let μ\mu be the Brown measure of XX. If e=0e=0,

(62) μ\displaystyle\mu =12π2logΔ(zX)\displaystyle=\frac{1}{2\pi}\nabla^{2}\log\Delta(z-X)
=τ(e00)δα+iβ+τ(e01)δα+iβ+τ(e10)δα+iβ+τ(e11)δα+iβ.\displaystyle=\tau(e_{00})\delta_{\alpha+i\beta}+\tau(e_{01})\delta_{\alpha+i\beta^{\prime}}+\tau(e_{10})\delta_{\alpha^{\prime}+i\beta}+\tau(e_{11})\delta_{\alpha^{\prime}+i\beta^{\prime}}.

If e0e\neq 0, let λ1,λ2:(0,π/2)\lambda_{1},\lambda_{2}:(0,\pi/2)\to\mathbb{C} be as in Proposition 4.2. Then,

(63) μ\displaystyle\mu =12π2logΔ(zX)\displaystyle=\frac{1}{2\pi}\nabla^{2}\log\Delta(z-X)
=τ(e00)δα+iβ+τ(e01)δα+iβ+τ(e10)δα+iβ+τ(e11)δα+iβ\displaystyle=\tau(e_{00})\delta_{\alpha+i\beta}+\tau(e_{01})\delta_{\alpha+i\beta^{\prime}}+\tau(e_{10})\delta_{\alpha^{\prime}+i\beta}+\tau(e_{11})\delta_{\alpha^{\prime}+i\beta^{\prime}}
+τ(e)μ,\displaystyle\quad+\tau(e)\mu^{\prime}\,,

where

(64) μ=(λ1)(ν)+(λ2)(ν)2.\mu^{\prime}=\frac{(\lambda_{1})_{*}(\nu)+(\lambda_{2})_{*}(\nu)}{2}\,.

Additionally, μ({α+iβ,α+iβ,α+iβ,α+iβ})=0\mu^{\prime}(\{\alpha+i\beta,\alpha^{\prime}+i\beta,\alpha+i\beta^{\prime},\alpha^{\prime}+i\beta^{\prime}\})=0.

Proof.

If e=0e=0, the result follows from directly applying 12π2log|λ|=δλ\frac{1}{2\pi}\nabla^{2}\log\left\lvert\cdot-\lambda\right\rvert=\delta_{\lambda} to the expression for logΔ(zX)\log\Delta(z-X) in Proposition 4.2.

If e0e\neq 0, we take the distributional Laplacian of the result from Proposition 4.2:

(65) 12π2logΔ(zX)\displaystyle\frac{1}{2\pi}\nabla^{2}\log\Delta(z-X) =12π2120log(x)𝑑νz(x)\displaystyle=\frac{1}{2\pi}\nabla^{2}\frac{1}{2}\int_{0}^{\infty}\log(x)\,d\nu_{z}(x)
=12π2(τ(e00)log|z(α+iβ)|\displaystyle=\frac{1}{2\pi}\nabla^{2}\biggl{(}\tau(e_{00})\log\left\lvert z-(\alpha+i\beta)\right\rvert
+τ(e01)log|z(α+iβ)|\displaystyle\quad+\tau(e_{01})\log\left\lvert z-(\alpha+i\beta^{\prime})\right\rvert
+τ(e10)log|z(α+iβ)|\displaystyle\quad+\tau(e_{10})\log\left\lvert z-(\alpha^{\prime}+i\beta)\right\rvert
+τ(e11)log|z(α+iβ)|\displaystyle\quad+\tau(e_{11})\log\left\lvert z-(\alpha^{\prime}+i\beta^{\prime})\right\rvert
+τ(e)0π/2log|zλ1(θ)|+log|zλ2(θ)|2dν(θ)).\displaystyle\quad+\tau(e)\int_{0}^{\pi/2}\frac{\log\left\lvert z-\lambda_{1}(\theta)\right\rvert+\log\left\lvert z-\lambda_{2}(\theta)\right\rvert}{2}\,d\nu(\theta)\biggr{)}.

As in the case when e=0e=0, applying 12π2log|λ|=δλ\frac{1}{2\pi}\nabla^{2}\log\left\lvert\cdot-\lambda\right\rvert=\delta_{\lambda} directly to the first 44 atomic terms of logΔ(zX)\log\Delta(z-X) produces the weighted sum of the 44 atoms in μ\mu.

To apply 12π2log|λ|=δλ\frac{1}{2\pi}\nabla^{2}\log\left\lvert\cdot-\lambda\right\rvert=\delta_{\lambda} for the final integral, we apply Fubini’s theorem. Consider fCc()f\in C_{c}^{\infty}(\mathbb{C}). Since log|zw|Lloc1()\log\left\lvert z-w\right\rvert\in L^{1}_{\text{loc}}(\mathbb{C}) for all ww\in\mathbb{C}, then we may apply Fubini’s theorem for i=1,2i=1,2:

(66) f,12π20π/2log|zλi(θ)|𝑑ν(θ)\displaystyle\langle f,\frac{1}{2\pi}\nabla^{2}\int_{0}^{\pi/2}\log\left\lvert z-\lambda_{i}(\theta)\right\rvert\,d\nu(\theta)\rangle
=2f,12π0π/2log|zλi(θ)|𝑑ν(θ)\displaystyle=\langle\nabla^{2}f,\frac{1}{2\pi}\int_{0}^{\pi/2}\log\left\lvert z-\lambda_{i}(\theta)\right\rvert\,d\nu(\theta)\rangle
=2f(z)(12π0π/2log|zλi(θ)|𝑑ν(θ))𝑑λ(z)\displaystyle=\int_{\mathbb{C}}\nabla^{2}f(z)\left(\frac{1}{2\pi}\int_{0}^{\pi/2}\log\left\lvert z-\lambda_{i}(\theta)\right\rvert\,d\nu(\theta)\right)\,d\lambda(z)
=0π/2(12π2f(z)log|zλi(θ)|𝑑dλ(z))𝑑ν(θ)\displaystyle=\int_{0}^{\pi/2}\left(\int_{\mathbb{C}}\frac{1}{2\pi}\nabla^{2}f(z)\log\left\lvert z-\lambda_{i}(\theta)\right\rvert\,dd\lambda(z)\right)\,d\nu(\theta)
=0π/2f(λi(θ))𝑑ν(θ).\displaystyle=\int_{0}^{\pi/2}f(\lambda_{i}(\theta))\,d\nu(\theta)\,.

Hence, for i=1,2i=1,2,

(67) 12π20π/2log|zλi(θ)|𝑑ν(θ)=(λi)(ν).\frac{1}{2\pi}\nabla^{2}\int_{0}^{\pi/2}\log\left\lvert z-\lambda_{i}(\theta)\right\rvert\,d\nu(\theta)=(\lambda_{i})_{*}(\nu)\,.

Thus, the Laplacian of the integral is:

(68) 12π20π/2log|zλ1(θ)|+log|zλ2(θ)|2𝑑ν(θ)=(λ1)(ν)+(λ2)(ν)2=μ.\frac{1}{2\pi}\nabla^{2}\int_{0}^{\pi/2}\frac{\log\left\lvert z-\lambda_{1}(\theta)\right\rvert+\log\left\lvert z-\lambda_{2}(\theta)\right\rvert}{2}\,d\nu(\theta)=\frac{(\lambda_{1})_{*}(\nu)+(\lambda_{2})_{*}(\nu)}{2}=\mu^{\prime}.

Combining this with the atomic terms gives the desired Brown measure for XX.

For the final note, it follows from Lemma 2.3 and the fact that λi({0,π/2})={α+iβ,α+iβ,α+iβ,α+iβ}\lambda_{i}(\{0,\pi/2\})=\{\alpha+i\beta,\alpha^{\prime}+i\beta,\alpha+i\beta^{\prime},\alpha^{\prime}+i\beta^{\prime}\} that

(69) μ({α+iβ,α+iβ,α+iβ,α+iβ})=ν({0,π/2})=0.\mu^{\prime}(\{\alpha+i\beta,\alpha^{\prime}+i\beta,\alpha+i\beta^{\prime},\alpha^{\prime}+i\beta^{\prime}\})=\nu(\{0,\pi/2\})=0\,.

Even though the weights τ(eij),τ(e)\tau(e_{ij}),\tau(e) and the measure ν\nu have not been determined yet, we can already say something about the support of the Brown measure in general.

First, we prove a lemma about a relevant hyperbola and rectangle:

Lemma 4.4.

Let α,α,β,β\alpha,\alpha^{\prime},\beta,\beta^{\prime}\in\mathbb{R}, where αα\alpha\neq\alpha^{\prime} and ββ\beta\neq\beta^{\prime}. Let 𝒜=αα\mathscr{A}=\alpha^{\prime}-\alpha and =ββ\mathscr{B}=\beta^{\prime}-\beta.

Let

(70) H\displaystyle H ={z=x+iy:(xα+α2)2(yβ+β2)2=𝒜224}\displaystyle=\left\{z=x+iy\in\mathbb{C}:\left(x-\frac{\alpha+\alpha^{\prime}}{2}\right)^{2}-\left(y-\frac{\beta+\beta^{\prime}}{2}\right)^{2}=\frac{\mathscr{A}^{2}-\mathscr{B}^{2}}{4}\right\}
R\displaystyle R ={z=x+iy:x[αα,αα],y[ββ,ββ]}.\displaystyle=\left\{z=x+iy\in\mathbb{C}:x\in[\alpha\wedge\alpha^{\prime},\alpha\vee\alpha^{\prime}],y\in[\beta\wedge\beta^{\prime},\beta\vee\beta^{\prime}]\right\}\,.

The equation of HH is equivalent to:

(71) (xα)(xα)=(yβ)(yβ).(x-\alpha)(x-\alpha^{\prime})=(y-\beta)(y-\beta^{\prime})\,.

The equation of HH in coordinates

(72) x\displaystyle x^{\prime} =xα+α2\displaystyle=x-\frac{\alpha+\alpha^{\prime}}{2}
y\displaystyle y^{\prime} =yβ+β2\displaystyle=y-\frac{\beta+\beta^{\prime}}{2}

is

(73) (x)2𝒜24=(y)224.(x^{\prime})^{2}-\frac{\mathscr{A}^{2}}{4}=(y^{\prime})^{2}-\frac{\mathscr{B}^{2}}{4}.

It follows that for (x,y)H(x,y)\in H,

(74) (x,y)R(73)0x[αα,αα] or y[ββ,ββ].(x,y)\in R\iff(\ref{eqn:lem:hyperbola_rectangle})\leq 0\iff x\in[\alpha\wedge\alpha^{\prime},\alpha\vee\alpha^{\prime}]\text{ or }y\in[\beta\wedge\beta^{\prime},\beta\vee\beta^{\prime}]\,.

Similarly,

(75) (x,y)R(73)<0x(αα,αα) or y(ββ,ββ).(x,y)\in\accentset{\circ}{R}\iff(\ref{eqn:lem:hyperbola_rectangle})<0\iff x\in(\alpha\wedge\alpha^{\prime},\alpha\vee\alpha^{\prime})\text{ or }y\in(\beta\wedge\beta^{\prime},\beta\vee\beta^{\prime})\,.

Alternatively, the equation of the hyperbola is:

(76) Re((zα+α2iβ+β2)2)=𝒜224.\text{{Re}}\left(\left(z-\frac{\alpha+\alpha^{\prime}}{2}-i\frac{\beta+\beta^{\prime}}{2}\right)^{2}\right)=\frac{\mathscr{A}^{2}-\mathscr{B}^{2}}{4}\,.

If zHz\in H, then zRz\in R if and only if

(77) |Im((zα+α2iβ+β2)2)||𝒜|2.\left\lvert\text{{Im}}\left(\left(z-\frac{\alpha+\alpha^{\prime}}{2}-i\frac{\beta+\beta^{\prime}}{2}\right)^{2}\right)\right\rvert\leq\frac{\left\lvert\mathscr{A}\mathscr{B}\right\rvert}{2}\,.
Proof.

The equivalent equations for the hyperbola are straightforward to check.

The equivalences for the closed conditions follow from the following equivalences and the equation of the hyperbola in x,yx^{\prime},y^{\prime} coordinates

(78) (x,y)R\displaystyle(x,y)\in R (x)2𝒜240 and (y)2240\displaystyle\iff(x^{\prime})^{2}-\frac{\mathscr{A}^{2}}{4}\leq 0\text{ and }(y^{\prime})^{2}-\frac{\mathscr{B}^{2}}{4}\leq 0
(79) (x)2𝒜240\displaystyle(x^{\prime})^{2}-\frac{\mathscr{A}^{2}}{4}\leq 0 |x||𝒜|2\displaystyle\iff\left\lvert x^{\prime}\right\rvert\leq\frac{\left\lvert\mathscr{A}\right\rvert}{2} x[αα,αα]\displaystyle\iff x\in[\alpha\wedge\alpha^{\prime},\alpha\vee\alpha^{\prime}]
(y)2240\displaystyle(y^{\prime})^{2}-\frac{\mathscr{B}^{2}}{4}\leq 0 |y|||2\displaystyle\iff\left\lvert y^{\prime}\right\rvert\leq\frac{\left\lvert\mathscr{B}\right\rvert}{2} y[ββ,ββ].\displaystyle\iff y\in[\beta\wedge\beta^{\prime},\beta\vee\beta^{\prime}]\,.

The equivalences for the open conditions follow from similar equivalences with the closed conditions replaced by open conditions.

The last equation of the hyperbola follows from direct computation. For the inequality of the rectangle, observe that

(80) Im((zα+α2iβ+β2)2)=2xy.\text{{Im}}\left(\left(z-\frac{\alpha+\alpha^{\prime}}{2}-i\frac{\beta+\beta^{\prime}}{2}\right)^{2}\right)=2x^{\prime}y^{\prime}\,.

In light of what was previously shown,

(81) xy|𝒜|4x|𝒜|2 or y||2zR.x^{\prime}y^{\prime}\leq\frac{\left\lvert\mathscr{A}\mathscr{B}\right\rvert}{4}\Longrightarrow x^{\prime}\leq\frac{\left\lvert\mathscr{A}\right\rvert}{2}\text{ or }y^{\prime}\leq\frac{\left\lvert\mathscr{B}\right\rvert}{2}\Longrightarrow z\in R\,.

Conversely,

(82) zRx|𝒜|2 and y||2xy|𝒜|4.z\in R\Longrightarrow x^{\prime}\leq\frac{\left\lvert\mathscr{A}\right\rvert}{2}\text{ and }y^{\prime}\leq\frac{\left\lvert\mathscr{B}\right\rvert}{2}\Longrightarrow x^{\prime}y^{\prime}\leq\frac{\left\lvert\mathscr{A}\mathscr{B}\right\rvert}{4}\,.

Now, we will show that the support of the Brown measure is contained in HRH\cap R:

Corollary 4.5.

Let 𝒜=αα\mathscr{A}=\alpha^{\prime}-\alpha and =ββ\mathscr{B}=\beta^{\prime}-\beta.

The continuous functions λ1,λ2:[0,π/2]\lambda_{1},\lambda_{2}:[0,\pi/2]\to\mathbb{C} in Proposition 4.2 parameterize the intersection of the hyperbola

(83) H={z=x+iy:(xα+α2)2(yβ+β2)2=𝒜224}H=\left\{z=x+iy\in\mathbb{C}:\left(x-\frac{\alpha+\alpha^{\prime}}{2}\right)^{2}-\left(y-\frac{\beta+\beta^{\prime}}{2}\right)^{2}=\frac{\mathscr{A}^{2}-\mathscr{B}^{2}}{4}\right\}

with the rectangle

(84) R={z=x+iy:x[αα,αα],y[ββ,ββ]}.R=\left\{z=x+iy\in\mathbb{C}:x\in[\alpha\wedge\alpha^{\prime},\alpha\vee\alpha^{\prime}],y\in[\beta\wedge\beta^{\prime},\beta\vee\beta^{\prime}]\right\}\,.

When |𝒜|||\left\lvert\mathscr{A}\right\rvert\geq\left\lvert\mathscr{B}\right\rvert, λ1\lambda_{1} parameterizes the left component of HRH\cap R and λ2\lambda_{2} parameterizes the right component of HRH\cap R. When |𝒜|<||\left\lvert\mathscr{A}\right\rvert<\left\lvert\mathscr{B}\right\rvert, λ1\lambda_{1} parameterizes the bottom component of HRH\cap R and λ2\lambda_{2} parameterizes the top component of HRH\cap R.

The support of the Brown measure is contained in HRH\cap R.

Proof.

From Lemma 4.4, z=x+iyz=x+iy is on HRH\cap R if and only if

(85) (zα+α2iβ+β2)2=𝒜22+2i𝒜cos(2θ)4 for θ[0,π/2].\left(z-\frac{\alpha+\alpha^{\prime}}{2}-i\frac{\beta+\beta^{\prime}}{2}\right)^{2}=\frac{\mathscr{A}^{2}-\mathscr{B}^{2}+2i\mathscr{A}\mathscr{B}\cos(2\theta)}{4}\quad\text{ for }\theta\in[0,\pi/2]\,.

Note that λ1(θ),λ2(θ)\lambda_{1}(\theta),\lambda_{2}(\theta) are exactly the solutions to this equation. Hence, the λi(θ)\lambda_{i}(\theta) parameterize the intersection of the hyperbola and rectangle.

For the cases of λi\lambda_{i} parameterizing the left/right or top/bottom components, it is easy to see from the formulas that the λi\lambda_{i} map into the left/right or top/bottom components, and since the λi(θ)\lambda_{i}(\theta) parameterize all of HRH\cap R, then the λi\lambda_{i} have to parameterize the entire left/right or top/bottom components.

As the λi\lambda_{i} parameterize HRH\cap R, then μ\mu^{\prime} is supported on HRH\cap R. The 4 atoms in the Brown measure are on (R)H(\partial R)\cap H (at the 4 corners of the rectangle).

Thus, we conclude that the Brown measure is supported on HRH\cap R. ∎

Figure 1 illustrates the conclusion of Corollary 4.5 where the Brown measure of X=p+iqX=p+iq is approximated by the ESD of Xn=Pn+iQnX_{n}=P_{n}+iQ_{n} for deterministic Pn,QnMn()P_{n},Q_{n}\in M_{n}(\mathbb{C}).

Refer to caption
Figure 1. ESD of Xn=Pn+iQnX_{n}=P_{n}+iQ_{n}
μPn=(1/2)δ0+(1/2)δ1\mu_{P_{n}}=(1/2)\delta_{0}+(1/2)\delta_{1}
μQn=(1/2)δ0+(1/2)δ4/5\mu_{Q_{n}}=(1/2)\delta_{0}+(1/2)\delta_{4/5}
n=1000n=1000

Motivated by the hyperbola and rectangle appearing in Corollary 4.5, we introduce the definition of the hyperbola and rectangle associated with XX:

Definition 4.6.

Let p,q(M,τ)p,q\in(M,\tau) be Hermitian with laws:

(86) μp\displaystyle\mu_{p} =aδα+(1a)δα\displaystyle=a\delta_{\alpha}+(1-a)\delta_{\alpha^{\prime}}
μq\displaystyle\mu_{q} =bδβ+(1b)δβ.\displaystyle=b\delta_{\beta}+(1-b)\delta_{\beta^{\prime}}\,.

Let 𝒜=αα\mathscr{A}=\alpha^{\prime}-\alpha and =ββ\mathscr{B}=\beta^{\prime}-\beta.

The hyperbola associated with XX is

(87) H={z=x+iy:(xα+α2)2(yβ+β2)2=𝒜224}.H=\left\{z=x+iy\in\mathbb{C}:\left(x-\frac{\alpha+\alpha^{\prime}}{2}\right)^{2}-\left(y-\frac{\beta+\beta^{\prime}}{2}\right)^{2}=\frac{\mathscr{A}^{2}-\mathscr{B}^{2}}{4}\right\}\,.

The rectangle associated with XX is

(88) R={z=x+iy:x[αα,αα],y[ββ,ββ]}.R=\left\{z=x+iy\in\mathbb{C}:x\in[\alpha\wedge\alpha^{\prime},\alpha\vee\alpha^{\prime}],y\in[\beta\wedge\beta^{\prime},\beta\vee\beta^{\prime}]\right\}\,.

While the weights τ(eij),τ(e)\tau(e_{ij}),\tau(e) are as of yet undetermined in the case pp and qq are free, there are some general relationships between the weights and traces of the spectral projections of pp and qq:

Proposition 4.7.

Let μ\mu be the Brown measure of XX and let μ\mu^{\prime} be as in Proposition 4.3.

Then, the μ\mu^{\prime} measure of each of the two components of HRH\cap R is equal to 1/21/2. Additionally,

(89) τ(χ{α}(p))\displaystyle\tau\left(\chi_{\{\alpha\}}(p)\right) =μ({α+iβ,α+iβ})+τ(e)/2\displaystyle=\mu(\{\alpha+i\beta,\alpha+i\beta^{\prime}\})+\tau(e)/2
τ(χ{α}(p))\displaystyle\tau\left(\chi_{\{\alpha^{\prime}\}}(p)\right) =μ({α+iβ,α+iβ})+τ(e)/2\displaystyle=\mu(\{\alpha^{\prime}+i\beta,\alpha^{\prime}+i\beta^{\prime}\})+\tau(e)/2
τ(χ{β}(q))\displaystyle\tau\left(\chi_{\{\beta\}}(q)\right) =μ({α+iβ,α+iβ})+τ(e)/2\displaystyle=\mu(\{\alpha+i\beta,\alpha^{\prime}+i\beta\})+\tau(e)/2
τ(χ{β}(q))\displaystyle\tau\left(\chi_{\{\beta^{\prime}\}}(q)\right) =μ({α+iβ,α+iβ})+τ(e)/2.\displaystyle=\mu(\{\alpha+i\beta^{\prime},\alpha^{\prime}+i\beta^{\prime}\})+\tau(e)/2\,.

If |𝒜|||\left\lvert\mathscr{A}\right\rvert\geq\left\lvert\mathscr{B}\right\rvert, the Brown measure of the left component of HRH\cap R is τ(χ{αα}(p))\tau(\chi_{\{\alpha\wedge\alpha^{\prime}\}(p)}) and the Brown measure of the right component of HRH\cap R is τ(χ{αα}(p))\tau(\chi_{\{\alpha\vee\alpha^{\prime}\}}(p)).

If |𝒜|<||\left\lvert\mathscr{A}\right\rvert<\left\lvert\mathscr{B}\right\rvert, the Brown measure of the bottom component of HRH\cap R is τ(χ{ββ}(q))\tau(\chi_{\{\beta\wedge\beta^{\prime}\}}(q)) and the Brown measure of the top component of HRH\cap R is τ(χ{ββ}(q))\tau(\chi_{\{\beta\vee\beta^{\prime}\}}(q)).

Proof.

Recall that

(90) μ=(λ1)(ν)+(λ2)(ν)2,\mu^{\prime}=\frac{(\lambda_{1})_{*}(\nu)+(\lambda_{2})_{*}(\nu)}{2}\,,

where ν\nu is a probability measure on (0,π/2)(0,\pi/2). From Corollary 4.5, the λi\lambda_{i} each parameterize one of the components of HRH\cap R. Hence, the μ\mu^{\prime} measure of each component of HRH\cap R is 1/21/2.

For the equations of the traces of spectral projections of pp and qq, we will just prove the first equation, the others are similar. From (39),

(91) χ{a}(p)=e00+e01+eχ{a}(p)e.\chi_{\{a\}}(p)=e_{00}+e_{01}+e\chi_{\{a\}}(p)e\,.

From Proposition 4.3, μ({α+iβ})=τ(e00)\mu(\{\alpha+i\beta\})=\tau(e_{00}) and μ({α+iβ})=τ(e01)\mu(\{\alpha+i\beta^{\prime}\})=\tau(e_{01}). From Lemma 2.4, τ(eχ{a}(p)e)=τ(e)/2\tau(e\chi_{\{a\}}(p)e)=\tau(e)/2. The desired equation follows from taking the trace of the equation and using these facts.

For the last point, we consider the Brown measure of the left component of HRH\cap R when |𝒜|||\left\lvert\mathscr{A}\right\rvert\geq\left\lvert\mathscr{B}\right\rvert as the other cases are similar. Let LL be this left component. Then,

(92) μ(L)\displaystyle\mu(L) =μ({αα+iβ,αα+iβ})+τ(e)μ(L)\displaystyle=\mu(\{\alpha\wedge\alpha^{\prime}+i\beta,\alpha\wedge\alpha^{\prime}+i\beta^{\prime}\})+\tau(e)\mu^{\prime}(L)
=μ({αα+iβ,αα+iβ})+τ(e)/2\displaystyle=\mu(\{\alpha\wedge\alpha^{\prime}+i\beta,\alpha\wedge\alpha^{\prime}+i\beta^{\prime}\})+\tau(e)/2
=τ(χ{αα}(p)).\displaystyle=\tau(\chi_{\{\alpha\wedge\alpha^{\prime}\}}(p))\,.

5. Computation of Weights and Measure

In this section, we will determine the weights τ(eij),τ(e)\tau(e_{ij}),\tau(e) and the measure ν\nu in Proposition 4.3 under the assumption that pp and qq are freely independent. We also deduce that the operators of the form X=p+iqX=p+iq, where p,qp,q Hermitian, freely independent, and have 22 atoms in their spectra are not normal (Corollary 5.5).

We will use the functions ψμ\psi_{\mu}, χμ\chi_{\mu}, and SμS_{\mu} that were introduced in Subsection 2.3. Recall that for x(M,τ)x\in(M,\tau), we will use ψx\psi_{x}, χx\chi_{x}, and SxS_{x} to denote the respective functions with respect to μx\mu_{x}, the spectral measure of xx.

Recall that the measure ν\nu and the weights τ(eij),τ(e)\tau(e_{ij}),\tau(e) are computed in terms of projections p,qp^{\prime},q^{\prime} where

(93) p\displaystyle p^{\prime} =χ{α}(p)\displaystyle=\chi_{\{\alpha^{\prime}\}}(p)
q\displaystyle q^{\prime} =χ{β}(q).\displaystyle=\chi_{\{\beta^{\prime}\}}(q)\,.

Since pp and qq are freely independent, then pp^{\prime} and qq^{\prime} are also freely independent. This along with the traces τ(p)=1a\tau(p^{\prime})=1-a and τ(q)=1b\tau(q^{\prime})=1-b determine the joint law of pp^{\prime}, qq^{\prime}.

For notational convenience, in this section we will consider two general projections pp and qq that are freely independent and τ(p)=a\tau(p)=a, τ(q)=b\tau(q)=b for some a,b(0,1)a,b\in(0,1). This is a natural continuation of Subsection 2.2. It is easy to translate the results in this section to the p,qp^{\prime},q^{\prime} defined in Section 3.

Recall from Subsection 2.2 that the weights τ(eij),τ(e)\tau(e_{ij}),\tau(e) are:

(94) τ(e00)\displaystyle\tau(e_{00}) =τ((1p)(1q))\displaystyle=\tau((1-p)\wedge(1-q))
τ(e01)\displaystyle\tau(e_{01}) =τ((1p)q)\displaystyle=\tau((1-p)\wedge q)
τ(e10)\displaystyle\tau(e_{10}) =τ(p(1q))\displaystyle=\tau(p\wedge(1-q))
τ(e11)\displaystyle\tau(e_{11}) =τ(pq)\displaystyle=\tau(p\wedge q)
τ(e)\displaystyle\tau(e) =1(τ(e00)+τ(e01)+τ(e10)+τ(e11)).\displaystyle=1-(\tau(e_{00})+\tau(e_{01})+\tau(e_{10})+\tau(e_{11}))\,.

Recall that when e0e\neq 0, ((0,π/2),ν)((0,\pi/2),\nu) is the pushforward measure of ((0,1),ν)((0,1),\nu^{*}) under the inverse of of cos2(θ)\cos^{2}(\theta) and ν\nu^{*} is the spectral measure of exeexe, where x=pqp+(1p)(1q)(1p)x=pqp+(1-p)(1-q)(1-p).

Since τ((pqp)n)τ(pq)\tau((pqp)^{n})\to\tau(p\wedge q) as nn\to\infty, then understanding the laws of pqppqp and (1p)(1q)(1p)(1-p)(1-q)(1-p) are relevant for computing both ν\nu and the weights τ(eij),τ(e)\tau(e_{ij}),\tau(e). The first step to this is computing the relevant free probability functions:

Proposition 5.1.

Let p,q(M,τ)p,q\in(M,\tau) be two freely independent projections with τ(p)=a\tau(p)=a, τ(q)=b\tau(q)=b, a,b(0,1)a,b\in(0,1). Then,

(95) ψp(z)=az1z\displaystyle\psi_{p}(z)=\frac{az}{1-z} ψq(z)=bz1z\displaystyle\qquad\psi_{q}(z)=\frac{bz}{1-z}
χp(w)=ww+a\displaystyle\chi_{p}(w)=\frac{w}{w+a} χq(w)=ww+b\displaystyle\qquad\chi_{q}(w)=\frac{w}{w+b}
(96) χpqp(w)=w(1+w)(w+a)(w+b).\chi_{pqp}(w)=\frac{w(1+w)}{(w+a)(w+b)}\,.

Let

(97) f(z)=1+(4ab2(a+b))z+(ab)2z2.f(z)=1+(4ab-2(a+b))z+(a-b)^{2}z^{2}\,.

Then, ψpqp,ψ(1p)(1q)(1p)\psi_{pqp},\psi_{(1-p)(1-q)(1-p)} are analytic on [1,)\mathbb{C}\setminus[1,\infty) and

(98) ψpqp(z)\displaystyle\psi_{pqp}(z) =1(a+b)zf(z)2(z1)\displaystyle=\frac{1-(a+b)z-\sqrt{f(z)}}{2(z-1)}
ψ(1p)(1q)(1p)(z)\displaystyle\psi_{(1-p)(1-q)(1-p)}(z) =1(2ab)zf(z)2(z1),\displaystyle=\frac{1-(2-a-b)z-\sqrt{f(z)}}{2(z-1)}\,,

where f(z)\sqrt{f(z)} is an analytic branch of the square root of f(z)f(z) on [1,)\mathbb{C}\setminus[1,\infty) where 1=+1\sqrt{1}=+1. In particular, the root(s) of f(z)f(z) are in [1,)[1,\infty) and are distinct when ff is quadratic.

Proof.

The formula for ψp\psi_{p} follows from

(99) μp=(1a)δ0+aδ1\mu_{p}=(1-a)\delta_{0}+a\delta_{1}

and similarly for ψq\psi_{q}. Since ψp\psi_{p}, ψq\psi_{q} are linear fractional transformations, the formulas for their inverses (χp\chi_{p}, χq\chi_{q}, respectively) are easily computed.

Since pp and qq are freely independent, then the formula for χpqp\chi_{pqp} follows from the multiplicativity of the SS-transform and its relationship with χ\chi.

Since pqp1\left\lVert pqp\right\rVert\leq 1 then μpqp\mu_{pqp} is supported on [0,1][0,1]. So, the formula

(100) ψpqp(z)=0tz1tz𝑑μpqp(t)=01tz1tz𝑑μpqp(t)\psi_{pqp}(z)=\int_{0}^{\infty}\frac{tz}{1-tz}\,d\mu_{pqp}(t)=\int_{0}^{1}\frac{tz}{1-tz}\,d\mu_{pqp}(t)

defines an analytic function for z[1,)z\in\mathbb{C}\setminus[1,\infty).

To compute the formula for ψpqp\psi_{pqp}, recall that ψpqp=χpqp1\psi_{pqp}=\chi_{pqp}^{-1} in a neighborhood of 0. Then,

(101) z=χpqp(w)=w(1+w)(w+a)(w+b)z=\chi_{pqp}(w)=\frac{w(1+w)}{(w+a)(w+b)}

if and only if

(102) z(w+a)(w+b)=w(1+w).z(w+a)(w+b)=w(1+w)\,.

This is clearly true for wa,bw\neq-a,-b and the latter equation is not satisfied for w=a,b(1,0)w=-a,-b\in(-1,0) at any zz.

This last equation is true if and only if

(103) (z1)w2+(z(a+b)1)w+zab=0.(z-1)w^{2}+(z(a+b)-1)w+zab=0\,.

Fixing a zz and solving for ww, then from the quadratic formula and simplifying, we obtain the desired formula for ψpqp\psi_{pqp}. The sign of the square root follows from the general fact that ψμ(0)=0\psi_{\mu}(0)=0.

The formula for ψ(1p)(1q)(1p)\psi_{(1-p)(1-q)(1-p)} follows from the formula for ψpqp\psi_{pqp} and noting that 1p1-p and 1q1-q are freely independent projections with traces τ(p)=1a\tau(p)=1-a, τ(q)=1b\tau(q)=1-b. Observe that f(z)f(z) is invariant under changing the pair (a,b)(a,b) to (1a,1b)(1-a,1-b).

Finally, note that the f(z)\sqrt{f(z)} in both ψpqp\psi_{pqp} and ψ(1p)(1q)(1p)\psi_{(1-p)(1-q)(1-p)} are identical since they are defined on the domain [1,)\mathbb{C}\setminus[1,\infty) and agree at z=0z=0.

The fact that f(z)f(z) has root(s) in [1,)[1,\infty) follows from the fact that w\sqrt{w} is not analytic in any neighborhood of 0, so the root(s) of ff cannot be on [1,)\mathbb{C}\setminus[1,\infty).

For distinctness of the roots when ff is quadratic, the discriminant is 16ab(1a)(1b)>016ab(1-a)(1-b)>0 for a,b(0,1)a,b\in(0,1). ∎

Now, we proceed to determine τ(eij),τ(e)\tau(e_{ij}),\tau(e). First, we need the following Lemma [10, Proposition 8]:

Lemma 5.2.

Let μ\mu be a finite measure on the real line and ss\in\mathbb{R}. For a sequence znsz_{n}\to s non-tangentially to \mathbb{R}, (zns)Gμ(zn)μ({s})(z_{n}-s)G_{\mu}(z_{n})\to\mu(\{s\}).

Proof.

Let zn=xn+iynz_{n}=x_{n}+iy_{n}. The condition that znsz_{n}\to s non-tangentially to \mathbb{R} is equivalent to |(xns)/yn|M\left\lvert(x_{n}-s)/y_{n}\right\rvert\leq M for some MM. In particular, yn0y_{n}\neq 0.

Computation shows that

(104) (zns)Gμ(zn)=znsznt𝑑μ(t).(z_{n}-s)G_{\mu}(z_{n})=\int_{-\infty}^{\infty}\frac{z_{n}-s}{z_{n}-t}\,d\mu(t)\,.

Let fn:f_{n}:\mathbb{R}\to\mathbb{C} where

(105) fn(t)=znsznt.f_{n}(t)=\frac{z_{n}-s}{z_{n}-t}\,.

Note that fn(s)=1f_{n}(s)=1 for all nn and limnfn(t)=0\lim\limits_{n\to\infty}f_{n}(t)=0 for all tst\neq s. It suffices to show that the fn(t)f_{n}(t) are uniformly bounded, as then the result follows from the bounded convergence theorem.

First, rewrite fn(t)f_{n}(t):

(106) fn(t)=znsznt=(xns)+iyn(xnt)+iyn=(xns)/yn+i(xnt)/yn+i.f_{n}(t)=\frac{z_{n}-s}{z_{n}-t}=\frac{(x_{n}-s)+iy_{n}}{(x_{n}-t)+iy_{n}}=\frac{(x_{n}-s)/y_{n}+i}{(x_{n}-t)/y_{n}+i}\,.

Since |(xns)/yn+i||(xns)/yn|+1=1+M\left\lvert(x_{n}-s)/y_{n}+i\right\rvert\leq\left\lvert(x_{n}-s)/y_{n}\right\rvert+1=1+M and |(xnt)/yn+i|1\left\lvert(x_{n}-t)/y_{n}+i\right\rvert\geq 1,

(107) |fn(t)|=|(xns)/yn+i||(xnt)/yn+i|1+M1=1+M.\left\lvert f_{n}(t)\right\rvert=\frac{\left\lvert(x_{n}-s)/y_{n}+i\right\rvert}{\left\lvert(x_{n}-t)/y_{n}+i\right\rvert}\leq\frac{1+M}{1}=1+M\,.

as desired. ∎

Next, we determine τ(eij),τ(e)\tau(e_{ij}),\tau(e) using the free independence of pp and qq:

Proposition 5.3.

Let p,q(M,τ)p,q\in(M,\tau) be two freely independent projections with τ(p)=a\tau(p)=a, τ(q)=b\tau(q)=b, a,b(0,1)a,b\in(0,1). Then,

(108) τ(e00)\displaystyle\tau(e_{00}) =τ((1p)(1q))=max(0,(1a)+(1b)1)\displaystyle=\tau((1-p)\wedge(1-q))=\max(0,(1-a)+(1-b)-1)
τ(e01)\displaystyle\tau(e_{01}) =τ((1p)q)=max(0,(1a)+b1)\displaystyle=\tau((1-p)\wedge q)=\max(0,(1-a)+b-1)
τ(e10)\displaystyle\tau(e_{10}) =τ(p(1q))=max(0,a+(1b)1)\displaystyle=\tau(p\wedge(1-q))=\max(0,a+(1-b)-1)
τ(e11)\displaystyle\tau(e_{11}) =τ(pq)=max(0,a+b1)\displaystyle=\tau(p\wedge q)=\max(0,a+b-1)
τ(e)\displaystyle\tau(e) =1(τ(e00)+τ(e01)+τ(e10)+τ(e11)).\displaystyle=1-(\tau(e_{00})+\tau(e_{01})+\tau(e_{10})+\tau(e_{11}))\,.
Proof.

By replacing pp with 1p1-p and/or qq with 1q1-q, it suffices to just prove

(109) τ(pq)=max(0,a+b1).\tau(p\wedge q)=\max(0,a+b-1)\,.

Recall that τ((pqp)n)τ(pq)\tau((pqp)^{n})\to\tau(p\wedge q) as nn\to\infty. Since xnχ{1}(x)x^{n}\to\chi_{\{1\}}(x) on [0,1][0,1] and σ(pqp)[0,1]\sigma(pqp)\subset[0,1], then τ((pqp)n)μpqp({1})\tau((pqp)^{n})\to\mu_{pqp}(\{1\}). Hence,

(110) τ(pq)=μpqp({1}).\tau(p\wedge q)=\mu_{pqp}(\{1\})\,.

We proceed to use Proposition 5.1 and Lemma 5.2 to determine μpqp({1})\mu_{pqp}(\{1\}).

In general, for zσ(pqp)z\in\mathbb{C}\setminus\sigma(pqp),

(111) Gpqp(z)=1z(ψpqp(1z)+1).G_{pqp}(z)=\frac{1}{z}\left(\psi_{pqp}\left(\frac{1}{z}\right)+1\right)\,.

From Proposition 5.1, the right-hand side of this equation is:

(112) 1z(ψpqp(1z)+1)=z+(a+b2)+zf(1/z)2z.\frac{1}{z}\left(\psi_{pqp}\left(\frac{1}{z}\right)+1\right)=\frac{z+(a+b-2)+z\sqrt{f(1/z)}}{2z}\,.

and is defined on [0,1]\mathbb{C}\setminus[0,1]. Since σ(pqp)[0,1]\sigma(pqp)\subset[0,1], then GpqpG_{pqp} is also defined on [0,1]\mathbb{C}\setminus[0,1], so then the following equality holds for z[0,1]z\in\mathbb{C}\setminus[0,1]:

(113) Gpqp(z)=z+(a+b2)+zf(1/z)2z(z1).G_{pqp}(z)=\frac{z+(a+b-2)+z\sqrt{f(1/z)}}{2z(z-1)}\,.

Thus, we may use this formula and Lemma 5.2 to obtain:

(114) μpqp({1})\displaystyle\mu_{pqp}(\{1\}) =limz1(z1)Gpqp(z)\displaystyle=\lim\limits_{z\to 1}(z-1)G_{pqp}(z)
=limz1z+(a+b2)+zf(1/z)2z\displaystyle=\lim\limits_{z\to 1}\frac{z+(a+b-2)+z\sqrt{f(1/z)}}{2z}
=a+b1+|a+b1|2\displaystyle=\frac{a+b-1+\left\lvert a+b-1\right\rvert}{2}
=max(0,a+b1).\displaystyle=\max(0,a+b-1)\,.

Proposition 5.3 can be summarized by: “free projections intersect as little as possible.” For τ(pq)=max(0,a+b1)\tau(p\wedge q)=\max(0,a+b-1), the term max(0,a+b1)\max(0,a+b-1) is just the minimum trace of the intersection between two projections pp and qq where of τ(p)=a\tau(p)=a and τ(q)=b\tau(q)=b:

Recall from the parallelogram law that for projections p,q(M,τ)p,q\in(M,\tau),

(115) τ(pq)=τ(p)+τ(q)τ(pq)τ(p)τ(q)1=a+b1.\tau(p\wedge q)=\tau(p)+\tau(q)-\tau(p\vee q)\geq\tau(p)-\tau(q)-1=a+b-1\,.

As τ(pq)0\tau(p\wedge q)\geq 0, then for any projections p,q(M,τ)p,q\in(M,\tau), τ(pq)max(0,a+b1)\tau(p\wedge q)\geq\max(0,a+b-1).

As a corollary, we will no longer need to consider the possibility that e=0e=0:

Corollary 5.4.

Let p,q(M,τ)p,q\in(M,\tau) be two freely independent projections. Then, e=0e=0 if and only if one of p,1p,q,1qp,1-p,q,1-q is 0.

Proof.

If one of p,1p,q,1qp,1-p,q,1-q is 0, then it is clear that e=0e=0.

Suppose that none of p,1p,q,1qp,1-p,q,1-q is 0. Then, we may apply Proposition 5.3. As τ\tau is faithful, it suffices to conclude that τ(e)=0\tau(e)=0.

Consider the expressions of aa and bb in Proposition 5.3. Observe that the expressions inside τ(e00)\tau(e_{00}) and τ(e11)\tau(e_{11}) are negatives of each other and hence

(116) τ(e00)+τ(e11)=max(0,(a+b1))+max(0,a+b1)=|a+b1|.\tau(e_{00})+\tau(e_{11})=\max(0,-(a+b-1))+\max(0,a+b-1)=\left\lvert a+b-1\right\rvert.

Similarly,

(117) τ(e01)+τ(e10)=max(0,ba)+max(0,ab)=|ab|.\tau(e_{01})+\tau(e_{10})=\max(0,b-a)+\max(0,a-b)=\left\lvert a-b\right\rvert.

Thus,

(118) τ(e)=0|a+b1|+|ab|=1.\tau(e)=0\iff\left\lvert a+b-1\right\rvert+\left\lvert a-b\right\rvert=1\,.

By considering the four cases of |a+b1|=±(a+b1)\left\lvert a+b-1\right\rvert=\pm(a+b-1) and |ab|=±(ab)\left\lvert a-b\right\rvert=\pm(a-b), we see that |a+b1|+|ab|=1\left\lvert a+b-1\right\rvert+\left\lvert a-b\right\rvert=1 if and only if one of p,1p,q,1qp,1-p,q,1-q is 0. Hence, τ(e)0\tau(e)\neq 0. ∎

As a corollary, we deduce the fact that the operators of the form X=p+iqX=p+iq are not normal:

Corollary 5.5.

Let p,q(M,τ)p,q\in(M,\tau) be Hermitian, freely independent, and have 22 atoms in their spectra. Then, X=p+iqX=p+iq is not normal.

Proof.

First, we consider the case when pp and qq are projections with τ(p)=a\tau(p)=a, τ(q)=b\tau(q)=b, a,b(0,1)a,b\in(0,1).

It suffices to show that pp and qq do not commute. For this, recall that pp and qq commute if and only if pq=pqpq=p\wedge q. Applying this to the other three pairs of commuting projections, (1p,q)(1-p,q), (p,1q)(p,1-q), and (1p,1q)(1-p,1-q), then (1p)q=(1p)q(1-p)q=(1-p)\wedge q, p(1q)=p(1q)p(1-q)=p\wedge(1-q), and (1p)(1q)=(1p)(1q)(1-p)(1-q)=(1-p)\wedge(1-q).

Thus if pp and qq commute, the following equalities hold:

(119) 1\displaystyle 1 =(p+(1p))(q+(1q))\displaystyle=(p+(1-p))(q+(1-q))
=pq+(1p)q+p(1q)+(1p)(1q)\displaystyle=pq+(1-p)q+p(1-q)+(1-p)(1-q)
=pq+(1p)q+p(1q)+(1p)(1q).\displaystyle=p\wedge q+(1-p)\wedge q+p\wedge(1-q)+(1-p)\wedge(1-q)\,.

In the notation of the projections eij,ee_{ij},e, this means that e=0e=0. From Corollary 5.4, this cannot happen if a,b(0,1)a,b\in(0,1). Thus, X=p+iqX=p+iq is not normal.

The general case follows by considering the projections pp^{\prime} and qq^{\prime} defined in Section 3. Then, X=p+iqX=p+iq is normal if and only if X=p+iqX^{\prime}=p^{\prime}+iq^{\prime} is normal, since pp and qq commute if and only if pp^{\prime} and qq^{\prime} commute. From the previous case, XX^{\prime} is not normal. Hence, XX is not normal. ∎

We proceed to compute ν\nu with the knowledge that e0e\neq 0. This requires computing ν\nu^{*}, the spectral measure of exeexe in (eMe,τ|eMe)(eMe,{\left.\kern-1.2pt\tau\vphantom{\big{|}}\right|_{eMe}}), where x=pqp+(1p)(1q)(1p)x=pqp+(1-p)(1-q)(1-p), and then pushing forward the measure under the inverse of cos2\cos^{2} onto (0,π/2)(0,\pi/2). The result of the computation is the following:

Proposition 5.6.

Let p,q(M,τ)p,q\in(M,\tau) be two freely independent projections with τ(p)=a,τ(q)=b\tau(p)=a,\tau(q)=b, a,b(0,1)a,b\in(0,1). Let f:f:\mathbb{R}\to\mathbb{R} be given by:

(120) f(x)=1+(4ab2(a+b))x+(ab)2x2.f(x)=1+(4ab-2(a+b))x+(a-b)^{2}x^{2}\,.

The measure ν\nu on (0,π/2)(0,\pi/2) is a probability measure with density with respect to Lebesgue measure θ\theta:

(121) dνdθ=2π1τ(e)Im(f(sec2θ))cot(θ),\frac{d\nu}{d\theta}=\frac{2}{\pi}\frac{1}{\tau(e)}\text{{Im}}\left(\sqrt{f(\sec^{2}\theta)}\right)\cot(\theta)\,,

where the square root of a negative number is on the positive imaginary axis.

Proof.

To compute ν\nu^{*}, first note that

(122) pqp\displaystyle pqp =pq+e(pqp)e\displaystyle=p\wedge q+e(pqp)e
(1p)(1q)(1p)\displaystyle(1-p)(1-q)(1-p) =(1p)(1q)+e((1p)(1q)(1p))e.\displaystyle=(1-p)\wedge(1-q)+e((1-p)(1-q)(1-p))e\,.

Taking nn-th powers of each side,

(123) (pqp)n\displaystyle(pqp)^{n} =(pq)n+(e(pqp)e)n\displaystyle=(p\wedge q)^{n}+(e(pqp)e)^{n}
((1p)(1q)(1p))n\displaystyle((1-p)(1-q)(1-p))^{n} =((1p)(1q))n+(e(1p)(1q)(1p)e)n.\displaystyle=((1-p)\wedge(1-q))^{n}+(e(1-p)(1-q)(1-p)e)^{n}.

Computing ψe(pqp)e,ψe(1p)(1q)(1p)e\psi_{e(pqp)e},\psi_{e(1-p)(1-q)(1-p)e} in (eMe,τ|eMe)(eMe,{\left.\kern-1.2pt\tau\vphantom{\big{|}}\right|_{eMe}}) and using the power series expansions of general ψμ\psi_{\mu}, then in a neighborhood of 0,

(124) ψpqp(z)\displaystyle\psi_{pqp}(z) =ψpq(z)+τ(e)ψe(pqp)e(z)\displaystyle=\psi_{p\wedge q}(z)+\tau(e)\psi_{e(pqp)e}(z)
ψ(1p)(1q)(1p)(z)\displaystyle\psi_{(1-p)(1-q)(1-p)}(z) =ψ(1p)(1q)(z)+τ(e)ψe(1p)(1q)(1p)e(z).\displaystyle=\psi_{(1-p)\wedge(1-q)}(z)+\tau(e)\psi_{e(1-p)(1-q)(1-p)e}(z)\,.

Since the spectra of pqp,e(pqp)e,(1p)(1q)(1p),e(1p)(1q)(1p)e,pq,(1p)(1q)pqp,e(pqp)e,(1-p)(1-q)(1-p),e(1-p)(1-q)(1-p)e,p\wedge q,(1-p)\wedge(1-q) are all contained in [0,1][0,1], then this equality holds on [1,)\mathbb{C}\setminus[1,\infty).

From Corollary 5.4, e0e\neq 0, so

(125) ψe(pqp)e(z)\displaystyle\psi_{e(pqp)e}(z) =ψpqp(z)ψpq(z)τ(e)\displaystyle=\frac{\psi_{pqp}(z)-\psi_{p\wedge q}(z)}{\tau(e)}
=ψpqp(z)τ(e)τ(pq)τ(e)z1z\displaystyle=\frac{\psi_{pqp}(z)}{\tau(e)}-\frac{\tau(p\wedge q)}{\tau(e)}\frac{z}{1-z}
ψe(1p)(1q)(1p)e(z)\displaystyle\psi_{e(1-p)(1-q)(1-p)e}(z) =ψ(1p)(1q)(1p)(z)ψ(1p)(1q)(z)τ(e)\displaystyle=\frac{\psi_{(1-p)(1-q)(1-p)}(z)-\psi_{(1-p)\wedge(1-q)}(z)}{\tau(e)}
=ψ(1p)(1q)(1p)(z)τ(e)τ((1p)(1q))τ(e)z1z.\displaystyle=\frac{\psi_{(1-p)(1-q)(1-p)}(z)}{\tau(e)}-\frac{\tau((1-p)\wedge(1-q))}{\tau(e)}\frac{z}{1-z}\,.

Since ee is central and x=pqp+(1p)(1q)(1p)x=pqp+(1-p)(1-q)(1-p), then for n1n\geq 1,

(126) (exe)n=(e(pqp)e)n+(e(1p)(1q)(1p)e)n.(exe)^{n}=(e(pqp)e)^{n}+(e(1-p)(1-q)(1-p)e)^{n}.

From Proposition 5.1, in (eMe,τ|eMe)(eMe,{\left.\kern-1.2pt\tau\vphantom{\big{|}}\right|_{eMe}}),

(127) ψexe(z)\displaystyle\psi_{exe}(z) =ψe(pqp)e(z)+ψe(1p)(1q)(1p)e(z)\displaystyle=\psi_{e(pqp)e}(z)+\psi_{e(1-p)(1-q)(1-p)e}(z)
=1τ(e)(ψpqp(z)+ψ(1p)(1q)(1p)(z))\displaystyle=\frac{1}{\tau(e)}\left(\psi_{pqp}(z)+\psi_{(1-p)(1-q)(1-p)}(z)\right)
τ(pq)τ(e)z1zτ((1p)(1q))τ(e)z1z\displaystyle\qquad-\frac{\tau(p\wedge q)}{\tau(e)}\frac{z}{1-z}-\frac{\tau((1-p)\wedge(1-q))}{\tau(e)}\frac{z}{1-z}
=1τ(e)(1f(z)z1)τ(pq)+τ((1p)(1q))τ(e)z1z,\displaystyle=\frac{1}{\tau(e)}\left(-1-\frac{\sqrt{f(z)}}{z-1}\right)-\frac{\tau(p\wedge q)+\tau((1-p)\wedge(1-q))}{\tau(e)}\frac{z}{1-z}\,,

where

(128) f(z)=1+(4ab2(a+b))z+(ab)2z2.f(z)=1+(4ab-2(a+b))z+(a-b)^{2}z^{2}.

From Proposition 5.1, f(z)\sqrt{f(z)} is analytic on [1,)\mathbb{C}\setminus[1,\infty), so this formula for ψexe\psi_{exe} is valid on [1,)\mathbb{C}\setminus[1,\infty).

Hence, the following formula for Gν=GexeG_{\nu^{*}}=G_{exe} is valid on [0,1]\mathbb{C}\setminus[0,1]:

(129) Gν(z)\displaystyle G_{\nu^{*}}(z) =Gexe(z)\displaystyle=G_{exe}(z)
=1z(ψexe(1z)+1)\displaystyle=\frac{1}{z}\left(\psi_{exe}\left(\frac{1}{z}\right)+1\right)
=1τ(e)f(1/z)z11τ(e)zτ(pq)+τ((1p)(1q))τ(e)1z(z1).\displaystyle=\frac{1}{\tau(e)}\frac{\sqrt{f(1/z)}}{z-1}-\frac{1}{\tau(e)z}-\frac{\tau(p\wedge q)+\tau((1-p)\wedge(1-q))}{\tau(e)}\frac{1}{z(z-1)}\,.

Recall that ν\nu^{*} is supported on [0,1][0,1]. We observe that the measure ν\nu^{*} has no atoms: For t(0,1)t\in(0,1), this follows from Lemma 5.2 and computing that limzt(zt)Gν(z)=0\lim\limits_{z\to t}(z-t)G_{\nu^{*}}(z)=0. Recall from Subsection 2.2 that ν({0,1})=0\nu^{*}(\{0,1\})=0 is true in general. Hence, ν\nu^{*} has no atoms.

Thus, we may recover the measure completely with the formula (Proposition 2.8):

(130) limy0+ab1πImGν(x+iy)dx\displaystyle\lim\limits_{y\to 0^{+}}\int_{a}^{b}-\frac{1}{\pi}\text{{Im}}\,G_{\nu^{*}}(x+iy)\,dx =ν((a,b))+12(ν({a})+ν({b}))\displaystyle=\nu^{*}((a,b))+\frac{1}{2}\left(\nu^{*}(\{a\})+\nu^{*}(\{b\})\right)
=ν([a,b]).\displaystyle=\nu^{*}([a,b])\,.

where a,b(0,1)a,b\in(0,1).

On compact subsets of (0,1)(0,1), Gν(x+iy)G_{\nu^{*}}(x+iy) is uniformly bounded as y0+y\to 0^{+}. Given that the following pointwise limit exists for x(0,1)x\in(0,1):

(131) limy0+1πImGν(x+iy),\lim\limits_{y\to 0^{+}}-\frac{1}{\pi}\text{{Im}}\,G_{\nu^{*}}(x+iy)\,,

then this limit will be the density of ν\nu^{*} with respect to the Lebesgue measure (see (21)).

Using the formula for Gν(z)G_{\nu^{*}}(z), we notice that for any x(0,1)x\in(0,1), only one term is non-zero in this limit:

(132) limy0+1πImGν(x+iy)\displaystyle\lim\limits_{y\to 0^{+}}-\frac{1}{\pi}\text{{Im}}\,G_{\nu^{*}}(x+iy) =limy0+1πIm(1τ(e)f(1/(x+iy))1(x+iy))\displaystyle=\lim\limits_{y\to 0^{+}}\frac{1}{\pi}\text{{Im}}\left(\frac{1}{\tau(e)}\frac{\sqrt{f(1/(x+iy))}}{1-(x+iy)}\right)
=1πτ(e)limy0+Imf(1/(x+iy))1x.\displaystyle=\frac{1}{\pi\tau(e)}\frac{\lim\limits_{y\to 0^{+}}\text{{Im}}\sqrt{f(1/(x+iy))}}{1-x}\,.

We proceed by showing limy0+Imf(1/(x+iy))\lim\limits_{y\to 0^{+}}\text{{Im}}\sqrt{f(1/(x+iy))} exists. It suffices to show the following limit exists:

(133) limztIm(z)<0g(z),\lim\limits_{\begin{subarray}{c}z\to t\\ \text{{Im}}(z)<0\end{subarray}}g(z)\,,

where t(0,1)t\in(0,1) and gg is defined an analytic on the lower half-plane and g(z)2=f(z)g(z)^{2}=f(z). First, note that for any sequence zntz_{n}\to t, g(zn)g(z_{n}) is bounded, because g(zn)2=f(zn)g(z_{n})^{2}=f(z_{n}). Considering a convergent subsequence where g(znk)g(z_{n_{k}}) converges, then if g(znk)sg(z_{n_{k}})\to s, then s2=limnkg(znk)2=limnkf(znk)=f(t)s^{2}=\lim\limits_{n_{k}\to\infty}g(z_{n_{k}})^{2}=\lim\limits_{n_{k}\to\infty}f(z_{n_{k}})=f(t). Hence, g(znk)g(z_{n_{k}}) converges to a square root of f(t)f(t). When f(t)=0f(t)=0 this is enough to prove the desired limit. When f(t)0f(t)\neq 0, we consider if there are two sequences in the lower half-plane, zn,zntz_{n},z_{n}^{\prime}\to t, where g(zn),g(zn)g(z_{n}),g(z_{n}^{\prime}) converge to the different square roots of f(t)f(t). Taking derivatives of g(z)2=f(z)g(z)^{2}=f(z), then 2g(z)g(z)=f(z)2g(z)g^{\prime}(z)=f^{\prime}(z). As |f(z)|\left\lvert f^{\prime}(z)\right\rvert is bounded from above and |g(z)|\left\lvert g(z)\right\rvert bounded from below near tt where f(t)0f(t)\neq 0, then |g(z)|\left\lvert g^{\prime}(z)\right\rvert is bounded from above near tt. Then, g(z)g(z) is Lipschitz near tt. Thus, g(zn),g(zn)g(z_{n}),g(z_{n}^{\prime}) converging to different square roots is a contradiction.

Since the measure is positive, then limy0+Imf(1/(x+iy))\lim\limits_{y\to 0^{+}}\text{{Im}}\sqrt{f(1/(x+iy))} has to be non-negative, at least Lebesgue almost everywhere t(0,1)t\in(0,1). It is possible to extend this to all t(0,1)t\in(0,1) by noting this set of tt is dense and using gg being Lipschitz near tt where f(t)0f(t)\neq 0.

Thus, for x(0,1)x\in(0,1),

(134) dνdλ=1πτ(e)Imf(1/x)1x,\frac{d\nu^{*}}{d\lambda}=\frac{1}{\pi\tau(e)}\frac{\text{{Im}}\sqrt{f(1/x)}}{1-x}\,,

where the square root of a negative number is on the positive imaginary axis.

By applying the change of variables formula, the measure ν\nu on (0,π/2)(0,\pi/2) is given by:

(135) dνdθ=2π1τ(e)Im(f(sec2θ))cotθ.\frac{d\nu}{d\theta}=\frac{2}{\pi}\frac{1}{\tau(e)}\text{{Im}}\left(\sqrt{f(\sec^{2}\theta)}\right)\cot\theta\,.

Note that this is a probability measure, being the spectral measure of a non-zero element of eMeeMe. ∎

6. The Brown measure of XX

By combining Propositions 4.3, 5.3, and 5.6, we can state the Brown measure of XX:

Theorem 6.1.

Let p,q(M,τ)p,q\in(M,\tau) be Hermitian, freely independent and

(136) μp\displaystyle\mu_{p} =aδα+(1a)δα\displaystyle=a\delta_{\alpha}+(1-a)\delta_{\alpha^{\prime}}
μq\displaystyle\mu_{q} =bδβ+(1b)δβ,\displaystyle=b\delta_{\beta}+(1-b)\delta_{\beta^{\prime}}\,,

where a,b(0,1)a,b\in(0,1), αα\alpha\neq\alpha^{\prime}, ββ\beta\neq\beta^{\prime}, and α,α,β,β\alpha,\alpha^{\prime},\beta,\beta^{\prime}\in\mathbb{R}.

Let

(137) ϵ00\displaystyle\epsilon_{00} =max(0,a+b1)\displaystyle=\max(0,a+b-1)
ϵ01\displaystyle\epsilon_{01} =max(0,a+(1b)1)\displaystyle=\max(0,a+(1-b)-1)
ϵ10\displaystyle\epsilon_{10} =max(0,(1a)+b1)\displaystyle=\max(0,(1-a)+b-1)
ϵ11\displaystyle\epsilon_{11} =max(0,(1a)+(1b)1)\displaystyle=\max(0,(1-a)+(1-b)-1)
ϵ\displaystyle\epsilon =1(ϵ00+ϵ01+ϵ10+ϵ11).\displaystyle=1-(\epsilon_{00}+\epsilon_{01}+\epsilon_{10}+\epsilon_{11})\,.

Then, ϵ>0\epsilon>0.

Let 𝒜=αα\mathscr{A}=\alpha^{\prime}-\alpha and =ββ\mathscr{B}=\beta^{\prime}-\beta and z\sqrt{z} denote the principal branch of the square root defined on (,0)\mathbb{C}\setminus(-\infty,0). Then,

(138) λ1(θ)={α+α2+iβ+β212𝒜22+2i𝒜cos(2θ) when |𝒜|||α+α2+iβ+β2i22𝒜22i𝒜cos(2θ) when |𝒜|<||\lambda_{1}(\theta)=\begin{dcases}\frac{\alpha+\alpha^{\prime}}{2}+i\frac{\beta+\beta^{\prime}}{2}-\frac{1}{2}\sqrt{\mathscr{A}^{2}-\mathscr{B}^{2}+2i\mathscr{A}\mathscr{B}\cos(2\theta)}&\text{ when }\left\lvert\mathscr{A}\right\rvert\geq\left\lvert\mathscr{B}\right\rvert\\ \frac{\alpha+\alpha^{\prime}}{2}+i\frac{\beta+\beta^{\prime}}{2}-\frac{i}{2}\sqrt{\mathscr{B}^{2}-\mathscr{A}^{2}-2i\mathscr{A}\mathscr{B}\cos(2\theta)}&\text{ when }\left\lvert\mathscr{A}\right\rvert<\left\lvert\mathscr{B}\right\rvert\end{dcases}
(139) λ2(θ)={α+α2+iβ+β2+12𝒜22+2i𝒜cos(2θ) when |𝒜|||α+α2+iβ+β2+i22𝒜22i𝒜cos(2θ) when |𝒜|<||.\lambda_{2}(\theta)=\begin{dcases}\frac{\alpha+\alpha^{\prime}}{2}+i\frac{\beta+\beta^{\prime}}{2}+\frac{1}{2}\sqrt{\mathscr{A}^{2}-\mathscr{B}^{2}+2i\mathscr{A}\mathscr{B}\cos(2\theta)}&\text{ when }\left\lvert\mathscr{A}\right\rvert\geq\left\lvert\mathscr{B}\right\rvert\\ \frac{\alpha+\alpha^{\prime}}{2}+i\frac{\beta+\beta^{\prime}}{2}+\frac{i}{2}\sqrt{\mathscr{B}^{2}-\mathscr{A}^{2}-2i\mathscr{A}\mathscr{B}\cos(2\theta)}&\text{ when }\left\lvert\mathscr{A}\right\rvert<\left\lvert\mathscr{B}\right\rvert\,.\end{dcases}

Let f:f:\mathbb{R}\to\mathbb{R} be given by:

(140) f(x)=1+(4ab2(a+b))x+(ab)2x2.f(x)=1+(4ab-2(a+b))x+(a-b)^{2}x^{2}\,.

Let ν\nu be a probability measure on (0,π/2)(0,\pi/2) with density with respect to Lebesgue measure θ\theta:

(141) dνdθ=2π1ϵIm(f(sec2θ))cot(θ),\frac{d\nu}{d\theta}=\frac{2}{\pi}\frac{1}{\epsilon}\text{{Im}}\left(\sqrt{f(\sec^{2}\theta)}\right)\cot(\theta)\,,

where the square root of a negative number is on the positive imaginary axis.

Let μ\mu^{\prime} be a complex probability measure given by:

(142) μ=(λ1)(ν)+(λ2)(ν)2.\mu^{\prime}=\frac{(\lambda_{1})_{*}(\nu)+(\lambda_{2})_{*}(\nu)}{2}\,.

Then, the Brown measure of X=p+iqX=p+iq is:

(143) μ=ϵ00δα+iβ+ϵ01δα+iβ+ϵ10δα+iβ+ϵ11δα+iβ+ϵμ.\mu=\epsilon_{00}\delta_{\alpha+i\beta}+\epsilon_{01}\delta_{\alpha+i\beta^{\prime}}+\epsilon_{10}\delta_{\alpha^{\prime}+i\beta}+\epsilon_{11}\delta_{\alpha^{\prime}+i\beta^{\prime}}+\epsilon\mu^{\prime}\,.
Proof.

The Theorem follows from combining several previous results:

The general form of μ\mu is given in Proposition 4.3. The τ(eij),e\tau(e_{ij}),e are relabeled as ϵij,ϵ\epsilon_{ij},\epsilon in light of Proposition 5.3 and interchanging general pp, qq for p=χ{α}(p),q=χ{β}(q)p^{\prime}=\chi_{\{\alpha^{\prime}\}}(p),q^{\prime}=\chi_{\{\beta^{\prime}\}}(q). Additionally, the fact that ϵ0\epsilon\neq 0 comes from Corollary 5.4. Finally, the measure ν\nu comes from Proposition 5.6 (after switching p,qp,q for p,qp^{\prime},q^{\prime}). ∎

We observe that the measure ν\nu is only dependent on the weights of the measures of pp and qq (i.e. aa and bb), and the “shape” of the measure (positions of the atoms and λi\lambda_{i}) is only dependent on the positions of the atoms pp and qq (i.e. α,α,β,β\alpha,\alpha^{\prime},\beta,\beta^{\prime})

We will use the definition of μ\mu^{\prime} from Theorem 6.1 for what follows:

Definition 6.2.

Let X=p+iqX=p+iq, where p,q(M,τ)p,q\in(M,\tau) are Hermitian, freely independent, and have 22 atoms. Define μ\mu^{\prime} to be the measure as in 6.1.

Now, we make some observations about the properties of the Brown measure of XX in the following Corollaries. For the figures, we approximate the Brown measure of X=p+iqX=p+iq with the empirical spectral distribution of a specific Xn=Pn+iQnX_{n}=P_{n}+iQ_{n}, where Pn,QnMn()P_{n},Q_{n}\in M_{n}(\mathbb{C}) are independently Haar-rotated Hermitian matrices with the same spectral distributions as pp and qq, respectively. The justification for this approximation will be discussed in a subsequent paper, where we will prove the convergence of the empirical spectral distributions of the XnX_{n} to the Brown measure of XX.

First, we consider the atoms of the Brown measure:

Corollary 6.3.

Let μ\mu be the Brown measure of p+iqp+iq where pp and qq have 22 atoms. Then,

  1. (1)

    μ\mu^{\prime} does not have atoms.

  2. (2)

    μ\mu is never atomic.

  3. (3)

    μ\mu can have atoms only at the points α+iβ,α+iβ,α+iβ,α+iβ\alpha+i\beta,\alpha^{\prime}+i\beta,\alpha+i\beta^{\prime},\alpha^{\prime}+i\beta^{\prime}.

  4. (4)

    μ\mu has no atoms at α+iβ\alpha+i\beta and α+iβ\alpha^{\prime}+i\beta^{\prime} if and only if a+b=1a+b=1. If a+b1a+b\neq 1, then μ\mu has exactly 11 atom at either α+iβ\alpha+i\beta or α+iβ\alpha^{\prime}+i\beta^{\prime}, with size |a+b1|\left\lvert a+b-1\right\rvert.

  5. (5)

    μ\mu has no atoms at α+iβ\alpha+i\beta^{\prime} and α+iβ\alpha^{\prime}+i\beta if and only if a=ba=b. If aba\neq b, then μ\mu has exactly 11 atom at either α+iβ\alpha+i\beta^{\prime} or α+iβ\alpha^{\prime}+i\beta, with size |ab|\left\lvert a-b\right\rvert.

  6. (6)

    μ\mu has 0, 1, or 2 atoms. μ\mu has 0 atoms if and only if a=b=1/2a=b=1/2. μ\mu has 1 atom if and only if one of a+b=1a+b=1 or a=ba=b. μ\mu has 2 atoms if and only if a+b1a+b\neq 1 and aba\neq b.

  7. (7)

    Changing a1aa\mapsto 1-a and/or b1bb\mapsto 1-b permutes the ϵij\epsilon_{ij}.

Proof.
  1. (1)

    ν\nu is absolutely continuous and λi\lambda_{i} are injective, so (λi)(ν)(\lambda_{i})_{*}(\nu) does not have atoms. Hence, μ\mu^{\prime} does not have atoms.

  2. (2)

    Since ϵ0\epsilon\neq 0, then μ0\mu^{\prime}\neq 0, so μ\mu is never atomic

  3. (3)

    Since μ\mu^{\prime} does not have atoms, the only atoms of μ\mu can occur at the points α+iβ,α+iβ,α+iβ,α+iβ\alpha+i\beta,\alpha^{\prime}+i\beta,\alpha+i\beta^{\prime},\alpha^{\prime}+i\beta^{\prime}.

  4. (4)

    Note that

    (144) ϵ00+ϵ11=max(0,(a+b1))+max(0,a+b1)=|a+b1|.\epsilon_{00}+\epsilon_{11}=\max(0,-(a+b-1))+\max(0,a+b-1)=\left\lvert a+b-1\right\rvert.

    Hence, when a+b=1a+b=1, ϵ00+ϵ11=0\epsilon_{00}+\epsilon_{11}=0 so μ\mu has no atoms at α+iβ\alpha+i\beta and α+iβ\alpha^{\prime}+i\beta^{\prime}. If a+b1a+b\neq 1, then only one of a+b1a+b-1 and (a+b1)-(a+b-1) is positive and equal to |a+b1|\left\lvert a+b-1\right\rvert and hence one of ϵ00,ϵ11\epsilon_{00},\epsilon_{11} is equal to |a+b1|\left\lvert a+b-1\right\rvert.

  5. (5)

    Follows similarly to 4, with the equation

    (145) ϵ01+ϵ10=max(0,ba)+max(0,ab)=|ab|.\epsilon_{01}+\epsilon_{10}=\max(0,b-a)+\max(0,a-b)=\left\lvert a-b\right\rvert.
  6. (6)

    Directly follows from 4. and 5.

  7. (7)

    Follows directly from the formulas for the ϵij\epsilon_{ij}.

Now, we consider the symmetries of μ\mu^{\prime}:

Corollary 6.4.

Let μ\mu be the Brown measure of p+iqp+iq where pp and qq have 22 atoms

  1. (1)

    Swapping the roles of pp and qq fixes ν\nu.

  2. (2)

    Changing one of a1aa\mapsto 1-a or b1bb\mapsto 1-b changes ν\nu by changing θπ/2θ\theta\to\pi/2-\theta. These correspond changing pα+αpp\mapsto\alpha+\alpha^{\prime}-p and qβ+βqq\mapsto\beta+\beta^{\prime}-q, respectively.

  3. (3)

    Changing both a1aa\mapsto 1-a and b1bb\mapsto 1-b fixes ν\nu. This corresponds to changing both pα+αpp\mapsto\alpha+\alpha^{\prime}-p and qβ+βqq\mapsto\beta+\beta^{\prime}-q. This amounts to changing p+iq(α+α)+i(β+β)(p+iq)p+iq\mapsto(\alpha+\alpha^{\prime})+i(\beta+\beta^{\prime})-(p+iq).

Proof.
  1. (1)

    Swapping the roles of pp and qq amounts to swapping aa and bb in the formula for f(x)f(x). But, the formula for f(x)f(x) is symmetric with respect to a,ba,b, so ν\nu is fixed.

  2. (2)

    Since ff is symmetric with respect to changing aa and bb, it suffices to check just a1aa\mapsto 1-a. Denote ff by fa,bf_{a,b} to refer to the coefficients. For this, we just need to check the identity:

    (146) f1a,b(sec2θ)cotθ=fa,b(csc2θ)tanθ\sqrt{f_{1-a,b}(\sec^{2}\theta)}\cot\theta=\sqrt{f_{a,b}(\csc^{2}\theta)}\tan\theta

    for θ(0,π/2)\theta\in(0,\pi/2). As cotθ,tanθ>0\cot\theta,\tan\theta>0 for θ(0,π/2)\theta\in(0,\pi/2), it is equivalent to check

    (147) f1a,b(sec2θ)cot2θ=fa,b(csc2θ)tan2θ.f_{1-a,b}(\sec^{2}\theta)\cot^{2}\theta=f_{a,b}(\csc^{2}\theta)\tan^{2}\theta\,.

    Checking this identity is a straightforward calculation.

  3. (3)

    Follows from applying 2. twice.

Figure 2 illustrates the behavior in Corollary 6.4 where the Brown measure of X=p+iqX=p+iq is approximated by the ESD of Xn=Pn+iQnX_{n}=P_{n}+iQ_{n} for deterministic Pn,QnMn()P_{n},Q_{n}\in M_{n}(\mathbb{C}). Note the symmetry between the two ESDs.

Refer to caption
(a) ESD of Xn=Pn+iQnX_{n}=P_{n}+iQ_{n}
μPn=(4/5)δ0+(1/5)δ9/10\mu_{P_{n}}=(4/5)\delta_{0}+(1/5)\delta_{9/10}
μQn=(1/5)δ0+(4/5)δ1\mu_{Q_{n}}=(1/5)\delta_{0}+(4/5)\delta_{1}
n=1000n=1000
Refer to caption
(b) ESD of Xn=Pn+iQnX_{n}=P_{n}+iQ_{n}
μPn=(1/5)δ0+(4/5)δ9/10\mu_{P_{n}}=(1/5)\delta_{0}+(4/5)\delta_{9/10}
μQn=(1/5)δ0+(4/5)δ1\mu_{Q_{n}}=(1/5)\delta_{0}+(4/5)\delta_{1}
n=1000n=1000
Figure 2. ESDs for Corollary 6.4

Finally, we consider when the density of μ\mu^{\prime} extends all the way to the 4 corners of the intersection of the hyperbola with the boundary of the rectangle:

Corollary 6.5.

Let μ\mu be the Brown measure of p+iqp+iq where pp and qq have 22 atoms

  1. (1)

    μ\mu^{\prime} has density extending to α+iβ\alpha^{\prime}+i\beta and α+iβ\alpha+i\beta^{\prime} if and only if a=ba=b.

  2. (2)

    μ\mu^{\prime} has density extending to α+iβ\alpha+i\beta and α+iβ\alpha^{\prime}+i\beta^{\prime} if and only if a=1ba=1-b.

  3. (3)

    μ\mu^{\prime} has density extending to all 44 corners of the intersection of the hyperbola with the boundary of the rectangle if and only if a=b=1/2a=b=1/2. Hence, the support of μ\mu is equal to HRH\cap R if and only if a=b=1/2a=b=1/2.

Proof.

Let 𝒜=αα\mathscr{A}=\alpha^{\prime}-\alpha and =ββ\mathscr{B}=\beta^{\prime}-\beta.

  1. (1)

    For z=α+iβ,α+iβz=\alpha^{\prime}+i\beta,\alpha+i\beta^{\prime},

    (148) (zα+α2iβ+β2)2=𝒜222i𝒜4.\left(z-\frac{\alpha+\alpha^{\prime}}{2}-i\frac{\beta+\beta^{\prime}}{2}\right)^{2}=\frac{\mathscr{A}^{2}-\mathscr{B}^{2}-2i\mathscr{A}\mathscr{B}}{4}\,.

    Hence, z=λi(π/2)z=\lambda_{i}(\pi/2).

    Thus, it is equivalent to determine for which a,ba,b the measure ν\nu has density approaching π/2\pi/2. Note that ν\nu has density approaching π/2\pi/2 if and only if f(sec2θ)<0f(\sec^{2}\theta)<0 for θπ/2\theta\to\pi/2^{-}. Note that limθπ/2sec2θ=\lim\limits_{\theta\to\pi/2^{-}}\sec^{2}\theta=\infty. Recall that f(z)f(z) is a polynomial of degree at most 22. If ff is quadratic, then limzf(z)=\lim\limits_{z\to\infty}f(z)=\infty, so then ν\nu does not have density approaching π/2\pi/2. Hence, we must have a=ba=b. Conversely, if a=ba=b, then f(z)=1+4(a2a)z=1+4(a1)azf(z)=1+4(a^{2}-a)z=1+4(a-1)az, and the linear term has negative coefficient for a(0,1)a\in(0,1). Thus, f(z)<0f(z)<0 as zz\to\infty and ν\nu has density approaching π/2\pi/2.

  2. (2)

    For z=α+iβ,α+iβz=\alpha+i\beta,\alpha^{\prime}+i\beta^{\prime},

    (149) (zα+α2iβ+β2)2=𝒜22+2i𝒜4.\left(z-\frac{\alpha+\alpha^{\prime}}{2}-i\frac{\beta+\beta^{\prime}}{2}\right)^{2}=\frac{\mathscr{A}^{2}-\mathscr{B}^{2}+2i\mathscr{A}\mathscr{B}}{4}\,.

    Hence, z=λi(0)z=\lambda_{i}(0).

    Thus, it is equivalent to determine for which a,ba,b the measure ν\nu has density approaching 0. Note that ν\nu has density approaching 0 if and only if f(sec2θ)<0f(\sec^{2}\theta)<0 for θ0+\theta\to 0^{+}. Since limθ0+sec2θ=1\lim\limits_{\theta\to 0^{+}}\sec^{2}\theta=1, then we need that f(x)<0f(x)<0 for x1+x\to 1^{+}. Since f(1)=(a+b1)2f(1)=(a+b-1)^{2}, then f(1)=0f(1)=0, i.e. a=1ba=1-b. Conversely, if a=1ba=1-b, then f(1)=0f(1)=0 and since ff is either quadratic with positive leading coefficient or ff is linear with negative slope, then ff must be negative to the right of 11. Note that in the case where ff is quadratic, ff cannot have a double root at 11 (from Proposition 5.1).

  3. (3)

    Follows from 1. and 2.

Figure 3 illustrates the behavior in Corollary 6.5 where the Brown measure of X=p+iqX=p+iq is approximated by the ESD of Xn=Pn+iQnX_{n}=P_{n}+iQ_{n}, for deterministic Pn,QnMn()P_{n},Q_{n}\in M_{n}(\mathbb{C}). The left ESD does not have density approaching the corners of RR, but the right ESD does.

Refer to caption
(a) ESD of Xn=Pn+iQnX_{n}=P_{n}+iQ_{n}
μPn=(9/10)δ0+(1/10)δ4/5\mu_{P_{n}}=(9/10)\delta_{0}+(1/10)\delta_{4/5}
μQn=(1/2)δ0+(1/2)δ1\mu_{Q_{n}}=(1/2)\delta_{0}+(1/2)\delta_{1}
n=1000n=1000
Refer to caption
(b) ESD of Xn=Pn+iQnX_{n}=P_{n}+iQ_{n}
μPn=(4/5)δ0+(1/5)δ9/10\mu_{P_{n}}=(4/5)\delta_{0}+(1/5)\delta_{9/10}
μQn=(1/5)δ0+(4/5)δ1\mu_{Q_{n}}=(1/5)\delta_{0}+(4/5)\delta_{1}
n=1000n=1000
Figure 3. ESDs for Corollary 6.5

Finally, we conclude that the Brown measure of X=p+iqX=p+iq uniquely determines the laws of pp and qq. Here we allow pp and qq to be possibly constant.

Corollary 6.6.

Let p,q(M,τ)p,q\in(M,\tau) be Hermitian and freely independent and

(150) μp\displaystyle\mu_{p} =aδα+(1a)δα\displaystyle=a\delta_{\alpha}+(1-a)\delta_{\alpha^{\prime}}
μq\displaystyle\mu_{q} =bδβ+(1b)δβ.\displaystyle=b\delta_{\beta}+(1-b)\delta_{\beta^{\prime}}\,.

where a,b[0,1]a,b\in[0,1] and α,α,β,β\alpha,\alpha^{\prime},\beta,\beta^{\prime}\in\mathbb{R}. Let μ\mu be the Brown measure of X=p+iqX=p+iq. Then, the assignment (μp,μq)μ(\mu_{p},\mu_{q})\mapsto\mu is 1 to 1.

Proof.

From Corollary 6.3, μ\mu is atomic if and only if one of pp or qq is constant. In this case, it is easy to determine the weights and atoms of both μp\mu_{p} and μq\mu_{q} from μ\mu.

Thus, we may consider μ\mu which is the Brown measure of X=p+iqX=p+iq where pp and qq are not constant.

First, we show that μ\mu determines the positions of the atoms of pp and qq. Since the support of μ\mu^{\prime} on HRH\cap R contains at least 55 points, then the equation of HH is uniquely determined. If a=ba=b or a+b=1a+b=1, then from Corollary 6.5, we can determine the positions of the atoms of pp and qq. Thus, assume that aba\neq b and a+b1a+b\neq 1. From Corollary 6.3, μ\mu has 22 atoms at the points {α+iβ,α+iβ,α+iβ,α+iβ}\{\alpha+i\beta,\alpha^{\prime}+i\beta,\alpha+i\beta^{\prime},\alpha^{\prime}+i\beta^{\prime}\}. From these two points, at least 33 of α,α,β,β\alpha,\alpha^{\prime},\beta,\beta^{\prime} are determined. To determine the last one, we can use the equation of the hyperbola and look at either the coefficient of xx or yy. Thus, μ\mu determines the positions of the atoms of pp and qq.

We can determine the weights of the atoms of pp and qq directly from Proposition 4.7. ∎

References

  • [1] Greg W. Anderson, Alice Guionnet and Ofer Zeitouni “An introduction to random matrices”, Cambridge Studies in Advanced Mathematics Cambridge University Press, Cambridge, 2010
  • [2] Serban T. Belinschi, Piotr Śniady and Roland Speicher “Eigenvalues of non-Hermitian random matrices and Brown measure of non-normal operators: Hermitian reduction and linearization method” In Linear Algebra and its Applications 537, 2018, pp. 48–83
  • [3] L.. Brown “Lidskii’s theorem in the type II case” In Geometric methods in operator algebras (Kyoto, 1983) 123, Pitman Res. Notes Math. Ser. Longman Sci. Tech., Harlow, 1986, pp. 1–35
  • [4] Bent Fuglede and Richard V. Kadison “Determinant theory in finite factors” In Annals of Mathematics. Second Series 55, 1952, pp. 520–530
  • [5] Alice Guionnet, Manjunath Krishnapur and Ofer Zeitouni “The single ring theorem” In Annals of Mathematics. Second Series 174.2, 2011, pp. 1189–1217
  • [6] Uffe Haagerup and Flemming Larsen “Brown’s spectral distribution measure for RR-diagonal elements in finite von Neumann algebras” In Journal of Functional Analysis 176.2, 2000
  • [7] Uffe Haagerup and Hanne Schultz “Brown measures of unbounded operators affiliated with a finite von Neumann algebra” In Mathematica Scandinavica 100.2, 2007, pp. 209–263
  • [8] Ching-Wei Ho “The Brown measure of the sum of a self-adjoint element and an elliptic element” In Electron. J. Probab. 27, 2022, pp. Paper No. 123\bibrangessep32
  • [9] Ching-Wei Ho and Ping Zhong “Brown measures of free circular and multiplicative Brownian motions with self-adjoint and unitary initial conditions” In J. Eur. Math. Soc. (JEMS) 25.6, 2023, pp. 2163–2227
  • [10] James A. Mingo and Roland Speicher “Free probability and random matrices” 35, Fields Institute Monographs Springer, New York; Fields Institute for Research in Mathematical Sciences, Toronto, ON, 2017
  • [11] Piotr Śniady “Random Regularization of Brown Spectral Measure” In Journal of Functional Analysis 193.2, 2002, pp. 291–313
  • [12] Terence Tao “Topics in random matrix theory”, Graduate Studies in Mathematics American Mathematical Society, Providence, RI, 2012
  • [13] Terence Tao and Van Vu “Random matrices: universality of ESDs and the circular law” With an appendix by Manjunath Krishnapur In The Annals of Probability 38.5, 2010, pp. 2023–2065
  • [14] Dan Voiculescu “The analogues of entropy and of Fisher’s information measure in free probability theory. VI. Liberation and mutual free information” In Advances in Mathematics 146.2, 1999, pp. 101–166