The Brown measure of non-Hermitian sums of projections
Abstract.
We compute the Brown measure of the non-normal operators , where and are Hermitian, freely independent, and have spectra consisting of atoms. The computation relies on the model of the non-trivial part of the von Neumann algebra generated by 2 projections as 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 be a von Neumann algebra with a faithful, normal, tracial state . Let . The Brown measure of , introduced in [3], is a complex Borel probability measure supported on the spectrum of . When is normal, the Brown measure of is the spectral measure. When is a random matrix, the Brown measure of is the empirical spectral distribution of .
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 “-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 , where are Hermitian, freely independent, and have spectra consisting of atoms. Note that when or only has atom in their spectra (i.e. is a real number), then is normal, so the Brown measure can be trivially computed. In the course of studying these operators, we will deduce that when and have atoms, then is non-normal.
The key tool is the model of the von Neumann algebra generated by projections from [14] to study the non-trivial part of this space. This reduces the computation in to for some Borel probability measure on . To completely determine the Brown measure, we use freeness of and to explicitly compute 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 , suppose that converges in -moments to . The Brown measure of is typically the limit of the empirical spectral distributions of the (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 corresponding to and prove that the empirical spectral distributions of the converge to the Brown measure of .
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 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 up to some parameters that depend on the joint law of and (Proposition 4.3). In Section 5, we use the freeness of and to compute these parameters (Propositions 5.3 and 5.6) and also deduce that is non-normal (Corollary 5.5). In Section 6, we combine these results to state the Brown measure of (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 and describe how to compute it.
First, we recall the definition of the Fuglede-Kadison determinant, first introduced in [4]:
Definition 2.1.
Let . Let be the spectral measure of . Then, the Fuglede-Kadison determinant of , , is given by:
(1) |
When , then . 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 converging to , we have:
(2) |
In practice, it is more convenient to compute the right-hand side to compute .
Then, the Brown measure is the following distribution:
Definition 2.2.
Let . The Brown measure of is defined as:
(3) |
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 is a probability measure supported on the spectrum of . When is normal, then the Brown measure is the spectral measure. When is a random matrix, then the Brown measure is the empirical spectral distribution.
2.2. The von Neumann algebra generated by projections
In this subsection, we recall some results from [[]Section 12]VoiculescuPaper characterizing the tracial von Neumann algebra generated by projections. Let be the von Neumann algebra generated by two projections and .
Let be the von Neumann algebra generated by two projections and .
Consider the following elements of :
(4) | ||||
These elements are central, mutually orthogonal projections and
(5) |
If we consider the matrix of with respect to the (i.e. the matrix with entries of the form where ), then this matrix is diagonal. Further, the 4 diagonal terms are constants. Hence, the only interesting part of is the subalgebra . In particular,
(6) | ||||
When , consider as a von Neumann subalgebra with identity and . Then,
(7) |
where is a Borel measure on .
For any , let . The isomorphism has the following correspondence between and matrix-valued functions of :
(8) | ||||
The measure is recovered by noting that for the central element ,
(9) |
Hence, is the spectral measure of the element in .
Observe that with the change of variables , ,
(10) |
where is the rotation matrix with angle .
In practice, we will work with instead of . Let be the inverse of and consider the measure on given by:
(11) |
Then, we will use the isomorphism:
(12) |
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 does not have atoms at or , i.e. . This is equivalent to not having atoms at or .
Proof.
Let and . Since commutes with and , and .
Consider as . Recall that and are mutually orthogonal, so . Under the isomorphism, computation shows that
(13) |
As the integral on the right-hand side decreases to . Hence, .
A similar argument considering using that and are mutually orthogonal shows that .
Finally, we note that from the definition of pushforward measure, . ∎
Next, we note that on , and have trace . This follows from being mutually orthogonal to all of the :
Lemma 2.4.
Let be the trace on . Then, .
Proof.
If this is false, then we may choose one of or and one of or so that the sum of the traces is greater than . Without loss of generality assume that . Since commutes with and , and . Then, from the parallelogram law, we have the following contradiction:
(14) | ||||
∎
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 :
Definition 2.5.
Let be a tracial von Neumann algebra, and let be a family of unital subalgebras of . are freely independent if for any with , and , then
(15) |
Let be positive integers. The sets of non-commutative random variables are free if the algebras they generate are free.
If generate and are freely independent, then determines .
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 , we will abuse notation by using the subscript to denote the integral transform with respect to the spectral measure .
First, we define the Stieltjes transform of a real probability measure:
Definition 2.6.
Let be a probability measure on . Then, the Stieltjes transform of is the function given by:
(16) |
We list some of the well-known facts about in the following Proposition:
Proposition 2.7.
Let be the Stieltjes transform of . Then,
-
•
is analytic on .
-
•
Let be the upper/lower half-planes of . Then, . In particular, if and only if .
-
•
for .
-
•
.
-
•
is analytic on .
-
•
is compactly supported, has the following Laurent series expansion for :
(17)
Thus, for compact measures, the Stieltjes transform of a measure contains the same information as the moments of .
We highlight one special fact about the Stieltjes transform. For , note that
(18) |
Recall that the Poisson kernel on the upper half-plane is given by:
(19) |
Combining these two facts shows that
(20) |
Recall that are approximations to the identity as , so then
(21) |
in the vague topology on .
By exploiting the conjugate symmetry of , we can also write this in terms of the discontinuity of across :
(22) |
in the vague topology on .
Additionally, there is an explicit formula for intervals ([10], Theorem 6):
Proposition 2.8.
For and ,
(23) |
When is a compactly supported real measure, then
(24) |
Computation using shows that is analytic in a neighborhood of , as the is exactly the pole of at .
For non-compactly supported measures , then can still be defined on wedge-shaped domains containing [[, see]Theorem 33]SpeicherBook, but we will not need this fact.
Similar to the Stieltjes transform, we can define :
Definition 2.9.
Let be a probability measure on . Let . Define by:
(25) |
We list some well-known properties of in the following Proposition:
Proposition 2.10.
Let be as above. Then,
-
•
.
-
•
is analytic on .
-
•
for all .
-
•
Let be the upper/lower half-planes of . If , then . In particular, if and only if .
-
•
If is compactly supported, then has the following Taylor series expansion for :
(26)
We highlight the following equation, valid for :
(27) |
To consider the inverse of at , note that
(28) |
As long as , then is invertible in a neighborhood of . In this situation, define the following functions:
Definition 2.11.
Let be a compactly supported probability measure on such that . Then, is invertible in a neighborhood of in . Define
(29) | ||||
is called the -transform of .
In particular, since then is well-defined in a neighborhood of .
Finally, we recall the addition (resp. multiplication) laws for free additive (resp. free multiplicative) convolution:
Theorem 2.12.
Let be Hermitian and freely independent. Let be the spectral measure of . Then, where the functions are defined,
(30) |
If are positive, let be the spectral measure of . Then, where the functions are defined,
(31) |
3. Notation and Outline of Computation
Recall that we consider operators of the form , where and are Hermitian, freely independent, and have atoms in their spectra.
We will use the following notation for the atoms of and and their weights:
(32) | ||||
where , , , and .
We fix (resp. ) to be the following spectral projection of (resp. ):
Definition 3.1.
Let be Hermitian with laws:
(33) | ||||
Let be the following projections:
(34) | ||||
As a consequence,
(35) | |||
Hence,
(36) | ||||
or equivalently
(37) | ||||
Recall the Brown measure of is defined as:
(38) |
where is the spectral measure of .
To compute the Brown measure, we need to complete the following steps:
-
(1)
Compute .
-
(2)
Compute .
-
(3)
Compute .
In Section 4, we will compute the Brown measure of up to some parameters that come from the von Neumann algebra generated by projections in Subsection 2.2. In Section 5, we use freeness of and to compute these parameters and also deduce that is non-normal. In Section 6, we combine these results to state the Brown measure of 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 as a convex combination of 4 atoms and another probability measure, . Applying the model of the von Neumann algebra generated by projections with the projections and defined in the previous section, observe that:
(39) | ||||
The measure depends on the measure in the isomorphism from the von Neumann algebra generated by projections: . The weights in the convex combination are the weights from Subsection 2.2. We will determine the Brown measure up to determining the weights and the measure . These parameters depend on the joint law of and . In particular, the results in this section are valid for arbitrary projections . In Section 5, we will compute these parameters when and are freely independent.
To compute the Brown measure of , recall we need to first compute , the spectral measure of .
The result of the computation is the following:
Proposition 4.1.
If ,
(40) |
If , there exists continuous functions such that are the singular values of the element of corresponding to .
Then,
(41) | ||||
Proof.
Let . Since the are central, mutually orthogonal projections that sum to , then
(42) | ||||
Taking traces of both sides,
(43) | ||||
From the definitions of the ,
(44) | ||||
Combining the previous two equations,
(45) | ||||
The right-hand side is the -th moment of the following convex combination of probability measures:
(46) | ||||
where is the spectral measure of in .
As moments determine compact measures on , then this is .
If , then is the desired convex combination of atoms.
If , consider using the isomorphism described in Section LABEL:sec:two_projections, i.e. we proceed by assuming that .
For any , let . Then, .
Now, we claim that it is possible to choose continuous functions that are the singular values of for . First, note that from the expressions for in Subsection 2.2 that is continuous in . Hence, the characteristic polynomial of is a monic polynomial with coefficients that are continuous in . From the quadratic formula, the eigenvalues of have expressions in terms of the coefficients of the characteristic polynomial of . As 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 . Thus, we may choose continuous expressions for the two eigenvalues of . Taking square roots of these two non-negative continuous functions produces .
Thus,
(47) | ||||
The right-hand side is the -th moment of the following measure:
(48) |
Since moments determine compact measures on , then
(49) |
By combining this with (46), we get the desired result:
(50) | ||||
∎
As an intermediate step in computing the Brown measure of , we compute :
Proposition 4.2.
If ,
(51) | ||||
If , there exists continuous functions such that are the eigenvalues of the element of corresponding to .
Let and and denote the principal branch of the square root defined on . Then,
(52) |
(53) |
Then,
(54) | ||||
Proof.
When , the formula for follows easily from Proposition 4.1.
For , from Proposition 4.1, for any continuous ,
(55) | ||||
Rewriting the final integral using the change of variables formula,
(56) | ||||
Let for . Then, are continuous and decrease to . Applying two previous equations for and using the monotone convergence theorem to take the limit as ,
(57) | ||||
For the rest of the proof, assume that .
Recall that are the singular values of . Then, the integrand of the final integral can be simplified as:
(58) | ||||
Given that we can verify the formulas for , , then are the eigenvalues for . Thus, combining the previous two equations,
(59) | ||||
Substituting this expression into (57) produces the desired formula for .
We return to verifying the formulas for , . Straightforward computation shows that the characteristic polynomial of is:
(60) |
The eigenvalues of are:
(61) |
where the square root is any branch of the square root.
When , then and the expression for the eigenvalues of 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 for .
When , then are also expressions for the square roots of , where the square root is the principal branch of the square root. Since , then and this expression is continuous and well-defined.. This gives the formulas for for . ∎
Finally, we compute the Brown measure of , in the following Proposition:
Proposition 4.3.
Let be the Brown measure of . If ,
(62) | ||||
If , let be as in Proposition 4.2. Then,
(63) | ||||
where
(64) |
Additionally, .
Proof.
If , the result follows from directly applying to the expression for in Proposition 4.2.
If , we take the distributional Laplacian of the result from Proposition 4.2:
(65) | ||||
As in the case when , applying directly to the first atomic terms of produces the weighted sum of the atoms in .
To apply for the final integral, we apply Fubini’s theorem. Consider . Since for all , then we may apply Fubini’s theorem for :
(66) | ||||
Hence, for ,
(67) |
Thus, the Laplacian of the integral is:
(68) |
Combining this with the atomic terms gives the desired Brown measure for .
Even though the weights and the measure 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 , where and . Let and .
Let
(70) | ||||
The equation of is equivalent to:
(71) |
The equation of in coordinates
(72) | ||||
is
(73) |
It follows that for ,
(74) |
Similarly,
(75) |
Alternatively, the equation of the hyperbola is:
(76) |
If , then if and only if
(77) |
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 coordinates
(78) |
(79) | |||||
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) |
In light of what was previously shown,
(81) |
Conversely,
(82) |
∎
Now, we will show that the support of the Brown measure is contained in :
Corollary 4.5.
Let and .
The continuous functions in Proposition 4.2 parameterize the intersection of the hyperbola
(83) |
with the rectangle
(84) |
When , parameterizes the left component of and parameterizes the right component of . When , parameterizes the bottom component of and parameterizes the top component of .
The support of the Brown measure is contained in .
Proof.
From Lemma 4.4, is on if and only if
(85) |
Note that are exactly the solutions to this equation. Hence, the parameterize the intersection of the hyperbola and rectangle.
For the cases of parameterizing the left/right or top/bottom components, it is easy to see from the formulas that the map into the left/right or top/bottom components, and since the parameterize all of , then the have to parameterize the entire left/right or top/bottom components.
As the parameterize , then is supported on . The 4 atoms in the Brown measure are on (at the 4 corners of the rectangle).
Thus, we conclude that the Brown measure is supported on . ∎
Figure 1 illustrates the conclusion of Corollary 4.5 where the Brown measure of is approximated by the ESD of for deterministic .

Motivated by the hyperbola and rectangle appearing in Corollary 4.5, we introduce the definition of the hyperbola and rectangle associated with :
Definition 4.6.
Let be Hermitian with laws:
(86) | ||||
Let and .
The hyperbola associated with is
(87) |
The rectangle associated with is
(88) |
While the weights are as of yet undetermined in the case and are free, there are some general relationships between the weights and traces of the spectral projections of and :
Proposition 4.7.
Let be the Brown measure of and let be as in Proposition 4.3.
Then, the measure of each of the two components of is equal to . Additionally,
(89) | ||||
If , the Brown measure of the left component of is and the Brown measure of the right component of is .
If , the Brown measure of the bottom component of is and the Brown measure of the top component of is .
Proof.
Recall that
(90) |
where is a probability measure on . From Corollary 4.5, the each parameterize one of the components of . Hence, the measure of each component of is .
For the equations of the traces of spectral projections of and , we will just prove the first equation, the others are similar. From (39),
(91) |
From Proposition 4.3, and . From Lemma 2.4, . 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 when as the other cases are similar. Let be this left component. Then,
(92) | ||||
∎
5. Computation of Weights and Measure
In this section, we will determine the weights and the measure in Proposition 4.3 under the assumption that and are freely independent. We also deduce that the operators of the form , where Hermitian, freely independent, and have atoms in their spectra are not normal (Corollary 5.5).
We will use the functions , , and that were introduced in Subsection 2.3. Recall that for , we will use , , and to denote the respective functions with respect to , the spectral measure of .
Recall that the measure and the weights are computed in terms of projections where
(93) | ||||
Since and are freely independent, then and are also freely independent. This along with the traces and determine the joint law of , .
For notational convenience, in this section we will consider two general projections and that are freely independent and , for some . This is a natural continuation of Subsection 2.2. It is easy to translate the results in this section to the defined in Section 3.
Recall from Subsection 2.2 that the weights are:
(94) | ||||
Recall that when , is the pushforward measure of under the inverse of of and is the spectral measure of , where .
Since as , then understanding the laws of and are relevant for computing both and the weights . The first step to this is computing the relevant free probability functions:
Proposition 5.1.
Let be two freely independent projections with , , . Then,
(95) | ||||
(96) |
Let
(97) |
Then, are analytic on and
(98) | ||||
where is an analytic branch of the square root of on where . In particular, the root(s) of are in and are distinct when is quadratic.
Proof.
The formula for follows from
(99) |
and similarly for . Since , are linear fractional transformations, the formulas for their inverses (, , respectively) are easily computed.
Since and are freely independent, then the formula for follows from the multiplicativity of the -transform and its relationship with .
Since then is supported on . So, the formula
(100) |
defines an analytic function for .
To compute the formula for , recall that in a neighborhood of . Then,
(101) |
if and only if
(102) |
This is clearly true for and the latter equation is not satisfied for at any .
This last equation is true if and only if
(103) |
Fixing a and solving for , then from the quadratic formula and simplifying, we obtain the desired formula for . The sign of the square root follows from the general fact that .
The formula for follows from the formula for and noting that and are freely independent projections with traces , . Observe that is invariant under changing the pair to .
Finally, note that the in both and are identical since they are defined on the domain and agree at .
The fact that has root(s) in follows from the fact that is not analytic in any neighborhood of , so the root(s) of cannot be on .
For distinctness of the roots when is quadratic, the discriminant is for . ∎
Now, we proceed to determine . First, we need the following Lemma [10, Proposition 8]:
Lemma 5.2.
Let be a finite measure on the real line and . For a sequence non-tangentially to , .
Proof.
Let . The condition that non-tangentially to is equivalent to for some . In particular, .
Computation shows that
(104) |
Let where
(105) |
Note that for all and for all . It suffices to show that the are uniformly bounded, as then the result follows from the bounded convergence theorem.
First, rewrite :
(106) |
Since and ,
(107) |
as desired. ∎
Next, we determine using the free independence of and :
Proposition 5.3.
Let be two freely independent projections with , , . Then,
(108) | ||||
Proof.
Proposition 5.3 can be summarized by: “free projections intersect as little as possible.” For , the term is just the minimum trace of the intersection between two projections and where of and :
Recall from the parallelogram law that for projections ,
(115) |
As , then for any projections , .
As a corollary, we will no longer need to consider the possibility that :
Corollary 5.4.
Let be two freely independent projections. Then, if and only if one of is .
Proof.
If one of is , then it is clear that .
Suppose that none of is . Then, we may apply Proposition 5.3. As is faithful, it suffices to conclude that .
Consider the expressions of and in Proposition 5.3. Observe that the expressions inside and are negatives of each other and hence
(116) |
Similarly,
(117) |
Thus,
(118) |
By considering the four cases of and , we see that if and only if one of is . Hence, . ∎
As a corollary, we deduce the fact that the operators of the form are not normal:
Corollary 5.5.
Let be Hermitian, freely independent, and have atoms in their spectra. Then, is not normal.
Proof.
First, we consider the case when and are projections with , , .
It suffices to show that and do not commute. For this, recall that and commute if and only if . Applying this to the other three pairs of commuting projections, , , and , then , , and .
Thus if and commute, the following equalities hold:
(119) | ||||
In the notation of the projections , this means that . From Corollary 5.4, this cannot happen if . Thus, is not normal.
The general case follows by considering the projections and defined in Section 3. Then, is normal if and only if is normal, since and commute if and only if and commute. From the previous case, is not normal. Hence, is not normal. ∎
We proceed to compute with the knowledge that . This requires computing , the spectral measure of in , where , and then pushing forward the measure under the inverse of onto . The result of the computation is the following:
Proposition 5.6.
Let be two freely independent projections with , . Let be given by:
(120) |
The measure on is a probability measure with density with respect to Lebesgue measure :
(121) |
where the square root of a negative number is on the positive imaginary axis.
Proof.
To compute , first note that
(122) | ||||
Taking -th powers of each side,
(123) | ||||
Computing in and using the power series expansions of general , then in a neighborhood of ,
(124) | ||||
Since the spectra of are all contained in , then this equality holds on .
From Corollary 5.4, , so
(125) | ||||
Since is central and , then for ,
(126) |
From Proposition 5.1, in ,
(127) | ||||
where
(128) |
From Proposition 5.1, is analytic on , so this formula for is valid on .
Hence, the following formula for is valid on :
(129) | ||||
Recall that is supported on . We observe that the measure has no atoms: For , this follows from Lemma 5.2 and computing that . Recall from Subsection 2.2 that is true in general. Hence, has no atoms.
On compact subsets of , is uniformly bounded as . Given that the following pointwise limit exists for :
(131) |
then this limit will be the density of with respect to the Lebesgue measure (see (21)).
Using the formula for , we notice that for any , only one term is non-zero in this limit:
(132) | ||||
We proceed by showing exists. It suffices to show the following limit exists:
(133) |
where and is defined an analytic on the lower half-plane and . First, note that for any sequence , is bounded, because . Considering a convergent subsequence where converges, then if , then . Hence, converges to a square root of . When this is enough to prove the desired limit. When , we consider if there are two sequences in the lower half-plane, , where converge to the different square roots of . Taking derivatives of , then . As is bounded from above and bounded from below near where , then is bounded from above near . Then, is Lipschitz near . Thus, converging to different square roots is a contradiction.
Since the measure is positive, then has to be non-negative, at least Lebesgue almost everywhere . It is possible to extend this to all by noting this set of is dense and using being Lipschitz near where .
Thus, for ,
(134) |
where the square root of a negative number is on the positive imaginary axis.
By applying the change of variables formula, the measure on is given by:
(135) |
Note that this is a probability measure, being the spectral measure of a non-zero element of . ∎
6. The Brown measure of
Theorem 6.1.
Let be Hermitian, freely independent and
(136) | ||||
where , , , and .
Let
(137) | ||||
Then, .
Let and and denote the principal branch of the square root defined on . Then,
(138) |
(139) |
Let be given by:
(140) |
Let be a probability measure on with density with respect to Lebesgue measure :
(141) |
where the square root of a negative number is on the positive imaginary axis.
Let be a complex probability measure given by:
(142) |
Then, the Brown measure of is:
(143) |
Proof.
The Theorem follows from combining several previous results:
We observe that the measure is only dependent on the weights of the measures of and (i.e. and ), and the “shape” of the measure (positions of the atoms and ) is only dependent on the positions of the atoms and (i.e. )
We will use the definition of from Theorem 6.1 for what follows:
Definition 6.2.
Let , where are Hermitian, freely independent, and have atoms. Define to be the measure as in 6.1.
Now, we make some observations about the properties of the Brown measure of in the following Corollaries. For the figures, we approximate the Brown measure of with the empirical spectral distribution of a specific , where are independently Haar-rotated Hermitian matrices with the same spectral distributions as and , 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 to the Brown measure of .
First, we consider the atoms of the Brown measure:
Corollary 6.3.
Let be the Brown measure of where and have atoms. Then,
-
(1)
does not have atoms.
-
(2)
is never atomic.
-
(3)
can have atoms only at the points .
-
(4)
has no atoms at and if and only if . If , then has exactly atom at either or , with size .
-
(5)
has no atoms at and if and only if . If , then has exactly atom at either or , with size .
-
(6)
has 0, 1, or 2 atoms. has 0 atoms if and only if . has 1 atom if and only if one of or . has 2 atoms if and only if and .
-
(7)
Changing and/or permutes the .
Proof.
-
(1)
is absolutely continuous and are injective, so does not have atoms. Hence, does not have atoms.
-
(2)
Since , then , so is never atomic
-
(3)
Since does not have atoms, the only atoms of can occur at the points .
-
(4)
Note that
(144) Hence, when , so has no atoms at and . If , then only one of and is positive and equal to and hence one of is equal to .
-
(5)
Follows similarly to 4, with the equation
(145) -
(6)
Directly follows from 4. and 5.
-
(7)
Follows directly from the formulas for the .
∎
Now, we consider the symmetries of :
Corollary 6.4.
Let be the Brown measure of where and have atoms
-
(1)
Swapping the roles of and fixes .
-
(2)
Changing one of or changes by changing . These correspond changing and , respectively.
-
(3)
Changing both and fixes . This corresponds to changing both and . This amounts to changing .
Proof.
-
(1)
Swapping the roles of and amounts to swapping and in the formula for . But, the formula for is symmetric with respect to , so is fixed.
-
(2)
Since is symmetric with respect to changing and , it suffices to check just . Denote by to refer to the coefficients. For this, we just need to check the identity:
(146) for . As for , it is equivalent to check
(147) Checking this identity is a straightforward calculation.
-
(3)
Follows from applying 2. twice.
∎
Figure 2 illustrates the behavior in Corollary 6.4 where the Brown measure of is approximated by the ESD of for deterministic . Note the symmetry between the two ESDs.


Finally, we consider when the density of extends all the way to the 4 corners of the intersection of the hyperbola with the boundary of the rectangle:
Corollary 6.5.
Let be the Brown measure of where and have atoms
-
(1)
has density extending to and if and only if .
-
(2)
has density extending to and if and only if .
-
(3)
has density extending to all corners of the intersection of the hyperbola with the boundary of the rectangle if and only if . Hence, the support of is equal to if and only if .
Proof.
Let and .
-
(1)
For ,
(148) Hence, .
Thus, it is equivalent to determine for which the measure has density approaching . Note that has density approaching if and only if for . Note that . Recall that is a polynomial of degree at most . If is quadratic, then , so then does not have density approaching . Hence, we must have . Conversely, if , then , and the linear term has negative coefficient for . Thus, as and has density approaching .
-
(2)
For ,
(149) Hence, .
Thus, it is equivalent to determine for which the measure has density approaching . Note that has density approaching if and only if for . Since , then we need that for . Since , then , i.e. . Conversely, if , then and since is either quadratic with positive leading coefficient or is linear with negative slope, then must be negative to the right of . Note that in the case where is quadratic, cannot have a double root at (from Proposition 5.1).
-
(3)
Follows from 1. and 2.
∎
Figure 3 illustrates the behavior in Corollary 6.5 where the Brown measure of is approximated by the ESD of , for deterministic . The left ESD does not have density approaching the corners of , but the right ESD does.


Finally, we conclude that the Brown measure of uniquely determines the laws of and . Here we allow and to be possibly constant.
Corollary 6.6.
Let be Hermitian and freely independent and
(150) | ||||
where and . Let be the Brown measure of . Then, the assignment is 1 to 1.
Proof.
From Corollary 6.3, is atomic if and only if one of or is constant. In this case, it is easy to determine the weights and atoms of both and from .
Thus, we may consider which is the Brown measure of where and are not constant.
First, we show that determines the positions of the atoms of and . Since the support of on contains at least points, then the equation of is uniquely determined. If or , then from Corollary 6.5, we can determine the positions of the atoms of and . Thus, assume that and . From Corollary 6.3, has atoms at the points . From these two points, at least of are determined. To determine the last one, we can use the equation of the hyperbola and look at either the coefficient of or . Thus, determines the positions of the atoms of and .
We can determine the weights of the atoms of and 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 -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