EXPLICIT BOUNDS FOR THE PSEUDOSPECTRA OF VARIOUS CLASSES OF MATRICES AND OPERATORS

FEIXUE GONG1, OLIVIA MEYERSON2, JEREMY MEZA3, ABIGAIL WARD4 MIHAI STOICIU5 (ADVISOR) SMALL 2014 - MATHEMATICAL PHYSICS GROUP

N×N Abstract. We study the ε-pseudospectra σε(A) of square matrices A ∈ C . We give a complete characterization of the ε- of any 2×2 and describe the asymptotic behavior (as ε → 0) of σε(A) for any square matrix A. We also present explicit upper and lower bounds for the ε-pseudospectra of bidiagonal and tridiagonal matrices, as well as for finite rank operators.

Contents

1. Introduction 2

2. Pseudospectra 2

2.1. Motivation and Definitions 2

2.2. Normal Matrices 6

2.3. Non-normal Diagonalizable Matrices 7

2.4. Non-diagonalizable Matrices 8

3. Pseudospectra of 2 2 Matrices 10

3.1. Non-diagonalizable× 2 2 Matrices 10

3.2. Diagonalizable 2 2 Matrices× 12

4. Asymptotic Union of Disks× Theorem 15 5. Pseudospectra of N N Jordan Block 20

6. Pseudospectra of bidiagonal× matrices 22 6.1. Asymptotic Bounds for Periodic Bidiagonal Matrices 22

6.2. Examples of Bidiagonal Matrices 29

7. Tridiagonal Matrices 32

Date: October 30, 2014. 1,2,5Williams College 3Carnegie Mellon University 4University of Chicago. 1 2 GONG, MEYERSON, MEZA, WARD, STOICIU

8. Finite Rank operators 33

References 35

1. Introduction

The pseudospectra of matrices and operators is an important mathematical object that has found applications in various areas of mathematics: , func- tional analysis, numerical analysis, and differential equations. An overview of the main results on pseudospectra can be found in [14].

In this paper we investigate a few classes of matrices and operators and provide explicit bounds on their ε-pseudospectra. We study 2 2 matrices, bidiagonal and tridiagonal matrices, as well as finite rank operators. We also describe the asymptotic behavior of the ε-pseudospectrum σε A of any× n n matrix A.

The paper is organized as follows: in Section 2 we( give) the three× standard equivalent definitions for the pseudospectrum and we present the “classical” results on ε- pseudospectra of normal and diagonalizable matrices (the Bauer-Fike theorems). Section 3 contains a detailed analysis of the ε-pseudospectrum of 2 2 matrices, including both the non-diagonalizable case (Subsection 3.1) and the diagonalizable case (Subsection 3.2). The asymptotic behavior (as ε 0) of the ε-pseudospectrum× of any n n matrix is described in Section 4, where we show (in Theorem 4.1) that, for any square matrix A, the ε-pseudospectrum converges,→ as ε 0 to a union of disks. × → The main result of Section 4 is applied to several classes of matrices: Jordan blocks (Section 5), bidiagonal matrices (Section 6), and tridiagonal matrices (Section 7). In Section 5 we derive explicit lower bounds for σε J , where J is any Jordan block, while Section 6 is dedicated to the analysis of arbitrary periodic bidiagonal matrices A. We derive explicit formulas (in terms the coefficients( ) of A) for the asymptotic radii of the ε-pseudospectrum of A, as ε 0. We continue with a brief investigation of tridiagonal matrices in Section 7, while in the last section (Section 8) we consider finite rank operators and show that the→ε-pseudospectrum of an operator of rank 1 m is at most as big as Cε m , as ε 0.

→ 2. Pseudospectra

2.1. Motivation and Definitions. The concept of the spectrum of a matrix A N×N C provides a fundamental tool for understanding the behavior of A. As is well- known, a complex number z C is in the spectrum of A (denoted σ A ) whenever∈ zI A (which we will denote as z A) is not invertible, i.e., the characteristic polynomial of A has z as a root.∈ As slightly perturbing the coefficients( ) of A will change− the roots of the characteristic− polynomial, the property of “membership in the set of eigenvalues” is not well-suited for many purposes, especially those in numerical analysis. We thus want to find a characterization of when a complex EXPLICIT BOUNDS FOR THE PSEUDOSPECTRA OF MATRICES AND OPERATORS 3 number is close to an eigenvalue, and we do this by considering the set of complex numbers z such that z A −1 is large, where the norm here is the usual operator norm induced by the Euclidean norm, i.e. Y( − ) Y A sup Av . YvY=1 The motivation for considering thisY Y question= Y comesY from the following observation: N×N Proposition 2.1. Let A C , λ an eigenvalue of A, and zn be a sequence of −1 complex numbers converging to λ with zn z for all n. Then zn A as n . ∈ ≠ Y( − ) Y → ∞ Proof.→ ∞ Let v be the eigenvector corresponding to λ with norm 1, so Av λv. Define

zn A v vn = zn A v ( − ) (note that zn A v 0 by assumption).= Now, note that Y( − ) Y z A −1 z A −1v ( − ) ≠ n n n z A −1 z A v Y( − ) Y ≥ Y( n− ) Yn zn A v ( − ) ( − ) = ] v ] Y( − ) Y zn A v = ] 1 ] Y( −, ) Y λ zn and since the last quantity gets arbitrarily= large for large n, we obtain the result.  S − S We call the operator z A −1 the resolvent of A. The observation that the norm of the resolvent is large when z is close to an eigenvalue of A leads us to the first definition of the ε-pseudospectrum( − ) of an operator. N×N Definition 2.1. Let A C , and let ε 0. The ε-pseudospectrum of A is the set of z C such that ∈ z A −1> 1 ε ∈ Note that the boundary of the εY(-pseudospectrum− ) Y > ~ is exactly the 1 ε level curve of the function z z A −1 . Fig. 2.1 depicts the behavior of this function near the eigenvalues, as described in Prop. 2.1. ~ ↦ Y( − ) Y The resolvent norm has singularities in the complex plane, and as we approach these points, the resolvent norm grows to infinity. Conversely, if z A −1 approaches infinity, then z must approach some eigenvalue of A [14, Thm 2.4]. Y( − ) Y (It is also possible to develop a theory of pseudo-spectrum for operators on Banach spaces, and it is important to note that this converse does not necessarily hold for such operators; that is, there are operators [4, 5] such that z A −1 approaches infinity, but z does not approach the spectrum of A.) Y( − ) Y The second definition of the ε-pseudospectrum arises from eigenvalue perturbation theory [8]. 4 GONG, MEYERSON, MEZA, WARD, STOICIU

Figure 2.1. Contour Plot of Resolvent Norm

N×N Definition 2.2. Let A C . The ε-pseudospectrum of A is the set of z C such that ∈ z σ A E ∈ for some E with E ε. ∈ ( + ) Finally, the thirdY definitionY < of the ε-pseudospectum is the following: N×N Definition 2.3. Let A C . The ε-pseudospectrum of A is the set of z C such that ∈ z A v ε ∈ for some unit vector v. Y( − ) Y < This definition is similar to our first definition in that it quantifies how close z is to an eigenvalue of A. In addition to this, it also gives us the notion of an ε-pseudoeigenvector. Theorem 2.1 (Equivalence of the definitions of pseudospectra). For any matrix N×N A C , the three definitions above are equivalent.

∈ The proof of this theorem follows [14, §2].

Proof. Define the following sets −1 −1 Mε z C z A ε N z z σ A E for some E N×N , E ε ε C = { ∈ S Y( − ) Y > C } N Pε z C z A v ε for some v C , v 1 = { ∈ S ∈ ( + ) ∈ Y Y < }

To show equivalence,= { we∈ willS Y( show− ) thatY

Nε Pε ⊆ ⊆ ⊆

Let⊆z Nε. Let E be such that z σ A E , with E ε. Let v be an eigenvector of norm one for z, that is v 1 and A E v zv. Then ∈ ∈ ( + ) Y Y < z A v z A E v Ev E v ε Y Y = ( + ) = and hence z P . ε Y( − ) Y = Y( − − ) + Y ≤ Y YY Y < ∈ EXPLICIT BOUNDS FOR THE PSEUDOSPECTRA OF MATRICES AND OPERATORS 5

Pε Mε

N Let⊆z Pε. Let v C be such that v 1 and z A v ε. Observe z A v ∈ ∈ z A v zY YA= v Y( − ) suY < z A v ( − ) (z−(A)−v ) = Y( − ) Y = −1 −1 where we define u Y(z−A)vY and s z A vY(. Then,− ) Y we see that s v z A u, which implies that = = Y( − ) Y = ( − ) z A −1 z A −1u s−1 v s−1 ε−1 and hence z Mε.Y( − ) Y ≥ Y( − ) Y = S SY Y = >

Mε Nε ∈

Let ⊆z Mε. We will construct an E with E ε and z σ A E . Since −1 −1 −1 z A supYvY=1 z A v ε , there exists some v with v 1 such that z A∈ −1v ε−1. Observe Y Y < ∈ ( + ) Y( − ) Y = Y( − ) Y > Y Y = z A −1v Y( − ) Y > z A −1v z A −1v z A −1v u z A −1v ( − ) ( − ) = Y((z−A−)−1)v Y v = Y( − ) Y where u is the unit vector Y(z−A)−1vY . SoY(Y(z−−A))−1vY Y z A u, from which we have v zu Au = ( − ) z A −1v Define E v u∗. Note Eu= + v and E ε. Then we have zu Y(z−A)−1vY Y(zY(−A)−1vY) Y Au Eu, and so z σ A E .  = = Y Y < = + ∈ ( + ) As all three definitions are now shown to be equivalent, we can unambiguously denote the ε-pseudospectrum of A as σε A .

Fig. 2.2 depicts an example of ε-pseudospectra( ) for a specific matrix and for var- ious ε. We see that the boundaries of ε-pseudospectra for a matrix are curves in the complex plane around the eigenvalues of the matrix. We are interested in understanding geometric and algebraic properties of these curves.

Figure 2.2. The curves bounding the ε-pseudospectra of an op- erator A, for different values of ε. 6 GONG, MEYERSON, MEZA, WARD, STOICIU

When a matrix A is the direct sum of smaller matrices, we can look at the pseu- dospectra of the smaller matrices to understand the ε-pseudospectrum of A. We get the following theorem from [14]: Theorem 2.2.

σε A1 A2 σε A1 σε A2 .

( ⊕ ) = ( ) ∪ ( )

2.2. Normal Matrices. Recall that a matrix A is normal if AA∗ A∗A, or equiv- alently if A can be diagonalized with an orthonormal basis of eigenvectors. = The pseudospectra of these matrices is particularly well-behaved: Thm. 2.3 shows that the ε-pseudospectrum of a is exactly a disk of radius ε around each eigenvalue, as in shown in Fig. 2.3. This is clear for diagonal matrices; it follows for normal matrices since as we shall see, the ε-pseudospectrum of a matrix is under a unitary change of basis.

Figure 2.3. The ε-pseudospectrum of a normal matrix. Note that each boundary is a perfect disk around an eigenvalue.

N×N Theorem 2.3. Let A C . Then,

(2.1) σ ∈A B 0, ε σε A for all ε 0 and if A is normal and( the) + operator( ) ⊆ norm( ) is the 2-norm, then> (2.2) σε A σ A B 0, ε for all ε 0

Conversely, (2.2) implies( that) = A( is) + normal.( ) >

Proof. Let z σ A B 0, ε . Write z λ w for some λ σ A and w C with w ε. Let v be an eigenvector for A with eigenvalue λ. Then, v is an eigenvector for the perturbed∈ ( system) + (A )wI with eigenvalue= + z. Thus, z ∈ σ(A) wI , and∈ hence zS S σ<ε A . This shows (2.1). + ∈ ( + ) To∈ show( ) (2.2), we now suppose A is a normal matrix. It suffices to show the backwards inclusion to (2.1). Write A UΛU ∗ for a unitary U and diagonal Λ. We have = 1 z A −1 UzU ∗ UΛU ∗ −1 U z Λ −1U ∗ z Λ −1 dist z, σ A Y( − ) Y = Y( − ) Y = Y ( − ) Y = Y( − ) Y = ( ( )) EXPLICIT BOUNDS FOR THE PSEUDOSPECTRA OF MATRICES AND OPERATORS 7 which implies that if z σε A , then dist z, σ A ε. Rewriting this, we get σε A σ A B 0, ε . ∈ ( ) ( ( )) < The( converse) ⊆ ( ) of+ this( proof) follows from the discussion below of the condition number and the full proof is postponed for now; see Corollary 2.1.



2.3. Non-normal Diagonalizable Matrices. Now suppose A is diagonalizable N but not normal, i.e. we cannot diagonalize A by an isometry of C . In this case we do not expect to get an exact characterization of the ε-pseudospectra as we did previously. The following example shows that there exist matrices with pseudospectra that are larger than the disk of radius ε: 1 1 Example 2.1. For the non-normal A , 0 2 − B 2, Cε σε A = Œ ‘ for any C 2 as ε 0. ( ) ⊆ ( ) √ < → Proof. Let z B 2, Cε , where C 2. We will construct E with E ε so that z is an eigenvalue of A E. √ ∈ ( ) < Y Y < 0 0 Let E where+ α √1 . Direct calculation gives E 2 α ε ε. We αε αε 2 √ find the= Œ eigenvalues‘ of A ES :S < Y Y = S S < − det A E zI 1 z 2 αε z αε z2 z 3 αε 2 0 + The roots of this equation are given by ( + − ) = ( − )( + − ) − = − ( + ) + = 3 αε 1 6αε α2ε2 z . √ 2 + ± + + Recall for x small, the Taylor= expansion 1 x p 1 px O x2 . This gives that the root in which we add the discriminant has the expansion ( + ) = + + ( ) 3 αε 1 3αε O ε2 z z 2 2αε O ε2 2 + + ( + + ( )) Since we may take any α with α √1 , we obtain the result. = 2 Ô⇒ = + + ( )

S S < 

We begin here the characterization of the behavior of non-normal, diagonalizable matrices. Write A VDV −1. Define the condition number of V as

−1 smax V = κ V V V smin V ( ) where s V and s V (are) = theY maximumYY Y = and minimum singular values of V , max min ( ) respectively. Note that there is some ambiguity when we define κ V , as V is not uniquely determined.( ) If( the) eigenvalues are distinct, then κ V becomes unique if the eigenvectors are normalized by vj 1.While this choice may( not) necessarily ( ) Y Y = 8 GONG, MEYERSON, MEZA, WARD, STOICIU be the one that minimizes κ V , [13] shows that this choice exceeds the minimal value by at most a factor of N. √( ) Note that κ V is necessarily greater than or equal to 1, and we claim κ V 1 if and only if A is normal. The backward direction follows from our unitarily equivalent norm:( ) if A is normal, then V is unitary, which implies that V (V )−1= 1 and hence κ V 1. To show the forward direction, suppose κ V smax(V ) 1. smin(V ) Thus, in the singular value decomposition of V , the singular values areY allY = theY same,Y = which implies that( ) =V sUW ∗ for s 0 and unitary matrices U, W .( Define) = V ′ V= s, and note that V ′ is a . Also note = > = ~ V ′DV ′−1 V s D sV −1 VDV −1 A and thus A is unitarily diagonalizable and hence normal. = ( ~ ) ( ) = = The condition number gives an upper bound on the size of the pseudo spectrum of a non-normal diagonalizable matrix:

N×N Theorem 2.4 (Bauer-Fike). Let A C and let A be diagonalizable, A VDV −1. Then for each ε 0, ∈ = σ A B 0, ε σ A σ A B 0, εκ V > ε where ( ) + ( ) ⊆ ( ) ⊆ (s ) + V( ( )) κ V V V −1 max . smin V ( ) ( ) = Y YY Y = ( ) Corollary 2.1. If σε A σ A B 0, ε , then A is normal.

( ) = ( ) + ( ) Proof (Thm. 2.4). The first inclusion was established in the previous theorem. For the second inclusion, let z σε A . Then z A −1 z VDV −1 −1 V z D V −1 −1 V z D −1V −1 ∈ ( ) which implies that ( − ) = ( − ) = [ ( − ) ] = ( − ) κ V ε−1 z A −1 κ V z D −1 , dist z, σ A ( ) whence < Y( − ) Y ≤ ( )Y( − ) Y = ( ( )) dist z, σ A εκ V .  ( ( )) < ( ) 2.4. Non-diagonalizable Matrices. So far we have considered normal matrices, and more generally diagonalizable matrices. We now relax our constraint that our matrix be diagonalizable, and provide similar bounds on the pseudospectra. While not every matrix is diagonalizable, every matrix can be put in . Below we give a brief review of the Jordan form.

N×N Let A C and suppose A has only one eigenvalue, λ with geometric multiplicity one. Writing A in Jordan form, there exists a matrix V such that AV VJ, where J is a∈ single Jordan block of size N. Write = V v1, v2, . . . , vn

= ‰ Ž EXPLICIT BOUNDS FOR THE PSEUDOSPECTRA OF MATRICES AND OPERATORS 9

Then,

AV Av1, Av2, . . . , Avn λv1, v1 λv2, . . . , vn−1 λvn V J, where v is the right eigenvector associated with λ and v , ..., v are generalized =1‰ Ž = ‰ + 2 n + Ž = right eigenvectors, that is right eigenvectors for A λI k for k 1.

We also know that there exists a matrix U ∗ V (−1 such− ) that U ∗>A JU ∗.

u1 u = = U 2 . ⎛ ⎞ ⎜ ⎟ = ⎜un⎟ ⎜ ⋮ ⎟ Then, ⎝ ⎠ λu1 u2 u1A λu2 u3 ∗ u2A ∗ U A ⎛ + ⎞ JU . ⎛ ⎞ ⎜λu + u ⎟ ⎜ ⎟ ⎜ n−1 n⎟ = ⎜unA⎟ = ⎜ ⋮ ⎟ = ⎜ ⋮ ⎟ ⎜ λun ⎟ ⎝ ⎠ ⎜ + ⎟ Thus, un is the left eigenvector associated⎝ with λ ⎠and u1, ..., un−1 are generalized left eigenvectors.

Similar to how κ V quantifies the normality of a matrix, we can also quantify the normality of an eigenvalue: ( ) Definition 2.4. For any simple eigenvalue λj of a matrix A, the condition number of λj is defined as uj vj κ λj ∗ , u vj Y jYY Y where v and u∗ are the right and( left) eigenvectors= associated with λ , respectively. j j S S j

∗ Note: The Cauchy-Schwarz inequality implies that uj vj uj vj , so κ λj 1, with equality when uj and vj are collinear. An eigenvalue for which κ λj 1 is called a normal eigenvalue; a matrix A with all eigenvaluesS S ≤ simpleY YY isY normal( if) and≥ only if κ λj 1 for all eigenvalues. ( ) =

With this( definition,) = we can find finer bounds for the pseudospectrum of a matrix; in particular, we can find bounds for the components of the pseduospectrum centered around each eigenvalue.

The following theorem can be found for example in [3].

N×N Theorem 2.5 (Asymptotic pseudospectra inclusion regions). Suppose A C has N distinct eigenvalues. Then, as ε 0, N ∈ 2 σε A B λ→j, εκ λj ε . j=1 ( ) ⊆ ‰ ( ) + O ‰ ŽŽ

We can drop the ε2 term, for which we get an increase in the radius of our inclusion disks by a factor of N [2, Thm. 4]. O( ) 10 GONG, MEYERSON, MEZA, WARD, STOICIU

N×N Theorem 2.6 (Bauer-Fike theorem based on κ λj ). Suppose A C has N distinct eigenvalues. Then ε 0, N ( ) ∈ σ∀ε A> B λj, εNκ λj . j=1 ( ) ⊆ ( ( )) The following has a proof in [14].

Theorem 2.7 (Asymptotic formula for the resolvent norm). Let λj σ A be an eigenvalue of index kj. For any z σε A , for small enough ε,

1 ∈ ( ) k z∈ λj( ) Cjε j ,

kj −1 ∗ where Cj VjTj Uj and T S J− λIS ≤. ( ) = Y Y = − 3. Pseudospectra of 2 2 Matrices

The following section presents a complete characterization× of the ε-pseudospectrum of any 2 2 matrix. We classify 2 2 matrices by whether they are diagonalizable or non-diagonalizable and then determine the ε-pseudospectra for each class. We begin with× an explicit formula for× computing the norm of a 2 2 matrix, and then analyze a 2 2 Jordan block. × a b Let A × , with a, b, c, d . Let s denote the largest singular value of A. c d C max = Œ ‘ ∈ Then, 2 A smax largest eigenvalue of A∗A SS SS = Tr A∗A Tr A∗A 2 4 det A∗A = . » 2 ( ) + ( ) − ( ) Letting ρ a 2 b 2 c=2 d 2, we can expand the norm as 1 = S SA+2S S + S S ρ+2S S4 a 2 d 2 b 2 c 2 adbc adbc ρ . 2 ¼ This formulaY forY the= norm‹ of− the‰S matrixS S S + allowsS S S S us− ( to compute+ ) theŽ + ε-pseudospectra of 2 2 matrices.

× 3.1. Non-diagonalizable 2 2 Matrices.

Proposition 3.1. Let A be any× non-diagonalizable 2 2 matrix, and let λ denote the eigenvalue of A. Write A VJV −1 where × a b λ 1 V ,= a, b, c, d , and J . c d C 0 λ

Then σε A is exactly= Œ a‘ disk. More specifically,∈ given any =ε,Œ ‘ σ A B λ, k ( ) ε ( ) = ( S S) EXPLICIT BOUNDS FOR THE PSEUDOSPECTRA OF MATRICES AND OPERATORS 11 where a 2 c 2 (3.1) k Cε ε2 and C . ad bc √ S S + S S S S = + = S − S Proof. Let z λ k where k C. Then, z A −1 z VJV −1 −1 V z J −1V −1. = + ∈ Let k z λ. We obtain that ( − ) = ( − ) = ( − ) = − 1 adk ac bck a2 z A −1 . k2 ad bc c2 bck ac adk − − Y( − ) Y = ]Œ ‘] Now, let S ( − )S − − + + adk ac bck a2 M . c2 bck ac adk − − = Œ ‘ A simple calculation gives − − + + Tr M ∗M a 2 c 2 2 2 k 2 bc ad 2 det M ∗M k 4 bc ad 4 ( ) = (S S + S S ) + S S S − S From the formula for the( norm,) we= S obtainS S − thatS z A −1 V z J −1V −1

Y( − ) Y = Y Tr( −M ∗)M Y Tr M ∗M 2 4 det M ∗M ¼ . »k 2 ad bc 2 ( ) + ( ) − ( ) Note that this function= depends only on k z √λ ; thus for any ε, σ A will S S S − S ε be a disk. Solving for k in the above equation to find the curve bounding the pseudospectrum, we obtain S S = S − S ( )

a 2 c 2 k ε ε2. ¿ ad bc Á S S + S S S S = ÀÁ + S − S  Example 3.1. Perhaps the simplest example of such a matrix is the 2 2 Jordan block. × λ 1 Let J where λ . In Jordan form, J decomposes as 0 λ C = Œ ‘ ∈ 1 0 λ 1 1 0 J . 0 1 0 λ 0 1 Plugging into Equation 3.1 with= Œ a 1‘, bŒ 0, c‘ Œ0, d ‘1, we get that ε2 1 0 1 0 ε 1 0 1 0 k = = = = ε ε2. » 1 0 ( − )( − ) + S − S( + ) √ S S = = + S − S 12 GONG, MEYERSON, MEZA, WARD, STOICIU

Hence, 2 σε J B λ, ε ε . √ We can check this result by direct( application) = Š of+ the first definition of pseudospectra.

Let z σε J . Note −1 z λ 1 1 z λ 1 ∈ ( ) z J −1 . 0 z λ z λ 2 0 z λ − − − Using our equation( − for) the= Œ norm of a 2‘ 2= matrix, weŒ get ‘ − ( − ) − 1 1 1 z J −1 × 2 z λ 2 1 4 z λ 2 1 . ε z λ 2 ¾2 » 4 < SS( − ) SS = Š 2Sw− S +2 + S −2 S +  Let w z λ , and square bothS sides− S to get ε2 2w 1 4w 1. This simplifies to w ε ε2. Thus, z σ J z λ ε ε2 √ z B λ, ε ε2 . = S − S ε < + + + √ √ √ < + ∈ ( ) ⇐⇒ S − S < + ⇐⇒ ∈ Š +  3.2. Diagonalizable 2 2 Matrices.

Proposition 3.2. Let A× be any diagonalizable 2 2 matrix and let λ1, λ2 be the eigenvalues of A and v1, v2 be the eigenvectors associated with the eigenvalues. Then σε A is the set of points z that satisfy the equation× 2 2 2 2 2 2 2 ( ) ε z λ1 ε z λ2 ε λ1 λ2 cot θ 0, where θ is the( angle− S between− S )( the− twoS − eigenvectors.S ) − S − S ( ) <

Proof. Since A is diagonalizable, we know that A VDV −1 where a b λ 0 V D 1 = . c d 0 λ2 = Œ ‘ = Œ ‘ Note that V is an with columns corresponding to the eigenvectors of A.

Without loss of generality, we can write z λ1 k.

Let γ λ1 λ2 and r ad bc. Then, = + −1 −1 −1 −1 −1 = − z =A − z VDV V z D V 1 1 a b k 0 d b ( − ) = ( − ) =1 ( − ) r c d 0 γ+k c a − = Œ 1 ‘ Œ adγ ‘rk Œ abγ‘ − . rk γ k cdγ bcγ rk + − = Œ ‘ Œ ‘ ( + ) − + adγ rk abγ Let M . cdγ bcγ rk + − = Œ −1 1 ‘ Then, z A S−rk(γ+k+)S M . Y( − ) Y = Y Y EXPLICIT BOUNDS FOR THE PSEUDOSPECTRA OF MATRICES AND OPERATORS 13

Calculating Tr M ∗M and det M ∗M gives Tr M ∗M cdγ 2 kr adγ 2 abγ 2 kr bcγ 2 ( ) ( ) 2 kr 2 2 r 2 Re kγ γ 2 a 2 c 2 b 2 d 2 ( ) = S S + S + S + S S + S − S a 2 c 2 b 2 d 2 = rS 2 S2+k 2S S 2 Re[ kγ] + S γS 2(S Sγ+2 S S )(S S + S S ) 1 . r 2 (S S + S S )(S S + S S ) Note that the first= S S partŒ S ofS + this expression[ ] + S S simplifies+ S S Œ to − ‘‘ S S 2 k 2 2 Re kγ γ 2 k γ 2 k 2. For the second part, we have S S + [ ] + S S = S + S + S S a 2 c 2 b 2 d 2 1 1 a 2 c 2 b 2 d 2 r 2 r 2 r 2 (S S + S S )(S S + S S ) 1 − = ((SabS2 + ScdS )(S2 bcS ad+ S S bcad) − S S. ) S S SrS2

Now, note that our eigenvectors v1, v2= are(S exactlyS + S theS columns+ + of V). So, we have 2 2 S S 2 2 v1 v2 ab¯ cd¯ ab cd adbc adbc. Thus, the second term in our simplifies to S ⋅ S = S + S = S S + S S + + v v 2 v 2 v 2 cos2 θ 1 2 1 2 cot2 θ, 2 2 2 2 r v1 v2 sin θ S ⋅ S S S S S where the denominator comes from= the fact that r= det V is the area of the paral- S S S S S S lelogram spanned by v1 and v2. Thus, we get that = Tr M ∗M r 2 γ k 2 k 2 γ 2 cot2 θ . For the , we have ( ) = S S ‰S + S + S S + S S Ž det M ∗M cdγ 2 kr adγ 2 abγ 2 kr bcγ 2 kr bcγ cdγ abγ kr adγ kr adγ abγ cdγ kr bcγ ( ) = (S S + S + S )(S S + S − S ) k2r2 kr2γ k2r2 kr2γ − (( − ) − ( + ))(−( + ) + ( − )) r 4 k 2 k γ 2 = ( + )( + ) We then plug= theseS S S S inS to+ theS equation for the norm of a 2 2 matrix and get

∗ ∗ 2 ∗ − − Tr M M Tr M M 4 det ×M M ε 1 z A 1 ¿ »2 r 2 k γ k 2 Á ( ) + ( ) − ( ) <2Y(k γ− k) 2 Y = ÀÁ γ k 2 k 2 γ 2 cot2 θ γ k 2 k 2 γ 2 cot2 θ 2 4 k γ k 2 2 S S S ( + )S ε ¼ S ( + )S 2 ⇐⇒ 2 k γ k 2< S + S + S S + S S + (S + S + S S + S S ) − S ( + )S γ k 2 k 2 γ 2 cot2 θ γ k 2 k 2 γ 2 cot2 θ 2 4 k γ k 2 ε2 S ( + )S ⇐⇒ Œ4 k γ k 4 −4(Sk γ+ Sk +2S S + S S )‘ < (S + S + S S + S S ) − S ( + )S γ k 2 k 2 γ 2 cot2 θ 4 k γ k 2 0 ε4 ε2 S ( + )S2 4S ( + )S2 2 2 2 2 ⇐⇒ 4 k γ k −ε γ k (S k+ S γ+ S cotS + Sθ Sε 0 ) + S ( + )S < ε2 k 2 ε2 k γ 2 ε2 γ 2 cot2 θ 0. ⇐⇒ S ( + )S + − (S + S + S S + S S ) <

⇐⇒ ( − S S )( − S + S ) − S S <  14 GONG, MEYERSON, MEZA, WARD, STOICIU

Note that for normal matrices, the eigenvectors are orthogonal, so θ π 2 and cot θ 0. Therefore the equation above reduces to 2 2 2 2 = ~ = ε k ε k γ 0 which describes two disks of( radius− S Sε)(centered− S + aroundS ) < λ1, λ2, as we expect. When the matrix only has one eigenvalue and is still diagonalizable (i.e. when it is a multiple of the identity), then we get ε2 k 2 ε2 k 0 2 0, which is a disk of radius ε centered( − S S around)( − S the+ eigenvalue.S ) <

One consequence to note of Proposition 3.2 is that the shape of σε A is dependent on both the eigenvalues and the eigenvectors of A. Another less obvious consequence is that the pseudospectrum of a 2 2 matrix approaches a union of( disks) as ε tends to 0. This is proven in the following proposition. × Proposition 3.3. Let A be a diagonalizable 2 2 matrix. Then, σε A asymptot- rmax ically tends toward a disk. In particular, 1 ε , where rmax, rmin are the rmin maximum and minimum distances from an eigenvalue× to the boundary( of) the closer connected component of ∂σε A . Moreover, for= A+ Odiagonalizable( ) but not normal, σε A is never a perfect disk. ( ) ( ) Proof. Let ε be small enough so that the ε-pseudospectrum is disconnected. With- out loss of generality, we will show that the component around λ1 tends towards a disk by showing that the ratio of the maximum distance from z to λ1 to the minimum distance from z to λ1 is 1 o ε .

Let zmax ∂σε A such that zmax λ+1 (is) a maximum. Consider the line joining λ1 and λ2. Suppose for contradiction that zmax did not lie on this line. Then, rotate ′ zmax in the∈ direction( ) of λ2 soS that− it isS on this line, and call this new point z . Note ′ ′ that z λ2 zmax λ2 , but z λ1 zmax λ1 . As such, we get that ′ 2 2 ′ 2 2 2 2 2 2 2 2 2 z Sλ1 − εS < Sz λ−2 Sε S zmax− Sλ=1S ε − zmaxS λ2 ε ε λ1 λ2 cot θ ′ ′ Thus, from Proposition 3.2 z σε A but z is not on the boundary of σε A . (S − S − )(S′ − S − ) < (S − S − )(S − S − ) = S −′′ S Starting from z and traversing the line joining λ1 and λ2, we can find z ∂σε A ′′ ′ such that z λ1 z λ1 z∈max( λ)1 . This contradicts our choice of zmax (and) so zmax must be on the line joining λ1 and λ2. A similar argument shows that∈ z(min) must also beS on− theS > lineS − joiningS = Sλ1 and− λ2S, where zmin ∂σε A such that zmin λ1 is a minimum. ∈ ( ) S − S Since zmax is on the line joining λ1 and λ2, we have the exact equality zmax λ2 zmax λ1 λ2 λ1 . Let rmax zmax λ1 and let y λ2 λ1 . The equation describing rmax becomes S − S = S − S + S − S 2 2 = S − 2 S 2 2 2 = S 2 − S rmax ε y rmax ε ε y cot θ

Solving this equation‰ for rmax− ,Ž we ‰( get− ) − Ž = 1 r y y2 4ε2 4yε csc θ max 2 » = Š − + −  EXPLICIT BOUNDS FOR THE PSEUDOSPECTRA OF MATRICES AND OPERATORS 15

Similarly, we get 1 r y2 4ε2 4yε csc θ y min 2 » p p(p−1) 2 3 For ε small, we can use the approximation= Š + 1+ ε 1 −pε 2 ε o ε . Using this, we see ( + ) = + + + ( ) 2 rmax 1 1 4 ε y 4 ε y csc θ rmin 1»4 ε y 2 4 ε y csc θ 1 − + ( ~ ) − ( ~ ) = »2 ε y csc θ 2 ε y 2 2 ε y 2 csc2 θ ε3 + ( ~ ) + ( ~ ) − 2 ε y csc θ 2 ε y 2 2 ε y 2 csc2 θ ε3 ( ~ ) − ( ~ ) + ( ~ ) + O( ) = csc θ ε y csc2 θ 1 ε2 ( ~ ) + ( ~ ) − ( ~ ) + O( ) csc θ ε y csc2 θ 1 ε2 + ( ~ )( − ) + O( ) = 1 ηε ε2 − ( ~ )( − ) + O( ) 1 ηε ε2 + + O( ) where η cos θ cot θ.= Recall the geometric series approximation 1 1 x x2 . − + O( ) 1−x Using this, we get our desired result = r = + +O( ) max 1 ηε ε2 1 ηε ε2 rmin = (1 +2ηε + O(ε2))( + + O( )) 1 2 cos θ cot θ ε ε2 = + + O( ) Thus, as ε 0, the farthest= point+ ( on the boundary) + O( of σ) ε A from λ1 tends towards the closest point on the boundary of σε A from λ1. Hence, σε A tends towards a disk. → ( ) ( ) ( ) Moreover, if A is diagonalizable but not normal, then the eigenvectors are linearly independent but not orthogonal, so θ is not a multiple of π 2 or π, and hence rmax cos θ cot θ 0 and r 1. min ~ ≠ ≠ 

4. Asymptotic Union of Disks Theorem

In the previous section, we showed that the ε-pseudospectrum for all 2 2 ma- trices are disks or asymptotically converge to a union of disks. We now explore whether this behavior holds in the general case. It is possible to find matrices× whose ε pseudospectra exhibit pathological properties for large ε; for example, the following non-diagonalizable matrix has, for larger ε, an ε-pseudospectrum that is not convex− and not simply connected.

Thus, pseudospectra may behave poorly for large enough ε; however, in the limit as ε 0, these properties disappear and the pseudospectra behave as disks centered around the eigenvalues with well-understood radii. The following theorem allows us→ to understand this asymptotic behavior.

We will use the following set-up (which follows [10]) for our theorem. 16 GONG, MEYERSON, MEZA, WARD, STOICIU

1 10 100 1000 10000 0 1 10 100 1000 ⎛−0− 0 − 1 − 10 − 100 ⎞ ⎜ 0− 0− 0 − 1 − 10 ⎟ ⎜ ⎟ ⎜ − − − ⎟ ⎜ 0 0 0 0 1 ⎟ ⎜ − − ⎟ ⎝ − ⎠

Figure 4.1. Pseudospectra of a

N×N For A C , find matrices that Jordanize A. That is, J Q Q ∈ A P Pˆ , P Pˆ I Jˆ Qˆ Qˆ ⎛ ⎞ ⎛ ⎞ ⎛ ⎞ = ‰ Ž ‰ Ž = ⎝ ⎠ ⎝ ⎠ ⎝ ⎠ and J consists of Jordan blocks J1,...,Jm corresponding to an eigenvalue λ. Jˆ is the part of the Jordan form containing the other eigenvalues of A.

Let n be the size of the largest Jordan block corresponding to λ, and suppose there are ` Jordan blocks of size n n. Arrange J1,...,Jm in J so that dim J dim J dim J dim J 1 × ` `+1 m where J ,...,J are n n. 1 ` ( ) = ⋯ = ( ) > ( ) ≥ ⋯ ≥ ( ) Further partition × P P1 ,...,P` ,...,Pm in a way that agrees with the above partition of J, so that the first column, xj, of = ‰ Ž each Pj is a right eigenvector of A associated with λ. We also partition Q likewise

Q1

⎛ Q ⎞ Q ⎜ ` ⎟ . ⎜ ⋮ ⎟ ⎜ ⎟ ⎜ ⎟ = ⎜ ⎟ ⎜ ⎟ ⎜Qm⎟ ⎜ ⋮ ⎟ The last row, yj, of each Qj is a left eigenvector⎝ ⎠ of A corresponding to λ. We now build the matrices

y1

y2 Y ⎛ ⎞ ,X x1 , x2 , . . . , x` , ⎜ ⎟ ⎜ ⎟ ⎜ ⎟ = ⎜y ⎟ = ‰ Ž ⎜ `⎟ ⎜ ⋮ ⎟ where X and Y are the⎝ matrices⎠ of right and left eigenvectors, respectively, corre- sponding to the Jordan blocks of maximal size for λ. EXPLICIT BOUNDS FOR THE PSEUDOSPECTRA OF MATRICES AND OPERATORS 17

With these matrices defined, we can understand the radii of the ε-pseudospectrum of A as ε 0. N×N Theorem→ 4.1. Let A C . Let ε 0. Then given λ σ A , for ε small enough, there exists a connected component U σε A such that U σ A λ; denote this component of the ε-pseudospectrum∈ σ>ε A λ. ∈ ( ) ⊆ ( ) ∩ ( ) = Then, as ε 0, ( ) ↾ B λ, Cε 1~n o ε1~n σ A B λ, Cε 1~n o ε1~n → ε λ where C XY , with X,Y defined above. ( ( ) + ( )) ⊆ ( ) ↾ ⊆ ( ( ) + ( ))

Proof. = Y Y

Lower Bound: We begin by proving the first containment, 1~n 1~n B λ, Cε o ε σε A λ . Let E N×N . From [10, Theorem 2.1], we know there is an eigenvalue of the C ( ( ) + ( )) ⊆ ( ) ↾ perturbed matrix A εE admitting a first order expansion ∈ λ ε λ γε 1~n o ε1~n , + where γ is any eigenvalue of YEX. [10] later shows in Theorem 4.2 that ( ) = + ( ) + ( ) α max γmax XY , YEY≤1 where γmax is the largest eigenvalue∶= of YEX=.Y In fact,Y the E that maximizes γ is given by E vu where v and u are the right and left singular vectors of the largest singular value of XY , normalized so v u 1. = 1~n 1~n We claim that B λ, α ε o ε Y Y =σYε AY =.

Fix θ 0, 2π , and( define(S S ) E˜ +eiθ(E. Considering)) ⊆ ( ) the perturbed matrix A εE˜, we can write ∈ [ ] λ=ε λ γε 1~n o ε1~n + where γ is now any eigenvalue of Y EX˜ . Solving for γ, we obtain: ( ) = + ( ) + ( ) 0 det eiθYEX γI det eiθ YEX e−iθγI , which implies 0 det YEX e−iθγI , and thus e−iθγ is any eigenvalue of YEX. = ( − ) = ‰ ( − )Ž Since α is an eigenvalue of YEX, we can take γ eiθα. Thus, we obtain = ( − ) λ eiθαε 1~n o ε1~n σ A . = ε Ranging θ from 0 to 2nπ, we get + ( ) + ( ) ∈ ( ) 1~n 1~n B λ, α ε o ε σε A .

Upper Bound: We now prove( (S theS ) second+ ( containment,)) ⊆ ( ) 1~n 1~n σε A λ B λ, Cε o ε .

From [14, §52, Theorem 52.3],( ) ↾ we⊆ have( that( ) asymptotically+ ( )) 1~n 1~n σε A λ B λ, βε o ε , where β PDn−1Q and J λI D. We claim β XY α. ( ) ↾ ⊆ ( ( ) + ( )) = Y Y = + = Y Y = 18 GONG, MEYERSON, MEZA, WARD, STOICIU

n−1 Note that D diag Γ1,... Γ`, 0 where Γk is a n n matrix with a 1 in the top right entry and zeros elsewhere. We find = [ ] ×

n−1 PD ⎛ x1 x2 x` ⎞ . ⎜ ⎟ ⎜ ⎟ = ⎜ ⋯ ⎟ ⎜ ⎟ ⎜ ⎟ This then gives ⎝ ⎠ XY 0 PDn−1Q . 0 0 Thus β PDn−1Q XY α. = Œ ‘

Combining= Y this withY = ourY lowerY = bound result, we see that as ε 0, 1~n 1~n σε A λ B λ, αε o ε . → ( ) ↾ = ( ( ) + ( )) 

We present special cases of matrices to explore the consequences of Theorem 4.1.

Special Cases:

(1) λ is simple. That is, λ has algebraic multiplicity 1. Then, n 1 and X and Y become the right and left eigenvectors x and y∗ for λ, respectively. Hence, C XY xy∗ x y κ λ where we normalize so= that y∗x 1. Then, Theorem 4.1 becomes = Y Y = Y Y = Y YY Y = ( ) S σS ε= A λ B λ, κ λ ε which matches with Theorem( ) ↾ 2.5.≈ ( ( ) ) (2) λ is semisimple. That is, the algebraic multiplicity of λ equals the geometric multiplicity. Then, n 1, and we obtain X P and Y Q, and C becomes the norm of the spectral projector for λ; see [14]. (3) λ has geometric= multiplicity 1 = = In this case we obtain the same result for when λ is simple, except n may not equal 1. In other words,

1~n σε A λ B λ, κ λ ε . 2×2 (4) A C . ( ) ↾ ≈ ( ( ( ) ) ) There are two cases, as in Section 3: Pseudospectra of 2 2 Matrices. ∈ First, assume A is non-diagonalizable. In this case, A only× has one eigen- value, λ. Let a b 1 d b V ,V −1 c d ad bc c a − =−1Œ ‘ = Œ ‘ where A VJV . − − = EXPLICIT BOUNDS FOR THE PSEUDOSPECTRA OF MATRICES AND OPERATORS 19

Then, we have that

a 1 X ,Y c a . c ad bc = Œ ‘ = ‰− Ž From Theorem 4.1, we then have that− as ε 0,

1~2 a 2 c 2 → σ A B λ, ε o ε1~2 . ε ad bc ⎛ S S + S S ⎞ ( ) ⊆ Œ ‘ + ‰ Ž ⎝ S − S ⎠ Equation 3.1 gives the explicit formula for σε A as ε 0, namely

1~2 a 2 c 2 ( ) → σ A B λ, ε ε2 ε ad bc ⎛ S S + S S ⎞ ( ) = Œ +1~2 ‘ a 2 c 2 B ⎝λ, S − S ε ⎠ε3~2 , ad bc ⎛ S S + S S ⎞ = Œ ‘ + O( ) which agrees with the above⎝ S theorem.− S ⎠

In the case where A is diagonalizable, A has two eigenvalues, λ1 and λ2. Without loss of generality, we will look at σε A around λ1.

a b λ 0 1 d( ) b A 1 . c d 0 λ2 ad bc c a − = Œ ‘ Œ ‘  Œ ‘ From this, we have − −

a 1 X ,Y d b . c ad bc = Œ ‘ = ‰ − Ž − a 2 c 2 b 2 d 2 XY csc θ. ad bc ‰S S + S S Ž ‰S S + S S Ž Y Y = = Thus, as ε 0, we have fromS Theorem− S 4.1:

B λ, csc→ θ ε o ε σε A B λ, csc θ ε o ε .

So, ( ( ) + ( )) ⊆ ( ) ⊆ ( ( ) + ( )) r csc θ ε o ε max . rmin csc θ ε o ε ( ) + ( ) = Note as ε 0, that this tends( to) 1+ faster( ) than ε. This agrees with the ratio we obtain from the explicit formula for diag- onalizable 2 →2 matrices. However, the formula for the 2 2 matrix gives us slightly more information on the o ε term. × × ( ) 20 GONG, MEYERSON, MEZA, WARD, STOICIU

1 r y y2 4ε2 4yε csc θ max 2 1 » r = Š −y 4ε2+ 4yε−csc θ y  min 2 » = Š + + −  2 rmax 1 1 4 ε y 4 ε y csc θ rmin 1»4 ε y 2 4 ε y csc θ 1 − + ( ~ ) − ( ~ ) = »2 ε y csc θ 2 ε y 2 2 ε y 2 csc2 θ o ε3 + ( ~ ) + ( ~ ) − 2 ε y csc θ 2 ε y 2 2 ε y 2 csc2 θ o ε3 ( ~ ) − ( ~ ) + ( ~ ) + ( ) = csc θ ε y csc2 θ 1 o ε2 ( ~ ) + ( ~ ) − ( ~ ) . + ( ) csc θ ε y csc2 θ 1 o ε2 + ( ~ )[ − ] + ( ) = − ( ~ )[ − ] + ( )

5. Pseudospectra of N N Jordan Block

As the Jordan blocks are the fundamental structural× elements of finite dimensional linear operators, it is especially important to understand their pseudospectra. Proposition 5.1 (Lower Bound). Let J be an N N Jordan block where

λ 1 × λ 1 J ⎛ ⎞ . ⎜ 1⎟ ⎜ ⎟ = ⎜ ⋱ ⋱ ⎟ ⎜ λ⎟ ⎜ ⋱ ⎟ ⎝ ⎠ Then, N B λ, ε σε J . √ Proof. By induction. This is clear( for N ) ⊆1; for( N) 2, we will show that

B λ, ε = σε J . = √ Let z B λ, ε . Write z λ ω,( where) ω⊆ ( ε). Define √ 0 0 √ ∈ ( ) = + E S S < . ω2 0

Note E ω 2 ε. Observe = Œ ‘ λ z 1 ω 1 SS SS = S detS

ω εN+1. Let E be a matrix with ωN+1 in the bottom-left corner entry and zeros elsewhere. Note that E ω N+1 ε. Then, S S < N+1 N N+1 N+1 N+1 N det J E zI Y Y ω= S S < 1 ω ω 1 1 0.

Thus, z (σ +J −E ,) and= (− so )z σε+J(−. ) = ‰(− ) + (− ) Ž = 

∈ ( + ) ∈ ( ) From the second definition of the pseudospectrum, using a perturbation matrix, we can get a better explicit lower bound on the ε-pseudospectra of an N N Jordan block. × Proposition 5.2 (Better Lower Bound). Let J be an N N Jordan block. Then,

N N−1 B λ, ε 1 ε σε J . × » Š ( + )  ⊆ ( ) Proof. We use the second definition for σε J . Let

0 k ( ) 0 k E ⎛ ⎞ ⎜ k⎟ ⎜ ⎟ = ⎜ ⋱ ⋱ ⎟ ⎜k 0⎟ ⎜ ⋱ ⎟ where k ε, and note that E ⎝ε. We take det ⎠J E zI and set it equal to zero to find the eigenvalues of J E. S S < Y Y < ( + − ) 0 det+ J E zI λ z k 1 = ( + − ) λ z k 1 det ⎛ − + ⎞ ⎜ − + k 1⎟ ⎜ ⎟ = ⎜ ⋱ ⋱ ⎟ ⎜ k λ z⎟ ⎜ ⋱ + ⎟ λ z N 1 N−1k 1 k N−1 ⎝ − ⎠ 1 N−1 z λ N k 1 k N−1 ; = ( − ) + (− ) ( + ) z λ N k 1 k N−1 = (− ) (( − ) + ( + ) ) N N−1 ⇐⇒ ( −z ) λ = ( k+1 ) k . » − = ( + ) N N−1 So, B λ, ε 1 ε σε J .  » Š ( + )  ⊆ ( ) From [14, pg. 470], we know that the ε-pseudospectrum of the Jordan block is a perfect disk about the eigenvalue of J of some radius. Although we do not find the explicit radius, we can use Theorem 4.1 to find the asymptotic behavior of σε J ,

Proposition 5.3 (Asymptotic Bound). Let J be an N N Jordan block. Then( ) 1~N 1~N σε J B λ, ε o ε .×

( ) ≈ ( + ( )) 22 GONG, MEYERSON, MEZA, WARD, STOICIU

Proof. The N N Jordan block has left and right eigenvectors uj and vj where uj 1 and vj 1. So, from Theorem 4.1, we find C XY vjuj 1. Thus, × 1~N 1~N Y Y = Y Y = σε J B λ, ε o ε =. Y Y = Y Y = ( ) ≈ ( + ( )) 

However, we can find the perturbation matrix that gives us the boundary of σε J .

Note: Let J be an n n Jordan block and E be the perturbation matrix ( )

× ε

E ⎛ ⎞ . ⎜ ⋰ ⎟ ⎜ ⎟ = ⎜ ⋰ ⎟ ⎜ε ⎟ ⎜ ⎟ We know that the ε-pseudospectrum of⋰ any Jordan block is a perfect disk around ⎝ ⎠ the eigenvalue. We observe that as we range θ from 0 to 2π, σ A eiθE numerically traces the boundary of σε J for ε 0. ( + ) For example, suppose J is( a) 5 5> Jordan block with 0 as an eigenvalue. Letting ε .025, we get through eigtool that the pseudospectral radius (defined as the farthest point on the boundary× of pseudo spectrum from 0) is 0.50765. We now use Mathematica= to compute the eigenvalue of J E and obtain 0.50765 as well.

( + ) 6. Pseudospectra of bidiagonal matrices

The Jordan block is a simple example of a bidiagonal matrix. We now extend our application of Theorem 4.1 to all bidiagonal matrices.

6.1. Asymptotic Bounds for Periodic Bidiagonal Matrices.

We investigate various classes of bidiagonal matrices. We describe the asymptotic behavior of the ε-pseudospectrum for the general periodic bidiagonal case, namely A is an N N matrix with period k on the main diagonal and nonzero superdiagonal entries × a1 b1

⎛ ak ⎞ ⎜ ⋱ ⋱ ⎟ ⎜ ⎟ A ⎜ ⎟ ⎜ ⋱ ⎟ ⎜ ⎟ ⎜ ⋱ ⋱ a1 ⎟ = ⎜ ⎟ ⎜ ⋱ ⋱ b − ⎟ ⎜ N 1⎟ ⎜ ⋱ ⎟ ⎜ ar ⎟ We have from Theorem 4.1,⎜ that ⋱ ⎟ ⎝ ⎠ k 1 n σε A B aj, Cjε j , j=1 ( ) ≈ ‹ ( )  EXPLICIT BOUNDS FOR THE PSEUDOSPECTRA OF MATRICES AND OPERATORS 23 where nj is the size of the Jordan block corresponding to aj and Cj XjYj . For matrices that are of the bidiagonal form, the geometric multiplicity of each eigenvalue is 1, meaning that all the eigenvalues fall in Case 3 of Theorem= Y 4.1.Y Thus, the constant Cj that multiplies any eigenvalue aj is simply Cj vj uj , where vj and uj denote the right and left eigenvectors, respectively. = Y YY Y We will begin by introducing ε-pseudospectrum for simple special cases which lead to the most general case.

The cases will be presented as follows:

(1) Let A be a kn kn matrix with a1, . . . , ak distinct and bi 1 for all i. (2) Let A be a kn kn matrix with a1, . . . , ak distinct. (3) Let A be an N× N matrix with a1, . . . , ak distinct. = (4) General Case:× Let A be an N N matrix with a1, . . . , ak not distinct. × We define: × x, x 0 f x . ⎧1, x 1 ⎪ ≠ ( ) = ⎨ ⎪ = Case 1: ⎩⎪

The size of A is kn kn. The ai’s are distinct. ● bi 1 for all i. × ● So, A●is of= the form:

a1 1

⎛ ak ⎞ ⎜ ⋱ ⋱ ⎟ ⎜ ⎟ A ⎜ ⎟ . ⎜ ⋱ ⎟ ⎜ ⎟ ⎜ ⋱ ⋱ a1 ⎟ = ⎜ ⎟ ⎜ ⋱ ⋱ 1 ⎟ ⎜ ⎟ ⎜ ⋱ ⎟ ⎜ ak⎟ ⎜ ⋱ ⎟ ⎝ ⎠ To find the asymptotic bound for the resolvent norm of A around each eigenvalue, we must determine the right and left eigenvectors associated with each eigenvalue. We claim: 1 f(aj −aj−1)⋯f(aj −a1) 1 ⎛ f(aj −aj−1)⋯f(aj −a2) ⎞ ⎜ ⎟ ⎜ ⎟ ⎜ 1 ⎟ vj ⎜ ⎟ , ⎜ f(aj −aj−1) ⎟ ⎜ ⋮ ⎟ ⎜ ⎟ ⎜ 1 ⎟ = ⎜ ⎟ ⎜ 0 ⎟ ⎜ ⎟ ⎜ ⎟ ⎜ ⎟ ⎜ ⎟ ⎜ 0 ⎟ ⎜ ⋮ ⎟ ⎝ ⎠ 24 GONG, MEYERSON, MEZA, WARD, STOICIU

0

⎛ 0 ⎞ ⎜ ⋮ ⎟ ⎜f aj+1 aj f ak aj ⎟ f a1 aj f aj−1 aj ⎜ ⎟ uj ⎜ ⎟ , f a a f a a f a a f a a n ⎜f aj+2 aj f ak aj ⎟ 1 j j−1 j j+1 j k j ⎜ ( − )⋯ ( − )⎟ ( − )⋯ ( − ) ⎜ ⎟ = ⎜ ⎟ ⎜ ( − )⋯ ( − )⎟ [ ( − )⋯ ( − ) ( − )⋯ ( − )] ⎜ ⎟ ⎜ f ak aj ⎟ ⎜ ⋮ ⎟ ⎜ ⎟ ⎜ 1 ⎟ ⎜ ( − ) ⎟ which can be verified by direction computation of the left⎝ and right eigenvectors⎠ for each eigenvalue a1, . . . , ak.

We will check that vj is a right eigenvector for eigenvalue aj by cases.

First observe that for i j, we have vj i 0. Therefore,

Avj >i Aii vj i (Ai)(i+=1) vj i+1 0 aj vj i.

Now suppose i j(. Then,) = we have( ) + ( ) = = ( )

≤Avj i Aii vj i Ai(i+1) vj i+1 ai vj i vj i+1.

For 1 i j 2,( we get) = ( ) + ( ) = ( ) + ( )

≤ ≤ − 1 1 Avj i ai aj aj−1 aj ai aj aj−1 aj ai+1

( ) = ⋅ ai aj ai + ( − )⋯( − ) ( − )⋯( − ) aj aj−1 aj ai + ( − ) = aj vj i. ( − )⋯( − ) For i j 1, we get= ( )

= − 1 Avj j−1 aj−1 1 aj vj j−1. aj aj−1 ( ) = ⋅ + = ( ) For i j we get −

= Avj j aj vj j 0.

Thus, in all cases we see that Av( j ) a=jvj,( and) + so vj is indeed a right eigenvector ∗ for the eigenvalue aj. The proof that uj is a left eigenvector for eigenvalue aj is similar. =

Case 2:

The size of A is kn kn. The ai’s are distinct. ● × Since● the superdiagonal of A is not constant, we now write the elements of the superdiagonal as b1, b2, . . . , bN−1. EXPLICIT BOUNDS FOR THE PSEUDOSPECTRA OF MATRICES AND OPERATORS 25

We claim: b1⋯bj−1 f(aj −a1)⋯f(aj −aj−1) b2⋯bj−1 ⎛ f(aj −a2)⋯f(aj −aj−1) ⎞ ⎜ ⎟ ⎜ ⎟ ⎜ ⎟ ⎜ bj−1 ⎟ vj ⎜ ⎟ , ⎜ f(aj −⋮aj−1) ⎟ ⎜ ⎟ ⎜ ⎟ = ⎜ 1 ⎟ ⎜ ⎟ ⎜ 0 ⎟ ⎜ ⎟ ⎜ ⎟ ⎜ ⎟ ⎜ 0 ⎟ ⎜ ⎟ ⎜ ⋮ ⎟ u∗ 0, 0, y, bp y,⎝ bp⋯bN−⎠2 y, bp⋯bN−1 y , j f(aj+1−aj ) f(aj+1−aj )⋯f(ak−1−aj ) f(aj+1−aj )⋯f(ak−aj ) = Š ⋯ ⋯  where p k n 1 j, and b b = ( −y ) + j p−1 . n−1 a1 aj aj−1 aj aj+1 aj ak aj ⋯ = Again, direct computation(( − will)⋯( show− that)( these are− indeed)⋯( left− and)) right eigenvectors associated with any eigenvalue aj. Case 3:

The size of A is N N. The ai’s are distinct. ● × We now● drop our assumption that the size of our matrix is kn kn, for period k on the diagonal. Let n, r be such that N kn r, where 0 r k. In other words, ar is the last entry on the main diagonal, so the period does not necessarily× complete. = + < ≤ For aj, the right eigenvector is given by

b1⋯bj−1 f(aj −a1)⋯f(aj −aj−1) b2⋯bj−1 ⎛ f(aj −a2)⋯f(aj −aj−1) ⎞ ⎜ ⎟ ⎜ ⎟ ⎜ ⎟ ⎜ bj−1 ⎟ vj ⎜ ⎟ . ⎜ f(aj −⋮aj−1) ⎟ ⎜ ⎟ ⎜ ⎟ = ⎜ 1 ⎟ ⎜ ⎟ ⎜ 0 ⎟ ⎜ ⎟ ⎜ ⎟ ⎜ ⎟ ⎜ 0 ⎟ ⎜ ⎟ ⎜ ⋮ ⎟ ⎝ ⎠ We split up the formula for the left eigenvectors into two cases:

(1) 1 j r (2) r j k. ≤ ≤ < ≤ 26 GONG, MEYERSON, MEZA, WARD, STOICIU

(1) 1 j r On the main diagonal, there are n complete blocks with entries a1, . . . , ak, and one partial block≤ ≤ at the end with entries a1, . . . ar. In the first case, when 1 j r, then aj is in this last partial block. In this case then, let p kn j. ≤ ≤ We claim = + 0

⎛ 0 ⎞ ⎜ ⋮ ⎟ ⎜ f aj+1 aj f ar aj ⎟ ⎜ ⎟ uj µj ⎜ ⎟ , ⎜bpf aj+2 aj f ar aj ⎟ ⎜ ( − )⋯ ( − ) ⎟ ⎜ ⎟ = ⎜ ⎟ ⎜ ( − )⋯ ( − )⎟ ⎜ ⎟ ⎜ bp bN−2f ar aj ⎟ ⎜ ⋮ ⎟ ⎜ ⎟ ⎜ bp bN−1 ⎟ ⎜ ⋯ ( − ) ⎟ ⎝ ⋯ ⎠ bj ⋯bp−1f(a1−aj )⋯f(aj−1−aj ) where µj n . [f(a1−aj )⋯f(aj−1−aj )f(aj+1−aj )⋯f(ak−aj )] f(a1−aj )⋯f(ar −aj ) (2) r= j k In this case, aj is in the last complete block. Now, let p k n 1 j. < ≤ Now, we claim = ( − ) + 0

⎛ 0 ⎞ ⎜ ⋮ ⎟ ⎜ f a1 aj f ar aj f aj+1 aj f ak aj ⎟ ⎜ ⎟ ⎜ ⎟ ⎜bpf a1 aj f ar aj f aj+2 aj f ak aj ⎟ ⎜ ( − )⋯ ( − ) ( − )⋯ ( − ) ⎟ ⎜ ⎟ ⎜ ⎟ uj µj ⎜ ( − )⋯ ( − ) ( − )⋯ ( − )⎟ , ⎜ b b f a a f a a ⎟ ⎜ p p+k−j−1 1 j r j ⎟ ⎜ ⋮ ⎟ = ⎜ ⎟ ⎜ bp bp+k−jf a2 aj f ar aj ⎟ ⎜ ⋯ ( − )⋯ ( − ) ⎟ ⎜ ⎟ ⎜ ⎟ ⎜ ⋯ ( − )⋯ ( − ) ⎟ ⎜ bp bN−2f ar aj ⎟ ⎜ ⎟ ⎜ ⋮ ⎟ ⎜ b b ⎟ ⎜ p N−1 ⎟ ⎜ ⋯ ( − ) ⎟ ⎝ bj ⋯bp−1f(a1⋯−aj )⋯f(aj−1−aj ) ⎠ again where µj n [f(a1−aj )⋯f(aj−1−aj )f(aj+1−aj )⋯f(ak−aj )] f(a1−aj )⋯f(ar −aj ) Case 4: General= Case.

The size of A is N N. The ai’s are not distinct for 1 i k ● × Let A●be a N N periodic bidiagonal≤ matrix≤ with period k on the main diagonal. Let n, r be such that N kn r, where 0 r k. Write a1, . . . , ak for the entries on the main diagonal× (ai’s not distinct) and b1, . . . , bN−1 for the entries on the superdiagonal. Let ar be= the last+ entry on< the≤ main diagonal. EXPLICIT BOUNDS FOR THE PSEUDOSPECTRA OF MATRICES AND OPERATORS 27

We can explicitly find the left and right eigenvectors for any eigenvalue, α. Sup- pose α first appears in position ` of the period k. Then the corresponding right eigenvector for α is the same form as v` in Case 4. That is,

b1⋯b`−1 f(a`−a1)⋯f(a`−a`−1) b2⋯b`−1 ⎛ f(a`−a2)⋯f(a`−a`−1) ⎞ ⎜ ⎟ ⎜ ⎟ ⎜ b`−1 ⎟ v` ⎜ ⎟ . ⎜ f(a`−a`−1) ⎟ ⎜ ⋮ ⎟ ⎜ ⎟ ⎜ 1 ⎟ = ⎜ ⎟ ⎜ 0 ⎟ ⎜ ⎟ ⎜ ⎟ ⎜ ⎟ ⎜ ⎟ ⎜ 0 ⎟ ⎜ ⋮ ⎟ ⎝ ⎠ The corresponding left eigenvector for α depends on the first and last positions of α. Let q m mod k . We split up the formula for the left eigenvector of α into two cases, which again mirror the formulas given in Case 4: ≡ ( ) (1) 1 ` r (2) r ` k. ≤ ≤ For both of< these≤ two cases, we define

bi, i p f bi . ⎧1, i p ⎪ ≥ ( ) = ⎨ ⎪ < (1) 1 ` r ⎩ In this case then, α appears in the partial block. Let p kn `. ≤ ≤ We claim = + 0

⎛ 0 ⎞ ⎜ ⋮ ⎟ ⎜ f bp+m−`−1 f am+1 a` f ar a` ⎟ ⎜ ⎟ u` µ` ⎜ ⎟ , ⎜f bp+m−`−1 f bp+m−l f am+2 a` f ar a` ⎟ ⎜ ( ) ( − )⋯ ( − ) ⎟ ⎜ ⎟ = ⎜ ⎟ ⎜ ( ) ( ) ( − )⋯ ( − )⎟ ⎜ ⎟ ⎜ f bp+m−`−1 f bN−2 bN−2f ar a` ⎟ ⎜ ⋮ ⎟ ⎜ ⎟ ⎜ f bp+m−`−1 f bN−1 ⎟ ⎜ ( )⋯ ( )⋯ ( − ) ⎟ ⎝ ( )⋯ ( ) ⎠ b`⋯bp−1f(a1−a`)⋯f(a`−1−a`) where µ` n . [f(a1−a`)⋯f(a`−1−aj )f(a`+1−a`)⋯f(ak−a`)] f(a1−a`)⋯f(ar −a`) (2) r= j k In this case, α is in the last complete block. Here, we let p k n 1 ` < ≤ = ( − ) + 28 GONG, MEYERSON, MEZA, WARD, STOICIU

Now, we claim 0

⎛ 0 ⎞ ⎜ ⋮ ⎟ ⎜ f bp+m−`−1 f a1 a` ar a` f am+1 a` f ak a` ⎟ ⎜ ⎟ ⎜ ⎟ ⎜f bp+m−`−1 f bp+m−l f a1 a` f ar a` f am+2 a` f ak a` ⎟ ⎜ ( ) ( − )⋯( − ) ( − )⋯ ( − ) ⎟ ⎜ ⎟ ⎜ ⎟ u` µ` ⎜ ( ) ( ) ( − )⋯ ( − ) ( − )⋯ ( − )⎟ , ⎜ f b f b f a a f a a ⎟ ⎜ p+m−`−1 p+k−`−1 1 ` r ` ⎟ ⎜ ⋮ ⎟ = ⎜ ⎟ ⎜ f bp+m−`−1 f bp+k−` f a2 a` f ar a` ⎟ ⎜ ( )⋯ ( ) ( − )⋯ ( − ) ⎟ ⎜ ⎟ ⎜ ⎟ ⎜ ( )⋯ ( ) ( − )⋯ ( − ) ⎟ ⎜ f bp+m−`−1 f bN−2 f ar a` ⎟ ⎜ ⎟ ⎜ ⋮ ⎟ ⎜ f b f b ⎟ ⎜ p+m−`−1 N−1 ⎟ ⎜ ( )⋯ ( ) ( − ) ⎟ where ⎝ ( )⋯ ( ) ⎠ b b f a a f a a µ ` p−1 1 ` `−1 ` . ` n f a1 a` f a`−1 a` f a`+1 aj f ak a` f a1 a` f ar a` ⋯ ( − )⋯ ( − ) = From these[ ( formulas,− )⋯ ( we can− find) ( the− eigenvectors,)⋯ ( − and)] hence( − the) asymptotic⋯ ( − ) be- havior of the ε-pseudospectrum for any bidiagonal matrix,

k 1 n σε A B aj, Cjε j j=1 ( ) ≈ ‹ ( )  where Cj vj uj and nj is the size of the Jordan block corresponding to aj.

Note 1: =LetY YYA beY a periodic, bidiagonal matrix and suppose bi 0 for some i. Then the matrix decouples into the direct sum of smaller matrices, call them A1,...,An. To find the ε-pseudospectrum of A, apply the same analysis= to these smaller matrices, and from Theorem 2.2, we have that n σε A σε Ai . i=1 ( ) = ( ) Note 2: It is also interesting that given the results in Case 4, we can find different matrices with the same asymptotic ε-pseudospectra. Based on the formulas for the right and left eigenvectors, we observe that we can change the arrangement of the period without changing the left and right eigenvectors for any eigenvalue if the following hold:

(1) The algebraic multiplicity of each ai remains the same. (2) The first position of each distinct eigenvalue remains the same. Addition- ally, suppose α first appears in position `. We can permute the previous entries in positions 1, 2, . . . , ` 1 as long as we fix the number of times each distinct entry appears. (3) The last position of each distinct− eigenvalue remains the same in a period. Additionally, suppose α last appears in position q on the main diagonal. Let q m mod k . We can permute the other entries in positions m 1, . . . , k as long as we fix the number of times each distinct entry appears. ≡ ( ) + EXPLICIT BOUNDS FOR THE PSEUDOSPECTRA OF MATRICES AND OPERATORS 29

Example 6.1. Let matrix A have period 13 on the main diagonal, with distinct eigenvalues a, b, c. Let A have period a, b, a, b, c, a, b, c, a, b, c, b, c . Note that the first entries of a, b, c are fixed in positions 1, 2, 5 respectively and the { } last entries are fixed in positions 9, 12, 13 respectively.

We can permute entries 3, 4 since permuting these entries does not change the num- ber of a or b entries before the first entry of c. We can also permute entries 6, 7, 8, as well as the entries 10, 11 for similar reasons. We now give possible permutations of said entries:

A1 has period a, b, b, a, c, a, b, c, a, b, c, b, c , permuting entries 3 and 4, A2 has period a, b, a, b, c, b, c, a, a, b, c, b, c , permuting entries 6, 7, 8, ● A3 has period {a, b, b, a, c, a, b, c, a, c, b, b, c}, permuting entries 10 and 11. ● { } With● these possible permutations,{ }

σε A σε A1 σε A2 σε A3

We will now present examples( ) ≈ of bidiagonal( ) ≈ ( matrices) ≈ ( with) explicit and asymptotic bounds.

6.2. Examples of Bidiagonal Matrices. Example 6.2. Let A be an n n bidiagonal matrix, a b 0 0 ... × 1 0 a b2 0 ... A ⎛ ... ⎞ ⎜ b ⎟ ⎜ n−1⎟ = ⎜ ⋮ ⋮ ⋱ ⋱ ⎟ ⎜0 0 0 0 a ⎟ ⎜ ⋮ ⋮ ⋮ ⋱ ⎟ ⎝ ⎠ where bi 0. The following propositions give explicit lower and upper bounds for the ε-pseudospectrum of A. ≠ Proposition 6.1 (Lower Bound).

1 B a, ε b1b2 bn−1 n σε A .

Š ( S ⋯ S)  ⊆ ( ) 1 Proof. Let z B a, ε b1b2 bn−1 n . We will show that z σε A . Let E be an n n matrix with ε in the lower-left corner and 0’s elsewhere. Then E ε. We take the determinant∈ Š ( ofS A⋯ E S)z and set it equal to 0 to∈ get( the) eigenvalues of A+E.× Y Y ≤ + − n n−1 0 A E z λ z 1 εb1b2 bn−1 n n λ z 1 εb b b − = S + − S = ( − ) +1(−2 ) n 1 ⋯ 1 n Ô⇒ (z −λ ) εb= 1(b−2 )bn−1 ⋯. Thus, z σ A . ε Ô⇒ = + ( ⋯ )  ∈ ( ) 30 GONG, MEYERSON, MEZA, WARD, STOICIU

Proposition 6.2 (Upper Bound). The Jordan decomposition of A is

1 a 1 1 1 b1 a 1 b1 1 A ⎛ b b ⎞ ⎛ a ⎞ ⎛ b1b2 ⎞ . ⎜ 1 2 ⎟ ⎜ ⎟ ⎜ 1⎟ ⎜ ⎟ ⎜ 1 ⎟ ⎜ ⎟ ⎜ ⎟ = ⎜ ⎟ ⎜ ⋱ a⎟ ⎜ b1 bn−1⎟ ⎜ b1⋯bn−1 ⎟ ⎜ ⎟ ⎜ ⎟ ⎜ ⋱ ⎟ ⎜ ⋱ ⎟ ⎜ ⋱ ⎟ ⎝ ⎠ ⎝ ⎠ ⎝ ⎠1 1~n ⋯ Thus σε A B a, b1 bn−1 ε n o ε .

( ) ⊆ Š (S ⋯ S ) + ‰ Ž Combining this upper bound with the lower bound from Prop. 6.1, we get that 1 asymptotically, σε A B a, b1 bn−1 ε n .

Note that if bi (0 for) ≈ someŠ (Si, then⋯ theS ) matrix decouples, such that its Jordan 1 1~n decomposition has multiple blocks and σε A B a, bk bj ε n o ε where = bk, bk+1, . . . bj correspond to the largest Jordan block. ( ) ⊆ Š (S ⋯ S ) + ‰ Ž

We now find bounds for a bidiagonal matrix of period 2 on the main diagonal.

Example 6.3. Restricting to period 2 on the main diagonal and constants on the superdiagonal, we can get explicit lower bounds for our pseudospectra.

Let A be a 2N 2N bidiagonal matrix with period 2 on the main diagonal, such that ×

a1 1 a2 1 A ⎛ ⎞ . ⎜ a 1 ⎟ ⎜ 1 ⎟ = ⎜ ⋱ ⋱ ⎟ ⎜ a2⎟ ⎜ ⎟ Proposition 6.3 (Lower Bound)⎝ . ⎠

2 1 1 B aj, ε N σε A . j=1 a1 a2 Œ Š ‘ ⊆ ( ) S − S

Proof. Fix θ 0, 2Nπ , and let E be the perturbation matrix with εeiθ on the lower left corner. ∈ [ ) N N Then, det A E z 0 a1 z a2 z ε 0. Solving for z, we find that

( + − ) = ⇒a ( a− ) (a −a ) 2 − 4 =εeiθ 1~N z 1 2 1 2 . » 2 + ± ( − ) + ( ) = EXPLICIT BOUNDS FOR THE PSEUDOSPECTRA OF MATRICES AND OPERATORS 31

Around the eigenvalue λ1 as ε 0, we have a a a a 2 4 εeiθ 1~N z 1 →2 1 2 » 2 + + ( − ) + ( )1 1~2 = iθ N a a a a 4 εe 1 2 1 2 1 2 2 ⎡ a a 2 ⎤ ⎢ 1 2 ⎥ + − ⎢ ‰ Ž ⎥ = + ⎢ + 1 ⎥ ⎢ N ⎥ a a a a ⎢ (2 εe−iθ ) ⎥ 1 2 1 2 ⎣1 ⎦ 2 2 ⎡ a a 2 ⎤ ⎢ 1 2 ⎥ + − ⎢ ‰ Ž ⎥ ≈ + ⎢ + ⎥ iθ 1~N ⎢ ⎥ εe ⎢ ( − ) ⎥ a1 . ⎣ ⎦ a1 a2 ‰ Ž 1 ε N iθ~N= + So, z a1 e . − Sa1−a2S

PerformingS − S ≈ the same calculation for the eigenvalue a2, we find that

1 ε N iθ~N z a2 e . a1 a2 S − S ≈ 1 Ranging θ over 0, 2Nπ , we get that theS − disksS of radius ε N , centered around Sa1−a2S a1 and a2 are contained in our ε-pseudospectrum. In other words, [ ) 2 1 1 B aj, ε N σε A . j=1 a1 a2 Œ Š ‘ ⊆ ( ) S − S 

We also have an asymptotic upper bound for σε A :

1 N 1 ε ( ) 2 2N z aj 1 a1 a2 a1 a2 S − S ≤ ‰ + S − S Ž 1 2 2N and the ratio of the upper and lowerS − boundsS is just 1 a1 a2 .

Proposition 6.4 (Asymptotic Lower Bound). ‰ + S − S Ž 2 1 1 N 1 B a , 1 a a 2ε o ε N σ A . j a a 1 2 ε j=1 1 2 » Œ Š + S − S  + ( )‘ ⊆ ( ) S − S Proof. From Theorem 4.1, we can find an asymptotic lower bound for the bidiagonal matrix of period 2.

We Jordan decompose matrix A such that QAP J where QP I. Then, we th multiply the first column of P , call it x1, by the N row of Q, call it y1 to get the constant that multiplies a1. We find that = = 1 x y 1 a a 2. 1 1 a a N 1 2 1 2 » Y Y = + S − S S − S 32 GONG, MEYERSON, MEZA, WARD, STOICIU

th For a2, we find x2y2 by multiplying the N column of P with the last row of Q, and get that x2y2 x1y1 . So, as ε 0 Y Y = Y Y 2 1 1 N 1 → B a , 1 a a 2ε o ε N σ A j a a 1 2 ε j=1 1 2 » Œ Š + S − S  + ( )‘ ⊆ ( ) S − S 

This lower bound differs from the lower bound given in Proposition 6.3 by a con- 2 1 1 stant, α 1 a1 a2 N and some o ε N . » = ( + S − S ) ( ) 7. Tridiagonal Matrices

The natural extension to bidiagonal matrices is to consider tridiagonal matrices. The study of tridiagonal matrices arises from a quantum mechanics problem known as the Anderson model, as introduced by Anderson [1]. Hatano, Nelson, and Shnerb [?] studied the non-Hermitian analog of the Anderson model, now known as the Hatano-Nelson model, which takes the form g −g a1 e e −g g e a2 e A ⎛ ⎞ ⎜ ⎟ ⎜ −g g ⎟ ⎜ e a − e ⎟ = ⎜ ⋱ ⋱N ⋱ 1 ⎟ ⎜ g −g ⎟ ⎜ e e aN ⎟ ⎜ ⎟ where g is a fixed real parameter⎝ and a1, . . . aN are iid random⎠ variables. We are interested in the nonperiodic Hatano-Nelson matrix g a1 e −g g e a2 e A ⎛ ⎞ ⎜ ⎟ ⎜ −g g ⎟ ⎜ e a − e ⎟ = ⎜ ⋱ ⋱N ⋱ 1 ⎟ ⎜ −g ⎟ ⎜ e aN ⎟ ⎜ ⎟ ⎝ ⎠ In this case, we can use a diagonal similarity transformation of D diag 1, eg, e2g, . . . , e(N−1)g to get that the nonperiodic Hatano-Nelson matrix is similar to the matrix = ( ) a1 1 1 a2 1 DAD−1 ⎛ ⎞ ⎜ ⎟ ⎜ ⎟ ⎜ 1 a − 1 ⎟ = ⎜ ⋱ ⋱N ⋱ 1 ⎟ ⎜ ⎟ ⎜ 1 aN ⎟ ⎜ ⎟ This matrix will only have one Jordan⎝ block per eigenvalue⎠ [7]. In fact, since this matrix is symmetric and tridiagonal with nonzero off-diagonal entries, then we know it has N distinct eigenvalues [12, 6]. EXPLICIT BOUNDS FOR THE PSEUDOSPECTRA OF MATRICES AND OPERATORS 33

For the case when a a1 a2 ... aN , then A is Toeplitz, and it is known that kπ the eigenvalues are of the form λk a 2 cos N+1 and eigenvectors take the form (k) kjπ = = = = uj sin n+1 [6, 15]. The Toeplitz case has been studied [11] and the sensitivity of the eigenvalues investigated; however= + by‰ findingŽ the eigenvectors, we can give asymptotic= results on the pseudospectrum.

For the case when the main diagonal is periodic with period 2, i.e.

a1 if j odd aj ⎧a2 if j even ⎪ = ⎨ then we can find an explicit formula⎪ for the eigenvalues and eigenvectors [9, 6]. ⎩⎪ With these formulae, we can then find an asymptotic upper bound on σε A , as we did for bidiagonal matrices. ( ) Further research would attempt to generalize these results to period k on the main diagonal, and then ultimately to arbitrary diagonal entries.

8. Finite Rank operators

The majority of this paper has focused on both explicit and asymptotic characteri- zations of ε-pseudospectra for various classes of finite dimensional linear operators. A natural next step is to consider finite rank operators on infinite dimensional space.

In section 2 we defined ε-pseudospectra for matrices, although our definitions are exactly the same in the infinite dimensional case. For our purposes, the only note- worthy difference between matrices and operators is that the spectrum of an oper- ator is no longer defined as the collection of eigenvalues, but rather σ A λ λI A does not have a bounded inverse As a result, we do not get the same properties for pseudospectra as we did previ- ( ) = { S − } ously; in particular, σε A is not necessarily bounded.

That being said, the following( ) theorem shows that finite rank operators behave similarly to matrices, in that asymptotically the radii ε-pseudospectra are bounded by roots of epsilon. The following theorem makes this precise. Theorem 8.1. Let V be a and A V V a finite rank operator on H. Then there exists C such that for sufficiently small ε, ∶ 1→ σε A σ A B 0, Cε m+1 . where m is the rank of A. Furthermore, this bound is sharp in the sense that there ( ) ⊆ ( ) + ( ) exists a rank-m operator A and a constant c such that

1 σε A σ A B 0, cε m+1 for sufficiently small ε. ( ) ⊇ ( ) + ( )

Proof. Since A has finite rank, there exists a finite dimensional subspace U such that V U W and A U U, and A W 0 . Choosing an orthonormal basis

= ⊕ ( ) ⊆ ( ) = { } 34 GONG, MEYERSON, MEZA, WARD, STOICIU for A which respects this decomposition we can write A A′ 0. Then the spectrum of A is σ A′ 0 , and we know that for any ε, ′ = ⊕ ( ) ∪ { } σε A σε A σε 0 .

The ε-pseudospectrum of the zero( ) operator= ( ) is∪ well-understood( ) since this operator is normal; for any ε, it is precisely the ball of radius ε. It thus suffices to consider the ε-pseudospectrum of the finite rank operator A′ U U, where U is finite dimensional. The ε-pseudospectrum of this operator goes like ε1~j, where j is the dimension of the largest Jordan block; we will prove that∶ j→ m 1. Note that the rank of the n n Jordan block given by ≤ + × λ 1 0 0 ... 0 λ 1 0 ... A ⎛ ... ⎞ ⎜ b ⎟ ⎜ n−1⎟ = ⎜ ⋮ ⋮ ⋱ ⋱ ⎟ ⎜0 0 0 0 λ ⎟ ⎜ ⋮ ⋮ ⋮ ⋱ ⎟ is n if λ 0, and n 1 if λ 0.⎝ Since we know that⎠ the rank of A is larger than or equal to the rank of the largest Jordan block, we have an upper bound on the dimension≠ of the largest− Jordan= block: it is of size m 1, with equality attained when λ 0. By previous theorems, we then know that σε A is contained, for small 1 enough ε, in the set σ A Cε m+1 . + = ( ) m+1 Note that this bound is( sharp;) + we can see this by taking V to be R and consid- ering the rank-m operator 0 1 0 0 ... 0 0 1 0 ... Am ⎛ ...⎞ ⎜ 1 ⎟ ⎜ ⎟ = ⎜ ⋮ ⋮ ⋱ ⋱ ⎟ ⎜0 0 0 0 0 ⎟ ⎜ ⋮ ⋮ ⋮ ⋱ ⎟ the pseudospectrum of which will⎝ contain the ball of⎠ radius ε1~m+1 by proposition 5.1. 

The natural question to ask now is whether we can extend this result to more arbitrary operators on Hilbert spaces. For operators on Banach spaces, we do still have a certain convergence of the ε-pseudospectrum to the spectrum [14, §4], namely ε>0 σε A σ A . Also, while the ε-pseudospectrum may be unbounded, any bounded component of it necessarily contains a component of the spectrum. ∩ ( ) = ( ) From these results, we see that we can no longer get a complete asymptotic result for σε A , as there could be an unbounded component of σε A that has no intersection with the spectrum. For example, there exist operators with bounded spectrum and unbounded( ) ε-pseudospectra for all ε 0. ( )

However, the above result states that the> bounded components of the ε-pseudospectrum must converge to the spectrum. If we restrict our attention to these bounded com- ponents, we can generalize Thms. 4.1 and 8.1 by asking whether the bounded components of σε A converge to the spectrum as a union of disks.

( ) EXPLICIT BOUNDS FOR THE PSEUDOSPECTRA OF MATRICES AND OPERATORS 35

References

[1] P. W. Anderson. Absence of diffusion in certain random lattices. Phys. Rev., 109:1492–1505, Mar 1958. [2] Friedrich L Bauer and CT Fike. Norms and exclusion theorems. Numerische Mathematik, 2(1):137–141, 1960. [3] Hellmut Baumg¨artel. Analytic perturbation theory for matrices and operators, volume 17. Birkh¨auserBasel, 1985. [4] EB Davies. Pseudo–spectra, the harmonic oscillator and complex resonances. Proceedings of the Royal Society of London. Series A: Mathematical, Physical and Engineering Sciences, 455(1982):585–599, 1999. [5] EB Davies. Semi-classical states for non-self-adjoint schr¨odingeroperators. Communications in mathematical physics, 200(1):35–41, 1999. [6] Joseph Frederick Elliott. The characteristic roots of certain real symmetric matrices. Master’s thesis, University of Tennessee, 1953. [7] Carla Ferreira, Beresford Parlett, and Froil´anM Dopico. Sensitivity of eigenvalues of an unsymmetric . Numerische Mathematik, 122(3):527–555, 2012. [8] Tosio Kato. Perturbation theory for linear operators, volume 132. springer, 1995. [9] S Kouachi. Eigenvalues and eigenvectors of some tridiagonal matrices with non-constant diagonal entries. Applicationes mathematicae, 35(1):107, 2008. [10] Julio Moro, James V Burke, and Michael L Overton. On the lidskii–vishik–lyusternik pertur- bation theory for eigenvalues of matrices with arbitrary jordan structure. SIAM Journal on Matrix Analysis and Applications, 18(4):793–817, 1997. [11] Silvia Noschese, Lionello Pasquini, and Lothar Reichel. Tridiagonal toeplitz matrices: prop- erties and novel applications. Numerical Linear Algebra with Applications, 20(2):302–326, 2013. [12] Beresford N Parlett. The symmetric eigenvalue problem, volume 7. SIAM, 1980. [13] AVD Sluis. Condition numbers and equilibration of matrices. Numerische Mathematik, 14(1):14–23, 1969. [14] Lloyd Nicholas Trefethen and Mark Embree. Spectra and pseudospectra: the behavior of nonnormal matrices and operators. Princeton University Press, 2005. [15] W. C. Yueh. Eigenvalues of Several Tridiagonal Matrices. Appl. Math. E-Notes, 5:66–74, 2005.