Min-max theorem

Summary

In linear algebra and functional analysis, the min-max theorem, or variational theorem, or Courant–Fischer–Weyl min-max principle, is a result that gives a variational characterization of eigenvalues of compact Hermitian operators on Hilbert spaces. It can be viewed as the starting point of many results of similar nature.

This article first discusses the finite-dimensional case and its applications before considering compact operators on infinite-dimensional Hilbert spaces. We will see that for compact operators, the proof of the main theorem uses essentially the same idea from the finite-dimensional argument.

In the case that the operator is non-Hermitian, the theorem provides an equivalent characterization of the associated singular values. The min-max theorem can be extended to self-adjoint operators that are bounded below.

Matrices edit

Let A be a n × n Hermitian matrix. As with many other variational results on eigenvalues, one considers the Rayleigh–Ritz quotient RA : Cn \ {0} → R defined by

 

where (⋅, ⋅) denotes the Euclidean inner product on Cn. Clearly, the Rayleigh quotient of an eigenvector is its associated eigenvalue. Equivalently, the Rayleigh–Ritz quotient can be replaced by

 

For Hermitian matrices A, the range of the continuous function RA(x), or f(x), is a compact interval [a, b] of the real line. The maximum b and the minimum a are the largest and smallest eigenvalue of A, respectively. The min-max theorem is a refinement of this fact.

Min-max theorem edit

Let   be Hermitian on an inner product space   with dimension  , with spectrum ordered in descending order  .

Let   be the corresponding unit-length orthogonal eigenvectors.

Reverse the spectrum ordering, so that  .

(Poincaré’s inequality) — Let   be a subspace of   with dimension  , then there exists unit vectors  , such that

 , and  .

Proof

Part 2 is a corollary, using  .

  is a   dimensional subspace, so if we pick any list of   vectors, their span   must intersect   on at least a single line.

Take unit  . That’s what we need.

 , since  .
Since  , we find  .

min-max theorem — 

 
Proof

Part 2 is a corollary of part 1, by using  .

By Poincare’s inequality,   is an upper bound to the right side.

By setting  , the upper bound is achieved.

Counterexample in the non-Hermitian case edit

Let N be the nilpotent matrix

 

Define the Rayleigh quotient   exactly as above in the Hermitian case. Then it is easy to see that the only eigenvalue of N is zero, while the maximum value of the Rayleigh quotient is 1/2. That is, the maximum value of the Rayleigh quotient is larger than the maximum eigenvalue.

Applications edit

Min-max principle for singular values edit

The singular values {σk} of a square matrix M are the square roots of the eigenvalues of M*M (equivalently MM*). An immediate consequence[citation needed] of the first equality in the min-max theorem is:

 

Similarly,

 

Here   denotes the kth entry in the increasing sequence of σ's, so that  .

Cauchy interlacing theorem edit

Let A be a symmetric n × n matrix. The m × m matrix B, where mn, is called a compression of A if there exists an orthogonal projection P onto a subspace of dimension m such that PAP* = B. The Cauchy interlacing theorem states:

Theorem. If the eigenvalues of A are α1 ≤ ... ≤ αn, and those of B are β1 ≤ ... ≤ βj ≤ ... ≤ βm, then for all jm,
 

This can be proven using the min-max principle. Let βi have corresponding eigenvector bi and Sj be the j dimensional subspace Sj = span{b1, ..., bj}, then

 

According to first part of min-max, αjβj. On the other hand, if we define Smj+1 = span{bj, ..., bm}, then

 

where the last inequality is given by the second part of min-max.

When nm = 1, we have αjβjαj+1, hence the name interlacing theorem.

Compact operators edit

Let A be a compact, Hermitian operator on a Hilbert space H. Recall that the spectrum of such an operator (the set of eigenvalues) is a set of real numbers whose only possible cluster point is zero. It is thus convenient to list the positive eigenvalues of A as

 

where entries are repeated with multiplicity, as in the matrix case. (To emphasize that the sequence is decreasing, we may write  .) When H is infinite-dimensional, the above sequence of eigenvalues is necessarily infinite. We now apply the same reasoning as in the matrix case. Letting SkH be a k dimensional subspace, we can obtain the following theorem.

Theorem (Min-Max). Let A be a compact, self-adjoint operator on a Hilbert space H, whose positive eigenvalues are listed in decreasing order ... ≤ λk ≤ ... ≤ λ1. Then:
 

A similar pair of equalities hold for negative eigenvalues.

Proof

Let S' be the closure of the linear span  . The subspace S' has codimension k − 1. By the same dimension count argument as in the matrix case, S' Sk has positive dimension. So there exists xS' Sk with  . Since it is an element of S' , such an x necessarily satisfy

 

Therefore, for all Sk

 

But A is compact, therefore the function f(x) = (Ax, x) is weakly continuous. Furthermore, any bounded set in H is weakly compact. This lets us replace the infimum by minimum:

 

So

 

Because equality is achieved when  ,

 

This is the first part of min-max theorem for compact self-adjoint operators.

Analogously, consider now a (k − 1)-dimensional subspace Sk−1, whose the orthogonal complement is denoted by Sk−1. If S' = span{u1...uk},

 

So

 

This implies

 

where the compactness of A was applied. Index the above by the collection of k-1-dimensional subspaces gives

 

Pick Sk−1 = span{u1, ..., uk−1} and we deduce

 

Self-adjoint operators edit

The min-max theorem also applies to (possibly unbounded) self-adjoint operators.[1][2] Recall the essential spectrum is the spectrum without isolated eigenvalues of finite multiplicity. Sometimes we have some eigenvalues below the essential spectrum, and we would like to approximate the eigenvalues and eigenfunctions.

Theorem (Min-Max). Let A be self-adjoint, and let   be the eigenvalues of A below the essential spectrum. Then

 .

If we only have N eigenvalues and hence run out of eigenvalues, then we let   (the bottom of the essential spectrum) for n>N, and the above statement holds after replacing min-max with inf-sup.

Theorem (Max-Min). Let A be self-adjoint, and let   be the eigenvalues of A below the essential spectrum. Then

 .

If we only have N eigenvalues and hence run out of eigenvalues, then we let   (the bottom of the essential spectrum) for n > N, and the above statement holds after replacing max-min with sup-inf.

The proofs[1][2] use the following results about self-adjoint operators:

Theorem. Let A be self-adjoint. Then   for   if and only if  .[1]: 77 
Theorem. If A is self-adjoint, then

 

and

 .[1]: 77 

See also edit

References edit

  1. ^ a b c d G. Teschl, Mathematical Methods in Quantum Mechanics (GSM 99) https://www.mat.univie.ac.at/~gerald/ftp/book-schroe/schroe.pdf
  2. ^ a b Lieb; Loss (2001). Analysis. GSM. Vol. 14 (2nd ed.). Providence: American Mathematical Society. ISBN 0-8218-2783-9.

External links and citations to related work edit

  • Fisk, Steve (2005). "A very short proof of Cauchy's interlace theorem for eigenvalues of Hermitian matrices". arXiv:math/0502408. {{cite journal}}: Cite journal requires |journal= (help)
  • Hwang, Suk-Geun (2004). "Cauchy's Interlace Theorem for Eigenvalues of Hermitian Matrices". The American Mathematical Monthly. 111 (2): 157–159. doi:10.2307/4145217. JSTOR 4145217.
  • Kline, Jeffery (2020). "Bordered Hermitian matrices and sums of the Möbius function". Linear Algebra and Its Applications. 588: 224–237. doi:10.1016/j.laa.2019.12.004.
  • Reed, Michael; Simon, Barry (1978). Methods of Modern Mathematical Physics IV: Analysis of Operators. Academic Press. ISBN 978-0-08-057045-7.