Eigenvectors From Eigenvalues
Eigenvectors From Eigenvalues
1. Introduction
If A is an n×n Hermitian matrix, we denote its n real eigenvalues by λ1 (A), . . . , λn (A).
The ordering of the eigenvalues will not be of importance in this survey, but for
sake of concreteness let us adopt the convention of non-decreasing eigenvalues:
λ1 (A) ≤ · · · ≤ λn (A).
If 1 ≤ j ≤ n, let Mj denote the n−1×n−1 minor formed from A by deleting the j th
row and column from A. This is again a Hermitian matrix, and thus has n − 1 real
eigenvalues λ1 (Mj ), . . . , λn−1 (Mj ), which for sake of concreteness we again arrange
in non-decreasing order. In particular we have the well known Cauchy interlacing
inequalities (see e.g., [Wil63, p. 103-104])
(1) λi (A) ≤ λi (Mj ) ≤ λi+1 (A)
for i = 1, . . . , n − 1.
By the spectral theorem, we can find an orthonormal basis of eigenvectors
v1 , . . . , vn of A associated to the eigenvalues λ1 (A), . . . , λn (A) respectively. For
any i, j = 1, . . . , n, let vi,j denote the j th component of vi . This survey paper is de-
voted to the following elegant relation, which we will call the eigenvector-eigenvalue
identity, relating this eigenvector component to the eigenvalues of A and Mj :
Theorem 1 (Eigenvector-eigenvalue identity). With the notation as above, we have
n
Y n−1
Y
(2) |vi,j |2 (λi (A) − λk (A)) = (λi (A) − λk (Mj )) .
k=1;k6=i k=1
one can observe the interlacing inequalities (1). One can then verify (2) for all
i, j = 1, 2, 3:
2 (0 − 2)(0 − 4)
= |v1,1 |2 =
3 (0 − 3)(0 − 4)
√ √
1 2 (0 − 2 − 2)(0 − 2 + 2)
= |v1,2 | =
6 (0 − 3)(0 − 4)
√ √
1 2 (0 − 2 − 2)(0 − 2 + 2)
= |v1,3 | =
6 (0 − 3)(0 − 4)
1 (3 − 2)(3 − 4)
= |v2,1 |2 =
3 (3 − 0)(3 − 4)
√ √
1 (3 − 2 − 2)(3 − 2 + 2)
= |v2,2 |2 =
3 (3 − 0)(3 − 4)
√ √
1 2 (3 − 2 − 2)(3 − 2 + 2)
= |v2,3 | =
3 (3 − 0)(3 − 4)
(4 − 2)(4 − 4)
0 = |v3,1 |2 =
(4 − 0)(4 − 3)
√ √
1 2 (4 − 2 − 2)(4 − 2 + 2)
= |v3,2 | =
2 (4 − 0)(4 − 3)
√ √
1 2 (4 − 2 − 2)(4 − 2 + 2)
= |v3,3 | = .
2 (4 − 0)(4 − 3)
One can also verify (4) for this example after computing
p0A (λ) = 3λ2 − 14λ + 12
pM1 (λ) = λ2 − 6λ + 8
pM2 (λ) = λ2 − 4λ + 2
pM3 (λ) = λ2 − 4λ + 2.
Numerical code to verify the identity can be found at [Den19].
Theorem 1 passes a number of basic consistency checks:
(i) (Dilation symmetry) If one multiplies the matrix A by a real scalar c, then
the eigenvalues of A and Mj also get multiplied by c, while the coefficients
vi,j remain unchanged, which does not affect the truth of (2). To put it
another way, if one assigns units to the entries of A, then the eigenvalues
of A, Mj acquire the same units, while vi,j remains dimensionless, and the
identity (2) is dimensionally consistent.
(ii) (Translation symmetry) If one adds a scalar multiple of the identity λIn to
A, then the eigenvalues of A and Mj are shifted by λ, while the coefficient
vi,j remains unchanged. Thus both sides of (2) remain unaffected by such
a transformation.
(iii) (Permutation symmetry) Permuting the eigenvalues of A or Mj does not
affect either side of (2) (provided one also permutes the index i accordingly).
Permuting the ordering of the rows (and colums), as well as the index j,
similarly has no effect on (2).
4 PETER B. DENTON, STEPHEN J. PARKE, TERENCE TAO, AND XINING ZHANG
(iv) (First degenerate case) If vi,j vanishes, then the eigenvector vi for A also
becomes an eigenvector for Mj with the same eigenvalue λi (A) after deleting
the j th coefficient. In this case, both sides of (2) vanish.
(v) (Second degenerate case) If the eigenvalue λi (A) of A occurs with multiplic-
ity greater than one, then by the interlacing inequalities (1) it also occurs
as an eigenvalue of Mj . Again in this case, both sides of (2) vanish.
(vi) (Compatibility with interlacing) More generally, the identity (2) is consis-
tent with the interlacing (1) because the component vi,j of the unit eigen-
vector vi has magnitude at most 1.
(vii) (Phase symmetry) One has the√ freedom to multiply each eigenvector vi by
an arbitrary complex phase e −1θi without affecting the matrix A or its
minors Mj . But both sides of (2) remain unchanged when one does so.
(viii) (Diagonal case) If A is a diagonal matrix with diagonal entries λ1 (A), . . . , λn (A),
then |vi,j | equals 1 when i = j and zero otherwise, while the eigenvalues of
Mj are formed from those of A by deleting one copy of λi (A). In this case
one can easily verify (2) by hand.
(ix) (Normalization) As the eigenvectors v1 , . . . , vn form Pn the columns of an
2
orthogonal matrix,Pone must have the identities |v
i=1 i,j | = 1 for all
n 2
j = 1, . . . , n and |v
j=1 i,j | = 1 for all i = 1, . . . , n. It is not imme-
diately obvious that these identities are consistent with (2), but we will
demonstrate the former identity in Remark 8. For the latter identity, we
use the translation symmetry P(ii) to normalize Pλni (A) = 0, and then observe
n
(e.g., from (7)) that (−1)n j=1 pMj (0) = j=1 det(Mj ) = tr adj(A) is
the (n − 1)th symmetricPn function of the eigenvalues and thus equal (since
λi (A) vanishes) to k=1;k6=i λk (A) = (−1)n p0A (0). Comparing this with
Pn
(4) we obtain j=1 |vi,j |2 = 1.
also, the form of the identity and the notation used varied widely from appear-
ance to appearance, making it difficult to search for occurrences of the identity
by standard search engines. The situation changed after a popular science article
[Wol19] reporting on the most recent rediscovery [DPZ19, DPTZ19] of the identity
by ourselves; in the wake of the publicity generated by that article, we received
many notifications (see Section 6) of the disparate places in the literature where
the eigenvector-eigenvalue identity, or one closely related to it, was discovered. Ef-
fectively, this crowdsourced the task of collating all these references together. In
this paper, we survey all the appearances of the eigenvector-eigenvalue identity
that we are aware of as a consequence of these efforts, as well as provide several
proofs, generalizations, and applications of the identity. Finally, we speculate on
some reasons for the limited nature of the dissemination of this identity in prior
literature.
6 PETER B. DENTON, STEPHEN J. PARKE, TERENCE TAO, AND XINING ZHANG
If one specializes to the case λ = λi (A) for some i = 1, . . . , n, then all but one of
the summands on the right-hand side vanish, and the adjugate matrix becomes a
scalar multiple of the rank one projection vi vi∗ :
n
Y
(8) adj(λi (A)In − A) = ( (λi (A) − λk (A)))vi vi∗ .
k=1;k6=i
Extracting out the jj component of this identity using (5), we conclude that
n
Y
(9) det(λi (A)In−1 − Mj ) = ( (λi (A) − λk (A)))|vi,j |2
k=1;k6=i
which is equivalent to (2). In fact this shows that the eigenvector-eigenvalue identity
holds for normal matrices A as well as Hermitian matrices (despite the fact that
the minor Mj need not be Hermitian or normal in this case). Of course in this case
the eigenvectors are not necessarily real and thus cannot be arranged in increasing
order, but the order of the eigenvalues plays no role in the identity (2).
leading to an extension
n
Y
(11) det(λi (A)In−1 − Mj ) = ( (λi (A) − λk (A)))vi,j wi,j
k=1;k6=i
2terrytao.wordpress.com/2019/08/13/eigenvectors-from-eigenvalues/#comment-519905
3In the case when A is a normal matrix and the v are unit eigenvectors, the dual eigenvector
i
wi would be the complex conjugate of vi .
8 PETER B. DENTON, STEPHEN J. PARKE, TERENCE TAO, AND XINING ZHANG
2.2. The Cramer’s rule proof. Returning now to the case of Hermitian matrices,
we give a variant of the above proof of (2) that still relies primarily on Cramer’s
rule, but makes no explicit mention of the adjugate matrix; as discussed in the next
section, variants of this argument have appeared multiple times in the literature.
We first observe that to prove (2) for Hermitian matrices A, it suffices to do so
under the additional hypothesis that A has simple spectrum (all eigenvalues occur
with multiplicity one), or equivalently that
This is because any Hermitian matrix with repeated eigenvalues can be approx-
imated to arbitrary accuracy by a Hermitian matrix with simple spectrum, and
both sides of (2) vary continuously with A (at least if we avoid the case when λi (A)
occurs with multiplicity greater than one, which is easy to handle anyway by the
second degenerate case (iv) noted in the introduction).
As before, we diagonalize A in the form (6). For any complex parameter λ not
equal to one of the eigenvalues λi (A), The resolvent (λIn − A)−1 can then also be
diagonalized as
n
−1
X vi vi∗
(12) (λIn − A) = .
i=1
λ − λi (A)
Extracting out the jj component of this matrix identity using Cramer’s rule [Cra50],
we conclude that
n
det(λIn−1 − Mj ) X |vi,j |2
=
det(λIn − A) i=1
λ − λi (A)
which we can express in terms of eigenvalues as
Qn−1 n
(λ − λk (Mj )) X |vi,j |2
(13) Qk=1
n = .
k=1 (λ − λk (A)) i=1
λ − λi (A)
EIGENVECTORS FROM EIGENVALUES 9
Both sides of this identity are rational functions in λ, and have a pole at λ = λi (A)
for any given i = 1, . . . , n. Extracting the residue at this pole, we conclude that
Qn−1
(λ (A) − λk (Mj ))
Qnk=1 i = |vi,j |2
k=1:k6=i (λ i (A) − λ k (A))
which rearranges to give (2).
Remark 6. One can view the above derivation of (2) from (13) as a special case
of the partial fractions decomposition
P (t) X P (α) 1
=
Q(t) Q0 (α) t − α
Q(α)=0
4twitter.com/quantum aram/status/1195185551667847170
10 PETER B. DENTON, STEPHEN J. PARKE, TERENCE TAO, AND XINING ZHANG
and Q
0 1≤l<k≤n−1 (λk (Mj ) − λl (Mj ))(λk (A) − λl (A))
det(S ) = Qn−1 Qn−1
l=1 k=1 (λl (Mj ) − λk (A))
2.3. Coordinate-free proof. We now give a proof that largely avoids the use
of coordinates or matrices, essentially due to Bo Berndtsson5. For this proof we
assume familiarity with exterior algebra (see e.g., [BM41, Chapter XVI]). The key
identity is the following statement.
Lemma 10 (Coordinate-free eigenvector-eigenvalue identity). Let T : Cn → Cn be
a self-adjoint linear map that annihilates a unit vector v. For each unit vector f ∈
Cn , let ∆T (f ) be the determinant of the quadratic form w 7→ (T w, w)Cn restricted
to the orthogonal complement f ⊥ := {w ∈ Cn : (f, w)Cn = 0}, where (, )Cn denotes
the Hermitian inner product on Cn . Then one has
(19) |(v, f )Cn |2 ∆T (v) = ∆T (f )
for all unit vectors f ∈ Cn .
Proof. The determinant of a quadratic form w 7→ (T w, w)Cn on a k-dimensional
subspace V of Cn can be expressed as (T α, α)Vk Cn /(α, α)Vk Cn for any non-degenerate
Vk Vk n
element α of the k th exterior power V ⊂ C (equipped with the usual Her-
Vk n
mitian inner product (, ) k Cn ), where the operator T is extended to
V C in the
n−1 n
usual fashion. If f ∈ Cn is a unit vector, then the Hodge dual ∗f ∈
V
C is a
Vn−1 ⊥
unit vector in (f ), so that we have the identity
(20) ∆T (f ) = (T (∗f ), ∗f )Vn−1 Cn .
To prove (19), it thus suffices to establish the more general identity
(21) (f, v)Cn ∆T (v)(v, g)Cn = (T (∗f ), (∗g))Vn−1 Cn
for all f, g ∈ Cn . If f is orthogonal
Vn−2 ton v then ∗f can be expressed as a wedge
product of v with an element of C , and hence T (∗f ) vanishes, so that (21)
holds in this case. If g is orthogonal to v then we again obtain (21) thanks to the
self-adjoint nature of T . Finally, when f = g = v the claim follows from (20). Since
the identity (21) is sesquilinear in f, g, the claim follows.
Now we can prove (2). Using translation symmetry we may normalize λi (A) = 0.
We apply Lemma 10 to the self-adjoint map T : w 7→ Aw, setting v to be the
null vector v = vi and f to be the standard basis vector Qn ej . Working in the
orthonormal eigenvector basis v1 , . . . , vn we have ∆(v) = k=1;k6=i λk (A); working
Qn−1
in the standard basis e1 , . . . , en we have ∆T (f ) = det(Mj ) = k=1 λk (Mj ). Finally
we have (v, f )Cn = vi,j . The claim follows.
Remark 11. In coordinates, the identity (20) may be rewritten as ∆T (f ) =
f ∗ adj(A)f . Thus we see that Lemma 10 is basically (8) in disguise.
2.4. Proof using perturbative analysis. Now we give a proof using perturba-
tion theory, which to our knowledge first appears in [MD89]. By the usual limiting
argument we may assume that A has simple eigenvalues. Let ε be a small param-
eter, and consider the rank one perturbation A + εej e∗j of A, where e1 , . . . , en is
the standard basis. From (3) and cofactor expansion, the characteristic polynomial
pA+εej e∗j (λ) of this perturbation may be expanded as
pA+εej e∗j (λ) = pA (λ) − εpMj (λ) + O(ε2 ).
5terrytao.wordpress.com/2019/08/13/eigenvectors-from-eigenvalues/#comment-519914
12 PETER B. DENTON, STEPHEN J. PARKE, TERENCE TAO, AND XINING ZHANG
On the other hand, from perturbation theory the eigenvalue λi (A + εej e∗j ) may be
expanded as
λi (A + εej e∗j ) = λi (A) + ε|vi,j |2 + O(ε2 ).
If we then Taylor expand the identity
pA+εej e∗j (λi (A + εej e∗j )) = 0
and extract the terms that are linear in ε, we conclude that
ε|vi,j |2 p0A (λi (A)) − εpMj (λi (A)) = 0
which gives (4) and hence (2).
2.5. Proof using a Cauchy-Binet type formula. Now we give a proof based on
a Cauchy-Binet type formula, which is also related to Lemma 10. This argument
first appeared in [DPTZ19].
Lemma 12 (Cauchy-Binet type formula). Let A be an n × n Hermitian matrix
with a zero eigenvalue λi (A) = 0. Then for any n × n − 1 matrix B, one has
2
det(B ∗ AB) = (−1)n−1 p0A (0) det B vi
where B vi denotes the n × n matrix with right column vi and all remaining
columns given by B.
Proof. We use a perturbative argument related to that in Section 2.4. Since Avi =
0, vi∗ A = 0, and vi∗ vi = 1, we easily confirm the identity
∗
−B ∗ AB + O(ε) O(ε)
B
(εIn − A) B v i =
vi∗ O(ε) ε
for any parameter ε, where the matrix on the right-hand side is given in block form,
with the top left block being an n − 1 × n − 1 matrix and the bottom right entry
being a scalar. Taking determinants of both sides, we conclude that
2
pA (ε) det B vi = (−1)n−1 det(B ∗ AB ∗ )ε + O(ε2 ).
Remark 13. In the case when vi is the basis vector en , we may write A in block
Mn 0n−1×1
form as A = , where 0i×j denotes the i × j zero matrix, and
0 01×n−1 0
B
write B = for some n − 1 × n − 1 matrix B 0 and n − 1-dimensional vector x,
x∗
in which case one can calculate
det(B ∗ AB) = det((B 0 )∗ Mn B 0 ) = det(Mn )|det(B 0 )|2
and
vi = det(B 0 ).
det B
Since p0A (0) = (−1)n−1 det(Mn ) in this case, this establishes (12) in the case vi =
en . The general case can then be established from this by replacing A by U AU ∗ and
B by U B, where U is any unitary matrix that maps vi to en .
EIGENVECTORS FROM EIGENVALUES 13
We now prove (4) and hence (2). Using the permutation and translation sym-
metries
we may normalize λi (A) = 0 and j = 1. If we then apply Lemma 12 with
01×n−1
B= , in which case
In−1
det(B ∗ AB) = det(M1 ) = (−1)n−1 pM1 (0)
and
det B vi = vi,1 .
Applying Lemma 12, we obtain (4).
Now we can give an alternate proof of (4) and hence (2). By permutation
symmetry (iii) it suffices to establish the j = 1 case. Using limiting arguments as
before we may assume that A has distinct eigenvalues; by further limiting arguments
we may also assume that the eigenvalues of M1 are distinct from those of A. By
translation symmetry (ii) we may normalize λi (A) = 0. Comparing (4) with (22),
our task reduces to establishing the identity
p0A (0) = pM1 (0)(1 + X ∗ M1−2 X).
However, for any complex number λ not equal to an eigenvalue of M1 , we may
apply Schur complementation [Cot74] to the matrix
−X ∗
λ − a11
λIn − A =
−X λIn−1 − M1
to obtain the formula
det(λIn − A) = det(λIn−1 − M1 )(λ − a11 − X ∗ (λIn−1 − M1 )−1 X)
or equivalently
pA (λ) = pM1 (λ)(λ − a11 − X ∗ (λIn−1 − M1 )−1 X)
which on Taylor expansion around λ = 0 using pA (0) = 0 gives
p0A (0)λ + O(λ2 ) = (pM1 (0) + O(λ))(λ − a11 + X ∗ M1−1 X + λX ∗ M1−2 X + O(λ2 )).
Setting λ = 0 and using pM1 (0) 6= 0, we conclude that a11 + X ∗ M1−1 X vanishes. If
we then extract the λ coefficient, we obtain the claim.
Remark 15. The same calculations also give the well known fact that the minor
eigenvalues λ1 (M1 ), . . . , λn−1 (M1 ) are precisely the roots for the equation
λ − a11 − X ∗ (λIn−1 − M1 )−1 X = 0.
Among other things, this can be used to establish the interlacing inequalities (1).
2.7. A generalization. The following generalization of the eigenvector-eigenvalue
identity was recently observed by Yu Qing Tang (private communication), relying
primarily on the Cauchy-Binet formula and a duality relationship (23) between the
various minors of a unitary matrix. If A is an n × n matrices and I, J are subsets of
{1, . . . , n} of the same cardinality m, let MI,J (A) denote the n − m × n − m minor
formed by removing the m rows indexed by I and the m columns indexed by J.
Proposition 16 (Generalized eigenvector-eigenvalue identity). Let A be a normal
n × n matrix diagonalized as A = U DU ∗ for some unitary U and diagonal D =
diag(λ1 , . . . , λn ), let 1 ≤ m < n, and let I, J, K ⊂ {1, . . . , n} have cardinality m.
Then
Y Y
detMJ c ,I c (U )(detMK c ,I c (U )) (λj − λi ) = detMJ,K ( (A − λi In ))
i∈I,j∈I c i∈I
Proof. We have Y Y
(A − λi In ) = U (D − λi In )U ∗
i∈I i∈I
and hence by the Cauchy-Binet formula
Y X Y
detMJ,K ( (A−λi In )) = (detMJ,L (U ))(detML,L0 ( (D−λi In )))(detML0 ,K (U ∗ ))
i∈I L,L0 i∈I
0
Qof {1, . . . , n} of cardinality m. A computation
where L, L range over subsets
0
reveals
that the quantity detML,L 0(
Q i∈I (D − λ i In )) vanishes unless L = L = I, in which
case the quantity equals i∈I,j∈I c (λj − λi ). Thus it remains to show that
detMJ c ,I c (U )(detMK c ,I c (U )) = detMJ,I (U )detMI,K (U ∗ ).
Since detMI,K (U ∗ ) = detMK,I (U ), it will suffice to show that
(23) detMJ,I (U ) = detMJ c ,I c (U )detU
for any J, I ⊂ {1, . . . , n} of cardinality m. By permuting rows and columns we may
assume that J = I = {1, . . . , m}. If we split the identity matrixIn into the
left
I m 0 m×n−m
m columns In1 := and the right n − m columns In2 := and
0n−m×m In−m
take determinants of both sides of the identity
U U ∗ In1 In2 = In1 U In2
we conclude that
det(U )detMI c ,J c (U ∗ ) = detMJ,I (U )
giving the claim.
the identity
(µ1 − λk )(µ2 − λk ) . . . (µn − λk )
(24) αk2 =
(λ1 − λk )(λ2 − λk ) . . . (λk−1 − λk )(λk+1 − λk ) . . . (λn − λk )
is established for k = 1, . . . , n, which closely resembles (2). The identity (24) is
obtained via “Eine einfache Rechnung” (an easy calculation) from the standard
relations
n
X αk2
=1
µi − λk
k=1
for i = 1, . . . , n (compare with (16)), after applying Cramer’s rule and the Cauchy
determinant identity (18); as such, it is very similar to the proof of (2) in Section
2.2 that is also based on (18). The identity (24) was used in [L3̈4] to help classify
monotone functions of matrices. It can be related to (2) as follows. For sake of
notation let us just consider the j = n case of (2). Let ε be a small parameter
and consider the perturbation en e∗n + εA of the rank one matrix en e∗n . Standard
perturbative analysis reveals that the eigenvalues of this perturbation consist of
n − 1 eigenvalues of the form ελi (Mn ) + O(ε2 ) for i = 1, . . . , n − 1, plus an outlier
eigenvalue at 1 + O(ε). Rescaling, we see that the rank one perturbation A + 1ε en e∗n
of A has eigenvalues of the form λi (Mn ) + O(ε) for i = 1, . . . , n − 1, plus an outlier
eigenvalue at 1ε + O(1). If we let An , Bn be the quadratic forms associated to
A, A + 1ε en e∗n expressed using the eigenvector basis v1 , . . . , vn , the identity (24)
becomes Qn
1 (λi (A + 1ε en e∗n ) − λk (A))
|vk,n |2 = i=1
Qn .
ε i=1;i6=k (λi (A) − λk (A))
Extracting the 1/ε component of both sides of this identity using the aforementioned
perturbative analysis, we recover (2) after a brief calculation.
A more complicated variant of (24) involving various quantities related to the
Rayleigh-Ritz method of bounding the eigenvalues of a symmetric linear operator
was stated by Weinberger [Wei60, (2.29)], where it was noted that it can be proven
much the same method as in [L3̈4].
The first appearance of the eigenvector-eigenvalue identity in essentially the form
presented here that we are aware of was by Thompson [Tho66, (15)], which does not
reference the prior work of Löwner or Weinberger. In the notation of Thompson’s
paper, A is a normal n × n matrix, and µ1 , . . . , µs are the distinct eigenvalues of A,
with each µi occurring with multiplicity ei . To avoid non-degeneracy it is assumed
that s ≥ 2. One then diagonalizes A = U DU −1 for a unitary U and diagonal
D = diag(λ1 , . . . , λn ), and then sets
X
θiβ = |Uij |2
j:λj =µβ
and
n
X λj − ξi,j−1 ξij − λj
(27) ≤1
i=1
λj − λ1 λn − λj
for 1 ≤ j ≤ n were proved (with most cases of these inequalities already established
in [Tho66]), with a key input being the identity
2 λj − ξi1 λj − ξi,j−1 ξij − λj ξi,n−1 − λj
(28) |uij | = ... ...
λj − λ1 λj − λj−1 λj+1 − λj λn − λj
where uij are the components of the unitary matrix U used in the diagonalization
H = U DU −1 of H. Note from the Cauchy interlacing inequalities that each of the
expressions in braces takes values between 0 and 1. It is not difficult to see that
this identity is equivalent to (2) (or (25)) in the case of Hermitian matrices with
simple eigenvalues, and the hypothesis of simple eigenvalues can then be removed
by the usual limiting argument. As in [Tho66], the identity is established using
adjugate matrices, essentially by the argument given in the previous section. How-
ever, the identity (28) is only derived as an intermediate step towards establishing
the inequalities (26), (27), and is not highlighted as of interest in its own right. The
identity (25) was then reproduced in a further followup paper [Tho69], in which the
identity (13) was also noted; this latter identity was also independently observed
in [DH78].
In the text of Šilov [Š69, Section 10.27], the identity (2) is established, essentially
by the Cramer rule method. Namely, if A(x, x) is a diagonal real quadratic form
on Rn with eigenvalues λ1 ≥ · · · ≥ λn , and Rn−1 is a hyperplane in Rn with unit
normal vector (α1 , . . . , αn ), and µ1 ≥ · · · ≥ µn−1 are the eigenvalues of A on Rn−1 ,
18 PETER B. DENTON, STEPHEN J. PARKE, TERENCE TAO, AND XINING ZHANG
using a version of the Cramer rule arguments in Section 2.2; they cite as reference
the earlier paper [Gol73, (3.6)], which also uses essentially the same proof. A
similar result was stated without proof by Galais, Kneller, and Volpe [GKV12,
Equations (6), (7)]. They provided expressions for both |vi,j |2 and the off-diagonal
eigenvectors as a function of cofactors in place of adjugate matrices. Their work
was in the context of neutrino oscillations.
The identities of Parlett and Golub-van Loan are cited in the thesis of Knyazev
[Kny86, (2.2.27)], again to analyze methods for computing eigenvalues and eigen-
vectors; the identities of Golub-van Loan and Šilov are similarly cited in the paper
of Knyazev and Skorokhodov [KS91] for similar purposes. Parlett’s result is also
reproduced in the text of Xu [Xu95, (3.19)]. In the survey [CG02, (4.9)] of Chu
and Golub on structured inverse eigenvalue problems, the eigenvector-eigenvalue
identity is derived via the adjugate method from the results of [TM68], and used
to solve the inverse eigenvalue problem for Jacobi matrices; the text [Par80] is also
cited.
In the paper [DLNT86, page 210] of Deift, Li, Nanda, and Tomei, the eigenvector-
eigenvalue identity (2) is derived by the Cramer’s rule method, and used to con-
struct action-angle variables for the Toda flow. The paper cites [BG78], which also
reproduces (2) as equation (1.5) of that paper, and in turn cites [Gol73].
In the paper of Mukherjee and Datta [MD89] the eigenvector-eigenvalue identity
was rediscovered, in the context of computing eigenvectors of graphs that arise in
chemistry. If G is a graph on n vertices v1 , . . . , vn , and G − vr is the graph on
n − 1 vertices formed by deleting a vertex vr , r = 1, . . . , n, then in [MD89, (4)] the
identity
(32) P (G − vr ; xj ) = P 0 (G; xj )Crj
2
eigenmode without having to compute the entire eigenmode. Here the generalized
eigenvalue problem
Kψj = λj Mψj
for j = 1, . . . , N is considered, where K is a positive semi-definite N × N real
symmetric matrix, M is a positive definite N × N real symmetric matrix, and the
matrix Ψ = (ψ1 . . . ψN ) of eigenfunctions is normalized so that ΨT M Ψ = I. For
any 1 ≤ α ≤ N , one also solves the constrained system
(α) (α)
Kα ψj = λαj Mα ψj
where Kα , Mα are the N − 1 × N − 1 minors of K, M respectively formed by
removing the αth row and column. Then in [Kau18, (18)] the Cramer rule method
is used to establish the identity
s v
u QN −1
ψαj = ± t Q k=1 (λj − λαk )
|Mα | u
|M| N −1
k=1;k6=j (λj − λk )
for the α component ψαj of ψj , where |M| is the notation in [Kau18] for the
determinant of M. Specializing to the case when M is the identity matrix, we
recover (2).
The eigenvector-eigenvalue identity was discovered by three of us [DPZ19] in
July 2019, initially in the case of 3 × 3 matrices, in the context of trying to find
a simple and numerically stable formula for the eigenvectors of the neutrino os-
cillation Hamiltonian, which form a separate matrix known as the PMNS lepton
mixing matrix. This identity was established in the 3 × 3 case by direct calcula-
tion. Despite being aware of the related identity (22), the four of us were unable
to locate this identity in past literature and wrote a preprint [DPTZ19] in August
2019 highlighting this identity and providing two proofs (the adjugate proof from
Section 2.1, and the Cauchy-Binet proof from Section 2.5). The release of this
preprint generated some online discussion6, and we were notified by Jiyuan Zhang
(private communication) of the prior appearance of the identity earlier in the year
in [FZ19]. However, the numerous other places in the literature in which some form
of this identity appeared did not become revealed until a popular science article
[Wol19] by Wolchover was written in November 2019. This article spread awareness
of the eigenvector-eigenvalue identity to a vastly larger audience, and generated a
large number of reports of previous occurrences of the identity, as well as other
interesting related observations, which we have attempted to incorporate into this
survey.
4. Further discussion
The eigenvector-eigenvalue identity (2) only yields information about the magni-
tude |vi,j | of the components of a given eigenvector vi , but does not directly reveal
the phase of these components. On one hand, this is to be expected, since (as al-
ready noted in the consistency check (vii) in the introduction) one has the freedom
to multiply vi by a phase; for instance, even if one restricts attention to real sym-
metric matrices A and requires the eigenvectors to be real vi , one has the freedom
to replace vi by its negation −vi , so the sign of each component vi,j is ambiguous.
6terrytao.wordpress.com/2019/08/13, www.reddit.com/r/math/comments/cq3en0
22 PETER B. DENTON, STEPHEN J. PARKE, TERENCE TAO, AND XINING ZHANG
However, relative phases, such as the phase of vi,j vi,j 0 are not subject to this ambi-
guity. There are several ways to try to recover these relative phases. One way is to
employ the off-diagonal analogue (10) of (2), although the determinants in that for-
mula may be difficult to compute in general. For small matrices, it was suggested
in [MD89] that the signs of the eigenvectors could often be recovered by direct
inspection of the components of the eigenvector equation Avi = λi (A)vi . In the ap-
plication in [DPZ19], the additional phase could be recovered by a further neutrino
specific identity [Tos91]. For more general matrices, one way to retrieve such phase
information is to apply (2) in multiple bases. For instance, suppose A was real sym-
metric and the vi,j were all real. If one were to apply the eigenvector-eigenvalue
identity after changing to a basis that involved the unit vector √12 (ej + ej 0 ), then
one could use the identity to evaluate the magnitude of √12 (vi,j + vi,j 0 ). Two fur-
ther applications of the identity in the original basis would give the magnitude of
vi,j , vi,j 0 , and this is sufficient information to determine the relative sign of vi,j and
vi,j 0 .
For large unstructured matrices, it does not seem at present that the identity
(2) provides a competitive algorithm to compute eigenvectors. Indeed, to use this
identity to compute all the eigenvector component magnitudes |vi,j |, one would
need to compute all n − 1 eigenvalues of each of the n minors M1 , . . . , Mn , which
would be a computationally intensive task in general; and furthermore, an addi-
tional method would then be needed to also calculate the signs or phases of these
components. However, if the matrix is of a special form (such as a tridiagonal form),
then the identity could be of more practical use, as witnessed by the uses of this
identity (together with variants such as (31)) in the literature to control the rate
of convergence for various algorithms to compute eigenvalues and eigenvectors of
tridiagonal matrices. Also, as noted recently in [Kau18], if one has an application
that requires only the component magnitudes |v1,j |, . . . , |vn,j | at a single location j,
then one only needs to compute the the characteristic polynomial of a single minor
Mj of A at a single value λi (A), and this may be more computationally feasible.
6. Acknowledgments
We thank Carlo Beenakker, Percy Deift, Laurent Demanet, Alan Edelman, Chris
Godsil, Aram Harrow, James Kneller, Andrew Knysaev, Manjari Narayan, Michael
Nielsen, Karl Svozil, Gang Tian, Carlos Tomei, Piet Van Mieghem, Fu Zhang,
Jiyuan Zhang, and Zhenzhong Zhang for pointing out a number of references where
24 PETER B. DENTON, STEPHEN J. PARKE, TERENCE TAO, AND XINING ZHANG
References
[AS64] Milton Abramowitz and Irene A. Stegun. Handbook of mathematical functions with
formulas, graphs, and mathematical tables, volume 55 of National Bureau of Stan-
dards Applied Mathematics Series. For sale by the Superintendent of Documents, U.S.
Government Printing Office, Washington, D.C., 1964.
[Bar01] Yu. Baryshnikov. GUEs and queues. Probab. Theory Related Fields, 119(2):256–274,
2001.
[BFdP11] N. Bebiano, S. Furtado, and J. da Providência. On the eigenvalues of principal sub-
matrices of J-normal matrices. Linear Algebra Appl., 435(12):3101–3114, 2011.
[BG78] D. Boley and G. H. Golub. Inverse eigenvalue problems for band matrices. In Numer-
ical analysis (Proc. 7th Biennial Conf., Univ. Dundee, Dundee, 1977), pages 23–31.
Lecture Notes in Math., Vol. 630, 1978.
[BM41] Garrett Birkhoff and Saunders MacLane. A Survey of Modern Algebra. Macmillan
Company, New York, 1941.
[BT13] Sara C. Billey and Bridget E. Tenner. Fingerprint databases for theorems. Notices
Amer. Math. Soc., 60(8):1034–1039, 2013.
[Cau41] Augustin Louis Cauchy. Exercices d’analyse et de physique mathématique. vol. 2.
Bachelier, page 154, 1841.
[CG02] Moody T. Chu and Gene H. Golub. Structured inverse eigenvalue problems. Acta
Numer., 11:1–71, 2002.
[Cot74] Richard W. Cottle. Manifestations of the Schur complement. Linear Algebra Appl.,
8:189–211, 1974.
[Cra50] Gabriel Cramer. Introduction à l’analyse des lignes courbes algébriques. Geneva, pages
657–659, 1750.
[CRS07] D. Cvetković, P. Rowlinson, and S.K. Simić. Star complements and exceptional graphs.
Linear Algebra and its Applications, 423(1):146 – 154, 2007. Special Issue devoted to
papers presented at the Aveiro Workshop on Graph Spectra.
[DE02] Ioana Dumitriu and Alan Edelman. Matrix models for beta ensembles. J. Math. Phys.,
43(11):5830–5847, 2002.
[Den19] Peter B. Denton. Eigenvalue-eigenvector identity code. https://github.com/
PeterDenton/Eigenvector-Eigenvalue-Identity, 2019.
[DH78] Emeric Deutsch and Harry Hochstadt. On Cauchy’s inequalities for Hermitian matri-
ces. Amer. Math. Monthly, 85(6):486–487, 1978.
[DLNT86] P. Deift, L. C. Li, T. Nanda, and C. Tomei. The Toda flow on a generic orbit is
integrable. Comm. Pure Appl. Math., 39(2):183–232, 1986.
[DPTZ19] Peter B. Denton, Stephen J. Parke, Terence Tao, and Xining Zhang. Eigenvectors from
Eigenvalues. arXiv e-prints, page arXiv:1908.03795v1, Aug 2019.
[DPZ19] Peter B Denton, Stephen J Parke, and Xining Zhang. Eigenvalues: the Rosetta Stone
for Neutrino Oscillations in Matter. arXiv:1907.02534. 2019.
[ESY09] László Erdős, Benjamin Schlein, and Horng-Tzer Yau. Semicircle law on short
scales and delocalization of eigenvectors for Wigner random matrices. Ann. Probab.,
37(3):815–852, 2009.
[FZ19] Peter J. Forrester and Jiyuan Zhang. Co-rank 1 projections and the randomised Horn
problem. arXiv e-prints, page arXiv:1905.05314, May 2019.
[Gan59] F. R. Gantmacher. The theory of matrices. Vol I. Translated by K. A. Hirsch. Chelsea
Publishing Co., New York, 1959.
[GGKL17] Chris Godsil, Krystal Guo, Mark Kempton, and Gabor Lippner. State transfer in
strongly regular graphs with an edge perturbation, 2017.
[GKV12] Sbastien Galais, James Kneller, and Cristina Volpe. The neutrino-neutrino interaction
effects in supernovae: the point of view from the matter basis. J. Phys., G39:035201,
2012.
[GM81] C. D. Godsil and B. D. McKay. Spectral conditions for the reconstructibility of a
graph. J. Combin. Theory Ser. B, 30(3):285–289, 1981.
EIGENVECTORS FROM EIGENVALUES 25
[God93] C. D. Godsil. Algebraic combinatorics. Chapman and Hall Mathematics Series. Chap-
man & Hall, New York, 1993.
[God12] Chris Godsil. When can perfect state transfer occur? Electron. J. Linear Algebra,
23:877–890, 2012.
[Gol73] Gene H. Golub. Some modified matrix eigenvalue problems. SIAM Rev., 15:318–334,
1973.
[GVL83] Gene H. Golub and Charles F. Van Loan. Matrix computations, volume 3 of Johns
Hopkins Series in the Mathematical Sciences. Johns Hopkins University Press, Balti-
more, MD, 1983.
[Hag02] Elias M. Hagos. Some results on graph spectra. Linear Algebra and its Applications,
356(1):103 – 111, 2002.
[Hal42] Paul R. Halmos. Finite Dimensional Vector Spaces. Annals of Mathematics Studies,
no. 7. Princeton University Press, Princeton, N.J., 1942.
[Kau18] Eduardo Kausel. Normalized modes at selected points without normalization. Journal
of Sound and Vibration, 420:261–268, April 2018.
[Kny86] Andrew Knyazev. Computation of eigenvalues and eigenvectors for mesh problems:
algorithms and error estimates (in Russian). PhD thesis, Department of Numerical
Mathematics, USSR Academy of sciences, Moscow, 1986.
[KS91] A. V. Knyazev and A. L. Skorokhodov. On exact estimates of the convergence rate
of the steepest ascent method in the symmetric eigenvalue problem. Linear Algebra
Appl., 154/156:245–257, 1991.
[L3̈4] Karl Löwner. Über monotone Matrixfunktionen. Math. Z., 38(1):177–216, 1934.
[LF79] Qiao Li and Ke Qin Feng. On the largest eigenvalue of a graph. Acta Math. Appl.
Sinica, 2(2):167–175, 1979.
[MD89] Asok K. Mukherjee and Kali Kinkar Datta. Two new graph-theoretical methods for
generation of eigenvectors of chemical graphs. Proceedings of the Indian Academy of
Sciences - Chemical Sciences, 101(6):499–517, Dec 1989.
[NTU93] Peter Nylen, Tin Yau Tam, and Frank Uhlig. On the eigenvalues of principal subma-
trices of normal, Hermitian and symmetric matrices. Linear and Multilinear Algebra,
36(1):69–78, 1993.
[Pai71] Christopher Conway Paige. The computation of eigenvalues and eigenvectors of very
large sparse matrices. PhD thesis, London University Institute of Computer Science,
1971.
[Par80] Beresford N. Parlett. The symmetric eigenvalue problem. Prentice-Hall, Inc., Engle-
wood Cliffs, N.J., 1980. Prentice-Hall Series in Computational Mathematics.
[Tho66] R. C. Thompson. Principal submatrices of normal and Hermitian matrices. Illinois J.
Math., 10:296–308, 1966.
[Tho69] R. C. Thompson. Principal submatrices. IV. On the independence of the eigenvalues
of different principal submatrices. Linear Algebra and Appl., 2:355–374, 1969.
[TM68] R.C. Thompson and P. McEnteggert. Principal submatrices ii: the upper and lower
quadratic inequalities. Linear Algebra and its Applications, 1:211–243, 1968.
[Tos91] S. Toshev. On T violation in matter neutrino oscillations. Mod. Phys. Lett., A6:455–
460, 1991.
[TV11] Terence Tao and Van Vu. Random matrices: Universality of local eigenvalue statistics.
Acta Math., 206(1):127–204, 2011.
[Van14] Piet Van Mieghem. Graph eigenvectors, fundamental weights and centrality metrics
for nodes in networks. arXiv e-prints, page arXiv:1401.4580, Jan 2014.
[VM11] Piet Van Mieghem. Graph spectra for complex networks. Cambridge University Press,
Cambridge, 2011.
[Š69] G. E. Šilov. Matematicheskiĭ analiz. Konechnomernye lineĭ nye prostranstva. Izdat.
“Nauka”, Moscow, 1969.
[Wei60] H. F. Weinberger. Error bounds in the Rayleigh-Ritz approximation of eigenvectors.
J. Res. Nat. Bur. Standards Sect. B, 64B:217–225, 1960.
[Wil63] J. H. Wilkinson. Rounding errors in algebraic processes. Prentice-Hall, Inc., Englewood
Cliffs, N.J., 1963.
[Wol19] Natalie Wolcholver. Neutrinos lead to unexpected discovery in basic math. Quanta
Magazine, Nov 2019.
26 PETER B. DENTON, STEPHEN J. PARKE, TERENCE TAO, AND XINING ZHANG
[Xu95] Shufang Xu. Theories and Methods of Matrix Calculations (in Chinese). Peking Uni-
versity Press, Beijing, 1995.