Next Article in Journal
Successive Approximation Technique in the Study of a Nonlinear Fractional Boundary Value Problem
Next Article in Special Issue
Geometry of Bi-Warped Product Submanifolds of Nearly Trans-Sasakian Manifolds
Previous Article in Journal
Integration of the Kenzo System within SageMath for New Algebraic Topology Computations
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Generalization of the Lieb–Thirring–Araki Inequality and Its Applications

1
College of Science, Zhengzhou University of Aeronautics, Zhengzhou 450015, China
2
School of Information, Beijing Wuzi University, Beijing 101149, China
3
School of Mathematics and Statistics, Beijing Institute of Technology, Beijing 100081, China
*
Author to whom correspondence should be addressed.
Mathematics 2021, 9(7), 723; https://doi.org/10.3390/math9070723
Submission received: 9 February 2021 / Revised: 20 March 2021 / Accepted: 22 March 2021 / Published: 26 March 2021
(This article belongs to the Special Issue Analytic and Geometric Inequalities: Theory and Applications)

Abstract

:
The matrix eigenvalue is very important in matrix analysis, and it has been applied to matrix trace inequalities, such as the Lieb–Thirring–Araki theorem and Thompson–Golden theorem. In this manuscript, we obtain a matrix eigenvalue inequality by using the Stein–Hirschman operator interpolation inequality; then, according to the properties of exterior algebra and the Schur-convex function, we provide a new proof for the generalization of the Lieb–Thirring–Araki theorem and Furuta theorem.

1. Introduction

As an important branch of mathematics, matrix theory has been widely applied in the fields of mathematics and technology, such as optimization theory ([1]), differential equations ([2]), numerical analysis, operations ([3]) and quantum theory ([4]).
In this manuscript, let C n be an n-dimensional complex vector space with the inner product x , y = x * y = i = 1 n x i * y i for x = ( x 1 , , x n ) , y = ( y 1 , , y n ) C n , where the superscripts x * and denote the conjugated transpose of x and the matrix transpose, respectively. Let M n denote the whole set of n × n matrices with complex entries, and we call x C n the eigenvector of A M n when A x = λ x (where λ is called the eigenvalue of A). We denote H n the set of all Hermitian matrices. For any A H n , we have A = i = 1 n λ i P i , where λ i is the eigenvalue of A and i = 1 n P i = I d , P i P j = 0 ( i j ) ; specially, when x * A x 0 for any x C n , we denote A H n + ( H n + is the set of n × n positive-definite Hermitian matrices whose eigenvalues are nonnegative). Let f be a function with the domain ( 0 , + ) ; for any A H n + , the matrix function is defined as
f ( A ) = i = 1 n f ( λ i ) P i .
On the basis of this definition, we have a formula relating the trace of matrix A and the eigenvalue of A:
Tr [ A ] = i = 1 n λ i ( A ) ,
where λ i ( A ) is the eigenvalue of A. It is well known that Tr [ A B ] = Tr [ B A ] for any A , B H n + . However, Tr [ e A + B ] = Tr [ e A e B ] only if A B = B A . Fortunately, in 1965, Thompson and Golden independently discovered an inequality called the Thompson–Golden theorem (refer to [5,6,7]):
Tr [ e A + B ] Tr [ e A e B ] .
In general, the following limit holds (called the Lie–Trotter formula [8]):
lim p 0 Tr [ e p A 2 e p B e p A 2 1 p ] = Tr [ e A + B ] .
Furthermore, the following inequality holds when p 1 :
Tr [ A 1 2 p B 1 p A 1 2 p ] = Tr [ e ln A 2 p e ln B p e ln A 2 p ] Tr [ e ln A 2 e ln B e ln A 2 1 p ] = Tr [ A 1 2 B A 1 2 1 p ] ,
which is the Lieb–Thirring–Araki theorem ([9,10]). Since the function F ( A ) = Tr e B + ln A is a Fr e ´ chet differential function for any A H n , the concavity of F ( A ) implies the Thompson–Golden theorem. At the same time, one can also obtain the Thompson–Golden theorem by using the relationship λ 1 ( A α 2 B α A α 2 ) 1 α and λ 1 ( A β 2 B β A β 2 ) 1 β ([11]). It is known that det ( A 1 2 p B 1 p A 1 2 p ) = [ det ( A 1 2 B A 1 2 ) ] 1 p . By using the matrix exterior algebra, we have
det ( A α 2 B α A α 2 ) 1 α = Tr n ( A α 2 B α A α 2 ) 1 α .
According to the convexity of Tr k e A , Huang proved the following inequality ([12]):
Tr k e A + B Tr k ( A α 2 B α A α 2 ) 1 α .
With this motivation, we utilize the Stein–Hirschman operator interpolation inequality to show that λ 1 ( A α 2 B α A α 2 ) 1 α is a monotone increasing function for any α > 0 . Then, we generalize the Lieb–Thirring–Araki theorem and provide a new proof of the Furuta theorem ([13]). The rest of the paper is organized as follows. In Section 2, some general definitions and important conclusions are introduced. In Section 3, a new proof of the monotonicity of λ 1 ( A α 2 B α A α 2 ) 1 α and some general results are offered.

2. Preliminary

In this section, we recall some notions and definitions from matrix analysis, and introduce some important results of the matrix-monotone function, which are used through the article (refer to [14,15,16,17]).

2.1. Tensor Product and Exterior Algebra

The tensor product, denoted by ⊗, is also called the Kronecker product. It is a generalization of the outer product from vectors to matrices, so the tensor product of matrices is referred to as the outer product as well in some contexts. For an m × n matrix A and a p × q matrix B, the tensor product of A and B is defined by
A B = a 11 B a 1 n B a m 1 B a m n B ,
where A = a i j 1 i m , 1 j n .
The tensor product is different from matrix multiplication, and one of the differences is commutativity:
( I B ) ( A I ) = ( A I ) ( I B ) = A B .
From this relation, one can obtain
A C B D = ( A C I ) ( I B D ) = ( A I ) ( C I ) ( I B ) ( I D ) = ( A I ) ( I B ) ( C I ) ( I D ) = ( A B ) ( C D ) .
For convenience, we denote
k A = A A A k .
In addition to the tensor product, there is another common product named exterior algebra ([18]). Exterior algebra, denoted by ∧, is a binary operation for any A n × n that is
( A 1 A 2 A k k ) ( ξ i 1 ξ i 2 ξ i k ) 1 i 1 < < i k n = ( A 1 ξ i 1 A 2 ξ i 2 A k ξ i k ) 1 i 1 < < i k n ,
where { ξ j } j = 1 n is an orthogonal basis of C n and
ξ i 1 ξ i 2 ξ i k = 1 n ! π σ n ( 1 ) π ξ π ( i 1 ) ξ π ( i 2 ) ξ π ( i k ) ,
where σ n is the family of all permutations on { 1 , 2 , , n } .
Let k C n be the span of the { ξ i 1 ξ i 2 ξ i k } 1 i 1 < < i k n ; a simple calculation shows that
λ 1 ( k A ) = i = 1 k λ i ( A ) .

2.2. Schur-Convex Function

Let x = ( x 1 , , x n ) , y = ( y 1 , , y n ) R n and denote
x [ 1 ] = max i = 1 , 2 , , n { x i } , , x [ i ] = max i = 1 , 2 , , n { x i / x [ i ] i = 1 , 2 , , i 1 } , , x [ n ] = min i = 1 , 2 , , n { x i } , y [ 1 ] = max i = 1 , 2 , , n { y i } , , y [ i ] = max i = 1 , 2 , , n { y i / y [ i ] i = 1 , 2 , , i 1 } , , y [ n ] = min i = 1 , 2 , , n { y i } .
If x and y satisfy
i = 1 k x [ i ] i = 1 k y [ i ] , k n 1 ; i = 1 n x [ i ] = i = 1 n y [ i ] , k = n .
then x is said to be majorized by y , denoted by x y . Meanwhile, we denote x w y if i = 1 k x [ i ] i = 1 k y [ i ] for any k n .
Suppose f is a real-valued function defined on a set A R n ; then, f is said to be a Schur-convex function on A if, for any x , y R n and x y , one obtains f ( x ) f ( y ) ([11]).
If f is differentiable and defined on I n ( I R being an open interval), then the following lemma holds (refer to [11]).
Lemma 1.
f is Schur-convex on I n R n if and only if
( x i x j ) ( f x i f x j ) 0 f o r a l l x I n .

2.3. The Matrix-Monotone Function

For a matrix A H n + , according to the spectral theorem ([19]), it can be decomposed as
A = P * Λ A P ,
where P is the unitary matrix and Λ A : = diag { λ 1 , , λ n } is a diagonal matrix with eigenvalues as elements. When x * ( A B ) x > 0 for any x C n , we denote A B .
Associated with a function f ( x ) on ( 0 , + ) , the matrix function f ( A ) is defined as
f ( A ) = P * f ( Λ A ) P ,
where f ( Λ A ) = diag { f ( λ 1 ) , , f ( λ n ) } . Then, the function f is said to be matrix-monotonic if it satisfies
f ( A ) f ( B ) for all A B > 0 .
Since the matrix-monotone function is a special type of operator monotone function, we present the following general conclusion about the operator-monotone function, which can be found in [20,21].
Lemma 2.
The following statements for a real-valued continuous function f on ( 0 , + ) are equivalent:
(i) 
f is operator-monotone;
(ii) 
f admits an integral representation
f ( λ ) = α + β λ + 0 ( 1 + λ t ) ( t λ ) 1 d μ ( t ) , f o r a n y λ > 0 ,
where α is a real number, β is non-negative and μ is a finite positive measure on ( , 0 ) .

3. The Main Results

For any A , B M n , it is known ln A B is not equal to ln A + ln B in general when A B B A . Generally, ( A 1 2 B A 1 2 ) α is not equal to A α 2 B α A α 2 . Therefore, many people pay much attention to studying the relation between ( A 1 2 B A 1 2 ) α and A α 2 B α A α 2 . A famous result regarding the trace inequality is the Lieb–Thirring–Araki theorem:
Tr [ ( A 1 2 B A 1 2 ) α ] Tr [ ( A α 2 B α A α 2 ) ] ,
where α 1 . In the following, we further study the relation between ( A 1 2 B A 1 2 ) α and A α 2 B α A α 2 and provide the main results.
Theorem 1.
For any 0 < α β and A , B H n + , the following inequality holds
λ 1 ( A α 2 B α A α 2 ) 1 α λ 1 ( A β 2 B β A β 2 ) 1 β .
Proof. 
By using the Cauchy inequality, we have
x * A B x 2 = < x , A B x > 2 ( x * A * A x ) ( x * B * B x ) , x C n .
If we denote
λ 1 ( A ) = max x C n { < x , A x > } < x , x > = max x C n , < x , x > = 1 { < x , A x > }
as the maximum eigenvalue of A, then we can obtain
λ 1 ( A 1 2 B A 1 2 ) = max x C n { < x , A 1 2 B A B A 1 2 x > } 1 2 = max x C n { < A B x , B A x > } 1 2 max x C n { < A B x , A B x > 1 4 < B A x , B A x > 1 4 } max x C n { < x , B A 2 B x > 1 4 } max x C n { < x , A B 2 A x > 1 4 } = λ 1 ( A B 2 A ) 1 2 ,
Here, we use the fact that λ 1 ( A B ) = λ 1 ( B A ) .
Through a simple deformation, we have λ 1 ( A k 1 2 B k 1 A k 1 2 ) 1 k 1 λ 1 ( A k 2 B k A k 2 ) 1 k . From this inequality, we have the expression
λ 1 ( A k 2 B k A k 2 ) = λ 1 ( A k + 1 2 B k + 1 2 B k 1 2 A k 1 2 ) λ 1 ( A k + 1 2 B k + 1 A k + 1 2 ) 1 2 λ 1 ( A k 1 2 B k 1 A k 1 2 ) 1 2 λ 1 ( A k + 1 2 B k + 1 A k + 1 2 ) 1 2 λ 1 ( A k 2 B k A k 2 ) k 1 2 k .
Furthermore,
λ 1 ( A k 2 B k A k 2 ) 1 k λ 1 ( A k + 1 2 B k + 1 A k + 1 2 ) 1 k + 1 .
This implies
λ 1 ( A m 1 n 2 2 B m 1 n 2 A m 1 n 2 2 ) 1 m 1 n 2 λ 1 ( A m 2 n 1 2 B m 2 n 1 A m 2 n 1 2 ) 1 m 2 n 1 ,
for any m 1 , m 2 , n 1 , n 2 > 0 and m 1 n 2 m 2 n 1 . Let A = A 1 n 1 n 2 , B = B 1 n 1 n 2 ; we obtain
λ 1 ( A m 1 2 n 1 B m 1 n 1 A m 1 2 n 1 ) n 1 m 1 λ 1 ( A m 2 2 n 2 B m 2 n 2 A m 2 2 n 2 ) n 2 m 2 .
Namely, for 0 < α β , we obtain
λ 1 ( A α 2 B α A α 2 ) 1 α λ 1 ( A β 2 B β A β 2 ) 1 β .
This completes the proof of Theorem 1. □
Although Theorem 1 has been obtained from the Cauchy inequality, the frequency of retractions improves the inequality. In the following, we obtain Theorem 1 by using operator interpolation. First, let us introduce the Stein–Hirschman operator interpolation inequality ([12]).
Lemma 3.
Supposing G z tobe an analytic family of linear operators of admissible growth defined in the strip 0 R ( z ) 1 and 1 p 1 , p 2 , q 1 , q 2 , when 1 p = 1 t p 1 + t p 2 , 1 q = 1 t q 1 + t q 2 ( 0 t 1 ), G i y ( f ) q 1 A 0 ( y ) f p 1 and G 1 + i y ( f ) q 2 A 1 ( y ) f p 2 , we can obtain
ln G t ( f ) q R ( ω ( 1 t , y ) ln A 0 ( y ) + ω ( t , y ) ln A 1 ( y ) ) d y + ln f p ,
where ln A 0 ( y ) , ln A 1 ( y ) A e α y ( α < π ) and ω ( t , y ) = t a n ( t π 2 ) 2 [ t a n 2 ( t π 2 ) + t a n h 2 ( y π 2 ) ] c o s h 2 ( y π 2 ) .
From Lemma 3, we can improve the result in Theorem 1 and obtain the following theorem.
Theorem 2.
For any A , B H n + , the following inequality holds
λ 1 ( A t 2 B t A t 2 ) 1 t R ( ω ( t , y ) t λ 1 ( ( A 1 2 B 1 i y A B 1 + i y A 1 2 ) ) 1 2 ) d y λ 1 ( A 1 2 B A 1 2 ) .
Proof. 
Firstly, let f be an analytic function in C . When A , B H n + , let G z ( A , B ) = λ 1 ( A z 2 B z A z 2 ) ; obviously, G z ( A , B ) is an analytic function for any z in strip 0 R ( z ) 1 .
Since
G i y ( A , B ) = λ 1 ( A i y 2 B i y A i y 2 ) λ 1 ( ( A i y 2 B i y A i y 2 ) ( A i y 2 B i y A i y 2 ) ) 1 2 = 1 ,
and
G 1 + i y ( A , B ) = λ 1 ( A 1 + i y 2 B 1 + i y A 1 + i y 2 ) = λ 1 ( A 1 + i y B 1 + i y ) λ 1 ( ( A 1 i y 2 B 1 i y A 1 i y 2 ) ( A 1 + i y 2 B 1 + i y A 1 + i y 2 ) ) 1 2 = λ 1 ( ( A 1 2 B 1 i y A B 1 + i y A 1 2 ) ) 1 2 = λ 1 ( ( B 1 2 A B 1 2 B 1 2 i y A B 1 2 + i y ) ) 1 2 λ 1 ( B 1 2 A B 1 2 ) λ 1 ( B 1 2 i y A B 1 2 + i y ) 1 2 = λ 1 ( A 1 2 B A 1 2 ) ,
and the formula
R ω ( 1 t , y ) d y = 1 t ,
then, for any f L 1 t ( C ) , we can obtain
ln G t ( f ) 1 t = ln λ 1 ( A t 2 B t A t 2 ) + ln f 1 t R ( ω ( 1 t , y ) ln A 0 ( y ) + ω ( t , y ) ln A 1 ( y ) ) d y + ln f 1 t R ( ω ( 1 t , y ) ln A 0 ( y ) + ω ( t , y ) ln λ 1 ( ( A 1 2 B 1 i y A B 1 + i y A 1 2 ) ) 1 2 ) d y + ln f 1 t R ( ω ( 1 t , y ) ln 1 + ω ( t , y ) ln λ 1 ( A 1 2 B A 1 2 ) d y + ln f 1 t = t ln λ 1 ( A 1 2 B A 1 2 ) + ln f 1 t .
This implies
λ 1 ( A t 2 B t A t 2 ) 1 t R ( ω ( t , y ) t λ 1 ( ( A 1 2 B 1 i y A B 1 + i y A 1 2 ) ) 1 2 ) d y λ 1 ( A 1 2 B A 1 2 )
or
λ 1 ( A t 2 B t A t 2 ) R ( ω ( t , y ) t λ 1 ( ( A 1 2 B 1 i y A B 1 + i y A 1 2 ) ) t 2 ) d y λ 1 ( A 1 2 B A 1 2 ) t ,
for any 0 < t < 1 , and the first ”≤“ is obtained by the Jensen inequality ([22]). This completes the proof of Theorem 2. □
Theorem 2 is very useful. On one hand, when α < β , letting t = α β , we can obtain Theorem 1. On the other hand, using the matrix exterior algebra, we obtain
i = 1 k λ i ( A α 2 B α A α 2 ) 1 α i = 1 k λ i ( A β 2 B β A β 2 ) 1 β .
Furthermore, we can deduce the following inequality
Tr k ( A α 2 B α A α 2 ) 1 α Tr [ k ( A β 2 B β A β 2 ) 1 β ] ,
whether it is true or not for any k n , and this inequality can be regarded as a generalization of the Lieb–Thirring–Araki theorem.

3.1. Generalization of Lieb–Thirring–Araki Theorem

According to Theorem 1 and Formula (1), we can show that
i = 1 k ln λ i ( A α 2 B α A α 2 ) 1 α i = 1 k ln λ i ( A β 2 B β A β 2 ) 1 β , k n 1 ; i = 1 n ln λ i ( A α 2 B α A α 2 ) 1 α = i = 1 n ln λ i ( A β 2 B β A β 2 ) 1 β .
Furthermore, we have
ln λ 1 ( A α 2 B α A α 2 ) 1 α , , ln λ n ( A α 2 B α A α 2 ) 1 α ln λ 1 ( A β 2 B β A β 2 ) 1 β , , ln λ n ( A β 2 B β A β 2 ) 1 β .
Let f ( x 1 , x 2 , x n ) = i = 1 n e x i ; direct calculations show that
( x i x j ) ( f x i f x j ) = ( x i x j ) ( e x i e x j ) 0 ,
which implies f ( x 1 , x 2 , x n ) = i = 1 n e x i is a Schur-convex function. Hence, we have the Thompson–Golden theorem
Tr [ ( A α 2 B α A α 2 ) 1 α ] = i = 1 n λ i ( A α 2 B α A α 2 ) 1 α i = 1 n λ i ( A β 2 B β A β 2 ) 1 β = Tr [ ( A β 2 B β A β 2 ) 1 β ] ,
when α β . Specially, when β = 1 and A = A 1 α , B = B 1 α , we have the Lieb–Thirring–Araki theorem
Tr [ ( A 1 2 B A 1 2 ) 1 α ] Tr [ ( A 1 2 α B 1 α A 1 2 α ) ] .
We know that the Lieb–Thirring–Araki theorem can be obtained from the Schur-convex function f ( x 1 , x 2 , x n ) = i = 1 n e x i . Generally, we can prove the following conclusion.
Theorem 3.
For any x i 0 , the function
g ( x 1 , x 2 , x n ) = 1 i 1 < i 2 < < i k n e γ x i 1 e γ x i 2 e γ x i k
is Schur-convex for any γ > 0 .
Proof. 
Since
g ( x 1 , x 2 , x n ) = 1 i 1 < i 2 < < i k n e γ x i 1 e γ x i 2 e γ x i k ,
we have
g x i = γ e γ x i 1 i 1 < < < i k n , x i ^ e γ x i 1 e γ x i 2 e γ x i k ,
where x i ^ denotes the removal of x i .
This implies
( x i x j ) ( g x i g x j ) = γ ( x i x j ) e γ x i e γ x j 1 i 1 < < i k n , x i ^ , x j ^ e γ x i 1 e γ x i 2 e γ x i k 0 .
This completes the proof of Theorem 3. □
From Theorem 3, we can deduce the following inequality immediately.
Corollary 1.
For any α 1 and A , B > 0 , the following inequality holds
Tr k ( A 1 2 B A 1 2 ) α γ Tr k ( A α 2 B α A α 2 ) γ ,
or
Tr k ( A 1 2 B A 1 2 ) α γ Tr k e α R ω ( 1 α , y ) ln ( A α 2 B α i α y A α B α + i α y A α 2 ) ) γ 2 d y , Tr k ( A α 2 B α A α 2 ) γ
for any γ > 0 .
From (8), it can be seen that, when k = 1 , Corollary 1 is just the Lieb–Thirring–Araki theorem. Especially, Tr k ( A 1 2 B A 1 2 ) α Tr k ( A α 2 B α A α 2 ) when γ = 1 ([12]).
Using Theorem 1, we can obtain
Tr k ( A α 2 B α A α 2 ) γ α Tr k ( A β 2 B β A β 2 ) γ β .
for any 0 α β , and this is a generalization of the Thompson–Golden theorem. For some other generalizations of the Thompson–Golden theorem, see [8,23]. Moreover, since
λ ( A r 2 ( A r 2 B r A r 2 ) α A r 2 ) λ [ ( A 1 2 ( A 1 2 B A 1 2 ) α A 1 2 ) r ] ,
where 0 < α 1 and r 1 , we can obtain the following corollary.
Corollary 2.
For any r 1 and A , B > 0 , the following inequality holds:
Tr k ( A 1 2 ( A 1 2 B A 1 2 ) A 1 2 ) r Tr k ( A r 2 ( A r 2 B r A r 2 ) A r 2 ) .

3.2. Applications in Matrix-Monotone Function

In this subsection, we obtain some other corollaries from Theorem 1 associated with the matrix-monotone function. Since
λ 1 ( A α 2 B α A α 2 ) 1 α λ 1 ( A β 2 B β A β 2 ) 1 β ,
we can obtain λ 1 ( A 1 2 B A 1 2 ) β λ 1 ( A β 2 B β A β 2 ) for any 0 < β 1 . While A I d is equivalent to λ 1 ( A ) 1 , using this fact, for A B , we have B 1 2 A B 1 2 I . For 0 < β 1 , we have
λ 1 ( B β 2 A β B β 2 ) λ 1 ( B 1 2 A B 1 2 ) β 1 .
Hence, we obtain the L o ¨ wner–Heinz Theorem ([4]).
Corollary 3.
For 0 < α 1 and A B , A α B α .
Unfortunately, when α > 1 , Corollary 3 is false. However, we can use Theorem 1 to obtain the Furuta theorem [13].
Corollary 4.
Let 0 B A ; then
A p + 2 r q ( A r B p A r ) 1 q ,
where p , r 0 , q 1 and q p + 2 r 1 + 2 r .
Proof. 
When 0 p 1 , we know that B p A p from Corollary 3; this implies
( A r B p A r ) 1 q ( A r A p A r ) 1 q = A p + 2 r q .
Hence, we suppose p 1 . Meanwhile, we know p + 2 r ( 1 + 2 r ) q 1 , so we only consider p + 2 r 1 + 2 r = q . Firstly, when 0 r 1 2 , we have
λ 1 B 1 2 A r ( A r B p A r ) 1 q A r B 1 2 = λ 1 B 1 2 B p 2 ( B p 2 A 2 r B p 2 ) 1 1 q B p 2 B 1 2 = λ 1 ( B p + 2 r 2 ) p 1 p + 2 r ( B p 2 A 2 r B p 2 ) p 1 p + 2 r ( B p + 2 r 2 ) p 1 p + 2 r λ 1 B p + 2 r 2 ( B p 2 A 2 r B p 2 ) B p + 2 r 2 p 1 p + 2 r = λ 1 ( A r B 2 r A r ) p 1 p + 2 r 1 .
This implies
A r ( A r B p A r ) 1 q A r B A .
That is,
( A r B p A r ) 1 q A 1 + 2 r .
Secondly, let A 1 = A 1 + 2 r , B 1 = ( A r B p A r ) 1 q ; then
( A 1 1 2 B 1 q A 1 1 2 ) 1 q 1 A 1 2 .
That is, ( A 2 r + 1 2 B p A 2 r + 1 2 ) 1 q A 2 ( 1 + 2 r ) , where q 1 = p + 4 r + 1 1 + 4 r + 1 .
This implies
( A s B p A s ) 1 q A 1 + 2 s ,
where s = 2 r + 1 2 [ 1 2 , 3 2 ] .
Repeating this process, we have finished the proof. □

3.3. Some Other Applications

In this subsection, we obtain a corollary associated with the matrix determinant. We suppose A , B H n and λ n ( e A 2 e B e A 2 ) 1 ; then, from Theorem 1, we have
ln λ 1 ( e α A 2 e α B e α A 2 ) 1 α , , ln λ n ( A α 2 B α A α 2 ) 1 α ln λ 1 ( e β A 2 e β B e β A 2 ) 1 β , , ln λ n ( e β A 2 e β B e β A 2 ) 1 β ,
where 0 < α β and ln λ i ( e A 2 e B e A 2 ) 0 ( i = 1 , 2 , n ). Let
d ( x 1 , , x n ) = i = 1 n x i , ( x i 0 ) .
Then, a straightforward calculation indicates
( x i x j ) ( d ( x 1 , , x n ) x i d ( x 1 , , x n ) x j ) 0 .
Hence, d ( x 1 , , x n ) is a Schur-concave function and the following inequality holds ([8]).
Corollary 5.
Supposing A , B H n and λ n ( e A 2 e B e A 2 ) 1 ,
det ( A + B ) det ( ln ( e α A 2 e α B e α A 2 ) 1 α ) det ( ln ( e β A 2 e β B e β A 2 ) 1 β ) ,
the first ”≥“ obtained by using the Lie–Trotter formula, and 0 < α β .
In fact, for any A H n , we have
det ( A ) = Tr [ n A ] .
Hence, Corollary 5 can be generalized as the following corollary.
Corollary 6.
Supposing A , B H n and λ n ( e A 2 e B e A 2 ) 1 ,
Tr [ k ln ( e α A 2 e α B e α A 2 ) 1 α ] Tr [ k ln ( e β A 2 e β B e β A 2 ) 1 β ] ,
where 1 k n and 0 < α β .
Proof. 
Since
ln λ 1 ( A α 2 B α A α 2 ) 1 α , , ln λ n ( A α 2 B α A α 2 ) 1 α ln λ 1 ( A β 2 B β A β 2 ) 1 β , , ln λ n ( A β 2 B β A β 2 ) 1 β ,
we can finish the proof if we show that the function
a ( x 1 , x 2 , x n ) = 1 i 1 < i 2 < < i k n x i 1 x i 2 x i k
is Schur-concave for any x i 0 . In fact, we have
( x i i x i j ) ( a ( x 1 , , x n ) x i i a ( x 1 , , x n ) x i j ) = ( x i k x i j ) ( 1 i 1 < i i ^ < < i k n x i 1 x i i ^ x i k 1 i 1 < i j ^ < < i k n x i 1 x i j ^ x i k ) = ( x i k x i j ) 2 ( 1 i 1 < i i ^ , i j ^ < < i k n x i 1 x i i ^ x i j ^ x i k ) 0 .
Hence, we know
Tr [ k ln ( e α A 2 e α B e α A 2 ) 1 α ] Tr [ k ln ( e β A 2 e β B e β A 2 ) 1 β ] .
This completes the proof of Corollary 6. □

4. Conclusions

In the paper, we discuss the relationship between λ 1 ( A 1 2 B A 1 2 ) α and λ 1 ( A α 2 B α A α 2 ) by using the Stein–Hirschman operator interpolation inequality. Through in-depth study, we obtain some eigenvalue inequalities such as the generalization Golden–Thompson theorem and Lieb–Thirring–Araki theorem. Moreover, the Furuta theorem is also shown by using the eigenvalue inequality. At last, we generalize an important determinant inequality by using the matrix exterior algebra.

Author Contributions

Conceptualization, H.S.; writing-original draft, Y.L.; writing-review and editing, J.W.; funding acquisition, J.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the General Project of Science and Technology Plan of Beijing Municipal Education Commission (Grant No. KM202010037003).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

J.W. gratefully thank Editors and Reviewers for their comments and suggestions.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

References

  1. Treanta, S.; Arana-Jimenez, M.; Antczak, T. A necessary and sufficient condition on the equivalence between local and global optimal solutions in variational control problems. Nonlinear Anal. 2020, 191, 111640. [Google Scholar] [CrossRef]
  2. Bellman, R. Introduction to Matrix Analysis; McGraw-Hill: New York, NY, USA, 1960. [Google Scholar]
  3. Anderson, W.N., Jr. Shorted operators. SIAM J. Appl. Math. 2006, 20, 520–525. [Google Scholar] [CrossRef]
  4. Carlen, E. Trace inequalities and quantum entropy: An introductory course. Contemp. Math. 2010, 529, 73–140. [Google Scholar]
  5. Golden, S. Lower bounds for Helmholtz function. Phys. Rev. 1965, 137, 1127–1128. [Google Scholar] [CrossRef]
  6. Thompson, C.J. Inequality with applications in statistical mechanics. J. Math. Phys. 1965, 6, 1812–1813. [Google Scholar] [CrossRef] [Green Version]
  7. Lenard, A. Generalization of the Golden-Thompson inequality Tr(eAeB) ≥ TreA+B. Zndiana Univ. Math. 1971, 21, 457–467. [Google Scholar] [CrossRef]
  8. Ando, T.; Hiai, F. Log majorization and complementary Golden-Thompson type inequalities. Linear Algebra Appl. 1994, 197–198, 113–131. [Google Scholar] [CrossRef] [Green Version]
  9. Araki, H. On an inequality of Lieb and Thirring. Lett. Math. Phys. 1990, 19, 167–170. [Google Scholar] [CrossRef]
  10. Lieb, E.H.; Thirring, W. Inequalities for the moments of the eigenvalues of the Schrodinger Hamiltonian and their relation to Sobolev inequalities. In Studies in Mathematical Physics; Lieb, E.H., Simon, B., Wightman, A., Eds.; Princeton University Press: Princeton, NJ, USA, 1976; pp. 269–303. [Google Scholar]
  11. Marshall, A.W.; Olkin, I.; Arnold, B.C. Inequalities: Theory of Majorization and Its Applications; Springer: New York, NY, USA, 2011. [Google Scholar]
  12. Huang, D. Generalizing Lieb’s concavity theorem via operator interpolation. Adv. Math. 2020, 369, 107208. [Google Scholar] [CrossRef]
  13. Furuta, T. AB ≥ 0 assures (BrApBr) 1 q B p + 2 r q for r ≥ 0, p > 0, q ≥ 1 with 1 + 2 r q p + 2r. Proc. Am. Math. Sot. 1987, 101, 85–88. [Google Scholar]
  14. Bhatia, R. Matrix Analysis; Springer: New York, NY, USA, 1997. [Google Scholar]
  15. Bhatia, R. Positive Definite Matrices; Princeton University Press: Princeton, NJ, USA; Oxford, UK, 2007. [Google Scholar]
  16. Hall, B. Lie Groups, Lie Algebras, and Representations: An Elementary Introduction; Springer: New York, NY, USA, 2003. [Google Scholar]
  17. Zhang, F. Matrix Theory: Basic Results and Techniques; Springer: New York, NY, USA, 1999. [Google Scholar]
  18. Simon, B. Trace Ideals and Their Applications, 2nd ed.; Math. Surveys and Monographs. A.M.S.: Cambridge, UK, 2005. [Google Scholar]
  19. Tropp, J.A. An introduction to matrix concentration inequalities. Found. Trends Mach. Learn. 2015, 8, 1–2. [Google Scholar] [CrossRef]
  20. Davis, C. Notions generalizing convexity for functions defined on spaces of matrices. In Proceedings of the Symposia in Pure Mathematics; American Mathematical Society: Providence, RI, USA, 1963; Volume 7, pp. 187–201. [Google Scholar]
  21. Donogue, W. Monotone Matrix Functions and Analytic. Continuution; Springer: New York, NY, USA, 1974. [Google Scholar]
  22. Bellman, R. Inequalities; Springer Verlag: Berlin/Gottingen/Heidelberg, Germany, 1961. [Google Scholar]
  23. Hansen, F. Golden-Thompson’s inequality for deformed exponentials. J. Stat. Phys. 2015, 159, 1300–1305. [Google Scholar] [CrossRef] [Green Version]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Li, Y.; Wang, J.; Sun, H. Generalization of the Lieb–Thirring–Araki Inequality and Its Applications. Mathematics 2021, 9, 723. https://doi.org/10.3390/math9070723

AMA Style

Li Y, Wang J, Sun H. Generalization of the Lieb–Thirring–Araki Inequality and Its Applications. Mathematics. 2021; 9(7):723. https://doi.org/10.3390/math9070723

Chicago/Turabian Style

Li, Yonggang, Jing Wang, and Huafei Sun. 2021. "Generalization of the Lieb–Thirring–Araki Inequality and Its Applications" Mathematics 9, no. 7: 723. https://doi.org/10.3390/math9070723

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop