1 Introduction
The characterization of Euclidean spaces among normed spaces, or Hilbert spaces among Banach spaces, is a classical theme in functional analysis. It can be traced back to the Jordan–von Neumann theorem [Reference Jordan and Von Neumann3], which states that a normed space is Euclidean if and only if it satisfies the parallelogram identity. This line of research has been very fruitful: for example, the monograph [Reference Amir1] compiles about 350 characterizations of inner product spaces.
We consider vector spaces over a field
$\mathbf {K}$
which is either
$\mathbf {R}$
or
$\mathbf {C}$
. Recall that an inner product on a vector space X is a map
$\langle \cdot , \cdot \rangle : X \times X \to \mathbf {K}$
that satisfies the usual axioms of conjugate symmetry, linearity in one variable and positive-definiteness. If
$\|\cdot \|$
is a norm on X, we say that the normed space
$(X,\|\cdot \|)$
is an inner product space if there exists an inner product
$\langle \cdot , \cdot \rangle $
on X such that the identity
$\|x\|^2 = \langle x , x\rangle $
holds for every
$x \in X$
. A finite-dimensional inner product space over the real field is also called a Euclidean space.
In this note, we consider a characterization of finite-dimensional inner product spaces by a special property of their endomorphisms. Recall that the operator norm of a linear operator
$u : X \to X$
, denoted
$\| u \|_{\mathrm {op}}$
, is defined as the smallest
$C \geq 0$
such that the inequality
$\|u(x) \| \leq C \|x\|$
holds for every
$x \in X$
. We consider the set
$\mathcal {N}(u)$
of norming vectors, defined as
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20250117094950355-0492:S0008439524000985:S0008439524000985_eqnu1.png?pub-status=live)
The following theorem has been proved by Sain and Paul (see [Reference Sain and Paul5, Theorem 2.2]) in the real case only. Our goal is to provide a completely different and self-contained proof, which extends naturally to the complex case.
Theorem. Let X be a finite-dimensional normed space over the real or complex field. The following are equivalent.
-
(1) The space X is an inner product space.
-
(2) For every linear operator
$u:X \to X$ , the set
$\mathcal {N}(u)$ is a linear subspace.
The implication (1)
$\Rightarrow $
(2) can be shown as follows. If X is an inner product space, we may consider the adjoint operator
$u^* : X \to X$
. Let
$\lambda _{\max }(u^*u)$
denote the largest eigenvalue of the self-adjoint operator
$u^*u$
and E be the associated eigenspace. It is elementary to check then for every
$x \in X$
,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20250117094950355-0492:S0008439524000985:S0008439524000985_eqnu2.png?pub-status=live)
with equality if and only if
$x \in E$
. Since
$\lambda _{\max }(u^*u)=\|u\|_{\mathrm {op}}^2$
, it follows that the set
$\mathcal {N}(u)$
coincides with E. In particular,
$\mathcal {N}(u)$
is a linear subspace.
In the following sections, we prove the harder implication (2)
$\Rightarrow $
(1). Our arguments are inherently finite-dimensional; we leave open the question of whether the theorem extends to infinite-dimensional normed spaces.
2 The main proposition
In this section, we introduce our main tool. It is strongly related to the concept of positive John position of convex bodies which is studied in [Reference Artstein-Avidan and Putterman2]. We use a slightly different approach that allows us to cover also the complex case in a natural way.
Consider a norm
$\|\cdot \|$
on
$\mathbf {K}^n$
and equip the algebra
$\mathsf {M}_n(\mathbf {K})$
of
$n \times n$
matrices with the corresponding operator norm
$\| \cdot \|_{\mathrm {op}}$
. We denote by
$\mathsf {GL}_n(\mathbf {K})$
the group of invertible matrices. Given
$Q \in \mathsf {GL}_n(\mathbf {K})$
, we consider the set
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20250117094950355-0492:S0008439524000985:S0008439524000985_eqnu3.png?pub-status=live)
Let
$\mathsf {M}_n^+(\mathbf {K})$
be the cone of positive semi-definite matrices. The set
$\mathsf {M}_n^+(\mathbf {K}) \cap \mathcal {C}_Q$
is compact and therefore contains an element of maximal determinant. (This element is unique but we do not need this information.) Such an element A satisfies
$\| AQ \|_{\mathrm {op}} = 1$
.
Proposition. Let
$Q \in \mathsf {GL}_n(\mathbf {K})$
and A of maximal determinant in
$\mathsf {M}_n^+(\mathbf {K}) \cap \mathcal {C}_Q$
. Then the set
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20250117094950355-0492:S0008439524000985:S0008439524000985_eqnu4.png?pub-status=live)
spans
$\mathbf {K}^n$
as a vector space.
We show in the next section how the Proposition implies our Theorem. In the real case, the Proposition follows easily from [Reference Artstein-Avidan and Putterman2, Theorem 1.2]. For completeness, we include a self-contained proof.
Proof of the Proposition. Introduce the unit ball
$\mathcal {B} = \{ x \in \mathbf {K}^n \ : \ \|x\| \leq 1\}$
. If we identify the dual space with
$\mathbf {K}^n$
, the unit ball for the dual norm is
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20250117094950355-0492:S0008439524000985:S0008439524000985_eqnu5.png?pub-status=live)
By the Hahn–Banach theorem, we have for
$x \in \mathbf {K}^n$
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20250117094950355-0492:S0008439524000985:S0008439524000985_eqnu6.png?pub-status=live)
Consider the compact set
$T = \mathcal {B} \times \mathcal {B}^*$
and let
$C(T)$
be the Banach space of continuous functions from T to
$\mathbf {K}$
. We define a map
$\alpha : \mathsf {M}_n(\mathbf {K}) \to C(T)$
as follows: if
$M \in \mathsf {M}_n(\mathbf {K})$
,
$x \in \mathcal {B}$
and
$y \in \mathcal {B}^*$
, then
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20250117094950355-0492:S0008439524000985:S0008439524000985_eqnu7.png?pub-status=live)
The map
$\alpha $
is linear and satisfies
$\| \alpha (M) \| = \| M \|_{\mathrm {op}}$
for every
$M \in \mathsf {M}_n(\mathbf {K})$
.
Denote
$\mathcal {N}=\mathcal {N}(AQ)$
. Assume by contradiction that the set
$\mathcal {N}$
does not span
$\mathbf {K}^n$
. Its linear image
$(A^{1/2}Q)(\mathcal {N})$
does not span
$\mathbf {K}^n$
either, and therefore there exists a nonzero orthogonal projection P such that
$PA^{1/2}Qx = 0$
for every
$x \in \mathcal {N}$
. Let H be the matrix defined as
$H= \lambda P- \mathrm {Id}$
, where
$\lambda $
is a positive number chosen so that
$\operatorname {\mathrm {Tr}} H>0$
.
Introduce functions f,
$g \in C(T)$
defined as
$f = \alpha (AQ)$
and
$g = \alpha (A^{1/2}HA^{1/2}Q)$
. Observe that
$\|f\|=\| AQ \|_{\mathrm {op}} =1$
. For
$t=(x,y) \in T$
, we have a chain of implications
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20250117094950355-0492:S0008439524000985:S0008439524000985_eqnu8.png?pub-status=live)
Consequently, the functions f and g satisfy the hypothesis of the following lemma, whose proof is postponed.
Lemma. Let T be a nonempty compact topological space and
$f,g \in C(T)$
. Assume that for every
$t \in T$
such that
$|f(t)|=\|f\|$
, we have
$\Re f(t)\overline {g(t)} <0$
. Then, for
$\delta>0$
small enough, we have
$\|f + \delta g\| < \|f\|$
.
The lemma implies that for
$\delta>0$
small enough, we have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20250117094950355-0492:S0008439524000985:S0008439524000985_eqnu9.png?pub-status=live)
and thus
$A^{1/2}(\mathrm {Id} + \delta H)A^{1/2} \in \mathcal {C}_Q$
. As
$\delta $
goes to zero, we have
$\det (\mathrm {Id} + \delta H) = 1 + \delta \operatorname {\mathrm {Tr}} H+ o(\delta )$
and therefore
$\det (\mathrm {Id} + \delta H)>1$
for
$\delta>0$
small enough. The inequality
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20250117094950355-0492:S0008439524000985:S0008439524000985_eqnu10.png?pub-status=live)
contradicts the maximality of A.
It remains to prove the lemma.
Proof of the Lemma. Both f and g are nonzero (otherwise the hypothesis fails) and we may assume by rescaling that
$\|f\|=\|g\|=1$
. Let
$T_1$
be the nonempty closed subset of T defined as
$T_1 = \{t \in T \ : \ |f(t)|=1 \}$
. Since the function
${t \mapsto \Re f(t)\overline {g(t)}}$
is continuous, it achieves its maximum on
$T_1$
and therefore there exists
$\varepsilon>0$
such that
$\Re f(t)\overline {g(t)} \leq -\varepsilon $
for every
$t \in T_1$
. Denote by
$T_2$
the closed subset of T defined as
$T_2 = \{t \in T \ : \ \Re f(t)\overline {g(t)} \geq -\varepsilon \}$
. Since f is continuous, there exists
$\eta>0$
such that
$|f(t)| \leq 1-\eta $
for every
$t \in T_2$
. For
$t \in T$
and
$\delta>0$
, we compute
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20250117094950355-0492:S0008439524000985:S0008439524000985_eqnu11.png?pub-status=live)
It follows that
$\| f +\delta g\|^2 \leq \max ((1-\eta )^2 + 2 \delta + \delta ^2,1 - 2 \delta \varepsilon + \delta ^2)$
and therefore
${\| f +\delta g\|<1}$
for
$\delta>0$
small enough.
3 Proof of the Theorem
The implication (1)
$\Longrightarrow $
(2) has been proved in the introduction. Conversely, let X be a finite-dimensional normed space satisfying condition (2) from the Theorem. Let
$\mathsf {Iso}(X)$
be the group of isometries of X, defined as
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20250117094950355-0492:S0008439524000985:S0008439524000985_eqnu12.png?pub-status=live)
There is an inner product on X which is invariant with respect to
$\mathsf {Iso}(X)$
(see for example [Reference Onishchik and Vinberg4, p. 131, Theorem 2]) and therefore, without loss of generality, we may assume that
$X=(\mathbf {R}^n,\|\cdot \|)$
and that
$\mathsf {Iso}(X)$
is a subgroup of the orthogonal group
$\mathsf {O}_n$
. (From now on we consider only the real case, the complex case is similar using the unitary group
$\mathsf {U}_n$
.)
Fix
$Q \in \mathsf {O}_n$
and let A be an element of maximal determinant in
$\mathsf {M}_n^+(\mathbf {K}) \cap \mathcal {C}_Q$
. The set
$\mathcal {N}(AQ)$
is a linear subspace of
$\mathbf {R}^n$
(by hypothesis (2) from the Theorem) which spans
$\mathbf {R}^n$
(by the conclusion of the Proposition). It follows that
$\mathcal {N}(AQ) = \mathbf {R}^n$
and therefore that
$AQ \in \mathsf {Iso}(X) \subset \mathsf {O}_n$
. The matrix
$A = (AQ)Q^{-1}$
is both orthogonal and positive semidefinite; it follows that A is the identity matrix and therefore
$Q \in \mathsf {Iso}(X)$
.
The previous paragraph shows that
$\mathsf {Iso}(X) = \mathsf {O}_n$
. As a consequence, for every x and y in the Euclidean unit sphere
$S^{n-1}$
, there exists
$u \in \mathsf {Iso}(X)$
such that
$u(x)=y$
. It follows that the norm
$\|\cdot \|$
is constant on
$S^{n-1}$
, hence is a multiple of the Euclidean norm. This shows that X is an inner product space.
Acknowledgments
We thank Miguel Martín for pointing to us the reference [Reference Sain and Paul5].