1 Introduction

In [16], the author succeeded in providing a useful representation of the orthogonally additive homogeneous polynomials on the spaces \(L^p([0,1])\) and \(\ell ^p\) with \(1\le p<\infty \). In [12] (see also [6]), the authors obtained a similar representation for the space C(K), for a compact Hausdorff space K. These results were generalized to Banach lattices [4] and Riesz spaces  [9]. Further, the problem of representing the orthogonally additive homogeneous polynomials has been also considered in the context of Banach function algebras [1, 19] and non-commutative Banach algebras [2, 3, 11]. Notably, [11] can be thought of as the natural non-commutative analogue of the representation of orthogonally additive polynomials on C(K)-spaces, and the purpose to this paper is to extend the results of [16] on the representation of orthogonally additive homogeneous polynomials on \(L^p\)-spaces to the non-commutative \(L^p\)-spaces.

The non-commutative \(L^p\)-spaces that we consider are those associated with a von Neumann algebra \({{\mathscr {M}}}\) equipped with a normal semifinite faithful trace \(\tau \). From now on, \(S({{\mathscr {M}}},\tau )\) stands for the linear span of the positive elements x of \({{\mathscr {M}}}\) such that \(\tau \bigl ({{\,\mathrm{supp}\,}}(x)\bigr )<\infty \); here \({{\,\mathrm{supp}\,}}(x)\) stands for the support of x. Then \(S({{\mathscr {M}}},\tau )\) is a \(*\)-subalgebra of \({{\mathscr {M}}}\) with the property that \(\vert x\vert ^p\in S({{\mathscr {M}}},\tau )\) for each \(x\in S({{\mathscr {M}}},\tau )\) and each \(0<p<\infty \). For \(0<p<\infty \), we define \(\Vert \cdot \Vert _p:S({{\mathscr {M}}},\tau )\rightarrow {\mathbb {R}}\) by \(\Vert x\Vert _p=\tau (\vert x\vert ^p)^{1/p}\)\((x\in S({{\mathscr {M}}},\tau ))\). Then \(\Vert \cdot \Vert _p\) is a norm or a p-norm according to \(1\le p<\infty \) or \(0<p<1\), and the space \(L^p({{\mathscr {M}}},\tau )\) can be defined as the completion of \(S({{\mathscr {M}}},\tau )\) with respect to \(\Vert \cdot \Vert _p\). Nevertheless, for our purposes here, it is important to realize the elements of \(L^p({{\mathscr {M}}},\tau )\) as measurable operators. Specifically, the set \(L^0({{\mathscr {M}}},\tau )\) of measurable closed densely defined operators affiliated to \({{\mathscr {M}}}\) is a topological \(*\)-algebra with respect to the strong sum, the strong product, the adjoint operation, and the topology of the convergence in measure. The algebra \({{\mathscr {M}}}\) is a dense \(*\)-subalgebra of \(L^0({{\mathscr {M}}},\tau )\), the trace \(\tau \) extends to the positive cone of \(L^0({{\mathscr {M}}},\tau )\) in a natural way, and we can define

$$\begin{aligned}\begin{gathered} \Vert x\Vert _p=\tau \bigl (\vert x\vert ^p\bigr )^{1/p}\quad (x\in L^0({{\mathscr {M}}},\tau )), \\ L^p({{\mathscr {M}}},\tau )=\bigl \{x\in L^0({{\mathscr {M}}},\tau ) : \Vert x\Vert _p<\infty \bigr \}. \end{gathered}\end{aligned}$$

Also we set \(L^\infty ({{\mathscr {M}}},\tau )={{\mathscr {M}}}\) (with \(\Vert \cdot \Vert _\infty :=\Vert \cdot \Vert \), the operator norm). Operators \(x,y\in L^0({{\mathscr {M}}},\tau )\) are mutually orthogonal, written \(x\perp y\), if \(xy^*=y^*x=0\). This condition is equivalent to requiring that x and y have mutually orthogonal left, and right, supports. Further, for \(x,y\in L^p({{\mathscr {M}}},\tau )\) with \(0<p<\infty \), the condition \(x\perp y\) implies that \(\Vert x+y\Vert _p^p=\Vert x\Vert _p^p+\Vert y\Vert _p^p\), and conversely, if \(\Vert x\pm y\Vert _p^p=\Vert x\Vert _p^p+\Vert y\Vert _p^p\) and \(p\ne 2\), then \(x\perp y\) (see [14, Fact 1.3]). The orthogonal additivity considered in [16] for the spaces \(L^p([0,1])\) and \(\ell ^p\) can of course equally well be considered for the space \(L^p({{\mathscr {M}}},\tau )\). Let P be a map from \(L^p({{\mathscr {M}}},\tau )\) into a linear space X. Then P is:

  1. (i)

    orthogonally additive on a subset \({\mathscr {S}}\) of \(L^p({{\mathscr {M}}},\tau )\) if

    $$\begin{aligned} x,y \in {\mathscr {S}}, \, x\perp y=0 \ \Rightarrow \ P(x+y)=P(x)+P(y); \end{aligned}$$
  2. (ii)

    an m-homogeneous polynomial if there exists an m-linear map \(\varphi \) from \(L^p({{\mathscr {M}}},\tau )^m\) into X such that

    $$\begin{aligned} P(x)=\varphi \left( x,\dotsc ,x \right) \quad (x\in L^p({{\mathscr {M}}},\tau )). \end{aligned}$$

    Here and subsequently, \(m\in {\mathbb {N}}\) is fixed with \(m\ge 2\) and the superscript m stands for the m-fold Cartesian product. Such a map is unique if it is required to be symmetric. Further, in the case where X is a topological linear space, the polynomial P is continuous if and only if the symmetric m-linear map \(\varphi \) associated with P is continuous.

Given a continuous linear map \(\varPhi :L^{p/m}({{\mathscr {M}}},\tau )\rightarrow X\), where X is an arbitrary topological linear space, the map \(P_\varPhi :L^p({{\mathscr {M}}},\tau )\rightarrow X\) defined by

$$\begin{aligned} P_\varPhi (x)=\varPhi (x^m)\quad (x\in L^p({{\mathscr {M}}},\tau )) \end{aligned}$$

is a natural example of a continuous m-homogeneous polynomial which is orthogonally additive on \(L^p({{\mathscr {M}}},\tau )_{\mathrm{sa}}\) (Theorem 4), and we will prove that every continuous m-homogeneous polynomial which is orthogonally additive on \(L^p({{\mathscr {M}}},\tau )_{\mathrm{sa}}\) is actually of this special form (Theorem 5). Here and subsequently, the subscripts “sa” and \(+\) are used to denote the self-adjoint and the positive parts of a given subset of \(L^0({{\mathscr {M}}},\tau )\), respectively.

We require a few remarks about the setting of our present work. Throughout the paper we are concerned with m-homogeneous polynomials on the space \(L^p({{\mathscr {M}}},\tau )\) with \(0<p\), and thus one might wish to consider polynomials with values in the space \(L^q({{\mathscr {M}}},\tau )\), especially with \(q\le p\). Further, in the case case where \(p/m<1\) and the von Neumann algebra \({{\mathscr {M}}}\) has no minimal projections, there are no non-zero continuous linear functionals on \(L^{p/m}({{\mathscr {M}}},\tau )\); since one should like to have non-trivial “orthogonally additive” polynomials on \(L^p({{\mathscr {M}}},\tau )\), some weakening of the normability must be allowed to the range space (see Corollary 2). For these reasons, throughout the paper, X will be a (complex and Hausdorff) topological linear space. In the case where the von Neumann algebra \({{\mathscr {M}}}\) is commutative, the prototypical polynomials \(P_\varPhi \) mentioned above are easily seen to be orthogonally additive on the whole domain. In contrast, we will point out in Propositions 1 and 3 that this is not the case for the von Neumann algebra \({{\mathscr {B}}}(H)\) of all bounded operators on a Hilbert space H whenever \(\dim H\ge 2\).

We assume a basic knowledge of \(C^*\)-algebras and von Neumann algebras, tracial non-commutative \(L^p\)-spaces, and polynomials on topological linear spaces. For the relevant background material concerning these topics, see [5, 7, 8, 10, 13, 17, 18].

2 \(C^*\)-algebras and von Neumann algebras

Our approach to the problem of representing the orthogonally additive homogeneous polynomials on the non-commutative \(L^p\)-spaces relies on the representation of those polynomials on the von Neumann algebras.

Recall that two elements x and y of a \(C^*\)-algebra \({\mathscr {A}}\) are mutually orthogonal if \(xy^*=y^*x=0\), in which case the identity \(\Vert x+y\Vert =\max \{\Vert x\Vert ,\Vert y\Vert \}\) holds. The reader should be aware that we have chosen the standard definition of orthogonality in the setting of non-commutative \(L^p\)-spaces. This definition is slightly different from the one used in [11], which is the standard one in the setting of Banach algebras. In [11] the orthogonality of two elements x and y is defined by the relation \(xy=yx=0\), and, further, the orthogonally additive polynomials on the self-adjoint part of a \(C^*\)-algebra are automatically orthogonally additive on the whole algebra. The important point to note here is that both the definitions of orthogonality agree on the self-adjoint part of the \(C^*\)-algebra. Thus, for a polynomial on a \(C^*\)-algebra, the property of being orthogonally additive on the self-adjoint part according to our definition is the same as being orthogonally additive according to [11]. Nevertheless, in contrast to [11], there are no non-zero orthogonally additive polynomials from the von Neumann algebra \({{\mathscr {B}}}(H)\) into any topological Banach space according to our definition (Proposition 1).

Suppose that \({\mathscr {A}}\) is a linear space with an involution \(*\). Recall that for a linear functional \(\varPhi :{\mathscr {A}}\rightarrow {\mathbb {C}}\), the map \(\varPhi ^*:{\mathscr {A}}\rightarrow {\mathbb {C}}\) defined by \(\varPhi ^*(x)=\overline{\varPhi (x^{*})}\)\((x\in {\mathscr {A}})\) is a linear functional, and \(\varPhi \) is said to be hermitian if \(\varPhi ^*=\varPhi \). Similarly, for an m-homogeneous polynomial \(P:{\mathscr {A}}\rightarrow {\mathbb {C}}\), the map \(P^*:{\mathscr {A}}\rightarrow {\mathbb {C}}\) defined by \(P^*(x)=\overline{P(x^*)}\)\((x\in {\mathscr {A}})\) is an m-homogeneous polynomial, and we call P hermitian if \(P^*=P\).

Lemma 1

Let X and Y be linear spaces, and let \(P:X\rightarrow Y\) be an m-homogeneous polynomial. Suppose that P vanishes on a convex set \(C\subset X\). Then P vanishes on the linear span of C.

Proof

Set \(x_1,x_2,x_3,x_4\in C\). Let \(\eta :Y\rightarrow {\mathbb {C}}\) be a linear functional, and define \(f:{\mathbb {C}}^4\rightarrow {\mathbb {C}}\) by

$$\begin{aligned} f(\alpha _1,\alpha _2,\alpha _3,\alpha _4)= \eta \bigl (P(\alpha _1x_1+\alpha _2 x_2+\alpha _3x_3+\alpha x_4)\bigr )\quad (\alpha _1,\alpha _2,\alpha _3,\alpha _4\in {\mathbb {C}}). \end{aligned}$$

Then f is a complex polynomial function in four complex variables that vanishes on the set

$$\begin{aligned} \bigl \{(\rho _1,\rho _2,\rho _3,\rho _4)\in {\mathbb {R}}^4 : 0\le \rho _1,\rho _2,\rho _3,\rho _4, \ \rho _1+\rho _2+\rho _3+\rho _4=1\bigr \}. \end{aligned}$$

This implies that f is identically equal to 0 on \({\mathbb {C}}^4\), and, in particular,

$$\begin{aligned} \eta \bigl (P(\rho _1x_1-\rho _2x_2+i\rho _3x_3-i\rho _4x_4)\bigr )=f(\rho _1,-\rho _2,i\rho _3,-i\rho _4)=0 \end{aligned}$$

for all \(\rho _1,\rho _2,\rho _3,\rho _4\ge 0\). Since this identity holds for each linear functional \(\eta \), it may be concluded that \(P(\rho _1x_1-\rho _2x_2+i\rho _3x_3-i\rho _4x_4)=0\) for all \(\rho _1,\rho _2,\rho _3,\rho _4\ge 0\). Thus P vanishes on the set

$$\begin{aligned} \bigl \{ \rho _1x_1-\rho _2x_2+i\rho _3x_3-i\rho _4x_4 : \rho _j\ge 0, \ x_j\in C \ (j=1,2,3,4) \bigr \}, \end{aligned}$$

which is exactly the linear span of the set C. \(\square \)

Theorem 1

Let \({\mathscr {A}}\) be a \(C^*\)-algebra, let X be a topological linear space, and let \(\varPhi :{\mathscr {A}}\rightarrow X\) be a continuous linear map. Then:

  1. (i)

    the map \(P_\varPhi :{\mathscr {A}}\rightarrow X\) defined by \(P_\varPhi (x)=\varPhi (x^m)\)\((x\in {\mathscr {A}})\) is a continuous m-homogeneous polynomial which is orthogonally additive on \({\mathscr {A}}_{\mathrm{sa}}\);

  2. (ii)

    the polynomial \(P_\varPhi \) is uniquely specified by the map \(\varPhi \).

Suppose, further, that X is a q-normed space, \(0<q\le 1\). Then:

  1. (iii)

    \(2^{-1/q}\Vert \varPhi \Vert \le \Vert P_\varPhi \Vert \le \Vert \varPhi \Vert \).

Moreover, in the case where \(X={\mathbb {C}}\),

  1. (iv)

    the functional \(\varPhi \) is hermitian if and only if the polynomial \(P_\varPhi \) is hermitian, in which case \(\Vert P_\varPhi \Vert =\Vert \varPhi \Vert \).

Proof

(i) It is clear that the map \(P_\varPhi \) is continuous and that \(P_\varPhi \) is the m-homogeneous polynomial associated with the symmetric m-linear map \(\varphi :{\mathscr {A}}^m\rightarrow X\) defined by

$$\begin{aligned} \varphi (x_1,\ldots ,x_m)= \frac{1}{m!}\sum _{\sigma \in \mathfrak {S}_m} \varPhi \left( x_{\sigma (1)} \cdots x_{\sigma (m)} \right) \quad (x_1,\ldots ,x_m\in {\mathscr {A}}); \end{aligned}$$

here and subsequently, we write \(\mathfrak {S}_m\) for the symmetric group of order m.

Suppose that \(x,y\in {\mathscr {A}}_{\mathrm{sa}}\) are such that \(x\perp y\). Then \(xy=yx=0\), and so \((x+y)^m=x^m+y^m\), which gives

$$\begin{aligned} P_\varPhi (x+y)=\varPhi \bigl ((x+y)^m\bigr )= \varPhi \bigl (x^m+y^m\bigr )=\varPhi \bigl (x^m\bigr )+\varPhi \bigl (y^m\bigr )= P_\varPhi (x)+P_\varPhi (y). \end{aligned}$$

(ii) Assume that \(\Psi :{\mathscr {A}}\rightarrow X\) is a linear map with the property that \(P_\Psi =P_\varPhi \). If \(x\in {\mathscr {A}}_+\), then

$$\begin{aligned} \varPhi (x)=\varPhi \bigl ((x^{1/m})^{m}\bigr )=P(x^{1/m})=\Psi \bigl ((x^{1/m})^{m}\bigr )=\Psi (x). \end{aligned}$$

By linearity we also get \(\Psi (x) = \varPhi (x)\) for each \(x \in {\mathscr {A}}\).

(iii) Next, assume that X is a q-normed space. For each \(x\in {\mathscr {A}}\), we have

$$\begin{aligned} \Vert P_\varPhi (x)\Vert =\Vert \varPhi (x^m)\Vert \le \Vert \varPhi \Vert \Vert x^m\Vert \le \Vert \varPhi \Vert \Vert x\Vert ^m, \end{aligned}$$

which implies that \(\Vert P_\varPhi \Vert \le \Vert \varPhi \Vert \). Now take \(x\in {\mathscr {A}}\), and let \(\omega \in {\mathbb {C}}\) with \(\omega ^m=-1\). Then \(x=\mathfrak {R}x+i\mathfrak {I}x\), where

$$\begin{aligned} \mathfrak {R}x=\frac{1}{2}(x^*+x), \, \mathfrak {I}x=\frac{i}{2}(x^*-x)\in {\mathscr {A}}_{\mathrm{sa}}, \end{aligned}$$

and, further, \(\Vert \mathfrak {R}x\Vert ,\Vert \mathfrak {I}x\Vert \le \Vert x\Vert \). Moreover, \(\mathfrak {R}x=x_1-x_2\) and \(\mathfrak {I}x=x_3-x_4\), where \(x_1,x_2,x_3,x_4\in {\mathscr {A}}_+\), \(x_1\perp x_2\), and \(x_3\perp x_4\). Since \(x_1\perp x_2\) and \(x_3\perp x_4\), it follows that \(x_1^{1/m}\perp x_2^{1/m}\) and \(x_3^{1/m}\perp x_4^{1/m}\). Consequently,

$$\begin{aligned} {\begin{matrix} \Vert \mathfrak {R}x\Vert &{}=\max \bigl \{\Vert x_1\Vert ,\Vert x_2\Vert \bigr \},\\ \Vert \mathfrak {I}x\Vert &{}=\max \bigl \{\Vert x_3\Vert ,\Vert x_4\Vert \bigr \}, \end{matrix}} \end{aligned}$$
(1)

and

$$\begin{aligned} {\begin{matrix} {\big \Vert } x_1^{1/m}+\omega x_2^{1/m} {\big \Vert }&{}= \max \bigl \{ {\big \Vert } x_1^{1/m}{\big \Vert },{\big \Vert } x_2^{1/m} {\big \Vert } \bigr \},\\ {\big \Vert } x_3^{1/m}+\omega x_4^{1/m} {\big \Vert }&{}= \max \bigl \{ {\big \Vert }x_3^{1/m}{\big \Vert },{\big \Vert } x_4^{1/m} {\big \Vert } \bigr \}. \end{matrix}} \end{aligned}$$
(2)

Since

$$\begin{aligned} \big \Vert x_1^{1/m}\big \Vert = \Vert x_1\Vert ^{1/m}, \ \big \Vert x_2^{1/m}\big \Vert = \Vert x_2\Vert ^{1/m}, \ \big \Vert x_3^{1/m}\big \Vert = \Vert x_3\Vert ^{1/m}, \ \big \Vert x_4^{1/m}\big \Vert = \Vert x_4\Vert ^{1/m}, \end{aligned}$$

it follows, from (1) and (2), that

$$\begin{aligned} {\begin{matrix} {\big \Vert } x_1^{1/m}+\omega x_2^{1/m}{\big \Vert }^m &{}= \max \bigl \{\Vert x_1\Vert ,\Vert x_2\Vert \bigr \}=\Vert {\mathfrak {R}} x\Vert ,\\ {\big \Vert } x_3^{1/m}+\omega x_4^{1/m}{\big \Vert }^m &{}= \max \bigl \{\Vert x_3\Vert ,\Vert x_4\Vert \bigr \}=\Vert {\mathfrak {I}} x\Vert . \end{matrix}} \end{aligned}$$
(3)

On the other hand, we have

$$\begin{aligned} \bigl (x_1^{1/m}+\omega x_2^{1/m}\bigr )^m=x_1-x_2=\mathfrak {R}x,\quad \bigl (x_3^{1/m}+\omega x_4^{1/m}\bigr )^m=x_3-x_4=\mathfrak {I}x, \end{aligned}$$

and so

$$\begin{aligned} {\begin{matrix} \varPhi (x) &{}= \varPhi (\mathfrak {R}x)+i\varPhi (\mathfrak {I}x)= \varPhi \bigl (\bigl (x_1^{1/m}+\omega x_2^{1/m}\bigr )^m\bigr )+ i\varPhi \bigl (\bigl (x_3^{1/m}+\omega x_4^{1/m}\bigr )^m\bigr )\\ &{}= P_\varPhi \bigl (x_1^{1/m}+\omega x_2^{1/m}\bigr )+ iP_\varPhi \bigl (x_3^{1/m}+\omega x_4^{1/m}\bigr ). \end{matrix}} \end{aligned}$$

Hence, by (3),

$$\begin{aligned} \Vert \varPhi (x)\Vert ^q\le & {} \big \Vert P_\varPhi \bigl (x_1^{1/m}+\omega x_2^{1/m}\bigr )\big \Vert ^q+ \big \Vert P_\varPhi \bigl (x_3^{1/m}+\omega x_4^{1/m}\bigr )\big \Vert ^q\\\le & {} \Vert P_\varPhi \Vert ^q\big \Vert x_1^{1/m}+\omega x_2^{1/m}\big \Vert ^{mq}+ \Vert P_\varPhi \Vert ^q\big \Vert x_3^{1/m}+\omega x_4^{1/m}\big \Vert ^{mq}\\= & {} \Vert P_\varPhi \Vert ^q\left( \Vert \mathfrak {R}x\Vert ^q+\Vert \mathfrak {I}x\Vert ^q\right) \\\le & {} \Vert P_\varPhi \Vert ^q 2\Vert x\Vert ^q. \end{aligned}$$

This clearly forces \(\Vert \varPhi \Vert \le 2^{1/q}\Vert P_\varPhi \Vert \), as claimed.

(iv) It is straightforward to check that \(P_\varPhi ^*=P_{\varPhi ^*}\). Consequently, if \(\varPhi \) is hermitian, then \(P_\varPhi ^*=P_{\varPhi ^*}=P_\varPhi \) so that \(P_\varPhi \) is hermitian. Conversely, if \(P_\varPhi \) is hermitian, then \(P_{\varPhi ^*}=P_\varPhi ^*=P_\varPhi \) and (ii) implies that \(\varPhi ^*=\varPhi \). Finally, assume that \(\varPhi \) is a hermitian functional. For the calculation of \(\Vert P_\varPhi \Vert \) it suffices to check that \(\Vert \varPhi \Vert \le \Vert P_\varPhi \Vert \). For this purpose, let \(\varepsilon \in {\mathbb {R}}^+\), and choose \(x\in {\mathscr {A}}\) such that \(\Vert x\Vert =1\) and \(\Vert \varPhi \Vert -\varepsilon <\vert \varPhi (x)\vert \). We take \(\alpha \in {\mathbb {C}}\) with \(\vert \alpha \vert =1\) and \(\vert \varPhi (x)\vert =\alpha \varPhi (x)\), so that

$$\begin{aligned} \Vert \varPhi \Vert -\varepsilon <\vert \varPhi (x)\vert =\varPhi (\alpha x)= \overline{\varPhi (\alpha x)}=\varPhi \bigl ((\alpha x)^*\bigr ). \end{aligned}$$

Note that \(\Vert \mathfrak {R}(\alpha x)\Vert \le 1\) and \(\Vert \varPhi \Vert -\varepsilon <\varPhi (\mathfrak {R}(\alpha x))\). Now we consider the decomposition \(\mathfrak {R}(\alpha x)=x_1-x_2\) with \(x_1,x_2\in {\mathscr {A}}_+\) and \(x_1\perp x_2\) and take \(\omega \in {\mathbb {C}}\) with \(\omega ^m=-1\). As in (3), we see that \(\big \Vert x_1^{1/m}+\omega x_2^{1/m}\big \Vert =\Vert \mathfrak {R}(\alpha x)\Vert ^{1/m}\le 1\). Moreover, we have

$$\begin{aligned} P_\varPhi \bigl ( x_1^{1/m}+\omega x_2^{1/m}\bigr )= \varPhi \bigl (\bigl ( x_1^{1/m}+\omega x_2^{1/m}\bigr )^m\bigr )= \varPhi (\mathfrak {R}(\alpha x)), \end{aligned}$$

which gives \(\Vert \varPhi \Vert -\varepsilon <\Vert P_\varPhi \Vert \). \(\square \)

Lemma 2

Let \({\mathscr {A}}\) be a \(C^*\)-algebra, let \({\mathscr {R}}\) be a \(*\)-subalgebra of \({\mathscr {A}}\), let X be a topological linear space, and let \(\varPhi :{\mathscr {R}}\rightarrow X\) be a linear map. Suppose that the polynomial \(P:{\mathscr {R}}\rightarrow X\) defined by \(P(x)=\varPhi (x^m)\)\((x\in {\mathscr {R}})\) is continuous and that \({\mathscr {R}}\) satisfies the following conditions:

  1. (i)

    \(\vert x\vert \in {\mathscr {R}}\) for each \(x\in {\mathscr {R}}_{\mathrm{sa}}\);

  2. (ii)

    \(x^{1/m}\in {\mathscr {R}}\) for each \(x\in {\mathscr {R}}_+\).

Then \(\varPhi \) is continuous.

Proof

Let U be a neighbourhood of 0 in X. Let V be a balanced neighbourhood of 0 in X with \(V+V+V+V\subset U\). The set \(P^{-1}(V)\) is a neighbourhood of 0 in \({\mathscr {R}}\), which implies that there exists \(r\in {\mathbb {R}}^+\) such that \(P(x)\in V\) whenever \(x\in {\mathscr {R}}\) and \(\Vert x\Vert <r\). Take \(x\in {\mathscr {R}}\) with \(\Vert x\Vert <r^m\). Since \({\mathscr {R}}\) is a \(*\)-subalgebra of \({\mathscr {A}}\), we see that \(\mathfrak {R}x,\mathfrak {I}x\in {\mathscr {R}}_{\mathrm{sa}}\). We write \(\mathfrak {R}x=x_1-x_2\) and \(\mathfrak {I}x=x_3-x_4\), as in the proof of Theorem 1, where, on account of the condition (i),

$$\begin{aligned} x_1&=\tfrac{1}{2}\bigl (\vert \mathfrak {R}x\vert +\mathfrak {R}x\bigr )\in {\mathscr {R}}_+,&x_2&=\tfrac{1}{2}\bigl (\vert \mathfrak {R}x\vert -\mathfrak {R}x\bigr ) \in {\mathscr {R}}_+,\\ x_3&=\tfrac{1}{2}\bigl (\vert \mathfrak {I}x\vert +\mathfrak {I}x\bigr )\in {\mathscr {R}}_+,&x_4&=\tfrac{1}{2}\bigl (\vert \mathfrak {I}x\vert -\mathfrak {I}x\bigr ) \in {\mathscr {R}}_+. \end{aligned}$$

For each \(j\in \{1,2,3,4\}\), condition (ii) gives \(x_j^{1/m}\in {\mathscr {R}}\), and, further, we have \(\Vert x_{j}^{1/m}\big \Vert =\Vert x_j\Vert ^{1/m}\le \Vert x\Vert ^{1/m}<r\). Hence

$$\begin{aligned} \varPhi (x)&= \varPhi \bigl ( \bigl (x_{1}^{1/m}\bigr )^m-\bigl (x_{2}^{1/m}\bigr )^m+i\bigl (x_{3}^{1/m}\bigr )^m -i\bigl (x_{4}^{1/m}\bigr )^m\bigr ) \\&= \varPhi \bigl (\bigl (x_{1}^{1/m}\bigr )^m\bigr )- \varPhi \bigl (\bigl (x_{2}^{1/m}\bigr )^m\bigr )+ i\varPhi \bigl (\bigl (x_{3}^{1/m}\bigr )^m\bigr )- i\varPhi \bigl (\bigl (x_{4}^{1/m}\bigr )^m\bigr ) \\&= P\bigl (x_{1}^{1/m}\bigr )-P\bigl (x_{2}^{1/m}\bigr )+iP\bigl (x_{3}^{1/m}\bigr ) -iP\bigl (x_{4}^{1/m}\bigr )\in V+V+V+V\subset U, \end{aligned}$$

which establishes the continuity of \(\varPhi \). \(\square \)

Theorem 2

Let \({\mathscr {A}}\) be a \(C^*\)-algebra, let X be a locally convex space, and let \(P:{\mathscr {A}}\rightarrow X\) be a continuous m-homogeneous polynomial. Then the following conditions are equivalent:

  1. (i)

    there exists a continuous linear map \(\varPhi :{\mathscr {A}} \rightarrow X\) such that \(P(x)=\varPhi (x^m)\)\((x\in {\mathscr {A}})\);

  2. (ii)

    the polynomial P is orthogonally additive on \({\mathscr {A}}_{\mathrm{sa}}\);

  3. (iii)

    the polynomial P is orthogonally additive on \({\mathscr {A}}_+\).

If the conditions are satisfied, then the map \(\varPhi \) is unique.

Proof

Theorem 1 gives (i)\(\Rightarrow \)(ii), and obviously (ii)\(\Rightarrow \)(iii). The task is now to prove that (iii)\(\Rightarrow \)(i).

Suppose that (iii) holds. For each continuous linear functional \(\eta :X\rightarrow {\mathbb {C}}\), set \(P_\eta =\eta \circ P\). Then \(P_\eta \) is a complex-valued continuous m-homogeneous polynomial. We claim that \(P_\eta \) is orthogonally additive on \({\mathscr {A}}_{\mathrm{sa}}\). Take \(x,y\in {\mathscr {A}}_{\mathrm{sa}}\) with \(x\perp y\). Then we can write \(x=x_{+}-x_{-}\) and \(y=y_{+}-y_{-}\) with \(x_{+},x_{-},y_{+},y_{-}\in {\mathscr {A}}_{+}\) mutually orthogonal. Define \(f:{\mathbb {C}}^2\rightarrow {\mathbb {C}}\) by

$$\begin{aligned} f(\alpha ,\beta )\!= \!P_\eta (x_++\alpha x_-\!+\!y_++\beta y_-)\!-\! P_\eta (x_+\!+\!\alpha x_-)\!-\!P_\eta (y_+\!+\!\beta y_-)\quad (\alpha ,\beta \in {\mathbb {C}}^2). \end{aligned}$$

Then f is a complex polynomial function in two complex variables. If \(\alpha ,\beta \in {\mathbb {R}}^+\), then \(x_++\alpha x_-,y_++\beta y_-\in {\mathscr {A}}_+\) are mutually orthogonal, and so, by hypothesis, \(P(x_++\alpha x_-+y_++\beta y_-)=P(x_++\alpha x_-)+P(y_++\beta y_-)\). This shows that \(f(\alpha ,\beta )=0\). Since f vanishes on \({\mathbb {R}}^{+}\times {\mathbb {R}}^{+}\), it follows that f vanishes on \({\mathbb {C}}^2\), which, in particular, implies

$$\begin{aligned} P_\eta (x+y)-P_\eta (x)-P_\eta (y)=f(-1,-1)=0. \end{aligned}$$

Having proved that \(P_\eta \) is orthogonally additive on \({\mathscr {A}}_{\mathrm{sa}}\) we can apply [11, Theorem 2.8] to obtain a unique continuous linear functional \(\varPhi _\eta \) on \({\mathscr {A}}\) such that

$$\begin{aligned} \eta \bigl (P(x)\bigr )=\varPhi _\eta (x^m)\quad (x\in {\mathscr {A}}). \end{aligned}$$
(4)

Each \(x\in {\mathscr {A}}\) can be written in the form \(x_1^m+\cdots +x_k^m\) for suitable \(x_1,\ldots ,x_k\in {\mathscr {A}}\), and we define

$$\begin{aligned} \varPhi (x)=\sum _{j=1}^kP(x_j). \end{aligned}$$

Our next goal is to show that \(\varPhi \) is well-defined. Suppose that \(x_1,\ldots ,x_k\in {\mathscr {A}}\) are such that \(x_1^m+\cdots +x_k^m=0\). For each continuous linear functional \(\eta \) on X, (4) gives

$$\begin{aligned} \eta \left( \sum _{j=1}^kP(x_j)\right) = \sum _{j=1}^k\eta \bigl (P(x_j)\bigr )= \sum _{j=1}^k\varPhi _\eta (x_j^m)= \varPhi _\eta \left( \sum _{j=1}^k x_j^m\right) = 0. \end{aligned}$$

Since X is locally convex, we conclude that \(\sum _{j=1}^kP(x_j)=0\).

It is a simple matter to check that \(\varPhi \) is linear and, by definition, \(P(x)=\varPhi (x^m)\)\((x\in {\mathscr {A}})\). The continuity of \(\varPhi \) then follows from Lemma 2.

The uniqueness of the map \(\varPhi \) follows from Theorem 1(ii). \(\square \)

The assumption that the space X be locally convex can be removed by requiring that the \(C^*\)-algebra \({\mathscr {A}}\) be sufficiently rich in projections. The real rank zero is the most important existence of projections property in the theory of \(C^*\)-algebras. We refer the reader to [5, Section V.3.2] and [7, Section V.7] for the basic properties and examples of \(C^*\)-algebras of real rank zero. This class of \(C^*\)-algebras contains the von Neumann algebras and the \(C^*\)-algebras \({\mathscr {K}}(H)\) of all compact operators on any Hilbert space H. Let us remark that every \(C^*\)-algebra of real rank zero has an approximate unit of projections (but not necessarily increasing).

Theorem 3

Let \({\mathscr {A}}\) be a \(C^*\)-algebra of real rank zero, let X be a topological linear space, and let \(P:{\mathscr {A}}\rightarrow X\) be a continuous m-homogeneous polynomial. Suppose that \({\mathscr {A}}\) has an increasing approximate unit of projections. Then the following conditions are equivalent:

  1. (i)

    there exists a continuous linear map \(\varPhi :{\mathscr {A}} \rightarrow X\) such that \(P(x)=\varPhi (x^m)\)\((x\in {\mathscr {A}})\);

  2. (ii)

    the polynomial P is orthogonally additive on \({\mathscr {A}}_{\mathrm{sa}}\);

  3. (iii)

    the polynomial P is orthogonally additive on \({\mathscr {A}}_+\).

If the conditions are satisfied, then the map \(\varPhi \) is unique.

Proof

Theorem 1 gives (i)\(\Rightarrow \)(ii), and it is clear that (ii)\(\Rightarrow \)(iii). We will henceforth prove that (iii)\(\Rightarrow \)(i).

We first note that such a map \(\varPhi \) is necessarily unique, because of Theorem 1(ii).

Suppose that (iii) holds and that \({\mathscr {A}}\) is unital. Let \(\varphi :{\mathscr {A}}^m\rightarrow X\) be the symmetric m-linear map associated with P and define \(\varPhi :{\mathscr {A}}\rightarrow X\) by

$$\begin{aligned} \varPhi (x)=\varphi (x,1,\ldots ,1)\quad (x\in {\mathscr {A}}). \end{aligned}$$

Let \(Q:{\mathscr {A}}\rightarrow X\) be the m-homogeneous polynomial defined by

$$\begin{aligned} Q(x)=\varPhi (x^m)\quad (x\in {\mathscr {A}}). \end{aligned}$$

We will prove that \(P=Q\). On account of Lemma 1, it suffices to show that \(P(x)=Q(x)\) for each \(x\in {\mathscr {A}}_{\mathrm{sa}}\).

First, consider the case where \(x\in {\mathscr {A}}_{\mathrm{sa}}\) has finite spectrum, say \(\{\rho _1,\ldots ,\rho _k\}\subset {\mathbb {R}}\). This implies that x can be written in the form

$$\begin{aligned} x=\sum _{j=1}^k\rho _j e_j, \end{aligned}$$

where \(e_1,\ldots ,e_k\in {\mathscr {A}}\) are mutually orthogonal projections (specifically, the projection \(e_j\) is defined by using the continuous functional calculus for x by \(e_j=\chi _{\{\rho _j\}}(x)\) for each \(j\in \{1,\ldots ,k\}\)). We also set \(e_0=1-(e_1+\cdots +e_k)\), so that the projections \(e_0,e_1,\ldots ,e_k\) are mutually orthogonal, and \(\rho _0=0\). We claim that if \(j_1,\ldots ,j_m\in \{0,\ldots ,k\}\) and \(j_l\ne j_{l'}\) for some \(l,l'\in \{1,\ldots ,m\}\), then

$$\begin{aligned} \varphi (e_{j_1},\ldots ,e_{j_m})=0. \end{aligned}$$
(5)

Let \(\varLambda _1=\bigr \{n\in \{1,\ldots ,m\} : j_n=j_l\bigr \}\) and \(\varLambda _2=\bigr \{n\in \{1,\ldots ,m\} : j_n\ne j_l\bigr \}\). For each \(\alpha _1,\ldots ,\alpha _m\in {\mathbb {R}}^+\), the elements \(\sum _{n\in \varLambda _1}\alpha _{n}e_{j_n}\) and \(\sum _{n\in \varLambda _2}\alpha _{n}e_{j_n}\) are positive and mutually orthogonal, so that the orthogonal additivity of P on \({\mathscr {A}}_+\) gives

$$\begin{aligned} P\left( \sum _{n=1}^m\alpha _ne_{j_n} \right) = P\left( \sum _{n\in \varLambda _1}\alpha _n e_{j_n}\right) + P\left( \sum _{n\in \varLambda _2}\alpha _n e_{j_n}\right) . \end{aligned}$$

This implies that, for each linear functional \(\eta :X\rightarrow {\mathbb {C}}\), the function \(f:{\mathbb {C}}^m\rightarrow {\mathbb {C}}\) defined by

$$\begin{aligned} f(\alpha _1,\ldots ,\alpha _m)= \eta \left( P\left( \sum _{n=1}^m\alpha _n e_{j_n}\right) - P\left( \sum _{n\in \varLambda _1}\alpha _n e_{j_n}\right) - P\left( \sum _{n\in \varLambda _2}\alpha _n e_{j_n}\right) \right) , \end{aligned}$$

for all \(\alpha _1,\ldots ,\alpha _m\in {\mathbb {C}}\), is a complex polynomial function in m complex variables vanishing in \(\bigl ({\mathbb {R}}^+\bigr )^m\). Therefore f vanishes on \({\mathbb {C}}^m\). Moreover, we observe that the coefficient of the monomial \(\alpha _1\cdots \alpha _m\) is given by \(n!\eta \bigl (\varphi (e_{j_1},\ldots ,e_{j_m})\bigr )\), because both \(\varLambda _1\) and \(\varLambda _2\) are different from \(\{1,\ldots ,m\}\). We thus get

$$\begin{aligned}n!\eta \bigl (\varphi (e_{j_1},\ldots ,e_{j_m})\bigr )=0 .\end{aligned}$$

Since this identity holds for each linear functional \(\eta \), our claim follows. Property (5) now leads to

$$\begin{aligned} {\begin{matrix} P(x) &{} = \varphi \left( \sum _{j=1}^k\rho _{j} e_j,\ldots ,\sum _{j=1}^k\rho _{j}e_j\right) = \sum _{j_1, \ldots , j_m =1}^k\rho _{j_{1}}\cdots \rho _{j_{m}}\varphi \left( e_{j_1},\ldots , e_{j_m}\right) \\ &{}= \sum _{j=1}^{k} \rho _{j}^m\varphi \left( e_j,\ldots ,e_j\right) \end{matrix}} \end{aligned}$$

and

$$\begin{aligned} {\begin{matrix} Q(x) &{}=\! \varphi \left( \!\left( \!\sum _{j=0}^k\rho _j e_j \!\right) ^m,\sum _{j=0}^k e_j,\ldots , \sum _{j=0}^k e_j\!\right) \!=\! \varphi \left( \!\sum _{j=0}^k \rho _{j}^m e_j , \sum _{j=0}^k e_j,\ldots ,\sum _{j=0}^k e_j \!\right) \\ &{}= \sum _{j_1,\ldots , j_m =0}^k \rho _{j_1}^m \varphi \left( e_{j_1},\ldots ,e_{j_m}\right) = \sum _{j=1}^k \rho _{j}^m \varphi \left( e_j , \ldots ,e_j \right) . \end{matrix}} \end{aligned}$$

We thus get \(P(x)=Q(x)\).

Now suppose that \(x\in {\mathscr {A}}_{\mathrm{sa}}\) is an arbitrary element. Since \({\mathscr {A}}\) has real rank zero, it follows that there exists a sequence \((x_n)\) in \({\mathscr {A}}_{\mathrm{sa}}\) such that each \(x_n\) has finite spectrum and \(\lim x_n=x\). On account of the above case, we have \(P(x_n)=Q(x_n)\) (\(n\in {\mathbb {N}}\)), and the continuity of both P and Q now yields \(P(x)=\lim P(x_n)=\lim Q(x_n)=Q(x)\), as required.

We are now in a position to prove the non-unital case. By hypothesis, there exists an increasing approximate unit of projections \((e_\lambda )_{\lambda \in \varLambda }\). For each \(\lambda \in \varLambda \), set \({\mathscr {A}}_\lambda =e_\lambda {\mathscr {A}}e_\lambda \). Then \({\mathscr {A}}_\lambda \) is a unital \(C^*\)-algebra (with identity \(e_\lambda \)) and has real rank zero (because \({\mathscr {A}}_\lambda \) is a hereditary \(C^*\)-subalgebra of \({\mathscr {A}}\)). From what has previously been proved, it follows that there exists a unique continuous linear map \(\varPhi _\lambda :{\mathscr {A}}_\lambda \rightarrow X\) such that

$$\begin{aligned} P(x)=\varPhi _\lambda (x^m)\quad (x\in {\mathscr {A}}_\lambda ). \end{aligned}$$
(6)

Define

$$\begin{aligned} {\mathscr {R}}=\bigcup _{\lambda \in \varLambda }{\mathscr {A}}_\lambda \end{aligned}$$

and, for each \(x\in {\mathscr {R}}\), set

$$\begin{aligned} \varPhi (x)=\varPhi _\lambda (x), \end{aligned}$$

where \(\lambda \in \varLambda \) is such that \(x\in {\mathscr {A}}_\lambda \). We will show that \(\varPhi \) is well-defined. Suppose \(\lambda ,\mu \in \varLambda \) are such that \(x\in {\mathscr {A}}_\lambda \cap {\mathscr {A}}_\mu \). Then there exists \(\nu \in \varLambda \) with \(\lambda ,\mu \le \nu \). Since the net \((e_\lambda )_{\lambda \in \varLambda }\) is increasing, we see that \(e_\lambda ,e_\mu \le e_\nu \) and therefore \({\mathscr {A}}_\lambda ,{\mathscr {A}}_\mu \subset {\mathscr {A}}_\nu \). The uniqueness of the representation of P on both \({\mathscr {A}}_\lambda \) and \({\mathscr {A}}_\mu \) implies that \(\varPhi _\nu \mid _{{\mathscr {A}}_\lambda }=\varPhi _\lambda \) and \(\varPhi _\nu \mid _{{\mathscr {A}}_\mu }=\varPhi _\mu \), which implies that \(\varPhi _\lambda (x)=\varPhi _\nu (x)=\varPhi _\mu (x)\). We now show that \({\mathscr {R}}\) is a \(*\)-subalgebra of \({\mathscr {A}}\) and that \(\varPhi \) is linear. Take \(x,y\in {\mathscr {R}}\) and \(\alpha ,\beta \in {\mathbb {C}}\). We take \(\lambda ,\mu \in \varLambda \) such that \(x\in {\mathscr {A}}_{\lambda }\) and \(y\in {\mathscr {A}}_\mu \). Then \(x^*\in {\mathscr {A}}_\lambda \subset {\mathscr {R}}\). Now set \(\nu \in \varLambda \) with \(\lambda ,\mu \le \nu \). Hence \(x,y\in {\mathscr {A}}_\nu \), so that \(\alpha x+\beta y,xy\in {\mathscr {A}}_\nu \subset {\mathscr {R}}\), which shows that \({\mathscr {R}}\) is a subalgebra of \({\mathscr {A}}\). Further, we have

$$\begin{aligned} \varPhi (\alpha x+\beta y)=\varPhi _\nu (\alpha x+\beta y)= \alpha \varPhi _\nu (x)+\beta \varPhi _\nu (y)=\alpha \varPhi (x)+\beta \varPhi (y), \end{aligned}$$

which shows that \(\varPhi \) is linear.

From (6) we deduce that \(P(x)=\varPhi (x^m)\) for each \(x\in {\mathscr {R}}\).

Our next goal is to show that \({\mathscr {R}}\) satisfies the conditions of Lemma 2. If \(x\in {\mathscr {R}}_{\mathrm{sa}}\) (\(x\in {\mathscr {R}}_+\)), then there exists \(\lambda \in \varLambda \) such that \(x\in \bigl ({\mathscr {A}}_\lambda \bigr )_{\mathrm{sa}}\) (\(x\in \bigl ({\mathscr {A}}_\lambda \bigr )_{+}\), respectively) and therefore \(\vert x\vert \in {\mathscr {A}}_\lambda \subset {\mathscr {R}}\) (\(x^{1/m}\in {\mathscr {A}}_\lambda \subset {\mathscr {R}}\), respectively). Since the polynomial \(P\mid _{\mathscr {R}}\) is continuous, Lemma 2 shows that the map \(\varPhi \) is continuous.

Since \((e_\lambda )_{\lambda \in \varLambda }\) is an approximate unit, it follows that \({\mathscr {R}}\) is dense in \({\mathscr {A}}\), and hence that the map \(\varPhi \) extends uniquely to a continuous linear map from \({\mathscr {A}}\) into the completion of X. By abuse of notation we continue to write \(\varPhi \) for this extension. Since both P and \(\varPhi \) are continuous, it may be concluded that \(P(x)=\varPhi (x^m)\) for each \(x\in {\mathscr {A}}\). We next prove that the image of \(\varPhi \) is actually contained in X. Of course, it suffices to show that \(\varPhi \) takes \({\mathscr {A}}_+\) into X. If \(x\in {\mathscr {A}}_+\), then

$$\begin{aligned} \varPhi (x)= \varPhi \bigl (\bigl (x^{1/m}\bigr )^m\bigr )= P\bigl (x^{1/m}\bigr )\in X, \end{aligned}$$

as required. \(\square \)

Since every von Neumann algebra is unital and has real rank zero, Theorem 3 applies in this setting and gives the following.

Corollary 1

Let \({{\mathscr {M}}}\) be a von Neumann algebra, let X be a topological linear space, and let \(P:{{\mathscr {M}}}\rightarrow X\) be a continuous m-homogeneous polynomial. Then the following conditions are equivalent:

  1. (i)

    there exists a continuous linear map \(\varPhi :{{\mathscr {M}}} \rightarrow X\) such that \(P(x)=\varPhi (x^m)\)\((x\in {{\mathscr {M}}})\);

  2. (ii)

    the polynomial P is orthogonally additive on \({{\mathscr {M}}}_{\mathrm{sa}}\);

  3. (iii)

    the polynomial P is orthogonally additive on \({{\mathscr {M}}}_+\).

If the conditions are satisfied, then the map \(\varPhi \) is unique.

Proposition 1

Let H be a Hilbert space with \(\dim H \ge 2\), let X be a topological linear space, and let \(P:{\mathscr {B}}(H) \rightarrow X\) be a continuous m-homogeneous polynomial. Suppose that P is orthogonally additive in \({\mathscr {B}}(H)\). Then \(P=0\).

Proof

For each unitary \(v\in {\mathscr {B}}(H)\), the map \(P_v:{\mathscr {B}}(H)\rightarrow X\) defined by

$$\begin{aligned} P_v(x)=P(vx)\quad (x\in {\mathscr {B}}(H)) \end{aligned}$$

is easily seen to be a continuous m-homogeneous polynomial that is orthogonally additive on \({\mathscr {B}}(H)\). In particular, \(P_v\) is orthogonally additive on \({\mathscr {B}}(H)_{\mathrm{sa}}\), and Corollary 1 then gives a unique continuous linear map \(\varPhi _v:{\mathscr {B}}(H)\rightarrow X\) such that

$$\begin{aligned} P(vx)=\varPhi _v(x^m)\quad (x\in {\mathscr {B}}(H)). \end{aligned}$$

We claim that, if \(e,e'\in {\mathscr {B}}(H)\) are equivalent projections with \(e\perp e'\), then \(P(e)=P(e')=0\). Let \(u\in {\mathscr {B}}(H)\) be a partial isometry such that \(u^*u=e\) and \(uu^*=e'\). Then

$$\begin{aligned} \left\| u^2 \right\| ^4= \left\| (u^2)^*u^2 \right\| ^2= \left\| \left( (u^2)^*u^2\right) ^2\right\| = \left\| u^{*}ee^{\prime } eu \right\| =0, \end{aligned}$$

which gives \(u^2=0\). From this we see that \(u\perp u^*\), and therefore

$$\begin{aligned} P(vu+vu^*)= P_v(u+u^*)=P_v(u)+P_v(u^*)=\varPhi _v\left( u^m\right) +\varPhi _v\left( (u^*)^m\right) =0. \end{aligned}$$
(7)

We now take \(\omega \in {\mathbb {C}}\) with \(\omega ^m=-1\), and define

$$\begin{aligned} {\begin{matrix} v &{} = 1+u+u^*-e-e',\\ v_\omega &{} = 1+\omega u+u^*-e-e'. \end{matrix}} \end{aligned}$$

It is immediately seen that both v and \(v_\omega \) are unitary, and so applying (7) (and using the orthogonal additivity of P and that \(e\perp e'\)), we see that

$$\begin{aligned} 0&=P(vu+vu^*)=P(e+e')=P(e)+P(e'),\\ 0&=P(v_\omega u+v_\omega u^*)=P(e+\omega e')=P(e)+P(\omega e')=P(e)-P(e'). \end{aligned}$$

By comparing both identities, we conclude that \(P(e)=P(e')=0\), as claimed.

Our next objective is to prove that \(P(e)=0\) for each projection \(e\in {\mathscr {B}}(H)\). Suppose that \(e\in {\mathscr {B}}(H)\) is a rank-one projection. Since \(\dim H\ge 2\), it follows that there exists an equivalent projection \(e'\) such that \(e'\perp e\). Then it follows from the above claim that \(P(e)=0\). Let \(e\in {\mathscr {B}}(H)\) be a finite projection. Then there exist mutually orthogonal projections \(e_1,\ldots ,e_n\) such that \(e_1+\cdots +e_n=e\). Using the preceding observation and the orthogonal additivity of P we get \(P(e)=P(e_1)+\cdots +P(e_n)=0\). We now assume that \(e\in {\mathscr {B}}(H)\) is an infinite projection. Then there exist mutually orthogonal, equivalent projections \(e_1\) and \(e_2\) such that \(e_1+e_2=e\). By the claim, we have \(P(e)=P(e_1)+P(e_2)=0\).

We finally proceed to show that \(P=0\). By Lemma 1, it suffices to show that \(P(x)=0\) for each \(x\in {\mathscr {B}}(H)_+\). Suppose that \(x\in {\mathscr {B}}(H)_+\) can be written in the form \(x=\sum _{j=1}^k\rho _j e_j\), where \(e_1,\ldots ,e_k\in {\mathscr {B}}(H)\) are mutually orthogonal projections and \(\rho _1,\ldots ,\rho _k\in {\mathbb {R}}^+\). Then we have \(P(x)=\sum _{j=1}^k{\rho _j}^mP(e_j)=0\). Now let \(x\in {\mathscr {B}}(H)_{+}\) be an arbitrary element. From the spectral decomposition we deduce that there exists a sequence \((x_n)\) in \({\mathscr {B}}(H)_{+}\) such that each \(x_n\) is a positive linear combination of mutually orthogonal projections and \(\lim x_n=x\). On account of the preceding observation, \(P(x_n)=0\) (\(n\in {\mathbb {N}}\)), and the continuity of P implies that \(P(x)=\lim P(x_n)=0\), as required. \(\square \)

3 Non-commutative \(L^p\)-spaces

Before giving the next results we make the following preliminary remarks.

A fundamental fact for us is the behaviour of the product of \(L^0({{\mathscr {M}}},\tau )\) when restricted to the \(L^p\)-spaces. Specifically, if \(0<p,q,r\le \infty \) are such that \(\tfrac{1}{p}+\tfrac{1}{q}=\tfrac{1}{r}\), then the Hölder inequality states that

$$\begin{aligned} x\in L^p({{\mathscr {M}}},\tau ), \ y\in L^q({{\mathscr {M}}},\tau ) \ \Rightarrow \ xy\in L^r({{\mathscr {M}}},\tau ) \text { and } \Vert xy\Vert _r\le \Vert x\Vert _p\Vert y\Vert _q. \end{aligned}$$
(8)

Suppose that \(x,y\in L^p({{\mathscr {M}}},\tau )_+\), \(0<p<\infty \), are mutually orthogonal and that \(\omega \in {\mathbb {C}}\) with \(\vert \omega \vert =1\). Then it is immediately seen that \(\vert x+\omega y\vert =x+y\), and it follows, by considering the spectral resolutions of x, y, and \(x+y\), that \((x+y)^p=x^p+y^p\). Hence

$$\begin{aligned} \Vert x+\omega y\Vert _p^p=\Vert x\Vert _p^p+\Vert y\Vert _p^p. \end{aligned}$$
(9)

Each \(x\in L^{p}({{\mathscr {M}}},\tau )\) can be written in the form

$$\begin{aligned} \begin{aligned} x=x_1-x_2+i(x_3-x_4), \text { with }&x_1,x_2,x_3,x_4\in L^{p}({{\mathscr {M}}},\tau )_+, \\&x_1\perp x_2, \ x_3\perp x_4,\\&\Vert x_1\Vert _p^p+\Vert x_2\Vert _p^p=\Vert x_1-x_2\Vert _p^p\le \Vert x\Vert _p^p,\\&\Vert x_3\Vert _p^p+\Vert x_4\Vert _p^p=\Vert x_3-x_4\Vert _p^p\le \Vert x\Vert _p^p. \end{aligned} \end{aligned}$$
(10)

Indeed, first we write \(x=\mathfrak {R}x+i\mathfrak {I}x\), where

$$\begin{aligned} \mathfrak {R}x=\frac{1}{2}(x^*+x), \, \mathfrak {I}x=\frac{i}{2}(x^*-x)\in L^{p}({{\mathscr {M}}},\tau )_{\mathrm{sa}}, \end{aligned}$$

and, since \(\Vert x^*\Vert _{p}=\Vert x\Vert _{p}\), it follows that \(\Vert \mathfrak {R}x\Vert _{p},\Vert \mathfrak {I}x\Vert _{p}\le \Vert x\Vert _{p}\). Further, we take the positive operators

$$\begin{aligned} x_1 \!=\!\tfrac{1}{2}\left( \vert \mathfrak {R}x\vert +\mathfrak {R}x \right) ,\ x_2 \!=\!\tfrac{1}{2} \left( \vert \mathfrak {R}x\vert -\mathfrak {R}x \right) ,\ x_3 \!=\!\tfrac{1}{2} \left( \vert \mathfrak {I}x\vert +\mathfrak {I}x \right) ,\ x_4 \!=\!\tfrac{1}{2} \left( \vert \mathfrak {I}x\vert -\mathfrak {I}x \right) . \end{aligned}$$

Then \(x_1,x_2,x_3,x_4\in L^p({{\mathscr {M}}},\tau )\), \(\mathfrak {R}x=x_1-x_2\) with \(x_1\perp x_2\), so that (9) gives

$$\begin{aligned} \Vert \mathfrak {R}x\Vert _p^p=\Vert x_1\Vert _p^p+\Vert x_2\Vert _p^p, \end{aligned}$$

and \(\mathfrak {I}x= x_3-x_4\) with \(x_3\perp x_4\), so that (9) gives

$$\begin{aligned} \Vert \mathfrak {I}x\Vert _p^p=\Vert x_3\Vert _p^p+\Vert x_4\Vert _p^p. \end{aligned}$$

Theorem 4

Let \({{\mathscr {M}}}\) be a von Neumann algebra with a normal semifinite faithful trace \(\tau \), let X be a topological linear space, and let \(\varPhi :L^{p/m}({{\mathscr {M}}},\tau )\rightarrow X\) be a continuous linear map with \(0<p<\infty \). Then:

  1. (i)

    the map \(P_\varPhi :L^p({{\mathscr {M}}},\tau )\rightarrow X\) defined by \(P_\varPhi (x)=\varPhi (x^m)\)\((x\in L^p({{\mathscr {M}}},\tau ))\) is a continuous m-homogeneous polynomial which is orthogonally additive on \(L^p({{\mathscr {M}}},\tau )_{\mathrm{sa}}\);

  2. (ii)

    the polynomial \(P_\varPhi \) is uniquely specified by the map \(\varPhi \).

Suppose, further, that X is a q-normed space, \(0<q\le 1\). Then:

  1. (iii)

    \(2^{-1/q}\Vert \varPhi \Vert \le \Vert P_\varPhi \Vert \le \Vert \varPhi \Vert \).

Moreover, in the case where \(X={\mathbb {C}}\),

  1. (iv)

    the functional \(\varPhi \) is hermitian if and only if the polynomial \(P_\varPhi \) is hermitian, in which case \(\Vert P_\varPhi \Vert =\Vert \varPhi \Vert \).

Proof

The proof of this result is similar to that establishing Theorem 1.

(i) It follows immediately from (8) that, for each \(x_1,\ldots ,x_m\in L^p({{\mathscr {M}}},\tau )\),

$$\begin{aligned} x_1\cdots x_m\in L^{p/m}({{\mathscr {M}}},\tau ) \text { and } \Vert x_1\cdots x_m\Vert _{p/m}\le \Vert x_1\Vert _p\cdots \Vert x_m\Vert _p. \end{aligned}$$
(11)

On the one hand, this clearly implies that the map \(P_\varPhi \) is well-defined, on the other hand, the map \(x\mapsto x^m\) from \(L^p({{\mathscr {M}}},\tau )\) into \(L^{p/m}({{\mathscr {M}}},\tau )\) is continuous, and so \(P_\varPhi \) is continuous. Further, \(P_\varPhi \) is the m-homogeneous polynomial associated with the symmetric m-linear map \(\varphi :L^p({{\mathscr {M}}},\tau )^m\rightarrow X\) defined by

$$\begin{aligned} \varphi (x_1,\ldots ,x_m)= \frac{1}{m!}\sum _{\sigma \in \mathfrak {S}_m} \varPhi \left( x_{\sigma (1)} \cdots x_{\sigma (m)} \right) \quad (x_1,\ldots ,x_m\in L^p({{\mathscr {M}}},\tau )). \end{aligned}$$

Suppose that \(x,y\in L^p({{\mathscr {M}}},\tau )_{\mathrm{sa}}\) are such that \(x\perp y\). Then \(xy=yx=0\), and so \((x+y)^m=x^m+y^m\), which gives

$$\begin{aligned} P_\varPhi (x+y)=\varPhi \bigl ((x+y)^m\bigr )= \varPhi \bigl (x^m+y^m\bigr )=\varPhi \bigl (x^m\bigr )+\varPhi \bigl (y^m\bigr )= P_\varPhi (x)+P_\varPhi (y). \end{aligned}$$

(ii) Suppose that \(\Psi :L^{p/m}({{\mathscr {M}}},\tau )\rightarrow X\) is a linear map such that \(P_\Psi =P_\varPhi \). For each \(x\in L^{p/m}({{\mathscr {M}}},\tau )_+\), we have \(x^{1/m}\in L^{p}({{\mathscr {M}}},\tau )\) and

$$\begin{aligned} \varPhi (x)= \varPhi \bigl (\bigl (x^{1/m}\bigr )^m\bigr )= P\bigl (x^{1/m}\bigr )= \Psi \bigl (\bigl (x^{1/m}\bigr )^m\bigr )= \Psi (x). \end{aligned}$$

By linearity we obtain \(\varPhi =\Psi \).

(iii) Next, assume that X is a q-normed space. For each \(x\in L^p({{\mathscr {M}}},\tau )\), by (11), we have

$$\begin{aligned} \Vert P_\varPhi (x)\Vert =\Vert \varPhi (x^m)\Vert \le \Vert \varPhi \Vert \Vert x^m\Vert _{p/m}\le \Vert \varPhi \Vert \Vert x\Vert _p^m, \end{aligned}$$

which clearly implies that \(\Vert P_\varPhi \Vert \le \Vert \varPhi \Vert \). Now take \(x\in L^{p/m}({{\mathscr {M}}},\tau )\), and take \(\omega \in {\mathbb {C}}\) with \(\omega ^m=-1\). Write

$$\begin{aligned} x= \mathfrak {R}x+i\mathfrak {I}x= x_1-x_2+i(x_3-x_4) \end{aligned}$$

as in (10) (with p / m instead of p). Since \(x_1\perp x_2\) and \(x_3\perp x_4\), it follows that \(x_1^{1/m}\perp x_2^{1/m}\) and \(x_3^{1/m}\perp x_4^{1/m}\), so that (9) gives

$$\begin{aligned} {\begin{matrix} \Vert \mathfrak {R}x\Vert _{p/m}^{p/m} &{}= \Vert x_1\Vert _{p/m}^{p/m}+\Vert x_2\Vert _{p/m}^{p/m},\\ \Vert \mathfrak {I}x\Vert _{p/m}^{p/m} &{}= \Vert x_3\Vert _{p/m}^{p/m}+\Vert x_4\Vert _{p/m}^{p/m}, \end{matrix}} \end{aligned}$$
(12)

and

$$\begin{aligned} {\begin{matrix} {\big \Vert } x_1^{1/m}+\omega x_2^{1/m} {\big \Vert }_p^p &{} = {\big \Vert } x_1^{1/m}{\big \Vert }_p^p+{\big \Vert }x_2^{1/m} {\big \Vert }_p^p,\\ {\big \Vert } x_3^{1/m}+\omega x_4^{1/m} {\big \Vert }_p^p &{} = {\big \Vert } x_3^{1/m}{\big \Vert }_p^p+{\big \Vert } x_4^{1/m} {\big \Vert }_p^p. \end{matrix}} \end{aligned}$$
(13)

Further, we have \(x_1^{1/m},x_2^{1/m},x_3^{1/m},x_4^{1/m}\in L^p({{\mathscr {M}}},\tau )\) and

$$\begin{aligned} \bigl \Vert x_1^{1/m}\big \Vert _p\!= \!\Vert x_1\Vert _{p/m}^{1/m}, \ \big \Vert x_2^{1/m}\big \Vert _p\!\!=\! \Vert x_2\Vert _{p/m}^{1/m}, \ \big \Vert x_3^{1/m}\big \Vert _p\!\!=\! \Vert x_3\Vert _{p/m}^{1/m}, \ \big \Vert x_4^{1/m}\big \Vert _p\!\!=\! \Vert x_4\Vert _{p/m}^{1/m}, \end{aligned}$$

so that (12) and (13) give

$$\begin{aligned} {\begin{matrix} {\big \Vert } x_1^{1/m}+\omega x_2^{1/m}{\big \Vert }_p^p &{}= \Vert {\mathfrak {R}} x\Vert _{p/m}^{p/m},\\ {\big \Vert } x_3^{1/m}+\omega x_4^{1/m}{\big \Vert }_p^p &{}= \Vert {\mathfrak {I}} x\Vert _{p/m}^{p/m}. \end{matrix}} \end{aligned}$$
(14)

On the other hand, we have

$$\begin{aligned} \left( x_1^{1/m}+\omega x_2^{1/m}\right) ^m=x_1-x_2=\mathfrak {R}x,\quad \left( x_3^{1/m}+\omega x_4^{1/m}\right) ^m=x_3-x_4=\mathfrak {I}x, \end{aligned}$$

whence

$$\begin{aligned} \varPhi (x)&= \varPhi (\mathfrak {R}x)+i\varPhi (\mathfrak {I}x)= \varPhi \bigl (\bigl (x_1^{1/m}+\omega x_2^{1/m}\bigr )^m\bigr )+ i\varPhi \bigl (\bigl (x_3^{1/m}+\omega x_4^{1/m}\bigr )^m\bigr ) \\&= P_\varPhi \bigl (x_1^{1/m}+\omega x_2^{1/m}\bigr )+ iP_\varPhi \bigl (x_3^{1/m}+\omega x_4^{1/m}\bigr ). \end{aligned}$$

Hence, by (14),

$$\begin{aligned} \Vert \varPhi (x)\Vert ^q&\le \big \Vert P_\varPhi \bigl (x_1^{1/m}+\omega x_2^{1/m}\bigr )\big \Vert ^q+ \big \Vert P_\varPhi \bigl (x_3^{1/m}+\omega x_4^{1/m}\bigr )\big \Vert ^q\\&\le \Vert P_\varPhi \Vert ^q\big \Vert x_1^{1/m}+\omega x_2^{1/m}\big \Vert _p^{mq}+ \Vert P_\varPhi \Vert ^q\big \Vert x_3^{1/m}+\omega x_4^{1/m}\big \Vert _p^{mq}\\&= \Vert P_\varPhi \Vert ^q\left( \Vert \mathfrak {R}x\Vert ^q+\Vert \mathfrak {I}x\Vert ^q\right) \\&\le \Vert P_\varPhi \Vert ^q 2\Vert x\Vert ^q. \end{aligned}$$

This clearly forces \(\Vert \varPhi \Vert \le 2^{1/q}\Vert P_\varPhi \Vert \), as claimed.

(iv) It is straightforward to check that \(P_\varPhi ^*=P_{\varPhi ^*}\). From this deduce that \(\varPhi \) is hermitian if and only if \(P_\varPhi \) is hermitian as in the proof of Theorem 1(iv). Suppose that \(\varPhi \) is a hermitian functional. By direct calculation, we see that \(P_\varPhi \) is hermitian, and it remains to prove that \(\Vert P_\varPhi \Vert =\Vert \varPhi \Vert \). We only need to show that \(\Vert \varPhi \Vert \le \Vert P_\varPhi \Vert \). To this end, let \(\varepsilon \in {\mathbb {R}}^+\), and choose \(x\in L^{p/m}({{\mathscr {M}}},\tau )\) such that \(\Vert x\Vert _{p/m}=1\) and \(\Vert \varPhi \Vert -\varepsilon <\vert \varPhi (x)\vert \). We take \(\alpha \in {\mathbb {C}}\) with \(\vert \alpha \vert =1\) and \(\vert \varPhi (x)\vert =\alpha \varPhi (x)\), so that

$$\begin{aligned} \Vert \varPhi \Vert -\varepsilon <\vert \varPhi (x)\vert =\varPhi (\alpha x)= \overline{\varPhi (\alpha x)}=\varPhi \bigl ((\alpha x)^*\bigr ). \end{aligned}$$

We see that \(\mathfrak {R}(\alpha x)\in L^{p/m}({{\mathscr {M}}},\tau )_{\mathrm{sa}}\), \(\Vert \mathfrak {R}(\alpha x)\Vert _{p/m}\le 1\), and \(\Vert \varPhi \Vert -\varepsilon <\varPhi (\mathfrak {R}(\alpha x))\). Now we consider the decomposition \(\mathfrak {R}(\alpha x)=x_1-x_2\) as in (10) (with p / m instead of p), and take \(\omega \in {\mathbb {C}}\) with \(\omega ^m=-1\). As in (14), we see that \(\big \Vert x_1^{1/m}+\omega x_2^{1/m}\big \Vert =\Vert \mathfrak {R}(\alpha x)\Vert ^{1/m}\le 1\). Moreover, we have

$$\begin{aligned} P_\varPhi \bigl ( x_1^{1/m}+\omega x_2^{1/m}\bigr )= \varPhi \bigl (\bigl ( x_1^{1/m}+\omega x_2^{1/m}\bigr )^m\bigr )= \varPhi (\mathfrak {R}(\alpha x)), \end{aligned}$$

and so \(\Vert \varPhi \Vert -\varepsilon <\Vert P_\varPhi \Vert \). \(\square \)

Theorem 5

Let \({{\mathscr {M}}}\) be a von Neumann algebra with a normal semifinite faithful trace \(\tau \), let X be a topological linear space, and let \(P:L^p({{\mathscr {M}}},\tau )\rightarrow X\) be a continuous m-homogeneous polynomial with \(0<p<\infty \). Then the following conditions are equivalent:

  1. (i)

    there exists a continuous linear map \(\varPhi :L^{p/m}({{\mathscr {M}}},\tau )\rightarrow X\) such that \(P(x)=\varPhi (x^m)\)\((x\in L^p({{\mathscr {M}}},\tau ))\);

  2. (ii)

    the polynomial P is orthogonally additive on \(L^p({{\mathscr {M}}},\tau )_{\mathrm{sa}}\);

  3. (iii)

    the polynomial P is orthogonally additive on \(S({{\mathscr {M}}},\tau )_+\).

If the conditions are satisfied, then the map \(\varPhi \) is unique.

Proof

Theorem 4 shows that (i)\(\Rightarrow \)(ii), and it is obvious that (ii)\(\Rightarrow \)(iii). We proceed to prove that (iii)\(\Rightarrow \)(i).

Suppose that (iii) holds. Let \(e\in {\mathscr {M}}\) be a projection such that \(\tau (e)<\infty \), and consider the von Neumann algebra \({{\mathscr {M}}}_e=e{{\mathscr {M}}}e\). We claim that \({\mathscr {M}}_e\subset S({{\mathscr {M}}},\tau )\) and that there exists a unique continuous linear map \(\varPhi _e:{\mathscr {M}}_e\rightarrow X\) such that

$$\begin{aligned} P(x)=\varPhi _e(x^m)\quad (x\in {\mathscr {M}}_e). \end{aligned}$$
(15)

Set \(x\in {\mathscr {M}}_e\), and write \(x=(x_1-x_2)+i(x_3-x_4)\) with \(x_1,x_2,x_3,x_4\in {\mathscr {M}}_{e \, +}\). Then \({{\,\mathrm{supp}\,}}(x_j)\le e\) and therefore \(\tau \bigl ({{\,\mathrm{supp}\,}}(x_j)\bigr )\le \tau (e)<\infty \)\((j\in \{1,2,3,4\})\). This shows that \(x_j\in S({\mathscr {M}},\tau )\)\((j\in \{1,2,3,4\})\), whence \(x\in S({\mathscr {M}},\tau )\). Our next goal is to show that the restriction \(P\mid _{{\mathscr {M}}_e}\) is continuous (with respect to the norm that \({\mathscr {M}}_e\) inherits as a closed subspace of \({\mathscr {M}}\)). Let \(x\in {\mathscr {M}}_e\), and let \(U\subset X\) be a neighbourhood of P(x). Since P is continuous, the set \(P^{-1}(U)\) is a neighbourhood of x in \(L^p({\mathscr {M}},\tau )\), which implies that there exists \(r\in {\mathbb {R}}^+\) such that \(P(y)\in U\) whenever \(y\in L^p({\mathscr {M}},\tau )\) and \(\Vert y- x\Vert _p<r\). If \(y\in {\mathscr {M}}_e\) is such that \(\Vert y- x\Vert <r/\Vert e\Vert _p\), then, from (8), we obtain

$$\begin{aligned} \Vert y-x\Vert _p=\Vert e(y-x)\Vert _p\le \Vert e\Vert _p\Vert y- x\Vert <r \end{aligned}$$

and therefore \(P(y)\in U\). Hence \(P\mid _{{\mathscr {M}}_e}\) is continuous. Since, by hypothesis, the polynomial \(P\mid _{{\mathscr {M}}_e}\) is orthogonally additive on \({\mathscr {M}}_{e \, +}\), Corollary 1 states that there exists a unique continuous linear map \(\varPhi _e:{\mathscr {M}}_e\rightarrow X\) such that (15) holds.

For each \(x\in S({\mathscr {M}},\tau )\), define

$$\begin{aligned} \varPhi (x)=\varPhi _e(x), \end{aligned}$$

where \(e\in {\mathscr {M}}\) is any projection such that

$$\begin{aligned} ex=xe=x\quad \text {and}\quad \tau (e)<\infty . \end{aligned}$$
(16)

We will show that \(\varPhi \) is well-defined. For this purpose we first check that, if \(x\in S({\mathscr {M}},\tau )\), then there exists a projection e such that (16) holds. Indeed, we write \(x=\sum _{j=1}^k\alpha _j x_j\) with \(\alpha _1,\ldots ,\alpha _k\in {\mathbb {C}}\) and \(x_1,\ldots ,x_k\in S({\mathscr {M}},\tau )_+\), and define \(e={{\,\mathrm{supp}\,}}(x_1)\vee \cdots \vee {{\,\mathrm{supp}\,}}(x_k)\). Then \(ex=xe=x\) and \(\tau (e)\le \sum _{j=1}^k\tau \bigl ({{\,\mathrm{supp}\,}}(x_j)\bigr )<\infty \), as required. Suppose that \(x\in S({\mathscr {M}},\tau )\) and that \(e_1,e_2\in {\mathscr {M}}\) are projections satisfying (16). Then the projection \(e=e_1\vee e_2\) satisfies (16) and \({\mathscr {M}}_{e_1},{\mathscr {M}}_{e_2}\subset {\mathscr {M}}_e\). The uniqueness of the representation (15) on both \({\mathscr {M}}_{e_1}\) and \({\mathscr {M}}_{e_2}\) gives \(\varPhi _{e}\mid _{{\mathscr {M}}_{e_1}}=\varPhi _{e_1}\) and \(\varPhi _{e}\mid _{{\mathscr {M}}_{e_2}}=\varPhi _{e_2}\), which implies that \(\varPhi _{e_1}(x)=\varPhi _{e}(x)=\varPhi _{e_2}(x)\).

We now show that \(\varPhi \) is linear. Take \(x_1,x_2\in S({\mathscr {M}},\tau )\) and \(\alpha ,\beta \in {\mathbb {C}}\). Let \(e_1,e_2\in {\mathscr {M}}\) be projections such that \(e_jx_j=x_je_j=x_j\) and \(\tau (e_j)<\infty \)\((j\in \{1,2\})\). Then the projection \(e=e_1\vee e_2\) satisfies

$$\begin{aligned}\begin{gathered} ex_j=x_je=x_j\quad (j\in \{1,2\}), \\ e(\alpha x_1+\beta x_2)=(\alpha x_1+\beta x_2)e=\alpha x_1+\beta x_2, \end{gathered}\end{aligned}$$

and

$$\begin{aligned} \tau (e)\le \tau (e_1)+\tau (e_2)<\infty . \end{aligned}$$

Thus

$$\begin{aligned} \varPhi (x_j)=\varPhi _e(x_j)\quad (j\in \{1,2\}) \end{aligned}$$

and

$$\begin{aligned} \varPhi (\alpha x_1+\beta x_2)= \varPhi _e(\alpha x_1+\beta x_2)= \alpha \varPhi _e(x_1)+\beta \varPhi _e(x_2)= \alpha \varPhi (x_1)+\beta \varPhi (x_2). \end{aligned}$$

We see from the definition of \(\varPhi \) that

$$\begin{aligned} P(x)=\varPhi (x^m) \quad (x\in S({\mathscr {M}},\tau )). \end{aligned}$$
(17)

Our next concern will be the continuity of \(\varPhi \) with respect to the norm \(\Vert \cdot \Vert _{p/m}\). Let U be a neighbourhood of 0 in X. Let V be a balanced neighbourhood of 0 in X with \(V+V+V+V\subset U\). The set \(P^{-1}(V)\) is a neighbourhood of 0 in \(L^p({\mathscr {M}},\tau )\), which implies that there exists \(r\in {\mathbb {R}}^+\) such that \(P(x)\in V\) whenever \(x\in L^p({\mathscr {M}},\tau )\) and \(\Vert x\Vert _p<r\). Take \(x\in S({\mathscr {M}},\tau )\) with \(\Vert x\Vert _{p/m}<r^m\), and write \(x=(x_1-x_2)+i(x_3-x_4)\) as in (10) (with p / m instead of p). Then it is immediate to check that actually \(x_1,x_2,x_3,x_4\in S({\mathscr {M}},\tau )_+\) and, further, \(\Vert x_j\Vert _{p/m}\le \Vert x\Vert _{p/m}\)\((j\in \{1,2,3,4\})\). For each \(j\in \{1,2,3,4\}\), we have

$$\begin{aligned} \big \Vert x_{j}^{1/m}\big \Vert _p&= \tau \bigl (x_{j}^{p/m}\bigr )^{1/p}= \bigl (\tau \bigl (x_{j}^{p/m}\bigr )^{m/p}\bigr )^{1/m}= {\Vert x_j\Vert _{p/m}}^{1/m} \\&\le \Vert x\Vert _{p/m}^{1/m}<r, \end{aligned}$$

whence

$$\begin{aligned} \varPhi (x)&= \varPhi \left( \bigl (x_{1}^{1/m}\bigr )^m-\bigl (x_{2}^{1/m}\bigr )^m+i\bigl (x_{3}^{1/m}\bigr )^m-i\bigl (x_{4}^{1/m}\bigr )^m\right) \\&= \varPhi \left( \bigl (x_{1}^{1/m}\bigr )^m\right) - \varPhi \left( \bigl (x_{2}^{1/m}\bigr )^m\right) + i\varPhi \left( \bigl (x_{3}^{1/m}\bigr )^m\right) - i\varPhi \left( \bigl (x_{4}^{1/m}\bigr )^m\right) \\&= P\bigl (x_{1}^{1/m}\bigr )-P\bigl (x_{2}^{1/m}\bigr ) \\&\quad {}+iP\bigl (x_{3}^{1/m}\bigr )-iP\bigl (x_{4}^{1/m}\bigr )\in V+V+V+V\subset U, \end{aligned}$$

which establishes the continuity of \(\varPhi \). Since \(S({\mathscr {M}},\tau )\) is dense in \(L^{p/m}({\mathscr {M}},\tau )\), the map \(\varPhi \) extends uniquely to a continuous linear map from \(L^{p/m}({\mathscr {M}},\tau )\) into the completion of X. By abuse of notation we continue to write \(\varPhi \) for this extension. Since both P and \(\varPhi \) are continuous, (17) gives \(P(x)=\varPhi (x^m)\) for each \(x\in L^p({\mathscr {M}})\). The task is now to show that the image of \(\varPhi \) is actually contained in X. Of course, it suffices to show that \(\varPhi \) takes \(L^{p/m}({\mathscr {M}},\tau )_+\) into X. Let \(x\in L^{p/m}({\mathscr {M}},\tau )_+\). Then \(x^{1/m}\in L^{p}({\mathscr {M}},\tau )_+\) and

$$\begin{aligned} \varPhi (x)= \varPhi \bigl (\bigl (x^{1/m}\bigr )^m\bigr )= P\bigl (x^{1/m}\bigr )\in X, \end{aligned}$$

as required.

The uniqueness of the map \(\varPhi \) is given by Theorem 4(ii). \(\square \)

Let us note that the space of all continuous m-homogeneous polynomials from \(L^p({\mathscr {M}},\tau )\) into any topological linear space X which are orthogonally additive on \(S({\mathscr {M}},\tau )_+\) is sufficiently rich in the case where \(p/m\ge 1\), because of the existence of continuous linear functionals on \(L^{p/m}({\mathscr {M}},\tau )\). However, some restriction on the space X must be imposed when we consider the case \(p/m<1\) and the von Neumann algebra \({\mathscr {M}}\) has no minimal projections, because in this case the dual of \(L^{p/m}({\mathscr {M}},\tau )\) is trivial ([15]). In fact, there are no non-zero continuous linear maps from \(L^p({\mathscr {M}},\tau )\) into any q-normed space X with \(q>p\). We think that this property is probably well-known, but we have not been able to find any reference, so that we next present a proof of this result for completeness.

Proposition 2

Let \({\mathscr {M}}\) be a von Neumann algebra with a normal semifinite faithful trace \(\tau \) and with no minimal projections, let X be a q-normed space, \(0<q\le 1\), and let \(\varPhi :L^p({\mathscr {M}},\tau )\rightarrow X\) be a continuous linear map with \(0<p<q\). Then \(\varPhi =0\).

Proof

The proof will be divided in a number of steps.

Our first step is to show that for each projection \(e_0\in {\mathscr {M}}\) with \(\tau (e_0)<\infty \) and each \(0\le \rho \le \tau (e_0)\), there exists a projection \(e\in {\mathscr {M}}\) such that \(e\le e_0\) and \(\tau (e)=\rho \). Set

$$\begin{aligned} {\mathscr {P}}_1=\bigl \{e\in {\mathscr {M}} : e\text { is a projection, } e\le e_0, \ \tau (e)\ge \rho \bigr \}. \end{aligned}$$

Note that \(e_0\in {\mathscr {P}}_1\), so that \({\mathscr {P}}_1\) is non-empty. Let \({\mathscr {C}}\) be a chain in \({\mathscr {P}}_1\), and let \(e'=\wedge _{e\in {\mathscr {C}}}e\). Then \(e'\) is a projection and \(e'\le e_0\). For each \(e\in {\mathscr {C}}\), since \(\tau (e_0)<\infty \), it follows that \(\tau (e_0)-\tau (e)=\tau (e_0-e)\). From the normality of \(\tau \) we now deduce that

$$\begin{aligned} \tau (e_0)-\inf _{e\in {\mathscr {C}}}\tau (e)&= \sup _{e\in {\mathscr {C}}}\bigl (\tau (e_0)-\tau (e)\bigr ) = \sup _{e\in {\mathscr {C}}}\tau (e_0-e) \\&= \tau \left( \vee _{e\in {\mathscr {C}}}(e_0-e)\right) = \tau (e_0-e'). \end{aligned}$$

Hence \(\tau (e')=\inf _{e\in {\mathscr {C}}}\tau (e)\ge \rho \), which shows that \(e'\) is a lower bound of \({\mathscr {C}}\), and so, by Zorn’s lemma, \({\mathscr {P}}_1\) has a minimal element, say \(e_1\). We now consider the set

$$\begin{aligned} {\mathscr {P}}_2=\bigl \{e\in {\mathscr {M}} : e\text { is a projection, } e\le e_1, \ \tau (e)\le \rho \bigr \}. \end{aligned}$$

Note that \(0\in {\mathscr {P}}_2\), so that \({\mathscr {P}}_2\) is non-empty. Let \({\mathscr {C}}\) be a chain in \({\mathscr {P}}_2\), and let \(e'=\vee _{e\in {\mathscr {C}}}e\). Then \(e'\le e_1\), and the normality of \(\tau \) yields

$$\begin{aligned} \tau (e')=\sup _{e\in {\mathscr {C}}}\tau (e)\le \rho . \end{aligned}$$

This implies that \(e'\) is an upper bound of \({\mathscr {C}}\), and so, by Zorn’s lemma, \({\mathscr {P}}_2\) has a maximal element, say \(e_2\). Assume towards a contradiction that \(e_1\ne e_2\). Since, by hypothesis, \({\mathscr {M}}\) has no minimal projections, it follows that there exists a non-zero projection \(e<e_1-e_2\). Since \(e\perp e_2\), we see that \(e_2+e\) is a projection. Further, we have \(e_2<e_2+e<e_1\). The maximality of \(e_2\) implies that \(\tau (e_2+e)>\rho \), which implies that \(e_2+e\in {\mathscr {P}}_1\), contradicting the minimality of \(e_1\). Thus \(e_1=e_2\), and this clearly implies that \(\tau (e_1)=\tau (e_2)=\rho \).

Our next goal is to show that \(\varPhi (e_0)=0\) for each projection \(e_0\) with \(\tau (e_0)<\infty \). From the previous step, it follows that there exists a projection \(e\le e_0\) with \(\tau (e)=\frac{1}{2}\tau (e_0)\). Set \(e'=e_0-e\). Then \(\tau (e')=\frac{1}{2}\tau (e_0)\). Further,

$$\begin{aligned} \Vert \varPhi (e_0)\Vert ^q= \Vert \varPhi (e)+\varPhi (e')\Vert ^q\le \Vert \varPhi (e)\Vert ^q+\Vert \varPhi (e')\Vert ^q, \end{aligned}$$

and therefore either \(\Vert \varPhi (e)\Vert ^q\ge \frac{1}{2}\Vert \varPhi (e_0)\Vert ^q\) or \(\Vert \varPhi (e')\Vert ^q\ge \frac{1}{2}\Vert \varPhi (e_0)\Vert ^q\). We define \(e_1\) to be any of the projections \(e,e'\) for which the inequality holds. We thus get \(e_1\le e_0\), \(\tau (e_1)=\frac{1}{2}\tau (e_0)\), and \(\Vert \varPhi (e_1)\Vert \ge 2^{-1/q}\Vert \varPhi (e_0)\Vert \). By repeating the process, we get a decreasing sequence of projections \((e_n)\) such that

$$\begin{aligned} \tau (e_n)=2^{-n}\tau (e_0)\quad \text {and}\quad \Vert \varPhi (e_n)\Vert \ge 2^{-n/q}\Vert \varPhi (e_0)\Vert \quad (n\in {\mathbb {N}}). \end{aligned}$$

Then

$$\begin{aligned} \big \Vert 2^{n/q}e_n\big \Vert _p=2^{n/q}\tau (e_n)^{1/p}=2^{n(1/q-1/p)}\tau (e_0)^{1/p}, \end{aligned}$$

which converges to zero, because \(p<q\). Since \(\varPhi \) is continuous and \(\Vert \varPhi (e_0)\Vert \le \big \Vert \varPhi \bigl (2^{n/q}e_n\bigr )\big \Vert _p\)\((n\in {\mathbb {N}})\), it may be concluded that \(\varPhi (e_0)=0\), as claimed.

Our next concern is to show that \(\varPhi \) vanishes on \(S({\mathscr {M}},\tau )\). Of course, it suffices to show that \(\varPhi \) vanishes on \(S({\mathscr {M}},\tau )_+\). Take \(x\in S({\mathscr {M}},\tau )_+\), and let \(e={{\,\mathrm{supp}\,}}(x)\), so that \(\tau (e)<\infty \). The spectral decomposition implies that there exists a sequence \((x_n)\) in \({\mathscr {M}}_+\) such that \(\lim x_n=x\) with respect to the operator norm and each \(x_n\) is of the form \(x_n=\sum _{j=1}^k\rho _j e_j\), where \(\rho _1,\ldots ,\rho _k\in {\mathbb {R}}^+\) and \(e_1,\ldots ,e_k\in {\mathscr {M}}\) are mutually orthogonal projections with \(e_je=ee_j=e_j\)\((j\in \{1,\ldots ,k\})\). From the previous step, we conclude that \(\varPhi (x_n)=0\)\((n\in {\mathbb {N}})\). Further, from (8) we deduce that

$$\begin{aligned} \Vert x-x_n\Vert _p=\Vert e(x-x_n)\Vert _p\le \Vert e\Vert _p\Vert x-x_n\Vert \rightarrow 0, \end{aligned}$$

and the continuity of \(\varPhi \) implies that \(\varPhi (x)=0\), as required.

Finally, since \(S({\mathscr {M}},\tau )\) is dense in \(L^p({\mathscr {M}},\tau )\) and \(\varPhi \) is continuous, it may be concluded that \(\varPhi =0\). \(\square \)

Corollary 2

Let \({\mathscr {M}}\) be a von Neumann algebra with a normal semifinite faithful trace \(\tau \) and with no minimal projections, let X be a q-normed space, \(0<q\le 1\), and let \(P:L^p({\mathscr {M}},\tau )\rightarrow X\) be a continuous m-homogeneous polynomial with \(0<p/m<q\). Suppose that P is orthogonally additive on \(S({\mathscr {M}},\tau )_+\). Then \(P=0\).

Proof

This is a straightforward consequence of Theorem 5 and Proposition 2. \(\square \)

We now turn our attention to the complex-valued polynomials. In this setting the representation given in Theorem 5 has a particularly significant integral form, because of the well-known representation of the dual of the \(L^p\)-spaces. The trace gives rise to a distinguished contractive positive linear functional on \(L^1({\mathscr {M}},\tau )\), still denoted by \(\tau \). By (8), if \(\frac{1}{p}+\frac{1}{q}=1\), for each \(\zeta \in L^q({\mathscr {M}},\tau )\), the formula

$$\begin{aligned} \varPhi _\zeta (x)=\tau (\zeta x)\quad (x\in L^p({\mathscr {M}},\tau )) \end{aligned}$$
(18)

defines a continuous linear functional on \(L^p({\mathscr {M}},\tau )\). Further, in the case where \(1\le p<\infty \), the map \(\zeta \mapsto \varPhi _\zeta \) is an isometric isomorphism from \(L^q({\mathscr {M}},\tau )\) onto the dual space of \(L^p({\mathscr {M}},\tau )\). It is immediate to see that \(\varPhi _\zeta ^*=\varPhi _{\zeta ^*}\), so that \(\varPhi _\zeta \) is hermitian if and only if \(\zeta \) is self-adjoint.

Corollary 3

Let \({\mathscr {M}}\) be a von Neumann algebra with a normal semifinite faithful trace \(\tau \), and let \(P:L^p({\mathscr {M}},\tau )\rightarrow {\mathbb {C}}\) be a continuous m-homogeneous polynomial with \(m\le p<\infty \). Then the following conditions are equivalent:

  1. (i)

    there exists \(\zeta \in L^{r}({\mathscr {M}},\tau )\) such that \(P(x)=\tau (\zeta x^m)\)\((x\in L^p({\mathscr {M}},\tau ))\), where \(r=p/(p-m)\) (with the convention that \(p/0=\infty \));

  2. (ii)

    the polynomial P is orthogonally additive on \(L^p({\mathscr {M}},\tau )_{\mathrm{sa}}\);

  3. (iii)

    the polynomial P is orthogonally additive on \(S({\mathscr {M}},\tau )_+\).

If the conditions are satisfied, then \(\zeta \) is unique and \(\Vert P\Vert \le \Vert \zeta \Vert _{r}\le 2\Vert P\Vert \); moreover, if P is hermitian, then \(\zeta \) is self-adjoint and \(\Vert \zeta \Vert _{r}=\Vert P\Vert \).

Proof

This follows from Theorems 4 and 5. \(\square \)

Let H be a Hilbert space. We denote by \({{\,\mathrm{Tr}\,}}\) the usual trace on the von Neumann algebra \({\mathscr {B}}(H)\). Then \(L^p({\mathscr {B}}(H),{{\,\mathrm{Tr}\,}})\), with \(0<p<\infty \), is the Schatten class \(S^p(H)\). In the case where \(0<p<q\), we have \(S^p(H)\subset S^q(H)\subset {\mathscr {K}}(H)\) and \(\Vert x\Vert \le \Vert x\Vert _q\le \Vert x\Vert _p\)\((x\in S^p(H))\). It is clear that \(S({\mathscr {B}}(H),{{\,\mathrm{Tr}\,}})={\mathscr {F}}(H)\), the two-sided ideal of \({\mathscr {B}}(H)\) consisting of the finite-rank operators. Thus, the following result is an immediate consequence of Corollary 3.

Corollary 4

Let H be a Hilbert space, and let \(P:S^p(H)\rightarrow {\mathbb {C}}\) be a continuous m-homogeneous polynomial with \(m< p<\infty \). Then the following conditions are equivalent:

  1. (i)

    there exists \(\zeta \in S^{r}(H)\) such that \(P(x)={{\,\mathrm{Tr}\,}}(\zeta x^m)\)\((x\in S^p(H))\), where \(r=p/(p-m)\);

  2. (ii)

    the polynomial P is orthogonally additive on \(S^p(H)_{\mathrm{sa}}\);

  3. (iii)

    the polynomial P is orthogonally additive on \({\mathscr {F}}(H)_+\).

If the conditions are satisfied, then \(\zeta \) is unique and \(\Vert P\Vert \le \Vert \zeta \Vert _{r}\le 2\Vert P\Vert \); moreover, if P is hermitian, then \(\zeta \) is self-adjoint and \(\Vert \zeta \Vert _{r}=\Vert P\Vert \).

Corollary 5

Let H be a Hilbert space, and let \(P:{\mathscr {K}}(H)\rightarrow {\mathbb {C}}\) be a continuous m-homogeneous polynomial. Then the following conditions are equivalent:

  1. (i)

    there exists \(\zeta \in S^1(H)\) such that \(P(x)={{\,\mathrm{Tr}\,}}(\zeta x^m)\)\((x\in {\mathscr {K}}(H))\);

  2. (ii)

    the polynomial P is orthogonally additive on \({\mathscr {K}}(H)_{\mathrm{sa}}\);

  3. (iii)

    the polynomial P is orthogonally additive on \({\mathscr {F}}(H)_+\).

If the conditions are satisfied, then \(\zeta \) is unique and \(\Vert P\Vert \le \Vert \zeta \Vert _{1}\le 2\Vert P\Vert \); moreover, if P is hermitian, then \(\zeta \) is self-adjoint and \(\Vert \zeta \Vert _{1}=\Vert P\Vert \).

Proof

In order to prove the equivalence of the conditions we are reduced to prove that (iii)\(\Rightarrow \)(i). Suppose that (iii) holds. Let \(x,y\in {\mathscr {K}}(H)_+\) such that \(x\perp y\). From the spectral decomposition of both x and y we deduce that there exist sequences \((x_n)\) and \((y_n)\) in \({\mathscr {F}}(H)_+\) such that \(\lim x_n=x\), \(\lim y_n=y\), and \(x_m\perp y_n\)\((m,n\in {\mathbb {N}})\). Then

$$\begin{aligned} P(x+y)=\lim P(x_n+y_n)=\lim \bigl (P(x_n)+P(y_n)\bigr )=P(x)+P(y). \end{aligned}$$

This shows that P is orthogonally additive on \({\mathscr {K}}(H)_+\). Since the \(C^*\)-algebra \({\mathscr {K}}(H)\) has real rank zero and the net consisting of all finite-rank projections is an increasing approximate unit, Theorem 3 applies and gives a continuous linear functional \(\varPhi \) on \({\mathscr {K}}(H)\) such that \(P(x)=\varPhi (x^m)\)\((x\in {\mathscr {K}}(H))\). It is well-known that the map \(\zeta \mapsto \varPhi _\zeta \), as defined in (18), gives an isometric isomorphism from \(S^1(H)\) onto the dual of \({\mathscr {K}}(H)\), so that there exists \(\zeta \in S^1(H)\) such that \(\varPhi (x)={{\,\mathrm{Tr}\,}}(\zeta x)\)\((x\in {\mathscr {K}}(H))\) and \(\Vert \zeta \Vert _1=\Vert \varPhi \Vert \). Thus we obtain (i). The additional properties of the result follow from Theorem 1. \(\square \)

Corollary 6

Let H be a Hilbert space, and let \(P:S^p(H)\rightarrow {\mathbb {C}}\) be a continuous m-homogeneous polynomial with \(0<p\le m\). Then the following conditions are equivalent:

  1. (i)

    there exists \(\zeta \in {\mathscr {B}}(H)\) such that \(P(x)={{\,\mathrm{Tr}\,}}(\zeta x^m)\)\((x\in S^p(H))\);

  2. (ii)

    the polynomial P is orthogonally additive on \(S^p(H)_{\mathrm{sa}}\);

  3. (iii)

    the polynomial P is orthogonally additive on \({\mathscr {F}}(H)_+\).

If the conditions are satisfied, then \(\zeta \) is unique and \(\Vert P\Vert \le \Vert \zeta \Vert \le 2\Vert P\Vert \); moreover, if P is hermitian, then \(\zeta \) is self-adjoint and \(\Vert \zeta \Vert =\Vert P\Vert \).

Proof

By Theorems 4 and 5, it suffices to show that the map \(\zeta \mapsto \varPhi _\zeta \), as defined in (18), gives isometric isomorphism from \({\mathscr {B}}(H)\) onto the dual of \(S^{p/m}(H)\). This is probably well-known, but we are not aware of any reference. Consequently, it may be helpful to include a proof of this fact. If \(\zeta \in {\mathscr {B}}(H)\) and \(x\in S^{p/m}(H)\), then, by (8), \(\zeta x\in S^{p/m}(H)\), so that \(\zeta x\in S^1(H)\) and

$$\begin{aligned} \big \Vert {{\,\mathrm{Tr}\,}}(\zeta x)\big \Vert \le \Vert \zeta x\Vert _1\le \Vert \zeta \Vert \Vert x\Vert _{1}\le \Vert \zeta \Vert \Vert x\Vert _{p/m}, \end{aligned}$$

which shows that \(\varPhi _\zeta \) is a continuous linear functional on \(S^{p/m}(H)\) with \(\Vert \varPhi _\zeta \Vert \le \Vert \zeta \Vert \). Conversely, assume that \(\varPhi \) is a continuous linear functional on \(S^{p/m}(H)\). For each \(\xi ,\eta \in H\), let \(\xi \otimes \eta \in {\mathscr {F}}(H)\) defined by

$$\begin{aligned} \bigl (\xi \otimes \eta \bigr )(\psi )=\langle \psi \vert \eta \rangle \xi \quad (\psi \in H), \end{aligned}$$

and define \(\varphi :H\times H\rightarrow {\mathbb {C}}\) by

$$\begin{aligned} \varphi (\xi ,\eta )=\varPhi (\xi \otimes \eta )\quad (\xi ,\eta \in H). \end{aligned}$$

It is easily checked that \(\varphi \) is a continuous sesquilinear functional with \(\Vert \varphi \Vert \le \Vert \varPhi \Vert \). Therefore there exists \(\zeta \in {\mathscr {B}}(H)\) such that \(\langle \zeta (\xi )\vert \eta \rangle =\varphi (\xi ,\eta )\) for all \(\xi ,\eta \in H\) and \(\Vert \zeta \Vert \le \Vert \varPhi \Vert \). The former condition implies that

$$\begin{aligned} \varPhi _\zeta (\xi \otimes \eta )= {{\,\mathrm{Tr}\,}}(\zeta \xi \otimes \eta )= \langle \zeta (\xi )\vert \eta \rangle = \varphi (\xi ,\eta )= \varPhi (\xi \otimes \eta ) \end{aligned}$$

for all \(\xi ,\eta \in H\), which gives \(\varPhi _\zeta (x)=\varPhi (x)\) for each \(x\in {\mathscr {F}}(H)\). Since \({\mathscr {F}}(H)\) is dense in \(S^{p/m}(H)\), it follows that \(\varPhi _\zeta =\varPhi \). Further, we have \(\Vert \zeta \Vert \le \Vert \varPhi \Vert =\Vert \varPhi _\zeta \Vert \le \Vert \zeta \Vert \). Finally, it is immediate to see that \(\varPhi _\zeta ^*=\varPhi _{\zeta ^*}\), so that \(\varPhi _\zeta \) is hermitian if and only if \(\zeta \) is self-adjoint. \(\square \)

Proposition 3

Let H be a Hilbert space with \(\dim H \ge 2\), let X be a topological linear space, and let \(P:S^p(H) \rightarrow X\) be a continuous m-homogeneous polynomial with \(0<p<\infty \). Suppose that P is orthogonally additive on \(S^p(H)\). Then \(P=0\).

Proof

Since \({\mathscr {F}}(H)\) is dense in \(S^p(H)\) and P is continuous, it suffices to prove that P vanishes on \({\mathscr {F}}(H)\). On account of Lemma 1, we are also reduced to prove that P vanishes on \({\mathscr {F}}(H)_{\mathrm{sa}}\). We continue to use the notation \(\xi \otimes \eta \) which was introduced in the proof of Corollary 6.

Let \(x\in {\mathscr {F}}(H)_{\mathrm{sa}}\). Then \(x=\sum _{j=1}^k\alpha _j\xi _j\otimes \xi _j\), where \(k\ge 2\), \(\alpha _1,\ldots ,\alpha _k\in {\mathbb {R}}\), and \(\{\xi _1,\ldots ,\xi _k\}\) is an orthonormal subset of H. It is clear that the subalgebra \({\mathscr {M}}\) of \({{\mathscr {B}}}(H)\) generated by \(\bigl \{\xi _i\otimes \xi _j : i,j\in \{1,\ldots ,k\}\bigr \}\) is contained in \({{\mathscr {F}}}(H)\) and it is \(*\)-isomorphic to the von Neumann algebra \({{\mathscr {B}}}(K)\), where K is the linear span of the set \(\{\xi _1,\ldots ,\xi _k\}\). By Proposition 1, \(P\mid _{{\mathscr {M}}}=0\), and therefore \(P(x)=0\). We thus get \(P\mid _{{{\mathscr {F}}}(H)_{\mathrm{sa}}}=0\), as required. \(\square \)