1 Introduction

In past decades, several partial orders on matrix algebras and operator algebras have been considered. For example, star partial order as well as diamond order were introduced in [2, 7] and studied by many authors(cf. [1, 3, 5, 6, 8, 11, 14, 15]). As we know, the definition of star partial order in the algebra \({\mathscr {B}}({\mathscr {H}})\) of all bounded linear operators on a complex Hilbert space \({\mathscr {H}}\) is given by :

$$\begin{aligned} A\overset{*}{\le }B\Leftrightarrow A^{*}A=A^{*}B,~AA^{*}=BA^{*}\Leftrightarrow A=P_{A}B=BQ_{A}, \end{aligned}$$

where \(P_{A}\) and \(Q_{A}\) denote the left as well as right support projections of A respectively. In [3, 5], the star supremum and infimum were studied and the forms of star partial order-hereditary subspaces in \({\mathscr {B}}({\mathscr {H}})\) were characterized in [14]. Very recently, we given a type decomposition for operators in \({\mathscr {B}}({\mathscr {H}})\) with respect to the star partial order and characterized the star partial order automorphisms on the class of type 1 operators in [11]. On the other hand, Cīrulis in [2] introduced the diamond order in strong Rickart rings and the definition of it in \({\mathscr {B}}({\mathscr {H}})\) was given by:

$$\begin{aligned}&A \le ^{\diamond } B\Leftrightarrow \overline{R(A)} \subseteq \overline{R(B)},\\&\overline{R(A^{*})} \subseteq \overline{R(B^{*})},~AA^{*}A = AB^{*}A\Leftrightarrow A=P_{A}BQ_{A}=P_{B}AQ_{B}. \end{aligned}$$

We characterized the form of order automorphisms under the diamond order \(\le ^{\diamond }\) on the set of all products of two orthogonal projections in [10] and bounds for the diamond order in \({\mathscr {B}}({\mathscr {H}})\) were considered in [13].

We note that these two partial orders may extend to von Neumann algerbas as well as some unbounded operator spaces such as \(L^p\) spaces associated with semi-finite von Neumann algebras. For example, the star partial order in a von Neumann algebra was considered in [15]. We now extend these two partial orders to \(L^p({\mathscr {M}})(1\le p<\infty )\) for a semi-finite von Neumann algebra \({\mathscr {M}}\) with a faithful normal semi-finite trace \(\tau \).

Next, we recall some notions. Let \({\mathscr {M}}\) be a von Neumann algebra on \({\mathscr {H}}\) with a faithful normal semi-finite trace \(\tau \) and \(L^p({\mathscr {M}})(1\le p<\infty )\) the non-commutative \(L^p\) space associated with \(\tau \)(cf. [12]). Denote by \({\mathscr {M}}_p\) the set of all the projections in \({\mathscr {M}}\). For any \(A\in {\mathscr {M}}_p\), we denote by \(C_A\) the central carrier of A. Let \(E,~F\in {\mathscr {M}}_p\), \([E,F]=\{U\in {\mathscr {M}}: UU^{*}\le E,~U^{*}U\le F\}\). Moreover, For any \(A\in L^p({\mathscr {M}})\), we have \(\Vert A\Vert =(\tau (|A|^p))^{1/p}\) and \(P_{A}\), \(Q_{A}\) denote the left as well as right support projections of A in \({\mathscr {M}}\) respectively, that is, \(P_A=\inf \{P\in {\mathscr {M}}_p:PA=A\}\) and \(Q_A=\inf \{ Q\in {\mathscr {M}}_p: AQ=A\}\). If \(A=U|A|\) is the polar decomposition of A, then we have \(P_A=UU^*\) and \(Q_A=U^*U\).

2 The Star Partial Order in \(L^p({\mathscr {M}})\)

In this section, we extend the notion of the star partial order to non-commutative \(L^p\) spaces. We present some properties about the bounds of star partial order and then characterize forms of all hereditary subspaces with respect to star partial order.

As we know, in \({\mathscr {B}}({\mathscr {H}})\), \(A\overset{*}{\le }B\Leftrightarrow A^{*}A=A^{*}B,~AA^{*}=BA^{*}\Leftrightarrow A=P_{A}B=BQ_{A}\). We give the definition of star partial order in \(L^p({\mathscr {M}})\) by the same way.

Definition 2.1

Let \(A, B\in L^p({\mathscr {M}})\), we say \(A\overset{*}{\le }B\) if \(A^{*}A=A^{*}B\) and \(AA^{*}=BA^{*}\). \(\overset{*}{\le }\) is said to be the star partial order of \(L^p({\mathscr {M}})\).

We easily know that \(\overset{*}{\le }\) is a partial order in \(L^p({\mathscr {M}})\) and \( A \overset{*}{\le }B \Longleftrightarrow A=P_{A}B=BQ_{A}\). In [1, Lemma 2.6], Antezana gave a important result which has been used in many problems about bounds of star partial order. In non-commutative \(L^p\) spaces, we also have a similar result.

Definition 2.2

Let \(B\in L^p({\mathscr {M}})\).

  1. (1)

    If \(P\in {\mathscr {M}}_p\) and \(P\le P_{B}\), then \(PB\overset{*}{\le }B\Leftrightarrow PBB^{*}=BB^{*}P\).

  2. (2)

    If \(Q\in {\mathscr {M}}_p\) and \(Q\le Q_{B}\), then \(BQ\overset{*}{\le }B\Leftrightarrow QB^{*}B=B^{*}BQ\).

Proof

(1) Put \(A=PB\). If \(A\overset{*}{\le }B\), then \(PBB^{*}=AB^{*}=BA^{*}=BB^{*}P\). Conversely, if \(PBB^{*}=BB^{*}P\), then \(BA^{*}=BB^{*}P=PBB^{*}P=AA^{*}\) and \(A^{*}A=B^{*}PPB=B^{*}PB=B^{*}A\). Thus \(A\overset{*}{\le }B\).

(2) The proof is similar to (1). \(\square \)

In [3], the star lower as well as star upper bounds were characterized. Djikic gave a necessary and sufficient condition for the existence of star supremum of two arbitrary operators on a Hilbert space in [5]. In the next theorem, we present some properties about the star infimum and star supremum and prove that if a subset of \(L^p({\mathscr {M}})\) has a upper bound, then it must have supremum in \(L^p({\mathscr {M}})\). Let \({\mathscr {A}}\subseteq L^p({\mathscr {M}})\) and \(D\in L^p({\mathscr {M}})\). If \(A\overset{*}{\le }D\) for all \(A\in {\mathscr {A}}\), then we say that D is an upper bound of \({\mathscr {A}}\). If D is an upper bound of \({\mathscr {A}}\) and for any upper bound C of \({\mathscr {A}}\), we have that \(D\overset{*}{\le }C\), then we say that D is the supremum of \({\mathscr {A}}\) and denote by \(\sup {\mathscr {A}}\). We similarly may define lower bounds and the infimum \(\inf {\mathscr {A}}\) of \({\mathscr {A}}\).

Theorem 2.3

Let \({\mathscr {A}}\subseteq L^p({\mathscr {M}})\) be a nonempty subset.

  1. (1)

    There exists the infimum \(\inf {\mathscr {A}}\) in \(L^p({\mathscr {M}})\).

  2. (2)

    If there is a upper bound \(D\in L^p({\mathscr {M}})\), then there exists the supremum \(\sup {\mathscr {A}}\) in \(L^p({\mathscr {M}})\).

Proof

(1) Put \( P_{0}=\inf \{P_{A}: A\in {\mathscr {A}}\}\), \( Q_{0}=\inf \{Q_{A}: A\in {\mathscr {A}}\}\) and

$$\begin{aligned} {\mathscr {P}}=\{P\in {\mathscr {M}}_p: P\le P_{0}, P|A^*|=|A^*|P \text{ and } PA=PB, \forall A, B\in {\mathscr {A}}\}. \end{aligned}$$

For any \(P\in {\mathscr {P}}\) and \( A\in {\mathscr {A}}\), since \(P|A^*|=|A^*|P\), we have \(PAA^{*}=AA^{*}P\). It follows that \(PAA^{*}P=AA^{*}P\). Put \(G=PA\), then \(GG^{*}=AG^{*}\), \(G^{*}G=A^{*}PPA=A^{*}PA=G^{*}A\). This means that \(G=PA \overset{*}{\le }A\) for all \(A\in {\mathscr {A}}\). That is, \(G=PA\) is a lower bound of \({\mathscr {A}}\). Note that \(G=PA\) and \(P\in {\mathscr {P}}\), we can see that

$$\begin{aligned} \forall A, B\in {\mathscr {A}},~PA=PB \end{aligned}$$

by the definition of \({\mathscr {P}}\). This implies that G is independent of the choice of A.

Put \(P_1=\sup \{P:P\in {\mathscr {P}}\}\). Then \(P_1\in {\mathscr {P}}\). Put \(D=P_1A\) for all \(A\in {\mathscr {A}}\). Similarly, we can see that D is independent of the choice of A and is a lower bound of \({\mathscr {A}}\). Now for any \(C\overset{*}{\le }A \) for all \(A\in {\mathscr {A}}\), we have that \(C=P_{C}A=AQ_{C}\) for all \(A\in {\mathscr {A}}\). This implies that \(P_{C}\le P_{A}\) and \(P_{C}A=P_{C}B\) for all \(A, B \in {\mathscr {A}}\) and hence \(P_{C}\le P_{0}\). On the other hand, since \(CC^*=CA^{*}\), \(P_{C}AA^*P_{C}=CC^*=CA^{*}=P_{C}AA^*\) for all \(A\in {\mathscr {A}}\). It follows that \(P_{C}AA^*=AA^*P_{C}\), that is, \(P_{C}|A^*|=|A^*|P_{C}\). Thus \(P_{C}\in {\mathscr {P}}\). This implies that \(C=P_{C}A=P_{C}P_{1}A=P_{C}D\). It is elementary that \(C\overset{*}{\le }D\). Thus \(D=\inf {\mathscr {A}}\). (2) Put \({\mathscr {D}}=\{D\in L^p({\mathscr {M}}): A \overset{*}{\le } D,\ \forall A\in {\mathscr {A}}\}\) and \(M=\inf {\mathscr {D}}\). We claim that \(\sup {\mathscr {A}}=M\). Note that for any \(A\in {\mathscr {A}}\), \(A \overset{*}{\le }D , \forall D\in {\mathscr {D}}\). This means that for any \(A\in {\mathscr {A}}\), A is a lower bound of \({\mathscr {D}}\) and thus \(A\overset{*}{\le }M\), \(\forall A\in {\mathscr {A}}\). Hence \(M\in {\mathscr {D}}\) and \(M=\sup {\mathscr {A}}\). \(\square \)

We recall that a subspace \({\mathscr {A}}\subseteq {\mathscr {M}}\) is said to be star partial order-hereditary if for any \(B\in {\mathscr {M}}\) and \(A\in {\mathscr {A}}\), we have \(B\in {\mathscr {A}}\) whenever \(B\overset{*}{\le } A\)(cf. [14, 15]). In [15], all forms of \(\sigma -\)weakly closed star partial order-hereditary subspaces in a von Neumann algebra \({\mathscr {M}}\) were given. We may consider star partial order-hereditary subspaces in \(L^p({\mathscr {M}})\) by the same way. Next, we characterize the form of norm closed star partial order-hereditary subspace in \(L^{p}({\mathscr {M}})\). We note that the idea follows from [15]. We need to treat unbounded operators in \(L^p({\mathscr {M}})\) in this case, so we still give the detailed proof.

Lemma 2.4

Let \({\mathscr {A}}\) be a norm closed star partial order-hereditary subspace of \(L^{p}({\mathscr {M}})\). Then \(P_{T}L^{p}({\mathscr {M}})Q_{T}\subseteq {\mathscr {A}}\) for all \(T\in {\mathscr {A}}\).

Proof

Let \(T=W|T|\) be the polar decomposition of T and \(|T|=\int _{0}^{+\infty }\lambda dE_{\lambda }\) the spectral decomposition of |T|. Then we have \(W\in {\mathscr {M}}\) and \( |T|\in L^{p}({\mathscr {M}})\). Note that \(\forall \varepsilon > 0\), \(\varepsilon E[\varepsilon ,+\infty )\le E[\varepsilon ,+\infty )|T|\le |T|\). It follows that

$$\begin{aligned} \varepsilon ^p\tau ((E[\varepsilon , +\infty ))\le \tau (E[\varepsilon ,+\infty )|T|^p)\le \tau (|T|^p)< +\infty \end{aligned}$$

and therefore \(\tau (E[\varepsilon , +\infty ))<\infty \). It follows that \(E[\varepsilon ,+\infty )\mathscr {M}E[\varepsilon ,+\infty )\subseteq L^p({\mathscr {M}})\).

For any bounded Borel subset \(\Delta \subseteq [\varepsilon , +\infty )\), we have \(WE(\Delta )|T|\overset{*}{\le } T\). Then we easily see that \(WE(\Delta )|T| \in {\mathscr {A}}\). This implies that \(W\sum _{k=1}^ma_kE(\Delta _k)|T|\in {\mathscr {A}}\) for any complex \(a_k\in {\mathbb {C}}\) and pairwise disjoint bounded Borel subsets \(\Delta _k\) for any \(1 \le k\le m\). \(\forall a>\varepsilon \) and let f be a continuous function on \([\varepsilon ,a]\). It follows that \(WE[\varepsilon , a]f(|T|)|T| \in {\mathscr {A}}\). Put \(f(\lambda )=\lambda ^{-1}\) on \([\varepsilon ,a]\), then \(WE[\varepsilon , a]\in {\mathscr {A}}\). For any projection \(E\in {\mathscr {M}}_p\) such that \(E\le E[\varepsilon , a]\), it is easy to know that \(WE \overset{*}{\le } WE[\varepsilon , a]\). Then \(WE \in {\mathscr {A}}\). This means that \(WE[\varepsilon , a]{\mathscr {M}}E[\varepsilon , a]\subseteq {\mathscr {A}}\). Since \({\mathscr {A}}\) is norm closed and \(E[\varepsilon , a]{\mathscr {M}}E[\varepsilon , a]\) is dense in \(E[\varepsilon , a]L^p({\mathscr {M}})E[\varepsilon , a]\) in norm topology, \(WE[\varepsilon , a]L^{p}({\mathscr {M}})E[\varepsilon , a]\subseteq {\mathscr {A}}\). Note that \(Q_{T}=W^*W=E(0, +\infty )\). Putting \(\varepsilon \rightarrow 0 \) and \(a\rightarrow +\infty \), we have \(WL^{p}({\mathscr {M}})Q_{T}\subseteq {\mathscr {A}}\). Note that \(P_T=WW^*\). Thus

$$\begin{aligned} P_{T}L^{p}({\mathscr {M}})Q_{T}=P_{T}WL^{p}({\mathscr {M}})Q_{T}=WL^{p}({\mathscr {M}})Q_{T}\subseteq {\mathscr {A}}. \end{aligned}$$

\(\square \)

Next, we characterize the form of a norm closed star partial order-hereditary subspace of \(L^{p}({\mathscr {M}})\).

Theorem 2.5

Let \({\mathscr {A}}\) be a norm closed star partial order-hereditary subspace in \(L^{p}({\mathscr {M}})\). Then there exists unique projection pair (EF) with the same central carriers such that \({\mathscr {A}}=EL^{p}({\mathscr {M}})F\).

Proof

(i) Suppose \({\mathscr {A}}=EL^{p}({\mathscr {M}})F=E_{0}L^{p}({\mathscr {M}})F_{0}\) and \(C_{E}=C_{F}, C_{E_{0}}=C_{F_{0}}\), then we have \(E(L^{p}({\mathscr {M}})\bigcap {\mathscr {M}})F=E_{0}(L^{p}({\mathscr {M}})\bigcap {\mathscr {M}})F_{0}\). Clearly, \(\overline{E(L^{p}({\mathscr {M}})\bigcap {\mathscr {M}})F}^{\sigma -w}=\overline{E_{0}(L^{p}({\mathscr {M}})\bigcap {\mathscr {M}})F_{0}}^{\sigma -w}\), that is, \(E{\mathscr {M}}F=E_{0}{\mathscr {M}}F_{0}\). It follows from [15, Theorem 3.2] that \(E=E_0\) and \(F=F_0\).

(ii) If \({\mathscr {A}}=0\), then the conclusion is obviously true. Now suppose \({\mathscr {A}}\ne 0\). Put

$$\begin{aligned} \Lambda =\{(P,Q): P, Q \in {\mathscr {M}}_p,~ C_{P}=C_{Q},~ PL^{p}({\mathscr {M}})Q\subseteq {\mathscr {A}}\}. \end{aligned}$$

Since \((0,0)\in {\mathscr {M}}\), \(\Lambda \ne \emptyset \). Define a partial order on \(\Lambda \) by: \((P_{1}, Q_{1})\le (P_{2}, Q_{2})\Leftrightarrow P_{1}\le P_{2}\) and \(Q_{1}\le Q_{2}\). Let \(\{(P_{\alpha }, Q_{\alpha })\}_{\alpha \in I}\) be a totally ordered family of \(\Lambda \), then \(\{P_{\alpha }\}_{\alpha \in I}\) and \(\{Q_{\alpha }\}_{\alpha \in I}\) are monotone totally ordered sets respectively. Assume that \(P_{\alpha }\rightarrow P\) and \(Q_{\alpha }\rightarrow Q\) \(\sigma \)-weakly. It is clear that \(\{C_{P_{\alpha }}\}_{\alpha \in I}\) and \(\{C_{Q_{\alpha }}\}_{\alpha \in I}\) are also monotone totally ordered sets and \(C_{P_{\alpha }}\rightarrow C_{P},~C_{Q_{\alpha }}\rightarrow C_{Q}\) \(\sigma \)-weakly. Obviously, \(C_{P}=C_{Q}\). As \(P_{k}L^{p}({\mathscr {M}})Q_{k}\subseteq P_{l}L^{p}({\mathscr {M}})Q_{l}\subseteq {\mathscr {A}},~\forall k< l\), it is easy to see that \(PL^{p}({\mathscr {M}})Q\subseteq {\mathscr {A}}\). Thus \((P,Q)\in \Lambda \). By Zorn’s lemma, there is a maximal element (EF) in \(\Lambda \). Then we have \(EL^{p}({\mathscr {M}})F\subseteq {\mathscr {A}}\).

Next, we prove that \({\mathscr {A}} \subseteq EL^{p}({\mathscr {M}})F\). We give two assertions firstly.

Assertion 1 Let \((P, Q)\in \Lambda \). If \(P\le I-E,~ Q\le I-F\), then \(P=Q=0\).

Suppose that \(P \ne 0\). Take any monotonically increasing \(\tau -\)finite projection families \(\{P_{i}\}_{i\in I}\) and \(\{Q_{i}\}_{i\in I}\) in \({\mathscr {M}}_p\) such that \(P_{i} \rightarrow P\) and \(Q_{i} \rightarrow Q\) \(\sigma -\)weakly. If \(C_{P_{i} } \ne C_{Q_{i} }\) for some \(i\in I\), then \(\{P_{i} '\}_{i\in I}\) and \(\{Q_{i} '\}_{i\in I}\) are also monotonically increasing \(\tau -\)finite projection families with \(C_{P_{i} '}=C_{Q_{i} '}\) by putting \(P_{i} '=P_{i} C_{Q_{i} }\) and \(Q_{i} '=Q_{i} C_{P_{i} }\). It is easy to check that \(P_{i} '\rightarrow PC_{Q}=PC_{P}=P\) and \(Q_{i} '\rightarrow QC_{P}=QC_{Q}=Q\) \(\sigma -\)weakly. Without loss of generality, we may assume that \(C_{P_{i} }= C_{Q_{i} }\) for all \(i\in I\). Similarly, let \(\{E_{j}\}_{j\in J}\) and \(\{F_{j}\}_{j\in J}\) be monotonically increasing \(\tau -\)finite projection families satisfying \(E_{j}\rightarrow E\), \(F_{j}\rightarrow F\) \(\sigma -\)weakly and \(C_{E_{j}}= C_{F_{j}}\) for any \(j\in J\). Since all of these projections are \(\tau -\)finite, we have \([P_{i} , Q_{i} ]\subseteq P_i\mathscr {M}Q_i\subseteq P_{i} L^{p}({\mathscr {M}})Q_{i} \subseteq PL^{p}({\mathscr {M}})Q\subseteq {\mathscr {A}}\) and \([E_{j}, F_{j}]\subseteq E_j\mathscr {M}F_j\subseteq E_{j}L^{p}({\mathscr {M}})F_{j}\subseteq EL^{p}({\mathscr {M}})F\subseteq {\mathscr {A}}\).

For any \(U\in [E_{j}, F_{j}]\) and \(V\in [P_{i} , Q_{i} ]\), we have \(U+V\in {\mathscr {A}}\). It follows from Lemma 2.4 that \(P_{U+V}L^{p}({\mathscr {M}})Q_{U+V}\subseteq {\mathscr {A}}\). It is easy to see that \(P_{U}\le E_{j},~P_{V}\le P_{i} \le P \le I-E\le I-E_{j},~Q_{U}\le F_{j},~Q_{V}\le Q_{i} \le Q \le I-F \le I-F_{j}\), then \(P_{U}\bot P_{V},~Q_{U}\bot Q_{V}\). Similarly with [15, Theorem 3.2], \(\forall T\in L^{p}({\mathscr {M}})\), one obtains \(P_{U}TQ_{V}+P_{V}TQ_{U}\in {\mathscr {A}}\) and

$$\begin{aligned} P_{U}TQ_{V}\overset{*}{\le }P_{U}TQ_{V}+P_{V}TQ_{U}, \\ P_{V}TQ_{U}\overset{*}{\le }P_{U}TQ_{V}+P_{V}TQ_{U}. \end{aligned}$$

Since \({\mathscr {A}}\) is star partial order-hereditary, \(P_{U}TQ_{V},~P_{V}TQ_{U}\in {\mathscr {A}}\). By the arbitrariness of UVT, one obtains \(E_{j}L^{p}({\mathscr {M}})Q_{i} \subseteq {\mathscr {A}},~P_{i} L^{p}({\mathscr {M}})F_{j}\subseteq {\mathscr {A}}\), so \((E_{j}+P_{i} )L^{p}({\mathscr {M}})(F_{j}+Q_{i} )\subseteq {\mathscr {A}}\).

Put \(E_{j}'=(E_{j}+P_{i} )C_{(F_{j}+Q_{i} )},~F_{j}'=(F_{j}+Q_{i} )C_{(E_{j}+P_{i} )}\), then \(E_{j}'L^{p}({\mathscr {M}})F_{j}'\subseteq {\mathscr {A}}\). Since \(E_{j}\le C_{E_{j}}=C_{F_{j}}\le C_{(F_{j}+Q_{i} )}\) and \(F_{j}\le C_{F_{j}}=C_{E_{j}}\le C_{(E_{j}+P_{i} )}\), it follows that \(E_{j}\le E_{j}',~F_{j}\le F_{j}'\). Assume that \(E_{j}'\rightarrow E'\) and \( F_{j}'\rightarrow F'\) \(\sigma -\)weakly. Then \(E\le E' \) and \(F\le F'\). As \(E_{j}'L^{p}({\mathscr {M}})F_{j}'\subseteq {\mathscr {A}}\), it is easy to see that \(E'L^{p}({\mathscr {M}})F'\subseteq {\mathscr {A}}\). A direct calculation yields \(C_{E'}=C_{(E+P)}C_{(F+Q)}=C_{F'}\), so \((E', F')\in \Lambda \) and \((E,F)\le (E',F')\). We have \((E,F)=(E',F')\) because (EF) is the maximal element in \(\Lambda \). Note that

$$\begin{aligned} E'=(E+P)C_{(F+Q)}= & {} EC_{(F+Q)}+PC_{(F+Q)}\ge EC_{F}+PC_{Q}\\= & {} EC_{E}+PC_{P}=E+P, \end{aligned}$$

then \(E' > E\), which is a contradiction. Thus \(P=0\) and \(Q=0\).

Assertion 2 Let \((P,Q)\in \Lambda \). If \(P\le E,~Q\le I-F\) or \(P\le I-E,~Q\le F\), then \(P=Q=0\).

Suppose that \(P\le E,~Q\le I-F\). Put \(P'=E-P\), then \(E=P\oplus P'\). If \(P=E\), then \(C_{E}=C_{F}=C_{P}=C_{Q}=C_{(F+Q)}\). It is easy to check that \((E, F)\le (E, F+Q)\in \Lambda \), then \(Q=0\), which is a contradiction. Thus \(P< E\).

If \(P'C_{Q}L^{p}({\mathscr {M}})C_{P'}Q=P'L^{p}({\mathscr {M}})Q=0\), similarly with [15, Theorem 3.2], we easily get \(Q=0\), \(P=0\). If \(P'C_{Q}L^{p}({\mathscr {M}})C_{P'}Q=P'L^{p}({\mathscr {M}})Q\ne 0\), then we may take monotonically increasing \(\tau -\)finite projection families \(\{P_{i}\}_{i\in I}\) and \(\{Q_{i} \}_{i\in I}\) in \({\mathscr {M}}\) such that \(P_{i} \rightarrow P\) and \(Q_{i}\rightarrow Q\) \(\sigma -\)weakly with \(C_{P_{i}}=C_{Q_{i}}\) for any \(i\in I\). Similarly, we take monotonically increasing \(\tau -\)finite projection families \(\{P_{j}'\}_{j\in J}\) such that \(P_{j}'\rightarrow P'\) \(\sigma -\)weakly. Since \(C_{P_{j}'}\le C_{P'}\le C_{E}=C_{F}\), there exists a \(j_0\) such that \(P_{j}'L^{p}({\mathscr {M}})C_{P_{j}'}F=P_{j}'L^{p}({\mathscr {M}})F\ne 0\) for all \(j\ge j_0\). By [9, Proposition 5.5.3], we know, if T is a central projection in a von Neumann algebra \({\mathscr {M}}\), then \(TC_{A}=C_{TA}\) for each A in \({\mathscr {M}}\). As we know, \(T=C_{P_{j}'}\) is a central projection, then we have \(C_{C_{P_{j}'}F}=C_{P_{j}'}C_{F}\). Note that \(C_{P_{j}'}\le C_{P'}\le C_{E}=C_{F}\), then \(C_{P_{j}'}\le C_{F}\), thus we have \(C_{C_{P_{j}'}F}=C_{P_{j}'}C_{F}=C_{P_{j}'}\).

For any \(U \in [P_{i}, Q_{i}],~V\in [P_{j}',C_{P_{j}'}F]\), we easily see that \(U+V\in P_{i}L^{p}({\mathscr {M}})Q_{i}+P_{j}'L^{p}({\mathscr {M}})C_{P_{j}'}F\subseteq PL^{p}({\mathscr {M}})Q+EL^{p}({\mathscr {M}})F\subseteq {\mathscr {A}}\). It follows from Lemma 2.4 that \(P_{U+V}L^{p}({\mathscr {M}})Q_{U+V}\subseteq {\mathscr {A}}\). It is obvious that \(P_{U+V}=P_{U}+P_{V},~Q_{U+V}=Q_{U}+Q_{V}\). Using similar method in Assertion 1, we can prove that \(P_{V}L^{p}({\mathscr {M}})Q_{U}\subseteq {\mathscr {A}}\). By the arbitrariness of UV, we have \(P_{j}'L^{p}({\mathscr {M}})Q_{i}\subseteq {\mathscr {A}}\). Therefore, \(P'L^{p}({\mathscr {M}})Q\subseteq {\mathscr {A}}\). Then \(EL^{p}({\mathscr {M}})Q=(P+P')L^{p}({\mathscr {M}})Q\subseteq {\mathscr {A}}\). Similarly, one obtains \(Q=0\), \(P=0\).

If \(P\le I-E,~Q\le F\), we have the same conclusion.

We next prove that \({\mathscr {A}}\subseteq EL^{p}({\mathscr {M}})F\). \(\forall T\in {\mathscr {A}}\), by Lemma 2.4 we can see that \(P_{T}L^{p}({\mathscr {M}})Q_{T}\subseteq {\mathscr {A}}\) and \(C_{P_{T}}=C_{Q_{T}}\), so \((P_{T}, Q_{T})\in \Lambda \). Note that \(T\in P_{T}L^{p}({\mathscr {M}})Q_{T}\). So we just need to prove that \(P_{T}L^{p}({\mathscr {M}})Q_{T}\subseteq EL^{p}({\mathscr {M}})F\).

Let \(\{P_{i}\}_{i\in I}\) and \(\{Q_{i}\}_{i\in I}\) be monotonically increasing \(\tau -\)finite projection families with \(C_{P_{i}}=C_{Q_{i}}\) in \({\mathscr {M}}\) such that \(P_{i}\rightarrow P_{T}\) and \(Q_{i}\rightarrow Q_{T}\) \(\sigma -\)weakly. Then \(P_{i}L^{p}({\mathscr {M}})Q_{i}\subseteq P_{T}L^{p}({\mathscr {M}})Q_{T}\subseteq {\mathscr {A}}\), so \((P_{i}, Q_{i})\in \Lambda \). We can easily see that if we have proven that \(P_{i}L^{p}({\mathscr {M}})Q_{i}\subseteq EL^{p}({\mathscr {M}})F\) for all \(i\in I\), then the conclusion follows. Thus without loss of generality, we may assume that \(P=P_T\) and \(Q=Q_T\) are \(\tau \)-finite projections.

Let \(L_{1}=R(E)\cap R(P),~L_{2}=R(E)\cap N(P),~L_{3}=N(E)\cap R(P )\), \(L_{4}=N(E)\cap N(P),~L_{5}=R(E)\ominus (L_{1}\oplus L_{2}),~L_{6}={\mathscr {H}}\ominus (\oplus _{i=1}^{5}L_{i})\). Denote by \(P_{kk}=P_{L_{k}}\) and let \(I_{k}\) be the identity operator on \(L_{k}(1\le k \le 6)\). Let

$$\begin{aligned} P_{56}=0\oplus 0 \oplus 0 \oplus 0 \oplus \left( \begin{array}{cc} P_{0} &{} (P_{0}(I_{5}-P_{0}))^{\frac{1}{2}}D \\ D^{*}(P_{0}(I_{5}-P_{0}))^{\frac{1}{2}} &{} D^{*}(I_{5}-P_{0})D \\ \end{array} \right) , \end{aligned}$$

where \(P_{0}\in {\mathscr {B}}(L_{5})\) such that \(P_{0}\) and \(I_{5}-P_{0}\) are injective positive contraction operators, \(D\in {\mathscr {B}}(L_{6}, L_{5})\) is a unitary operator. Note that \(E,P\in {\mathscr {M}}\). Thus \(P_0,P_{ii}, P_{56},D\in {\mathscr {M}}\). By [4], we have that \({\mathscr {H}}=\bigoplus _{k=1}^6L_k \) and

$$\begin{aligned} E=P_{11}+P_{22}+P_{55},~P=P_{11}+P_{33}+P_{56}. \end{aligned}$$
(1)

Using the same method, let \(L_{1}'=R(F)\cap R(Q),~L_{2}'=R(F)\cap N(Q),~L_{3}'=N(F)\cap R(Q)\), \(L_{4}'=N(F)\cap N(Q),~L_{5}'=R(F)\ominus (L_{1}'\oplus L_{2}'),~L_{6}'={\mathscr {H}}\ominus (\oplus _{i=1}^{5}L_{i}')\). Denote by \(Q_{kk}=P_{L_{k}'}\) and let \(I_{k}\) be the identity operator on \(L_{k}'(1\le k \le 6)\). Let

$$\begin{aligned} Q_{56}=0\oplus 0 \oplus 0 \oplus 0 \oplus \left( \begin{array}{cc} Q_{0} &{} (Q_{0}(I_{5}-Q_{0}))^{\frac{1}{2}}K \\ K^{*}(Q_{0}(I_{5}-Q_{0}))^{\frac{1}{2}} &{} K^{*}(I_{5}-Q_{0})K \\ \end{array} \right) , \end{aligned}$$

where \(Q_{0}\in {\mathscr {B}}(L_{5}')\) such that \(Q_{0}\) and \(I_{5}-Q_{0}\) are injective positive contraction operators and \(K\in {\mathscr {B}}(L_{6}', L_{5}')\) is a unitary operator. In this case we have

$$\begin{aligned} F=Q_{11}+Q_{22}+Q_{55},~Q=Q_{11}+Q_{33}+Q_{56}. \end{aligned}$$
(2)

Since P and Q are \(\tau \)-finite projections, \(P_{ii}\), \(Q_{ii}(i=1,3)\), \(P_{56}\) and \(Q_{56}\) are all \(\tau \)-finite projections. It is known that

$$\begin{aligned} {\mathscr {A}}\supseteq PL^{p}({\mathscr {M}})Q=P_{11}L^{p}({\mathscr {M}})Q_{11}+P_{11}L^{p}({\mathscr {M}})Q_{33}+P_{11}L^{p}({\mathscr {M}})Q_{56}+\nonumber \\ P_{33}L^{p}({\mathscr {M}})Q_{11}+P_{33}L^{p}({\mathscr {M}})Q_{33}+P_{33}L^{p}({\mathscr {M}})Q_{56}+\nonumber \\ P_{56}L^{p}({\mathscr {M}})Q_{11}+P_{56}L^{p}({\mathscr {M}})Q_{33}+P_{56}L^{p}({\mathscr {M}})Q_{56}. \end{aligned}$$
(3)
  1. (1)

    It is easy to see that \(P_{11}L^{p}({\mathscr {M}})Q_{11}\subseteq EL^{p}({\mathscr {M}})F\).

  2. (2)

    \(P_{33}C_{Q_{33}}L^{p}({\mathscr {M}})C_{P_{33}}Q_{33}=P_{33}L^{p}({\mathscr {M}})Q_{33}\subseteq {\mathscr {A}}\), so \((P_{33}C_{Q_{33}}, C_{P_{33}}Q_{33})\in \Lambda \). It follows from Assertion 1 that \(P_{33}L^{p}({\mathscr {M}})Q_{33}=0\).

  3. (3)

    Similarly, by \(P_{11}L^{p}({\mathscr {M}})Q_{33}\subseteq {\mathscr {A}}\), \(P_{33}L^{p}({\mathscr {M}})Q_{11}\subseteq {\mathscr {A}}\) and Assertion 2, we have \(P_{11}L^{p}({\mathscr {M}})Q_{33}=0\) and \(P_{33}L^{p}({\mathscr {M}})Q_{11}=0\).

  4. (4)

    Suppose that \(P_{56}L^{p}({\mathscr {M}})Q_{56}\ne 0\). Since \(P_{55}L^{p}({\mathscr {M}})Q_{55}\subseteq EL^{p}({\mathscr {M}})F\subseteq {\mathscr {A}}\), for any \(T_{55}\in P_{55}L^{p}({\mathscr {M}})Q_{55}\), we have \( P_{56}T_{55}Q_{56}\in {\mathscr {A}}\) and

    $$\begin{aligned} P_{56}T_{55}Q_{56}&=P_{55}P_{0}T_{55}Q_{0}Q_{55} +P_{55}P_{0}T_{55}(Q_{0}(I_{5}-Q_{0}))^{\frac{1}{2}}KQ_{66}\nonumber \\&\quad + P_{66}D^{*}(P_{0}(I_{5}-P_{0}))^{\frac{1}{2}}T_{55}Q_{0}Q_{55}\\&\quad + P_{66}D^{*}(P_{0}(I_{5}-P_{0}))^{\frac{1}{2}}T_{55}(Q_{0}(I_{5}-Q_{0}))^{\frac{1}{2}}KQ_{66}.\end{aligned}$$

    Let

    $$\begin{aligned} G= & {} P_{55}P_{0}T_{55}(Q_{0}(I_{5}-Q_{0}))^{\frac{1}{2}}KQ_{66}\\&+P_{66}D^{*}(P_{0}(I_{5}-P_{0}))^{\frac{1}{2}}T_{55}(Q_{0}(I_{5}-Q_{0}))^{\frac{1}{2}}KQ_{66}, \\ H= & {} P_{66}D^{*}(P_{0}(I_{5}-P_{0}))^{\frac{1}{2}}T_{55}Q_{0}Q_{55}\\&+P_{66}D^{*}(P_{0}(I_{5}-P_{0}))^{\frac{1}{2}}T_{55}(Q_{0}(I_{5}-Q_{0}))^{\frac{1}{2}}KQ_{66}, \end{aligned}$$

    and let

    $$\begin{aligned} J&=-P_{55}(I_{5}-P_{0})T_{55}Q_{0}Q_{55}+P_{55}P_{0}T_{55}(Q_{0}(I_{5}-Q_{0}))^{\frac{1}{2}}KQ_{66}\nonumber \\&\quad +P_{66}D^{*}(P_{0}(I_{5}-P_{0}))^{\frac{1}{2}}T_{55}Q_{0}Q_{55}+P_{66}D^{*}(P_{0}(I_{5}-P_{0}))^{\frac{1}{2}}\\&\quad T_{55}(Q_{0}(I_{5}-Q_{0}))^{\frac{1}{2}}KQ_{66}, \end{aligned}$$
    $$\begin{aligned} K&=-P_{55}P_{0}T_{55}(I_{5}-Q_{0})Q_{55}+P_{55}P_{0}T_{55}(Q_{0}(I_{5}-Q_{0}))^{\frac{1}{2}}KQ_{66}\nonumber \\&\quad +P_{66}D^{*}(P_{0}(I_{5}-P_{0}))^{\frac{1}{2}}T_{55}Q_{0}Q_{55}+P_{66}D^{*}(P_{0}(I_{5}-P_{0}))^{\frac{1}{2}}\\&\quad T_{55}(Q_{0}(I_{5}-Q_{0}))^{\frac{1}{2}}KQ_{66}. \end{aligned}$$

    It is easy to see that \(J, K\in {\mathscr {A}}\). By an elementary calculation, we have \(G\overset{*}{\le }J,~H\overset{*}{\le }K\). Then \(G, H, H-G\in {\mathscr {A}}\). Let \(G_{0}=-P_{55}P_{0}T_{55}(Q_{0}(I_{5}-Q_{0}))^{\frac{1}{2}}KQ_{66}\), \(H_{0}=P_{66}D^{*}(P_{0}(I_{5}-P_{0}))^{\frac{1}{2}}T_{55}Q_{0}Q_{55}\). Since \(G_{0}\overset{*}{\le }H-G,~H_{0}\overset{*}{\le }H-G\), we have \(G_{0}, H_{0}, G+G_{0}\in {\mathscr {A}}\). Note that \(G_{0}\in EL^{p}({\mathscr {M}})(I-F),~H_{0}\in (I-E)L^{p}({\mathscr {M}})F,~G+G_{0}\in (I-E)L^{p}({\mathscr {M}})(I-F)\). By Assertion 1 and Assertion 2, we have \(G_{0}=H_{0}=G+G_{0}=0\). Since \(P_{0}, (Q_{0}(I_{5}-Q_{0}))^{\frac{1}{2}}\) are injective positive operators and K is unitary operator, \(T_{55}=0\), that is, \(P_{55}L^{p}({\mathscr {M}})Q_{55}=0\). Thus \(C_{P_{55}}C_{Q_{55}}=0\). Note that \(C_{P_{55}}=C_{P_{66}},~C_{Q_{55}}=C_{Q_{66}}\), then

    $$\begin{aligned} P_{56}L^{p}({\mathscr {M}})Q_{56}\subseteq (P_{55}+P_{66})L^{p}({\mathscr {M}})(Q_{55}+Q_{66})=0, \end{aligned}$$

    which is a contradiction. Therefore, \(P_{56}L^{p}({\mathscr {M}})Q_{56}=0\).

  5. (5)

    Since \(P_{56}L^{p}({\mathscr {M}})Q_{11}=(P_{55}+P_{66})P_{56}L^{p}({\mathscr {M}})Q_{11}\) and \(P_{55}P_{56}L^{p}({\mathscr {M}})Q_{11}\subseteq EL^{p}({\mathscr {M}})F\subseteq {\mathscr {A}}\), one obtains \(P_{66}P_{56}L^{p}({\mathscr {M}})Q_{11}\subseteq {\mathscr {A}}\). It follows from Assertion 2 that \(P_{66}P_{56}L^{p}({\mathscr {M}})Q_{11}=0\). Then for any \(T_{51}\in P_{55}L^{p}({\mathscr {M}})Q_{11}\), we have \(D^{*}(P_{0}(I_{5}-P_{0}))^{\frac{1}{2}}T_{51}=0\). This implies that \(T_{51}=0\). Then \(0=C_{P_{55}}C_{Q_{11}}=C_{P_{66}}C_{Q_{11}}=C_{P_{56}}C_{Q_{11}}\). Thus we have \(P_{56}L^{p}({\mathscr {M}})Q_{11}=0\). Symmetrically, we can see that \(P_{11}L^{p}({\mathscr {M}})Q_{56}=0\).

  6. (6)

    Suppose that \(P_{56}L^{p}({\mathscr {M}})Q_{33}\ne 0\). Without loss of generality, we may assume that \(C_{P_{56}}=C_{Q_{33}}\). At this time, \(C_{P_{55}}=C_{P_{56}}=C_{P_{66}}=C_{Q_{33}}\ne 0\). Then there exists a nonzero partial isometry U such that \(U^{*}U\le Q_{33},~UU^{*}\le P_{55}\). Since \(Q_{33}\) is \(\tau \)-finite, \(U\in P_{55}L^{p}({\mathscr {M}})Q_{33}\). Obviously, \(0\ne A=P_{56}UQ_{33}\in P_{A}L^{p}({\mathscr {M}})Q_{A}\subseteq P_{56}L^{p}({\mathscr {M}})Q_{33}\subseteq {\mathscr {A}}\). Similarly with [15, Theorem 3.2], we may assume that \(P_{A}\) is equivalent to a subprojection of F.

Using the same method, E and \(P_{A}\) can be represented in the form of equation (1). \(E=P_{11}'+P_{22}'+P_{55}',~P_{A}=P_{11}'+P_{33}'+P_{56}'\). Then

$$\begin{aligned} {\mathscr {A}}\supseteq P_{A}L^{p}({\mathscr {M}})Q_{A}=P_{11}'L^{p}({\mathscr {M}})Q_{A}+P_{33}'L^{p}({\mathscr {M}})Q_{A}+P_{56}'L^{p}({\mathscr {M}})Q_{A}. \end{aligned}$$

Thus \(P_{11}'L^{p}({\mathscr {M}})Q_{A},~P_{33}'L^{p}({\mathscr {M}})Q_{A},~P_{56}'L^{p}({\mathscr {M}})Q_{A}\subseteq {\mathscr {A}}\). Since \(P_{11}'\le E,~Q_{A}\le I-F,~P_{33}'\le I-E\), we have \(P_{11}'L^{p}({\mathscr {M}})Q_{A}=0,~P_{33}'L^{p}({\mathscr {M}})Q_{A}=0\) by Assertion 1 and Assertion 2. Thus \(P_{56}'L^{p}({\mathscr {M}})Q_{A}\ne 0\).

Let \(P_{0}'\) and \(D'\) be the injective positive contraction operators and unitary operator which corresponding to \(P_{56}'\), respectively. Similarly with [15, Theorem 3.2], we may assume that there exist \(\alpha , \beta \in (0, 1)\) such that \(0<\alpha \le P_{0}' \le \beta <1\).

Since \(P_{56}'\le P_{A}\thicksim Q_{A}\) and \(P_{A}\) is equivalent to a subprojection of F, there exist \(Q_{33}'\le Q_{A},~F_{1}\le F\) such that \(P_{55}' \thicksim P_{56}' \thicksim Q_{33}' \thicksim F_{1}\). Then there exists a nonzero partial isometry U such that \(U^{*}U= Q_{33}',~UU^{*}= P_{55}'\). Because \(Q_{A}\le Q\) and Q is \(\tau -\)finite, \(P_{55}',~P_{56}',~ Q_{33}'\) and \(F_{1}\) are all \(\tau -\)finite. Then \(U\in P_{55}'L^{p}({\mathscr {M}})Q_{33}'\). Clearly, \(A'=P_{56}'UQ_{33}'\ne 0\) and \(A'\in P_{56}'L^{p}({\mathscr {M}})Q_{33}'\subseteq P_{A}L^{p}({\mathscr {M}})Q_{A}\subseteq {\mathscr {A}}\). It is easy to check that \(A'=P_{55}'P_{0}'UQ_{33}'+P_{66}'D'^{*}(P_{0}'(I_{5}'-P_{0}'))^{\frac{1}{2}}UQ_{33}'\). Note that \(P_{55}'\sim F_1\). Then there exist some invertible elements in \(P_{55}'\mathscr {M}F_1\). Since \(0<\alpha \le P_{0}' \le \beta <1\), for any invertible \(X\in P_{55}'\mathscr {M}F_1\subseteq P_{55}'L^{p}({\mathscr {M}})F_{1}\subseteq {\mathscr {A}}\), \(B_{X}=X+A'\in {\mathscr {A}}\) is also invertible and

$$\begin{aligned} B_{X}=P_{55}'XF_{1}+P_{55}'P_{0}'UQ_{33}'+P_{66}'D'^{*}(P_{0}'(I_{5}'-P_{0}'))^{\frac{1}{2}}UQ_{33}'. \end{aligned}$$

It is easy to see that for any \(0<\lambda <1\),

$$\begin{aligned} E_{\lambda }=\frac{1}{1+\lambda ^{2}}(P_{55}'+\lambda P_{55}'D'P_{66}'+\lambda P_{66}'D'^{*}P_{55}'+\lambda ^{2}P_{66}') \end{aligned}$$

is a projection. It follows form Lemma 2.2 that if \(E_{\lambda }B_{X}B_{X}^{*}=B_{X}B_{X}^{*}E_{\lambda }\), then \(E_{\lambda }B_{X}\overset{*}{\le } B_{X}\).

In fact, put \(P(\lambda )=\frac{1-\lambda ^{2}}{\lambda }(P_{0}')^{\frac{3}{2}}(I_{5}'-P_{0}')^{\frac{1}{2}}+P_{0}'-2(P_{0}')^{2}\). Since \(0<\alpha \le P_{0}' \le \beta <1\), there exits \(\lambda >0\) such that \(P(\lambda )\) is positive invertible operator. Note that \(P_{55}'\) and \(F_{1}\) are \(\tau -\)finite projections. Let W be a partial isometry with \(WW^{*}=P_{55}',~W^{*}W=F_{1}\) and let \(X=(P(\lambda ))^{\frac{1}{2}}W\). Similarly with [15, Theorem 3.2], we can easily see that X and \(\lambda \) satisfy the equation \(E_{\lambda }B_{X}B_{X}^{*}=B_{X}B_{X}^{*}E_{\lambda }\), so \(E_{\lambda }B_{X}\overset{*}{\le } B_{X}\). Since \(B_{X}\in {\mathscr {A}}\), we have

$$\begin{aligned} {\mathscr {A}}\ni (1+\lambda ^{2})E_{\lambda }B_{X}&=P_{55}'XF_{1}+P_{55}'(P_{0}'U+\lambda (P_{0}'(I_{5}'-P_{0}'))^{\frac{1}{2}}U)Q_{33}'\nonumber \\&\quad +\lambda P_{66}' D'^{*}XF_{1}+P_{66}'(\lambda D'^{*}P_{0}'U+\lambda ^{2}D'^{*}(P_{0}'(I_{5}'-P_{0}'))^{\frac{1}{2}}U)Q_{33}'. \end{aligned}$$

Let

$$\begin{aligned} A_{1}=\lambda P_{66}' D'^{*}XF_{1}+P_{66}'(\lambda D'^{*}P_{0}'U+\lambda ^{2}D'^{*}(P_{0}'(I_{5}'-P_{0}'))^{\frac{1}{2}}U)Q_{33}', \\ Y=-P_{55}'(P_{0}'+\lambda (P_{0}'(I_{5}'-P_{0}'))^{\frac{1}{2}})^{2}X^{*-1}F_{1}\in P_{55}'L^{p}({\mathscr {M}})F_{1}\subseteq {\mathscr {A}}. \end{aligned}$$

Then \(B_{1}=(1+\lambda ^{2})E_{\lambda }B_{X}+Y-X\in {\mathscr {A}}\). It is easy to check that \(A_{1}\overset{*}{\le }B_{1}\), which implies that \(A_{1}\in {\mathscr {A}}\). Thus \(B_{1}-A_{1}-Y\in {\mathscr {A}}\). In fact, \(B_{1}-A_{1}-Y\in P_{55}'L^{p}({\mathscr {M}})Q_{33}'\), by Assertion 2 we have \(B_{1}-A_{1}-Y=0\). Then

$$\begin{aligned} 0=(P_{0}'+\lambda (P_{0}'(I_{5}'-P_{0}'))^{\frac{1}{2}})U, \end{aligned}$$

which implies \(U=0\). It is a contradiction. Therefore, \(P_{56}L^{p}({\mathscr {M}})Q_{33}= 0\). Symmetrically, we easily get \(P_{33}L^{p}({\mathscr {M}})Q_{56}= 0\).

From \((1)\thicksim (6)\), we know that \(PL^{p}({\mathscr {M}})Q=P_{11}L^{p}({\mathscr {M}})Q_{11}\subseteq EL^{p}({\mathscr {M}})F\). Thus we have \({\mathscr {A}}\subseteq EL^{p}({\mathscr {M}})F\). Therefore, \({\mathscr {A}}= EL^{p}({\mathscr {M}})F\). \(\square \)

If \({\mathscr {A}}\subseteq L^p({\mathscr {M}})\) is a norm closed star partial order-hereditary subspace of \(L^{p}({\mathscr {M}})\), then we have that \({\mathscr {A}}\cap {\mathscr {M}}\) is a star partial order-hereditary liner manifold in \({\mathscr {M}}\). By Theorem 2.5, we have that \(\overline{{\mathscr {A}}\cap {\mathscr {M}}}^{\sigma -w}=E\mathscr {M}F\) for some \(E,F\in {\mathscr {M}}_p\) with the same central carriers. In general, let \({\mathfrak {M}}\subseteq {\mathscr {M}}\) be a star partial order-hereditary liner manifold in \({\mathscr {M}}\). Is the \(\sigma \)-weak closure \(\overline{{\mathfrak {M}}}^{\sigma -w}\) of \({\mathfrak {M}}\) a star partial order-hereditary subspace?

3 The Diamond Order in \(L^{p}({\mathscr {M}})\)

In this section, we consider the diamond order in non-commutative \(L^p\) spaces. The definition is similar to the case in \({\mathscr {B}}({\mathscr {H}})\).

Definition 3.1

Let \(A, B\in L^p({\mathscr {M}})\). Then we say that \(A \le ^{\diamond } B\) if \( P_{A}\le P_{B},~ Q_{A}\le Q_{B},~AA^{*}A = AB^{*}A\) and \(\le ^{\diamond }\) is said to be diamond order.

As in \({\mathscr {B}}({\mathscr {H}})\), it is also known that \(A \le ^{\diamond } B\Leftrightarrow A=P_{A}BQ_{A}=P_{B}AQ_{B}\). We now may consider the upper bounds as well as lower bounds of a subset \({\mathscr {A}}\subseteq L^p({\mathscr {M}})\). However, there is no supremum and infimum in general. It is known that there exist minimal upper bounds for a subset with an upper bound in \({\mathscr {B}}({\mathscr {H}})\) (cf. [13]). Now for a finite von Neumann algebra \({\mathscr {M}}\), we also have the following result.

Theorem 3.2

Let \({\mathscr {M}}\) be a finite von Neumann algebra with a faithful normal finite trace \(\tau \) and \(L^p({\mathscr {M}})\) the non-commutative \(L^p\)-space associated with \(\tau \). If \({\mathscr {A}}\subseteq L^p({\mathscr {M}})\) has an upper bound, then \({\mathscr {A}}\) has a minimal upper bound in \(L^p({\mathscr {M}})\).

Proof

Let \({\mathscr {A}}\) be a nonempty subset in \(L^p({\mathscr {M}})\) and \(B\in L^p({\mathscr {M}})\) an upper bound of \({\mathscr {A}}\) with respect to the diamond order. Put \(P=\bigvee \{P_A: A\in {\mathscr {A}}\}\) and \(Q=\bigvee \{Q_A:A\in {\mathscr {A}}\}\). Then \(P\le P_B\) and \(Q\le Q_B\). Put \(B_{11}=PBQ\), \(P_1=P_{B_{11}}\) and \(Q_1=Q_{B_{11}}\). Then \(P_1\le P\le P_B\) and \(Q_1\le Q\le Q_B\). It follows that \(B_{11}=P_1B_{11}Q_1=P_1BQ_1=P_{B_{11}}BQ_{B_{11}}\). Thus \(B_{11}\le ^{\diamond } B\).

If \( P_1=P\) and \( Q_1=Q\), then it is easy to see that \(B_{11}\) is a minimal upper bound of \({\mathscr {A}}\). Otherwise, we let \(Q_2=I-Q\), \(Q_3=Q-Q_1\), \(P_2=P-P_1\) and \(P_3=I-P\). Put \(B_{ij}=P_iBQ_j\) for \(1\le i,j\le 3\). Since \(PBQ=P_1BQ_1\), we have \(B_{13}=0\), \(B_{21}=0 \) and \(B_{23}=0\). Note that \(P\le P_B\) as well as \(Q\le Q_B\). We also have that \(P_{B_{22}}=P_2\) and \(Q_{B_{33}}=Q_3\). Next we may assume that \(P_{1}<P\) and \(Q_{1}<Q\). The case that either \(P_{1}=P\) or \(Q_{1}=Q\) is similar. Let \(B_{22}=U|B_{22}|\) and \(B_{33}=V|B_{33}|\) be the polar decompositions of \(B_{22} \) and \(B_{33}\). Then \(UU^*=P_2\), \(U^*U\le Q_2\) and \(V^*V=Q_3\), \(VV^*\le P_3\). Put \(D=B_{11}+U+V\). Then \(D\in L^p({\mathscr {M}})\) since \( {\mathscr {M}}\subseteq L^p({\mathscr {M}})\). We next show that D is a minimal upper bound of \({\mathscr {A}}\).

We note that \(P_D=P_1+P_2+VV^*=P+VV^*\ge P_A\) and \(Q_D=Q_1+U^*U+Q_3=Q+U^*U\ge Q_A\) for all \(A\in {\mathscr {A}}\). Then \(A=P_DAQ_D\) for all \(A\in {\mathscr {A}}\). On the other hand, for any \(A\in {\mathscr {A}}\), \(P_A\le P\) and \(Q_A\le Q\). It follows that \(P_ADQ_A=P_A(PBQ+U+V)Q_A=P_APBQQ_A=P_ABQ_A=A\). Thus \(A\le ^{\diamond }D\).

Now for any \(C\in L^p({\mathscr {M}})\) with \(A\le ^{\diamond }C\le ^{\diamond }D\) for any \(A\in {\mathscr {A}}\), we have \(P\le P_C\) and \(Q\le Q_C\). These mean that \(P_C=P+E\) and \(Q_C=Q+F\) for some projections \(E,F \in {\mathscr {M}}_p\) with \(E\le I-P\) and \(F\le I-Q\). Note that \(C\le ^{\diamond }D\). We have \(C=P_CDQ_C=B_{11}+UF+EV\). It is known that both UF and \(V^*E\) have dense ranges since \(P\le P_C\) and \(Q\le Q_C\). By an elementary calculation, we have \( UF\le ^{\diamond }U\) and \(EV\le ^{\diamond }V\). It follows that \(UF=U\) and \(EV=V\) by [13, Corollary 2.4] since both U and \(V^{*}\) are surjective. Hence D is a minimal upper bound of \({\mathscr {A}}\). \(\square \)

However, if \({\mathscr {M}}\) is not finite, then the result in Theorem 3.2 may fail. We note that for a separable infinite dimensional Hilbert space \({\mathscr {H}}\), \({\mathscr {B}}({\mathscr {H}})\) is a semi-finite von Neumann algebra with the classical faithful normal semi-finite trace \(\tau \) and the non-commutative \(L^p\) space associated with \({\mathscr {B}}({\mathscr {H}})\) is the Schatten p-class \(C_p({\mathscr {H}})=\{ T\in {\mathscr {B}}({\mathscr {H}}): \tau (|T|^p)<\infty \}\). We next give a subset \({\mathscr {A}}\subseteq C_p({\mathscr {H}})\) with an upper bound but has no minimal upper bounds.

Example 3.3

Let \({\mathscr {K}}\) be a complex separable infinite dimensional Hilbert space with an orthonormal basis \(\{e_n:n\ge 1\}\) and \({\mathscr {H}}={\mathscr {K}}\oplus {\mathscr {K}}\). Let \(H_1=\vee \{e_1\}\) and \(T_1e_1=e_1\) on \(H_1\). For any \(n\ge 2\), let \(K_{n-1}=\vee \{e_{2^{n-2}+j}:1\le j\le 2^{n-2}\}\) and \(H_{n}=H_{n-1}\oplus K_{n-1}\). We define \(B_{n-1}e_{2^{n-2}+j}=ne_j\) for \(1\le j\le 2^{n-2}\) from \(K_{n-1}\) to \(H_{n-1}\) and

$$\begin{aligned} T_n=\left( \begin{array}{cc} T_{n-1} &{} B_{n-1} \\ 0 &{} I \\ \end{array} \right) \end{aligned}$$

on \(H_n\). Then \(T_n \) is an invertible operator on \(H_n\). Let \(P_n \) and \(Q_n\) be the projections from \({\mathscr {H}}\) onto \(H_n\oplus \{0\}\) and \(M_n=\{x\oplus T_nx:x\in H_n\}\) respectively. Note that both \(\{P_n\}\) and \(\{Q_n\}\) are increasing sequences. It is elementary that \(\lim _nP_n=I\oplus 0=P\). On the other hand, for any \(j\ge 1\), put \(x_n^j=0\oplus \frac{1}{n}e_{2^{n-2}+j}\) for \(n\ge 2\). we have \(T_nx_n^j=e_j+ \frac{1}{n}e_{2^{n-2}+j}\). Then \(\lim _n(x_n^j\oplus T_nx_n^j)=0\oplus e_j\). It follows that \(\{0\}\oplus {\mathscr {K}}\subseteq \vee \{M_n:n\ge 1\}\). Thus \(\lim _nQ_n=I\oplus I\).

We now take a positive injective operator \(A\in C_p({\mathscr {H}})\) such that \(P_nA=AP_n\) for all n. Let \(A_n=P_nAQ_n=AP_nQ_n\) for \(n\ge 1\) and \({\mathscr {A}}=\{A_n:n\ge 1\}\). Note that \(T_n\) and \(A_n|_{P_n({\mathscr {H}})}\) are invertible on \(H_n\) and \(H_n\oplus \{0\}\) for all n. Then \(P_{A_n}=P_n\) as well as \(Q_{A_n}=Q_n\). It follows that \(A_n\le ^{\diamond }A\) in \(C_p({\mathscr {H}})\) for all n and thus A is an upper bound of \({\mathscr {A}}\). It is clear that \(\lim _nA_n=PA=AP=PAP\). Take any upper bound \(D\in C_p({\mathscr {H}})\) of \({\mathscr {A}}\). Then \(P\le P_D\), \(Q_D=I\) and \(A_n=P_nDQ_n\). Thus \(PA=PD=PDP\). This means that

$$\begin{aligned} A=\left( \begin{array}{cc} A_{11} &{} 0 \\ 0 &{} A_{22} \\ \end{array} \right) \text{ and } D=\left( \begin{array}{cc} A_{11} &{} 0 \\ D_{21} &{} D_{22} \\ \end{array} \right) . \end{aligned}$$

Note that \(D_{22}\) is injective and \(P_D=P_{A_{11}}\oplus P_{[D_{21},D_{22}]}\). Since \(R(D_{22}^*)\) is not closed and dense in \({\mathscr {K}}\), there exists an injective operator \(G_{22}\) on \({\mathscr {K}}\) such that \(G_{22}\le ^{\diamond }D_{22}\) by [13, Lemma 2.5]. Thus \(G_{22}=P_{G_{22} }D_{22}\). Put

$$\begin{aligned} G=\left( \begin{array}{cc} I &{} 0 \\ 0 &{} P_{G_{22}} \\ \end{array} \right) \left( \begin{array}{cc} A_{11} &{} 0 \\ D_{21} &{} D_{22} \\ \end{array} \right) =\left( \begin{array}{cc} A_{11} &{} 0 \\ P_{G_{22}}D_{21} &{} G_{22} \\ \end{array} \right) . \end{aligned}$$

Thus \(P\le P_G \le P_D\) and \(Q_G=Q_D=I\). By an elementary calculation, we have \(A_n\le ^{\diamond }G\le ^{\diamond }D\) for all \(n\ge 1\) in \(C_p({\mathscr {H}})\). Note that \(G\ne D\). Thus \({\mathscr {A}}\) has not any minimal upper bound.

We now may consider the diamond order-hereditary subspaces in \(L^{p}({\mathscr {M}})\).

Proposition 3.4

Let \({\mathscr {M}}\) be a semi-finite von Neumann algebra with a faithful normal semi-finite trace \(\tau \) and \(L^p({\mathscr {M}})\) the non-commutative \(L^p\) space associated with \(\tau \). If \({\mathscr {A}}\) is a norm closed diamond order-hereditary subspace in \(L^{p}({\mathscr {M}})\), then there exists unique projection pair (EF) with same central carriers such that \({\mathscr {A}}=EL^{p}({\mathscr {M}})F\).

Proof

Suppose that \({\mathscr {A}}\) is a norm closed diamond order-hereditary subspace in \(L^{p}({\mathscr {M}})\). For any \(A\in L^p({\mathscr {M}})\) and \(B\in {\mathscr {A}}\), if \(A\overset{*}{\le }B\), then we easily see that \(A\le ^{\diamond }B\). Thus we have \(A\in {\mathscr {A}}\). Therefore, a diamond order-hereditary subspace must be a star partial order-hereditary subspace. The conclusion follows from Theorem 2.5. \(\square \)