1 Introduction

The theory of noncommutative symmetric spaces has been rapidly developed. Many of the noncommutative martingale results have been transferred to the noncommutative symmetric case. Especially, in [1], J. Yong proved Burkholder-Gundy inequalities for symmetric Banach spaces of noncommutative martingales. In [9], T. N. Bekjan proved the duality for conditional Hardy spaces of martingales in noncommutative symmetric Banach spaces.

The quasi-martingales are generalizations of martingales and play important roles in many different areas of mathematics. In [15], we studied duality theorems for \(L_p\)-spaces of noncommutative quasi-martingales. In this paper, we will extend the above results to the noncommutative symmetric case. Let E be a symmetric Banach space on [0, 1] with the Fatou property and \(1<p_E\le q_E<p\). Then

$$\begin{aligned} (_p\widehat{E}(\mathcal {M}))^*= {_{p'}{\widehat{E}^\times }(\mathcal {M})}, \end{aligned}$$

where \(_p\widehat{E}(\mathcal {M})\) and \({_{p'}{\widehat{E}^\times }(\mathcal {M})}\) denote the symmetric Banach spaces of noncommutative quasi-martingales which we refer to the next section for formal definitions. As applications of this result, we obtain the description of the symmetric space \(_p\widehat{E}(\mathcal {M})\) as interpolations of noncommutative quasi-martingale \(L_p\)-spaces.

The organization of the paper is as follows. In Section 2, we give some preliminaries and notations on symmetric Banach spaces, quasi-martingale spaces and interpolations. We prove the main results in Section 3.

2 Preliminaries

Let E be a symmetric Banach space on [0, 1]. The K\({\ddot{\text{ o }}}\)the dual of E is the function space defined by setting:

$$\begin{aligned} E^{\times }=\{f\in L_0([0,1]) :\int ^1_0|f(t)g(t)|dt<\infty , \forall g\in E \}. \end{aligned}$$

When equipped with the norm \(\Vert f\Vert _{E^\times }:=\sup \{\int ^1_0|f(t)g(t)|dt:\Vert g\Vert _E\le 1\},\) \(E^{\times }\) is a symmetric Banach space.

A symmetric Banach space E on [0, 1] is said to have the Fatou property if for every sequence \((x_n)_{n}\) in E satisfying \(0\le x_n \uparrow \) and \(\sup _{n} \Vert x_n\Vert _E<\infty \), the supremum \(x=\sup _{n}x_n\) belongs to E and \(\Vert x_n\Vert _E\uparrow \Vert x\Vert _E\). Note that E has the Fatou property if and only if \(E=E^{\times \times }\) isometrically. Examples of symmetric spaces with the Fatou property are separable symmetric spaces and duals of separable symmetric spaces.

For any \(s > 0\) we define the dilation operator \(D_s\) on \(L_0[0,1]\) by

$$\begin{aligned} (D_sf)(t)= f(st)\chi _{[0,1]}(st),\ \ t\in [0,1]. \end{aligned}$$

If E is a symmetric Banach space on [0, 1], then \(D_s\) is a bounded linear operator. Define the lower and upper Boyd indices of E by

$$\begin{aligned} p_E:=\lim \limits _{s\rightarrow \infty }\frac{\log s}{\log \Vert D_s\Vert } \ \ \ \ \ \text{ and } \ \ \ \ \ q_E:=\lim \limits _{s\rightarrow 0^{+}}\frac{\log s}{\log \Vert D_s\Vert }, \end{aligned}$$

respectively. It is well known that \(1\le p_E\le q_E\le \infty \) and E has non-trivial Boyd indices, whenever \(1<p_E\le q_E<\infty \). We shall need the following duality for Boyd indices:

$$\begin{aligned} \frac{1}{p_E}+\frac{1}{q_{E^{\times }}}=1,\ \ \frac{1}{q_{E}}+\frac{1}{p_{E^{\times }}}=1. \end{aligned}$$

Let E be a symmetric Banach space on [0, 1]. For \(0<r<\infty \), we define \(E^{(r)}\) and \(E_{(r)}\) by

$$\begin{aligned}&E^{(r)}:=\{x: |x|^r\in E\}, \ \ \Vert x\Vert _{E^{(r)}}:=\big \Vert |x|^r\big \Vert ^{\frac{1}{r}}_E, \\&E_{(r)}:=\{x: |x|^{\frac{1}{r}}\in E\}, \ \ \Vert x\Vert _{E_{(r)}}:=\big \Vert |x|^{\frac{1}{r}}\big \Vert ^{r}_E, \end{aligned}$$

respectively. It is clear from the definitions that \(E^{(r)}\), \(E_{(r)}\) are symmetric and

$$\begin{aligned}&p_{E_{(r)}}={\frac{1}{r}}p_E,\ \ q_{E_{(r)}}={\frac{1}{r}}q_E,\ \ p_{E^{(r)}}=rp_E, \ \ q_{E^{(r)}}=rq_E. \end{aligned}$$

Let \(E_i\) be a quasi Banach idea space on [0, 1], \(i=1,2.\) The pointwise product space of \(E_1\) and \(E_2\) is defined as

$$\begin{aligned} E_1\odot E_2=\{x: x=x_1x_2, x_i\in E_i, i=1,2\} \end{aligned}$$

with a functional \(\Vert x\Vert _{E_1\odot E_2}\) defined by

$$\begin{aligned} \Vert x\Vert _{E_1\odot E_2}=\inf \{\Vert x\Vert _{E_1}\Vert x\Vert _{E_2}: x=x_1x_2, x_i\in E_i, i=1,2\}. \end{aligned}$$

Note that if E and F are symmetric Banach spaces on [0, 1], then we have the following results (see Theorem 1 in [1]).

  1. (i)

    If \(0<p<\infty \), then \((E\odot F)^{(p)}=E^{(p)}\odot F^{(p)}.\)

  2. (ii)

    If \(1<p<\infty \), then \((E^{(p)})^{\times }=(E^\times )^{(p)}\odot L_{p'}[0,1].\)

Let \(\mathcal {M}\) be a semi-finite von Neumann algebra with a faithful normal semi-finite trace \(\tau \). The set of all \(\tau \)-measurable operators is denoted by \(L_0(\mathcal {M})\). For \(x\in L_0(\mathcal {M})\), define its generalized singular number by

$$\begin{aligned} \mu _t(x)=\inf \{\lambda>0:\tau (\chi _{(\lambda ,\infty )}(|x|))\le t\}, \ \ t>0. \end{aligned}$$

For a given symmetric Banach function space E on [0, 1], we define the corresponding noncommutative space by setting:

$$\begin{aligned} E(\mathcal {M},\tau )=\{x\in L_0(\mathcal {M}):\mu _t(x)\in E\}. \end{aligned}$$

Equipped with the norm \(\Vert x\Vert _{E(\mathcal {M},\tau )}:=\Vert \mu _t(x)\Vert _E\), the space \(E(\mathcal {M},\tau )\) is a Banach space and is referred to as the noncommutative symmetric Banach space associated with \((\mathcal {M},\tau )\) corresponding to the function space \((E, \Vert \cdot \Vert _E).\) Note that if \(1\le p<\infty \) and \(E=L_p([0,1])\), then \(E(\mathcal {M},\tau )=L_p(\mathcal {M},\tau )\) is the usual noncommutative \(L_p\)-space associated with \((\mathcal {M},\tau )\).

2.1 Noncommutative quasi-martingales

We first recall the general setup for noncommutative martingales. Let \(({\mathcal {M}}_n)_{n\ge 1}\) be an increasing sequence of von Neumann subalgebras of \(\mathcal {M}\) such that the union of the \({\mathcal {M}}_n\)’s is \(\hbox {weak}^*\)-dense in \(\mathcal {M}\). For every \(n\ge 1\), the restriction \(\tau |_{\mathcal {M}_n}\) of \(\tau \) to \(\mathcal {M}_n\) remains semi-finite, still denoted by \(\tau \), and we assume that there exists a trace preserving conditional expectation \({\mathcal {E}}_n\) from \(\mathcal {M}\) onto \(\mathcal {M}_n\). In this case, \((\mathcal {M}_n)_{n\ge 1}\) is called a filtration of \(\mathcal {M}\). Note that \({\mathcal {E}}_n\) extends to a contractive projection from \(L_p(\mathcal {M})\) onto \(L_p(\mathcal {M}_n)\) for all \(1\le p\le \infty \). A noncommutative \(E(\mathcal {M})\)-martingale with respect to \(({\mathcal {M}}_n)_{n\ge 1}\) is a sequence \(x=(x_n)_{n\ge 1}\) such that \(x_n\in E(\mathcal {M}_n)\) and \({\mathcal {E}}_{n}(x_{n+1})=x_n\ \ \text{ for } \text{ any }\ \ n\ge 1.\) Let \(\Vert x\Vert _{E(\mathcal {M})}=\sup _{n\ge 1}\Vert x_n\Vert _{E(\mathcal {M})}\). If \(\Vert x\Vert _{E(\mathcal {M})}<\infty \), then x is called a bounded \(E(\mathcal {M})\)-martingale. The martingale difference sequence \(dx=(dx_n)_{n\ge 1}\) of x is defined by \(dx_n=x_n-x_{n-1}\) for \(n\ge 1\). Here and in the following, we set \(x_0=0\) and \(\mathcal {E}_0=\mathcal {E}_1\) for the sake of convenience.

In this paper, we are concerned with the following quasi-martingales in noncommutative symmetric Banach spaces.

Definition 2.1

Let E be a symmetric Banach space on [0, 1] and \(1\le p\le \infty \). A noncommutative \({_p}E(\mathcal {M})\)-quasi-martingale with respect to \(({\mathcal {M}}_n)_{n\ge 1}\) is a sequence \(x=(x_n)_{n\ge 1}\) such that \(x_n\in E(\mathcal {M}_n)\) for \(n\ge 1\) and (with \(\mathcal {E}_0 = 0, x_0=0\))

$$\begin{aligned} \sum \limits _{n=1}^\infty \Vert {\mathcal {E}}_{n-1}(dx_n)\Vert _{E(\mathcal {M})}^p<\infty . \end{aligned}$$

Let \(y_n=\sum \limits _{k=1}^n(dx_k-{\mathcal {E}}_{k-1}(dx_k))\) for \(n\ge 1\). We set

$$\begin{aligned} \Vert x\Vert _{{_p}\widehat{E}(\mathcal {M})}:=\sup _n\Vert y_n\Vert _{E(\mathcal {M})} +(\sum \limits _{n=1}^\infty \Vert {\mathcal {E}}_{n-1}(dx_n)\Vert _{E(\mathcal {M})}^p)^{\frac{1}{p}}. \end{aligned}$$

If \(\Vert x\Vert _{{_p}\widehat{E}(\mathcal {M})}<\infty \), then x is called a bounded \({_p}E(\mathcal {M})\)-quasi-martingale. The quasi-martingale space \({_p}\widehat{E}(\mathcal {M})\) is defined as the space of all bounded \({_p}E(\mathcal {M})\)-quasi-martingales, equipped with the norm \(\parallel \cdot \parallel _{{_p}\widehat{E}(\mathcal {M})}\). We remark that if \(1\le q\le \infty \) and \(E=L_q([0,1])\) then \({_p}\widehat{E}(\mathcal {M})={_p}\widehat{L_q}(\mathcal {M})\), where \({_p}\widehat{L_q}(\mathcal {M})\) consists of \(x=(x_n)_{n\ge 1}\subset L_q(\mathcal {M})\) for which

$$\begin{aligned} \Vert x\Vert _{{_p}\widehat{L_q}(\mathcal {M})}=\sup _n\Vert y_n\Vert _{L_q(\mathcal {M})} +(\sum \limits _{n=1}^\infty \Vert {\mathcal {E}}_{n-1}(dx_n)\Vert ^p_{L_q(\mathcal {M})})^{\frac{1}{p}}. \end{aligned}$$

Now we define the noncommutative space \({_p}G_E(\mathcal {M})\) which is used in the proof of our main results.

Definition 2.2

Let E be a symmetric Banach space on [0, 1] and \(1\le p\le \infty \). The noncommutative space \({_p}G_E(\mathcal {M})\) is defined as the subspace of \(l_p(E(\mathcal {M}))\) consisting of all sequences \(dx=(dx_n)_{n\ge 1}\) such that \(x=(x_n)_{n\ge 1}\) is a predictable \({_p}E(\mathcal {M})\)-quasi-martingale with \(x_1=0\), and is equipped with the norm

$$\begin{aligned} \Vert x\Vert _{{_p}G_E(\mathcal {M})}=(\sum \limits _{n=1}^\infty \Vert dx_n\Vert ^p_{E(\mathcal {M})})^{\frac{1}{p}}. \end{aligned}$$

Note that if \(1\le q\le \infty \) and \(E=L_q([0,1])\) then \({_p}G_E(\mathcal {M})={_p}G_q(\mathcal {M})\), where \({_p}G_q(\mathcal {M})\) denotes the space of \(x=(x_n)_{n\ge 1}\subset L_q(\mathcal {M})\) for which

$$\begin{aligned} \Vert dx\Vert _{{_p}G_q(\mathcal {M})}=(\sum \limits _{n=1}^\infty \Vert dx_n\Vert ^p_{L_q(\mathcal {M})})^{\frac{1}{p}}. \end{aligned}$$

The following theorem plays an important role in our paper which we call Doob’s decomposition.

Theorem 2.3

(Doob’s decomposition) Let E be a symmetric Banach space on [0, 1] and \(1\le p\le \infty \). Then each bounded \({_p}E(\mathcal {M})\)-quasi-martingale \(x=(x_n)_{n\ge 1}\) can be uniquely decomposed as a sum of two sequences \(y=(y_n)_{n\ge 1}\) and \(z=(z_n)_{n\ge 1}\), where \(y=(y_n)_{n\ge 1}\) is a bounded \(E(\mathcal {M})\)-martingale and \(z=(z_n)_{n\ge 1}\) is a predicable \({_p}E(\mathcal {M})\)-quasi-martingale with \(z_1=0\).

Proof

The proof is similar with Lemma 2.2 in [15].

2.2 Interpolations

For a compatible Banach couple \((X_0,X_1)\), we define the K-functional by setting for any \(x\in X_0+X_1\) and \(t>0\),

$$\begin{aligned} K_t(x;X_0,X_1) = \inf \{\Vert x_0\Vert _{X_0}+t\Vert x_1\Vert _{ X_1}:x=x_0+x_1, x_0\in X_0,x_1\in X_1\}. \end{aligned}$$

The interpolation space \((X_0,X_1)_{E,K}\) is defined as the space of all elements \(x\in X_0+X_1\) such that

$$\begin{aligned} \Vert x\Vert _{(X_0,X_1)_{E,K}}:=\left\| \frac{K_t(x;X_0,X_1)}{t}\right\| _E<\infty . \end{aligned}$$

We may state the following interpolation result which is needed in the sequel (see Theorem 2.2 in [19] and [20]).

Theorem 2.4

Let E be a symmetric Banach space on [0, 1] with the Fatou property and \(1<p<p_E\le q_E<q<\infty .\) Then there exists a symmetric Banach space F with nontrivial Boyd indices such that

$$\begin{aligned} E(\mathcal {M})=(L_p(\mathcal {M}), L_q(\mathcal {M}))_{F,K}\ \ \text{(with } \text{ equivalent } \text{ norms) }. \end{aligned}$$
(2.1)

Proof

By Theorem 2.2 in [19], there is a symmetric Banach function space F on [0, 1] such that \(f\in E\) if and only if \(t\rightarrow K_t(f;L_p[0,1],L_q[0,1]) \in F\) and there exist a constant C such that

$$\begin{aligned} C^{-1}\Vert t\rightarrow K_t(f;L_p[0,1],L_q[0,1])\Vert _F\le \Vert f\Vert _E\le C\Vert t\rightarrow K_t(f;L_p[0,1],L_q[0,1]). \end{aligned}$$
(2.2)

For any \(x\in E(\mathcal {M})\), using the results \(K_t(\mu (x);L_p[0,1],L_q[0,1])\approx K_t(x;L_p(\mathcal {M}),L_p(\mathcal {M})\) and \(\Vert \mu (x)\Vert _E=\Vert x\Vert _{E(\mathcal {M})}\), we can extend (2.2) to the noncommutative setting. The proof is completed.

Throughout the paper \(p'\) will denote the conjugate index of p. \(\square \)

3 Main results

Our first result in this section is concerned with the dual space of \(_p\widehat{E}(\mathcal {M})\) which is the symmetric Banach space of noncommutative quasi-martingales.

Theorem 3.1

Let E be a symmetric Banach space on [0, 1] with the Fatou property and \(1<p_E\le q_E<p\). Then

$$\begin{aligned} \big ({_p}\widehat{E}(\mathcal {M})\big )^*={_{p'}}\widehat{E}^{\times }(\mathcal {M}) \end{aligned}$$

isometrically, with associated duality bracket given by

$$\begin{aligned} \forall x\in {_p}\widehat{E}(\mathcal {M}), \ \ \forall u\in {_{p'}}\widehat{E}^{\times }(\mathcal {M}),\ \ (x,u)=\tau (\nu y)+\sum \limits _{n=1}^\infty \tau (d\omega _ndz_n), \end{aligned}$$

where \(\mu _n=\nu _n+\omega _n\) and \(x_n=y_n+z_n(n\ge 1)\) are the Doob’s decomposition of u and x respectively.

For the proof we need the following Lemma.

Lemma 3.2

Let E be a symmetric Banach space on [0, 1] with the Fatou property and \(1<p_E\le q_E<p\). Then

$$\begin{aligned} E=F\odot {E^{\times }}^{(\frac{p}{p'})}, \end{aligned}$$

where \(F=(E^{\times (\frac{1}{p'})})^{\times }\) is separable.

Proof

From the proof of Lemma 2.1 in [1], we know that \(E^{\times (\frac{1}{p'})}\) is reflexive and F is separable. By (ii) of the properties of pointwise product spaces, we have

$$\begin{aligned} E=(E^\times )^\times =([E^{\times (\frac{1}{p'})}]^{(p')})^\times = ([E^{\times (\frac{1}{p'})}]^\times )^{(p')}\odot L_{p}[0,1]=F^{(p')}\odot L_{p}[0,1]. \end{aligned}$$

Using the equality \(L_{1}[0,1]=E\odot E^\times \) (see Theorem 1.2 in [1]) and (i) of the properties of pointwise product spaces, we obtain that

$$\begin{aligned} E=F^{(p')}\odot (F\odot E^{\times (\frac{1}{p'})})^{(p)} =F^{(p')}\odot (F^{(p)}\odot E^{\times (\frac{1}{p'})(p)}) =F\odot E^{\times (\frac{p}{p'})} . \end{aligned}$$

The proof is completed. \(\square \)

We also require the following duality result (see Theorem 5.6 in [6]).

Lemma 3.3

Let E be a symmetric Banach space on [0, 1] with the Fatou property, then \(\big (E(\mathcal {M})\big )^*=E^\times (\mathcal {M})\) isometrically, with associated duality bracket given by

$$\begin{aligned} (x,y)=\tau (xy), \ \ x\in E(\mathcal {M}), y\in E^\times (\mathcal {M}). \end{aligned}$$

Now, we concern the dual space of \(l_p(E(\mathcal {M}))\) which is the main ingredient in the proof of Theorem 3.1.

Lemma 3.4

Let E be a symmetric Banach space on [0, 1] with the Fatou property and \(1<p_E\le q_E<p\). Then

$$\begin{aligned} \big (l_p({E}(\mathcal {M}) )\big )^*=l_{p'}({E}^{\times }(\mathcal {M})) \end{aligned}$$

with equivalent norms.

Proof

Let \(x=(x_n)_{n\ge 1}\in l_p({E}(\mathcal {M}))\) and \(y=(y_n)_{n\ge 1}\in l_{p'}({E}^{\times }(\mathcal {M}))\). Now, we define a linear functional on \(l_p({E}(\mathcal {M}))\) by

$$\begin{aligned} l_y(x)=\sum \limits _{n=1}^\infty \tau (x_ny_n). \end{aligned}$$

Then by Lemma 3.3 and Hölder’s inequality,

$$\begin{aligned} \left| \sum \limits _{n=1}^\infty \tau (x_ny_n)\right|\le & {} \sum \limits _{n=1}^\infty |\tau (x_ny_n)|\\\le & {} \sum \limits _{n=1}^\infty \Vert x_n\Vert _{{E}(\mathcal {M})}\Vert y_n\Vert _{{E}^{\times }(\mathcal {M})}\\\le & {} (\sum \limits _{n=1}^\infty \Vert x_n\Vert ^p_{{E}(\mathcal {M})})^{\frac{1}{p}} (\sum \limits _{n=1}^\infty \Vert y_n\Vert ^{p'}_{{E}^{\times }(\mathcal {M})})^{\frac{1}{{p'}}}\\= & {} \Vert x\Vert _{ l_p({E}(\mathcal {M}))}\Vert y\Vert _{l_{p'}({E}^{\times }(\mathcal {M}))}. \end{aligned}$$

Thus the series \(\sum \limits _{n=1}^\infty \tau (x_ny_n)\) converges absolutely. Therefore, \(l_y(x)\) is continuous on \(l_p({E}(\mathcal {M}))\) and \(\Vert l_y\Vert \le \Vert y\Vert _{l_{p'}({E}^{\times }(\mathcal {M}))}.\)

We pass to the converse inclusion. Let \(l \in \big (l_p({E}(\mathcal {M}) )\big )^*\) of norm one. For every \(n\ge 1\), set

$$\begin{aligned} l_n(x_n)=l(\theta ),\ \ x_n\in E(\mathcal {M}), \end{aligned}$$

where \(\theta =(\mathop {\underbrace{{0,\ldots ,0,x_n,}}}\limits _n0,\ldots )\).Then

$$\begin{aligned} |l_n(x_n)|=|l(\theta )|\le \Vert l\Vert \Vert \theta \Vert _{l_p({E}(\mathcal {M}))}=\Vert x_n\Vert _{E(\mathcal {M})}. \end{aligned}$$

This implies that \(l_n\in (E(\mathcal {M}))^*\). Since \(\big ({E}(\mathcal {M})\big )^*={E}^{\times }(\mathcal {M})\), the representation theorem allows us to find an element \(y_n\in {E}^{\times }(\mathcal {M})\) such that

$$\begin{aligned} l_n(x_n)=\tau (x_ny_n),\ \ x_n\in E(\mathcal {M}). \end{aligned}$$

Thus we have that

$$\begin{aligned} l(x)=\sum \limits _{n=1}^\infty l_n(x_n)=\sum \limits _{n=1}^\infty \tau (x_ny_n) \end{aligned}$$
(3.1)

for any finite sequence \(x=(x_n)_{n\ge 1}\in l_p({E}(\mathcal {M}))\). We must show that \(y=(y_n)_{n\ge 1}\in l_{p'}(E^{\times }(\mathcal {M}))\) and is of norm \(\le 1\). Now, fix an n. Note that for any \(k\le n\)

$$\begin{aligned} \Vert y_k\Vert ^{p'}_{{E}^{\times }(\mathcal {M})} =\Vert (|y_k|^{p'})^{\frac{1}{{p'}}}\Vert ^{p'}_{E^{\times }(\mathcal {M})} =\Vert |y_k|^{p'}\Vert _{E^{\times (\frac{1}{{p'}})}(\mathcal {M})} =\sup \{\tau (a_k|y_k|^{p'}):a_k\in F(\mathcal {M}),\Vert a_k\Vert _{F(\mathcal {M})}\le 1\}, \end{aligned}$$

where \(\big (F(\mathcal {M})\big )^*={E}^{\times (\frac{1}{{p'}})}(\mathcal {M}).\) Thus for an arbitrarily given \(\varepsilon >0\), there exists \(a_k^\varepsilon \in F(\mathcal {M})\) and \(\Vert a_k^\varepsilon \Vert _{F(\mathcal {M})}\le 1\) such that

$$\begin{aligned} \Vert y_k\Vert ^{p'}_{{E}^{\times }(\mathcal {M})}\le \tau (a_k^\varepsilon |y_k|^{p'})+\frac{\varepsilon }{2^k}. \end{aligned}$$
(3.2)

Set \(z_k=\frac{1}{\gamma _n}a_k^\varepsilon |y_k|^{{p'}-2}y^*,\) where \(\gamma _n=(\sum \limits _{k=1}^n\Vert y_k\Vert _{E^{\times }(\mathcal {M})}^{p'})^{\frac{1}{{p}}}.\) Then noting that \(|y_k|^{{p'}-2}y^*\in E^{\times (\frac{p}{p'})}\) and by Lemma 3.2, we get \(z_k\in E(\mathcal {M})\) and

$$\begin{aligned} \Vert z_k\Vert _{E(\mathcal {M})}\le \frac{1}{\gamma _n}\Vert a_k^\varepsilon \Vert _{F(\mathcal {M})}\Vert |y_k|^{{p'}-2}y^*\Vert _{E^{\times (\frac{p}{p'})}}\le \frac{1}{\gamma _n}\Vert y_k\Vert ^{{p'}-1}_{E^{\times }(\mathcal {M})}. \end{aligned}$$
(3.3)

Thus we have that

$$\begin{aligned} \left( \sum \limits _{k=1}^n\Vert z_k\Vert _{E(\mathcal {M})}^p\right) ^{\frac{1}{p}} \le \frac{1}{\gamma _n}(\sum \limits _{k=1}^n\Vert y_k\Vert _{E^\times (\mathcal {M})}^{((p'-1)p)})^{\frac{1}{p}} =\frac{1}{\gamma _n}\left( \sum \limits _{k=1}^n\Vert y_k\Vert _{E^{\times }(\mathcal {M})}^{p'}\right) ^{\frac{1}{{p}}}=1. \end{aligned}$$

Let \(z^{(n)}=(z_1,\ldots ,z_n,0,\ldots ).\) Then \(z^{(n)}\in l_p({E}(\mathcal {M}))\) and \(\Vert z^{(n)}\Vert _{l_p ({E}(\mathcal {M}))}\le 1.\) Using (3.1) and (3.2), we obtain that

$$\begin{aligned} l(z^{(n)})= & {} \sum \limits _{k=1}^n\tau (z_ky_k)\\= & {} \frac{1}{\gamma _n}\sum \limits _{k=1}^n\tau (a_k^\varepsilon |y_k|^{p'})\\\ge & {} \frac{1}{\gamma _n}\sum \limits _{k=1}^n(\Vert y_k\Vert ^{p'}_{{E}^{\times }(\mathcal {M})}-\frac{\varepsilon }{2^k})\\\ge & {} (\sum \limits _{k=1}^n\Vert y_k\Vert ^{p'}_{{E}^{\times }(\mathcal {M})})^{\frac{1}{{p'}}}- \frac{1}{\gamma _n}\varepsilon . \end{aligned}$$

Thus we have that

$$\begin{aligned} \left( \sum \limits _{k=1}^n\Vert y_k\Vert _{E^{\times }(\mathcal {M})}^{p'})^{\frac{1}{{p'}}}\le \frac{1}{\gamma _n}\varepsilon +l(z^{(n)}\right) \le 1. \end{aligned}$$

It follows that

$$\begin{aligned} \left( \sum \limits _{n=1}^\infty \Vert y_n\Vert _{E^{\times }(\mathcal {M})}^{p'}\right) ^{\frac{1}{{p'}}}\le 1 \end{aligned}$$

as \(n\rightarrow \infty \) which implies

$$\begin{aligned} y\in l_{p'}(E^{\times }(\mathcal {M}))\ \ \text{ and } \ \ \ \Vert y\Vert _{l_{p'}(E^{\times }(\mathcal {M}))}\le 1. \end{aligned}$$

For any \(x=(x_n)_{n\ge 1}\in l_p({E}(\mathcal {M}))\), let \(x^{(n)}=(x_1,\ldots ,x_n,0,\ldots )\) \((n\ge 1)\). Then

$$\begin{aligned} \Vert x-x^{(n)}\Vert _{l_p({E}(\mathcal {M}))}\rightarrow 0\ \ (n\rightarrow \infty ). \end{aligned}$$

Using (3.1), we have

$$\begin{aligned} l(x)=\lim \limits _{n\rightarrow \infty }l(x^{(n)})=\lim \limits _{n\rightarrow \infty }\sum \limits _{i=1}^n\tau (x_iy_i) =\sum \limits _{i=1}^\infty \tau (x_iy_i). \end{aligned}$$

The proof is completed. \(\square \)

The proof of Theorem 3.1

Let \(\mu =(\mu _n)_{n\ge 1}\in {_{p'}}\widehat{E}^{\times }(\mathcal {M})\) and \( x=(x_n)_{n\ge 1}\in {_p}\widehat{E}(\mathcal {M})\). Let \(\mu _n=\nu _n+\omega _n\) and \(x_n=y_n+z_n(n\ge 1)\) be the Doob’s decomposition of \(\mu \) and x respectively. Then \(y=(y_n)_{n\ge 1}\) is a bounded \(E(\mathcal {M})\)-martingale and \(\nu = (\nu _n)_{n\ge 1}\) is a bounded \(E^{\times }(\mathcal {M})\)-martingale. Thus there exist \(y_\infty \in E(\mathcal {M})\) and \(\nu _\infty \in E^{\times }(\mathcal {M})\) such that \(y_n\overset{E(\mathcal {M})}{\longrightarrow }y_\infty ,\ \ \nu _n\overset{E^{\times }(\mathcal {M})}{\longrightarrow }\nu _\infty .\)

Now we define a linear functional on \(_{p}\widehat{{E}}(\mathcal {M})\) by

$$\begin{aligned} l_\mu (x)=\tau (\nu _\infty y_\infty )+\sum \limits _{n=1}^\infty \tau (d\omega _ndz_n). \end{aligned}$$

Then by Lemma 3.3 and Hölder’s inequality,

$$\begin{aligned} |l_\mu (x)|\le & {} \Vert \nu _{\infty }\Vert _{E^{\times }(\mathcal {M})}\Vert y_{\infty }\Vert _{E(\mathcal {M})}+ \sum \limits _{n=1}^\infty \Vert d\omega _n\Vert _{E^{\times }(\mathcal {M})}\Vert dz_n\Vert _{E(\mathcal {M})}\\\le & {} \Vert \nu _{\infty }\Vert _{E^{\times }(\mathcal {M})}\Vert y_{\infty }\Vert _{E(\mathcal {M})}+ \left( \sum \limits _{n=1}^\infty \Vert d\omega _n\Vert ^{p'}_{E^{\times }(\mathcal {M})}\right) ^{\frac{1}{{p'}}} \left( \sum \limits _{n=1}^\infty \Vert dz_n\Vert ^{p}_{E^(\mathcal {M})}\right) ^{\frac{1}{p}}\\\le & {} \Bigg (\Vert \nu _\infty \Vert _{E^{\times }(\mathcal {M})}+\left( \sum \limits _{n=1}^\infty \Vert d\omega _n\Vert ^{p'}_{E^{\times }(\mathcal {M})}\right) ^{\frac{1}{{p'}}}\Bigg ) \big (\Vert y_\infty \Vert _{E(\mathcal {M})}+\left( \sum \limits _{n=1}^\infty \Vert dz_n\Vert ^p_{E^(\mathcal {M})}\right) ^{\frac{1}{p}}\big )\\= & {} \Vert \mu \Vert _{{_{p'}\widehat{E}^{\times }(\mathcal {M})}}\Vert x\Vert _{_{p}\widehat{E}(\mathcal {M})} . \end{aligned}$$

Thus \(l_\mu (x)\) is continuous on \(_{p}\widehat{E}(\mathcal {M})\) and \(\Vert l_\mu \Vert \le \Vert \mu \Vert _{{_{p'}\widehat{E}^{\times }(\mathcal {M})}}.\)

We pass to the converse inclusion. Let \(l\in \big (_{p}\widehat{E}(\mathcal {M})\big )^{*}\) of norm one. Let \(l_1\) be the restriction of l on \(E(\mathcal {M})\). Noting that \(\big ({E}(\mathcal {M})\big )^*={E}^{\times }(\mathcal {M})\), there exists \(\nu \in E^{\times }(\mathcal {M})\) and \(\Vert \nu \Vert _{E^{\times }(\mathcal {M})}\le 1\) such that

$$\begin{aligned} l_1(a)=\tau (av),\ \ a\in E(\mathcal {M}). \end{aligned}$$
(3.4)

On the other hand, define a functional on \(_{p}G_E(\mathcal {M})\) by

$$\begin{aligned} l_2(db)=l(b),\ \ db=(db_n)_{n\ge 1}\in {_{p}}G_E(\mathcal {M}). \end{aligned}$$

Then \(|l_2(db)|\le \Vert l\Vert \Vert b\Vert _{_{p}\widehat{E}(\mathcal {M})}=\Vert db\Vert _{_{p}G_E(\mathcal {M})}.\) Thus we have that \(l_2\) is a continuous linear functional on \(_{p}G_E(\mathcal {M})\) and \(\Vert l_2\Vert \le 1\). Recall that \(_{p}G_E(\mathcal {M})\) is the closed subspace of \(l_p(E(\mathcal {M}))\). By the Hahn-Banach theorem, \(l_2\) extends to a norm one functional \(\widetilde{l}_2\) on \(l_p(E(\mathcal {M}))\). Consequently, by Lemma 3.4, \(\widetilde{l}_2\) is given by a norm one element \(\omega '=(\omega _n')_{n\ge 1}\) of \(l_{p'}(E^\times (\mathcal {M}))\). Thus

$$\begin{aligned} l_2(db)=\sum \limits _{n=1}^\infty \tau (d\omega _n'db_n)\ \ db=(db_n)_{n\ge 1}\in {_{p}G_E}(\mathcal {M}). \end{aligned}$$
(3.5)

Set \(\omega _1=0\) and \(\omega _n=\sum \limits _{k=1}^{n}{\mathcal {E}}_{k-1}(\omega _k') (n\ge 2)\). For any \(db=(db_n)_{n\ge 1}\in {_p}G_E(\mathcal {M})\), noting that \(db=(db_n)_{n\ge 1}\) is predicable, it follows from (3.5) that

$$\begin{aligned} l_2(db)=\sum \limits _{n=1}^\infty \tau ({\mathcal {E}}_{n-1}(\omega _n'db_n)) =\sum \limits _{n=1}^\infty \tau (db_n{\mathcal {E}}_{n-1}(\omega _n')) =\sum \limits _{n=1}^\infty \tau (d\omega _ndb_n). \end{aligned}$$
(3.6)

It is easy to see that \(\omega =(\omega _n)_{n\ge 1}\) is predicable with \(\omega _1=0\) and

$$\begin{aligned} \left( \sum \limits _{n=1}^\infty \Vert d\omega _n\Vert ^{p'}_{E^{\times }(\mathcal {M})}\right) ^{\frac{1}{{p'}}} =\left( \sum \limits _{n=1}^\infty \Vert \mathcal {E}_{n-1}(\omega _n')\Vert ^{p'}_{E^{\times }(\mathcal {M})}\right) ^{\frac{1}{{p'}}}\le \left( \sum \limits _ {n=1}^\infty \Vert \omega _n'\Vert ^{p'}_{E^{\times }(\mathcal {M})}\right) ^{\frac{1}{{p'}}}=1. \end{aligned}$$

Set \(\mu _n=\nu _n+\omega _n(n\ge 1)\), where \(\nu _n=\mathcal {E}_n(\nu )(n\ge 1)\). Then \(\mu =(\mu _n)_{n\ge 1}\in {_{p'}}\widehat{E}^{\times }(\mathcal {M})\) and

$$\begin{aligned} \Vert \mu \Vert _{_{p'}\widehat{E}^{\times }(\mathcal {M})} =\Vert \nu \Vert _{E^{\times }(\mathcal {M})}+ \left( \sum \limits _{n=1}^\infty \Vert d\omega _n\Vert ^{p'}_{E^{\times }(\mathcal {M})}\right) ^{\frac{1}{{p'}}}\le 2. \end{aligned}$$

For any \(x=(x_n)_{n\ge 1}\in {_p}\widehat{{E}}(\mathcal {M})\), let \(x_n=y_n+z_n(n\ge 1)\) be its Doob’s decomposition. Noting that \(y=(y_n)_{n\ge 1}\) is a bounded \(E(\mathcal {M})\)-martingale and \(dz=(dz_n)_{n\ge 1}\in {_p}G_E(\mathcal {M})\), it follows from (3.4) and (3.6) that

$$\begin{aligned} l(x)=l_1(y)+l_2(dz)=\tau (y_{\infty }\upsilon _{\infty })+\sum \limits _{n=1}^\infty \tau (d\omega _ndz_n), \end{aligned}$$

where \(y_\infty \) is the limit of \((y_n)_{n\ge 1}\) in \(E(\mathcal {M})\). The proof is completed. \(\square \)

As applications of Theorem 3.1, we shall consider the symmetric space \({_p}\widehat{E}(\mathcal {M})\) as interpolations of noncommutative quasi-martingale \(L_p\)-spaces, which is a generalization of Theorem 2.4.

Theorem 3.5

Let E be a symmetric Banach space on [0, 1] with the Fatou property and \(1<p<p_E\le q_E<q<\infty .\) Then there exists a symmetric Banach space F with nontrivial Boyd indices such that

$$\begin{aligned} _s\widehat{E}(\mathcal {M})=(_s\widehat{L}_p(\mathcal {M}), {_s\widehat{L}}_q(\mathcal {M}))_{F,K}, \end{aligned}$$

where \(1<s<p_E.\)

For the proof of Theorem 3.5, we need the following lemmas (see [2]).

Lemma 3.6

Let E be a symmetric Banach space on [0, 1] with the Fatou property, and let \((X_1, X_2)\) be a compatible Banach couple. Then

$$\begin{aligned} (X_1, X_2)^*_{F,K}=(X_2^*, X_1^*)_{F^\times ,K}. \end{aligned}$$

Lemma 3.7

Let E be a symmetric Banach space on [0, 1] with the Fatou property and \(1<p<p_E\le q_E<q<\infty .\) Then there exists a symmetric Banach space F with nontrivial Boyd indices such that

$$\begin{aligned} l_s({E}(\mathcal {M}))=\big (l_s({L}_p(\mathcal {M})), l_s({L}_q(\mathcal {M}))\big )_{F,K}, \end{aligned}$$

where \(1<s<p_E.\)

Proof

Let \(x=(x_n)_{n\ge 1}\in l_s(E(\mathcal {M}))\). Then by Theorem 2.4, there exists a symmetric Banach space F with nontrivial Boyd indices such that \(x_n\in ({L}_p(\mathcal {M}), {L}_q(\mathcal {M}))_{F,K}\) for any \(n\ge 1\). For any \(a,b>0\), \(\alpha _s(a^s+b^s)\le (a+b)^s\le \beta _s (a^s+b^s)\) for some constants \(\alpha _s,\beta _s\) depending only on s. Using this fact, it is easy to show that

$$\begin{aligned}&\big ( K_t(x;l_s(L_{p}(\mathcal {M})),l_s(L_{q}(\mathcal {M})))\big )^s\\&\quad \le \beta _s\sum \limits _{n=1}^\infty \inf \limits _{x_n=x_n^0+x_n^1}(\Vert x_n^0\Vert ^s_{L_{p}(\mathcal {M})}+t^s\Vert x_n^1\Vert ^s_{L_{q}(\mathcal {M})})\\&\quad \le \frac{\beta _s}{\alpha _s}\sum \limits _{n=1}^\infty \big ( K_t(x_n;L_{p}(\mathcal {M}),L_{q}(\mathcal {M}))\big )^s. \end{aligned}$$

Noting that\(F_{(s)}\) is a quasi-Banach space, we have that

$$\begin{aligned}&\Vert x\Vert ^s_{(l_s(L_{p}(\mathcal {M})),l_s(L_{q}(\mathcal {M})))_{F,K}}\\&\quad = \left\| \frac{ K_t(x;l_s(L_{p}(\mathcal {M})),l_s(L_{q}(\mathcal {M})))^s}{t^s}\right\| _{F_{(s)}}\\&\quad \le C_s \sum \limits _{n=1}^\infty \left\| \frac{ K_t(x_n;L_{p}(\mathcal {M}),L_{q}(\mathcal {M}))^s}{t^s}\right\| _{F_{(s)}}\\&\quad = C_s \sum \limits _{n=1}^\infty \left\| \frac{ K_t(x_n;L_{p}(\mathcal {M}),L_{q}(\mathcal {M}))}{t}\right\| ^s_{F}, \end{aligned}$$

where \(C_s\) is a constant depending on s. This means that

$$\begin{aligned} \Vert x\Vert _{\big (l_s({L}_p(\mathcal {M})), l_s({L}_q(\mathcal {M}))\big )_{F,K}}\le C_s\Vert x\Vert _{l_s({E}(\mathcal {M}))} \end{aligned}$$

and

$$\begin{aligned} l_s({E}(\mathcal {M}))\subset \big (l_s({L}_p(\mathcal {M})), l_s({L}_q(\mathcal {M}))\big )_{F,K}. \end{aligned}$$
(3.7)

Similarly, we have that \(l_{s'}({E^\times }(\mathcal {M}))\subset \big (l_{s'}({L}_{q'}(\mathcal {M})), l_{s'}({L}_{p'}(\mathcal {M}))\big )_{F^\times ,K}. \) It follows that

$$\begin{aligned} (l_{s'}({E^\times }(\mathcal {M})))^*\supset \big ((l_{s'}({L}_{q'}(\mathcal {M})), l_{s'}({L}_{p'}(\mathcal {M})))\big )^*_{F^\times ,K}. \end{aligned}$$

Observe that \(p_{E^\times }\le q_{E^\times }\le s'\). Thus by Lemma  3.4 and Lemma 3.6, we have that

$$\begin{aligned} l_s({E}(\mathcal {M}))\supset \big (l_s({L}_p(\mathcal {M})), l_s({L}_q(\mathcal {M}))\big )_{F,K}. \end{aligned}$$
(3.8)

Putting (3.7) and (3.8) together, we obtain that

$$\begin{aligned} l_s({E}(\mathcal {M}))=\big (l_s({L}_p(\mathcal {M})), l_s({L}_q(\mathcal {M}))\big )_{F,K}. \end{aligned}$$

The proof is completed.

The following is an interpolation result on the space \(_pG_E(\mathcal {M})\).

Lemma 3.8

Let E be a symmetric Banach space on [0, 1] with the Fatou property and \(1<p<p_E\le q_E<q<\infty .\) Then there exists a symmetric Banach space F with nontrivial Boyd indices such that

$$\begin{aligned} _sG_E(\mathcal {M})=\big ({_sG_p}(\mathcal {M}), {_sG_q}(\mathcal {M})\big )_{F,K}, \end{aligned}$$

where \(1<s<p_E.\)

Proof

Note that \({_sG_p}(\mathcal {M})\) consists of quasi-martingale difference sequences in \(l_s(L_p(\mathcal {M}))\). So \({_sG_p}(\mathcal {M})\) is 1-complemented in \(l_s(L_p(\mathcal {M}))\) via the projection

$$\begin{aligned} P: {\left\{ \begin{array}{ll} \ l_s(L_p(\mathcal {M}))\rightarrow {_sG_p}(\mathcal {M});\\ \ (a_n)_{n\ge 1} \rightarrow (\mathcal {E}_{n-1}(a_n))_{n\ge 1}. \end{array}\right. } \end{aligned}$$

It follows that for any \(x\in \big ({_sG_p}(\mathcal {M}), {_sG_q}(\mathcal {M})\big )_{F,K}\),

$$\begin{aligned} K_t(x;{_sG_p}(\mathcal {M}),{_sG_q}(\mathcal {M}))= K_t(x;l_s(L_{p}(\mathcal {M})),l_s(L_{q}(\mathcal {M})))\ \ t>0. \end{aligned}$$

Thus

$$\begin{aligned} \Vert x\Vert _{\big ({_sG_p}(\mathcal {M}), {_sG_q}(\mathcal {M})\big )_{F,K}}=\Vert x\Vert _{\big (l_s(L_{p}(\mathcal {M})),l_s(L_{q}(\mathcal {M}))\big )_{F,K}}. \end{aligned}$$

Therefore, using Lemma 3.6, we have finished the proof of the theorem.

Proof of Theorem 3.5

Let \(x\in ({_s\widehat{L}_p}(\mathcal {M}), {_s\widehat{L}}_q(\mathcal {M}))_{F,K}\) and \(x=x^0+x^1\) be a decomposition of x where \( x^0\in {_s\widehat{L}_p}(\mathcal {M})\) and \(x^1\in {_s\widehat{L}}_q(\mathcal {M}). \) Let \(x_n^k=y_n^k+z_n^k \ (n\ge 1) \) be the Doob’s decomposition of \(x^k \ (k=0,1)\). Then we have that \(y^0\in L_p(\mathcal {M})\), \(y^1\in L_q(\mathcal {M})\) and \(dz^0\in {_sG_p}(\mathcal {M})\), \(dz^1\in {_sG_q}(\mathcal {M})\). Set \(y=y^0+y^1\) and \(z=z^0+z^1\). Then

$$\begin{aligned}&K_t(y;L_{p}(\mathcal {M}),L_{q}(\mathcal {M}))+K_t(dz; {_sG_p}(\mathcal {M}), {_sG_q}(\mathcal {M}))\\&\quad \le \Vert y^0\Vert _{L_{p}(\mathcal {M})}+t\Vert y^1\Vert _{L_{q}(\mathcal {M})} +\Vert dz^0\Vert _{{_sG_p}(\mathcal {M}))}+t\Vert dz^1\Vert _{{_sG_q}(\mathcal {M})}\\&\quad = \Vert x^0\Vert _{_s\widehat{L}_{p}(\mathcal {M})}+t\Vert x^1\Vert _{_s\widehat{L}_{q}(\mathcal {M})}. \end{aligned}$$

Thus we get that

$$\begin{aligned} K_t(y;L_{p}(\mathcal {M}),L_{q}(\mathcal {M}))+K_t(dz; {_sG_p}(\mathcal {M}), {_sG_q}(\mathcal {M})) \le K_t(x;{_s\widehat{L}}_p(\mathcal {M}),{_s\widehat{L}}_q(\mathcal {M})), \end{aligned}$$

where the infimum runs over all decomposition of x. Using the equality \(\Vert x\Vert _{(X_0,X_1)_{F,K}}=\Vert \frac{K_t(x;X_0,X_1)}{t}\Vert _F, \) we have that

$$\begin{aligned} \Vert y\Vert _{(L_{p}(\mathcal {M}),L_{q}(\mathcal {M}))_{F,K}}+\Vert dz\Vert _{({_sG_p}(\mathcal {M}), {_sG_q}(\mathcal {M}))_{F,K}} \le 2\Vert x\Vert _{({_s\widehat{L}}_{p}(\mathcal {M}),{_s\widehat{L}}_{q}(\mathcal {M}))_{F,K}}. \end{aligned}$$

By Lemma 2.4 and Lemma 3.8, we get that

$$\begin{aligned} \Vert y\Vert _{E(\mathcal {M})}+\Vert dz\Vert _{{_sG_E}(\mathcal {M})} \le 2\Vert x\Vert _{({_s\widehat{L}}_{p}(\mathcal {M}),{_s\widehat{L}}_{q}(\mathcal {M}))_{F,K}} \end{aligned}$$

which implies that \(\Vert x\Vert _{_s\widehat{E}(\mathcal {M})} \le 2\Vert x\Vert _{({_s\widehat{L}}_{p}(\mathcal {M}),{_s\widehat{L}}_{q}(\mathcal {M}))_{F,K}}\) and

$$\begin{aligned} _s\widehat{E}(\mathcal {M})\supset ({_s\widehat{L}}_{p}(\mathcal {M}),{_s\widehat{L}}_{q}(\mathcal {M}))_{F,K}. \end{aligned}$$

By Theorem 3.1 and Lemma 3.6, we have that

$$\begin{aligned} _s\widehat{E}(\mathcal {M})= (_{s'}\widehat{E^\times }(\mathcal {M}))^*\subset \big (({_{s'}\widehat{L}}_{q'}(\mathcal {M}),{_{s'}\widehat{L}}_{p'}(\mathcal {M}))_{F^\times ,K}\big )^* =({_s\widehat{L}}_{p}(\mathcal {M}),{_s\widehat{L}}_{q}(\mathcal {M}))_{F,K}. \end{aligned}$$

Therefore,

$$\begin{aligned} _s\widehat{E}(\mathcal {M})= ({_s\widehat{L}}_{p}(\mathcal {M}),{_s\widehat{L}}_{q}(\mathcal {M}))_{F,K}. \end{aligned}$$

The proof is completed.