1 Introduction and preliminaries

Throughout this paper X is a Banach space which is assumed not to have the Schur property, i.e., X has a weakly convergent sequence that is not norm convergent. \(S(X)\) and \(B(X)\) denote the unit sphere and the unit ball of X, respectively and \(l^{0}\) denotes the set of all real sequences.

A Banach space X is said to have the fixed point property (FPP, for short) if every nonexpansive mapping \(T :C\rightarrow C\), i.e., the mapping satisfying

$$\Vert Tx-Ty\Vert \leq \Vert x-y\Vert ,\quad \forall x,y\in C, $$

and acting on a nonempty bounded closed and convex subset C of X has a fixed point in C. A natural generalization of FPP is the weak fixed point property (WFPP, for short). A Banach space X is said to have the WFPP whenever it satisfies the above condition from the definition of FPP with ‘weakly compact’ in place of ‘bounded closed’. In 1965 Kirk [2] proved that any reflexive Banach space with normal structure has the FPP. In 1989 Prus [3] introduced a property of a Banach space X, called nearly uniformly smoothness. In 1992 Prus [4] also proved that a weakly nearly uniformly smooth Banach space X with the weak Opial property has the FPP. To obtain the weak fixed point property in Banach spaces, García-Falset introduced in [5] the following coefficient:

$$ R(X)=\sup \Bigl\{ \liminf_{n\rightarrow\infty}\Vert x_{n}-x\Vert:\{ x_{n}\} \subset B(X), {x_{n}} \stackrel{w}{\rightarrow} 0, x\in B(X) \Bigr\} . $$

He has proved that a Banach space X with \(R(X)<2\) has the weak fixed point property, i.e., every nonexpansive mapping T from a weakly compact nonempty and convex set \(A\subset X\) into itself has a fixed point in A (see [6] and [7]).

A Banach space X is said to be NUS provided that for every \(\varepsilon>0\) there is \(\eta>0\) such that if \(t\in(0,\eta)\) and \(\{ x_{n}\}\) is a basic sequence in \(B(X)\), then there exists \(k>1\) so that \(\Vert x_{1}+tx_{k}\Vert \leq1+t\varepsilon\) (see [3]).

A natural generalization of this notion is said to be WNUS. A Banach space X is WNUS whenever it satisfies the above condition with ‘for some \(\varepsilon\in(0,1)\)’ in place of ‘for every \(\varepsilon>0\)’ (see [6]).

It is well known that a Banach space X is WNUS if and only if X is reflexive and \(R(X)<2\) (see [6]).

The coefficient \(R(X,a)\) of a Banach space X was defined in 1996 by Benavides [1], as a generalization of the coefficient \(R(X)\), which also plays an important role in the fixed point theory for nonexpansive mappings. Benavides defined the coefficient \(R(X,a)\) for a Banach space X as follows.

Definition 1.1

For a given \(a\geq0\),

$$ R(X,a)=\sup \Bigl\{ \liminf_{n\rightarrow\infty} \Vert x_{n}-x \Vert :x_{n}\in B(X), x_{n} \stackrel{w}{\rightarrow} 0,D\bigl[(x_{n})\bigr]\leq1,\Vert x\Vert \leq a \Bigr\} , $$

where \(D[(x_{n})]={ \liminf_{n\rightarrow\infty} } \{ \Vert x_{i}-x_{j}\Vert :i\neq j,i,j\geq n \}\).

He also defined the following coefficient:

$$M(X)=\sup \biggl\{ \frac{1+a}{R(X,a)}:a\geq0 \biggr\} . $$

Moreover, the coefficient \(R(X,a)\) remains unaltered if in the definition we replace lim inf by lim sup. He obtained the following: for a given \(a\geq0\), if \(R(X,a)<1+a\), then the Banach space X has the fixed point property; the result means that the condition \(M(X)>1\) implies that X has the weak fixed point property for nonexpansive mappings [1].

In this paper, we introduce a new geometric coefficient that is a related García-Falset coefficient and a weak fixed point property, written \(R^{\ast} ( X^{\ast},a ) \). For a given \(a\geq0\), let

$${R^{* }\bigl(X^{* },a\bigr)= \sup\Bigl\{ \lim _{n\rightarrow\infty}\inf\Vert x_{n}-x\Vert:x_{n}\in B \bigl(X^{* }\bigr), x_{n} \stackrel{w^{\ast}}{ \rightarrow} 0, D\bigl[(x_{n})\bigr]\leq1,\Vert x\Vert \leq a, x\in X\Bigr\} } , $$

where \(D[(x_{n})]= \lim_{n\rightarrow\infty}\inf \{ \Vert x_{i}-x_{j}\Vert :i\neq j,i,j\geq n \} \).

Similarly to [1], we can prove that if \(R^{\ast} ( X^{\ast},a ) <1+a\) for some \(a\in(0,1]\) then \(T:C\rightarrow C\) has a fixed points \(a\in C\), if C is a weak-compact and weak sequentially complete subset of \(X^{*}\) and T is a nonexpansive mapping. It is clear that \(R(X,a)=R^{\ast} ( X^{\ast},a ) \) if the Banach space X is reflexive.

A Banach space X is called a Köthe sequence space if it is a subspace of \(l^{0}\) and for every \(x\in l^{0}\) and \(y\in X\) satisfying \(\vert x(i)\vert \leq \vert y(i)\vert \) for all \(i\in \mathcal{N}\), we have \(x\in X\) and \(\Vert x\Vert \leq \Vert y\Vert \) and if there is a \(x=(x(i))\in X\) with \(x(i)>0\) for all \(i\in\mathcal{N}\) (see [8, 9] and [10]).

Let

$$ X_{a}= \Bigl\{ x\in X:\lim_{n\rightarrow\infty}\bigl\Vert \bigl(0,0, \ldots ,0,x(n+1),x(n+2), \ldots \bigr)\bigr\Vert =0 \Bigr\} . $$

A Köthe sequence space X is said to have an absolutely continuous norm if \(X_{a}=X\). A Köthe sequence space X is said to have the semi-Fatou property if for every sequence \(\{x_{n}\}\subset X\) and \(x\in X\) satisfying \(\vert x_{n}(i)\vert \uparrow \vert x(i)\vert \) for all \(i\in N\) we have \(\Vert x_{n}\Vert \rightarrow \Vert x\Vert \).

A mapping \(\Phi:\mathcal{R}\rightarrow\mathcal{R}_{+}\) is said to be an Orlicz function if Φ vanishes only at 0, Φ is even and convex on the whole line R. For every Orlicz function Φ we define its complementary function \(\Psi:\mathcal{R}\rightarrow{}[0,\infty)\) by the formula

$$\Psi ( v ) = {\sup_{{u>0}}} \bigl\{ u\vert v\vert -\Phi ( u ) \bigr\} $$

for every \(v\in\mathcal{R}\).

Denote by \(\Phi= \{ \Phi_{i} \} _{i=1}^{\infty}\) a sequence Orlicz function. Such a sequence is called a Musielak-Orlicz function on \(N\times R\).

We say that Φ satisfies the \(\delta_{2}\)-condition (\(\Phi\in \delta_{2}\) for short) if there exist \(k>0\), \(u_{0}>0\),and \(c_{i}\geq 0\), with \({{\sum_{i\geq1}c_{i}}}<\infty\) such that we have the inequality

$$\Phi_{i} ( 2u ) \leq k\Phi_{i} ( u ) +c_{i} \quad \bigl( i\geq1,\Phi_{i} ( u ) \leq u_{0} \bigr). $$

We now introduce a new definition, namely \(\overline{\delta}_{2}(k)\). We say that Φ satisfies the \(\overline{\delta}_{2}(k)\)-condition (\(\Phi \in\overline{\delta}_{2}(k)\) for short) if there exist \(\varepsilon \in ( 0,1 ) \) and \(i_{\varepsilon}\in\mathcal{N}\) such that

$$I_{\Phi} \biggl( \frac{x}{2} \biggr) \leq\frac{1-\varepsilon}{2}I_{\Phi } ( x ) $$

whenever \(I_{\Phi} ( x ) =k\) and \(N(x)\geq i_{\varepsilon}\), where \(N(x)=\{i\in\mathcal{N}:x(i)\neq0\}\) and \(N(x)\geq i_{\varepsilon}\), which means that \(\min \{ i:i\in N(x) \} \geq i_{\varepsilon}\).

Proposition 1.2

The following are equivalent (see [11]):

  1. (1)

    \(\Phi\in\delta_{2}\);

  2. (2)

    for any \(\varepsilon>0\), there exist \(k_{\varepsilon }>0\), \(u_{\varepsilon}>0\), and \(c_{i}\geq0\) (\(i\geq1 \)), \({\sum_{i\geq1}}c_{i}<\infty\) such that

    $$\Phi_{i} \biggl( \frac{u}{\varepsilon} \biggr) \leq k_{\varepsilon}\Phi _{i} ( u ) +c_{i}\quad \bigl( i\geq1,\Phi_{i} ( u ) \leq u_{\varepsilon} \bigr) ; $$
  3. (3)

    there exist \(\varepsilon\in ( 0,1 ) \), \(i_{\varepsilon}\in M\), \(c_{i}\geq0\) (\(i\geq1\)), \({\sum_{i\geq1}c_{i}}<\infty\) and \(v_{\varepsilon}>0\) such that

    $$\Psi_{i} \biggl( \frac{v}{2} \biggr) \leq\frac{1-\varepsilon}{2}\Psi _{i} ( v ) +c_{i} \quad \bigl( i\geq1,\Psi_{i} ( v ) \leq v_{\varepsilon} \bigr) . $$

The Musielak-Orlicz sequence space \(l_{\Phi}\) is defined to be the set \(\{x\in l_{0}:I_{\Phi}(\lambda x)=\sum_{i=1}^{\infty}\Phi _{i}(\lambda x(i))<\infty\mbox{ for some }\lambda>0\}\) and its subspace \(h_{\Phi}\) is defined to be the set \(\{x\in l_{0}:I_{\Phi}(\lambda x)=\sum_{i=1}^{\infty}\Phi _{i}(\lambda x(i))<\infty\mbox{ for any }\lambda>0\}\) both equipped with the Luxemburg norm

$$\Vert x\Vert =\inf \biggl\{ k>0:I_{\Phi} \biggl( \frac {x}{k} \biggr) \leq1 \biggr\} . $$

To simplify notations, we put \(l_{\Phi}=(l_{\Phi}, \Vert \, \Vert _{\Phi})\) and \(h_{\Phi}=(h_{\Phi}, \Vert \, \Vert _{\Phi})\).

We say that a Musielak-Orlicz function Φ satisfies the \(\overline{\delta}_{2}\)-condition if its complementary function Ψ satisfies the \(\delta_{2}\)-condition.

The basic information on Musielak-Orlicz spaces can be found in [1113], and [14].

2 Results

The idea of Theorem 2.1 is similar to Corollary 8.2 in [15]. In order to keep the consistency of this paper, we accept it in the following.

Theorem 2.1

Let X be a Köthe sequence space with the Fatou property. If X has no absolute continuous norm, then \(R(X,a)=1+a \).

Proof

Suppose that X does not have an absolutely continuous norm. Then there exists \(\varepsilon_{0}>0\) and \(x_{0}\in S(X)\) such that

$$\lim_{n\rightarrow\infty}\Biggl\Vert \sum_{i=n+1}^{\infty }x_{0}(i)e_{i} \Biggr\Vert =\varepsilon_{0}, $$

where \(e_{i}=(0,0, \ldots , \stackrel{i\mathrm{th}}{1},0, \ldots )\).

Take a sequence of positive numbers \(\{\varepsilon_{n}\}\) such that \(\varepsilon_{n}\downarrow0\). By \(\lim_{n\rightarrow\infty} \Vert \sum_{i=n+1}^{\infty}x_{0}(i)e_{i}\Vert =\varepsilon _{0}\), there exists \(n_{1}\in N\) such that

$$\Biggl\Vert \sum_{i=n_{1}}^{\infty}x_{0}(i)e_{i} \Biggr\Vert \leq (1+\varepsilon_{1})\varepsilon_{0}. $$

Notice that

$${\lim_{m\rightarrow\infty} }\Biggl\Vert \sum_{i=n_{1}+1}^{m}x_{0}(i)e_{i} \Biggr\Vert =\varepsilon_{0}, $$

so there exists \(n_{2}>n_{1}\) such that

$$(1-\varepsilon_{1})\varepsilon_{0}\leq\Biggl\Vert \sum _{i=n_{1}+1}^{n_{2}}x_{0}(i)e_{i} \Biggr\Vert \leq(1+\varepsilon _{1})\varepsilon_{0}. $$

In this way, we get by induction a sequence \(\{n_{i}\}\) of natural numbers such that

$$(1-\varepsilon_{i})\varepsilon_{0}\leq\Biggl\Vert \sum _{j=n_{i}+1}^{n_{i+1}}x_{0}(i)e_{i} \Biggr\Vert \leq (1+\varepsilon_{i})\varepsilon_{0}, \quad i=1,2, \ldots . $$

Put \(x_{i}=\sum_{j=n_{i}+1}^{n_{i+1}}x_{0}(i)e_{i}\) and \(v_{k}=\sum_{j=n_{k}+1}^{\infty}x_{0}(i)e_{i}\). Then:

(a) \(\Vert x_{i}\Vert \rightarrow\varepsilon_{0}\) as \(i\rightarrow \infty\).

(b) \(x_{i} \stackrel{w}{\rightarrow} 0\) as \(i\rightarrow\infty\). It is well known that for any Köthe space X we have

$$X^{*}=X'\oplus S, $$

where S is the space of all singular functionals over X, i.e., functionals which vanish on the subspace \(X_{a}=\{x\in X: x \mbox{ has absolutely continuous norm}\}\) and \(X'=\{y\in l^{0}:\sum_{i=1}^{\infty}x(i)y(i)<\infty\mbox{ for all }x\in X\}\) (see [11]). This means that every \(f\in X^{*}\) is uniquely represented in the form

$$f=T_{y}+\varphi, $$

where \(\varphi\in S\) and for \(y\in X'\) the function \(T_{y}\) is defined by

$$T_{y}(x)= \sum_{i=1}^{\infty}x(i)y(i) $$

for all \(x\in X\).

Taking any \(y\in X\), we have

$${\lim_{i\rightarrow\infty} }\sum_{j=1}^{\infty}x_{n}(j)y(j)=\lim_{i\rightarrow\infty} \sum _{j=n_{i}+1}^{n_{i+1}}x_{i}(j)y(j)=0. $$

(c) Put \(z_{i}=\frac{x_{i}}{\Vert x_{i}\Vert }\) and \(w_{k}=\frac{v_{k}}{\Vert v_{k}\Vert }\) for all \(i,k\in\mathcal{N}\). It is easy to check \(D(z_{k}) \leq1\). Then

$$\begin{aligned}& \liminf_{i\rightarrow\infty} \Vert z_{i}+aw_{k}\Vert \\& \quad = \liminf_{i\rightarrow\infty} \biggl\Vert \frac{x_{i}}{\Vert x_{i}\Vert }+a \frac{v_{k}}{\Vert v_{k}\Vert }\biggr\Vert \\& \quad = \liminf_{i\rightarrow\infty}\frac{1}{\Vert x_{i}\Vert \Vert v_{k}\Vert }\bigl\Vert \Vert v_{k}\Vert x_{i}+a\Vert x_{i}\Vert v_{k}\bigr\Vert \\& \quad \geq\frac{1}{\varepsilon_{0}(1+\varepsilon_{k})\varepsilon_{0}} \liminf_{i\rightarrow\infty} \bigl\Vert \Vert v_{k}\Vert x_{i}+a\Vert x_{i}\Vert v_{k}\bigr\Vert \\& \quad =\frac{1}{\varepsilon_{0}(1+\varepsilon_{kj})\varepsilon_{0}} \liminf_{i\rightarrow\infty}\bigl\Vert \bigl( \Vert v_{k}\Vert +a\Vert x_{i}\Vert \bigr) x_{i}+a\Vert x_{i}\Vert (v_{k}-x_{i}) \bigr\Vert \\& \quad \geq\frac{1}{\varepsilon_{0}(1+\varepsilon_{k})\varepsilon_{0}} \Bigl( \liminf_{i\rightarrow\infty}\bigl( \varepsilon_{0}+a(1-\varepsilon _{i})\varepsilon_{0} \bigr)\Vert x_{i}\Vert - \limsup _{i\rightarrow\infty} \bigl( \Vert v_{k}\Vert -\Vert x_{i}\Vert \bigr) \Vert x_{i}\Vert \Bigr) \\& \quad \geq\frac{1}{\varepsilon_{0}(1+\varepsilon_{k})\varepsilon_{0}} \Bigl( \liminf_{i\rightarrow\infty}\bigl( \varepsilon_{0}+a(1-\varepsilon _{i})\varepsilon_{0} \bigr) (1-\varepsilon_{i})\varepsilon_{0}- \limsup _{i\rightarrow\infty} \bigl( \Vert v_{k}\Vert -\Vert x_{i} \Vert \bigr) (1+\varepsilon_{i})\varepsilon _{0} \Bigr) \\& \quad \geq\frac{1}{\varepsilon_{0}(1+\varepsilon_{k})\varepsilon_{0}} \Bigl( \liminf_{i\rightarrow\infty}\bigl( \varepsilon_{0}+a(1-\varepsilon _{i})\varepsilon_{0} \bigr) (1-\varepsilon_{i})\varepsilon_{0} \\& \qquad {}- \limsup _{i\rightarrow\infty} \bigl( (1+\varepsilon _{k})\varepsilon _{0}-(1-\varepsilon_{i})\varepsilon_{0} \bigr) \varepsilon_{0} \Bigr) \\& \quad =\frac{1}{\varepsilon_{0}(1+\varepsilon_{k})\varepsilon_{0}} \bigl( (\varepsilon_{0}+a \varepsilon_{0})\varepsilon _{0}- \bigl((1+\varepsilon _{k})\varepsilon_{0}-\varepsilon_{0} \bigr) \varepsilon_{0} \bigr) \\& \quad =\frac{1}{\varepsilon_{0}(1+\varepsilon_{k})} \bigl( (\varepsilon _{0}+a\varepsilon _{0})-\varepsilon_{k}\varepsilon_{0} \bigr) \\& \quad =\frac{1}{(1+\varepsilon_{k})} \bigl( (1+a)-\varepsilon_{k} \bigr). \end{aligned}$$

By the arbitrariness of k and \(\lim_{{k\rightarrow\infty}}\varepsilon _{k}=0\), we get \(R(X,a)\geq1+a\). It is clear that \(R(X,a)\leq1+a\). Therefore \(R(X,a)=1+a\). □

Corollary 2.2

If \(\Phi\notin{\delta}_{2}\) then \(R(l_{\Phi },a)=R^{\ast} ( l_{\Phi},a ) =1+a\) for any \(0< a\leq1\).

Proof

Since \(\Phi\notin\overline{\delta}_{2}\), we see that \(l_{\Phi}\) has no absolutely continuous norm. So we have \(R(l_{\Phi},a)=1+a\). Since \(h_{\Psi}\) is separable and \(( h_{\Psi} ) ^{\ast}=l_{\Phi}\), we have \(R^{\ast} ( l_{\Phi},a ) =1+a\). □

For any \(x\in\ell_{\Phi}\) with \(\Vert x\Vert =a\) and \(N(x)= \{ i\in N:x(i)\neq0 \} \) being finite, we define \(c_{x}\) by the formula

$$ c_{x}={\lim_{n\rightarrow\infty} }\sup\biggl\{ c_{x,y}>0:I_{\Phi } \biggl( \frac{x}{c_{x,y}} \biggr) +I_{\Phi} \biggl( \frac {y}{c_{x,y}} \biggr) =1: { y}\in \ell_{\Phi}, I_{\Phi} ( y ) \leq \frac{1}{2},n\leq N(y)< \infty\biggr\} . $$

Theorem 2.3

Suppose that \(\Phi\in{\delta}_{2}\). Then for the Musielak-Orlicz sequence space \(\ell_{\Phi}\) we have

$$ R^{\ast} ( \ell_{\Phi},a ) =\sup \bigl\{ c_{x}:x\in \ell _{\Phi}\textit{ with }\Vert x\Vert =a \textit{ and } N(x) \textit{ being finite} \bigr\} . $$

Proof

Let

$$d_{\Phi}=\sup \bigl\{ c_{x}:x\in\ell_{\Phi}\mbox{ with }\Vert x\Vert =a\mbox{ and }N(x)\mbox{ being finite} \bigr\} . $$

Then for any \(\varepsilon\in ( 0,d_{\Phi} ) \), there exists \(\Vert x\Vert =a\) with finite \(N(x)\) such that

$$c_{x}>d_{\Phi}-\varepsilon. $$

By the definition of \(c_{x}\) there exists \(n_{1}\in\mathcal{N}\) such that

$$\sup\biggl\{ c_{x,y}>0:I_{\Phi} \biggl( \frac{x}{c_{x,y}} \biggr) +I_{\Phi } \biggl( \frac{y}{c_{x,y}} \biggr) =1\mbox{ for }I_{\Phi} ( y ) \leq \frac{1}{2}\mbox{ and }N(y)\geq n_{1}\biggr\} > d_{\Phi}-\varepsilon, $$

whenever \(n\geq n_{1}\). By the definition of the supremum, there exists \(y_{1}\in S(\ell_{\Phi})\) with \(N(y_{1})>n_{1}\) such that \(c_{x,y_{1}}>d_{\Phi}-\varepsilon\), i.e., \(I_{\Phi} ( \frac{x}{d_{\Phi }-\varepsilon} ) +I_{\Phi} ( \frac{y_{1}}{d_{\Phi }-\varepsilon} ) >1\). Hence there exists \(n_{2}>n_{1}\) such that \(I_{\Phi} ( \frac{x}{d_{\Phi}-\varepsilon} ) +\sum_{i=n_{1}+1}^{n_{2}}\Phi _{i} ( \frac{y_{1} ( i ) }{d_{\Phi}-\varepsilon} ) >1\). Since \(n_{2}>n_{1}\), we also have

$$\sup\biggl\{ c_{x,y}>0:I_{\Phi} \biggl( \frac{x}{c_{x,y}} \biggr) +I_{\Phi} \biggl( \frac{y}{c_{x,y}} \biggr) =1\mbox{ for }I_{\Phi} ( y ) \leq\frac{1}{2}\mbox{ and }N(y)\geq n_{2}\biggr\} > d_{\Phi }-\varepsilon. $$

There exists \(y_{2}\in\ell_{\Phi}\) with \(I_{\Phi} ( y ) \leq \frac{1}{2}\) and \(N(y_{2})>n_{1}\) such that \(c_{x,y_{2}}>d_{\Phi }-\varepsilon\), i.e., \(I_{\Phi} ( \frac{x}{d_{\Phi }-\varepsilon} ) +I_{\Phi} ( \frac{y_{2}}{d_{\Phi}-\varepsilon} ) >1\). Hence there exists \(n_{2}>n_{1}\) such that \(I_{\Phi} ( \frac {x}{d_{\Phi }-\varepsilon} ) +\sum_{i=n_{1}+1}^{n_{2}}\Phi_{i} ( \frac{y_{2} ( i ) }{d_{\Phi}-\varepsilon} ) >1\). Furthermore, there exists \(n_{3}>n_{2}\) such that \(I_{\Phi} ( \frac{x}{d_{\Phi }-\varepsilon} ) +\sum_{i=n_{2}+1}^{n_{3}}\Phi_{i} ( \frac{y_{2} ( i ) }{d_{\Phi}-\varepsilon} ) >1\). In such a way, we can prove by induction that there exist a sequence \(\{ y_{k } \} _{k=1}^{\infty }\subset\ell_{\Phi}\) with \(I_{\Phi} ( y_{k} ) \leq\frac {1}{2}\) for any natural k and a sequence of natural numbers \(n_{1}< n_{2 }< n_{3}<\cdots \) such that \(I_{\Phi} ( \frac{x}{d_{\Phi}-\varepsilon} ) + \sum_{i=n_{k}+1}^{n_{k+1}}\Phi_{i} ( \frac{y_{k} ( i ) }{d_{\Phi}-\varepsilon} ) >1\) for all \(k\in\mathcal{N}\). It is clear that \(y_{k}\) is weakly star convergent to 0. Since the supports of \(y_{k}\) are pairwise disjoint, we have \(I_{\Phi} ( y_{i}-y_{i} ) =I_{\Phi} ( y_{i} ) +I_{\Phi} ( y_{i} ) \leq\frac {1}{2}+\frac{1}{2}=1\) for \(i,j\in N\) with \(i\neq j\). Therefore, \(D[(y_{k})]\leq1\).

For any \(k>i_{0}\), we have

$$I_{\Phi}\biggl(\frac{y_{k}-x}{d_{\Phi}-\varepsilon}\biggr)=\sum _{i=n_{k}+1}^{n_{k+1}} \Phi_{i} \biggl( \frac{y_{k} ( i ) }{d_{\Phi}-\varepsilon} \biggr) +I_{\Phi} \biggl( \frac{x ( i ) }{d_{\Phi}-\varepsilon} \biggr) >1, $$

i.e., \(\Vert y_{k}-x\Vert >d_{\Phi}-\varepsilon\). Therefore, \(R^{\ast} ( \ell_{\Phi},a ) \geq d_{\Phi}-\varepsilon\) and by the arbitrariness of \(\varepsilon>0 \), we have \(R^{\ast} ( \ell _{\Phi},a ) \geq d_{\Phi}\).

Now, we will prove that \(R^{\ast} ( \ell_{\Phi},a ) \leq d_{\Phi}\). By the definition of \(d_{\Phi}\), we always have

$${\lim_{n\rightarrow\infty} }\sup \biggl\{ c_{x,y}>0:I_{\Phi} \biggl( \frac{x}{c_{x,y}} \biggr) +I_{\Phi} \biggl( \frac{y}{c_{x,y}} \biggr) =1\mbox{ for }y\in S(\ell_{\Phi})\mbox{ and }N(y)\geq n \biggr\} \leq d_{\Phi} $$

for any \(\Vert x\Vert =a\) with finite \(N(x)\).

First of all, we want to prove that for any weak star null sequence \(\{x_{n} \} \subset l_{\Phi}\) and \(\varepsilon>0\) there exists a subsequence \(\{ x_{n_{i}} \} \subset \{ x_{n} \} \) such that \(I_{\Phi} ( x_{n_{i}} ) \leq\frac{1}{2}+\varepsilon\) for each \(i\in N\). Otherwise there exists \(\varepsilon_{0}>0\). Without loss of generality, we may assume that \(I_{\Phi} ( x_{n} ) >\frac {1}{2}+\varepsilon_{0}\) for all \(n\in\dot{N}\). By \(\Phi\in\delta_{2} \) there exists a \(\delta_{1}>0\) such that \(\Vert x\Vert >1+5\delta_{1}\) whenever \(I_{\Phi} ( x ) >1+\frac{2\varepsilon_{0}}{3}\). Using \(\Phi\in\delta_{2} \)again, there exists a \(\delta_{2}>0\) such that \(I_{\Phi} ( x ) <\delta_{2}\) whenever \(\Vert x \Vert <\frac{\varepsilon_{0}}{3}\). Set \(\delta_{0}=\min \{ \delta_{1}, \delta_{2} \} \).

Put \(n_{1}=1\). Then there exists a \(i_{1}>1\) such that \(\Vert \sum_{i=i_{1}+1}^{\infty}x_{n_{1}}(i)e_{i}\Vert <\delta _{0}\leq\delta_{2}\). Since the sequence \(\{ x_{n} \} \) is a weakly null sequence, there exists \(n_{2}>n_{1}\) such that

$$\Biggl\Vert \sum_{i=1}^{i_{1}}x_{n}(i)e_{i} \Biggr\Vert < \delta_{0}\leq \delta_{2}\quad \mbox{whenever }n\geq n_{2}. $$

Using \(\Phi\in\delta_{2}\) again, we see that there exists a \(i_{2}>i_{1}\) such that \(\Vert \sum_{i=i_{2}+1}^{\infty}x_{n_{2}}(i)e_{i}\Vert <\delta_{0}\leq\delta_{2}\). Hence

$$I_{\Phi} \Biggl( \sum_{i=1}^{i_{1}}x_{n_{2}}(i)e_{i} \Biggr) < \frac{\varepsilon_{0}}{3}\quad \mbox{and}\quad I_{\Phi} \Biggl( \sum _{i=i_{2}+1}^{\infty }x_{n_{2}}(i)e_{i} \Biggr) < \frac{\varepsilon_{0}}{3}. $$

Therefore \(I_{\Phi} ( \sum_{i=i_{1}+1} ^{i_{2}}x_{n_{2}}(i)e_{i} ) >\frac{1}{2}+\frac{\varepsilon_{0}}{3}\).

In such a way, we get a subsequence \(\{ x_{n_{j}} \} \subset \{ x_{n} \} \) such that

$$I_{\Phi} \Biggl( \sum_{i=1}^{i_{j-1}}x_{n_{j}}(i)e_{i} \Biggr) < \frac{\varepsilon_{0}}{3},\qquad I_{\Phi} \Biggl( \sum _{i=i_{j}+1}^{\infty }x_{n_{2}}(i)e_{i} \Biggr) < \frac{\varepsilon_{0}}{3} $$

and

$$I_{\Phi } \Biggl( \sum_{i=i_{j-1}+1}^{i_{j}}x_{n_{2}}(i)e_{i} \Biggr) >\frac{1}{2}+ \frac{\varepsilon_{0}}{3} $$

for all \(j\in N\).

So

$$\begin{aligned} \Vert x_{n_{k}}-x_{n_{j}}\Vert = & \Biggl\Vert \sum _{i=1}^{i_{k-1}}x_{n_{k}}(i)e_{i}+\sum_{i=i_{k-1}+1}^{i_{k}}x_{n_{k}}(i)e_{i}+ \sum_{i=i_{k}+1}^{\infty }x_{n_{k}}(i)e_{i} \\ &{}- \sum_{i=1}^{i_{j-1}}x_{n_{j}}(i)e_{i}-\sum_{i=i_{j-1}+1}^{i_{j}}x_{n_{j}}(i)e_{i}- \sum_{i=i_{j}+1}^{\infty }x_{n_{j}}(i)e_{i} \Biggr\Vert \\ \geq& \Biggl\Vert \sum_{i=i_{k-1}+1}^{i_{k}}x_{n_{k}}(i)e_{i}-\sum_{i=i_{j-1}+1}^{i_{j}}x_{n_{j}}(i)e_{i} \Biggr\Vert -4\delta_{0}. \end{aligned}$$

By

$$\begin{aligned}& I_{\Phi} \Biggl( \sum_{i=i_{k-1}+1}^{i_{k}}x_{n_{k}}(i)e_{i}-\sum_{i=i_{j-1}+1}^{i_{j}}x_{n_{j}}(i)e_{i} \Biggr) \\& \quad =I_{\Phi} \Biggl( \sum_{i=i_{k-1}+1}^{i_{k}}x_{n_{k}}(i)e_{i} \Biggr) +I_{\Phi} \Biggl( \sum_{i=i_{j-1}+1}^{i_{j}}x_{n_{j}}(i)e_{i} \Biggr) \\& \quad \geq\frac{1}{2}+\frac{\varepsilon_{0}}{3}+\frac{1}{2}+ \frac{\varepsilon _{0}}{3}=1+\frac{2\varepsilon_{0}}{3}, \end{aligned}$$

we have

$$\Biggl\Vert \Biggl( \sum_{i=i_{k-1}+1}^{i_{k}}x_{n_{k}}(i)e_{i}-\sum_{i=i_{j-1}+1}^{i_{j}}x_{n_{j}}(i)e_{i} \Biggr) \Biggr\Vert \geq 1+5\delta_{1}\geq1+5\delta_{0}. $$

Hence

$$\Vert x_{n_{k}}-x_{n_{j}}\Vert \geq\Biggl\Vert \sum _{i=i_{k-1}+1}^{i_{k}}x_{n_{k}}(i)e_{i}-\sum_{i=i_{j-1}+1}^{i_{j}}x_{n_{j}}(i)e_{i} \Biggr\Vert -4\delta_{0}\geq 1+\delta_{0}. $$

This contradicts with the inequality \(D[ ( x_{n} ) ]\leq1\).

For convenience, we may assume that \(I_{\Phi} ( x_{n} ) \leq \frac{1}{2}+\varepsilon\) for all \(n\in N\). Hence

$$I_{\Phi} \biggl( \frac{1-\varepsilon}{1+\varepsilon}x_{n} \biggr) \leq \frac{1-\varepsilon}{1+\varepsilon}I_{\Phi} ( x_{n} ) \leq\frac{1-\varepsilon}{1+\varepsilon} \biggl(\frac{1}{2}+\varepsilon\biggr)=\frac{1}{2} $$

for all \(n\in N\).

Let us take an element \(x\in l_{\Phi}\) with \(\Vert x\Vert =a\). By \(\Phi\in\delta_{2}\), for any \(\varepsilon>0\) there exists \(i_{0}>0\) such that \(\Vert \sum_{i=i_{0}+1}^{\infty}x(i)e_{i}\Vert <\varepsilon\). Put \(x_{0}=a\frac{\sum_{i=i_{0}+1}^{\infty }x(i)e_{i}}{\Vert \sum_{i=i_{0}+1}^{\infty}x(i)e_{i}\Vert }\). Since \({x_{n}}\stackrel{w^{\ast}}{\rightarrow} 0\), there is a \(n_{0}\in N\) such that \(\Vert \sum_{i=1}^{i_{0}}x_{n} ( i ) e_{i}\Vert <\varepsilon\) when \(n\geq n_{0}\).

Hence

$$\begin{aligned} \Vert x_{n}-x\Vert &\leq\Biggl\Vert \sum _{i=1}^{i_{0}}\bigl(x_{n}(i)-x(i) \bigr)e_{i}+\sum_{i=i_{0}+1}^{\infty} \bigl(x_{n}(i)-x(i)\bigr)e_{i}\Biggr\Vert \\ &\leq\Biggl\Vert \sum_{i=1}^{i_{0}}x(i)e_{i}+ \sum_{i=i_{0}+1}^{\infty }x_{n}(i)e_{i} \Biggr\Vert +2\varepsilon. \end{aligned}$$

We next estimate the \(\Vert \sum_{i=1}^{i_{0}}x(i)e_{i}+\sum_{i=i_{0}+1}^{\infty }x_{n}(i)e_{i}\Vert \). Put \(z_{n}=\frac{1-\varepsilon }{1+\varepsilon}\sum_{i=i_{0}+1}^{\infty}x_{n}(i)e_{i}\) for \(n\geq n_{0}\). We have

$$\begin{aligned}& I_{\Phi} \biggl( \frac{a\frac{\sum_{i=1}^{i_{0}}x(i)e_{i}}{\Vert \sum_{i=1}^{i_{0}}x(i)e_{i}\Vert }-\frac{1-\varepsilon }{1+\varepsilon }\sum_{i=i_{0}+1}^{\infty}x_{n}(i)e_{i}}{d_{\Phi}+\varepsilon} \biggr) \\& \quad =I_{\Phi} \biggl( \frac{x_{0}}{d_{\Phi}+\varepsilon} \biggr) +I_{\Phi } \biggl( \frac{z_{n}}{d_{\Phi}+\varepsilon} \biggr) \\& \quad \leq I_{\Phi} \biggl( \frac{x_{0}}{c_{x_{0},z_{n}}} \biggr) +I_{\Phi } \biggl( \frac{z_{n}}{c_{x_{0},z_{n}}} \biggr) \\& \quad =1, \end{aligned}$$

whence

$$\Biggl\Vert a\frac{\sum_{i=1}^{i_{0}}x(i)e_{i}}{\Vert \sum_{i=1}^{i_{0}}x(i)e_{i}\Vert }-\frac{1-\varepsilon }{1+\varepsilon }\sum _{i=i_{0}+1}^{\infty}x_{n}(i)e_{i}\Biggr\Vert \leq d_{\Phi }+\varepsilon $$

for \(n\geq n_{0}\). Therefore, we obtain the inequalities

$$\begin{aligned}& \Biggl\Vert \sum_{i=1}^{i_{0}}x(i)e_{i}- \sum_{i=i_{0}+1}^{\infty }x_{n}(i)e_{i} \Biggr\Vert \\& \quad \leq\Biggl\Vert x_{0}\frac{\Vert \sum_{i=1}^{i_{0}}x(i)e_{i}\Vert }{a}-\frac{1-\varepsilon}{1+\varepsilon}\sum _{i=i_{0}+1}^{\infty}x_{n}(i)e_{i}- \frac {2\varepsilon}{1+\varepsilon}\sum_{i=i_{0}+1}^{\infty}x_{n}(i)e_{i} \Biggr\Vert \\& \quad \leq a \biggl( \frac{\Vert \sum_{i=1}^{i_{0}}x(i)e_{i}\Vert }{a}-1 \biggr) +\biggl\Vert a \frac{\sum_{i=1}^{i_{0}}x(i)e_{i}}{\Vert \sum_{i=1}^{i_{0}}x(i)e_{i}\Vert }-z_{n}\biggr\Vert +\frac{2\varepsilon}{1+\varepsilon}\Biggl\Vert \sum_{i=i_{0}+1}^{\infty }x_{n}(i)e_{i} \Biggr\Vert \\& \quad \leq a \biggl( \frac{a+\varepsilon}{a}-1 \biggr) +\biggl\Vert a \frac{\sum_{i=1}^{i_{0}}x(i)e_{i}}{\Vert \sum_{i=1}^{i_{0}}x(i)e_{i}\Vert }-z_{n}\biggr\Vert +\frac{2\varepsilon}{1+\varepsilon} \\& \quad \leq a\varepsilon+d_{\Phi}+\varepsilon+\frac{2\varepsilon }{1+\varepsilon }=d_{\Phi}+ \biggl( \frac{2+(1+a)(1+\varepsilon)}{1+\varepsilon} \biggr) \varepsilon. \end{aligned}$$

Therefore, we have \(\Vert x_{n}-x_{0}\Vert \leq d_{\Phi }+ ( \frac{2+(1+a)(1+\varepsilon)}{1+\varepsilon} ) \varepsilon +2\varepsilon\). By the arbitrariness of \(\varepsilon>0 \), we get the inequality \(R ( \ell _{\Phi} ) \leq d_{\Phi}\).

Summing up, we see that the equality \(R ( \ell_{\Phi} ) =d_{\Phi}\) holds. □

For any \(x\in\ell_{\Phi}\) with \(\Vert x\Vert = 1\) and \(N(x)= \{ i\in N:x(i)\neq0 \} \) being finite we define \(c_{x}\) as follows:

$$\begin{aligned} \tilde{c}_{x} =&{\lim_{n\rightarrow\infty} }\sup\biggl\{ c_{x,y}>0:I_{\Phi } \biggl( \frac{x}{c_{x,y}} \biggr) +I_{\Phi} \biggl( \frac {y}{c_{x,y}} \biggr) =1\mbox{ for } y\in \ell_{\Phi} \mbox{ with } I_{\Phi} ( y ) \leq 1, \\ & n\leq N(y)< \infty\biggr\} . \end{aligned}$$

Corollary 2.4

If \(\Phi\in\delta_{2}\) then \(R(l_{\Phi})= \sup \{ \tilde{c}_{x}:x\in\ell_{\Phi}\textit { with }\Vert x\Vert =1 \textit{ and } N(x) \textit{ being finite} \}\).

Proof

The proof is similar to the proof of Theorem 2.3. □

Corollary 2.5

If \(\Phi\in\delta_{2}\) and \(\Phi\in \overline{\delta}_{2}\) then \(R(l_{\Phi},a)=R^{\ast} ( l_{\Phi},a ) \) for any \(0< a\leq1\).

Theorem 2.6

\(R^{\ast}(\ell_{\Phi},1)<2\) if and only if \(\ell_{\Phi}\in\delta_{2}\) and \(\Phi\in\overline{\delta}_{2}(1)\).

Proof

Necessity. We only need to prove the necessity of \(\Phi \in\overline{\delta}_{2}(1)\). Suppose that \(\Phi\notin\overline{\delta}_{2}(1)\). Then for any natural number k there exists \(x_{k}\in S ( \ell_{\Phi} ) \) with \(N(x_{k})\geq k\) such that

$$I_{\Phi} \biggl( \frac{x_{k}}{2} \biggr) >\frac{1-\frac{1}{k}}{2}I_{\Phi } ( x_{k} ) . $$

For any fixed \(k\in N\) and for any \(\varepsilon>0\), there exists \(i_{k}\in N\) such that \(\sum_{i=k}^{i_{k}}\Phi_{i}(x_{k}(i))>1-\varepsilon\). Put \(\overline{x}_{k}=\frac{\sum_{i=k}^{i_{k}}x_{k}(i)e_{i}}{\Vert \sum_{i=k}^{i_{k}}x_{k}(i)e_{i}\Vert }\). Then

$$\begin{aligned} I_{\Phi} \biggl( \frac{\overline{x}_{k}}{2} \biggr) +I_{\Phi} \biggl( \frac {x_{i_{k}}}{2} \biggr) &=I_{\Phi} \biggl( \frac{\sum _{i=k}^{i_{k}}x_{k}(i)e_{i}}{2\Vert \sum_{i=k}^{i_{k}}x_{k}(i)e_{i}\Vert } \biggr) +I_{\Phi} \biggl( \frac{x_{i_{k}}}{2} \biggr) \\ &\geq I_{\Phi} \biggl( \frac{\sum_{i=k}^{i_{k}}x_{k}(i)e_{i}}{2} \biggr) +I_{\Phi} \biggl( \frac{x_{i_{k}}}{2} \biggr) +I_{\Phi} \biggl( \frac{\sum_{i=i_{k}+1}^{\infty}x_{k}(i)e_{i}}{2} \biggr) -\varepsilon \\ &=I_{\Phi} \biggl( \frac{x_{k}}{2} \biggr) +I_{\Phi} \biggl( \frac {x_{i_{k}}}{2} \biggr) -\varepsilon\geq\frac{1-\frac{1}{k}}{2}+ \frac{1-\frac {1}{i_{k}}}{2}-\varepsilon \\ &=1-\frac{1}{k}-\varepsilon, \end{aligned}$$

which shows that \(\Vert \overline{x}_{k}+x_{i_{k}}\Vert \geq 2(1-\frac{1}{k}-\varepsilon)\). Hence \(c_{\overline{x}_{k},x_{i_{k}}}\geq2(1-\frac {1}{k}-\varepsilon)\). By the arbitrariness of \(\varepsilon>0 \) and \(k\in N\), we see that \(R(\ell_{\Phi})=2\).

Sufficiency. Since \(\Phi\in\overline{\delta}_{2}(1)\), there exist \(\varepsilon_{0}\in ( 0,1 ) \) and \(i_{\varepsilon}\in \mathcal{N}\) such that

$$I_{\Phi} \biggl( \frac{x}{2} \biggr) \leq\frac{1-\varepsilon _{0}}{2}I_{\Phi } ( x ), $$

whenever \(I_{\Phi} ( x ) =1\) and \(N(x)\geq i_{\varepsilon}\). For any \(x\in S(\ell_{\Phi})\) with finite \(N(x)\) and \(y\in S(\ell_{\Phi})\) with \(N(y)\geq n\), we may assume without loss of generality that \(\max \{ i:i\in N(x) \} < n\) and \(n\geq i_{\varepsilon}\). Hence

$$I_{\Phi} \biggl( \frac{x}{2} \biggr) +I_{\Phi} \biggl( \frac{y}{2} \biggr) \leq \frac{1}{2}+\frac{1-\varepsilon}{2}= \frac{2-\varepsilon}{2}. $$

Since \(\Phi\in\delta_{2}\), there exists \(0<\alpha<1\) such that

$$\Vert z\Vert \leq\alpha\quad \mbox{whenever }I_{\Phi} ( z ) \leq \frac{2-\varepsilon}{2}. $$

Therefore \(\Vert \frac{x+y}{2}\Vert \leq\alpha\), i.e., \(\Vert x+y\Vert \leq2\alpha\). Note that \(c_{x,y}=\Vert x+y\Vert \). Hence \(R(\ell_{\Phi})\leq2\alpha<2\). □

Theorem 2.7

\(R^{\ast}(\ell_{\Phi},a)<1+a\) for \(0< a<1\) if and only if \(\Phi \in\delta_{2}\).

Proof

We only need to prove the sufficiency. For any \(0<\varepsilon<\frac{1}{2}\), by \(\Phi\in\delta_{2}\), there exists a \(d_{0}>0\) such that \(\Vert x\Vert \leq1-d\) whenever \(I_{\Phi} ( x ) \leq\frac {1}{2}+\varepsilon\). Hence \(\Vert x_{n}\Vert \leq1-d\) if n large enough for any weakly star null sequence \(\{ x_{n} \} \subset B ( l_{\Phi} ) \) with \(D[ ( x_{n} ) ]\leq1\). Hence

$$\liminf_{n\rightarrow\infty} \Vert x_{n}-x\Vert \leq1-d+a< 1+a, $$

that is, \(R^{\ast} ( l_{\Phi},a ) <1+a\). □

Example 2.8

Let \(\Phi_{n}(u)=\bigl \{\scriptsize{ \begin{array}{l@{\quad}l} u^{2}& \mbox{if }u\leq\frac{1}{n}, \\ a_{n}u+b_{n}& \mbox{if }\frac{1}{n}\leq u\leq\infty, \end{array}} \bigr.\) where \(a_{n}=\frac{2}{n}\), \(b_{n}=-\frac{1}{n^{2}}\). Then \(\Phi _{n}\) is an Orlicz function for each \(n\in N\).

If \(u\leq\frac{1}{n}\) and \(2u>\frac{1}{n}\), then \(\frac{1}{2n}< u\leq \frac{1}{n}\). Hence

$$\Phi_{n} ( 2u ) =\frac{2}{n}2u-\frac{1}{n^{2}}\leq \frac {4}{n^{2}}-\frac{1}{n^{2}}=\frac{3}{n^{2}}\leq12 \frac{1}{4n^{2}}\leq 12u^{2}=12\Phi _{n} ( u ) . $$

If \(0< u\leq\frac{1}{2n}\) then \(\Phi_{n} ( 2u ) =4\Phi _{n} ( u ) \). If \(\frac{1}{n}< u\leq1\) then \(\Phi_{n} ( 2u ) \leq 2\Phi_{n} ( u ) \). If we put \(K=24\) and \(u_{0}=1\), then

$$\Phi_{n}(2u)\leq K\Phi_{n} ( u )\quad \mbox{for all }n\in N, $$

that is, \(\Phi\in\delta_{2}\).

If \(0< a <1\) then \(R^{\ast} ( l_{\Phi},a ) <1+a\).

Let us take \(x_{n}= ( 0,0,\ldots,0, \overbrace{\tfrac{n}{2}+\tfrac{1}{2n}}^{n\mathrm{th}},0,\ldots ) \) for any \(n\in N\). Then \(I_{\Phi} ( x_{n} ) =1\) and

$$I_{\Phi} \biggl( \frac{x_{n}}{2} \biggr) =\frac{2}{n} \biggl( \frac {n}{4}+\frac{1}{4n} \biggr) -\frac{1}{n^{2}}= \frac{1}{2}-\frac{1}{2n^{2}}, $$

whence \({\lim_{n\rightarrow\infty} }I_{\Phi} ( \frac{x_{n}}{2} ) =\frac{1}{2}\). Therefore \(\Phi\notin\delta_{2} ( 1 ) \), which implies that \(R^{\ast} ( l_{\Phi},1 ) =2\).

Let \(\{ p_{k} \} _{k=1}^{\infty}\) be a sequence of real increasing numbers with \(1< p_{1}\) and \({\lim_{n\rightarrow\infty} }p_{n}=p<\infty\). Then we have the following.

Theorem 2.9

Let \(l^{(p_{i})}\) be a Nakano sequence space equipped with the Luxemburg norm. Then \(R(l^{(p_{i})})=2^{\frac{1}{p}}\) and \(R(l_{\Phi},a)= ( \frac{1}{2}+a^{p} ) ^{\frac{1}{p}}\), \(0< a\leq1\).

Proof

Since \(1<\inf \{ p_{i} \} \leq\sup \{ p_{i} \} =p<\infty\), the Nakano space equipped with the Luxemburg norm is reflexive. For any \(x,y\in S ( l^{(p_{i})} ) \) with \(N(x)\), \(N(y)\) being finite. We now consider the following equation:

$$I_{\Phi} \biggl( \frac{x}{c} \biggr) +I_{\Phi} \biggl( \frac{y}{c} \biggr) =1, $$

that is,

$$\sum_{i=1}^{\infty}\biggl\vert \frac{x(i)}{c}\biggr\vert ^{p_{i}}+\sum_{i=1}^{\infty} \biggl\vert \frac{y ( i ) }{c}\biggr\vert ^{p_{i}}=1. $$

Then

$$\frac{1}{c^{p}}\sum_{i=1}^{\infty}\bigl\vert x(i)\bigr\vert ^{p_{i}}+\frac{1}{c^{p}}\sum _{i=1}^{\infty}\bigl\vert x_{n}(i)\bigr\vert ^{p_{i}}=\frac{2}{c^{p}}\leq1, $$

i.e., \(c\leq2^{\frac{1}{p}}\). This shows the inequality \(R(l^{(p_{i})},1)\leq2^{\frac{1}{p}}\). Take the classical basic sequence \(\{ e_{n} \} \subset S(l^{(p_{i})})\). If \(c_{n,m}\) is a solution of the equation

$$I_{\Phi} \biggl( \frac{e_{n}}{c} \biggr) +I_{\Phi} \biggl( \frac{e_{m}}{c} \biggr) =1, $$

and assuming without loss of generality that we may take \(n>m\), we have the inequality

$$\frac{2}{c^{p_{n}}}\geq1. $$

Hence \(R(l^{(p_{i})})\geq2^{\frac{1}{p}}\). Together with the opposite inequality proved already, we have \(R(l^{(p_{i})})=2^{\frac{1}{p}}\).

If \(0< a<1\), for \(x\in B ( l^{(p_{i})} ) \) with finite \(N(x)\) and \(I_{\Phi} ( x ) =\frac{1}{2}\) and \(y\in B ( l^{(p_{i})} ) \) with finite \(N(y)\) and \(\Vert x\Vert =a\), we consider the following equation:

$$I_{\Phi} \biggl( \frac{x}{c_{x,y}} \biggr) +I_{\Phi} \biggl( \frac {y}{c_{x,y}} \biggr) =1, $$

that is, the equation

$$\sum_{i=1}^{\infty}\biggl\vert \frac{x(i)}{c_{x,y}}\biggr\vert ^{p_{i}}+\sum _{i=1}^{\infty}\biggl\vert \frac{y ( i ) }{c_{x,y}}\biggr\vert ^{p_{i}}=1. $$

Hence

$$\frac{1}{c_{x,y}^{p}}\sum_{i=1}^{\infty}\bigl\vert x(i)\bigr\vert ^{p_{i}}+\frac{a^{p_{\max \{ i:i\in N(y) \} }}}{c_{x,y}^{p}}\sum_{i=1}^{\infty}\biggl\vert \frac{y(i)}{a}\biggr\vert ^{p_{i}}=\frac {1}{2c_{x,y}^{p}}+ \frac{a^{p_{\max \{ i:i\in N(y) \} }}}{c_{x,y}^{p}}\geq1, $$

where \(c_{x,y}^{p}\leq\frac{1}{2}+a^{p_{\max \{ i:i\in N(y) \} }}\). Therefore \(R(l_{\Phi},a)\leq ( \frac{1}{2}+a^{p} ) ^{\frac{1}{p}}\).

Taking \(x_{n}=(0,\ldots,0, ( \frac{1}{2} ) ^{\frac{1}{p_{n}}},0,\ldots)\) and \(x_{m}=(0,\ldots,0, \stackrel{m\mathrm{th}}{a} ,0,\ldots)\), we get \(I_{\Phi} ( x_{n} ) =\frac{1}{2}\) and \(I_{\Phi} ( \frac{x_{m}}{a} ) =1\), which implies the equality \(\Vert x_{m}\Vert =a\).

For any \(n\neq m\), if \(c>0\) is such that

$$I_{\Phi} \biggl( \frac{x_{n}+x_{m}}{c} \biggr) =1, $$

then

$$\biggl( \frac{1}{2}+a^{p_{m}} \biggr) ^{\frac{1}{\max \{ p_{n},p_{m} \} }}\leq c. $$

Letting \(n,m\rightarrow\infty\), we get \(R(l_{\Phi},a)\leq ( \frac {1}{2}+a^{p} ) ^{\frac{1}{p}}\), that is, \(R(l_{\Phi},a)= ( \frac {1}{2}+a^{p} ) ^{\frac{1}{p}}\). □