1 Introduction

Invariant means on amenable groups are an important tool in many parts of Mathematics, especially in Harmonic analysis (see [10, 11]). For the basic properties of invariant means, we refer the reader to [10]. Invariant means and their generalizations for vector-valued functions also play an important role in the stability of functional equations and selections of set-valued functions (see [1, 6, 7, 20]).

The space of all bounded functions from a set S into a Banach space X is denoted by \(\ell _\infty (S,X)\). Let us recall the definition of an amenable semigroup (see [5]).

Definition 1.1

A semigroup \((S,+)\) is called left [resp. right] amenable if and only if there exists a linear map \(L:\ell _\infty (S,{\mathbb {R}}) \rightarrow {\mathbb {R}}\) such that

$$\begin{aligned} \inf f(S)&\le L(f)\le \sup f(S),\ f\in \ell _\infty (S,{\mathbb {R}})\\ L(_a f)&= L(f),\ a\in S,\ f\in \ell _\infty (S,{\mathbb {R}}),\\ [L(f_a)&= L(f),\ a\in S,\ f\in \ell _\infty (S,{\mathbb {R}})], \end{aligned}$$

where

$$\begin{aligned} a f(x)&= f(a+x),\ a,x\in S,\ f\in \ell _\infty (S,{\mathbb {R}}),\\ [f_a(x)&= f(x+a),\ a,x\in S,\ f\in \ell _\infty (S,{\mathbb {R}})]. \end{aligned}$$

If both left and right invariant means exist, then there exists a two-sided invariant mean and S is called amenable.

Remark 1.2

In the above definition the first condition

$$\begin{aligned} \inf f(S)\le L(f)\le \sup f(S) \end{aligned}$$

is equivalent to conditions \(L(1_S)=1\) and \(|L(f)|\le \Vert f \Vert :=\sup |f(S)|\).

It is known that every commutative semigroup is amenable (an easy consequence of the Markov–Kakutani fixed point theorem, see [18, Theorem 5.23]).

Certain generalizations of invariant means were investigated for vector-valued functions in [6] and the existence thereof appears to be related to properties such as reflexivity.

Some generalized definition of an invariant mean has been used by many mathematicians as a folklore (e.g. by Pełczyński [17]). The explicit form of this definition can be found e.g. in the work of Ger [7].

Definition 1.3

Let \((S,+)\) be a left [right] amenable semigroup and X be a Banach space. A linear map \({M:\ell _{\infty } (S,X) \rightarrow X}\) is called a left [right] X-valued invariant mean if

$$\begin{aligned} \Vert M \Vert&\le 1,\\ M(c 1_S)&=c,\ c\in X,\\ M(_a f)&=M(f),\ a\in S,f\in \ell _\infty (S,X),\\ [M(f_a)&=M(f),\ a\in S,f\in \ell _\infty (S,X),] \end{aligned}$$

where

$$\begin{aligned} {}_a f(x)&= f(a+x),\ a,x\in S,\ f\in \ell _\infty (S,X),\\ [f_a(x)&= f(x+a),\ a,x\in S,\ f\in \ell _\infty (S,X).] \end{aligned}$$

If M is a left and right invariant mean, then M is called an X-valued invariant mean.

If in the above definition the norm of map M is equal to at most \(\lambda \ge 1\), then M is called an X-valued invariant \(\lambda \)-mean.

The existence of such invariant means for a fixed Banach space and for all amenable semigroups has been studied by Domecq [3, Theorem 1 and 2], Kania [12], Goucher and Kania [9], and by the author in [15, 16]. The most general result obtained is as follows (recall that for \(\lambda \ge 1\) a subspace X of a Banach space Y is called \(\lambda -\)complemented if there is a projection of norm \(\lambda \) from Y on X).

Corollary 1.4

(see [16, Corollary 3.3]) Let X be an infinite-dimensional Banach space, \(\lambda \ge 1\). The following assertions are equivalent:

  1. (1)

    X is complemented in \(X^{**}\) by a projection of norm at most \(\lambda \);

  2. (2)

    for every amenable semigroup S there exists an X-valued invariant \(\lambda \)-mean on S;

  3. (3)

    for any commutative cancellative semigroup S of torsion-free rank \(\delta \), \({\text {dens}}{X^{**}}=\max (\delta , \omega )\), there exists an X-valued invariant \(\lambda \)-mean on S.

It is also demonstrated ([9, Remark 1.1]) that there exists a commutative noncancellative semigroup S (that could be chosen as large as one wishes) such that there exists an X-valued invariant mean on S.

Definition 1.5

A Banach space X is said to have the finite-infinite intersection property (\(IP_{f,\infty }\)) if every family of closed balls in X with empty intersection contains a finite subfamily with empty intersection.

It is known that dual spaces and their constrained subspaces have \(IP_{f,\infty }\). It can be shown (see [8]) that a Banach space X has \(IP_{f,\infty }\) if and only if any family of closed balls centered at points of X that intersects in \(X^{**}\) also intersects in X. With this in mind Bandyopadhyay and Dutta ([2]) defined

Definition 1.6

A subspace X of a Banach space Y is said to be an almost constrained (AC) subspace of Y if any family of closed balls centered at points of X that intersects in Y also intersects in X.

Clearly, any constrained subspace of a Banach space is an AC-subspace. In the case of \(IP_{f,\infty }\), whether the converse is also true remains an open question. However, they give an example (from [13]) that an AC-subspace need not, in general, be constrained.

In this paper we will study a connection between some generalization of AC-subspaces, vector valued invariant \(\lambda \)-means and \(\lambda \)-complementability.

2 Preliminaries

First we recall the definition of torsion-free rank (see [4]).

Definition 2.1

Let S be a commutative cancellative semigroup. A set \(A\subset S\) is independent if \(\sum _{i=1}^n k_i a_i=\sum _{i=1}^n m_i a_i\) for any \(n\in {\mathbb {N}}\) and \(a_i\in A\), \(k_i,m_i \in {\mathbb {N}}_0\), \(i\in \{ 1,\ldots ,n \}\) implies \(k_i=m_i\) for \(i\in \{ 1,\ldots ,n \}\).

Let further \({\mathcal {A}}_0\) be the family of all independent sets L in S consisting only of elements whose order is infinite and such that L is maximal with respect to these properties. The cardinal number of any set in \({\mathcal {A}}_0\) is called the torsion-free rank of S and is denoted by \(r_0 (S)\) (all the sets in \({\mathcal {A}}_0\) have the same cardinal number).

The density character of a Banach space X, denoted \({\text {dens}}X\), is the smallest cardinal \(\kappa \) for which X has a dense subset of cardinality \(\kappa \).

We will also require the version of the Principle of Local Reflexivity due to Lindenstrauss and Rosenthal [14]. We will identify any element of a Banach space X with its canonical image in \(X^{**}\).

Theorem 2.2

Let X be a Banach space. Then for every finite-dimensional subspace \(F\subset X^{**}\) and each \(\varepsilon \in (0, 1]\) there exists a linear map \(P_F^\varepsilon :F\rightarrow X\) such that

  1. (1)

    \((1-\varepsilon )\Vert x \Vert \le \Vert P_F^\varepsilon (x) \Vert \le (1+\varepsilon )\Vert x \Vert \), \(x\in F\);

  2. (2)

    \(P_F^\varepsilon (x) =x\) for \(x\in F\cap X\).

We also recall some result from [16].

Theorem 2.3

(see [16, Theorems 3.1 and 3.2]) Let X be an infinite-dimensional Banach space, \(\lambda \ge 1\), S be a commutative cancellative semigroup of torsion-free rank \(\delta \), \(\gamma = \max (\delta ,\omega )\). If there exists an X-valued invariant \(\lambda \)-mean \(M:\ell _{\infty } (S,X)\rightarrow X\), then for every subspace E of \(X^{**}\) such that \({\text {dens}}E=\gamma \) there exists a linear map \(P:E\rightarrow X\) such that \(\Vert P \Vert \le \lambda \) and \(P(x)=x\) for \(x\in X\cap E\).

Now, we introduce generalizations of the finite-infinite intersection property and almost constrained subspaces.

Definition 2.4

Let \(\gamma \ge \omega \) be a cardinal number, \(\lambda \ge 1\). We say that a Banach space X has the \(\lambda -\)finite-\(\gamma \) intersection property (\(\lambda -IP_{f,\gamma }\)) if for every collection \({\mathcal {F}} = \{ B_X (x_\alpha ,r_\alpha ):\alpha < \gamma \}\) of closed balls in X such that \(\bigcap \nolimits _{\alpha <\gamma } B_X(x_\alpha ,\lambda r_\alpha ) = \emptyset \), there exists a finite set \(J\subset \gamma \) such that \(\bigcap \nolimits _{\alpha \in J } B_X(x_\alpha ,r_\alpha ) = \emptyset \).

If X has \(\lambda -IP_{f,\gamma }\) for every cardinal number \(\gamma \), then X is said to haves the \(\lambda \) -finite-infinite intersection property (\(\lambda -IP_{f,\infty }\)).

Definition 2.5

Let \(\gamma \ge \omega \) be a cardinal number, \(\lambda \ge 1\). A subspace X of a Banach space Y is said to be a \((\lambda ,\gamma )\) almost constrained (\((\gamma ,\lambda )-AC\)) subspace of X if for every family \({\mathcal {F}} = \{ B_X (x_\alpha ,r_\alpha ):\alpha < \gamma \}\) of closed balls in X, if \({\mathcal {F}}\) intersects in Y, then \(\{ B_X (x_\alpha ,\lambda r_\alpha ):\alpha < \gamma \}\) intersects in X.

If X is a \(( \lambda ,\gamma )-AC\) subspace of Y for every cardinal number \(\gamma \), then X is said to be a \(\lambda \) -almost constrained (\(\lambda -AC\)) subspace of Y.

Lemma 2.6

Let X, Y be Banach spaces, \(X\subset Y\). Then X is a \((2+\varepsilon )-AC\) subspace of Y for every \(\varepsilon >0\).

Proof

Let \({\mathcal {F}} = \{ B_X (x_\alpha ,r_\alpha ): \alpha \in I \}\) intersects in Y. Let further \(r:=\inf \{ r_\alpha : \alpha \in I \}\) and \(\varepsilon >0\). If \(r=0\), then there exists a sequence \((x_{\alpha _n})_{n\in {\mathbb {N}}}\) such that \(\lim \nolimits _{n\rightarrow \infty } r_{\alpha _n} =0\). Hence \((x_{\alpha _n})_{n\in {\mathbb {N}}}\) is convergent to some \(x\in X\). We have

$$\begin{aligned}&\Vert x_\alpha -x \Vert \le \Vert x_\alpha - x_{\alpha _n} \Vert +\Vert x_{\alpha _n} -x \Vert \le r_\alpha +r_{\alpha _n} +\Vert x_{\alpha _n} -x \Vert ,\ n\in {\mathbb {N}}, \end{aligned}$$

so

$$\begin{aligned}&\Vert x_\alpha -x \Vert \le r_\alpha +\lim \limits _{n\rightarrow \infty } (r_{\alpha _n} +\Vert x_{\alpha _n} -x \Vert ) = r_\alpha , \end{aligned}$$

which means that \(x\in \bigcap {\mathcal {F}}\).

Assume that \(r>0\). Let \(x_{\alpha _0}\) be such that \(r_{\alpha _0} <r(1+\varepsilon )\) and \(y\in Y\) be a point from the intersection of \({\mathcal {F}}\) in Y. Then

$$\begin{aligned} \Vert x_\alpha -x_{\alpha _0} \Vert&\le \Vert x_\alpha -y \Vert + \Vert y -x_{\alpha _0} \Vert \le r_\alpha + r_{\alpha _0}\\&\le r_\alpha + (1+\varepsilon )r\le r_\alpha + (1+\varepsilon )r_\alpha = (2+\varepsilon )r_\alpha . \end{aligned}$$

\(\square \)

Lemma 2.7

Let XY be Banach spaces, \(\lambda \ge 1\). If X is \(\lambda \)-complemented in Y, then X is a \(\lambda -AC\) subspace of Y.

Proof

Let \(P:Y\rightarrow X\) be a norm \(\lambda \) projection, \(\gamma \) be a cardinal number, \({\mathcal {F}} = \{ B_X (x_\alpha ,r_\alpha ): \alpha < \gamma \}\) be a family of closed balls in X such that \({\mathcal {F}}\) intersects in Y. Let \(z\in Y\) be any point from this intersection. Then

$$\begin{aligned}&\Vert P(z) -x_\alpha \Vert = \Vert P(z-x_\alpha ) \Vert \le \lambda \Vert z-x_\alpha \Vert \le \lambda r_\alpha ,\ \alpha < \gamma , \end{aligned}$$

so \(P(z) \in \bigcap \nolimits _{\alpha < \gamma } B_X (x_\alpha ,\lambda r_\alpha )\). \(\square \)

Lemma 2.8

Let X be a Banach space, \(\lambda \ge 1\), \(\gamma \ge \omega \) be a cardinal number. Then the following are equivalent:

  1. (a)

    X has \(\lambda -IP_{f,\gamma }\).

  2. (b)

    For every Banach space Y, \(X\subset Y\subset X^{**}\), X is a \((\lambda ,\gamma ) -AC\) subspace of Y.

  3. (c)

    X is a \((\lambda ,\gamma ) -AC\) subspace of \(X^{**}\).

Proof

(a)\(\Rightarrow \)(b) Assume that X has \(\lambda -IP_{f,\gamma }\). Let Y be a Banach space such that \(X\subset Y\subset X^{**}\). Let further \({\mathcal {F}} = \{ B_X (x_\alpha ,r_\alpha ): \alpha < \gamma \}\), \(z\in Y\subset X^{**}\) be a point from the intersection of \({\mathcal {F}}\) in Y. Let J be a finite subset of \(\gamma \) and \(\varepsilon >0\). From the Principle of Local Reflexivity there exists a linear map \(P:{\text {span}}\left\{ \{ x_\alpha :\alpha \in J \} \cup \{ z\} \right\} \rightarrow X\) such that \(P(x_\alpha )=x_\alpha \), \(\alpha \in J\) and \(\Vert P \Vert \le 1+\varepsilon \).

$$\begin{aligned}&\Vert P(z)-x_\alpha \Vert =\Vert P(z-x_\alpha ) \Vert \le (1+\varepsilon ) \Vert z-x_\alpha \Vert \le (1+\varepsilon )r_\alpha . \end{aligned}$$

Hence \(P(z)\in B_X (x_\alpha ,(1+\varepsilon )r_\alpha )\), \(\alpha \in J\) so \(\bigcap \nolimits _{\alpha \in J } B_X (x_\alpha ,(1+\varepsilon )r_\alpha )\ne \emptyset \). Thus, from \(\lambda -IP_{f,\gamma }\) we have

$$\begin{aligned}&\bigcap \limits _{\alpha< \gamma } B_X (x_\alpha ,\lambda r_\alpha ) = \bigcap \limits _{\alpha < \gamma } \bigcap \limits _{n\in {\mathbb {N}}} B_X (x_\alpha ,\frac{n+1}{n}\lambda r_\alpha ) \ne \emptyset . \end{aligned}$$

(b)\(\Rightarrow \)(c) is trivial.

(c)\(\Rightarrow \)(a) Let \({\mathcal {F}} = \{ B_X (x_\alpha ,r_\alpha ): \alpha < \gamma \}\) be such that for every finite subset J of \(\gamma \) we have \(\bigcap \nolimits _{\alpha \in J } B_X (x_\alpha ,r_\alpha )\ne \emptyset \). Then for every finite subset J of \(\gamma \) we have \(\bigcap \nolimits _{\alpha \in J } B_{X^{**}}(x_\alpha ,r_\alpha )\ne \emptyset \) and since \(X^{**}\) is a dual space, it has \(IP_{f,\infty }\), then \(\bigcap \nolimits _{\alpha < \gamma } B_{X^{**}}(x_\alpha ,r_\alpha )\ne \emptyset \). Since X is a \((\lambda ,\gamma )-AC\) subspace of \(X^{**}\), we obtain that \(\bigcap \nolimits _{\alpha < \gamma } B_X ( x_\alpha ,\lambda r_\alpha )\ne \emptyset \). \(\square \)

Lemma 2.9

Let X, Y be Banach spaces, \(\lambda \ge 1\). If X is a \((\lambda ,{\text {dens}}(X) )-AC\) subspace of Y, then X is a \(\lambda -AC\) subspace of Y.

Proof

Assume that X is a \((\lambda ,{\text {dens}}(X))-AC\) subspace of Y. Let \(D\subset X\) be a dense set in X such that \(\vert D\vert ={\text {dens}}(X)\). For every \(x_{\alpha }\in X\) and \(r_\alpha >0\) there exist \((x_{\alpha ,n})_{n\in {\mathbb {N}}} \subset D\), \((q_{\alpha ,n})_{n\in {\mathbb {N}}} \subset {\mathbb {Q}}_+\) such that \(\lim \nolimits _{n\rightarrow \infty } x_{\alpha ,n} =x_\alpha \), \(q_{\alpha ,n}\in [r_\alpha +\Vert x-x_{\alpha ,n} \Vert ,r_\alpha +\Vert x-x_{\alpha ,n} \Vert +\frac{1}{n}]\). Hence

$$\begin{aligned}&\bigcap \limits _{\alpha \in I} B_X (x_\alpha , \lambda r_\alpha ) = \bigcap \limits _{\alpha \in I} \bigcap \limits _{n\in {\mathbb {N}}} B_X (x_{\alpha ,n}, \lambda q_{\alpha ,n} )\ne \emptyset , \end{aligned}$$

because the intersection on the right side contains at most \({\text {dens}}(X)\) different balls. \(\square \)

3 Main results

We will use the following notation:

Let X be a subspace of a Banach space Y, \(\lambda \ge 1\). For all \(y\in Y\)

$$\begin{aligned} \beta _\lambda (y):= \bigcap _{x\in X} B_X (x,\lambda \Vert y-x \Vert ). \end{aligned}$$

Clearly \(\beta _\lambda (y) = \{ y \}\) for \(y\in X\).

$$\begin{aligned}&O_\lambda (X,Y) := \{ y\in Y: \lambda \Vert y+x \Vert \ge \Vert x \Vert \text { for all } x\in X \}. \end{aligned}$$

\(O_\lambda (X,X^{**})\) is denoted by \(O_\lambda (X)\).

Clearly if \(x\in O_\lambda (X,Y)\) and \(\alpha \in {\mathbb {K}}\) (\({\mathbb {K}}\) is a scalar field of X and Y), then \(\alpha x\in O_\lambda (X,Y)\).

Theorem 3.1

For a subspace X of a Banach space Y and \(\lambda \ge 1\) the following are equivalent:

  1. (a)

    X is a \(\lambda -AC\) subspace of Y.

  2. (b)

    For every \(y\in Y\) \(\beta _\lambda (y)\ne \emptyset \).

  3. (c)

    For every \(y\in Y\) there exist \(x\in X\) and \(u\in O_\lambda (X,Y)\) such that \(y=x+u\).

  4. (d)

    For every subspace Z of Y, \(X\subset Z\), \({\text {dim}} (Z/X)=1\), X is \(\lambda -\)complemented in Z.

  5. (e)

    There exists a map (not necessarily linear) \(P:Y\rightarrow X\) such that

    • \(P(\alpha y)=\alpha P(y)\), \(\alpha \in {\mathbb {K}}\), \(y\in Y\);

    • \(P(x+y)=x+P(y)\), \(x\in X\), \(y\in Y\);

    • \(\Vert P(y) \Vert \le \lambda \Vert x \Vert \), \(y\in Y\).

Proof

(a)\(\Rightarrow \)(b) is trivial.

(b)\(\Rightarrow \)(c) Let \(y\in Y\), \(z\in \beta _\lambda (y)\). Then

$$\begin{aligned}&\Vert z-x \Vert \le \lambda \Vert y-x \Vert ,\ x\in X, \end{aligned}$$

so

$$\begin{aligned}&\Vert x \Vert \le \lambda \Vert y-z+x \Vert , x\in X. \end{aligned}$$

Hence \(u:=y-z\in O_\lambda (X,Y)\).

(c)\(\Rightarrow \)(d) Let Z be a subspace of Y such that \(X\subset Z\) and \({\text {dim}} (Z/X)=1\). We have \(Z={\text {span}}{\{ X\cup \{y_0 \}\}}\) for some \(y_0 \in Y\). From (c) there exist \(x_0\in X\), \(u_0\in O_\lambda (X,Y)\) such that \(y_0 =x_0 +u_0\). Hence \(Z=X\oplus {\mathbb {K}}u_0\). Let \(P:Z\rightarrow X\) be a map given by \(P(x+\alpha u_0)=x\), \(x\in X\), \(\alpha \in {\mathbb {K}}\). Since

$$\begin{aligned}&\Vert x+\alpha u_0 \Vert =\Vert x \Vert \le \lambda \Vert x+\alpha u_0 \Vert ,\ x\in X,\ \alpha \in {\mathbb {K}}, \end{aligned}$$

P is a norm \(\lambda \) projection.

(d)\(\Rightarrow \)(b) Let \(y\in Y{\setminus } X\), \(Z={\text {span}}\{ X\cup \{ y \} \}\). From (d) there exists a norm \(\lambda \) projection \(P:Z\rightarrow X\). Hence

$$\begin{aligned}&\Vert P(y)-x \Vert =\Vert P(y-x) \Vert \le \lambda \Vert y-x \Vert ,\ x\in X, \end{aligned}$$

so \(P(y)\in \beta _\lambda (y)\).

(c)\(\Rightarrow \)(e) Let \(z\in O_\lambda (X,Y)\), \(X_z:= X\oplus {\mathbb {K}}z\) and \(P_z:X_z\rightarrow X\) be a norm \(\lambda \) projection. From (c) we have \(Y=\bigcup \nolimits _{z\in O_\lambda (X,Y)} X_z\). For \(z_1,z_2\in O_\lambda (X,Y)\) we also have \(X_{z_1} = X_{z_2}\) or \(X_{z_1} \cap X_{z_2} = X\), so if \(X_{z_1} = X_{z_2}\), then we can assume that \(P_{z_1}=P_{z_2}\). Then we can define \(P:Y\rightarrow X\) by the formula \(P(y):=P_z (y)\) if \(y\in X_z\). It is easy to observe that \(\Vert P(y) \Vert \le \lambda \Vert y \Vert \), \(y\in Y\) and \(P(\alpha y)=\alpha P(y)\), \(\alpha \in {\mathbb {K}}\), \(y\in Y\). We also have

$$\begin{aligned} P(x+y)&=P_z (x+y)=P_z (x) +P_z (y)=x+P_z(y)\\&=x+P(y),\ x\in X,\ y\in X_z. \end{aligned}$$

(e)\(\Rightarrow \)(a) Let \(x_\alpha \in X\), \(r_\alpha >0\) for \(\alpha \in I\), a family \(\{ B_Y (x_\alpha ,r_\alpha ):\alpha \in I \}\) intersects in Y, and let \(z\in \bigcap \nolimits _{\alpha \in I} B_Y (x_\alpha ,r_\alpha )\). Let \(P:Y\rightarrow X\) be a map from (e). Then

$$\begin{aligned}&\Vert P(z)-x_\alpha \Vert =\Vert P(z-x_\alpha ) \Vert \le \lambda \Vert z-x_\alpha \Vert \le \lambda r_\alpha ,\ \alpha \in I, \end{aligned}$$

so \(P(z)\in \bigcap \nolimits _{\alpha \in I } B_X (x_\alpha ,\lambda r_\alpha )\). \(\square \)

From Theorems 2.33.1 (d) and Lemma 2.8 we get

Corollary 3.2

Let X be a Banach space, \(\lambda \ge 1\), S be a commutative cancellative semigroup of torsion-free rank \(\delta \), \({\text {dens}}(X)=\max \{ \delta ,\omega \}\). If there exists an X-valued invariant \(\lambda \)-mean \(M:\ell _{\infty } (S,X)\rightarrow X\), then X has \(\lambda -IP_{f,\infty }\).

Remark 3.3

It is known (Sobczyk [19]) that there is no projection of \(\ell ^{\infty }\) on \(c_0\). On the other hand for every separable subspace Y of \(\ell ^{\infty }\), \(c_0 \subset Y\), there exists a projection of norm not greater than 2 from Y onto \(c_0\) (and 2 is a best constant for the norm of this projection). From Lemma 2.8, Theorems 3.1, 2.3 we obtain that \(c_0\) has \(2-IP_{f,\infty }\), so in general \(\lambda -IP_{f,\infty }\) and \(\lambda \)-complementability aren’t equivalent. Moreover, there doesn’t exist any \(c_0\)-valued invariant \(\lambda \)-mean from any semigroup of countable non-zero torsion-free rank for any \(\lambda \ge 1\) (so the converse of Theorem 2.3 doesn’t hold). Indeed, if \(M:\ell _{\infty } (S_\omega ,c_0)\rightarrow c_0\) is a \(c_0\)-valued invariant \(\lambda \)-mean, then we define the map \(P:\ell ^\infty \rightarrow c_0\) given by \(P(y):=M({\widetilde{y}})\), where \({\widetilde{y}}(A)=\sum \nolimits _{n\in A} y_n e_n\), \(y\in \ell ^\infty \), A is a finite subset of \({\mathbb {N}}\). It is easy to see that P is a norm \(\lambda \) projection which is impossible.

Remark 3.4

Now we can see that we have the following implications and open problems:

figure a