Abstract

We study the interplay between the order structure and the 𝑝-operator space structure of Figà-Talamanca-Herz algebra 𝐴𝑝(𝐺) of a locally compact group 𝐺. We show that for amenable groups, an order and algebra isomorphism of Figà-Talamanca-Herz-algebras yields an isomorphism or anti-isomorphism of the underlying groups. We also give a bound for the norm of a 𝑝-completely positive linear map from Figà-Talamanca-Herz algebra to its dual space.

1. Introduction and Preliminaries

Throughout this paper, 𝐺 is a locally compact group, 𝑝 is a real number in (1,) and 𝑞(1,) is the conjugate of 𝑝, that is, 1/𝑝+1/𝑞=1. The Fourier algebra 𝐴(𝐺) consists of all coefficient functions of the left regular representation 𝜆 of 𝐺𝐴(𝐺)=𝑤=(𝜆𝜉,𝜂)𝜉,𝜂𝐿2(𝐺).(1.1) This is a Banach algebra with the norm 𝑤𝐴(𝐺)=inf{𝜉2𝜂2𝑤=(𝜆𝜉,𝜂)} [1]. When 𝐺 is abelian, the Fourier transform yields an isometric isomorphism from 𝐴(𝐺) onto 𝐿1(𝐺), where 𝐺 is the Pontryagin dual of 𝐺. In general, 𝐴(𝐺) is a two-sided closed ideal of the Fourier-Stieltjes algebra 𝐵(𝐺) [1]. This is the linear span of the set 𝑃(𝐺) of all positive definite continuous functions on 𝐺.

In [2], Figà-Talamanca introduced a natural generalization of the Fourier algebra, for a compact abelian group 𝐺, by replacing 𝐿2(𝐺) by 𝐿𝑝(𝐺). In [3], Herz extended the notion to an arbitrary group, leading to the commutative Banach algebra 𝐴𝑝(𝐺), called the Figà-Talamanca-Herz algebra. In many ways, this algebra behaves like the Fourier algebra. For example, Leptin's theorem remains valid, namely, 𝐺 is amenable if and only if 𝐴𝑝(𝐺) has a bounded approximate identity [4]. The 𝑝-analog, 𝐵𝑝(𝐺) of the Fourier-Stieltjes algebra is defined as the multiplier algebra of 𝐴𝑝(𝐺), by some authors in [5, 6]. In this paper, we follow [7] for the definition of 𝐵𝑝(𝐺).

This paper investigates the order structure of 𝐴𝑝(𝐺). In an earlier paper, the authors studied the order structure of the Fourier algebra 𝐴(𝐺) [8]. Here, we first introduce a positive cone on 𝐴𝑝(𝐺), then we show that for locally compact amenable groups 𝐺1 and 𝐺2, if 𝐴𝑝(𝐺1) and 𝐴𝑝(𝐺2) are order and algebra isomorphic, then 𝐺1 and 𝐺2 are isomorphic or anti-isomorphic (Theorem 3.2(ii)). This extends a result of Arendt and Cannière on the Fourier algebra in the amenable group case [9].

1.1. 𝑝-Operator Spaces

In this section, we give a brief introduction to the notion of 𝑝-operator spaces [7]. Let 𝑛,  𝑝(1,), and let 𝐸 be a vector space. We denote the vector space of 𝑛×𝑚 matrices with entries from 𝐸 by 𝕄𝑛,𝑚(𝐸). We put simply 𝕄𝑛,𝑚=𝕄𝑛,𝑚(). The space 𝕄𝑛=𝕄𝑛,𝑛 is equipped with the operator norm ||𝑛 from its canonical action on 𝑛-dimensional 𝐿𝑝-space, 𝑛𝑝. Clearly, 𝕄𝑛 acts on 𝕄𝑛(𝐸) by matrix multiplication. For a square matrix 𝑎=(𝑎𝑖𝑗)𝕄𝑛, we have 𝑎𝐵(𝑛𝑝)=sup𝑛𝑖=1|||||𝑛𝑗=1𝑎𝑖𝑗𝑥𝑗|||||𝑝1/𝑝𝑥𝑗,𝑛𝑗=1||𝑥𝑗||𝑝1.(1.2)

Definition 1.1. Let 𝐸 be a vector space. A 𝑝-matricial norm on 𝐸 is a family (𝑛)𝑛=1 such that for each 𝑛,𝑛 is a norm on 𝕄𝑛(𝐸) satisfying ||𝜆||𝜆𝑥𝜇𝑥𝑛||𝜇||,𝑥𝑦𝑛+𝑚=max𝑥𝑛,𝑦𝑚,(1.3) for each 𝜆𝕄𝑚,𝑛,𝜇𝕄𝑛,𝑚,𝑥𝕄𝑛(𝐸), and 𝑦𝕄𝑚(𝐸). Here, 𝜆𝑥𝜇 is the obvious matrix product, and |𝜆| and |𝜇| are the norms of 𝜆 and 𝜇 as the members of (𝑛𝑝,𝑚𝑝) and (𝑚𝑝,𝑛𝑝), respectively.
The vector space 𝐸 equipped with a 𝑝-matricial norm (𝑛)𝑛=1 is called a 𝑝-matricial normed space. If moreover, each (𝕄𝑛(𝐸),𝑛) is a Banach space, 𝐸 is called an (abstract) 𝑝-operator space.

Clearly, 2-operator spaces are the same as classical operator spaces. For more details about operator spaces see [1012].

Definition 1.2. Let 𝐸 and 𝐹 be 𝑝-operator spaces, and let 𝑇(𝐸,𝐹), then for each 𝑛, 𝑇(𝑛)𝕄𝑛(𝐸)𝕄𝑛(𝐹),𝑇(𝑛)𝑥𝑖𝑗=𝑇𝑥𝑖𝑗(1.4) is the 𝑛th amplification of 𝑇. The map 𝑇 is called 𝑝-completely bounded if 𝑇𝑝cb𝑇=sup(𝑛)<.(1.5) If 𝑇𝑝cb1, we say that 𝑇 is a 𝑝-complete contraction, and if 𝑇(𝑛) is an isometry, for each 𝑛, we call 𝑇 a 𝑝-complete isometry.

By [13, Section 4], the collection 𝒞𝑝(𝐸,𝐹) of all 𝑝-completely bounded maps from 𝐸 to 𝐹 is a Banach space under 𝑝cb and a 𝑝-operator space through the identification 𝕄𝑛𝒞𝑝(𝐸,𝐹)=𝒞𝑝𝐸,𝕄𝑛(𝐹)(𝑛).(1.6)

Figà-Talamanca-Herz algebras are our main examples of 𝑝-operator spaces, studied in [13, 14]. For any function 𝑓𝐺 we define 𝑓𝐺 by 𝑓(𝑥)=𝑓(𝑥1),  𝑥𝐺. The Figà-Talamanca-Herz algebra 𝐴𝑝(𝐺) consists of those functions 𝑓𝐺 for which there are sequences (𝜉𝑛)𝑛=1 and (𝜂𝑛)𝑛=1 in 𝐿𝑞(𝐺) and 𝐿𝑝(𝐺), respectively, such that 𝑓=𝑛=1𝜉𝑛̃𝜂𝑛 and 𝑛=1𝜉𝑛𝑞𝜂𝑛𝑝<.(1.7)

The norm 𝑓𝐴𝑝(𝐺) of 𝑓𝐴𝑝(𝐺) is defined as the infimum of the above sums over all possible representations of 𝑓. Then 𝐴𝑝(𝐺) is a Banach space which is embedded contractively in 𝐶0(𝐺). It was shown by Herz that 𝐴𝑝(𝐺) is a Banach algebra under pointwise multiplication. When 𝑝=2, we get the Fourier algebra 𝐴(𝐺).

Let 𝜆𝑝𝐺(𝐿𝑝(𝐺)) be the left regular representation of 𝐺 on 𝐿𝑝(𝐺), defined by 𝜆𝑝(𝑠)(𝑓)(𝑡)=𝑓(𝑠1𝑡). Then 𝜆𝑝 can be lifted to a representation of 𝐿1(𝐺) on 𝐿𝑝(𝐺). The algebra of pseudomeasures PM𝑝(𝐺) is defined as the 𝑤-closure of 𝜆𝑝(𝐿1(𝐺)) in (𝐿𝑝(𝐺)). There is a canonical duality PM𝑝(𝐺)𝐴𝑝(𝐺) via𝜉̃𝜂,𝑇=𝜉,𝑇(𝜂)𝜉𝐿𝑝(𝐺),𝜂𝐿𝑞(𝐺),𝑇PM𝑝(𝐺).(1.8) In particular, PM2(𝐺) is the group von Neumann algebra VN(𝐺). If the map Λ𝑝 from the projective tensor product 𝐿𝑞(𝐺)𝛾𝐿𝑝(𝐺) to 𝐶0(𝐺) is defined byΛ𝑝(𝑔𝑓)(𝑠)=𝑔,𝜆𝑝(𝑠)𝑓,(1.9) for 𝑔𝐿𝑞(𝐺),𝑓𝐿𝑝(𝐺), and 𝑠𝐺, then 𝐴𝑝(𝐺) is isometrically isomorphic to 𝐿𝑞(𝐺)𝛾𝐿𝑝(𝐺)/kerΛ𝑝 and 𝐴𝑝(𝐺)=𝐿𝑇𝑝(𝐺)𝑇|kerΛ𝑝=0.(1.10)

1.2. Complexification of Ordered Vector Spaces

We often consider vector spaces and algebras over the complex field. It is therefore desirable to have the notion of a complex ordered vector space. This is usually done through the complexification of real ordered spaces. We recall some basic constructions in the theory of complexification. For more details, see [15].

If 𝐸 is a real vector space, then the complexification 𝐸 of 𝐸 is the additive group 𝐸×𝐸 with scalar multiplication defined by (𝛼,𝛽)(𝑥,𝑦)=(𝛼𝑥𝛽𝑦,𝛽𝑥+𝛼𝑦), for (𝛼,𝛽)=𝛼+𝑖𝛽. Each 𝑧𝐸×𝐸 is uniquely represented as 𝑧=𝑥+𝑖𝑦, where 𝑥,𝑦𝐸. Thus, 𝐸 can be written as 𝐸+𝑖𝐸.

For real vector spaces 𝐸 and 𝐹 with complexifications 𝐸 and 𝐹, every -linear map 𝑇𝐸𝐹 has a unique -linear extension 𝑇 given by 𝑇(𝑧)=𝑇𝑥+𝑖𝑇𝑦, for 𝑧=𝑥+𝑖𝑦𝐸, where 𝑥,𝑦𝐸. The map 𝑇 is the canonical extension of 𝑇 (usually again denoted by 𝑇).

A real vector space 𝐸, endowed with an order relation ≤, is called a real ordered space if(i)𝑥𝑦 implies 𝑥+𝑧𝑦+𝑧, for all 𝑥,𝑦,𝑧𝐸(ii)𝑥𝑦 implies 𝛼𝑥𝛼𝑦, for all 𝑥,𝑦𝐸 and 𝛼+.

The subset 𝐸+={𝑥𝐸0𝑥} is called the positive cone of the real ordered space 𝐸. In general, a cone is a subset 𝑃 of 𝐸 such that 𝑥+𝑦𝑃 and 𝛼𝑥𝑃, for 𝑥,𝑦𝑃 and 𝛼+. Then 𝑃 defines an order structure on 𝐸 by 𝑥𝑦 if and only if 𝑦𝑥𝑃. A cone 𝑃 is called proper if 𝑃(𝑃)={0}. If a normed space 𝐸 is an ordered space with the positive cone 𝐸+, then there is a natural order structure on its dual space 𝐸. The positive cone of 𝐸 is defined as 𝐸+=𝑓𝐸𝑓(𝑝)0𝑝𝐸+.(1.11)

A complex vector space is called an ordered space, if it is the complexification of a real ordered vector space. A (complex) Banach space 𝐴 is called a Banach ordered space if it is the complexification of a real ordered space 𝐵 which is also a real Banach space, such that:(i)the inclusion map 𝑖𝐵𝐴 is an isometry,(ii)each element 𝑎𝐴 can be written as 𝑎=𝑎1𝑎2+𝑖(𝑎3𝑎4), where 𝑎1,,𝑎4 are positive in 𝐴 and 𝑎𝑗𝑎, for 𝑗=1,,4.

Banach lattices, 𝐶-algebras, and their duals are Banach ordered spaces.

Given ordered spaces 𝐸 and 𝐹, a -linear map 𝑇𝐸𝐹 is positive if 𝑇(𝐸)𝐹 and the restriction 𝑇|𝐸𝐸𝐹 is positive.

1.3. Order Structure of 𝑝-Operator Spaces

Each operator space 𝐸 can be embedded in (), for some Hilbert space , by Ruan's Theorem [10, Theorem 2.3.5], and the order structure of 𝐸 is induced by () [8] (see also [16]). We have a 𝑝-analog of Ruan's theorem for 𝑝-operator spaces [17, Theorem 4.1] which asserts that each 𝑝-operator space 𝐸 is 𝑝-complete isometrically embedded in () for some QS𝐿𝑝 space , which again induces an order structure on 𝐸. The main challenge is of course that the powerful methods from 𝐶-algebras and von Neumann algebras are no longer at one's disposal for 𝑝2.

In this paper, we confine ourselves to the special case of =𝐿𝑝(𝑋), where 𝑋 is a measure space and 1<𝑝<. We say that 𝑇(𝐿𝑝(𝑋)) is positive if 𝑇𝑓,𝑓0 for each 𝑓𝐿𝑝(𝑋)𝐿𝑞(𝑋), where the pairing is the canonical dual action of 𝐿𝑞(𝑋) on 𝐿𝑝(𝑋). For 𝑝=2, this order is the natural order on the 𝐶-algebra (𝐿2(𝑋)). We also put an order on subspaces of 𝕄𝑛((𝐿𝑝(𝑋))) for 𝑛1. We say 𝑇𝕄𝑛((𝐿𝑝(𝑋)))(𝐿𝑛𝑝(𝑋)) is positive if 𝑛𝑖,𝑗=1𝑇𝑓𝑖,𝑓𝑗0, for each 𝑓1,,𝑓𝑛𝐿𝑝(𝑋)𝐿𝑞(𝑋). It is easy to see that these orders define proper cones on (𝐿𝑝(𝑋)) and 𝕄𝑛((𝐿𝑝(𝑋))), respectively. Also for 𝑇=[𝑇𝑖𝑗]𝕄𝑛((𝐿𝑝(𝑋))), we say 𝑇 is positive, if for each 𝑚 and 𝜙=[𝜙𝑖𝑗]𝕄𝑚((𝐿𝑝(𝑋)))+, the natural matrix action 𝜙,𝑇 is positive.

2. Order Structure of Some 𝑝-Operator Spaces

2.1. The 𝑝-Fourier Stieltjes Algebra

The Fourier-Stieltjes algebra 𝐵(𝐺) was defined by Eymard as the algebra of coefficient functions 𝑥𝜋(𝑥)𝜉,𝜂 of unitary representations 𝜋 of 𝐺 on a Hilbert space , where 𝜉,𝜂 [1]. In this section, we consider the 𝑝-analog 𝐵𝑝(𝐺) of the Fourier-Stieltjes algebra, introduced by Runde [7] and study its order structure given by the 𝑝-analog 𝑃𝑝(𝐺) of positive definite continuous functions.

Definition 2.1. A representation of 𝐺 on a Banach space is a pair (𝜋,), where 𝜋 is a group homomorphism from 𝐺 into the group of invertible isometries on which is continuous with respect to the given topology on 𝐺 and the strong operator topology on ().

Definition 2.2. A Banach space is called (i)an 𝐿𝑝-space if it is of the form 𝐿𝑝(𝑋), for some measure space 𝑋,(ii)a QS𝐿𝑝-space if it is isometrically isomorphic to a quotient of a subspace of an 𝐿𝑝-space (or equivalently, a subspace of a quotient of an 𝐿𝑝-space [7, Section 1, Remark 1]).

We denote by Rep𝑝(𝐺), the collection of all (equivalence classes of) representations of 𝐺 on QS𝐿𝑝-spaces.

Definition 2.3. Let 𝐺 be a locally compact group and let (𝜋,) be a representation of 𝐺. A coefficient function of (𝜋,) is a function 𝑓𝐺 of the form 𝑓(𝑥)=𝜋(𝑥)𝜉,𝜙(𝑥𝐺),(2.1) where 𝜉 and 𝜙. Define 𝐵𝑝(𝐺)=𝑓𝐺𝑓isacoecientfunctionofsome(𝜋,𝐸)Rep𝑝(𝐺).(2.2) For 𝑓𝐵𝑝(𝐺), put 𝑓=inf{𝜉𝜙𝑓()=𝜋()𝜉,𝜙}.

Using a suitable definition of tensor product of QS𝐿𝑝-spaces, it is shown in [7] that 𝐵𝑝(𝐺) is a commutative unital Banach algebra with the pointwise multiplication which contains 𝐴𝑝(𝐺) contractively as a closed ideal. Also we know that 𝐴𝑝(𝐺) can be embedded in 𝐵𝑝(𝐺) isometrically if 𝐺 is amenable [7, Corollary 5.3].

A compatible couple of Banach spaces in the sense of interpolation theory (see [18]) is a pair (0,1) of Banach spaces such that both 0 and 1 are embedded continuously in some (Hausdorff) topological vector space. In this case, the intersection 01 is again a Banach space under the norm (0,1)=max{0,1}. For example, the pairs (𝐴𝑝(𝐺),𝐴𝑞(𝐺)) and (𝐿𝑝(𝐺),𝐿𝑞(𝐺)) are compatible couples.

Definition 2.4. Let (𝜋,) be a representation of 𝐺 such that (,) is a compatible couple. Then a 𝜋-positive definite function on 𝐺 is a function which has a representation as 𝑓(𝑥)=𝜋(𝑥)𝜉,𝜉(𝑥𝐺), where 𝜉. We call each element in the closure of the set of all 𝜋-positive definite functions on 𝐺 in 𝐵𝑝(𝐺), where 𝜋 is a representation of 𝐺 on an 𝐿𝑝-space, a 𝑝-positive definite function on 𝐺, and the set of all 𝑝-positive definite functions on 𝐺 will be denoted by 𝑃𝑝(𝐺).

It follows from [7, Lemma 4.3] and the definition of 𝑃𝑝(𝐺) that for each 𝑓𝑃𝑝(𝐺), associated to a representation (𝜋,), there exist a sequence (𝜋𝑛,𝑛)𝑛=1 of cyclic representations of 𝐺 on closed subspaces 𝑛 of , and {𝜉𝑛} in 𝑛, such that 𝑓(𝑥)=𝑛=1𝜋𝑛(𝑥)𝜉𝑛,𝜉𝑛(𝑥𝐺).(2.3)

The closed subspace 𝐵𝑝𝑝(𝐺) of 𝐵𝑝(𝐺) is the closure of the set of all functions 𝑓𝐵𝑝(𝐺) of the form 𝑓(𝑥)=𝜋(𝑥)𝜉,𝜂,𝑥𝐺, for some representation (𝜋,), where is an 𝐿𝑝-space, 𝜉, and 𝜂.

Proposition 2.5. The linear span of 𝑃𝑝(𝐺) is dense in 𝐵𝑝𝑝(𝐺).

Proof. Let 𝑢𝐵𝑝𝑝(𝐺) have a representation as 𝑢(𝑥)=𝜋(𝑥)𝜉,𝜂,𝑥𝐺, where 𝜋 is a representation on some 𝐿𝑝-space ,𝜉 and 𝜂. It is clear that is dense in both and . Hence, there exist sequences {𝜉𝑛} and {𝜂𝑛} in converging to 𝜉 and 𝜂 in and , respectively. Put 𝑢𝑛(𝑥)=𝜋(𝑥)𝜉𝑛,𝜂𝑛, then 𝑢𝑢𝑛𝜋(𝑥)=(𝑥)𝜉,𝜂𝜂𝑛+𝜋(𝑥)𝜉𝜉𝑛,𝜂𝑛.(2.4) By the definition of the norm in 𝐵𝑝(𝐺), we have 𝑢𝑢𝑛𝜉𝑝𝜂𝜂𝑛𝑞+𝜉𝜉𝑛𝑝𝜂𝑛𝑞,(2.5) which tends to zero, as 𝑛.
Now, it is enough to consider the following decomposition for 𝑢𝑛, 𝜋(𝑥)𝜉𝑛,𝜂𝑛1=43𝑗=0𝜉𝜋(𝑥)𝑛+𝑖𝑗𝜂𝑛,𝜉𝑛+𝑖𝑗𝜂𝑛,(2.6) where 𝑖=1, and note that each 𝜉𝑛+𝑖𝑗𝜂𝑛, for 𝑗=0,1,2,3, belongs to . So 𝑢 belongs to the closed linear span of 𝑃𝑝(𝐺).

Each representation (𝜋,) of 𝐺 induces a representation of the group algebra 𝐿1(𝐺) on via 𝜋(𝑓)=𝐺𝑓(𝑥)𝜋(𝑥)𝑓𝐿1(𝐺).(2.7) Let (𝜋𝑝,) be the 𝑝-universal representation of 𝐺. The Banach space UPF𝑝(𝐺) of all pseudofunctions on 𝐺 is the closure of 𝜋𝑝(𝐿1(𝐺)) in 𝐵(). By [7, Theorem 6.6], we know that 𝐵𝑝(𝐺) is the dual space of UPF𝑝(𝐺) via the pairing 𝜋𝑝(=𝑓),𝑔𝐺𝑓(𝑥)𝑔(𝑥)𝑓𝐿1(𝐺),𝑔𝐵𝑝(𝐺).(2.8)

We say that 𝑓UPF𝑝(𝐺) is positive if 𝜙𝑔(𝑓)0, for all 𝑔𝐵𝑝(𝐺)+=𝑃𝑝(𝐺), where 𝜙𝑔 is the corresponding linear functional of 𝑔. Let 𝐶(𝐺) be the full group 𝐶-algebra of 𝐺, which is the enveloping 𝐶-algebra of 𝐿1(𝐺) [1].

Proposition 2.6. There exists a positive contraction from UPF𝑝(𝐺) to 𝐶(𝐺).

Proof. Since each Hilbert space is a QS𝐿𝑝-space [3] and for each 𝑓𝐿1(𝐺),𝑓Rep2𝑓Rep𝑝, where 𝑓Rep𝑝=sup𝜋Rep𝑝(𝐺),𝜋1𝜋(𝑓), it follows that the identity map 𝐿𝑖1(𝐺),Rep𝑝𝐿1(𝐺),Rep2(2.9) is continuous. Thus, it induces a continuous map 𝑖𝑐UPF𝑝(𝐺)𝐶(𝐺).(2.10) Consider the conjugate map 𝑖𝑐𝐵(𝐺)𝐵𝑝(𝐺).(2.11) Since for each 𝑢𝐵(𝐺) and 𝑓𝐿1(𝐺), 𝑖𝑐(𝑢),𝑓=𝑢𝑖,𝑓=𝑢,𝑓,(2.12)(𝑖𝑐) is the inclusion map. By the definition of positivity in 𝐵𝑝(𝐺) and 𝐵(𝐺),(𝑖𝑐) is a positive map. Therefore (𝑖𝑐) is positive, and so is 𝑖𝑐=(𝑖𝑐)UPF𝑝(𝐺).

2.2. Figà-Talamanca-Herz Algebra

In this section, we study the order structure of the Figà-Talamanca-Herz algebra 𝐴𝑝(𝐺). Since 𝐴𝑝(𝐺) is the set of coefficient functions of the left regular representation of 𝐺 on 𝐿𝑝(𝐺), we have 𝐴𝑝(𝐺)𝐵𝑝𝑝(𝐺). We define the positive cone of 𝐴𝑝(𝐺) as the closure in 𝐴𝑝(𝐺), of the set of all function of the form 𝑓=𝑛𝑖=1𝜉𝑖̃𝜉𝑖, for a sequence (𝜉𝑖) in 𝐿𝑝(𝐺)𝐿𝑞(𝐺), and denote it by 𝐴𝑝(𝐺)+. It is clear that 𝐴𝑝(𝐺)+ is contained in 𝐵𝑝(𝐺)+. Since 𝐶𝑐(𝐺)𝑃(𝐺)=𝐴(𝐺)𝑃(𝐺), this order structure, in the case where 𝑝=2, is the same as the order structure of 𝐴(𝐺), induced by the set 𝑃(𝐺)𝐴(𝐺) as a positive cone.

Clearly, 𝑇PM𝑝(𝐺) is positive as an element of 𝐵(𝐿𝑝(𝐺)) if and only if it is positive as an element of 𝐴𝑝(𝐺)+, where 𝐴𝑝(𝐺)+ is the dual cone induced by the positive cone 𝐴𝑝(𝐺)+. Also, since 𝐴𝑝(𝐺)+ is closed, 𝑢𝐴𝑝(𝐺)+ if and only if for each 𝑇PM𝑝(𝐺)+,  𝑇(𝑢)0 [19].

Theorem 2.7. Let 𝐺 be a locally compact amenable group. Then 𝐴𝑝(𝐺)+=𝑃(𝐺)𝐴(𝐺)={𝑢𝐴(𝐺)𝑢𝐴𝑝(𝐺)=𝑢(𝑒)}, where 𝑒 is the identity of 𝐺. In particular, 𝐴𝑝(𝐺)+ is a proper cone.

Proof. Since 𝐿𝑝(𝐺)𝐿𝑞(𝐺)𝐿2(𝐺), it follows that 𝐴𝑝(𝐺)+𝑃(𝐺)𝐴(𝐺)𝐴𝑝(𝐺). But when 𝐺 is amenable, the identity map 𝑖𝐴(𝐺)𝐴𝑝(𝐺) is an embedding with 𝐴𝑝(𝐺)𝐴(𝐺) [3, Theorem C]. For each 𝑢𝑃(𝐺)𝐴(𝐺), we have 𝑢𝐴(𝐺)=𝑢(𝑒)𝑢𝑢𝐴𝑝(𝐺)𝑢𝐴(𝐺),(2.13) where is the supremum norm on 𝐴(𝐺), that is, 𝑢𝐴(𝐺)=𝑢𝐴𝑝(𝐺). In particular, 𝑃(𝐺)𝐴(𝐺) is closed in 𝐴𝑝(𝐺) and consequently 𝐴𝑝(𝐺)+𝑃(𝐺)𝐴(𝐺). Now, let 𝑣 be an arbitrary element of 𝐴(𝐺)𝑃(𝐺). Then there exists a sequence in 𝑃(𝐺)𝐶𝑐(𝐺) of the form {𝑓𝑖𝑓𝑖},𝑓𝑖𝐶𝑐(𝐺),𝑖 converging to 𝑣 in 𝐴(𝐺). Since for each 𝑢𝐴(𝐺),𝑢𝐴𝑝(𝐺)𝑢𝐴(𝐺), it follows that {𝑓𝑖𝑓𝑖} converges to 𝑣 in 𝐴𝑝(𝐺). Clearly, {𝑓𝑖𝑓𝑖} is contained in 𝐴𝑝(𝐺)+ and since 𝐴𝑝(𝐺)+ is closed, 𝑣𝐴𝑝(𝐺)+.
The second equality follows immediately from the above-mentioned fact that for each 𝑢𝑃(𝐺)𝐴(𝐺),  𝑢𝐴(𝐺)=𝑢𝐴𝑝(𝐺).

3. Order Maps between 𝑝-Operator Spaces

In this section, we study the positive maps between various 𝑝-operator spaces. Also, we give the general form of the algebra and order isomorphisms between Figà-Talamanca-Herz algebras.

Proposition 3.1. Let 𝐸 be a Banach ordered space with a closed positive cone 𝐸+, and, 𝐹 be a normed space and an ordered space such that for 𝑥,𝑦𝐹, 0𝑥𝑦 implies 𝑥𝑦. Then every positive linear map from 𝐸 into 𝐹 is continuous.

Proof. Assume towards a contradiction that 𝑇𝐸𝐹 is a positive linear map that is not continuous. Then 𝑇 is unbounded on the unit ball 𝑈 of 𝐸, and hence, on 𝑈+=𝑈𝐸+ since 𝑈𝑈+𝑈++𝑖(𝑈+𝑈+). This implies that there exists a sequence {𝑥𝑛}𝑛=1 in 𝑈+ such that 𝑇𝑥𝑛𝑛3, for each 𝑛. Since 𝐸+ is closed, 𝑥𝑧=𝑛/𝑛2 is in 𝐸+. Hence, 𝑇𝑧𝑇𝑥𝑛/𝑛2>0, for each 𝑛. Therefore, 𝑇𝑧𝑛, for each 𝑛, which is impossible.

We note that the above proposition remains true for a Banach space 𝐸 with a closed positive cone 𝐸+ satisfying the following property: for each 𝑥𝐸, there is a sequence {𝑥𝑛} converging to 𝑥, such that 𝑥𝑛=𝑥1𝑛𝑥2𝑛+𝑖(𝑥3𝑛𝑥4𝑛) with 𝑥𝑖𝑛𝐸+ and 𝑥𝑖𝑛𝑥𝑛.

A linear map 𝑇between two ordered spaces is called an order isomorphism if 𝑇 is one-to-one and onto, and moreover 𝑇 and 𝑇1 are both positive maps.

Theorem 3.2. Let 𝐺1 and 𝐺2 be amenable locally compact groups with identities 𝑒1 and 𝑒2, respectively, and let 𝑇𝐴𝑝(𝐺1)𝐴𝑝(𝐺2) be an order and algebra isomorphism. Then (i)𝑇(𝜆𝑒2)=𝜆𝑒1, where 𝜆𝑒𝑖 is the evaluation homomorphism at 𝑒𝑖,  𝑖=1,2,(ii)there exists an isomorphism or anti-isomorphism 𝜑 from 𝐺2 onto 𝐺1 such that 𝑇(𝑓)=𝑓𝜑 for all 𝑓𝐴𝑝(𝐺1).

Proof. (i) Consider the adjoint map 𝑇𝐴𝑝(𝐺2)𝐴𝑝(𝐺1), which is clearly an order isomorphism. Since 𝑇 is an algebra isomorphism, 𝑇(𝜆𝑒2) is a multiplicative linear functional on 𝐴𝑝(𝐺1), and since for each locally compact group 𝐺, any (non zero) multiplicative linear functional on 𝐴𝑝(𝐺)is an evaluation homomorphism at some point of 𝐺, it follows that, 𝑇(𝜆𝑒2)=𝜆𝑥, for some 𝑥𝐺1. We note that 𝜆𝑥 is positive because 𝜆𝑒2 is a positive element of 𝐴𝑝(𝐺), hence, 𝑥=𝑒1 by [20, Proposition 4.3].
(ii) We first note that 𝐴𝑝(𝐺𝑖)+=𝑃(𝐺𝑖)𝐴(𝐺𝑖),  𝑖=1,2, by Theorem 2.7. Hence 𝑇 maps 𝑃(G1)𝐴(𝐺1) onto 𝑃(𝐺2)𝐴(𝐺2), and since for 𝑖=1,2,  𝑃(𝐺𝑖)𝐴(𝐺𝑖) spans 𝐴(𝐺𝑖), it follows that the restriction map 𝑇1=𝑇|𝐴(𝐺1) is an order and algebra isomorphism from 𝐴(𝐺1) onto 𝐴(𝐺2). Now, [9, Theorem 3.2] implies that there exists an isomorphism or anti-isomorphism 𝜓𝐺2𝐺1 such that 𝑇1(𝑓)=𝑓𝜓 for all 𝑓𝐴(𝐺1). Now, the result follows from the density of 𝐴(𝐺1) in 𝐴𝑝(𝐺1) and continuity of 𝑇.

Proposition 3.3. Let 𝐺 and 𝐻 be locally compact groups, and let 𝑇𝐴(𝐺)VN(𝐻) be a completely positive linear map. Then there are infinitely many 𝑛 such that 𝑇(𝑛)𝑛2.

Proof. We first note that 𝑇 is continuous by Proposition 3.1. Now, assume on the contrary that there exists 𝑛0 such that for each 𝑛𝑛0,𝑇(𝑛)>𝑛2. Fix 𝑛𝑛0. Then there is some 𝑥(𝕄𝑛(𝐴(𝐺)))1, the unit ball of 𝕄𝑛(𝐴(𝐺)), such that 𝑇(𝑛)(𝑥)>𝑛2. Since 𝕄𝑛(𝐴(𝐺))=𝒞𝜎(VN(𝐺),𝕄𝑛), where 𝒞𝜎(VN(𝐺),𝕄𝑛) is the algebra of all 𝑤-continuous, completely bounded linear maps from VN(𝐺) to 𝕄𝑛, and since 𝕄𝑛 is an injective von Neumann algebra, by [7, Corollary 2.6] or [4, Theorem 2.1], it follows that 𝑥 can be decomposed into positive elements such that the norm of each element in this decomposition is less than or equal to the norm of 𝑥. Considering 𝑦=(𝑥+𝑥𝑡)/2 instead of 𝑥, without loss of generality, we may assume that 𝑥 is positive, 𝑥1,𝑥=𝑥𝑡, and 𝑇(𝑛)(𝑥)𝑛2/8. Let 𝑇𝑛=𝑛𝑖=1𝑇𝑛𝑖=1𝐴(G)𝑛𝑖=1VN(𝐻) be defined by (𝑧𝑖)(𝑇𝑧𝑖). Then clearly, 𝑇𝑛=𝑛𝑖=1𝑇=sup(𝑧𝑖)1(𝑛𝑖=1𝑇𝑧𝑖2)1/2𝑛𝑇.
Now, 𝑇(𝑛)(𝑥)=[𝑇(𝑥𝑖𝑗)] is positive in 𝕄𝑛(VN(𝐻)), hence, [𝑇(𝑥𝑖𝑗)]=[𝑇(𝑥𝑖𝑗)], and for each 𝑖,𝑗,  𝑇(𝑥𝑖𝑗) is self-adjoint. Hence, 𝑇𝑥𝑖𝑗=𝑇(𝑇𝑥1𝑖𝑗+𝑇𝑥2𝑖𝑗)/2, where for each 𝑖,𝑗{1,,𝑛} and 𝑘=1,2,  𝑇𝑥𝑘𝑖𝑗 is a unitary operator in (𝐿2(𝐺)). For each unit vector (𝑓1,𝑓2,,𝑓𝑛) in 𝑛𝑖=1𝐿2(𝐻), set 𝑇𝑥𝑘𝑖𝑗(𝑓𝑖)=𝑒𝑘𝑖𝑗,  𝑘=1,2,1𝑖,𝑗𝑛. Then 𝑗𝑖𝑇𝑥1𝑖𝑗+𝑇𝑥2𝑖𝑗2𝑓𝑖21/2=𝑗𝑖𝑒1𝑖𝑗+𝑒2𝑖𝑗221/2=12𝑗𝑖𝑒1𝑖𝑗+𝑒2𝑖𝑗,𝑖𝑒1𝑖𝑗+𝑒2𝑖𝑗1/212𝑖,𝑗𝑒1𝑖,𝑗+𝑒2𝑖,𝑗,𝑒1𝑖𝑗+𝑒2𝑖𝑗+4𝑛21/2=12𝑖,𝑗𝑒1𝑖,𝑗+𝑒2𝑖,𝑗2+4𝑛21/2𝑇𝑛2𝑥𝑖𝑗.+𝑛(3.1) Hence, we have 𝑇(𝑛)𝑥𝑖𝑗𝑇𝑇𝑛2𝑥𝑖𝑗+𝑛.(3.2) On the other hand, 𝑥𝕄𝑛(𝐴(𝐺))=sup𝑆𝕄𝑛(VN(𝐺))1𝑆𝑘𝑙𝑥𝑖𝑗=sup𝑆𝕄𝑛(VN(𝐺))1𝑖,𝑗,𝑘,𝑙||𝑆𝑘𝑙𝑥𝑖𝑗||21/2.(3.3) By the Hahn-Banach theorem, for each 𝑥𝑖𝑗, there is some 𝑆𝑖𝑗VN(𝐺)1 such that 𝑆𝑖𝑗(𝑥)=𝑥𝑖𝑗. Now, put 𝑆=[𝑆𝑖𝑗], then 𝑆𝕄𝑛(VN(𝐺))𝑛. Hence, 𝑆/𝑛𝕄𝑛(VN(𝐺))1 and (𝑆𝑖𝑗/𝑛)(𝑥𝑖𝑗)=𝑥𝑖𝑗/𝑛. Therefore, 𝑥𝕄𝑛(𝐴(𝐺))=sup𝑅𝕄𝑛(VN(𝐺))1𝑅𝑘𝑙𝑥𝑖𝑗𝑥𝑖𝑗𝑛21/2=1𝑛𝑥𝑖𝑗21/2.(3.4) This means that the norm of 𝑥 as an element of 𝑛2𝑖=1𝐴(𝐺) is less than or equal to 𝑛. Therefore, by inequality (3.2), 𝑛2𝑇(8𝑇)𝑛𝑛2𝑥𝑖𝑗𝑥𝑇𝑖𝑗𝑛2𝑖=1𝐴(𝐺)𝑛𝑇.(3.5) Hence, 𝑇 is unbounded, which is a contradiction.

Let 1<𝑝<, and let 𝑌 be an arbitrary measure space. For subspaces (𝐿𝑝(𝑋)) and 𝒩(𝐿𝑝(𝑌)), we say that a linear map 𝑇𝒩 is (𝑝,𝑝)-completely positive, if for all 𝑛, the 𝑛th amplification 𝑇(𝑛)𝕄𝑛()𝕄𝑛(𝒩),  𝑇(𝑛)[𝑥𝑖𝑗]=[𝑇(𝑥𝑖𝑗)] of 𝑇 is a positive map. For simplicity in the case where 𝑝=𝑝, we call such maps 𝑝-completely positive.

Recall that for a locally compact group 𝐺, 𝐴𝑝(𝐺) has the natural 𝑝-operator space structure as the predual of PM𝑝(𝐺) [13]. Now, we define an order structure on 𝕄𝑛(𝐴𝑝(𝐺)) as follows. Let 𝑚,𝑛, we say that 𝑇=[𝑇𝑖𝑗]𝕄𝑛(𝐴𝑝(𝐺)) is positive if for every 𝑝-completely positive and 𝑝-completely bounded linear map 𝜙PM𝑝(𝐺)𝕄𝑚, the natural action 𝜙,𝑇 is a positive scalar matrix.

For =𝐴𝑝(𝐺)orPM𝑝(𝐺) and 𝒩=𝐴𝑝(𝐺) or PM𝑝(𝐺), we say that a linear map 𝑇𝒩 is (𝑝,𝑝)-completely positive if for each positive integer 𝑛, the 𝑛th amplification 𝑇𝑛𝕄𝑛()𝕄𝑛(𝒩) of 𝑇 is a positive map. If this is the case for 𝑝=𝑝, we say that 𝑇 is 𝑝- completely positive.

Let 𝑝2. Then since 2 and 𝑝 are equivalent norms on 𝑛, for each 𝑛, there exists a positive scalar 𝛼𝑛 such that 𝛼𝑛2𝑝2 on 𝑛. For the case where 1<𝑝2, for each 𝑛, we choose 𝛽𝑛>0 such that 𝛽𝑛2𝑝 on 𝑛.

Proposition 3.4. Let 𝐺 be an amenable locally compact group, let 𝑝(1,) and let 𝑇𝐴𝑝(𝐺)PMp(𝐺) be a linear map. Let 𝑇1=𝑇|𝐴(𝐺) be the restriction of 𝑇 to 𝐴(𝐺). (i)If 𝑇 is 𝑝-completely positive, then 𝑇1 is a completely positive linear map from 𝐴(𝐺) to VN(𝐺). (ii)If 𝑝2 and 𝑇 is a bounded linear map, then for each 𝑛, we have 𝑇1(𝑛)𝑇(𝑛).(iii)If 1<𝑝2 and 𝑇 is bounded, then for each 𝑛, we have 𝑇(𝑛)𝛽2𝑛𝑇1(𝑛).(iv)If 1<𝑝2 and 𝑇 is a 𝑝-completely positive map, then there are infinitely many 𝑛 such that 𝑇(𝑛)𝛽2𝑛𝑛2.

Proof. (i) Let 𝑇 be 𝑝-completely positive. Let 𝑛 and [𝑥𝑖𝑗]𝕄𝑛(𝐴(𝐺)). Since 𝕄𝑛(𝐴(𝐺))+=𝒞𝒫𝜎(VN(𝐺),𝕄𝑛), where 𝒞𝒫𝜎(VN(𝐺),𝕄𝑛) is the set of all 𝑤-continuous completely positive linear maps from VN(𝐺) to 𝕄𝑛, it follows that [𝑥𝑖𝑗]𝕄𝑛(𝐴(𝐺))+ if and only if for each [𝑇𝑖𝑗]𝕄𝑛[VN(𝐺)]+, we have [𝑇𝑘𝑙(𝑥𝑖𝑗)]𝕄𝑛2+. Since 𝐺 is amenable, the embedding 𝑖𝐴(𝐺)𝐴𝑝(𝐺) is norm decreasing [3]. Therefore, each element [𝑆𝑖𝑗] in 𝕄𝑛[PM𝑝(𝐺)] can be considered as an element of 𝕄𝑛[VN(𝐺)]. Moreover, [𝑆𝑖𝑗]𝕄𝑛[PM𝑝(𝐺)]+ if and only if [𝑆𝑖𝑗][𝑓],[𝑓]0, for all [𝑓]=(𝑓1,,𝑓𝑛) with 𝑓1,,𝑓𝑛𝐶𝑐(𝐺). It is clear that if [𝑆𝑖𝑗]𝕄𝑛[PM𝑝(𝐺)]+, then it also belongs to 𝕄𝑛[VN(𝐺)]+. Therefore, 𝑖𝐴(𝐺)𝐴𝑝(𝐺) is a (2,𝑝)-completely positive map. This implies that 𝑇1𝐴(𝐺)VN(𝐺) is a completely positive linear map.
(ii) As we noted before, the embedding 𝑖𝐴(𝐺)𝐴𝑝(𝐺) is a norm decreasing map with dense range. So, PM𝑝(𝐺) can be considered as a subspace of VN(𝐺). We first show that for each 𝑛, 𝑖(𝑛) is continuous and has dense range. Let [𝑎𝑖𝑗]𝕄𝑛(𝐴(𝐺)), then [𝑆(𝑎𝑖𝑗)]𝐵(𝑛𝑝), for each 𝑆PM𝑝(𝐺) and𝑎𝑖𝑗𝕄𝑛(𝐴𝑝(𝐺))=sup𝑆PM𝑝(𝐺)1𝑆𝑎𝑖𝑗=sup𝑆PM𝑝(𝐺)1sup𝑛𝑖=1|||||𝑛𝑗=1𝑆𝑎𝑖𝑗𝑥𝑗|||||𝑝1/𝑝,𝑥𝑗𝑛,𝑛𝑗=1||𝑥𝑗||𝑝11𝛼𝑛sup𝑆PM𝑝(𝐺)1sup𝑛𝑖=1|||||𝑛𝑗=1𝑆𝑎𝑖𝑗𝑥𝑗|||||21/2,𝑥𝑗𝑛,𝑛𝑗=1||𝑥𝑗||2=11𝛼𝑛sup𝑆PM𝑝(𝐺)1𝑆𝑎𝑖𝑗𝐵(2)1𝛼𝑛𝑎𝑖𝑗𝕄𝑛(𝐴(𝐺)),(3.6) which shows that 𝑖(𝑛) is continuous. Consider now an element [𝑎𝑖𝑗]𝕄𝑛(𝐴𝑝(𝐺)). Then for each 𝑖,𝑗, there exists a sequence (𝑎𝑚𝑖𝑗)𝑚 in 𝐴(𝐺) converging to 𝑎𝑖𝑗 in 𝐴𝑝(𝐺). Hence, 𝑎𝑚𝑖𝑗𝑎𝑖𝑗𝕄𝑛(𝐴𝑝(𝐺))=𝑎𝑚𝑖𝑗𝑎𝑖𝑗𝕄𝑛(𝐴𝑝(𝐺))=sup𝑆PM𝑝(𝐺)1sup𝑛𝑖=1|||||𝑛𝑖=1𝑆𝑎𝑖𝑗𝑎𝑆𝑚𝑖𝑗𝑥𝑗|||||𝑝1/𝑝,𝑥𝑗𝑛,𝑛𝑗=1||𝑥𝑗||𝑝,1(3.7) which clearly converges to zero as 𝑚. This implies that 𝑖(𝑛) has dense range. Hence, for each 𝑛, since 𝛼𝑛𝑀𝑛(VN(𝐺))𝑀𝑛(PM𝑝(𝐺)) we have 𝛼𝑛𝑇(𝑛)=𝛼𝑛𝑇𝑥sup𝑖𝑗𝑀𝑛(PM𝑝(𝐺))𝑥𝑖𝑗𝕄𝑛𝐴𝑝𝑥(𝐺),𝑖𝑗𝕄𝑛(𝐴𝑝(𝐺))1𝛼𝑛𝛼sup𝑛𝑇1𝑥𝑖𝑗𝑀𝑛(VN(𝐺))𝑥𝑖𝑗𝕄𝑛𝑥(𝐴(𝐺)),𝑖𝑗𝕄𝑛(𝐴𝑝(𝐺))𝛼1sup𝑛𝑇1𝑥𝑖𝑗𝑀𝑛(VN(𝐺))𝑥𝑖𝑗𝕄𝑛𝑥(𝐴(𝐺)),𝑖𝑗𝕄𝑛(𝐴(𝐺))1=𝛼𝑛𝑇1(𝑛),(3.8) that is, 𝑇1(𝑛)𝑇(𝑛).(iii)Using the weak-star density of PM𝑝(𝐺)1 in VN(𝐺)1, the same proof as in (ii) can be applied.(iv)Let 𝑇𝐴𝑝(𝐺)PM𝑝(𝐺) be a 𝑝-completely positive linear map. By part (i), the restriction map 𝑇1 is a completely positive map. Hence, by Proposition 3.3, there are infinitely many 𝑛 such that 𝑇1(𝑛)𝑛2. Now, since by part (iii), for each 𝑛,  𝑇(𝑛)𝛽2𝑛𝑇1(𝑛) the statement is clear.