1 Introduction

Let G be a topological group which is compact and Hausdorff. Consider a homogeneous space G/H where H is a closed subgroup of G. If we denote the normalized Haar measures on G and H by m and dh respectively, then there is an induced left invariant Radon measure \(\widetilde{m}\) on G/H satisfying Weil’s formula:

$$\begin{aligned} \int _{G/H} \int _H f(th) \, dh \, d\widetilde{m}(tH) = \int _G f \,dm \qquad (f \in C(G)). \end{aligned}$$

In this setting, Farashahi [5] introduced a method to obtain many of the well-known results on G/H from the ones on G. This method relies unavoidably on an extension of the operator \(T_H:C(G) \rightarrow C(G/H)\) given by \(T_Hf(tH)=\int _H f(th) \, dh\). The extension is in fact a norm-decreasing operator from \(L^p(G,m)\) onto \(L^p(G/H,\widetilde{m})\) where \(1 \le p < \infty \). The crucial property for this method is the surjectivity of the extension as it provides a connection to all \(L^p\)-functions on G/H to those on G. The extension was used to study abstract Fourier analysis on homogeneous spaces in various aspects such as convolutions, Fourier transform operators, Fourier series and measure algebras, see [6,7,8,9].

A vector measure is a measure taking values in a Banach space. There are many studies about functions in \(L^p\)-spaces of a compact group associated to a vector measure and invariant properties under the group operations of the vector measure itself. For example, the Fourier transform and the convolution along with invariant properties were studied in [1,2,3] under the condition that G is an abelian compact group. Then they were generalized to a non-abelian case in [13, 14].

Let \(\nu \) be a vector measure on G. In this paper, we initiate a study of an extension of the operator \(T_H: C(G) \rightarrow C(G/H)\) to an operator with the domain \(L^p(G,\nu )\). However, the codomain C(G/H) must be extended as well. For this purpose, we will construct a corresponding vector measure \({\tilde{\nu }} \) on G/H and show that the codomain of the extended operator is \(L^p(G/H,{\tilde{\nu }} )\). We investigate whether the extended operator is surjective and whether Weil’s formula is valid. It turns out that these are true for some vector measure \(\nu \). Fortunately, it is sufficient for the study of functions in \(L^p(G/H,\mu )\) for any vector measure \(\mu \) on G/H. We will employ the extension to obtain properties of invariant vector measures on G/H. Moreover, we introduce a new definition of a Fourier transform of functions in \(L^1(G,\nu )\) which is a variant definition of [13]. In our definition, \(\nu \) is taking values in a Banach space while in [13] \(\nu \) is taking values in an operator space. The uniqueness theorem of the Fourier transform and the Riemann–Lebesgue lemma are considered. Finally, we provide an analogous definition for a Fourier transform of functions in \(L^1(G/H,\mu )\) and once more employ the extension to obtain relations between the Fourier transforms of functions on G and G/H.

This paper is organized as follows. We give preliminary background in Sect. 2. In Sect. 3, an extension of the operator \(T_H\) to the space \(L^p(G,\nu )\) is studied along with its properties. Then the obtained properties of the extension will be used to derive properties of invariant vector measures on G/H in Sect. 4. There are three types of invariant vector measures we consider in this paper: translation invariant, norm integral invariant and semivariation invariant measures. Section 5 concerns the Fourier transforms of functions on G and G/H.

2 Preliminaries

2.1 Fourier Analysis with Haar Measures

Let G be a compact group with the normalized Haar measure m. The dual space \(\widehat{G}\) of G is the set of all unitary equivalence classes of irreducible unitary representations of G. For each \([\pi ] \in \widehat{G}\), the representation space of \(\pi \) is denoted by \({\mathcal {H}}_\pi \) with the dimension \(d_\pi = \dim {\mathcal {H}}_\pi \). For \([\pi ] \in \widehat{G}\) and \(u,v \in {\mathcal {H}}_\pi \), the function \(\pi _{u,v}:G \rightarrow {\mathbb {C}}\) given by \(\pi _{u,v}(t)=\langle \pi (t)v,u \rangle \) is called a matrix element of \(\pi \). We write \(\pi _{ij}\) for \(\pi _{e_i,e_j}\). Denote by \(\text {Trig}(G)\) the set of all finite linear combinations of matrix elements of irreducible representations. Note that \(\text {Trig}(G)\) is dense in C(G) in the uniform norm. For \(f \in L^1(G,m)\) and \([\pi ] \in \widehat{G}\), the Fourier transform of f is defined in the weak sense as

$$\begin{aligned} {\mathcal {F}}_G(f)(\pi )=\widehat{f}(\pi )=\int _{G} f(t)\pi (t)^* \, dm(t) \in {\mathcal {B}}({\mathcal {H}}_\pi ). \end{aligned}$$

Given any collection \(\{X_i\}_{i \in I}\) of Banach spaces where each \(X_i\) is equipped with the norm \(\Vert \cdot \Vert _i\). The space \(\ell ^\infty (I;X_i) =\{ x \in \prod _{i \in I} X_i: \sup _{i \in I} \Vert x_i\Vert _i < \infty \}\) is a Banach space with the norm \(\Vert x\Vert _\infty = \sup _{i \in I} \Vert x_i\Vert _i\). The set \(c_0(I;X_i)\) of all \(x=(x_i)\) for which \(\{i \in I: \Vert x_i\Vert _i > \varepsilon \}\) is finite for any \(\varepsilon > 0\) is a closed subspace of \(\ell ^\infty (I;X_i)\). By [12, Theorem 28.40], the Fourier transform operator \({\mathcal {F}}_G\) is a bounded linear operator from \(L^1(G,m)\) into \(c_0(\widehat{G};{\mathcal {B}}({\mathcal {H}}_\pi ))\) with \(\Vert \widehat{f}(\pi )\Vert \le \Vert f\Vert _{L^1(G,m)}\). For more details, see [11].

Let H be a closed subgroup of G and G/H the homogeneous space of left cosets equipped with the quotient topology. We denote the quotient map by \(q: G \rightarrow G/H\). For \(\varphi :G/H \rightarrow {\mathbb {C}}\), we write \(\varphi _q:G \rightarrow {\mathbb {C}}\) for a function given by \(\varphi _q(t)=\varphi (tH)\). Let dh be the normalized Haar measure on H. It is well-known that there is a unique (up to scalar) invariant Radon measure \(\widetilde{m}\) on G/H satisfying Weil’s formula:

$$\begin{aligned} \int _{G/H} \int _H f(th) \, dh \, d\widetilde{m}(tH) = \int _G f \,dm \quad (f \in C(G)). \end{aligned}$$

In fact, \(\widetilde{m}\) is the pushforward measure of m by the quotient map q. Define a bounded operator \(T_H:C(G) \rightarrow C(G/H)\) by

$$\begin{aligned} T_Hf(tH)=\int _H f(th) \, dh \quad (tH \in G/H, \ f \in C(G)). \end{aligned}$$

According to [5], for any \(1 \le p < \infty \), the operator \(T_H\) can be extended to a norm-decreasing operator from \(L^p(G,m)\) onto \(L^p(G/H,\widetilde{m})\) (still denoted by \(T_H\)) for which the extended Weil’s formula holds:

$$\begin{aligned} \int _{G/H} T_H f \, d\widetilde{m} = \int _G f \,dm \quad \big (f \in L^1(G,m)\big ). \end{aligned}$$
(1)

For more details on Weil’s formula, see [16]. The dual space of G/H is given by \(\widehat{G/H}:=\big \{ [\pi ]\in \widehat{G}: T_H^\pi \ne 0 \big \}\) where \(T_H^\pi \) is defined in the weak sense as the operator \(T_H^\pi :=\int _H \pi (h) \,dh \in {\mathcal {B}}({\mathcal {H}}_\pi )\). For \(\varphi \in L^1(G/H,\widetilde{m})\) and \([\pi ] \in \widehat{G/H}\), the Fourier transform of \(\varphi \) is defined in the weak sense as

$$\begin{aligned} {\mathcal {F}}_{G/H}(\varphi )(\pi ) = \widehat{\varphi }(\pi ) = \int _{G/H} \varphi (tH)\Gamma _{\pi }(tH)^*\,d\widetilde{m}(tH) \in {\mathcal {B}}({\mathcal {H}}_\pi ), \end{aligned}$$

where \(\Gamma _{\pi }(tH)=\pi (t)T^\pi _H\). Then the Fourier transform operator \({\mathcal {F}}_{G/H}\) is a bounded linear operator from \(L^1(G/H,\widetilde{m})\) into \(c_0(\widehat{G/H};{\mathcal {B}}({\mathcal {H}}_\pi ))\) with \(\Vert \widehat{\varphi }(\pi )\Vert \le \Vert \varphi \Vert _{L^1(G/H,\widetilde{m})}\), see [5, Theorem 5.5].

2.2 Vector Measures

Let \((\Omega ,\mathfrak {B}(\Omega ))\) be a Borel measurable space and X a Banach space. The closed unit ball in the dual space \(X^*\) is denoted by \(B_{X^*}\). A (countably additive) vector measure \(\nu \) on \((\Omega ,\mathfrak {B}(\Omega ))\) is an X-valued function \(\nu :\mathfrak {B}(\Omega ) \rightarrow X\) such that \(\nu (\cup _{n=1}^\infty E_n) = \sum _{n=1}^\infty \nu (E_n)\) in the norm topology for any sequence \((E_n)\) of pairwise disjoint sets in \(\mathfrak {B}(\Omega )\). Given \(x^* \in X^*\), let \({\langle \nu , x^*\rangle } :\mathfrak {B}(\Omega ) \rightarrow {\mathbb {C}}\) be the complex measure given by \({\langle \nu , x^*\rangle } (E)=\langle \nu (E),x^* \rangle \) for \(E \in \mathfrak {B}(\Omega )\). The semivariation \(\Vert \nu \Vert \) of \(\nu \) is the set function defined by \(\Vert \nu \Vert (E)=\sup _{x^* \in B_{X^*}} \vert {\langle \nu , x^*\rangle } \vert (E)\) for \(E \in \mathfrak {B}(\Omega )\). A vector measure \(\nu \) is said to be regular if for each \(E \in \mathfrak {B}(\Omega )\) and \(\varepsilon > 0\) there exist a compact set K and an open set O such that \(K \subset E \subset O\) and \(\Vert \nu \Vert (O\setminus K) < \varepsilon \). We denote by \({\mathcal {M}}(\Omega ,X)\) the set of all regular X-valued measures on \(\Omega \).

A measurable function \(f:\Omega \rightarrow {\mathbb {C}}\) is said to be \(\nu \)-integrable if \(f \in L^1({\langle \nu , x^*\rangle } )\) for every \(x^* \in X^*\) and for each \(E \in \mathfrak {B}(\Omega )\) there is an \(x_E \in X\) such that \(\langle x_E, x^* \rangle = \int _E f \,d{\langle \nu , x^*\rangle } \) for every \(x^* \in X^*\). We denote \(x_E\) by \(\int _E f \, d\nu \). For a measurable function \(f:\Omega \rightarrow {\mathbb {C}}\), define

$$\begin{aligned} \Vert f\Vert _{\nu } = \sup _{x^* \in B_{X^*}} \int _\Omega \vert f \vert \, d\vert {\langle \nu , x^*\rangle } \vert \end{aligned}$$

and \(\Vert f\Vert _{\nu ,p}:= \Vert \vert f \vert ^p\Vert ^{1/p}_{\nu }\). The space \(L^1(\Omega ,\nu )\) of all \(\nu \)-integrable functions is a Banach space with the norm \(\Vert \cdot \Vert _{\nu }\). We say that \(f=g\) \(\nu \)-a.e. if \(\Vert f-g\Vert _{\nu }=0\). For each \(1 \le p < \infty \), the space \(L^p(\Omega ,\nu ):=\{f \in L^1(\Omega ,\nu ): \vert f \vert ^p \in L^1(\Omega ,\nu ) \}\) is a Banach space with the norm \(\Vert \cdot \Vert _{L^p(\Omega ,\nu )}:= \Vert \cdot \Vert _{\nu ,p} \). We denote by \(S(\Omega )\) the set of all simple functions on \(\Omega \). The integral operator \(I_\nu :L^1(\Omega ,\nu ) \rightarrow X\) is defined by \(I_\nu (f)=\int _\Omega f \, d\nu \) for \(f \in L^1(\Omega ,\nu ).\) Then \(I_\nu \) is bounded with \(\Vert I_\nu (f)\Vert _X \le \Vert f\Vert _{L^1(\Omega ,\nu )}\).

Theorem 2.1

[15] Let \(f:\Omega \rightarrow {\mathbb {C}}\) be a complex function. Then f is \(\nu \)-integrable if and only if there is a sequence \((f_n)\) of simple functions which converges pointwise to f and for which \((\int _E f_n \, d\nu )\) is Cauchy for any \(E \in \mathfrak {B}(\Omega )\).

Theorem 2.2

[13] Let \(\nu \in {\mathcal {M}}(G,X)\). Then C(G) is dense in \(L^p(G,\nu )\) for all \(1\le p < \infty \).

For Banach spaces X and Y, a linear operator \(T:X \rightarrow Y\) is said to be weakly compact if T(B) is a relatively weakly compact subset of Y whenever B is a bounded subset of X. By [4, Corollary VI.2.14], we have that on a compact Hausdorff space there is a one-to-one correspondence between the set of all regular vector measures and the set of all weakly compact operators. To be precise, given a regular vector measure \(\nu :\mathfrak {B}(\Omega ) \rightarrow X\), there is a weakly compact operator \(T:C(\Omega ) \rightarrow X\) representing \(\nu \), that is, \(T(f)=\int _\Omega f \, d\nu \) for all \(f \in C(\Omega )\), and vice versa.

A vector measure \(\nu \) is said to be absolutely continuous with respect to a positive scalar measure \(\lambda \), denoted by \(\nu \ll \lambda \), if \(\nu (E) \rightarrow 0\) in norm as \(\lambda (E) \rightarrow 0\) where \(E \in \mathfrak {B}(\Omega )\). Note that \(\nu \ll \lambda \) if and only if \(\nu \) vanishes on all sets of \(\lambda \)-measure zero, by [4, Theorem I.2.1]. Moreover, \(\nu \) vanishes on all sets of \(\lambda \)-measure zero if and only if \(\Vert \nu \Vert \) vanishes on all sets of \(\lambda \)-measure zero. By Rybakov’s theorem [4], there is a linear functional \(x^* \in X^*\) such that \(\nu \ll \vert {\langle \nu , x^*\rangle } \vert \). This functional is called a Rybakov functional. For \(k \in [0,\infty )\), a vector measure \(\nu \) is said to be k-scalarly bounded by m if for any \(x^* \in X^*\) and \(E \in \mathfrak {B}(\Omega )\), we have \(\vert {\langle \nu , x^*\rangle } \vert (E) \le km(E)\).

Let \(\tau :G \rightarrow G\) be a homeomorphism. For a measurable function \(f:G \rightarrow {\mathbb {C}}\), we denote \(f \circ \tau ^{-1}\) by \(f_\tau \). For \(a \in G\), we define the left translation \(L_a\) and the right translation \(R_a\) by \(L_a(t)=at\) and \(R_a(t)=ta^{-1}\) for \(t \in G\). In the case that \(\tau = L_a\) or \(R_a\), we shall write \(L_af\) or \(R_af\) instead of \(f_\tau \). Hence \((L_a f)(t)=f(a^{-1}t)\) and \((R_a f)(t)=f(ta)\) for each \(t \in G\).

Definition 1

Let \(\tau :G \rightarrow G\) be a homeomorphism and \(\nu \) a vector measure on G. We say that \(\nu \) is \(\tau \)-invariant if

$$\begin{aligned} I_\nu (f_\tau ) =I_\nu (f) \quad \text { for all } f \in S(G). \end{aligned}$$

Given a collection \({\mathcal {T}}\) of homeomorphisms on G\(\nu \) is said to be \({\mathcal {T}}\)-invariant if it is \(\tau \)-invariant for all \(\tau \in {\mathcal {T}}\). In particular, if \({\mathcal {T}}=\{L_a: a\in G\}\) (or \({\mathcal {T}}=\{R_a: a\in G\}\)), we say that \(\nu \) is left (or right) invariant.

Definition 2

Let \(\tau :G \rightarrow G\) be a homeomorphism and \(\nu \) a vector measure on G. We say that \(\nu \) is norm integral \(\tau \)-invariant if

$$\begin{aligned} \Vert I_\nu (f_\tau ) \Vert =\Vert I_\nu (f) \Vert \quad \text { for all } f \in S(G). \end{aligned}$$

Given a collection \({\mathcal {T}}\) of homeomorphisms on G\(\nu \) is said to be norm integral \({\mathcal {T}}\)-invariant if it is norm integral \(\tau \)-invariant for all \(\tau \in {\mathcal {T}}\). In particular, if \({\mathcal {T}}=\{L_a: a\in G\}\) (or \({\mathcal {T}}=\{R_a: a\in G\}\)), we say that \(\nu \) is norm integral left (or right) invariant.

Definition 3

Let \(\tau :G \rightarrow G\) be a homeomorphism and \(\nu \) a vector measure on G. We say that \(\nu \) is semivariation \(\tau \)-invariant if

$$\begin{aligned} \Vert f_\tau \Vert _{L^1(G,\nu )} =\Vert f \Vert _{L^1(G,\nu )} \quad \text { for all } f \in S(G). \end{aligned}$$

Given a collection \({\mathcal {T}}\) of homeomorphisms on G\(\nu \) is said to be semivariation \({\mathcal {T}}\)-invariant if it is semivariation \(\tau \)-invariant for all \(\tau \in {\mathcal {T}}\). In particular, if \({\mathcal {T}}=\{L_a: a\in G\}\) (or \({\mathcal {T}}=\{R_a: a\in G\}\)), we say that \(\nu \) is semivariation left (or right) invariant.

2.3 Tensor Integration

Let X and Y be any Banach spaces. Recall that the space \({\mathcal {B}}(Y^* \times X^*)\) of bounded bilinear forms on \(Y^* \times X^*\) is a Banach space equipped with the norm

$$\begin{aligned} \Vert b\Vert = \sup \{\vert b(y^*,x^*) \vert : y^* \in B_{Y^*}, x^* \in B_{X^*}\}. \end{aligned}$$

Note that we can realize \(Y \otimes X\) as a subspace of \({\mathcal {B}}(Y^* \times X^*)\) by considering \(u=\sum _{i=1}^n y_i \otimes x_i \in Y \otimes X\) as a bilinear form given by \(b_u(y^*,x^*) = \sum y^*(y_i)x^*(x_i)=(y^* \otimes x^*)(u)\) for \(y^* \in Y^*\) and \(x^* \in X^*\). The injective norm \(\Vert \cdot \Vert _\vee \) on \(Y \otimes X\) is the norm induced by this embedding, i.e.,

$$\begin{aligned} \Vert u \Vert _\vee = \sup _{y^* \in B_{Y^*}, x^* \in B_{X^*}} \vert (y^* \otimes x^*)(u)\vert . \end{aligned}$$

Moreover, we have alternative formulas for the injective norm

$$\begin{aligned} \Vert u \Vert _\vee = \sup _{y^* \in B_{Y^*}} \left\| \sum y^*(y_i) x_i \right\| _X = \sup _{x^* \in B_{X^*}} \left\| \sum x^*(x_i) y_i \right\| _Y. \end{aligned}$$

The completion of the tensor product space \(Y \otimes X\) with the injective norm is called the injective tensor product of Y and X, denoted by \(Y \check{\otimes } X\). For more details, see [17].

Now we summarize the concept of tensor integration introduced by [18]. Let \(\nu \) be an X-valued vector measure. A function \(f:\Omega \rightarrow Y\) is said to be \(\nu \)-measurable if there is a sequence of Y-valued simple functions \((f_n)\) with \(\lim _{n \rightarrow \infty } \Vert f_n - f\Vert _Y = 0\) \(\nu \)-a.e. We say that a function \(f:\Omega \rightarrow Y\) is weakly \(\nu \)-measurable if for each \(y^* \in Y^*\) the function \(y^*f\) is \(\nu \)-measurable. Note that a function \(f:\Omega \rightarrow Y\) is \(\nu \)-measurable if and only if f is \(\vert {\langle \nu , x^*\rangle } \vert \)-measurable for some Rybakov functional \(x^* \in X^*\).

Theorem 2.3

(Pettis’s measurability theorem [4]) Let \(\lambda \) be a finite positive measure. A function \(f:\Omega \rightarrow Y\) is \(\lambda \)-measurable if and only if f is weakly \(\lambda \)-measurable and \(\lambda \)-essentially separably valued.

Let \(E \in \mathfrak {B}(\Omega )\) and \(\phi =\sum _{i=1}^n y_i \chi _{A_i}\) be a Y-valued simple function on \(\Omega \), where \(y_i \in Y\) and \(A_i \in \mathfrak {B}(\Omega )\). We define \(\int _E \phi \, d\nu = \sum y_i \otimes \nu (E \cap A_i) \in Y \otimes X.\) Then it can be shown that \((y^* \otimes x^*)(\int _E \phi \, d\nu ) = \int _E y^*\phi \,d{\langle \nu , x^*\rangle } \) for \(y^* \in Y^*\) and \(x^* \in X^*\), hence \(\left\| \int _E \phi \, d\nu \right\| _\vee \le \sup _{x^* \in B_{X^*}} \int _E \Vert \phi \Vert \, d\vert \langle \nu ,x^* \rangle \vert .\) For a \(\nu \)-measurable function \(f:\Omega \rightarrow Y\), we let

$$\begin{aligned} \text {N}(f) = \sup _{x^* \in B_{X^*}} \int _\Omega \Vert f\Vert \, d\vert {\langle \nu , x^*\rangle } \vert . \end{aligned}$$

Definition 4

A \(\nu \)-measurable function \(f:\Omega \rightarrow Y\) is \(\check{\otimes }\)-integrable if there exists a sequence \((f_n)\) of simple functions such that

$$\begin{aligned} \lim _{n \rightarrow \infty } \text {N}(f-\phi _n) =0. \end{aligned}$$

In this case, the sequence \(( \int _E \phi _n \,d\nu )\) is a Cauchy sequence in \(Y \check{\otimes } X\) for each \(E \in \mathfrak {B}(\Omega )\). By the completeness of \(Y \check{\otimes } X\), the limit of \(( \int _E \phi _n \,d\nu )\) is denoted by \(\int _E f \, d\nu \) and is called the \(\check{\otimes }\)-integral of f over E with respect to \(\nu \).

Note that if f is \(\check{\otimes }\)-integrable, then \((y^* \otimes x^*)(\int _E f \, d\nu ) = \int _E y^*f \,d{\langle \nu , x^*\rangle } \) for \(E \in \mathfrak {B}(\Omega )\), \(y^* \in Y^*\) and \(x^* \in X^*\) and \(\Vert \int _\Omega f \,d\nu \Vert _\vee \le N(f)\).

Theorem 2.4

[18] A \(\nu \)-measurable function f is \(\check{\otimes }\)-integrable if and only if \(\Vert f\Vert \) is \(\nu \)-integrable.

3 Extensions of the Operator \(T_H\)

In this section, we study extensions of the operator \(T_H: C(G) \rightarrow C(G/H)\). Given a vector measure \(\nu \in {\mathcal {M}}(G,X)\), we can naturally construct a vector measure on G/H as follows. Let \(T_\nu :C(G) \rightarrow X\) be the corresponding weakly compact operator for \(\nu \), i.e.,

$$\begin{aligned} T_\nu (f)=\int _{G} f \, d\nu \quad (f \in C(G)). \end{aligned}$$

Define \(T_{\tilde{\nu }} :C(G/H) \rightarrow X\) by

$$\begin{aligned} T_{\tilde{\nu }} (\varphi )=T_\nu (\varphi _q)=\int _G \varphi _q \,d\nu \quad (\varphi \in C(G/H)). \end{aligned}$$

Then \(T_{\tilde{\nu }} \) is weakly compact since \(\Vert \varphi \Vert _{\sup }=\Vert \varphi _q\Vert _{\sup }\) for all \(\varphi \in C(G/H)\). Hence there is a representing vector measure \({\tilde{\nu }} \in {\mathcal {M}}(G/H,X)\). Moreover, we immediately have that

$$\begin{aligned} \int _{G/H} \varphi \, d{\tilde{\nu }} = \int _G \varphi _q \, d\nu \quad (\varphi \in C(G/H)). \end{aligned}$$
(2)

We shall begin with some basic properties of \({\tilde{\nu }} \).

Proposition 3.1

Let \(\nu \in {\mathcal {M}}(G,X)\).

  1. 1.

    The vector measure \({\tilde{\nu }} \) is the pushforward (vector) measure of \(\nu \) by the quotient map q, i.e., \({\tilde{\nu }} (E)=\nu (q^{-1}(E))\) for all \(E \in \mathfrak {B}(G/H)\). Moreover, the Eq. (2) holds for all \(\varphi \in L^1(G/H,{\tilde{\nu }} )\) provided that \(\varphi _q \in L^1(G,\nu )\).

  2. 2.

    For any \(x^* \in X^*\) and \(E \in \mathfrak {B}(G/H)\), \(\vert {\langle {\tilde{\nu }} , x^*\rangle } \vert (E) \le \vert {\langle \nu , x^*\rangle } \vert (q^{-1}(E))\). Then \(\Vert \varphi \Vert _{L^p(G/H,{\tilde{\nu }} )} \le \Vert \varphi _q\Vert _{\nu ,p}\) for any \(1\le p < \infty \) and \(\varphi \in L^p(G/H,{\tilde{\nu }} )\).

Proof

  1. 1.

    Let \(\lambda \) be the pushforward measure of \(\nu \) by the quotient map q. It follows from Eq. (2) that

    $$\begin{aligned} \int _{G/H} \varphi \, d\lambda =\int _G \varphi _q \, d\nu =\int _{G/H} \varphi \, d{\tilde{\nu }} \end{aligned}$$

    for all \(\varphi \in C(G/H)\). Hence \({\tilde{\nu }} =\lambda \). Next observe that Eq. (2) holds for all \(\varphi \in S(G/H)\). Let \(0 \le \varphi \in L^1(G/H,{\tilde{\nu }} )\). Then there is a sequence of positive simple functions \(\varphi _n \uparrow \varphi \) pointwise. By the monotone convergence theorem, for each \(x^* \in X^*\)

    $$\begin{aligned} \int _{G/H} \varphi \, d{\langle {\tilde{\nu }} , x^*\rangle }&= \lim _{n\rightarrow \infty } \int _{G/H} \varphi _n \, d{\langle {\tilde{\nu }} , x^*\rangle } \\&= \lim _{n\rightarrow \infty } \int _G (\varphi _n)_q \, d{\langle \nu , x^*\rangle } \\&= \int _G \varphi _q \, d{\langle \nu , x^*\rangle } . \end{aligned}$$

    This identity easily extends to \(\varphi \in L^1(G/H,{\tilde{\nu }} )\). If we assume that \(\varphi _q \in L^1(G,\nu )\), then

    $$\begin{aligned} \left\langle \int _{G/H} \varphi \, d{\tilde{\nu }} , x^* \right\rangle = \int _{G/H} \varphi \, d{\langle {\tilde{\nu }} , x^*\rangle } = \int _G \varphi _q \, d{\langle \nu , x^*\rangle } = \left\langle \int _{G} \varphi _q \, d\nu , x^* \right\rangle \end{aligned}$$

    for all \(x^* \in X^*\), which proves the Eq. (2).

  2. 2.

    Let \(E \in \mathfrak {B}(G/H)\). Consider any disjoint partition \(\{E_n\}_{n=1}^{k}\) of E where \(E_n \in \mathfrak {B}(G/H)\). Since \(\{q^{-1}(E_n)\}_{n=1}^{k}\) forms a disjoint partition of \(q^{-1}(E)\),

    $$\begin{aligned} \sum _{n=1}^k \vert {\langle {\tilde{\nu }} , x^*\rangle } (E_n)\vert = \sum _{n=1}^k \vert {\langle \nu , x^*\rangle } (q^{-1}(E_n))\vert \le \vert {\langle \nu , x^*\rangle } \vert (q^{-1}(E)). \end{aligned}$$

    Hence \(\vert {\langle {\tilde{\nu }} , x^*\rangle } \vert (E) \le \vert {\langle \nu , x^*\rangle } \vert (q^{-1}(E))\). Consequently,

    $$\begin{aligned} \int _{G/H} \varphi \, d\vert {\langle {\tilde{\nu }} , x^*\rangle } \vert \le \int _{G} \varphi _q \, d\vert {\langle \nu , x^*\rangle } \vert \end{aligned}$$

    holds for any simple function \(\varphi \ge 0\) on G/H. Then for any \(\varphi \in L^1(G/H,\mu )\), the monotone convergence theorem implies that

    $$\begin{aligned} \int _{G/H} \vert \varphi \vert \, d\vert {\langle {\tilde{\nu }} , x^*\rangle } \vert \le \int _{G} \vert \varphi _q\vert \, d\vert {\langle \nu , x^*\rangle } \vert . \end{aligned}$$

    Therefore, \(\Vert \varphi \Vert _{L^p(G/H,{\tilde{\nu }} )} \le \Vert \varphi _q\Vert _{\nu ,p}\) for any \(1\le p < \infty \) and \(\varphi \in L^p(G/H,{\tilde{\nu }} )\).

\(\square \)

Example 1

Let \(1 \le p < \infty \) and \(S:L^p(G,m) \rightarrow X\) be any bounded linear map, where m is the normalized Haar measure on G. Define a vector measure \(\nu :\mathfrak {B}(G) \rightarrow X\) corresponding to S by \(\nu (E) = S(\chi _E)\) for \(E \in \mathfrak {B}(G)\). Then by Proposition 3.1.1 the vector measure \({\tilde{\nu }} \) is given by \({\tilde{\nu }} (F)=S(\chi _{q^{-1}(F)})\) for \(F \in {\mathcal {B}}(G/H)\).

  1. 1.

    Let \(X = {\mathbb {C}}\) and \(S:L^1(G,m) \rightarrow {\mathbb {C}}\) be given by \(S(f)=\int _G f \,dm\) for any \(f \in L^1(G,m)\). In this case, \(\nu = m\). Moreover, \({\tilde{\nu }} = \widetilde{m}\) since \({\tilde{\nu }} (F) = \int _G \chi _{q^{-1}(F)} \, dm = \int _{G/H} \chi _F \, d\widetilde{m} = \widetilde{m}(F)\) for all \(F \in \mathfrak {B}(G/H)\), where \(\widetilde{m}\) is the pushforward measure of m.

  2. 2.

    Let \(X = L^1(G,m)\) and \(S = \text {Id}_{L^1(G,m)}\). Then \({\tilde{\nu }} (F)=\chi _{q^{-1}(F)}\) for \(F \in \mathfrak {B}(G/H)\).

  3. 3.

    Let \(\lambda \) be a complex regular measure on G and \(1 \le p < \infty \). We define \(S:L^p(G,m) \rightarrow L^p(G,m)\) by \(S(f) = f * \lambda \) where \((f * \lambda )(t)=\int _G f(ts^{-1}) \,d\lambda (s)\) for \(t \in G\). Then \({\tilde{\nu }} (F)=\chi _{q^{-1}(F)}* \lambda \) for \(F \in \mathfrak {B}(G/H)\).

  4. 4.

    Let \(1 \le p \le 2\) and \(S:L^p(G,m) \rightarrow \ell ^{p'}(\widehat{G};{\mathcal {B}}({\mathcal {H}}_\pi ))\) be defined by \(S(f)={\mathcal {F}}_G(f)\). Then \({\tilde{\nu }} (F)={\mathcal {F}}_G(\chi _{q^{-1}(F)})={\mathcal {F}}_{G/H}(\chi _F)\) for \(F \in \mathfrak {B}(G/H)\).

Let \(\tau :G/H \rightarrow G/H\) be a homeomorphism. For example, one can consider a left translation \(L_a:G/H \rightarrow G/H\) by \(a \in G\) given by \(L_a(tH)=atH\) for each \(tH \in G/H\). For a measurable function \(\varphi :G/H \rightarrow {\mathbb {C}}\), we denote \(\varphi \circ \tau ^{-1}\) by \(\varphi _\tau \). In the case that \(\tau =L_a\) where \(a \in G\), we shall denote \(\varphi \circ (L_a)^{-1}\) by \(L_a \varphi \) and by definition we have \((L_a \varphi )(tH)=\varphi (a^{-1}tH)\) for all \(tH \in G/H\).

Definition 5

Let \(\tau :G/H \rightarrow G/H\) be a homeomorphism. For any vector measure \(\mu \) on G/H, we say that \(\mu \) is norm integral \(\tau \)-invariant if

$$\begin{aligned} \Vert I_\mu (\varphi _\tau ) \Vert =\Vert I_\mu (\varphi ) \Vert \quad \text { for all } \varphi \in S(G/H). \end{aligned}$$

Given a collection \({\mathcal {T}}\) of homeomorphisms on G/H\(\mu \) is said to be norm integral \({\mathcal {T}}\)-invariant if it is norm integral \(\tau \)-invariant for all \(\tau \in {\mathcal {T}}\). In particular, if \({\mathcal {T}}=\{L_a: a\in G\}\), we say that \(\mu \) is norm integral left invariant.

This proposition is merely a consequence of Proposition 3.1.

Proposition 3.2

Let \(\nu \in {\mathcal {M}}(G,X)\).

  1. 1.

    For \(a \in G\), if \(\nu \) is norm integral \(L_a\)-invariant, then so is \({\tilde{\nu }} \).

  2. 2.

    If \(\nu \ll m\), then \({\tilde{\nu }} \ll \widetilde{m}\).

  3. 3.

    If \(\nu \) is k-scalarly bounded by m, then \({\tilde{\nu }} \) is k-scalarly bounded by \(\widetilde{m}\).

Proof

  1. 1.

    For \(\varphi \in S(G/H)\), by Proposition 3.1.1, Eq. (2) holds for simple functions, we have

    $$\begin{aligned} \Vert I_{\tilde{\nu }} (L_a\varphi ) \Vert = \Vert I_\nu ((L_a\varphi )_q)\Vert = \Vert I_\nu (L_a(\varphi _q))\Vert = \Vert I_\nu (\varphi _q)\Vert = \Vert I_{\tilde{\nu }} (\varphi )\Vert . \end{aligned}$$
  2. 2.

    For any \(F \in \mathfrak {B}(G/H)\), \(\widetilde{m}(F) = m(q^{-1}(F))\) and \({\tilde{\nu }} (F)=\nu (q^{-1}(F))\). If \(\widetilde{m}(F) \rightarrow 0\), then also \(m(q^{-1}(F)) \rightarrow 0\), and hence \({\tilde{\nu }} (F)=\nu (q^{-1}(F)) \rightarrow 0\) since \(\nu \ll m\).

  3. 3.

    It follows immediately from the fact that

    $$\begin{aligned} \vert {\langle {\tilde{\nu }} , x^*\rangle } \vert (E)\le \vert {\langle \nu , x^*\rangle } \vert (q^{-1}(E)) \le k m(q^{-1}(E)) = k\widetilde{m}(E) \end{aligned}$$

    for any \(E \in \mathfrak {B}(G/H)\).

\(\square \)

Now we prove an existence of an extension of \(T_H\) to an operator from \(L^p(G,\nu )\) into \(L^p(G/H,{\tilde{\nu }} )\) for each \(1\le p < \infty \). This is a generalization of Theorem 3.2 in [5].

Theorem 3.3

Let \(1\le p < \infty \) and \({\mathcal {R}} = \{R_h: h \in H\}\). Suppose that \(\nu \) is semivariation \({\mathcal {R}}\)-invariant. Then the operator \(T_H:C(G) \rightarrow C(G/H)\) satisfies

$$\begin{aligned} \Vert T_H f\Vert _{L^p(G/H,{\tilde{\nu }} )} \le \Vert f\Vert _{L^p(G,\nu )} \quad \text { for all } f \in C(G), \end{aligned}$$

hence it has a unique extension to a norm-decreasing operator \(T_{H,\nu }:L^p(G,\nu ) \rightarrow L^p(G/H,{\tilde{\nu }} )\).

Proof

Let \(f \in C(G)\). By Proposition 3.1.2 and \(\nu \) being semivariation \({\mathcal {R}}\)-invariant,

$$\begin{aligned} \Vert T_H f\Vert _{L^p(G/H,{\tilde{\nu }} )}^p&\le \Vert (T_H f)_q\Vert _{L^p(G,\nu )}^p \\&= \sup _{x^* \in B_{X^*}} \int _{G} \vert (T_H f)(tH)\vert ^p \, d\vert {\langle \nu , x^*\rangle } \vert (t) \\&\le \sup _{x^* \in B_{X^*}} \int _G \int _H \vert f(th)\vert ^p \, dh \, d\vert {\langle \nu , x^*\rangle } \vert (t) \\&= \sup _{x^* \in B_{X^*}} \int _H \int _G \vert f(th)\vert ^p \, d\vert {\langle \nu , x^*\rangle } \vert (t) \, dh \\&\le \int _H \bigg (\sup _{x^* \in B_{X^*}} \int _G \vert f(th)\vert ^p \, d\vert {\langle \nu , x^*\rangle } \vert (t)\bigg ) \, dh \\&= \int _H \Vert R_h f\Vert _{L^p(G,\nu )}^p \, dh \\&= \int _H \Vert f\Vert _{L^p(G,\nu )}^p \, dh \\&= \Vert f\Vert _{L^p(G,\nu )}^p. \end{aligned}$$

By the density of C(G) in \(L^p(G,\nu )\), the operator \(T_H\) can be extended uniquely to a bounded linear map from \(L^p(G,\nu )\) to \(L^p(G/H,{\tilde{\nu }} )\). To verify that \(T_{H,\nu }\) is norm-decreasing, let \(f \in L^p(G,\nu )\) with \(f_n \rightarrow f\) in \(L^p(G,\nu )\) where \(f_n \in C(G)\). Then

$$\begin{aligned} \Vert T_{H,\nu } f\Vert _{L^p(G/H,{\tilde{\nu }} )} = \lim _{n\rightarrow \infty } \Vert T_H f_n\Vert _{L^p(G/H,{\tilde{\nu }} )} \le \lim _{n\rightarrow \infty } \Vert f_n\Vert _{L^p(G,\nu )} = \Vert f\Vert _{L^p(G,\nu )} \end{aligned}$$

as desired. \(\square \)

Remark 1

If there is no ambiguity, we shall denote \(T_{H,\nu }\) by \(T_H\). Secondly, it is worth noting that even though the extensions of \(T_H:C(G) \rightarrow C(G/H)\) to \(L^p(G,\nu )\) and \(L^q(G,\nu )\) might be different operators if \(p \ne q\), they coincide on the intersection of the domains. Suppose that we denote the extension of \(T_H\) to \(L^p(G,\nu )\) by \(T_{H,p}\) for \(1 \le p < \infty \). Consider \(1 \le p< q < \infty \). Note that it follows from [15, Proposition 3.31(ii)] that for any vector measure \(\mu \) on \(\Omega \), \(L^q(\Omega ,\mu ) \subset L^p(\Omega ,\mu )\) with \(\Vert f\Vert _{L^p(\Omega ,\mu )} \le K \Vert f\Vert _{L^q(\Omega ,\mu )}\) for some constant \(K>0\). Now we show that the extensions \(T_{H,p}\) and \(T_{H,q}\) coincide on \(L^q(G,\nu ) \subset L^p(G,\nu )\). Let \(f_n \rightarrow f\) in \(L^q(G,\nu )\) where \(f_n \in C(G)\). Then \(T_{H,p} f_n \rightarrow T_{H,p} f\) in \(L^p(G/H,{\tilde{\nu }} )\) and also \(T_{H,q} f_n \rightarrow T_{H,q} f\) in \(L^p(G/H,{\tilde{\nu }} )\). Since \(T_{H,p}\) and \(T_{H,q}\) agree on C(G), we have that \(T_{H,p} f = T_{H,q} f\) in \(L^p(G/H,{\tilde{\nu }} )\) which implies \(T_{H,p} f = T_{H,q} f\) \({\tilde{\nu }} \)-a.e. Thus there is no ambiguity to denote any extension \(T_{H,p}\) for any \(1\le p < \infty \) by \(T_H\).

Now we prove that the extension \(T_H\) is norm-decreasing in the sense of the norm in X.

Theorem 3.4

Let \(\nu \) be norm integral \({\mathcal {R}}\)-invariant where \({\mathcal {R}} = \{R_h: h \in H\}\). Then

$$\begin{aligned} \left\| \int _{G/H} T_H f \, d{\tilde{\nu }} \right\| _X \le \left\| \int _G f \, d\nu \right\| _X \quad \big (f \in L^1(G,\nu )\big ). \end{aligned}$$

Proof

Let \(f \in C(G)\). For \(x^* \in B_{X^*}\), by Eq. (2)

$$\begin{aligned} \int _{G/H} T_H f \, d{\langle {\tilde{\nu }} , x^*\rangle }&= \int _G( T_H f)_q \, d{\langle \nu , x^*\rangle } \quad \\&= \int _G \int _H f(th) \,dh \,d{\langle \nu , x^*\rangle } (t) \\&= \int _H \int _G (R_h f)(t) \,d{\langle \nu , x^*\rangle } (t) \,dh. \end{aligned}$$

Hence

$$\begin{aligned} \left\| \int _{G/H} T_H f \, d{\tilde{\nu }} \right\| _X&=\sup _{x^* \in B_{X^*}} \left| \int _{G/H} T_H f \, d{\langle {\tilde{\nu }} , x^*\rangle } \right| \\&=\sup _{x^* \in B_{X^*}} \left| \int _H \int _G (R_h f)(t) \,d{\langle \nu , x^*\rangle } (t) \,dh \right| \\&\le \int _H \bigg ( \sup _{x^* \in B_{X^*}} \left| \int _G (R_h f)(t) \,d{\langle \nu , x^*\rangle } (t) \right| \bigg ) \,dh \\&=\int _H \left\| \int _G R_h f \, d\nu \right\| _X dh \\&=\left\| \int _G f \, d\nu \right\| _X. \end{aligned}$$

For any \(f \in L^1(G,\nu )\), let \(f_n\) be a sequence of continuous functions converging to f in \( L^1(G,\nu )\). Then

$$\begin{aligned} \Vert I_{\tilde{\nu }} (T_H f) \Vert _X = \lim _{n \rightarrow \infty } \Vert I_{\tilde{\nu }} (T_H f_n) \Vert _X \le \lim _{n \rightarrow \infty } \Vert I_\nu (f_n) \Vert _X = \Vert I_\nu (f) \Vert _X \end{aligned}$$

that is \(\Vert \int _{G/H} T_H f \, d{\tilde{\nu }} \Vert _X \le \Vert \int _G f \, d\nu \Vert _X\) as desired. \(\square \)

We have investigated the properties of the extension \(T_H:L^p(G,\nu ) \rightarrow L^p(G/H,{\tilde{\nu }} )\) for a given vector measure \(\nu \in {\mathcal {M}}(G,X)\). However, in general, to study Fourier analysis on homogeneous spaces, it is essential to consider the space \(L^p(G/H,\mu )\) for a given vector measure \(\mu \) on G/H instead of the space \(L^p(G/H,{\tilde{\nu }} )\). To deal with this situation, we will define a corresponding measure \({\breve{\mu }} \) on G and study the extension \(T_H:L^p(G,{\breve{\mu }} ) \rightarrow L^p(G/H,\mu )\).

Let \(\mu \in {\mathcal {M}}(G/H,X)\) and \(T_\mu :C(G/H) \rightarrow X\) be the corresponding weakly compact operator given by

$$\begin{aligned} T_\mu (\varphi ) = \int _{G/H} \varphi \, d\mu \quad (\varphi \in C(G/H)). \end{aligned}$$

Observe that the operator \(T_\mu \circ T_H:C(G) \rightarrow X\) is weakly compact since \(T_H\) is bounded and \(T_\mu \) is weakly compact. Then there is a representing regular vector measure on G. Denote the representing vector measure by \({\breve{\mu }} \in {\mathcal {M}}(G,X)\) and \(T_\mu \circ T_H\) by \(T_{\breve{\mu }} \). Hence we immediately have that

$$\begin{aligned} \int _G f \,d{\breve{\mu }} = \int _{G/H} T_Hf \, d\mu \quad (f \in C(G)). \end{aligned}$$
(3)

Remark 2

Let \(\Phi :{\mathcal {M}}(G,X) \rightarrow {\mathcal {M}}(G/H,X)\) be defined by \(\Phi (\nu )={\tilde{\nu }} \) for \(\nu \in {\mathcal {M}}(G,X)\) and \(\Psi :{\mathcal {M}}(G/H,X) \rightarrow {\mathcal {M}}(G,X)\) by \(\Psi (\mu )={\breve{\mu }} \) for \(\mu \in {\mathcal {M}}(G/H,X)\). Then the following diagram commutes:

figure a

In other words, \(\Phi \circ \Psi = \text {Id}_{{\mathcal {M}}(G/H,X)}\) or equivalently \(\tilde{{\breve{\mu }} }=\mu \) for any \(\mu \in {\mathcal {M}}(G/H,X)\). This can be proved by observing that

$$\begin{aligned} \int _{G/H} \varphi \, d\tilde{{\breve{\mu }} } = \int _G \varphi _q \, d{\breve{\mu }} = \int _{G/H} T_H(\varphi _q) \, d\mu = \int _{G/H} \varphi \, d\mu \end{aligned}$$

for all \(\varphi \in C(G/H)\). Note that the commutativity of the diagram also implies that \(\Phi \) is surjective and \(\Psi \) is injective.

Proposition 3.5

Let \({\mathcal {R}} = \{R_h: h \in H\}\) and \(x^* \in B_{X^*}\). Then \({\breve{\mu }} \) and \(\vert {\langle {\breve{\mu }} , x^*\rangle } \vert \) are \({\mathcal {R}}\)-invariant.

Proof

To show that \({\breve{\mu }} \) is \({\mathcal {R}}\)-invariant, let \(h \in H\) and \(f\in C(G)\). Observe that

$$\begin{aligned} T_H(R_h f)(tH) = \int _H (R_hf)(th') \, dh' = \int _H f(th') \, dh' = T_H(f)(tH). \end{aligned}$$

Hence

$$\begin{aligned} T_{{\breve{\mu }} _{R_h}}( f) = T_{\breve{\mu }} (R_h f) = T_\mu (T_H(R_h f)) = T_\mu (T_H f) = T_{\breve{\mu }} (f). \end{aligned}$$

Hence \({\breve{\mu }} _{R_h}={\breve{\mu }} \), that is, \({\breve{\mu }} \) is \(R_h\)-invariant.

Now let \(x^* \in B_{X^*}\), \(E \in \mathfrak {B}(G)\) and \(h \in H\). For any disjoint partition \(\{E_n\}_{n=1}^{k}\) of E where \(E_n \in \mathfrak {B}(G)\), note that \(\{R_hE_n\}_{n=1}^{k}\) forms a disjoint partition of \(R_hE\) and

$$\begin{aligned} \sum _{n=1}^{k}\vert {\langle {\breve{\mu }} , x^*\rangle } (E_n)\vert = \sum _{n=1}^{k}\vert {\langle {\breve{\mu }} , x^*\rangle } (R_hE_n)\vert \le \vert {\langle {\breve{\mu }} , x^*\rangle } \vert (R_hE). \end{aligned}$$

Hence \(\vert {\langle {\breve{\mu }} , x^*\rangle } \vert (E) \le \vert {\langle {\breve{\mu }} , x^*\rangle } \vert (R_hE)\). Taking E as \(R_hE\) and h as \(h^{-1}\), we also get \(\vert {\langle {\breve{\mu }} , x^*\rangle } \vert (R_h E) \le \vert {\langle {\breve{\mu }} , x^*\rangle } \vert (E)\). \(\square \)

This proposition particularly implies that \({\breve{\mu }} \) is semivariation \({\mathcal {R}}\)-invariant. Hence we can apply Theorem 3.3 to get that the operator \(T_H\) has an extension to a norm-decreasing operator from \(L^p(G,{\breve{\mu }} )\) to \(L^p(G/H,\mu )\) for any \(1 \le p < \infty \). Moreover, Eq. (3) extends to \(L^1(G,{\breve{\mu }} )\)

$$\begin{aligned} \int _G f \,d{\breve{\mu }} = \int _{G/H} T_H f \, d\mu \quad \big (f \in L^1(G,{\breve{\mu }} )\big ). \end{aligned}$$
(4)

Indeed, if \(f_n \rightarrow f\) in \(L^1(G,{\breve{\mu }} )\) where \(f_n \in C(G)\), then \(I_\mu (T_H f) = \lim _{n\rightarrow \infty } T_\mu (T_H(f_n)) =\lim _{n\rightarrow \infty } T_{\breve{\mu }} (f_n)=I_{\breve{\mu }} (f)\). Now we prove that the Eq. (4) is also true for the total variation of the associated complex measures.

Lemma 3.6

For \(x^* \in B_{X^*}\) and \(f \in L^1(G,{\breve{\mu }} )\),

$$\begin{aligned} \int _G f \, d\vert {\langle {\breve{\mu }} , x^*\rangle } \vert = \int _{G/H} T_H f \, d\vert {\langle \mu , x^*\rangle } \vert . \end{aligned}$$

In particular, \(\Vert T_H \vert f \vert \Vert _{L^1(G/H,\mu )}=\Vert f\Vert _{L^1(G,{\breve{\mu }} )}\) for any \(f \in L^1(G,{\breve{\mu }} )\).

Proof

It suffices to prove that for any \(f \in L^1(G,{\breve{\mu }} )\) and \(x^* \in B_{X^*}\)

$$\begin{aligned} \int _G \vert f \vert \, d\vert {\langle {\breve{\mu }} , x^*\rangle } \vert = \int _{G/H} T_H \vert f\vert \, d\vert {\langle \mu , x^*\rangle } \vert . \end{aligned}$$

We first claim that for each \(E \in \mathfrak {B}(G)\), \(T_H(\chi _E) \ge 0\) \(\vert {\langle \mu , x^*\rangle } \vert \)-a.e that is the set \(F=\{tH \in G/H: T_H(\chi _E) < 0\}\) is \(\vert {\langle \mu , x^*\rangle } \vert \)-null. Let \(f_n \rightarrow \chi _E\) in \(L^1(G,{\breve{\mu }} )\) where \(f_n \in C(G)\) is positive (which exists by using Urysohn’s lemma together with the regularity of \({\breve{\mu }} \)). Since \(T_H(f_n) \ge 0\) for all \(n \in {\mathbb {N}}\),

$$\begin{aligned} \int _F \vert T_H(\chi _E)\vert \, d\vert {\langle \mu , x^*\rangle } \vert&\le \int _F \vert T_H(\chi _E)-T_H(f_n)\vert \, d\vert {\langle \mu , x^*\rangle } \vert \\&\le \Vert T_H(\chi _E - f_n)\Vert _{L^1(G/H,\mu )} \end{aligned}$$

which implies that \(\vert {\langle \mu , x^*\rangle } \vert (F) = 0\) as desired. Now fix \(E \in \mathfrak {B}(G)\) and consider any disjoint partition \(\{E_n\}_{n=1}^{k}\) of E where \(E_n \in \mathfrak {B}(G)\). By Eq. (4) and the claim,

$$\begin{aligned} \sum _{n=1}^k \vert {\langle {\breve{\mu }} , x^*\rangle } (E_n)\vert&= \sum _{n=1}^k \left| \int _{G/H} T_H(\chi _{E_n}) \, d{\langle \mu , x^*\rangle } \right| \\&\le \sum _{n=1}^k \int _{G/H} T_H(\chi _{E_n}) \, d\vert {\langle \mu , x^*\rangle } \vert \\&= \int _{G/H} T_H(\chi _E) \, d\vert {\langle \mu , x^*\rangle } \vert . \end{aligned}$$

Hence \(\int _G \chi _E \, d\vert {\langle {\breve{\mu }} , x^*\rangle } \vert \le \int _{G/H} T_H(\chi _E) \, d\vert {\langle \mu , x^*\rangle } \vert .\) It follows immediately that for any \(f \in S(G)\),

$$\begin{aligned} \int _G \vert f\vert \, d\vert {\langle {\breve{\mu }} , x^*\rangle } \vert \le \int _{G/H} T_H\vert f\vert \, d\vert {\langle \mu , x^*\rangle } \vert \end{aligned}$$

which can be extended to any \(f\in L^1(G,{\breve{\mu }} )\) by using the density of S(G) in \(L^1(G,{\breve{\mu }} )\).

Conversely, by Propositions 3.1.2 and 3.5, for \(f \in C(G)\)

$$\begin{aligned} \int _{G/H} T_H\vert f\vert \, d\vert {\langle \mu , x^*\rangle } \vert&\le \int _G (T_H\vert f\vert )_q \, d\vert {\langle {\breve{\mu }} , x^*\rangle } \vert \\&= \int _G \int _H \vert f(th)\vert \,dh \, d\vert {\langle {\breve{\mu }} , x^*\rangle } \vert (t) \\&= \int _H \int _G \vert f(th)\vert \, d\vert {\langle {\breve{\mu }} , x^*\rangle } \vert (t) \,dh \\&= \int _H \int _G \vert f(t)\vert \, d\vert {\langle {\breve{\mu }} , x^*\rangle } \vert (t) \,dh \\&= \int _G \vert f(t)\vert \, d\vert {\langle {\breve{\mu }} , x^*\rangle } \vert (t). \end{aligned}$$

Hence by the density of C(G) in \(L^1(G,{\breve{\mu }} )\), for any \(f \in L^1(G,{\breve{\mu }} )\)

$$\begin{aligned} \int _{G/H} T_H\vert f\vert \, d\vert {\langle \mu , x^*\rangle } \vert \le \int _G \vert f\vert \, d\vert {\langle {\breve{\mu }} , x^*\rangle } \vert . \end{aligned}$$

\(\square \)

For any \(\nu \in {\mathcal {M}}(G,X)\), we cannot find an example of an operator \(T_{H,\nu }\) constructed in the manner of Theorem 3.3 which is not surjective. However, we know that if \(\nu \) is in the form of \({\breve{\mu }} \), where \(\mu \in {\mathcal {M}}(G/H,X)\), then the operator \(T_{H,{\breve{\mu }} }\) is certainly surjective as shown in the following theorem.

Theorem 3.7

Let \(1 \le p < \infty \). The extension \(T_H:L^p(G,{\breve{\mu }} ) \rightarrow L^p(G/H,\mu )\) satisfies the formula \(T_H f(tH)=\int _H f(th) \, dh\) \(\mu \)-a.e. for all \(f \in L^p(G,{\breve{\mu }} )\). Moreover, the extension \(T_H:L^p(G,{\breve{\mu }} ) \rightarrow L^p(G/H,\mu )\) is surjective.

Proof

Claim that for a lower semicontinuous function \(\phi \ge 0\) and \(x^* \in B_{X^*}\),

$$\begin{aligned} \int _G \phi \,d\vert {\langle {\breve{\mu }} , x^*\rangle } \vert =\int _{G/H}\int _H \phi (th) \, dh \, d\vert {\langle \mu , x^*\rangle } \vert (tH). \end{aligned}$$

Let \(\Phi = \{g \in C(G): 0 \le g \le \phi \}\). By [10, Proposition 7.12] and Lemma 3.6,

$$\begin{aligned} \int _G \phi \, d\vert {\langle {\breve{\mu }} , x^*\rangle } \vert&= \sup _{g \in \Phi } \int _G g \, d\vert {\langle {\breve{\mu }} , x^*\rangle } \vert \\&= \sup _{g \in \Phi } \int _{G/H} T_H g \, d\vert {\langle \mu , x^*\rangle } \vert \ \\&= \int _{G/H} \left( \sup _{g \in \Phi } \, T_H g \right) d\vert {\langle \mu , x^*\rangle } \vert \\&= \int _{G/H} \left( \sup _{g \in \Phi } \int _H g(th) \, dh \right) d\vert {\langle \mu , x^*\rangle } \vert (tH) \\&= \int _{G/H} \left( \sup _{\tilde{g} \in \Phi (tH)} \int _H \tilde{g}(h) \, dh \right) d\vert {\langle \mu , x^*\rangle } \vert (tH) \\&= \int _{G/H} \int _H \phi (th) \, dh \, d\vert {\langle \mu , x^*\rangle } \vert (tH) \end{aligned}$$

where \(\Phi (tH):= \{\tilde{g} \in C(H): 0 \le \tilde{g}(h) \le \phi (th) \text { for } h \in H\}\). Hence for any measurable function F and any lower semicontinuous function \(\phi \ge \vert F\vert \),

$$\begin{aligned} \int _{G/H} \int _H \vert F(th)\vert \, dh \, d\vert {\langle \mu , x^*\rangle } \vert (tH)&\le \int _{G/H} \int _H \phi (th) \, dh \, d\vert {\langle \mu , x^*\rangle } \vert (tH) \\&= \int _G \phi \, d\vert {\langle {\breve{\mu }} , x^*\rangle } \vert . \end{aligned}$$

Hence by [10, Proposition 7.14]

$$\begin{aligned} \int _{G/H} \int _H \vert F(th)\vert \, dh \, d\vert {\langle \mu , x^*\rangle } \vert (tH) \le \int _G \vert F\vert \, d\vert {\langle {\breve{\mu }} , x^*\rangle } \vert . \end{aligned}$$
(5)

Let \(f \in L^p(G,{\breve{\mu }} )\) and \(f_n \rightarrow f\) in \(L^p(G,{\breve{\mu }} )\) where \(f_n \in C(G)\). Define a function \(\tilde{f}:G/H \rightarrow {\mathbb {C}}\) by \(\tilde{f}(tH) =\int _H f(th) \, dh\) for \(tH \in G/H\). By taking \(F=\vert f-f_n\vert ^p\) in (5), we have

$$\begin{aligned} \Vert \tilde{f} - T_H f_n \Vert ^p_{L^p(G/H,\mu )}&= \sup _{x^* \in B_{X^*}} \int _{G/H} \big \vert \tilde{f} - T_H f_n\big \vert ^p \,d\vert {\langle \mu , x^*\rangle } \vert \\&\le \sup _{x^* \in B_{X^*}} \int _{G/H} \int _H \vert f - f_n\vert ^p(th)\,dh \,d\vert {\langle \mu , x^*\rangle } \vert (tH) \\&\le \sup _{x^* \in B_{X^*}} \int _G \vert f-f_n\vert ^p \, d\vert {\langle {\breve{\mu }} , x^*\rangle } \vert \\&= \Vert f-f_n\Vert ^p_{L^p(G,{\breve{\mu }} )} \end{aligned}$$

which shows that \(\tilde{f}\) is well-defined and \(\tilde{f}=T_Hf\) \(\mu \)-a.e.

To show that \(T_H\) is surjective, we first claim that \(\Vert \phi _q\Vert _{L^p(G,{\breve{\mu }} )} = \Vert \phi \Vert _{L^p(G/H,\mu )}\) for \(\phi \in L^p(G/H,\mu )\). Let \(\phi _n \uparrow \vert \phi \vert \) pointwise where \(\phi _n \in S(G/H)\). Then \((\phi _n)_q \uparrow \vert \phi \vert _q\) pointwise. Applying the monotone convergence theorem and Lemma 3.6 to each \(x^* \in B_{X^*}\), we get

$$\begin{aligned} \int _G \vert \phi _q\vert ^p \, d\vert {\langle {\breve{\mu }} , x^*\rangle } \vert&= \lim _{n\rightarrow \infty } \int _G \vert (\phi _n)_q\vert ^p \, d\vert {\langle {\breve{\mu }} , x^*\rangle } \vert \\&= \lim _{n\rightarrow \infty } \int _{G/H} \vert \phi _n\vert ^p \, d\vert {\langle \mu , x^*\rangle } \vert \\&= \int _G \vert \phi \vert ^p \, d\vert {\langle \mu , x^*\rangle } \vert \end{aligned}$$

which proves the claim. Now let \(\varphi \in L^p(G/H,\mu )\). If we can show that \(\varphi _q \in L^p(G,{\breve{\mu }} )\), then by the formula of \(T_H\) we have \(T_H(\varphi _q)=\varphi \). Let \(\varphi _n \rightarrow \varphi \) in \(L^p(G/H,\mu )\) where \(\varphi _n \in S(G/H)\). Then it follows by the claim that \(\Vert \varphi _q - (\varphi _n)_q\Vert _{L^p(G,{\breve{\mu }} )} = \Vert \varphi -\varphi _n\Vert _{L^p(G/H,\mu )} \rightarrow 0\). Hence \(\varphi _q \in L^p(G,{\breve{\mu }} )\) by the completeness of \(L^p(G,{\breve{\mu }} )\). \(\square \)

Corollary 3.8

  1. 1.

    Weil’s formula holds for all \(f \in L^1(G,{\breve{\mu }} )\)

    $$\begin{aligned} \int _{G/H}\int _H f(th) \, dh \, d\mu (tH) = \int _G f \,d{\breve{\mu }} . \end{aligned}$$

    Moreover, for all \(x^* \in X^*\) and \(f \in L^1(G,{\breve{\mu }} )\)

    $$\begin{aligned} \int _{G/H}\int _H f(th) \, dh \, d\vert {\langle \mu , x^*\rangle } \vert (tH) = \int _G f \,d\vert {\langle {\breve{\mu }} , x^*\rangle } \vert . \end{aligned}$$
  2. 2.

    For \(1 \le p < \infty \), if \(\varphi \in L^p(G/H,\mu )\), then \(\varphi _q \in L^p(G,{\breve{\mu }} )\) with \(\Vert \varphi _q\Vert _{L^p(G,{\breve{\mu }} )} = \Vert \varphi \Vert _{L^p(G/H,\mu )}\).

Proof

The first two equations follow by applying the formula of \(T_H\) to Eq. (4) and Lemma 3.6 while the last assertion is in the proof of the theorem. \(\square \)

Corollary 3.9

For \(1 \le p < \infty \), C(G/H) is dense in \(L^p(G/H,\mu )\).

Proof

Let \(\varphi \in L^p(G/H,\mu )\). Then \(\varphi _q \in L^p(G,{\breve{\mu }} )\). By the density of C(G) in \(L^p(G,{\breve{\mu }} )\), there is a sequence \(f_n \rightarrow \varphi _q\) in \(L^p(G,{\breve{\mu }} )\) with \(f_n \in C(G)\). Hence \(T_H f_n \rightarrow T_H(\varphi _q)=\varphi \) in \(L^p(G/H,\mu )\). \(\square \)

It is straightforward to see that \(\breve{\widetilde{m}}=m\). Hence \(T_{H,m} = T_{H,\breve{\widetilde{m}}}\) is the same operator \(T_H\) given by Farashahi in [5]. Now we provide a relation between the extensions \(T_{H,m}\) and \(T_{H,{\breve{\mu }} }\).

Proposition 3.10

If \(\mu \ll \widetilde{m}\), then \(T_{H,m} f = T_{H,{\breve{\mu }} } f\) \(\mu \)-a.e. for all \(f \in L^1(G,m) \cap L^1(G,{\breve{\mu }} )\), and hence \({\breve{\mu }} \ll m\).

Proof

Let \(f \in L^1(G,m) \cap L^1(G,{\breve{\mu }} )\). Then \(T_{H,m} f(tH) = \int _H f(th) \, dh\) \(\widetilde{m}\)-a.e.; in particular, \(T_{H,m} f(tH) = \int _H f(th) \, dh\) \(\mu \)-a.e. since \(\mu \ll \widetilde{m}\). By Theorem 3.7, we also have \(T_{H,{\breve{\mu }} } f(tH) = \int _H f(th) \, dh\) \(\mu \)-a.e., so \(T_{H,m} f = T_{H,{\breve{\mu }} } f\) \(\mu \)-a.e.

Given \(E \in \mathfrak {B}(G)\) with \(m(E)=0\). Then \(T_{H,m} \chi _E = 0\) \(\widetilde{m}\)-a.e. and hence \(T_{H,{\breve{\mu }} } \chi _E = 0\) \(\mu \)-a.e. By Lemma 3.6, we get \(\Vert {\breve{\mu }} \Vert (E)=0\). We conclude that \({\breve{\mu }} \ll m\). \(\square \)

Example 2

Let \(1 \le p < \infty \) and \(S:L^p(G/H,\widetilde{m}) \rightarrow X\) be any bounded linear map. Define a vector measure \(\mu :\mathfrak {B}(G/H) \rightarrow X\) corresponding to S by \(\mu (E) = S(\chi _E)\) for \(E \in \mathfrak {B}(G/H)\). Then the vector measure \({\breve{\mu }} \) is given by \({\breve{\mu }} (F) = \int _{G/H} T_{H,{\breve{\mu }} }\chi _F \, d\mu \) for \(F \in {\mathcal {B}}(G)\). Note that for \(\varphi \in L^p(G/H,\widetilde{m})\), \(\varphi \) is \(\mu \)-integrable and \(\int _{G/H} \varphi \,d\mu = S(\varphi )\), see [15, Proposition 4.4]. Hence it follows from Proposition 3.10 that \({\breve{\mu }} (F) = \int _{G/H} T_{H,m}\chi _F \, d\mu = S(T_{H,m}\chi _F)\).

  1. 1.

    Let \(X = {\mathbb {C}}\) and \(S:L^1(G/H,\widetilde{m}) \rightarrow {\mathbb {C}}\) be given by \(S(\varphi )=\int _{G/H} \varphi \,d\widetilde{m}\) for any \(\varphi \in L^1(G/H,\widetilde{m})\). In this case, \(\mu = \widetilde{m}\). Moreover, \({\breve{\mu }} = m\) since \({\breve{\mu }} (F) = \int _{G/H} T_{H,m}\chi _F \, d\widetilde{m} = \int _G \chi _F \,dm = m(F)\) for all \(F \in \mathfrak {B}(G)\).

  2. 2.

    If \(X = L^1(G/H,\widetilde{m})\) and \(S = Id_{L^1(G/H,\widetilde{m})}\), then \({\breve{\mu }} (F)=T_{H,m}\chi _F\) for \(F \in \mathfrak {B}(G)\).

  3. 3.

    If we let \(1 \le p \le 2\) and define \(S:L^p(G/H,\widetilde{m}) \rightarrow \ell ^{p'}(\widehat{G/H};{\mathcal {B}}({\mathcal {H}}_\pi ))\) by \(S(\varphi )={\mathcal {F}}_{G/H}(\varphi )\), then \({\breve{\mu }} (F)(\pi )={\mathcal {F}}_{G/H}(T_{H,m}\chi _F)(\pi )=T^\pi _H\widehat{\chi _F}(\pi )\) for \(F \in \mathfrak {B}(G)\) and \([\pi ] \in \widehat{G/H}\), by [5, Proposition 5.3].

Finally, we give relations between \(\mu \) and \({\breve{\mu }} \) in terms of invariant properties.

Definition 6

Let \(\mu \) be a vector measure on G/H. For each \(a \in G\), \(\mu \) is said to be \(L_a\)-invariant if \(\mu (aE)=\mu (E)\) for all \(E \in \mathfrak {B}(G/H)\). We say that \(\mu \) is left invariant if it is \(L_a\)-invariant for all \(a \in G\).

Definition 7

Let \(\tau :G/H \rightarrow G/H\) be a homeomorphism. For any vector measure \(\mu \) on G/H, we say that \(\mu \) is semivariation \(\tau \)-invariant if

$$\begin{aligned} \Vert \varphi _\tau \Vert _{L^1(G/H,\mu )} =\Vert \varphi \Vert _{L^1(G/H,\mu )} \quad \text { for all } \varphi \in S(G/H). \end{aligned}$$

Given a collection \({\mathcal {T}}\) of homeomorphisms on G/H\(\mu \) is said to be semivariation \({\mathcal {T}}\)-invariant if it is semivariation \(\tau \)-invariant for all \(\tau \in {\mathcal {T}}\). In particular, if \({\mathcal {T}}=\{L_a: a\in G\}\), we say that \(\mu \) is semivariation left invariant.

Proposition 3.11

Let \(a \in G\).

  1. 1.

    \(\mu \) is \(L_a\)-invariant if and only if \({\breve{\mu }} \) is \(L_a\)-invariant.

  2. 2.

    \(\mu \) is norm integral \(L_a\)-invariant if and only if \({\breve{\mu }} \) is norm integral \(L_a\)-invariant.

  3. 3.

    \(\mu \) is semivariation \(L_a\)-invariant if and only if \({\breve{\mu }} \) is semivariation \(L_a\)-invariant.

Proof

  1. 1.

    Suppose that \(\mu \) is \(L_a\)-invariant. Then by the Weil formula (4), for any \(f \in C(G)\),

    $$\begin{aligned} \int _G L_a f \,d{\breve{\mu }}&= \int _{G/H} T_H (L_a f) \, d\mu = \int _{G/H} L_a (T_H f) \, d\mu \\&= \int _{G/H} T_H f \, d\mu = \int _G f \,d{\breve{\mu }} . \end{aligned}$$

    Hence \({\breve{\mu }} \) is \(L_a\)-invariant. Conversely, suppose that \({\breve{\mu }} \) is \(L_a\)-invariant. Then for any \(\varphi \in S(G/H)\)

    $$\begin{aligned} \int _{G/H} L_a \varphi \, d\mu = \int _G L_a \varphi _q \, d{\breve{\mu }} = \int _G \varphi _q \, d{\breve{\mu }} = \int _{G/H} \varphi \, d\mu \end{aligned}$$

    Hence \(\mu \) is \(L_a\)-invariant.

  2. 2.

    Suppose that \(\mu \) is norm integral \(L_a\)-invariant. Then by [2, Theorem 3.3], we have \(\Vert I_\mu (L_a \varphi )\Vert = \Vert I_\mu (\varphi )\Vert \) for all \(\varphi \in L^1(G/H,\mu )\). Hence by the Weil formula (4)

    $$\begin{aligned} \Vert I_{\breve{\mu }} (L_a f) \Vert = \Vert I_\mu (T_H(L_a f))\Vert = \Vert I_\mu (L_a T_Hf)\Vert = \Vert I_\mu (T_H f)\Vert = \Vert I_{\breve{\mu }} f \Vert \end{aligned}$$

    for any \(f \in S(G)\). Hence \({\breve{\mu }} \) is norm integral left invariant. The converse is proved in Proposition 3.2.

  3. 3.

    Suppose that \(\mu \) is semivariation \(L_a\)-invariant. It is routine to check that \(\Vert L_a \varphi \Vert _{L^1(G/H,\mu )} = \Vert \varphi \Vert _{L^1(G/H,\mu )}\) for all \(\varphi \in L^1(G/H,\mu )\). So

    $$\begin{aligned} \Vert L_a f\Vert _{L^1(G,{\breve{\mu }} )} = \Vert T_H\vert L_a f \vert \Vert _{L^1(G/H,\mu )} = \Vert T_H\vert f \vert \Vert _{L^1(G/H,\mu )} = \Vert f\Vert _{L^1(G,{\breve{\mu }} )} \end{aligned}$$

    for any \(f \in S(G)\). Conversely, if \({\breve{\mu }} \) is semivariation \(L_a\)-invariant then

    $$\begin{aligned} \Vert L_a \varphi \Vert _{L^1(G/H,\mu )} = \Vert L_a \varphi _q\Vert _{L^1(G,{\breve{\mu }} )} = \Vert \varphi _q\Vert _{L^1(G,{\breve{\mu }} )} = \Vert \varphi \Vert _{L^1(G/H,\mu )} \end{aligned}$$

    for any \(\varphi \in S(G/H)\).

\(\square \)

4 Invariant Measures

In this section, we provide properties of invariant measures on G and their analogies on G/H. The following proposition generalizes Proposition 5.2 in [1].

Proposition 4.1

Let \(\nu \in {\mathcal {M}}(G,X)\). The following are equivalent:

  1. 1.

    \(\nu \) is left (or right) invariant

  2. 2.

    \({\langle \nu , x^*\rangle } \) is left (or right) invariant for all \(x^* \in X^*\)

  3. 3.

    \(\nu = \nu (G)m\).

Proof

We only show that 2 implies 3; the other directions are trivial. Assume that \({\langle \nu , x^*\rangle } \) is left invariant for all \(x^* \in X^*\). Then the real part \({\langle \nu , x^*\rangle } _r\) is left invariant. Let \(G = P \cup N\) be a Hahn decomposition for \({\langle \nu , x^*\rangle } _r\) where P is positive and N is negative. Note that \(G = aP \cup aN\) is also a Hahn decomposition for \({\langle \nu , x^*\rangle } _r\) for any \(a \in G\). Hence \({\langle \nu , x^*\rangle } _r^+(aE) = {\langle \nu , x^*\rangle } _r(aE \cap aP) = {\langle \nu , x^*\rangle } _r(E \cap P) = {\langle \nu , x^*\rangle } _r^+(E)\) for any \(a \in G\) and \(E \in \mathfrak {B}(G)\). This shows that \({\langle \nu , x^*\rangle } _r^+\) is left invariant. By the uniqueness of the left Haar measure, \({\langle \nu , x^*\rangle } _r^+ =\alpha _r^+(x^*) m\) for some \(\alpha _r^+(x^*) \ge 0\). Applying the same argument to all parts of \({\langle \nu , x^*\rangle } \), we obtain that \({\langle \nu , x^*\rangle } =\alpha (x^*) m\) for some \(\alpha (x^*) \in {\mathbb {C}}\). Hence \(\langle \nu (E), x^* \rangle = \alpha (x^*) m(E)=\langle \nu (G), x^* \rangle m(E) =\langle \nu (G)m(E), x^* \rangle \) for any \(E \in \mathfrak {B}(G)\). Since this equation holds for all \(x^* \in X^*\), we have that \(\nu = \nu (G)m\). A similar argument can be applied to the case of right invariance. \(\square \)

Proposition 4.2

Let \(\mu \in {\mathcal {M}}(G/H,X)\). The following are equivalent:

  1. 1.

    \(\mu \) is left invariant

  2. 2.

    \({\langle \mu , x^*\rangle } \) is left invariant for all \(x^* \in X^*\)

  3. 3.

    \(\mu = \mu (G/H)\widetilde{m}\).

Proof

The first two assertions follow from the fact that \({\langle {\breve{\mu }} , x^*\rangle } = {\langle \mu , x^*\rangle } ^\smallsmile \) for all \(x^* \in X^*\). Next, assume that \(\mu \) is left invariant. Then \({\breve{\mu }} \) is left invariant. By Proposition 4.1, \({\breve{\mu }} ={\breve{\mu }} (G) m\). Since \(\mu \) and \(\widetilde{m}\) are the pushforward measures of \({\breve{\mu }} \) and m, we have \(\mu ={\breve{\mu }} (G) \widetilde{m}=\mu (G/H) \widetilde{m}\). This finishes the proof. \(\square \)

The following proposition improves Lemma 3.4 in [3].

Proposition 4.3

Let \(\nu \) be a vector measure on G. The following are equivalent.

  1. 1.

    \(\nu \) is norm integral left (or right) invariant.

  2. 2.

    For each \(x^* \in X^*\) and \(a \in G\), there exists \(x^*_a \in X^*\) such that \(\Vert x^*_a\Vert \le \Vert x^*\Vert \) and \({\langle \nu , x^*\rangle } (aE)=\langle \nu , x^*_{a}\rangle (E)\) (or \({\langle \nu , x^*\rangle } (Ea)=\langle \nu , x^*_{a}\rangle (E)\)) for all \(E \in {\mathcal {B}}(G)\).

Moreover, \(x^*_a \in X^*\) is unique in the sense that if there is another such functional then they must agree on \(I_\nu (S(G))\).

Proof

We shall prove only for the case of norm integral left invariance as the other case is similar. The proof of 1 implies 2 follows by the same argument of [3, Lemma 3.4]. For the converse, let \(f \in S(G)\) and \(a \in G\). Then

$$\begin{aligned} \Vert I_\nu (L_a f)\Vert = \sup _{x^* \in B_{X^*}} \vert x^*I_\nu (L_a f)\vert = \sup _{x^* \in B_{X^*}} \vert x^*_a I_\nu (f)\vert \le \Vert I_\nu (f)\Vert . \end{aligned}$$

This also implies \(\Vert I_\nu (f)\Vert = \Vert I_\nu (L_{a^{-1}} (L_a f))\Vert \le \Vert I_\nu (L_a f)\Vert \).

For the uniqueness, suppose there is another functional \(y^* \in X^*\) such that \(\Vert y^*\Vert \le \Vert x^*\Vert \) and \({\langle \nu , x^*\rangle } (aE)=\langle \nu , y^*\rangle (E)\) for all \(E \in {\mathcal {B}}(G).\) Then \(x^*_a(\nu (E))=\langle \nu , x^*_{a}\rangle (E)=\langle \nu , y^*\rangle (E)=y^*(\nu (E))\) for all \(E \in {\mathcal {B}}(G).\) By the linearity of \(x^*_a\) and \(y^*\), we have that \(x^*_a = y^*\) on \(I_\nu (S(G))\). \(\square \)

Proposition 4.4

Let \(\mu \) be a vector measure on G/H. The following are equivalent.

  1. 1.

    \(\mu \) is norm integral left invariant.

  2. 2.

    For each \(x^* \in X^*\) and \(a \in G\), there exists \(x^*_a \in X^*\) such that \(\Vert x^*_a\Vert \le \Vert x^*\Vert \) and \({\langle \mu , x^*\rangle } (aE)=\langle \mu , x^*_{a}\rangle (E)\) for all \(E \in {\mathcal {B}}(G/H)\).

Moreover, \(x^*_a \in X^*\) is unique in the sense that if there is another such functional then they must agree on \(I_\mu (S(G/H))\).

Proof

It can be proven by the same argument as in Proposition 4.3. However, if \(\mu \) is also assumed to be regular, we can employ Proposition 4.3 with \({\breve{\mu }} \) and obtain the result immediately. \(\square \)

The following result can be proved by the same argument as in [13, Theorem 5.6] and [1, Theorem 5.10]. Hence the proof is omitted.

Proposition 4.5

Let \(1 \le p < \infty \). Suppose that \(\nu \in {\mathcal {M}}(G,X)\) is semivariation left (or right) invariant with \(\nu (G) \ne 0\). Then \(L^p(G,\nu ) \subset L^p(G,m)\) with \(\Vert f\Vert _{L^p(G,m)} \le \Vert \nu (G)\Vert ^{-1/p} \Vert f\Vert _{L^p(G,\nu )}\) for \(f \in L^p(G,\nu )\).

Proposition 4.6

Let \(1 \le p < \infty \). Suppose that \(\mu \in {\mathcal {M}}(G/H,X)\) is semivariation left invariant with \(\mu (G/H) \ne 0\). Then \(L^p(G/H,\mu ) \subset L^p(G/H,\widetilde{m})\) with \(\Vert \varphi \Vert _{L^p(G/H,\widetilde{m})} \le \Vert \mu (G/H)\Vert ^{-1/p} \Vert \varphi \Vert _{L^p(G/H,\mu )}\) for \(\varphi \in L^p(G/H,\mu )\).

Proof

Since \({\breve{\mu }} \) is semivariation left invariant, by Proposition 4.5, \(L^p(G,{\breve{\mu }} ) \subset L^p(G,m)\) with \(\Vert f \Vert ^p_{L^p(G,m)} \le \Vert {\breve{\mu }} (G)\Vert ^{-1} \Vert f\Vert ^p_{L^p(G,{\breve{\mu }} )}\) for \(f \in L^p(G,{\breve{\mu }} ).\) Hence

$$\begin{aligned} \Vert \varphi \Vert ^p_{L^p(G/H,\widetilde{m})} = \Vert \varphi _q \Vert ^p_{L^p(G,m)}&\le \Vert {\breve{\mu }} (G)\Vert ^{-1} \Vert \varphi _q\Vert ^p_{L^p(G,{\breve{\mu }} )} \\&= \Vert \mu (G/H)\Vert ^{-1} \Vert \varphi \Vert ^p_{L^p(G/H,\mu )} \end{aligned}$$

for \(\varphi \in L^p(G/H,\mu )\). \(\square \)

5 Fourier Transforms

In this section, we define a Fourier transform of functions in \(L^1(G,\nu )\) and \(L^1(G/H,\mu )\). Our definition is motivated by Definition 4.1 in [13]; however, X is not considered as an operator space. Let \(\nu \) be a vector measure on G.

Definition 8

For \(f \in L^1(G,\nu )\) and \([\pi ] \in \widehat{G}\), we define the Fourier transform of f as

$$\begin{aligned} \widehat{f}^\nu (\pi ) = \int _{G} f(t)\pi (t)^* \, d\nu \in {\mathcal {B}}({\mathcal {H}}_\pi )\check{\otimes } X. \end{aligned}$$

To see that the definition is well-defined, we have to show that the function \(g:G \rightarrow {\mathcal {B}}({\mathcal {H}}_\pi )\) given by \(g(t) = f(t)\pi (t)^*\) is \(\nu \)-measurable and \(\check{\otimes }\)-integrable. Let \(x^* \in X^*\) be a Rybakov functional. Clearly, g is weakly \(\vert {\langle \nu , x^*\rangle } \vert \)-measurable since \(y^*g(\cdot ) = f(\cdot )y^*\pi (\cdot )^*\) is a product of \(\vert {\langle \nu , x^*\rangle } \vert \)-measurable functions for all \(y^* \in {\mathcal {B}}({\mathcal {H}}_\pi )^*\). Moreover, \({\mathcal {B}}({\mathcal {H}}_\pi )\) is separable. Thus, by Pettis’s measurability theorem, g is \(\vert {\langle \nu , x^*\rangle } \vert \)-measurable and hence is \(\nu \)-measurable. Since the function \(\Vert g\Vert =\vert f \vert \) is \(\nu \)-integrable, g is \(\check{\otimes }\)-integrable. This immediately implies the following proposition.

Proposition 5.1

Define the operator \({\mathcal {F}}^\nu _G: L^1(G,\nu ) \rightarrow \ell ^\infty (\widehat{G};{\mathcal {B}}({\mathcal {H}}_\pi )\check{\otimes } X )\) by \({\mathcal {F}}^\nu _G(f)(\pi )=\widehat{f}^\nu (\pi )\) for \(f \in L^1(G,\nu )\) and \([\pi ] \in \widehat{G}\). Then the Fourier transform operator \({\mathcal {F}}^\nu _G\) is bounded with \(\Vert \widehat{f}^\nu (\pi )\Vert _\vee \le \Vert f\Vert _{L^1(G,\nu )}\).

Remark 3

If we take \(\nu \) to be \({\langle \nu , x^*\rangle } \), then

$$\begin{aligned} \widehat{f}^{\, {\langle \nu , x^*\rangle } }(\pi ) = \int _{G} f(t)\pi (t)^* \, d{\langle \nu , x^*\rangle } = (Id_{{\mathcal {B}}({\mathcal {H}}_\pi )} \otimes x^*)\big (\widehat{f}^\nu (\pi )\big ). \end{aligned}$$

This can be considered as a generalization of Definition 4.6 in [13].

Remark 4

If G is abelian, then \({\mathcal {B}}({\mathcal {H}}_\pi ) \cong {\mathbb {C}}\) for any \([\pi ] \in \widehat{G}\). In this case, note that \({\mathbb {C}} \check{\otimes } X \cong X\) isometrically via the map \(\alpha \otimes x \mapsto \alpha x\) and \(\text {N}(\cdot )=\Vert \cdot \Vert _{L^1(G,\nu )}\). Hence our definition generalizes Definition 2.1 in [2].

Definition 9

We say that the Fourier transform \({\mathcal {F}}^\nu _G\) satisfies the Riemann–Lebesgue lemma if \({\mathcal {F}}^\nu _G(f) \in c_{0}(\widehat{G};{\mathcal {B}}({\mathcal {H}}_\pi )\check{\otimes } X )\) for all \(f \in L^1(G,\nu )\).

The Fourier transform \({\mathcal {F}}^\nu _G\) need not satisfy the Riemann–Lebesgue lemma even if G is abelian as shown in [2, Example 2.4]. Now we give a necessary condition for \({\mathcal {F}}^\nu _G\) to satisfy the Riemann–Lebesgue lemma and also a stronger condition for the sufficiency.

Theorem 5.2

Let \({\mathcal {M}}= \{ \pi _{ij}: [\pi ] \in \widehat{G}, 1 \le i,j \le d_\pi \}\).

  1. 1.

    If \({\mathcal {F}}^\nu _G\) satisfies the Riemann–Lebesgue lemma, then the set \(\{ \psi \in {\mathcal {M}}: \Vert \int _G \phi (t) \overline{\psi (t)} \, d\nu \Vert _X > \varepsilon \}\) is finite for any \(\varepsilon > 0\) and \(\phi \in {\mathcal {M}}\).

  2. 2.

    Moreover, if \(\nu \) is regular and \(\{ \psi =\pi _{ij}\in {\mathcal {M}}: d_\pi ^2\Vert \int _G \phi (t) \overline{\psi (t)} \, d\nu \Vert _X > \varepsilon \}\) is finite for any \(\varepsilon > 0\) and \(\phi \in {\mathcal {M}}\), then \({\mathcal {F}}^\nu _G\) satisfies the Riemann–Lebesgue lemma.

Proof

Observe that for \(F:G \rightarrow Y={\mathcal {B}}({\mathcal {H}}_\pi )\)

$$\begin{aligned} \left\| \int _G F \,d\nu \right\| _\vee = \sup _{y^* \in B_{Y^*}} \left\| \int _G y^*F \,d\nu \right\| _X. \end{aligned}$$

Hence if we write \(y^* \in B_{Y^*}\) as \(y^* = \sum _{1 \le i,j \le d_\pi } \alpha _{ij} e^*_{ij}\), we have

$$\begin{aligned} \max _{i,j}\left\| \int _G e^*_{ij}F \,d\nu \right\| _X \le \left\| \int _G F \,d\nu \right\| _\vee \le d_\pi ^2 \max _{i,j}\left\| \int _G e^*_{ij}F \,d\nu \right\| _X. \end{aligned}$$
  1. 1.

    Let \(\varepsilon > 0\) and \(\phi \in {\mathcal {M}} \subset L^1(G,\nu )\). Suppose that \({\mathcal {F}}^\nu _G\) satisfies the Riemann–Lebesgue lemma. If \(\pi _{ij} \in {\mathcal {M}}\) satisfies \(\Vert \int _G \phi (t) \overline{\pi _{ij}(t)} \, d\nu \Vert _X > \varepsilon \), by the observation above with \(F(t) = \phi (t) \pi (t)^*\), we have \(\Vert \widehat{\phi }^\nu (\pi )\Vert _\vee > \varepsilon \). Hence if \(\{ \psi \in {\mathcal {M}}: \Vert \int _G \phi (t) \overline{\psi (t)} \, d\nu \Vert _X > \varepsilon \}\) is infinite, then so does the set \(\{[\pi ] \in \widehat{G}: \Vert \widehat{\phi }^\nu (\pi ) \Vert _\vee > \varepsilon \}\), which is a contradiction.

  2. 2.

    Let \(\phi \in {\mathcal {M}}\) and \(\varepsilon > 0\). Suppose that \(\{ \psi \in {\mathcal {M}}: d_\pi ^2\Vert \int _G \phi (t) \overline{\psi (t)} \, d\nu \Vert _X > \varepsilon \}\) is finite. If \([\pi ] \in \widehat{G}\) satisfies \(\Vert \widehat{\phi }^\nu (\pi )\Vert _\vee > \varepsilon \), then \(d_\pi ^2 \Vert \int _G \phi (t) \overline{\pi _{ji}(t)} \, d\nu \Vert _X > \varepsilon \) for some ij. Hence we must have that \(\widehat{\phi }^\nu \in c_{0}(\widehat{G};{\mathcal {B}}({\mathcal {H}}_\pi )\check{\otimes } X )\). Note that the linear span of \({\mathcal {M}}\) is \(\text {Trig}(G)\) and \(\text {Trig}(G)\) is dense in \(L^1(G,\nu )\). By the continuity, \({\mathcal {F}}^\nu _G\) satisfies the Riemann–Lebesgue lemma.

\(\square \)

Remark 5

If G is abelian and \(\nu \) is regular, then \({\mathcal {F}}^\nu _G\) satisfies the Riemann–Lebesgue lemma if and only if the set \(\{ \psi \in {\mathcal {M}}: \Vert \int _G \phi (t) \overline{\psi (t)} \, d\nu \Vert _X > \varepsilon \}\) is finite for any \(\varepsilon > 0\) and \(\phi \in {\mathcal {M}}\).

We now prove the uniqueness theorem for the Fourier transform \({\mathcal {F}}^\nu _G\).

Theorem 5.3

Let \(\nu \in {\mathcal {M}}(G,X)\) and \(f \in L^1(G,\nu )\). If \(\widehat{f}^\nu (\pi )=0\) for all \([\pi ] \in \widehat{G}\), then \(f=0\) \(\nu \)-a.e.

Proof

Suppose that \(\widehat{f}^\nu (\pi )=0\) for all \([\pi ] \in \widehat{G}\). Fix a Rybakov functional \(x^* \in X^*\) and write \(d{\langle \nu , x^*\rangle } = g \, d\vert {\langle \nu , x^*\rangle } \vert \) where \(g \in L^1(G,\vert {\langle \nu , x^*\rangle } \vert )\). Then \(\int _G f(t) y^*\pi (t)^* \, d{\langle \nu , x^*\rangle } = 0\) for any \(y^* \in {\mathcal {B}}({\mathcal {H}}_\pi )^*\) and \([\pi ] \in \widehat{G}\). In particular, \(\int _G \overline{\pi _{ij}(t)} (fg)(t) \, d\vert {\langle \nu , x^*\rangle } \vert = 0\) for any \([\pi ] \in \widehat{G}\) and \(1 \le i,j \le d_\pi \). Since \(\overline{\pi _{ij}}\) is a matrix element of the contragradient representation of \(\pi \), \(\int _G \phi (f g) \, d\vert {\langle \nu , x^*\rangle } \vert =0\) for any \(\phi \in \text {Trig}(G)\). By the density of \(\text {Trig}(G)\) in C(G) in the uniform norm, \(fg \, d\vert {\langle \nu , x^*\rangle } \vert =0\) as a measure. Hence \(fg=0\) \(\vert {\langle \nu , x^*\rangle } \vert \)-a.e. However \(\vert g\vert = 1\) \(\vert {\langle \nu , x^*\rangle } \vert \)-a.e. Then it must be the case that \(f = 0\) \(\vert {\langle \nu , x^*\rangle } \vert \)-a.e. Therefore \(f=0\) \(\nu \)-a.e. since \(\nu \ll \vert {\langle \nu , x^*\rangle } \vert \). \(\square \)

Now we give a definition of a Fourier transform of functions on G/H with a vector measure. This definition is motivated by [5]. Let \(\mu \) be a vector measure on G/H.

Definition 10

For \(\varphi \in L^1(G/H,\mu )\) and \([\pi ] \in \widehat{G/H}\), we define the Fourier transform of \(\varphi \) at \([\pi ]\) as

$$\begin{aligned} \widehat{\varphi }^\mu (\pi ) = \int _{G/H} \varphi (tH)\Gamma _{\pi }(tH)^* \, d\mu (tH) \in {\mathcal {B}}({\mathcal {H}}_\pi )\check{\otimes } X, \end{aligned}$$

where \(\Gamma _{\pi }(tH)=\pi (t)T^\pi _H\).

Let \(g:G/H \rightarrow {\mathcal {B}}({\mathcal {H}}_\pi )\) be defined by \(g(tH)=\varphi (tH)\Gamma _{\pi }(tH)^*\) for \(tH \in G/H\). Then the \(\mu \)-measurability of g can be verified similarly to case of compact groups. Moreover, \( \Vert \Gamma _{\pi }(tH)\Vert ^2 = \Vert \Gamma _{\pi }(tH)^* \Gamma _{\pi }(tH)\Vert = \Vert (T^\pi _H)^*T^\pi _H\Vert = \Vert T^\pi _H\Vert ^2 = 1, \) so \(\Vert g\Vert \) is \(\mu \)-integrable. Hence the definition is well-defined.

Proposition 5.4

Define the operator \({\mathcal {F}}^\mu _{G/H}: L^1(G/H,\mu ) \rightarrow \ell ^\infty (\widehat{G/H};{\mathcal {B}}({\mathcal {H}}_\pi )\check{\otimes } X )\) by \({\mathcal {F}}^\mu _{G/H}(\varphi )(\pi )=\widehat{\varphi }^\nu (\pi )\) for \(\varphi \in L^1(G/H,\mu )\) and \([\pi ] \in \widehat{G/H}\). Then the Fourier transform operator \({\mathcal {F}}^\mu _{G/H}\) is bounded with \(\Vert {\mathcal {F}}^\mu _{G/H}(\varphi )(\pi )\Vert _\vee \le \Vert \varphi \Vert _{L^1(G/H,\mu )}\).

Proposition 5.5

Let \(\mu \in {\mathcal {M}}(G/H,X)\), \(\varphi \in L^1(G/H,\mu )\). Then \(\widehat{\varphi }^\mu (\pi ) = \widehat{\varphi _q}^{\breve{\mu }} (\pi )\) for each \([\pi ] \in \widehat{G/H}\).

Proof

Recall that \(T^\pi _H =\int _H \pi (h) \,dh\) is a bounded linear operator on \({\mathcal {H}}_\pi \) defined in the weak sense that is \(\langle T^\pi _H u,v \rangle = \int _H \langle \pi (h) u,v \rangle \,dh\) for \(u,v \in {\mathcal {H}}_\pi \). For \(y^* \in {\mathcal {B}}({\mathcal {H}}_\pi )^*\), write \(y^* = \sum _{i,j} \alpha _{ij} e_{ij}^*\). Since \(\int _H e_{ij}^*\pi (th)^* \,dh = e_{ij}^* \int _H \pi (th)^* \,dh = e_{ij}^* (T^\pi _H \pi (t)^*)\) for any ij, we have

$$\begin{aligned} T_H(y^*\pi (t)^*)&= \int _H y^*\pi (th)^* \,dh \\&= \sum _{i,j} \alpha _{ij} \int _H e_{ij}^*\pi (th)^* \,dh \\&= \sum _{i,j} \alpha _{ij} e_{ij}^* (T^\pi _H \pi (t)^* ) \\&= y^* (T^\pi _H \pi (t)^*) \end{aligned}$$

for any \(t \in G\). Hence \(T_H(y^*\pi (\cdot )^*) = y^* (T^\pi _H \pi (\cdot )^*)\). Consider for \(x^* \in X^*\) and \(y^* \in {\mathcal {B}}({\mathcal {H}}_\pi )^*\),

$$\begin{aligned} (y^* \otimes x^*)(\widehat{\varphi }^\mu (\pi ))&= \int _{G/H} \varphi (tH) y^*(T^\pi _H \pi (t)^*) \, d{\langle \mu , x^*\rangle } (tH) \\&= \int _{G/H} T_H(\varphi _q (\cdot ) y^*\pi (\cdot )^* )\, d{\langle \mu , x^*\rangle } (tH) \\&= \int _{G} \varphi _q(t) y^*\pi (t)^* \, d{\langle {\breve{\mu }} , x^*\rangle } (t) \\&= (y^* \otimes x^*)(\widehat{\varphi _q}^{\breve{\mu }} (\pi )). \end{aligned}$$

Hence \(\widehat{\varphi }^\mu (\pi ) = \widehat{\varphi _q}^{\breve{\mu }} (\pi )\). \(\square \)

Definition 11

We say that the Fourier transform \({\mathcal {F}}^\mu _{G/H}\) satisfies the Riemann–Lebesgue lemma if \({\mathcal {F}}^\mu _{G/H}(\varphi ) \in c_{0}(\widehat{G/H};{\mathcal {B}}({\mathcal {H}}_\pi )\check{\otimes } X )\) for all \(\varphi \in L^1(G/H,\mu )\).

Corollary 5.6

If \({\mathcal {F}}^{\breve{\mu }} _G\) satisfies the Riemann–Lebesgue lemma, then so does \({\mathcal {F}}^\mu _{G/H}\).

The Fourier transform \({\mathcal {F}}^\mu _{G/H}\) also satisfies the uniqueness theorem.

Theorem 5.7

Let \(\mu \in {\mathcal {M}}(G/H,X)\) and \(\varphi \in L^1(G/H,\mu )\). If \(\widehat{\varphi }^\mu (\pi )=0\) for all \([\pi ] \in \widehat{G/H}\), then \(\varphi =0\) \(\mu \)-a.e.

Proof

Suppose that \(\widehat{\varphi }^\mu (\pi )=0\) for all \([\pi ] \in \widehat{G/H}\). Then \(\widehat{\varphi _q}^{\breve{\mu }} (\pi )=0\) for all \([\pi ] \in \widehat{G/H}\). Moreover, if \([\pi ] \in \widehat{G}\) but \([\pi ] \notin \widehat{G/H}\), then \(\widehat{\varphi _q}^{\breve{\mu }} (\pi )=0\). Indeed, for any \(x^* \in X^*\) and \(y^* \in {\mathcal {B}}({\mathcal {H}}_\pi )^*\),

$$\begin{aligned} (y^* \otimes x^*)(\widehat{\varphi _q}^{\breve{\mu }} (\pi ))&= \int _{G} \varphi _q(t) y^*\pi (t)^* \, d{\langle {\breve{\mu }} , x^*\rangle } (t) \\&= \int _{G/H} \varphi (tH) y^*(T^\pi _H \pi (t)^*) \, d{\langle \mu , x^*\rangle } (tH) = 0 \end{aligned}$$

since \(T_H(y^*\pi (\cdot )^*) = y^* (T^\pi _H \pi (\cdot )^*)=0\). Then one can apply Theorem 5.3 and obtains that \(\varphi _q = 0\) \({\breve{\mu }} \)-a.e. Hence \(\varphi = T_H(\varphi _q) = 0\) \(\mu \)-a.e. \(\square \)