
In this paper, the authors construct the Morrey spaces Mλp(V) on the metric measure spaces (V,d,μ) and explore their associated properties. Here, (V,d,μ) refers to metric measure spaces composed of infinite homogeneous trees T equipped with the standard graph metric d and a weighted measure μ. As applications, the boundedness of the maximal operator and the fractional maximal operator related to admissible trapezoids on the Morrey spaces Mλp(V) is established.
Citation: Xiaoyu Qian, Jiang Zhou. Morrey spaces on weighted homogeneous trees[J]. AIMS Mathematics, 2025, 10(4): 7664-7683. doi: 10.3934/math.2025351
[1] | Suixin He, Shuangping Tao . Boundedness of some operators on grand generalized Morrey spaces over non-homogeneous spaces. AIMS Mathematics, 2022, 7(1): 1000-1014. doi: 10.3934/math.2022060 |
[2] | Heng Yang, Jiang Zhou . Compactness of commutators of fractional integral operators on ball Banach function spaces. AIMS Mathematics, 2024, 9(2): 3126-3149. doi: 10.3934/math.2024152 |
[3] | Naqash Sarfraz, Muhammad Bilal Riaz, Qasim Ali Malik . Some new characterizations of boundedness of commutators of $p$-adic maximal-type functions on $p$-adic Morrey spaces in terms of Lipschitz spaces. AIMS Mathematics, 2024, 9(7): 19756-19770. doi: 10.3934/math.2024964 |
[4] | Yueping Zhu, Yan Tang, Lixin Jiang . Boundedness of multilinear Calderón-Zygmund singular operators on weighted Lebesgue spaces and Morrey-Herz spaces with variable exponents. AIMS Mathematics, 2021, 6(10): 11246-11262. doi: 10.3934/math.2021652 |
[5] | Ming Liu, Bin Zhang, Xiaobin Yao . Weighted variable Morrey-Herz space estimates for $ m $th order commutators of $ n- $dimensional fractional Hardy operators. AIMS Mathematics, 2023, 8(9): 20063-20079. doi: 10.3934/math.20231022 |
[6] | Mingquan Wei, Xiaoyu Liu . Sharp weak bounds for discrete Hardy operator on discrete central Morrey spaces. AIMS Mathematics, 2023, 8(2): 5007-5015. doi: 10.3934/math.2023250 |
[7] | Javeria Younas, Amjad Hussain, Hadil Alhazmi, A. F. Aljohani, Ilyas Khan . BMO estimates for commutators of the rough fractional Hausdorff operator on grand-variable-Herz-Morrey spaces. AIMS Mathematics, 2024, 9(9): 23434-23448. doi: 10.3934/math.20241139 |
[8] | Wanjing Zhang, Suixin He, Jing Zhang . Boundedness of sublinear operators on weighted grand Herz-Morrey spaces. AIMS Mathematics, 2023, 8(8): 17381-17401. doi: 10.3934/math.2023888 |
[9] | Kieu Huu Dung, Do Lu Cong Minh, Pham Thi Kim Thuy . Commutators of Hardy-Cesàro operators on Morrey-Herz spaces with variable exponents. AIMS Mathematics, 2022, 7(10): 19147-19166. doi: 10.3934/math.20221051 |
[10] | Yanlong Shi, Xiangxing Tao . Rough fractional integral and its multilinear commutators on $ p $-adic generalized Morrey spaces. AIMS Mathematics, 2023, 8(7): 17012-17026. doi: 10.3934/math.2023868 |
In this paper, the authors construct the Morrey spaces Mλp(V) on the metric measure spaces (V,d,μ) and explore their associated properties. Here, (V,d,μ) refers to metric measure spaces composed of infinite homogeneous trees T equipped with the standard graph metric d and a weighted measure μ. As applications, the boundedness of the maximal operator and the fractional maximal operator related to admissible trapezoids on the Morrey spaces Mλp(V) is established.
In 1741, Euler [1] laid the foundations of graph theory in solving the famous Königsberg Seven Bridges problem. A series of studies on graphs laid the foundation for the development of tree theory, in which trees are acyclic connected graphs. In 1857, Cayley [2] introduced the concept of trees while studying saturated hydrocarbon isomers and the counting problems of acyclic graphs. In the 20th century, trees were gradually applied to fields such as signal processing and image analysis, as referenced in [3,4].
In recent decades, the theory of function spaces and operators in harmonic analysis based on trees has attracted considerable attention. In 1986, Korányi and Picardello [5] investigated the Laplace operator on homogeneous trees and the boundary behavior of its eigenfunctions. The spectral properties of Schrödinger operators on homogeneous root metric trees were studied in 2002 by Sobolev and Solomyak [6].
The metric measure spaces (V,d,μ) grow exponentially and do not satisfy the doubling condition. Therefore, the classical Calderón-Zygmund decomposition theory is not applicable in this setting. In 2003, Hebisch and Steger [7] presented an abstract Calderón-Zygmund decomposition theory, which applied to spaces (V,d,μ). They applied this theory to obtain the weak (1,1) boundedness of the maximal operator related to admissible trapezoids, as well as the properties of multipliers and singular integrals. In 2020, Arditti, Tabacco, and Vallarino [8] introduced the atomic Hardy space H1 based on space (V,d,μ). In 2021, Arditti, Tabacco, and Vallarino [9] studied the BMO space defined on space (V,d,μ) and proved that BMO is the dual space of H1.
In 1938, the classical Morrey spaces were introduced by Morrey [10] during his study of the regularity of solutions to second order elliptic partial differential equations. They can be regarded as a significant and natural extension of Lebesgue spaces. For more work related to the classical Morrey spaces in this paper, see references [11,12,13]. In 1997, Arai and Mizuhara [14] defined Morrey spaces on spaces of homogeneous type, which are metric measure spaces (X,d,μ) that satisfy the doubling condition. That is, there exists a constant C such that
μ(B(x,2r))≤Cμ(B(x,r)),∀x∈X,r>0. |
Moreover, Morrey spaces have also been investigated in the setting of metric measure spaces [15,16,17,18] that do not satisfy the doubling condition.
Due to the important significance of Morrey spaces, many scholars have begun to extend them to discrete metric spaces in recent years. In 2019, Gunawan and Schwanke [19] proved the boundedness of the Hardy-Littlewood maximal operator on discrete Morrey spaces by using a discrete Fefferman-Stein inequality. In 2021, Zhang, Liu, and Zhang [20] presented the Morrey spaces on infinite connected graphs and established the boundedness of the Hardy-Littlewood maximal operators and their fractional variations on these spaces.
Let us discuss several open problems related to the theory of maximal operators on weighted homogeneous trees, which can guide further research. One significant direction involves investigating the Fefferman-Stein inequality on such trees, as well as norm estimates related to sharp maximal functions and maximal functions. Another research direction is to investigate the boundedness of Riesz transforms and spectral multipliers of the flow Laplacian in Morrey spaces on weighted homogeneous trees, where the Lp estimates for these operators on weighted homogeneous trees are already known.
The aim of this work is to study the Morrey spaces on (V,d,μ). The structure of this paper is as follows: In Section 2, we review the definitions of weighted homogeneous trees and admissible trapezoids, examine their corresponding geometric properties, and introduce Morrey spaces and fractional maximal operators associated with admissible trapezoids. In Section 3, we define the preduals of Morrey spaces in terms of block spaces and investigate the boundedness of the maximal operators on these block spaces. Finally, we investigate the boundedness of both the maximal operator and the fractional maximal operator on Morrey spaces.
The symbol N denotes the set of all positive integers. The statement A≲B indicates that A≤CB, where C is any independent positive constant, and A≈B indicates that A≲B≲A. The characteristic function of a measurable set E is denoted by χE. For 1≤p≤∞, p′ is the conjugate index of p, which satisfies the relation 1p+1p′=1.
In this subsection, we first review the definition of the infinite homogeneous tree, as well as the distance d and measure μ defined in this space.
Definition 2.1. [7] An infinite homogeneous tree of order q+1 is a graph T=(V,E) that satisfies the following properties:
(ⅰ) T is connected and acyclic;
(ⅱ) Each vertex has exactly q+1 neighbors, q∈N,
where V denotes the set of vertices and E denotes the set of edges. The distance d(x,y) of x,y∈V is the length of the shortest path connecting x and y.
Definition 2.2. [7] Let T=(V,E) be an infinite homogeneous tree, a doubly infinite geodesic g in T is a connected subset g⊂V such that
(ⅰ) For any element v∈g, there are precisely two neighbors of v in g;
(ⅱ) For any two vertices u, v∈g, the shortest path joining u and v is contained within g.
Definition 2.3. [7] Define a mapping F: g→Z such that
|F(x)−F(y)|=d(x,y)∀x,y∈g. |
The level function ℓ:V→Z is defined as
ℓ(x)=F(x′)−d(x,x′), |
where x′ is the only vertex in g for which
d(x,x′)=min{d(x,z):z∈g}. |
This implies choosing an orientation for g and a unique origin o∈g, where F(o)=0. In this way, a numbering system for vertices in g is established.
Definition 2.4. [7] For any x,y∈V, it is said that y lies above x or x lies below y, if
ℓ(x)=ℓ(y)−d(x,y). |
Definition 2.5. [7] T=(V,E) is an infinite homogeneous tree of order q+1; Each f: V→R is a non-negative measurable function; Let μ be the measure on V such that
∫Vfdμ=∑x∈Vf(x)qℓ(x). |
Remark 2.1. Based on the above, μ is a weighted counting measure in which a vertex's weight is only reliant on its level, and the weight associated with a certain level is given by q times the weight of the level immediately below it.
Lemma 2.1. [8] Let T=(V,E) be an infinite homogeneous tree of order q+1, x0∈V, and r∈N. Consider the closed ball Br(x0)={x∈V:d(x,x0)≤r}. Its measure is derived as
μ(Br(x0))=qℓ(x0)qr+1+qr−2q−1. |
It is observed that the measures depend on the level of the center x0 and increase exponentially with the radius r.
A weighted homogeneous tree (V,d,μ) does not satisfy the doubling condition, so it is useful to introduce a suitable class of subsets of V called admissible trapezoids. Admissible trapezoids and Calderón-Zygmund sets are introduced in [7]. In this subsection, we review their definitions and related properties.
Definition 2.6. [7] A subset R⊂V is called an admissible trapezoid if it satisfies one of the following conditions:
(ⅰ) R consists of just one vertex, i.e., R={xR}, where xR∈V;
(ⅱ) There exists xR∈V and h(R)∈N such that
R={x∈V:x lies below xR,h≤ℓ(xR)−ℓ(x)<2h}. |
We specify that in the first case h(R)=1 and in the second case h(R)=h. Here xR is referred to as the root of the trapezoid. In both cases mentioned above, h(R) is referred to as the height of the admissible trapezoid R and coincides with the number of levels spanned by the admissible trapezoid R. The quantity w(R)=qℓ(xR) is referred to as the width of the admissible trapezoid R. It follows that
μ(R)=h(R)qℓ(xR)=h(R)w(R). |
R is denoted as the set of all admissible trapezoids. The admissible trapezoid R of height h(R)=2 with xR as the root on a degree q=2 weighted homogeneous tree, as shown in Figure 1.
Definition 2.7. [7] Let R be an admissible trapezoid. Define its envelope ˜R as follows:
(ⅰ) If R consists of just one vertex, then ˜R=R;
(ⅱ) Otherwise,
˜R={x∈V:x lies below xR,h2≤ℓ(xR)−ℓ(x)<4h}. |
The envelope of an admissible trapezoid is known as a Calderón-Zygmund set. We specify that h(˜R)=h(R). ˜R is the set of all Calderón-Zygmund sets. The geometric relationship between the admissible trapezoid R and its envelope ˜R, as shown in Figure 2.
The measure of the set consisting of all points x∈R such that ℓ(x)=ℓ(xR)−hi is given by
∑x∈R,ℓ(x)=ℓ(xR)−hiqℓ(x)=qhiqℓ(xR)−hi=qℓ(xR). |
Therefore, the measure of each level within the trapezoid R is the same, and it is qℓ(xR). By definition, it can also be obtained that the measure of each level of ˜R is qℓ(xR).
Lemma 2.2. [8] Take an admissible trapezoid R. Then it holds
μ(˜R)≤4μ(R). |
Lemma 2.3. [8] Take two admissible trapezoids R1 and R2, if R1∩R2≠∅andw(R1)≥w(R2), then
R2⊂˜R1. |
Lemma 2.4. [21] For any R∈R and its envelope ˜R, there exist three admissible trapezoids S1,S2, and S3 such that ˜R⊂S1∪S2∪S3 and the following properties hold:
(ⅰ) w(Sk)>w(R) for k=1,2,3;
(ⅱ) μ(Sk)≤2qμ(˜R) for k=1,2,3.
In the following proposition, we will construct a covering of V using an increasing family of admissible trapezoids.
Proposition 2.1. There exists a family of admissible trapezoids {Rn}∞n=0 such that Rn⊂Rn+1 and ⋃nRn=V.
Proof. Consider the family {Rn}∞n=0 where
● R0 is an admissible trapezoid with the root at x0 and height h=1. Here, x0 represents the unique vertex in the double undirected geodesic g such that ℓ(o)=0.
● For any n≥1, Rn represents the admissible trapezoid with a root node xn. Here, xn serves as the grandfather node of xn−1, positioned such that xn lies above xn−1. Additionally, it satisfies the conditions ℓ(xn)−ℓ(xn−1)=3 and height h(Rn)−h(Rn−1)=2. Then there is h(Rn)=2n+1.
Firstly, we demonstrate that Rn⊂Rn+1. Take x∈Rn. According to the definition, if x is below xn, then it also lies below xn+1 by construction. Furthermore, we have ℓ(xn+1)−ℓ(x)=ℓ(xn)+3−ℓ(x), and it follows that
ℓ(xn+1)−ℓ(x)<2h(Rn)+3<2h(Rn+1), |
ℓ(xn+1)−ℓ(x)≥h(Rn)+3≥h(Rn+1). |
Therefore, we can conclude that x∈Rn+1.
To demonstrate that ⋃nRn=V, let us consider an element x∈V. We denote by k the smallest index such that x lies below xk (and so x lies below xj for all j≥k) and set ℓ(xk)−ℓ(x)=d. We are looking for an index j≥k such that x∈Rj, which can be expressed as
2j+1=h(Rj)≤ℓ(xj)−ℓ(x)<2h(Rj)=4j+2. |
Note that ℓ(xj)−ℓ(x)=ℓ(xj)−ℓ(xk)+ℓ(xk)−ℓ(x)=3(j−k)+d, which implies
2j+1≤3(j−k)+d<4j+2⟺{j>d−3k−2j≥1−d+3k. |
Therefore, it is sufficient to choose j≥max{k,d−3k−1,1−d+3k}.
In this subsection, we review the definition of the maximal operator and define the fractional maximal operator associated with the family of admissible trapezoids.
Definition 2.8. [8] Given f: V→R, the maximal operator M is defined as follows:
Mf(x)=supR:x∈R1μ(R)∫R|f(y)|dμ(y)=supR:x∈R1μ(R)∑x∈R|f(y)|qℓ(y), |
where the supremum is taken over all admissible trapezoids R containing x.
Definition 2.9. Let 0<α<1. Given f: V→R, we define the fractional maximal operator Mα as follows:
Mαf(x)=supR:x∈R1μ(R)1−α∫R|f(y)|dμ(y)=supR:x∈R1μ(R)1−α∑x∈R|f(y)|qℓ(y), |
where the supremum is taken over all admissible trapezoids R containing x.
Lemma 2.5. [8] The maximal operator M is of weak type (1,1). That is, for any f∈L1(V) and β>0,
μ({x∈V:Mf(x)>β})≤C‖f‖L1(V)β. |
Utilizing Marcinkiewicz interpolation theorem on measure spaces [22], it can be demonstrated that M is (p,p) bounded, where 1<p<∞.
In this subsection, we introduced two types of Morrey spaces. One is related to admissible trapezoids, and the other is associated with Calderón-Zygmund sets. Subsequently, we proved the equivalence of the norms of these two Morrey spaces. We also explored some properties of Morrey spaces. Let Lploc(V) denote the set of all functions f that satisfy ∫R|f(x)|pdμ(x)<∞ for any R∈R.
Definition 2.10. Let 1≤p,λ≤∞. The Morrey spaces are
Mλp(V)={f∈Lploc(V):‖f‖Mλp<∞}, |
where
‖f‖Mλp=supR∈Rμ(R)1λ−1p(∫R|f(x)|pdμ(x))1p=supR∈Rμ(R)1λ−1p(∑x∈R|f(x)|pqℓ(x))1p. |
Remark 2.2. If p=λ<∞, then Mλp(V)=Lp(V). When p≤λ=∞, Mλp(V)=L∞(V). Additionally, if λ<p, then Mλp(V)={0}. Therefore, we will only consider the case where 1≤p≤λ<∞.
Definition 2.11. Let 1≤p≤λ<∞. The Morrey spaces are
˜Mλp(V)={f∈Lploc(V):‖f‖˜Mλp<∞}, |
where
‖f‖˜Mλp=sup˜R∈˜Rμ(˜R)1λ−1p(∫˜R|f(x)|pdμ(x))1p=sup˜R∈˜Rμ(˜R)1λ−1p(∑x∈R|f(x)|pqℓ(x))1p. |
Proposition 2.2. Let 1≤p≤λ<∞ and f∈Lploc(V). Then,
‖f‖Mλp≈‖f‖˜Mλp. |
Proof. According to Lemma 2.2, it is clear that
μ(R)1λ−1p(∑x∈R|f(x)|pqℓ(x))1p≤41p−1λμ(˜R)1λ−1p(∑x∈˜R|f(x)|pqℓ(x))1p≲‖f‖˜Mλp. |
For the opposite estimate, by Lemma 2.4, we obtain
μ(˜R)1λ−1p(∑x∈˜R|f(x)|pqℓ(x))1p≤μ(˜R)1λ−1p(∑x∈S1∪S2∪S3|f(x)|pqℓ(x))1p≤3∑k=1(2q)1p−1λμ(Sk)1λ−1p(∑x∈Sk|f(x)|pqℓ(x))1p≲‖f‖Mλp, |
where Sk is the admissible trapezoid, k=1,2,3.
Lemma 2.6. For any 1≤p1≤p2<∞, q∈N,
(1μ(R)∑x∈R|f(x)|p1qℓ(x))1p1≤(1μ(R)∑x∈R|f(x)|p2qℓ(x))1p2, |
where f:V→R and R∈R.
Proof. Let 1≤p1≤p2<∞ and q∈N. According to Hölder's inequality, we have
∑x∈R|f(x)|p1qℓ(x)≤(∑x∈R|f(x)|p2qℓ(x))p1p2(∑x∈Rqℓ(x))1−p1p2=μ(R)1−p1p2(∑x∈R|f(x)|p2qℓ(x))p1p2=μ(R)(1μ(R)∑x∈R|f(x)|p2qℓ(x))p1p2. |
Further simplification gives
1μ(R)∑x∈R|f(x)|p1qℓ(x)≤(1μ(R)∑x∈R|f(x)|p2qℓ(x))p1p2. |
This concludes the proof.
Proposition 2.3. If 1≤p1≤p2≤λ<∞. Then Mλp2(V)⊂Mλp1(V) with
‖f‖Mλp1≤‖f‖Mλp2, |
for any f∈Mλp2(V).
Lemma 2.7. Let 1≤p≤λ<∞. Then, for any R0∈R, we have χR0∈Mλp(V).
Proof. For any R∈R, we will consider the following two distinct cases.
If μ(R)>1, we can derive
μ(R)1λ−1p(∑x∈R|χR0(x)|pqℓ(x))1p≤(∑x∈R|χR0(x)|pqℓ(x))1p≤(∑x∈V|χR0(x)|pqℓ(x))1p=μ(R0)1p. |
If μ(R)≤1, it can be inferred that
μ(R)1λ−1p(∑x∈R|χR0(x)|pqℓ(x))1p≤μ(R)1λ−1p(∑x∈V|χR(x)|pqℓ(x))1p=μ(R)1λ. |
Our proof is complete.
The predual spaces of classical Morrey spaces, weighted product Morrey spaces [13], and non-double Morrey spaces [15] are all block Spaces. In this subsection, we naturally define block spaces that are actually the predual spaces of Mλp(V). Furthermore, it is shown that the maximal operator M is bounded on the block spaces.
Definition 3.1. Suppose 1≤λ′≤p′<∞; A measurable function b is classified as a (λ′,p′)-block if it holds that suppb⊂R, R∈R, and
‖b‖Lp′≤μ(R)1p′−1λ′. |
Define the block space Bλ′p′(V) as
Bλ′p′(V)={∞∑k=1λkbk:∞∑k=1|λk|<∞,each bkis a (λ′,p′)−block}. |
The block space Bλ′p′(V) is equipped with the following norm:
‖f‖Bλ′p′=inf{∞∑k=1|λk|:f=∞∑k=1λkbk,each bkis a (λ′,p′)−block}. |
We call Bλ′p′(V) the block space associated with Lp′(V).
Theorem 3.1. Suppose 1<p≤λ<∞, 1p+1p′=1, then we have
(Bλ′p′(V))∗=Mλp(V). |
This indicates that Bλ′p′(V) represents the predual space of Mλp(V).
Proof. Let b be a (λ′,p′)-block, supported on an admissible trapezoid R. For any f∈Mλp(V), the Hölder's inequality of weighted homogeneous tree yields
∑x∈V|f(x)b(x)|qℓ(x)≤(∑x∈R|f(x)|pqℓ(x))1p(∑x∈R|b(x)|p′qℓ(x))1p′=μ(R)1λ−1p(∑x∈R|f(x)|pqℓ(x))1pμ(R)1p−1λ(∑x∈R|b(x)|p′qℓ(x))1p′≲‖f‖Mλp. |
Therefore, for any g=∑∞k=1λkbk∈Bλ′p′(V), we derive the following inequality:
∑x∈V|f(x)g(x)|qℓ(x)≲∞∑k=1|λk|∑x∈V|f(x)bk(x)|qℓ(x)≲‖g‖Bλ′p′‖f‖Mλp. | (3.1) |
Consequently, it follows that Mλp(V)↪(Bλ′p′(V))∗.
To prove the embedding in the reverse case, we only need to show that any element in (Bλ′p′(V))∗ can be regarded as a function of Mλp(V).
First, we prove that any bounded linear functional on Bλ′p′(V) can be represented by a function from Lploc(V). For g∈Lp′loc(V),
G=gχR‖gχR‖Lp′μ(R)1p−1λ. |
Then G is a (λ′,p′)-block. According to the definition of block space, for any (λ′,p′)-block b, we have ‖b‖Bλ′p′≤1, as a result
‖G‖Bλ′p′≤1. | (3.2) |
That means,
‖gχR‖Bλ′p′≤‖gχR‖Lp′μ(R)1p−1λ. | (3.3) |
In accordance with formula (3.3), L∈(Bλ′p′(V))∗,
|L(gχR)|≤‖L‖(Bλ′p′)∗‖gχR‖Bλ′p′≲‖gχR‖Lp′μ(R)1p−1λ. | (3.4) |
For R∈R, consider the set X={gχR:g∈Lp′(V)}, which defines X as a subspace of Lp′(V). For R∈R and L∈(Bλ′p′(V))∗, we can define a linear functional l:X→C by
l(h)=L(gχR), |
where h=gχR is an element of X and g belongs to Lp′(V).
In accordance with (3.4), it is established that l is bounded on X. The Hahn-Banach theorem guarantees the extension of l as a member of (Lp′(V))∗. The duality (Lp′(V))∗=Lp(V) yields a function fR∈Lp(V) such that
l(g)=∑x∈VfR(x)g(x)qℓ(x),∀g∈Lp′(V). |
Without loss of generality, we assume that suppfR⊂R. Based on Proposition 2.1, let ⋃nRn=V, such that R⊂Ri⊂Ri+1, define f(x)=fRi(x) if x∈Ri, which makes sense since
∑x∈R|fRi|qℓ(x)=l(χR)=∑x∈R|fRj|qℓ(x). |
Consequently, for x∈R⊂Ri⊂Ri+1, it follows that fRi=fRi+1. Furthermore, there exists a unique measurable function f=fR defined on any set R.
Next, we continue to prove that f∈Mλp(V). Let's consider a set R and an index j, such that R⊂Rj. For h∈Lp′(V),
H=hχR‖hχR‖Lp′μ(R)1p−1λ |
is a (λ′,p′)-block. In the light of (3.2), we have ‖H‖Bλ′p′≤1, which implies:
‖hχR‖Bλ′p′≤‖hχR‖Lp′μ(R)1p−1λ. |
We obtain
μ(R)1λ−1p‖fχR‖Lp=μ(R)1λ−1psup‖h‖Lp′=1|∑x∈Vf(x)h(x)qℓ(x)|=sup‖h‖Lp′=1|∑x∈RjfR(x)h(x)χR(x)μ(R)1p−1λqℓ(x)|≲‖L‖(Bλ′p′)∗sup‖h‖Lp′=1‖hχRμ(R)1p−1λ‖Bλ′p′≤‖L‖(Bλ′p′)∗. |
We can conclude that Lf=L and (Bλ′p′(V))∗↪Mλp(V). That completes the proof.
Corollary 3.1. Let 1<p≤λ<∞. Then f∈Mλp(V) if and only if for all g∈Bλ′p′(V), it holds that
∑x∈Vf(x)g(x)qℓ(x)<∞. |
Moreover, for all f∈Mλp(V),
‖f‖Mλp∼sup{∑x∈Vf(x)g(x)qℓ(x):g∈Bλ′p′(V),‖g‖Bλ′p′≤1}. |
Proposition 3.1. Let 1≤λ′≤p′<∞. If f∈ Bλ′p′(V) and |h|≤|f|, then h∈ Bλ′p′(V).
Proof. Given f∈Bλ′p′(V), for any ϵ>0, there exists a family of (λ′,p′)-blocks {bi}∞i=1 and a family of scalars {λi}∞i=1 such that f=∑∞i=1λibi and
∞∑i=1|λi|≤(1+ϵ)‖f‖Bλ′p′. |
We express g=∑∞i=1λici, where
ci(z)={h(z)f(z)bi(z),if f(z)≠0;0,if f(z)=0. |
Under the condition that |h|≤|f|, the sequence {ci}∞i=1 is a set of (λ′,p′)-blocks. Consequently, it can be inferred that h∈ Bλ′p′(V). Since ϵ is arbitrary, we can conclude that ‖h‖Bλ′p′(V)≤‖f‖Bλ′p′(V).
Theorem 3.2. Let 1<λ′≤p′<∞. Then, Bλ′p′(V)⊂Lloc(V) and Bλ′p′(V)is a Banach space.
Proof. By utilizing Lemma 2.7, it follows that χR∈Mλp(V) for all R∈R. Furthermore, Theorem 3.1 ensures that χR∈(Bλ′p′(V))∗. According to the norm conjugate formula (3.1), it yields
∑x∈R|h(x)|qℓ(x)≲‖χR‖Mλp‖h‖Bλ′p′, | (3.5) |
for any h∈Bλ′p′(V). This means that Bλ′p′(V)↪Lloc(V).
To prove Bλ′p′(V) is a banach space, take hi∈Bλ′p′(V), i∈N satisfying
∞∑i=1‖hi‖Bλ′p′<∞. |
For any ϵ>0, there exists a positive integer N such that for all integers n>N, it holds that
∞∑i=n‖hi‖Bλ′p′<ϵ. | (3.6) |
In light of (3.5), for any R∈R, it follows that
∑x∈R∞∑i=1|hi(x)|qℓ(x)≲‖χR‖Mλp(∞∑i=1‖hi‖Bλ′p′). |
Consequently, the function h=∑∞i=1hi is well-defined as a Lebesgue measurable function and belongs to Lloc(V).
Our next demonstration illustrates that h=∑∞i=1hi belongs to Bλ′p′(V). Based on the definition of Bλ′p′(V), for any ϵ>0, we have that
hi=∞∑k=1λk,ihk,i, |
where hk,i with i,k∈N are (λ′,p′)-blocks, along with
∞∑k=1|λk,i|≤(1+ϵ)‖hi‖Bλ′p′. |
In addition, for any 1≤i≤N, there exists a Ni∈N such that
‖hi−Ni∑k=1λk,ihk,i‖Bλ′p′≤∞∑k=Ni+1|λk,i|<2−iϵ. | (3.7) |
As a result, for any R∈R, we can deduce that
∑x∈R|h(x)−N∑i=1Ni∑k=1λk,ihk,i(x)|qℓ(x)≲∑x∈R|h(x)−N∑i=1hi(z)|qℓ(x)+∑x∈R|N∑i=1hi(x)−N∑i=1Ni∑k=1λk,ihk,i(x)|qℓ(x)≲∑x∈R∞∑i=N+1|hi(x)|qℓ(x)+N∑i=1∑x∈R|hi(x)−Ni∑k=1λk,ihk,i(x)|qℓ(x). |
Combining (3.5) and (3.6) with (3.7) yields that
∑x∈R|h(x)−N∑i=1Ni∑k=1λk,ihk,i(x)|qℓ(x)≲‖χR‖Mλp(∞∑i=N+1‖hi‖Bλ′p′+N∑i=1‖hi−Ni∑k=1λk,ihk,i‖Bλ′p′)≲‖χR‖Mλp(ϵ+N∑i=12−iϵ)≲‖χR‖Mλpϵ. |
Thus, ∑∞i=1∑∞k=1λk,ihk,i converges to h in Lloc(V). As a result, ∑∞i=1∑∞k=1λk,ihk,i converges to h in the sense of local measures. Therefore, there exists a subsequence {∑Ni=1∑Mk=1λk,ihk,i}N,M which converges to h a.e. Furthermore, λk,i with i,k∈N satisfies
∞∑i=1∞∑k=1|λk,i|≤∞∑i=1(1+ϵ)‖hi‖Bλ′p′. |
∑∞i=1hi converges to h in Bλ′p′(V). As a result, Bλ′p′(V) is a Banach space.
Theorem 3.3. Let 1<λ′≤p′<∞ and R∈R. Then, the maximal operator M is bounded on the space Bλ′p′(V).
In order to prove Theorem 3.3, it is only necessary to prove the following.
Lemma 3.1. Suppose b is a (λ′,p′)-block and R, S∈R. Then, there exists a sequence Sl,m∈R, where l∈N and m=1,2,⋯,N, along with a sequence of measurable functions hl,m that possess the following properties.
(ⅰ) The pointwise estimate Mb(x)≤∑l∈N∑Nm=1hl,m(x),x∈V∖˜R holds.
(ⅱ) μ(˜Sl,m)∼2lμ(R), where the implicit constants do not depend on R and l.
(ⅲ) 0≤hl,m(x)≤‖b‖12l−1μ(R)χ˜Sl,m(x)for x∈V.
In particular, there exists C=Cλ,p so that, for every (λ′,p′)-block b,
‖Mb‖Bλ′p′≤C. |
Indeed, upon accepting Lemma 3.1, we can establish Theorem 3.3 through the following approach. First, assume that f∈Bλ′p′(V). Consequently, we can identify a sequence of (λ′,p′)-block denoted as {bj}j∈N and a corresponding sequence of coefficients {λj}j∈N, such that
f=∞∑j=1λjbj,∞∑j=1|λj|≤2‖f‖Bλ′p′. |
By Lemma 3.1, the function g=∑∞j=1|λj|Mbj satisfies
|f(x)|≤g(x) and ‖g‖Bλ′p′≤C∞∑j=1|λj|≤C‖f‖Bλ′p′. |
According to Proposition 3.1, we conclude that
‖Mf‖Bλ′p′≤‖g‖Bλ′p′≤C‖f‖Bλ′p′. |
Proof of Lemma 3.1.
Proof. First construct the desired admissible trapezoid and functions. Let x∈V∖˜R and S,R∈R such that w(S)>w(R), with the condition that suppb⊂R. If S∩R≠∅, then R⊂˜S. Therefore, the following can be obtained:
Mb(x)≲supx∈S‖b‖1μ(S)≲supx∈S4‖b‖1μ(˜S), |
where x∈V∖˜R. Set
Sl:={S∈R:R⊂˜S,2l−1μ(R)≤μ(˜S)<2lμ(R)} |
for l∈N. Since supS∈Slμ(s)<∞, we can apply [23, Lemma 4] to get the existence of sets Sl,1,…,Sl,N∈Sl, such that
⋃S∈SlS⊂N⋃m=1˜Sl,m |
holds, where N is independent of l. We define
hl,m:=χ˜Sl,m∖˜RMb. |
Then we obtain
0≤hl,m(x)≤4‖b‖12l−1μ(R), |
for x∈V. Consequently, the decomposition stated in Lemma 3.1 is obtained.
We decomposed Mb into two parts by ˜R, that is, Mb=B1+B2, where B1=χ˜RMb and B2=χV∖˜RMb. This means that the estimate is divided into ‖B1‖Bλ′p′ and ‖B2‖Bλ′p′ estimates. Based on Lemma 2.5, we know that the operator M is bounded on Lp′(V). This implies that
‖B1‖p′≤‖Mb‖p′≤C0‖b‖p′≤C0μ(R)1p′−1λ′. |
Here, the operator norm is defined as C0:=‖M‖Lp′⟶Lp′. Then we see that B1C0 is a (λ′,p′)-block. Therefore, the estimate of ‖B1‖Bλ′p′ is deemed valid. By utilizing Hölder's inequality, we can discern that
‖hl,m‖p′≤C‖b‖1μ(˜Sl,m)1p′2l−1μ(R)≤Cμ(˜Sl,m)1p′μ(R)1λ2l−1μ(R)≤Cμ(˜Sl,m)1p′−1λ′2l−1λ. |
Thus, we deduce that 2l−1λhl,mC constitutes a (λ′,p′)-block. Given the pointwise estimate |B2(x)|≤∑l∈N∑Nm=1hl,m(x), it follows that B2 belongs to the space Bλ′p′(V), with its norm defined by an absolute constant.
In this subsection, we focus primarily on the maximal operator M and the fractional maximal operator Mα. We present proof of the boundedness of these operators in Morrey spaces. We select an addmissible trapezoid R and decompose the function space based on its envelope ˜R.
Theorem 3.4. Let 1<p≤λ<∞, for all f∈Mλp(V), it holds that Mf∈Mλp(V), and there exists a constant C>0 independent of f such that
‖Mf‖Mλp≲‖f‖Mλp. |
Proof. We just need to demonstrate that
μ(R)1λ−1p(∑x∈R|Mf(x)|pqℓ(x))1p≲‖f‖Mλp(R∈R), |
according to the definition. Write f(x)=f1(x)+f2(x), where f1=fχ˜R and f2=fχV∖˜R, we deduce that
μ(R)1λ−1p(∑x∈R|Mf(x)|pqℓ(x))1p≲μ(R)1λ−1p(∑x∈R|Mf1(x)|pqℓ(x))1p+μ(R)1λ−1p(∑x∈R|Mf2(x)|pqℓ(x))1p=:I+II. |
For the first part I, by expanding the integral domain and taking advantage of the boundedness of M on Lp(V), we can infer that
I≲μ(R)1λ−1p(∑x∈V|Mf1(x)|pqℓ(x))1p≲μ(R)1λ−1p(∑x∈V|f1(x)|pqℓ(x))1p=μ(R)1λ−1p(∑x∈˜R|f1(x)|pqℓ(x))1p. |
Based on Lemmas 2.2 and 2.4, we obtain
I≲μ(R)1λ−1p(∑x∈S1∪S2∪S3|f(x)|pqℓ(x))1p≲41p−1λμ(˜R)1λ−1p(∑x∈S1∪S2∪S3|f(x)|pqℓ(x))1p≲3∑k=1(8q)1p−1λμ(Sk)1λ−1p(∑x∈Sk|f(x)|pqℓ(x))1p≲‖f‖Mλp, |
and the proof of the first part is now complete.
For the second part II, we estimate
Mf2(x)=supS:x∈S1μ(S)∫S|f2(y)|dμ(y). |
When (V∖˜R)∩S=∅ or R∩S=∅, the norm result naturally holds. Let's discuss the alternative case in which (V∖˜R)∩S≠∅ and R∩S≠∅. Since the measure of every level on R and ˜R is the same as that of qℓ(xR), we have qℓ(xS)>qℓ(xR), that is, w(S)>w(R). With the help of Lemma 2.3, it can be deduced that R⊂˜S. Therefore,
Mf2(x)≲supS:x∈S∩Rw(S)>w(R)1μ(S)∑y∈S|f2(y)|qℓ(y)≲supS:x∈S∩Rw(S)>w(R)1μ(S)∑y∈S|f(y)|qℓ(y)≲supR⊂˜Sw(S)>w(R)1μ(S)∑y∈S|f(y)|qℓ(y), |
it follows that
II≲μ(R)1λ−1pμ(R)1psupR⊂˜Sw(S)>w(R)1μ(S)∑y∈S|f(y)|qℓ(y). |
By utilizing Hölder's inequality and Lemma 2.2, we obtain
II≲μ(R)1λsupR⊂˜Sw(S)>w(R)(1μ(S)∑y∈S|f(y)|pqℓ(y))1p≲supR⊂˜Sw(S)>w(R)μ(˜S)1λμ(S)−1p(∑y∈S|f(y)|pqℓ(y))1p≲supS41λμ(S)1λ−1p(∑y∈S|f(y)|pqℓ(y))1p≲‖f‖Mλp, |
and all the proofs have been completed.
Remark 3.1. In [19], H. Gunawan and C. Schwanke investigated the boundedness of maximal operators on Morrey spaces over the set of Zd (d=1,2,…). However, it is noteworthy that in the context of our discrete tree, when the degree of the tree is 2, it corresponds to their case of Z, i.e., the results for d=1. Therefore, our results are more general in the one-dimensional setting.
Lemma 3.2. If 1<p≤λ<∞, 0<α<1 and λ<1α, then
Mαf(x)≲‖f‖MλpMf(x)1−λα, |
for every f∈Mλp(V).
Proof. For each x∈V, we denote tx=‖f‖λMλp/Mf(x)λ. We divide the estimate into two parts, namely
Mαf(x)≲supR:x∈Rμ(R)≤tx1μ(R)1−α∑y∈R|f(y)|qℓ(y)+supR:x∈Rμ(R)>tx1μ(R)1−α∑y∈R|f(y)|qℓ(y):=L+LL. |
For L, we have that
L≲supR:x∈Rμ(R)≤txtαx1μ(R)∑y∈R|f(y)|qℓ(y)≲tαxMf(x)=‖f‖λαMλpMf(x)1−λα. |
For LL, if μ(R)>tx, there exists i∈N such that 2i−1tx≤μ(R)≤2itx. Applying Hölder's inequality, we obtain
LL≲supR:x∈Rμ(R)>tx(2i−1tx)α−1∑y∈R|f(y)|qℓ(y)≲supR:x∈Rμ(R)>tx(2i−1tx)α−1μ(R)1−1p(∑y∈R|f(y)|pqℓ(y))1p. |
According to 1<λ and λ<1α, it can be inferred that
LL≲supR:x∈Rμ(R)>tx(2i−1tx)α−1μ(R)1λ−1p(∑y∈R|f(y)|pqℓ(y))1pμ(R)1−1λ≲‖f‖Mλpsupi∈N(2i−1tx)α−1(2itx)1−1λ≲‖f‖Mλpsupi∈N(2i)α−1λ tα−1λx≲‖f‖λαMλpMf(x)1−λα. |
The proof of the lemma is completed.
Theorem 3.5. If 1<s≤r<∞, 1<p≤λ<∞, α=1λ−1r and pλ=sr, then the operator Mα is bounded from Mλp(V) to Mrs(V).
Proof. We write
‖Mαf‖Mrs=supR∈Rμ(R)1r−1s(∑x∈R|Mαf(x)|sqℓ(x))1s, |
by Lemma 3.2, we obtain
‖Mαf‖Mrs≲‖f‖λαMλpsupR∈Rμ(R)1r−1s(∑x∈R|Mf(x)|(1−λα)sqℓ(x))1s≲‖f‖λαMλp{supR∈Rμ(R)1(1−λα)(1r−1s)(∑x∈R|Mf(x)|(1−λα)sqℓ(x))1s(1−λα)}1−λα≲‖f‖Mλp, |
since α=1λ−1r and pλ=sr.
This paper defines the Morrey space Mλp(V) associated with admissible trapezoids on weighted homogeneous trees and investigates the properties of these spaces. We also study the boundedness of the maximal operator and fractional maximal operator on the Morrey space Mλp(V). By constructing the predual space of Morrey space Mλp(V), we reveal their structure and analyze their applications in function spaces and operator theory. Our study extends the theory of classical Morrey spaces and applies it to metric measure spaces that do not satisfy the doubling condition.
Xiaoyu Qian: Writing-original draft, investigation, formal analysis; Jiang Zhou: Writing-review and editing, investigation. All authors have read and approved the final version of the manuscript for publication.
The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.
The authors want to express their sincere thanks to the editors and referees for the valuable remarks and suggestions. This work was supported by the National Natural Science Foundation of China (No.12461021).
The authors declare that there is no conflicts of interest.
[1] | L. Euler, Solutio problematis ad geometriam situs pertinentis, Comm. Acad. Sci. Petrop., 8 (1741), 128–140. |
[2] |
A. Cayley, On the theory of the analytical forms called trees, Coll. Math. Pap., 3 (1890), 242–246. https://doi.org/10.1017/CBO9780511703690.046 doi: 10.1017/CBO9780511703690.046
![]() |
[3] |
A. Haar, Zur Theorie der orthogonalen Funktionensysteme, Math. Ann., 69 (1910), 331–371. https://doi.org/10.1007/BF01456326 doi: 10.1007/BF01456326
![]() |
[4] | L. R. Foulds, Graph theory applications, New York: Springer, 1992. https://doi.org/10.1007/978-1-4612-0933-1 |
[5] | A. Korányi, M. A. Picardello, Boundary behaviour of eigenfunctions of the Laplace operator on trees, Ann. Scuola Norm. Sup. Pisa Cl. Sci., 13 (1986), 389–399. |
[6] |
A. V. Sobolev, M. Z. Solomyak, Schrödinger operators on homogeneous metric trees: Spectrum in gaps, Rev. Math. Phys., 14 (2002), 421–467. https://doi.org/10.1142/S0129055X02001235 doi: 10.1142/S0129055X02001235
![]() |
[7] |
W. Hebisch, T. Steger, Multipliers and singular integrals on exponential growth groups, Math. Z., 245 (2003), 37–61. https://doi.org/10.1007/s00209-003-0510-6 doi: 10.1007/s00209-003-0510-6
![]() |
[8] | L. Arditti, A. Tabacco, M. Vallarino, Hardy spaces on weighted homogeneous trees, Advances in microlocal and time-frequency analysis, Cham: Birkhäuser., 2020. https://doi.org/10.1007/978-3-030-36138-9_2 |
[9] |
L. Arditti, A. Tabacco, M. Vallarino, BMO spaces on weighted homogeneous trees, J. Geom. Anal., 31 (2021), 8832–8849. https://doi.org/10.1007/s12220-020-00435-w doi: 10.1007/s12220-020-00435-w
![]() |
[10] | C. B. Morrey, On the solutions of quasi-linear elliptic partial differential equations, Trans. Amer. Math. Soc., 43 (1938), 126–166. |
[11] | F. M. Chiarenza, M. Frasca, Morrey spaces and Hardy-Littlewood maximal function, Rend. Mat. Appl., 7 (1987), 273–279. |
[12] |
J. Tao, D. C. Yang, W. Yuan, A bridge connecting Lebesgue and Morrey spaces via Riesz norms, Banach J. Math. Anal., 15 (2021), 1–29. https://doi.org/10.1007/s43037-020-00106-6 doi: 10.1007/s43037-020-00106-6
![]() |
[13] |
M. Q. Wei, Extrapolation for weighted product Morrey spaces and some applications, Potential Anal., 60 (2024), 445–472. https://doi.org/10.1007/s11118-022-10056-3 doi: 10.1007/s11118-022-10056-3
![]() |
[14] |
H. Arai, T. Mizuhara, Morrey spaces on spaces of homogeneous type and estimates for ◻b and the Cauchy-Szegő projection, Math. Nachr., 185 (1997), 5–20. https://doi.org/10.1002/mana.3211850102 doi: 10.1002/mana.3211850102
![]() |
[15] |
Y. Sawano, H. Tanaka, Predual spaces of Morrey spaces with non-doubling measures, Tokyo J. Math., 32 (2009), 471–486. https://doi.org/10.3836/tjm/1264170244 doi: 10.3836/tjm/1264170244
![]() |
[16] |
G. H. Lu, S. P. Tao, Generalized Morrey spaces over nonhomogeneous metric measure spaces, J. Aust. Math. Soc., 103 (2017), 268–278. https://doi.org/10.1017/S1446788716000483 doi: 10.1017/S1446788716000483
![]() |
[17] |
Y. H. Cao, J. Zhou, Morrey spaces for nonhomogeneous metric measure spaces, Abstr. Appl. Anal., 2013 (2013), 1–8. https://doi.org/10.1155/2013/196459 doi: 10.1155/2013/196459
![]() |
[18] |
L. Liu, Y. Sawano, D. C. Yang, Morrey-type spaces on Gauss measure spaces and boundedness of singular integrals, J. Geom. Anal., 24 (2014), 1007–1051. https://doi.org/10.1007/s12220-012-9362-9 doi: 10.1007/s12220-012-9362-9
![]() |
[19] |
H. Gunawan, C. M. Schwanke, The Hardy-Littlewood maximal operator on discrete Morrey spaces, Mediterr. J. Math., 16 (2019), 1–12. https://doi.org/10.1007/s00009-018-1277-7 doi: 10.1007/s00009-018-1277-7
![]() |
[20] |
X. Zhang, F. Liu, H. Zhang, Mapping properties of maximal operators on infinite connected graphs, J. Math. Inequal., 15 (2021), 1613–1636. https://doi.org/10.7153/jmi-2021-15-111 doi: 10.7153/jmi-2021-15-111
![]() |
[21] | L. Arditti, Analysis on weighted homogeneous trees, Master's thesis, 2018. |
[22] | L. Grafakos, Classical Fourier analysis, 3 Eds., New York: Springer, 2024.10. https://doi.org/10.1007/978-1-4939-1194-3 |
[23] |
Y. Zhang, J. Zhou, Variable Lebesgue space over weighted homogeneous tree, Symmetry, 16 (2024), 1–15. https://doi.org/10.3390/sym16101283 doi: 10.3390/sym16101283
![]() |