1 Introduction

Consider a Riemannian manifold M with gradient \(\nabla \) and Laplace–Beltrami operator \(\varDelta \). The Riesz transform \(\nabla \varDelta ^{-1/2}\), with \(\varDelta ^{-1/2}\) defined via the spectral theorem, maps \(L^2(M)\) boundedly to the space of square integrable vector fields \(L^2(M;TM)\). Much attention has been given to the question of whether this operator extends to a bounded map from \(L^p(M)\) to \(L^p(M;TM)\) for \(p \ne 2\), or equivalently, whether the estimate

$$\begin{aligned} (R_p): \quad \Vert |\nabla f| \Vert _p \lesssim \Vert \varDelta ^{1/2} f \Vert _p \qquad \hbox {for all } f \in C_c^\infty (M) \end{aligned}$$

holds. It is conjectured that for \(p \in (1,2)\) the estimate \((R_p)\) holds whenever M is complete, with implicit constant depending only on p; in [1] it is shown that the failure of this uniformity among manifolds of a fixed dimension would imply the existence of a manifold for which \((R_p)\) fails.

One is naturally led to consider also the ‘reverse’ estimate

$$\begin{aligned} (RR_p): \quad \Vert \varDelta ^{1/2} f \Vert _p \lesssim \Vert |\nabla f| \Vert _p \qquad \hbox {for all} f \in C_c^\infty (M). \end{aligned}$$

A duality argument shows that for \(p \in (1,\infty )\), \((R_p)\) implies \((RR_{p^\prime })\), where \(p^\prime = p/(p-1)\) is the Hölder conjugate of p. If \((R_p)\) and \((RR_p)\) both hold, then we have a norm equivalence

$$\begin{aligned} \Vert |\nabla f| \Vert _p \simeq \Vert \varDelta ^{1/2} f \Vert _p, \end{aligned}$$

which says that the homogeneous Sobolev space \(\dot{W}^p_1(M)\) may be defined either via the gradient or via the square root of the Laplace–Beltrami operator.

Generally \((R_p)\) holds only for some interval of \(p \in (1,\infty )\) including 2, and proving \((R_p)\) presents different difficulties depending on whether \(p < 2\) or \(p > 2\). When \(1< p < 2\), \((R_p)\) is known to follow from the volume doubling property and Gaussian or sub-Gaussian heat kernel upper estimates [9, 10] (see also [15] for examples which do not satisfy such kernel estimates). The volume doubling property and an appropriately scaled \(L^2\)-Poincaré inequality imply \((R_p)\) for some \(p > 2\) [2]. In [3] this is linked with gradient estimates for the heat kernel, and in [5] the \(L^2\)-Poincaré inequality is replaced by a relative Faber–Krahn inequality and a reverse Hölder inequality.

Some manifolds for which \((R_p)\) fails for some \(p > 2\) are known. If M is the connected sum of two copies of \(\mathbb {R}^n \setminus B(0,1)\) with \(n \ge 3\) —or more generally, an n-dimensional manifold with at least two (and finitely many) Euclidean ends—\((R_p)\) holds if and only if \(p \in (1,n)\) [7, 10]. Similar results are known for conical manifolds [14] and for 2-hyperbolic, p-parabolic manifolds with at least two ends [6].

The most relevant examples to this article are Vicsek manifolds, which are ‘thickenings’ of Vicsek graphs (pictured in the 2-dimensional case in Fig. 2). The Vicsek graph, being a graphical realisation of a Vicsek fractal, is a ‘fractal at infinity’. Locally a Vicsek manifold behaves like Euclidean space (it is, of course, a manifold), but at large scale it behaves like a fractal. In [9] it is shown that for a Vicsek manifold of any dimension, \((R_p)\) holds if and only if \(p \in (1,2]\). The result for \(p < 2\) is a consequence of volume doubling and sub-Gaussian heat kernel estimates. The proof that \((R_p)\) fails for \(p > 2\) directly uses the definition of the Vicsek graph [9, Theorem 5.1].

In this article we construct a class of manifolds of arbitrary dimension for which \((R_p)\) fails for all \(p > 2\).Footnote 1 These manifolds are thickenings of what we call spinal graphs, satisfying generalised dimension conditions defined in terms of the spinal structure along with a polynomial volume lower bound. The Vicsek graphs satisfy these conditions, but the proof of this exploits their fractal nature. We construct a large class of non-fractal spinal graphs with the desired dimension conditions and volume lower bounds, thus yielding manifolds of arbitrary dimension with no fractal structure that fail \((R_p)\) for all \(p > 2\).

2 Notation

The graphs we consider are non-directed, with at most one edge per pair of vertices, and with no edges from a vertex to itself. The set of vertices of a graph G is denoted by V(G), and if two vertices \(x,y \in V(G)\) are neighbours we write \(x \sim y\). The set of edges of G is denoted by E(G). For a connected graph G we let \(d_G(x,y)\) denote the combinatorial distance between x and y, given by the minimum length of a path from x to y, and for \(x \in V(G)\), \(r > 0\) let

$$\begin{aligned} B_G(x,r) := \{y \in V(G) : d_G(x,y) \le r\}. \end{aligned}$$

3 Spinal graphs

Definition 1

Let G be a connected graph, \(\varSigma \subset V(G)\), and let \(\pi :V(G) \rightarrow \varSigma \) be a function such that

  • \(\pi (x) = x\) for all \(x \in \varSigma \),

  • \(\pi ^{-1}(x)\) is finite for all \(x \in \varSigma \),

  • if \(a,b \in V(G)\) and \(\pi (a) \ne \pi (b)\), then every path from a to b contains a subpath from \(\pi (a)\) to \(\pi (b)\).

We refer to \((G,\varSigma ,\pi )\) as a spinal graph, and the set of vertices \(\varSigma \) is called the spine.

Remark 1

One could formulate this definition without the finiteness condition, but it will be convenient for us to keep it.

Fig. 1
figure 1

A spinal graph \((G,\varSigma ,\pi )\), with a few vertices labeled. The spine \(\varSigma \subset V(G)\) consists of the shaded vertices

An example of a spinal graph \((G,\varSigma ,\pi )\) is pictured in Fig. 1. There the vertices of the spine \(\varSigma \) are shaded black, while the other vertices are unshaded; for each vertex x, the point \(\pi (x)\) is the (uniquely determined) point on \(\varSigma \) of minimal distance to x. The dotted lines are not edges of G; if they were to be added to G, then the resulting graph would not be a spinal graph.

To help the reader familiarise themselves with the definition of a spinal graph we prove the following lemma (which will be useful later).

Lemma 1

Let \((G,\varSigma ,\pi )\) be a spinal graph, and suppose \(a,b \in V(G)\) with \(\pi (a) = \pi (b) =: x\). Then every minimal path from a to b is contained entirely in \(\pi ^{-1}(x)\).

Proof

Suppose this is false. Then there exist \(a,b \in V(G)\) with \(\pi (a) = \pi (b) =: x\) and a path \(\gamma \) from a to b of minimal length which passes through a vertex c with \(\pi (c) \ne x\). Since \(\pi (a) \ne \pi (c)\), by the third condition in the definition of a spinal graph, there exists a subpath \(\heartsuit \) of \(\gamma \) from \(\pi (a) = x\) to \(\pi ^2(c) = \pi (c)\), so we can write \(\gamma \) as a concatenation of paths \(\gamma = \alpha *\heartsuit *\beta \), where \(\alpha \) is a path from a to x and \(\beta \) is a path from \(\pi (c)\) to b; viewed as a commutative diagram,

Similarly, there is a subpath \(\heartsuit ^\prime \) of \(\beta \) from \(\pi (c)\) to \(\pi (b) = x\), and we can write \(\beta = \delta *\heartsuit ^\prime *\delta ^\prime \) as a concatenation of paths, summarised by the commutative diagram

Putting these commutative diagrams together we get

from which we can read that

$$\begin{aligned} \ell (\gamma ) = \ell (\alpha *\heartsuit *\delta *\heartsuit ^\prime *\delta ^\prime ) = \ell (\alpha ) + \ell (\heartsuit ) + \ell (\delta ) + \ell (\heartsuit ^\prime ) + \ell (\delta ^\prime ) \end{aligned}$$

where \(\ell (\cdot )\) denotes the length of a path. Since \(x \ne \pi (c)\), the paths \(\heartsuit \) and \(\heartsuit ^\prime \) have positive length, so we find that

$$\begin{aligned} \ell (\alpha * \delta ^\prime ) = \ell (\alpha ) + \ell (\delta ^\prime ) < \ell (\gamma ). \end{aligned}$$

Since \(\alpha * \delta ^\prime \) is a path from a to b, this contradicts the assumption that \(\gamma \) has minimal length. \(\square \)

Spinal graphs may be constructed by gluing a collection of finite graphs along another graph; this is made precise in the following example. In fact, this construction yields all spinal graphs (up to isomorphism, in the usual graph-theoretical sense), as will be shown in Proposition 1.

Example 1

Let \(\varGamma \) be a connected graph and let \((G_x)_{x \in V(\varGamma )}\) be a collection of finite connected graphs indexed by the vertices of \(\varGamma \). Suppose that for each \(x \in V(\varGamma )\) a distinguished vertex \(z_x \in V(G_x)\) is given. Then one can construct a graph G by gluing each \(G_x\) to \(\varGamma \) with the identification \(z_x \sim x\). More precisely we have

$$\begin{aligned} V(G) := \bigsqcup _{x \in V(\varGamma )} V(G_x) = \{(x,z) : x \in V(\varGamma ), z \in G_x\}, \end{aligned}$$
(1)

and two vertices \((x_1,y_1)\), \((x_2,y_2)\) are neighbours if and only if either

$$\begin{aligned} x_1 = x_2 = x \hbox { and } y_1 \sim y_2 \hbox { in } G_x, \end{aligned}$$

or

$$\begin{aligned} y_1 = z_{x_1} \hbox { and } y_2 = z_{x_2} \hbox { and } x_1 \sim x_2 in \varGamma . \end{aligned}$$

The graph \(\varGamma \), along with the graphs \((G_x)_{x \in V(\varGamma )}\), naturally embed into G. We set \(\varSigma := V(\varGamma )\) in this embedding; in the disjoint union representation (1) we have

$$\begin{aligned} \varSigma := \{(x,z_x) : x \in V(\varGamma )\}. \end{aligned}$$

Every vertex \(z \in V(G)\) belongs to precisely one of the embedded graphs \(G_x\) with \(x \in V(\varGamma )\), and we define \(\pi :V(G) \rightarrow \varSigma \) by the relation \(z \in G_{\pi (z)}\); in the disjoint union representation (1), we have \(\pi (x,z) = (x,z_x)\).

It is immediate that \(\pi (x) = x\) for all \(x \in \varSigma \), and that each \(\pi ^{-1}(x)\) is finite. Now suppose \(a,b \in V(G)\) with \(\pi (a) \ne \pi (b)\). By construction, every path including a that does not pass through \(\pi (a)\) must be entirely contained in \(G_{\pi (a)}\). Since \(b \notin G_{\pi (a)}\), every path from a to b must pass through \(\pi (a)\), and by symmetry such a path must also pass through \(\pi (b)\). That is, every path from a to b contains a subpath from \(\pi (a)\) to \(\pi (b)\). Therefore \((G,\varSigma ,\pi )\) is indeed a spinal graph.

Proposition 1

Suppose \((G,\varSigma ,\pi )\) is a spinal graph. Then there exists a connected graph \(\varGamma \) and a collection \((G_x)_{x \in V(\varGamma )}\) of finite connected graphs such that there exists an isomorphism between G and the graph constructed in Example 1, under which \(\varSigma \) corresponds to \(V(\varGamma )\) and \(\pi \) corresponds to the map \(\pi ^\prime \) given by the relation \(z \in G_{\pi '(z)}\).

Proof

Let \((G,\varSigma ,\pi )\) be a spinal graph. Let \(\varGamma \) be the full subgraph determined by \(\varSigma \), and for every \(x \in \varSigma = V(\varGamma )\) let \(G_x\) be the full subgraph determined by \(\pi ^{-1}(x)\). Then \(\varGamma \) is connected, each \(G_x\) is finite and connected (by Lemma 1), and we have a bijection

$$\begin{aligned} \varphi :V(G) \rightarrow \bigsqcup _{x \in V(\varGamma )} V(G_x), \quad a \mapsto (\pi (a),a). \end{aligned}$$

By the construction in Example 1, it suffices to show that \(a,b \in V(G)\) are neighbours if and only if either \(\pi (a) = \pi (b)\) and \(a \sim b\) in \(G_{\pi (a)}\), or \(a = \pi (a)\) and \(b = \pi (b)\) and \(\pi (a) \sim \pi (b)\) in \(\varGamma \).

By Lemma 1, if \(\pi (a) = \pi (b)\) then every shortest path from a to b is entirely contained in \(G_{\pi (a)}\), so in this case a and b are neighbours in G if and only if \(a \sim b\) in \(G_{\pi (a)}\). On the other hand, if \(\pi (a) \ne \pi (b)\), then every path from a to b contains a subpath from \(\pi (a)\) to \(\pi (b)\), and thus a and b are neighbours if and only if either \(a = \pi (a)\) and \(b = \pi (b)\) or \(a = \pi (b)\) and \(b = \pi (a)\). In the first case, since \(\varGamma \) is the full subgraph determined by \(\varSigma = \pi (V(G))\), we have \(\pi (a) \sim \pi (b)\) in \(\varGamma \), and we are done. The second case never occurs: since \(\pi (b) \in \varSigma \), we would have \(\pi (a) = \pi (\pi (b)) = \pi (b)\), which is a contradiction. \(\square \)

4 Dimensions of a spinal graph

For a spinal graph \((G,\varSigma ,\pi )\) we write \(d_\varSigma \) and \(B_\varSigma \) for the combinatorial distance and balls in the full subgraph determined by \(\varSigma \).

Definition 2

Let \((G,\varSigma ,\pi )\) be a spinal graph. For all \(x,y \in V(G)\) define the spinal distance [xy] by

$$\begin{aligned}{}[x,y] := d_\varSigma (\pi (x),\pi (y)), \end{aligned}$$

and for \(r > 0\) we define associated spinal sets by

$$\begin{aligned} D(x,r) := \{y \in G : [x,y] \le r\} = \pi ^{-1}(B_\varSigma (\pi (x),r)). \end{aligned}$$

The spinal distance is a pseudometric on V(G), and the quotient metric space is isometric to \((\varSigma ,d_\varSigma )\), but we will not use this fact in what follows.

Definition 3

Let \(\delta _\varSigma , \delta _G \ge 1\). We say that a spinal graph \((G,\varSigma ,\pi )\) has dimensions \((\delta _\varSigma , \delta _G)\) if there exists a point \(x_0 \in \varSigma \) and an increasing sequence \((n_k)_{k \in \mathbb {N}}\) of natural numbers such that for all \(k \in \mathbb {N}\),

$$\begin{aligned} |D(x_0,{2n_k})|\lesssim & {} |D(x_0,n_k)|, \end{aligned}$$
(2)
$$\begin{aligned} |B_\varSigma (x_0,2n_k)|\lesssim & {} n_k^{\delta _\varSigma }, \end{aligned}$$
(3)
$$\begin{aligned} |D(x_0,n_k)|\simeq & {} n_k^{\delta _G}. \end{aligned}$$
(4)

Note that the dimenions of a spinal graph need not be uniquely determined, and may vary for different choices of \(x_0\) and \((n_k)_{k \in \mathbb {N}}\).

Example 2

Let \(n \in \mathbb {N}\) and consider the Vicsek graph \(\mathcal {V}^n\) in \(\mathbb {R}^n\), the construction of which is given in [4, Proof of Theorem 4.1], [8, Chapter 5], and [9, Sec. 5]. One can consider \(\mathcal {V}^n\) as a graph with \(V(\mathcal {V}^n) \subset \mathbb {Z}^n\), defined as an increasing union of subgraphs \(\cup _{m=0}^\infty \mathcal {V}_m^n\). The subgraph \(\mathcal {V}_0^n\) consists of \(2^n + 1\) vertices: one at each corner of the unit n-cube, and a central vertex at the origin. Each corner vertex is connected to the central vertex. For \(m \ge 1\), \(\mathcal {V}_m^n\) is constructed inductively by connecting a copy of \(\mathcal {V}_{m-1}^n\) to each ‘corner’ of \(\mathcal {V}_{m-1}^n\). It follows that \(|V(\mathcal {V}_m^n)| \simeq (2^n + 1)^m\) (see [4, equation (4.10)]).

Let \(\varSigma \in V(\mathcal {V}^n)\) be the set of vertices along the \(2^n\) diagonals: with \(V(\mathcal {V}^n) \subset \mathbb {Z}^n\), we have

$$\begin{aligned} \varSigma = \{(\varepsilon _1 m,\varepsilon _2 m,\ldots ,\varepsilon _n m) \in \mathbb {Z}^n : m \in \mathbb {N}, \varepsilon _j \in \{1,-1\}\}. \end{aligned}$$

For every vertex \(x \in V(\mathcal {V}^n)\) there is a unique \(y \in \varSigma \) such that x and y are connected by a path which intersects \(\varSigma \) only at y. Setting \(\pi (x) := y\) makes \((\mathcal {V}^n,\varSigma )\) a spinal graph. Pictured in Fig. 2 are the first few steps of the construction of \(\mathcal {V}^2\), with the spine \(\varSigma \) emphasised.

Let \(o \in V(\mathcal {V})\) be the ‘center vertex’ of \(\mathcal {V}_0^n\), and for \(k \in \mathbb {N}\) let \(n_k := 3^k\). Then \(D(o,n_k) = \mathcal {V}_k^n\), and so

$$\begin{aligned} |D(o,n_k)| = (2^n + 1)^k = n_k^{\log _3(2^n + 1)}. \end{aligned}$$

We also have

$$\begin{aligned} |D(o,2n_k)| \le |D(o,n_{k+1})| = n_{k+1}^{\log _3(2^n + 1)} \simeq n_k^{\log _3(2^n + 1)} = |D(o,n_k)| \end{aligned}$$

and

$$\begin{aligned} |B_\varSigma (o, 2n_k)| = 2^n(2n_k) - 1 < 2^{n+1} 3^k \simeq n_k, \end{aligned}$$

which tells us that the spinal graph \((\mathcal {V}^n,\varSigma )\) has dimensions \((1,\log _3(2^n + 1))\). In addition, \(\mathcal {V}^n\) has polynomial volume growth of dimension \(\log _3(2^n + 1)\), that is

$$\begin{aligned} |B_\mathcal {V}(x,r)| \simeq r^{\log _3(2^n + 1)} \end{aligned}$$

for all \(x \in V(\mathcal {V}^n)\) and \(r \in \mathbb {N}\). (see [4, page 632]). The lower estimate will allow us to apply Corollary 2 to \(\mathcal {V}^n\).

Fig. 2
figure 2

The first three steps of the construction of the Vicsek graph \(\mathcal {V}^2\), with spine

In Section 6 we construct spinal graphs with global volume lower bounds and dimensions (1, D) with \(D > 1\) that do not arise from fractals.

5 Nash-type inequalities and spinal dimensional consequences

Now we assume that G is locally finite. For each vertex \(x \in V(G)\) let \(m(x) < \infty \) denote the number of neighbours of x, and for each \(f :V(G) \rightarrow \mathbb {C}\) define the length of the gradient \(|\nabla f(x)|\) by

$$\begin{aligned} |\nabla f(x)| := \left( \frac{1}{2} \sum _{\begin{array}{c} y \in V(G) \\ y \sim x \end{array}} \frac{1}{m(x)} |f(y) - f(x)|^2 \right) ^{1/2}. \end{aligned}$$

For \(1 < p \le \infty \) and \(\beta > 0\), we consider the Nash-type inequality

$$\begin{aligned} S(p,\beta ): \qquad \Vert f \Vert _p^{1 + \frac{p^\prime }{\beta }} \lesssim \Vert f \Vert _1^{\frac{p^\prime }{\beta }} \Vert |\nabla f| \Vert _p \qquad (f :V(G) \rightarrow \mathbb {C} \hbox { finitely supported}), \end{aligned}$$

which G may or may not satisfy.

In the presence of a spinal structure, the inequality \(S(p,\beta )\) gives quantitative information connecting the ‘spinal volume growth’ of G with the volume growth of \(\varSigma \). This is shown by constructing test functions, defined in terms of the spinal distance, which are constant on the fibres \(\pi ^{-1}(x)\). The gradients of these test functions are supported on the spine \(\varSigma \), while the functions themselves are supported on spinal sets.

Lemma 2

Let \((G,\varSigma ,\pi )\) be a spinal graph. Fix \(p \in (1,\infty )\) and suppose that G satisfies \(S(p,\beta )\). Then for every \(x_0 \in \varSigma \) and \(n \in \mathbb {N}\) we have

$$\begin{aligned} |D(x_0,n)|^{\frac{1}{p} \left( 1 + \frac{p^\prime }{\beta } \right) } \lesssim n^{-1} |D(x_0,2n)|^{\frac{p^\prime }{\beta }} |B_\varSigma (x_0,2n)|^{\frac{1}{p}}. \end{aligned}$$
(5)

Proof

For each \(x_0 \in \varSigma \) and \(n \in \mathbb {N}\) define \(g_n :V(G) \rightarrow [0,1]\) by

$$\begin{aligned} g_{n}(x) := \frac{\max (0,n-[x,x_0])}{n}. \end{aligned}$$

Note that \(g_{n}(x) = 0\) if and only if \([x,x_0] \ge n\), so that \({\text {supp}}g_{n} = D(x_0,n-1)\). Furthermore note that \(g_{n}\) is constant on each \(\pi ^{-1}(x)\).

Since \(|g_{2n}| \le 1\) and \({\text {supp}}g_{2n} \subset D(x_0,2n)\) we have

$$\begin{aligned} \Vert g_{2n} \Vert _1 \lesssim |D(x_0,2n)|. \end{aligned}$$
(6)

Next, since \(g_{2n}(x) \ge 1/2\) for \(x \in D(x_0,n)\), we have

$$\begin{aligned} \Vert g_{2n} \Vert _p \ge \left( \sum _{x \in D(x_0,n)} 2^{-p} \right) ^{1/p} \simeq |D(x_0,n)|^{1/p}. \end{aligned}$$
(7)

Finally, note that \(|\nabla g_{2n}(x)|= 0\) whenever \(x \in G {\setminus } \varSigma \) (since \(g_{2n}\) is constant on each connected component of \(G \setminus \varSigma \)) or \(x \in G \setminus D(x_0,2n)\) (since \({\text {supp}}g_{2n} = D(x_0,2n-1)\)). When \(x \in \varSigma \cap D(x_0,2n)\) and \(y \sim x\), we have

$$\begin{aligned} g_{2n}(x) - g_{2n}(y) = \left\{ \begin{array}{ll} (2n)^{-1} &{} \text {if } y \in \varSigma \cap D(x_0,2n) \\ 0 &{} \text {otherwise.} \end{array} \right. \end{aligned}$$

Therefore

$$\begin{aligned} \Vert |\nabla g_{2n}| \Vert _p&= \left( \sum _{x \in \varSigma \cap D(x_0,2n)} \left( \frac{1}{2} \sum _{y \sim x, y \in \varSigma \cap D(x_0,2n)} \frac{1}{m(x)} (2n)^{-2} \right) ^{p/2} \right) ^{1/p} \\&\lesssim n^{-1} |\varSigma \cap D(x_0,2n)|^{1/p} = n^{-1} |B_\varSigma (x_0,2n)|^{1/p} \end{aligned}$$

Therefore, applying \(S(p,\beta )\) to \(g_{2n}\), we get (5). \(\square \)

The previous lemma can be used to show that the Nash-type inequalities \(S(p,\beta )\) restrict the possible dimensions of a spinal graph.

Lemma 3

Let \((G,\varSigma ,\pi )\) be a spinal graph with dimensions \((\delta _\varSigma , \delta _G)\). Fix \(p > 1\) and \(\beta > 0\), and suppose G satisfies \(S(p,\beta )\). Then

$$\begin{aligned} \frac{\delta _G - \delta _\varSigma }{p} - \frac{\delta _G}{\beta } \le -1. \end{aligned}$$
(8)

Proof

Fix \(x_0 \in \varSigma \) and a sequence \((n_k)_{k \in \mathbb {N}}\) as in Definition 3. From Lemma 2 and assumptions (3) and (2), for all \(k \in \mathbb {N}\) we have

$$\begin{aligned} |D(x_0,n_k)|^{\frac{1}{p} \left( 1 + \frac{p^\prime }{\beta }\right) }&\lesssim n_k^{-1} |D(x_0,2n_k)|^{\frac{p^\prime }{\beta }} |B_\varSigma (x_0,2n_k)|^{\frac{1}{p}} \\&\lesssim n_k^{-1 + \frac{\delta _\varSigma }{p}} |D(x_0,n_k)|^{\frac{p^\prime }{\beta }}. \end{aligned}$$

Rearranging yields

$$\begin{aligned} |D(x_0,n_k)|^{\frac{1}{p} - \frac{1}{\beta }} \lesssim n_k^{-1 + \frac{\delta _\varSigma }{p}}, \end{aligned}$$

and then

$$\begin{aligned} n_k^{\delta _G\left( \frac{1}{p} - \frac{1}{\beta }\right) } \lesssim n_k^{-1 + \frac{\delta _\varSigma }{p}}. \end{aligned}$$

follows by (4). Since \(n_k\) is increasing, taking the limit as \(k \rightarrow \infty \) tells us that

$$\begin{aligned} \delta _G\left( \frac{1}{p} - \frac{1}{\beta }\right) \le -1 + \frac{\delta _\varSigma }{p}, \end{aligned}$$

which is equivalent to (8).\(\square \)

Corollary 1

Suppose the conditions of Lemma 3 are satisfied, with \(\delta _G > \delta _\varSigma \). Then \(\delta _G > \beta \) and

$$\begin{aligned} p \ge \beta \frac{\delta _G - \delta _\varSigma }{\delta _G - \beta }. \end{aligned}$$
(9)

Proof

Rearranging (8) gives

$$\begin{aligned} \delta _G \ge \beta \left( 1 + \frac{\delta _G - \delta _\varSigma }{p}\right) . \end{aligned}$$

Since \(\delta _G - \delta _\varSigma > 0\), we get \(\delta _G > \beta \). We can then rearrange further to get (9). \(\square \)

6 Riesz transform unboundedness for thickened spinal graphs

Definition 4

Let G be a uniformly locally finite graph (i.e. \(\sup _{x \in V(G)} m(x) < \infty \)) and \(n \in \mathbb {N}\). Then an n-dimensional thickening of G is a smooth Riemannian manifold M constructed by replacing each vertex \(x \in V(G)\) by an n-sphere, each edge \(e \in E(G)\) by an n-cylinder, and welding the cylinders smoothly to the balls according to the graph structure of G, in such a way that M has bounded geometry (i.e. M has positive injectivity radius, and Ricci curvature bounded from below).

More precisely: define

$$\begin{aligned} \widetilde{M} := \bigsqcup _{x \in V(G)} B_x \sqcup \bigsqcup _{e \in E(G)} C_e, \end{aligned}$$

where \(B_x\) is isometric to a round n-sphere \(S^n\) with m(x) disjoint open balls of fixed small radius \(\varepsilon \) removed, and where each \(C_e\) is isometric to a cylinder \(S_\varepsilon ^{n-1} \times [0,1]\), with \(S_\varepsilon ^{n-1} = \partial B(0,\varepsilon ) \subset \mathbb {R}^n\). A \(C^0\) Riemannian manifold \(M^\prime \) is constructed as a quotient of \(\widetilde{M}\) by gluing a cylinder \(C_e\) to two spheres \(B_x\) and \(B_y\) if and only if \(x \sim y\) in G (in such a way that every ‘hole’ in \(B_x\) has a cylinder attached to it). A thickening M with bounded geometry may then be defined by smoothing the interface between spheres and the cylinders in \(M^\prime \) arbitrarily (but uniformly among all the interfaces).

Remark 2

In what follows, we may replace a thickening of G (in the sense above) by any Riemannian manifold M of bounded geometry that is isometric to G at infinity in the sense of Coulhon–Saloff-Coste [12] (following Kanai [13]); our discretisation/thickening procedures only depend on results in [12].

The following proposition can be proven by directly following the proof of [11, Proposition 6.2] (see also the first part of [9, Theorem 5.1]). The proof involves the discretisation results of [12, Sec. 6].

Proposition 2

Let G be a locally uniformly finite graph, and let M be a thickening of G of any dimension. Fix \(p \in (1,\infty )\) and suppose that M satisfies \((RR_p)\). Furthermore, suppose that the heat kernel h of M satisfies

$$\begin{aligned} h_t(x,x) \lesssim t^{-\alpha /2} \qquad \textit{for all}\ t > 1, x \in M. \end{aligned}$$

Then G satisfies \(S(p,\alpha )\).

Since \(S(p,\alpha )\) restricts the possible dimensions of a spinal graph, we may argue by contraposition to show that dimension and volume information on a spinal graph implies unboundedness of the Riesz transform on \(L^p(M)\) for sufficiently large \(p > 2\) (in fact, we prove that \((RR_{p'})\) does not hold for sufficiently large \(p > 2\), which is strictly stronger).

Theorem 1

Let \((G,\varSigma ,\pi )\) be a locally uniformly finite spinal graph with dimensions \((\delta _\varSigma , \delta _G)\), with \(\delta _G > \delta _\varSigma \). Furthermore, suppose that \(B_G(x,r) > rsim r^\nu \) for all \(r \ge 1\). Let M be a thickening of G of any dimension. Then for all

$$\begin{aligned} p > 2\frac{\delta _G-\delta _\varSigma }{\frac{\delta _G}{\nu ^\prime } - 2\delta _\varSigma + 2} =: p_c(\delta _\varSigma , \delta _G, \nu ), \end{aligned}$$

M does not satisfy \((R_p)\).

Proof

The volume assumption on G implies a corresponding large-scale volume estimate

$$\begin{aligned} V(x,r) > rsim r^{\nu } \qquad (\hbox {for all } r > 1, x \in M) \end{aligned}$$

on M (this may be derived from the results of [12, Sec. 6]). Since M has bounded geometry, [4, Theorem 1.1] implies the heat kernel estimate

$$\begin{aligned} h_t(x,x) \lesssim t^{-\nu /(\nu + 1)} \qquad \hbox {for all } t > 1 \end{aligned}$$

on M. Fix \(q > 1\) and suppose that M satisfies \((R_q)\), hence also \((RR_{q^\prime })\). Proposition 2 then implies that G satisfies \(S(q^\prime , 2\nu /(\nu + 1))\), and Corollary 1 then yields

$$\begin{aligned} q^\prime \ge \frac{2\nu }{\nu +1} \left( \frac{\delta _G - \delta _\varSigma }{\delta _G - \frac{2\nu }{\nu +1}} \right) = p_c^\prime \end{aligned}$$

or equivalently that \(q \le p_c\). Therefore M does not satisfy \((R_p)\) for any \(p > p_c\). \(\square \)

Taking \(\delta _\varSigma = 1\) and \(\nu = \delta _G\) gives the following corollary.

Corollary 2

Let \((G,\varSigma ,\pi )\) be a locally uniformly finite spinal graph with dimensions \((1,\delta _G)\), with \(\delta _G > 1\), and suppose that \(B_G(x,r) > rsim r^{\delta _G}\) for all \(r \in \mathbb {N}\). Let M be a thickening of G of any dimension. Then the Riesz transform bound \((R_p)\) for M fails for all \(p > 2\).

As remarked in Example 2, these assumptions are satisfied by the Vicsek graphs \(\mathcal {V}^n\), reproving [9, Theorem 5.1].

7 Non-fractal spinal graphs with volume lower bounds

Fix \(D > 1\). In this section we show how to construct locally uniformly finite spinal graphs \((G,\varSigma ,\pi )\) with dimensions (1, D) and the volume lower bound

$$\begin{aligned} |B_G(x,r)| > rsim r^D \qquad (x \in V(G), r \in \mathbb {N}), \end{aligned}$$
(10)

but which need not possess any ‘fractal’ structure (in contrast with the Vicsek graph example). Corollary 2 applies to such spinal graphs, thus yielding many manifolds M for which \((R_p)\) fails for all \(p > 2\).

First we need a technical lemma on volumes of intersections of balls in doubling graphs. We defer the proof to Sec. 7. Recall that a graph G is doubling if there exist constants \(C_d,\nu > 0\) such that for all \(0< r< R < \infty \) and \(x \in V(G)\),

$$\begin{aligned} |B_G(x,R)| \le C_d(R/r)^\nu |B_G(x,r)|. \end{aligned}$$

Taking \(R = 1\) and \(r = 1-\varepsilon \) for \(\varepsilon \) arbitrarily small shows that a doubling graph is locally uniformly finite, with \(m(x) \le C_d\) for all \(x \in V(G)\).

Lemma 4

Let G be a doubling graph. Then there exists \(C > 0\), depending only on the doubling constants of G, such that for all \(y \in V(G)\) and \(R > 0\), and for all \(x \in B(y,R)\) and \(r \le 2R\), we have

$$\begin{aligned} |B_G(x,r) \cap B_G(y,R)| \ge C|B_G(x,r)|. \end{aligned}$$

Now we move on to our construction. This is inspired by the ‘plate’ construction in [4, Theorem 5.1].

Example 3

Fix \(\delta > D\), and let \((P_n)_{n \in \mathbb {N}}\) be a family of graphs satisfying

$$\begin{aligned} |B_{P_n}(x,r)| \simeq r^\delta \qquad (x \in V(P_n), r \in \mathbb {N}) \end{aligned}$$

with implicit constants independent of n. For simplicity one can take each \(P_n\) to be equal to a fixed graph P; one could even take \(\delta \in \mathbb {N}\) and \(P = \mathbb {Z}^\delta \). We allow for arbitrary choices to emphasise that self-similarity is not necessary. Let \(\alpha = (D-1)/\delta \) (so that \(\alpha \delta + 1 = D\) and \(\alpha < 1\)) and for each \(n \in \mathbb {N}\) choose an arbitrary vertex \(o_n \in V(P_n)\). Construct a spinal graph \((G,\varSigma ,\pi )\) with \(\varSigma = \mathbb {N}\) as in Example 1 by taking \(G_n\) to be the full subgraph of \(P_n\) determined by \(B_{P_n}(o_n,n^\alpha )\), and choosing as distinguished points \(z_n = o_n\).

To show that this spinal graph has dimensions (1, D), take the sequence \(n_k = k\) and observe that

$$\begin{aligned} |D(1,k)| = \sum _{n=1}^k |B_{P_n}(o_n,n^\alpha )| \simeq \sum _{n=1}^k n^{\alpha \delta } \simeq k^{\alpha \delta + 1} = k^D \end{aligned}$$

(the second sum may be estimated by comparing with integrals of the function \(t \mapsto t^{\alpha \delta }\)). In particular we have \(|D(1,2k)| \simeq (2k)^D \simeq |D(1,k)|\), and furthermore it is clear that \(|B_\mathbb {N}(x,r)| \simeq r\). Therefore the spinal graph has dimensions (1, D).

It is more difficult to show the global volume lower bound (10), but luckily the proof of [4, Theorem 5.1] already does this for a similar problem. Note that it suffices to assume \(r \ge 2\).

First we show that \(|B_G(n,r)| > rsim r^D\) for all \(n \in \mathbb {N}\). To see this, write

$$\begin{aligned} |B_G(n,r)|&\ge \sum _{k=0}^{r} |B_{P_{n+k}}(o_{n+k},\min (r-k,(n+k)^\alpha ))| \\&\ge \sum _{k=0}^{\lfloor r/2 \rfloor } |B_{P_{n+k}}(o_{n+k},\min (r-k,k^\alpha ))| \\&\simeq \sum _{k=0}^{\lfloor r/2 \rfloor } k^{\alpha \delta } \simeq \lfloor r/2 \rfloor ^{\alpha \delta + 1} \simeq r^D, \end{aligned}$$

using that \(k^\alpha < r-k\) for \(k \le r/2\) in the third line.

Now suppose \(x \in V(G)\) with \(\pi (x) = n\). After identifying x with the appropriate vertex \(z \in B_{P_n}(o_n, n^\alpha )\) (which, recall, is identified with \(\pi (x)\)), we get an identification of \(B_G(x,r) \cap \pi ^{-1}(n)\) with \(B_{P_n}(z,r) \cap B_{P_n}(o_n, n^\alpha )\). Thus for \(r \le 2n^\alpha \) we have

$$\begin{aligned} |B_G(x,r)|&\ge |B_G(x,r) \cap \pi ^{-1}(n)| \\&= |B_{P_n}(z,r) \cap B_{P_n}(o_n,n^\alpha )| \\& > rsim |B_{P_n}(z,r)| \simeq r^\delta > r^D \end{aligned}$$

using Lemma 4 in the third line. On the other hand, if \(r > 2n^\alpha \), then \(B_G(x,r)\) contains both \(\pi ^{-1}(n)\) and \(B_G(n+1,r-1-n^\alpha )\), so

$$\begin{aligned} |B_G(x,r)|&\ge \max (|\pi ^{-1}(n)|, |B_G(n+1,r-1-n^\alpha )|) \\& > rsim |\pi ^{-1}(n)| + |B_G(n+1, r-1-n^\alpha )| \\& > rsim n^{\alpha D} + (r-1-n^\alpha )^D \\&\simeq (n^\alpha + r - 1 - n^\alpha )^D \\&\simeq r^D. \end{aligned}$$

This completes the proof of (10).

The following corollary is then an immediate consequence of Corollary 2.

Corollary 3

Suppose M is a thickening of a spinal graph \((G,\varSigma ,\pi )\) as constructed as in Example 3. Then the Riesz transform bound \((R_p)\) for M fails for all \(p > 2\).

Remark 3

It is probably possible to construct spinal graphs with dimensions \((\delta _\varSigma , \delta _G)\) with \(1< \delta _\varSigma < \delta _G\) and with a polynomial volume lower bound of dimension \(\delta _G\), thus yielding manifolds M for which \((R_p)\) fails for all \(p> p_c > 2\). Since our construction exploits taking \(\varSigma = \mathbb {N}\), this is beyond the scope of this article. It may even be possible to show that \((R_p)\) holds on such manifolds for \(p \in (2,p_c)\), but this is very much beyond the scope of this article.

8 Proof of Lemma 4

Here we prove the technical lemma needed in the construction of the previous section. We write \(B(x,r) := B_G(x,r)\) and \(d(x,y) := d_G(x,y)\).

Proof

First note that if the result is true for \(r \le R\), then it holds for \(r \le 2R\), because in this case for \(r > R\) we have

$$\begin{aligned} |B(x,r) \cap B(y,R)| \ge |B(x,R) \cap B(y,R)|&\ge C|B(x,R)|\\&\ge CC_d^{-1}(R/r)^\nu |B(x,r)| \\&\ge CC_d^{-1}2^{-\nu } |B(x,r)| \end{aligned}$$

using the doubling condition and \(R \ge r/2\) in the last step. Thus we assume that \(r \le R\), and split the proof into two cases.

Case 1: \(r > 2d(x,y)\). By definition of the combinatorial distance, there exists a vertex z such that \(d(y,z) + d(z,x) = d(y,x)\) and \(d(y,z) = \lceil d(y,x)/2 \rceil \). Suppose that \(z^\prime \in B(z,r-d(y,z))\). Then

$$\begin{aligned} d(z^\prime ,x)&\le r - d(y,z) + d(z,x) \\&= r - 2\lceil d(y,x)/2 \rceil + d(y,x) \\&\le r - 2d(y,x) \le r \end{aligned}$$

and

$$\begin{aligned} d(z^\prime ,y) \le r - d(y,z) + d(z,y) = r \le R, \end{aligned}$$

so \(B(z,r-d(y,z)) \subset B(x,r) \cap B(y,R)\). Therefore

$$\begin{aligned} |B(x,r) \cap B(y,R)|&\ge |B(z,r-d(y,z))| \\& > rsim \left( \frac{r-d(y,z)}{r+d(x,z)} \right) ^\nu |B(z,r+d(x,z))| \\&\ge \left( \frac{r-d(y,z)}{r+d(x,z)} \right) ^\nu |B(x,r)| \end{aligned}$$

for some \(\nu > 0\) determined by the doubling constant of G. To see that the bracketed expression above is uniformly bounded below, estimate its reciprocal from above by

$$\begin{aligned} \frac{r+d(x,z)}{r-d(y,z)}&= \frac{r-d(y,z)+d(x,y)}{r-d(y,z)} \\&= 1 + \frac{d(x,y)}{r - d(y,z)} \\&< 1 + \frac{d(x,y)}{2d(x,y) - d(y,z)} \\&= 1 + \frac{d(x,y)}{d(x,y) + d(z,x)} \\&\le 2. \end{aligned}$$

using \(r > 2d(x,y)\) in the third line and \(d(z,x) = d(x,y) - d(y,z)\) in the fourth line.

Case 2: \(r \le 2d(x,y)\). Note that the estimate for \(r < 2\) follows from the fact that G is locally uniformly finite, so it suffices to consider \(r \ge 2\). As in the previous case, there exists a vertex z such that \(d(y,z) + d(z,x) = d(y,x)\) and \(d(x,z) = \lfloor r/2 \rfloor \) (here we use that \(r/2 \le d(x,y)\)). If \(z^\prime \in B(z,\lfloor r/2 \rfloor )\), then

$$\begin{aligned} d(z^\prime ,x) < 2\left\lfloor \frac{r}{2} \right\rfloor \le r \end{aligned}$$

and

$$\begin{aligned} d(z^\prime ,y) < \left\lfloor \frac{r}{2} \right\rfloor + d(z,y) = \left\lfloor \frac{r}{2}\right\rfloor + d(x,y) - \left\lfloor \frac{r}{2} \right\rfloor = d(x,y) \le R, \end{aligned}$$

using that \(x \in B(y,R)\) by assumption, and so \(B(z,\lfloor r/2 \rfloor ) \subset B(x,r) \cap B(y,R)\). Therefore

$$\begin{aligned} |B(x,r) \cap B(y,R)|&\ge |B(z,\lfloor r/2 \rfloor )| \\&\ge |B(z,r/4)| \\&\ge (8^\nu C_d)^{-1} |B(z,2r)| \\&\ge (8^\nu C_d)^{-1} |B(z,r+d(z,x))| \\&\ge (8^\nu C_d)^{-1} |B(x,r)| \end{aligned}$$

since \(\lfloor r/2 \rfloor \ge r/4\) holds whenever \(r \ge 2\), and using the doubling condition. \(\square \)