1 Introduction

1.1 Background

In addition to its rich stochastic geometric structure, first-passage percolation on \(\mathbb {Z}^d\) provides a model for the study of fluctuations of a non-linear function of a large number of independent random variables. For recent surveys, see [5, 15, 17].

In this paper, we are concerned with the variance of the passage time \(\tau (0,x)\) from \(0\) to \(x\in \mathbb {Z}^d\). The passage time is the random variable defined as

$$\begin{aligned} \tau (0,x) =\inf _{\gamma : 0\rightarrow x}\sum _{e\in \gamma } t_e, \end{aligned}$$
(1.1)

where the infimum is taken over all lattices paths \(\gamma =(v_0=0,e_0,v_1,\ldots ,e_N,v_N=x)\) joining \(0\) to \(x\). The collection \((t_e)_{e\in \mathcal {E}^d}\) consists of nonnegative independent random variables with common distribution \(\mu \) and \(\mathcal {E}^d\) is the set of nearest-neighbor edges.

When \(d=1\), (1.1) is simply a sum over i.i.d. random variables for each \(x\), and the variance of \(\tau (0,x)\) is of order \(\Vert x\Vert _1\). In contrast, when \(d\ge 2\), (1.1) is a minimum over a collection of correlated sums of i.i.d. random variables. This correlation structure has led physicists to conjecture a sublinear scaling of the form \(\Vert x\Vert _1^\alpha \), \(\alpha <1\) for the fluctuations. In the case \(d=2\), the model is expected to have KPZ scaling [19], with \(\alpha =\frac{2}{3}\), and the recentered passage time approximately follows the Tracy–Widom distribution. Except for Johansson’s work [18] on a related exactly solvable model, there has been little success in rigorously confirming these predictions.

In [21], Kesten showed that the variance of \(\tau (0,x)\) is at most linear in the distance of \(x\) to the origin:

$$\begin{aligned} {{\mathrm{Var}}}\,\tau (0,x)\le C\Vert x\Vert _1, \end{aligned}$$

for some constant \(C\). Kesten also showed that if \(\mu \) has exponential moments:

$$\begin{aligned} \int e^{\delta x}\,\mu (\mathrm {d}x)<\infty \quad \text { for some } \delta >0, \end{aligned}$$
(1.2)

then the passage time is exponentially concentrated around its mean:

$$\begin{aligned} \mathbb {P}(|\tau (0,x)-\mathbb {E}\tau (0,x)|\ge \uplambda \sqrt{\Vert x\Vert _1})\le Ce^{-c\uplambda }, \end{aligned}$$
(1.3)

for \(\uplambda \le C\Vert x\Vert _1\). Talagrand improved this result to Gausssian concentration on the scale \(\sqrt{\Vert x\Vert _1}\): see [31, Proposition 8.3]. These results have been used to derive concentration of the mean of the passage time around the “time constant.” Some relevant papers include [1, 27, 33]. In the other direction, lower bounds have been given for the variance of the passage time, but the strongest results are dimension-dependent; see [3, 21, 25, 34].

In a remarkable paper [4], Benjamini et al. used an inequality due to Talagrand [32] to prove that if the edge-weight distribution is uniform on a set of two positive values, the variance is sublinear in the distance:

$$\begin{aligned} {{\mathrm{Var}}}\, \tau (0,x) \le C(a,b)\frac{\Vert x\Vert _1}{\log \Vert x\Vert _1},~\quad d \ge 2 \end{aligned}$$

for \(0<a<b\) and \(\mathbb {P}(t_e=a)=\mathbb {P}(t_e = b) =\frac{1}{2}.\) Benaim and Rossignol [6] introduced their “modified Poincaré inequality,” itself based on an inequality of Falik and Samorodnitsky (a corresponding inequality appears in Rossignol [28, Equations (11)–(14)]), to extend the variance estimate to a class of continuous distributions which they termed “nearly gamma.” Nearly gamma distributions satisfy an entropy bound analogous to the logarithmic Sobolev inequality for the gamma distribution, which explains their name; for a nearly gamma \(\mu \) and, for simplicity, \(f\) smooth,

$$\begin{aligned} Ent_\mu f^2 :{=} \int f^2(x) \log \frac{f^2(x)}{\mathbb {E}_\mu f^2}\,\mu (\mathrm {d}x) \le C\int \left( \sqrt{x}f'(x)\right) ^2\,\mu (\mathrm {d}x). \end{aligned}$$
(1.4)

Benaim and Rossignol also show exponential concentration at scale \(\sqrt{\Vert x\Vert _1/\log \Vert x\Vert _1}\) for nearly gamma distributions with exponential moments: if \(\mu \) satisfies (1.4) and (1.2), then

$$\begin{aligned} \mathbb {P}_\mu (|\tau (0,x)-\mathbb {E}_\mu \tau (0,x)| \ge \uplambda \sqrt{\Vert x\Vert _1/\log \Vert x\Vert _1}) \le Ce^{-c\uplambda }. \end{aligned}$$
(1.5)

The nearly gamma condition excludes many natural distributions, including all power law distributions and distributions with infinite support which decay too quickly, mixtures of continuous and discrete distributions, singular continuous distributions, and continuous distributions with disconnected support, or whose density has zeros on its support.

1.2 Main result

The purpose of the present work is to extend the sublinear variance results mentioned above to general distributions with \(2+\log \) moments. We make two assumptions:

$$\begin{aligned}&\int x^2(\log x)_+~\mu (\text {d}x) < \infty , \end{aligned}$$
(1.6)
$$\begin{aligned}&\mu (\{0\})<p_c(d) , \end{aligned}$$
(1.7)

where \(p_c(d)\) is the critical parameter for bond percolation on \(\mathbb {Z}^d\).

Our main result is the following:

Theorem 1.1

Let \(\mu \) be a Borel probability measure supported on \([0,\infty )\) satisfying (1.6) and (1.7). In i.i.d. first-passage percolation on \((\mathbb {Z}^d,\mathcal {E}^d)\), \(d\ge 2\), with edge-weight distribution \(\mu \), there exists a constant \(C=C(\mu ,d)\) such that

$$\begin{aligned} {{\mathrm{Var}}}\, \tau (0,x) \le C\frac{\Vert x\Vert _1}{\log \Vert x\Vert _1}\quad \textit{for all}\,\, x \in \mathbb {Z}^d. \end{aligned}$$

Remark 1.2

When (1.7) fails, the passage time is known to be bounded by \(C\Vert x\Vert _1^\epsilon \) for any \(\epsilon \). See [9, 35] for more details. \(\square \)

Remark 1.3

The moment condition \(\mathbb {E}t_e^2(\log t_e)_+<\infty \) may be able to be weakened, perhaps as low as \(\mathbb {E}t_e^{(2/d)+a}<\infty \) for some \(a>0\) by tensorizing entropy over small blocks, as in [12, Lemma 2.6]. The main reason is that, due to [10, Lemma 3.1], \({{\mathrm{Var}}}\,\tau (x,y) < \infty \) for all \(x,y\) under the condition \(\mathbb {E} t_e^{(1/d)+a} < \infty \) for some \(a>0\). \(\square \)

Our method of proof may be of independent interest. Following [6], we use a martingale difference decomposition and the inequality of Falik and Samorodnitsky to control the variance of an averaged version of \(\tau (0,x)\) by the entropy times a \(1/\log \Vert x\Vert _1\) factor. Instead of representing the measure \(\mu \) as the pushfoward of a Gaussian by an invertible transformation and using the Gaussian logarithmic Sobolev inequality, we represent \(\mu \) as the image of an infinite sequence of uniform Bernoulli variables, and use Bonami and Gross’s [7, 16] two-point entropy inequality (the “discrete log-Sobolev inequality”) to control the entropy. A central part of the argument is then to estimate the discrete derivatives of \(\tau (0,x)\) with respect to variations of the Bernoulli variables.

1.3 Outline of the paper

The plan of the paper is as follows: in Sect. 2, we review some basic properties of the entropy functional with respect to a probability measure, and present the inequality of Falik and Samorodnitsky which we will use. In Sect. 3, we apply this inequality to first-passage percolation, using the martingale decomposition introduced in [6]. We then briefly explain Benaim and Rossignol’s approach based on the Gaussian log-Sobolev inequality (LSI) in Sect. 4, and show that a modification of their method using positive association already allows one to deal with a larger class of continuous distributions than the ones handled in [6]. The purpose of Sect. 4 is only to clarify the role of conditions appearing in [6]. This section is independent of the derivation of our main result.

In Sect. 5, we provide a lower bound for the quantity \(\sum _{k=1}^\infty (\mathbb {E} |V_k|)^2\) appearing in the variance bound, which will give the logarithmic factor in the final inequality. Next, in Sect. 6 we represent the passage time variables through a Bernoulli encoding and, after applying Bonami’s inequality, bound a sum of discrete derivatives with the help of estimates on greedy lattice animals.

1.4 Notation and preliminary results

We will work on the space \(\varOmega = [0,\infty )^{\mathcal {E}^d}\) and let \(\mu \) be a Borel probability measure on \([0,\infty )\). The product measure \(\prod _{e \in \mathcal {E}^d} \mu \) will be denoted by \(\mathbb {P}\). A realization of passage times (edge-weights) \(\omega \in \varOmega \) will be written as \(\omega = (t_e)\) with point-to-point passage time \(\tau (x,y)\) given by (1.1). Throughout the paper, the letter \(I\) will refer to the infimum of the support of \(\mu \): writing

$$\begin{aligned} F(x) = \mu ((-\infty ,x]) \end{aligned}$$
(1.8)

for the distribution function of \(\mu \), set

$$\begin{aligned} I = \inf \{x : F(x) > 0\} . \end{aligned}$$
(1.9)

A fundamental object in first-passage percolation is a geodesic, and we spend some time here giving some basic properties of geodesics. Any path \(\gamma \) from \(x\) to \(y\) with passage time \(\tau (\gamma ) = \sum _{e \in \gamma } t_e\) satisfying \(\tau (\gamma ) = \tau (x,y)\) will be called a geodesic from \(x\) to \(y\). From the shape theorem of Cox–Durrett [10] and the fact that under (1.7), the limiting shape for the model is bounded [20, Theorem 6.1], assumptions (1.6) and (1.7) ensure the existence of geodesics:

$$\begin{aligned} \mathbb {P}( \text {for all }x,y \in \mathbb {Z}^d \text { there exists a geodesic from }x \text { to } y) = 1. \end{aligned}$$
(1.10)

There is almost surely a unique geodesic between \(x\) and \(y\) if and only if \(\mu \) is continuous, so this need not be true in general. For any \(x,y \in \mathbb {Z}^d\) we then use the notation

$$\begin{aligned} Geo(x,y) = \{e \in \mathcal {E}^d : e \in \gamma \text { for all geodesics } \gamma \text { from } x \text { to } y\}. \end{aligned}$$
(1.11)

Central to the current proofs of variance bounds for the passage time are estimates on the length of geodesics. The key theorem is due to Kesten [21, (2.25)] and is listed below. We will need to derive two generalizations of this result. The first is Lemma 5.1 and concerns the number of intersections of \(Geo(0,x)\) with arbitrary edge sets. The second, Theorem 6.6, gives a bound on the number of edges of \(Geo(0,x)\) whose weight lies in a given Borel set.

Theorem 1.4

(Kesten) Assume \(\mathbb {E}t_e<\infty \) and (1.7). There exists \(\mathbf {C}_1\) such that for all \(x\),

$$\begin{aligned} \mathbb {E}\#Geo(0,x) \le \mathbf {C}_1\Vert x\Vert _1. \end{aligned}$$

The second tool we shall need is [20, Propsition 5.8] and shows that under assumption (1.7), it is unlikely that long paths have small passage time.

Theorem 1.5

(Kesten) Assuming (1.7), there exist constants \(a, \mathbf {C}_2>0\) such that for all \(n \in \mathbb {N}\),

$$\begin{aligned} \mathbb {P}\bigg (\exists \,\, { self}\text {-}{ avoiding} \,\, \gamma \,\,{ starting} \,\,{ at}\,\, 0 \,\,{ with} \,\,\#\gamma \!\ge \! n \,\,{ but}\,\, { with}\,\, \tau (\gamma ) < { an} \bigg ) \!\le \! \exp (-\mathbf {C}_2 n). \end{aligned}$$

1.5 Proof sketch

The setup Our argument begins with the setup of Benaim and Rossignol: to bound the variance, we use the inequality of Falik–Samorodnitsky. That is, if \(T = \tau (0,x)\) is the passage time, then we enumerate the edges of the lattice as \(\{e_1, e_2, \ldots \}\) and perform a martingale decomposition

$$\begin{aligned} T - \mathbb {E}T = \sum _{k=1}^\infty V_k , \end{aligned}$$

where \(V_k = \mathbb {E}[T \mid \mathcal {F}_k] - \mathbb {E}[T \mid \mathcal {F}_{k-1}]\) and \(\mathcal {F}_k\) is the sigma-algebra generated by the edge weights \(t_{e_1}, \ldots , t_{e_k}\). Then one has

$$\begin{aligned} {{\mathrm{Var}}}\,T~\log \left[ \frac{{{\mathrm{Var}}}\,T}{\sum _{k=1}^\infty (\mathbb {E}|V_k|)^2} \right] \le \sum _{k=1}^\infty Ent(V_k^2) . \end{aligned}$$

(See Lemma 3.3.) If \({{\mathrm{Var}}}\,T \le \Vert x\Vert ^{7/8}\), then the required bound holds; otherwise, one has \({{\mathrm{Var}}}\,T \ge \Vert x\Vert ^{7/8}\) and the bound is

$$\begin{aligned} {{\mathrm{Var}}}\,T~\log \left[ \frac{\Vert x\Vert ^{7/8}}{\sum _{k=1}^\infty (\mathbb {E}|V_k|)^2} \right] \le \sum _{k=1}^\infty Ent(V_k^2). \end{aligned}$$

By working with an averaged version \(F_m\) of \(T\) (similar to that used in [4], but a different definition that simplifies the analysis and requires a new argument) one can ensure that the sum in the denominator on the left is at most order \(\Vert x\Vert ^{3/4}\). (See Proposition 5.3.) Thus we begin our analysis with

$$\begin{aligned} {{\mathrm{Var}}}\,T \le \frac{C}{\log \Vert x\Vert } \sum _{k=1}^\infty Ent(V_k^2) . \end{aligned}$$
(1.12)

Step 1 Bernoulli encoding. If one knows a LSI of the form \(Ent~f^2 \le C \mathbb {E}\Vert \nabla f\Vert _2^2\), then the argument of Benaim–Rossignol would give \(\sum _{k=1}^\infty Ent(V_k^2) \le C\mathbb {E}\Vert \nabla T\Vert _2^2\) and the method of Kesten can give an upper bound on this term by \(C\Vert x\Vert _1\). Combining with (1.12) gives the sub-linear variance bound.

Unfortunately very few distributions satisfy a LSI of the above type. Benaim–Rossignol deal with this by exhibiting certain edge-weight distributions (those in the “nearly gamma” class) as images of Gaussian random variables and using the Gaussian LSI. This does not work for all distributions, so our main idea is to encode general edge-weights using infinite sequences of Bernoulli variables and use the Bernoulli (two-point) LSI.

For simplicity, assume that the edge-weights \(t_e\) are uniformly distributed on \([0,1]\), so that we can encode their values using the binary expansion and i.i.d. Bernoulli (1/2) sequences

$$\begin{aligned} t_e = \sum _{i=1}^\infty \omega _{e,i} 2^{-i}, \text { where } (\omega _{e,1}, \omega _{e,2}, \ldots ) \text { is i.i.d. Bernoulli}(1/2) . \end{aligned}$$

(For general distributions, we compose with the right-continuous inverse of the distribution function of \(t_e\).) Then using the Bernoulli LSI and the argument of Benaim–Rossignol,

$$\begin{aligned} \sum _{k=1}^\infty Ent(V_k^2) \le 2\sum _{k=1}^\infty \sum _{i=1}^\infty \mathbb {E}\left( \Delta _{e_k,i} T\right) ^2 , \end{aligned}$$
(1.13)

where \(\Delta _{e_k,i}\) is the discrete derivative of \(T\) relative to flipping the \(i\)-th bit in the binary expansion of \(t_{e_k}\). This is done in Lemma 6.3.

Step 2 The bulk of the paper is devoted to bounding these discrete derivatives: giving the inequality

$$\begin{aligned} \sum _{k=1}^\infty \sum _{i=1}^\infty \mathbb {E}\left( \Delta _{e_k,i} T\right) ^2 \le C \Vert x\Vert _1 . \end{aligned}$$

This is not a priori clear because flipping bits in the binary expansion can have a large change on \(t_e\) if there are gaps in the support of the edge-weight distribution. We deal with this by considering this influence “on average.” That is, letting \(\mathbb {E}_{< i}\) be the expectation over the binary variables \(\omega _{e_k,1}, \ldots , \omega _{e_k,i-1}\), one has

$$\begin{aligned} \mathbb {E}_{< i}\left( \Delta _{e_k,i}T\right) ^2 = \frac{1}{2^{i-1}} \sum _{\sigma \in \{0,1\}^{i-1}} (T(\sigma ,1)-T(\sigma ,0))^2 , \end{aligned}$$

where we have indicated dependence of \(T\) only on the first \(i\) binary variables. Because the weights are bounded in \([0,1]\), the differences above are at most 1 (and nonzero only when \(e_k\) is in a geodesic from \(0\) to \(x\) for some value of \(t_e\)), so we can telescope them, obtaining the upper bound

$$\begin{aligned}&\mathbf {1}_{\left\{ e_k \in Geo(0,x) \text { for some value of }t_{e_k}\right\} } \frac{1}{2^{i-1}} \sum _{\sigma \in \{0,1\}^{i-1}} (T(\sigma ,1) - T(\sigma ,0)) \\&\quad \le \frac{1}{2^{i-1}} \mathbf {1}_{\left\{ e_k \in Geo(0,x) \text { for some value of } t_{e_k}\right\} } . \end{aligned}$$

Pretending for the moment that the indicator is actually of the event that \(e_k \in Geo(0,x)\), we can sum over \(i\) to give the bound \(2 \mathbf {1}_{\{e_k \in Geo(0,x)\}}\), and sum over \(k\), using Theorem 1.4, to obtain

$$\begin{aligned} \sum _{k=1}^\infty Ent(V_k^2) \le C \sum _k \mathbb {P}(e_k \in Geo(0,x)) \le C\Vert x\Vert _1 . \end{aligned}$$

Step 3 General case. We are not using only uniform \([0,1]\) edge weights, so several complications arise, due both to large edge-weights and to edge-weights near the infimum of the support. The first problem forces the moment condition \(\mathbb {E}t_e^2(\log t_e)_+<\infty \) and the second is related to the change from \(\mathbf {1}_{\{e \in Geo(0,x) \text { for some } t_e\}}\) to \(\mathbf {1}_{\{e \in Geo(0,x)\}}\). However, careful bounding (for example, keeping track of the value \(D_{z,e}\) of the edge-weight above which the edge leaves the geodesic—see Lemma 5.2) leads to the inequality in Proposition 6.4:

$$\begin{aligned} \sum _{k=1}^\infty Ent(V_k^2) \le C \mathbb {E} \sum _e (1-\log F(t_e)) \mathbf {1}_{\{e \in Geo(0,x)\}} , \end{aligned}$$
(1.14)

where \(F(t_e)\) is the distribution function of the weight \(t_e\). Note that this is large when \(t_e\) is near its infimum. In a sense, (1.14) is our version of an LSI, with the penalties due to the fact that we do not have a traditional LSI.

For certain distributions, we can bound \((1-\log F(t_e)) \le C\) and sum as above. In particular, this is possible when there is an atom at the infimum of the support. But for general distributions, we must analyze the number of edges in the geodesic which have weight near the infimum. For this we use the theory of greedy lattice animals. Theorem 6.6 shows that without such an atom, for any \(\epsilon >0\), the expected number of edges in \(Geo(0,x)\) with weight within \(\epsilon \) of the infimum of the support \(I\) satisfies

$$\begin{aligned} \mathbb {E}\#\{e \in Geo(0,x) : t_e \in [I,I+\epsilon ]\} \le C \Vert x\Vert _1 \beta (\epsilon ) , \end{aligned}$$

where \(\beta (\epsilon ) \rightarrow 0\) as \(\epsilon \rightarrow 0\). Combining this with another dyadic partition of the interval \([I,\infty )\) (see Sect. 6.2.4) provides the required control on \((1-\log F(t_e))\) and allows the bound

$$\begin{aligned} \mathbb {E} \sum _e (1-\log F(t_e)) \mathbf {1}_{\{e \in Geo(0,x)\}} \le C\Vert x\Vert _1 . \end{aligned}$$

Along with (1.14), we obtain \(\sum _{k=1}^\infty Ent(V_k^2) \le C\Vert x\Vert \) and complete the proof.

2 Entropy

Recall the definition of entropy with respect to a probability measure \(\mu \):

Definition 2.1

Let \((\varOmega ,\mathcal {F},\mu )\) be a probability space and \(X\in L^1(\varOmega ,\mu )\) be nonnegative. Then

$$\begin{aligned} Ent_\mu X = \mathbb {E}_\mu X\log X - \mathbb {E}_\mu X \log \mathbb {E}_\mu X . \end{aligned}$$

Note that by Jensen’s inequality, \(Ent_\mu X \ge 0\). We will make use of the variational characterization of entropy (see [23, Section 5.2]):

Proposition 2.2

If \(X\) is nonnegative, then

$$\begin{aligned} Ent_\mu (X) = \sup \{\mathbb {E}_\mu XY : \mathbb {E}_\mu e^Y \le 1\}. \end{aligned}$$

This characterization will let us prove the “tensorization” of entropy.

Theorem 2.3

Let \(X\) be a non-negative \(L^2\) random variable on a product probability space

$$\begin{aligned} \left( \prod _{i=1}^\infty \varOmega _i,\mathcal {F}, \mu = \prod _{i=1}^\infty \mu _i \right) , \end{aligned}$$

where \(\mathcal {F} = \bigvee _{i=1}^\infty \mathcal {G}_i\) and each triple \((\varOmega _i,\mathcal {G}_i,\mu _i)\) is probability space. Then

$$\begin{aligned} Ent_\mu X \le \sum _{k=1}^\infty \mathbb {E}_\mu Ent_{i}X , \end{aligned}$$
(2.1)

where \(Ent_i X\) is the entropy of \(X(\omega )=X(\omega _1,\ldots , \omega _i, \ldots ))\) with respect to \(\mu _i\), as a function of the \(i\)-th coordinate \((\)with all other values fixed\()\).

Proof

We use a telescoping argument: write \(\mathcal {F}_k\) for the sigma algebra generated by \(\mathcal {G}_1\cup \cdots \cup \mathcal {G}_k\) (with \(\mathcal {F}_0\) trivial) and compute for any \(n\)

$$\begin{aligned} Ent_\mu X&= \mathbb {E}_\mu X \left[ \log X - \log \mathbb {E}_\mu X \right] \\&= \sum _{k=1}^n \mathbb {E}_\mu X \left[ \log \mathbb {E}_\mu [ X \mid \mathcal {F}_k] - \log \mathbb {E}_\mu [X \mid \mathcal {F}_{k-1}] \right] \\&\quad + \mathbb {E}_\mu X \left[ \log X - \log \mathbb {E}_\mu [X \mid \mathcal {F}_n] \right] \\&= \sum _{k=1}^n \mathbb {E}_\mu \mathbb {E}_{\mu _k} X \left[ \log \mathbb {E}_\mu [ X \mid \mathcal {F}_k] - \log \mathbb {E}_\mu [X \mid \mathcal {F}_{k-1}] \right] \\&\quad + \mathbb {E}_\mu X \left[ \log X - \log \mathbb {E}_\mu [ X \mid \mathcal {F}_n] \right] . \end{aligned}$$

Here \(\mathbb {E}_{\mu _k}\) is expectation with respect to the coordinate \(\omega _k\). Because for almost all realizations of \(\{(\omega _i) : i \ne k\}\),

$$\begin{aligned} \mathbb {E}_{\mu _k} \exp \left( \log \mathbb {E}_\mu [ X \mid \mathcal {F}_k] - \log \mathbb {E}_\mu [ X \mid \mathcal {F}_{k-1}] \right) = 1 , \end{aligned}$$

we use Proposition 2.2 to get the bound

$$\begin{aligned} Ent_\mu X \le \sum _{k=1}^n \mathbb {E}_\mu Ent_k X + \mathbb {E}_\mu X \log X - \mathbb {E}_\mu X \log \mathbb {E}_\mu [ X \mid \mathcal {F}_n]. \end{aligned}$$

Putting \(X_n = \mathbb {E}_\mu [ X \mid \mathcal {F}_n]\), one has

$$\begin{aligned} \mathbb {E}_\mu X \log \mathbb {E}_\mu [X \mid \mathcal {F}_n] = \mathbb {E}_\mu X_n \log X_n. \end{aligned}$$

By martingale convergence (since \(X \in L^1\)), one has \(X_n \rightarrow X\) almost surely. Furthermore, since \(X \in L^2\), the sequence \((X_n \log X_n)\) is uniformly integrable. Therefore

$$\begin{aligned} \mathbb {E}_\mu X \log X - \mathbb {E}_\mu X \log \mathbb {E}_\mu [ X \mid \mathcal {F}_n] \rightarrow 0 \end{aligned}$$

and the proof is complete. \(\square \)

We end this section with the lower bound from Falik and Samorodnitsky [13, Lemma 2.3].

Proposition 2.4

(Falik–Samorodnitsky) If \(X \ge 0\) almost surely,

$$\begin{aligned} Ent_\mu (X^2) \ge \mathbb {E}_\mu X^2 \log \frac{\mathbb {E}_\mu X^2}{(\mathbb {E}_\mu X)^2} . \end{aligned}$$

Proof

First assume \(X > 0\) almost surely and define \(Y = X/ \Vert X\Vert _2\). Then

$$\begin{aligned} Ent_\mu (Y^2) = \mathbb {E}_\mu Y^2 \log Y^2 - \mathbb {E}_\mu Y^2 \log \mathbb {E}_\mu Y^2&= \mathbb {E}_\mu Y^2 \log Y^2 \\&= -2 \mathbb {E}_\mu Y^2 \log (1/Y) . \end{aligned}$$

Apply Jensen to the measure \(\mathbf {E}(\cdot )=\mathbb {E}_\mu (\cdot ~ Y^2)\) and the function \({-}\log \) to obtain

$$\begin{aligned} Ent_\mu (Y^2) \ge -2\mathbb {E}_\mu Y^2 \log \frac{\mathbf {E}(1/Y)}{\mathbb {E}_\mu Y^2} = \mathbb {E}_\mu Y^2 \log \frac{\mathbb {E}_\mu Y^2}{(\mathbb {E}_\mu Y)^2} , \end{aligned}$$

proving the proposition for \(Y\). Now for \(X\),

$$\begin{aligned} Ent_\mu (X^2) = \Vert X\Vert _2^2 Ent_\mu (Y^2) \ge \Vert X\Vert _2^2 \mathbb {E}_\mu Y^2 \log \frac{\mathbb {E}_\mu Y^2}{(\mathbb {E}_\mu Y)^2} = \mathbb {E}_\mu X^2 \log \frac{\mathbb {E}_\mu X^2}{(\mathbb {E}_\mu X)^2} . \end{aligned}$$

If \(X = 0\) with positive probability, we can conclude by a limiting argument applied to \(X_n = \max \{1/n,X\}\).\(\square \)

3 Variance bound for \(\tau (0,x)\)

The mechanism for sublinear behavior of the variance which was identified in [4] can be understood as follows. Since a geodesic from the origin to \(x\) is “one-dimensional,” one expects that most edges in the lattice have small probability to lie in it: the edges have small influence. This is not true of edges very close to the origin. To circumvent this difficulty, Benjamini et al. considered an averaged version of the passage time (see [4, Lemma 3]), which they subsequently compare to the actual passage time from \(0\) to \(x\). It was brought to our attention by Sodin (see [29, Section 3]) that their argument can be replaced by a geometric average. This observation was made earlier by Alexander and Zygouras in [2] for polymer models. Let \(x \in \mathbb {Z}^d\) and \(B_m\) be a box of the form \([-m,m]^d\) for \(m=\lceil \Vert x\Vert _1\rceil ^{1/4}\). Define

$$\begin{aligned} F_m = \frac{1}{\# B_m} \sum _{z \in B_m} \tau (z,z+x) . \end{aligned}$$
(3.1)

Note that by (1.6), \({{\mathrm{Var}}}\,F_m < \infty \).

3.1 Approximating \(\tau (0,x)\) by \(F_m\)

Because of the choice of \(m\), the variance of \(F_m\) closely approximates that of \(\tau \):

Proposition 3.1

Assume \(\mathbb {E} t_e^2<\infty \). Then there exists \(\mathbf {C}_3>0\) such that

$$\begin{aligned} |{{\mathrm{Var}}}\,\tau (0,x) - {{\mathrm{Var}}}\,F_m| \le \mathbf {C}_3 \Vert x\Vert _1^{3/4}\quad \text { for all } x . \end{aligned}$$

Proof

By subadditivity, for each \(z \in B_m\), \(|\tau (0,x) - \tau (z,z+x)| \le \tau (0,z) + \tau (x,z+x)\). Therefore, writing \(M_x = \max \{\tau (0,z) : z \in B_m \}\) and \(\hat{X} = X - \mathbb {E} X\),

$$\begin{aligned} |{{\mathrm{Var}}}\,\tau (0,x)-{{\mathrm{Var}}}\,F_m| \le (\Vert \hat{\tau }(0,x)\Vert _2 + \Vert \hat{F}_m\Vert _2) \Vert \hat{\tau }(0,x) - \hat{F}_m\Vert _2 . \end{aligned}$$

Using \(\Vert \hat{F}_m\Vert _2 \le \Vert \hat{\tau }(0,x)\Vert _2\), we get the bound

$$\begin{aligned} 2\Vert \hat{\tau }(0,x)\Vert _2 ( \Vert \tau (0,x) - F_m\Vert _2 + \mathbb {E} |\tau (0,x) - F_m|) \le 4\Vert \hat{\tau }(0,x)\Vert _2 \Vert M_x\Vert _2. \end{aligned}$$

Since we assume (1.6), [21, Theorem 1] gives \(\Vert \hat{\tau }(0,x)\Vert _2 \le \mathbf {C}_4 \Vert x\Vert _1^{1/2}\). On the other hand, we can bound \(M_x\) using the following lemma.

Lemma 3.2

If \(\mathbb {E}t_e^2<\infty \), there exists \(\mathbf {C}_5\) such that for all finite subsets \(S\) of \(\mathbb {Z}^d\),

$$\begin{aligned} \mathbb {E} \left[ \max _{x,y \in S} \tau (x,y)\right] ^2 \le \mathbf {C}_5 (\text {diam }S)^2. \end{aligned}$$

Proof

We start with the argument of [20, Lemma 3.5]. Given \(x,y \in S\), we can build \(2d\) disjoint (deterministic) paths from \(x\) to \(y\) of length at most \(\mathbf {C}_6 \Vert x-y\Vert _1\) for some integer \(\mathbf {C}_6\). This means that \(\tau (y,z)\) is bounded above by the minimum of \(2d\) variables \(T_1, \ldots , T_{2d}\), the collection being i.i.d. and each variable distributed as the sum of \(\mathbf {C}_6 \text {diam}(S)\) i.i.d. variables \(t_e\), so

$$\begin{aligned} \mathbb {P}(\tau (x,y) \ge \uplambda ) \le \prod _{i=1}^{2d} \mathbb {P}(T_i \ge \uplambda ) \le \left[ \frac{\mathbf {C}_6\text {diam}(S){{\mathrm{Var}}}\,t_e }{(\uplambda -\mathbf {C}_6 \text {diam}(S) \mathbb {E} t_e)^2} \right] ^{2d}. \end{aligned}$$

Therefore if we fix some \(x_0 \in S\), for \(M = \lceil 2\mathbf {C}_6 \mathbb {E} t_e \rceil \),

$$\begin{aligned} \sum _{\uplambda = M \text {diam}(S)}^\infty \uplambda \max _{y \in S} \mathbb {P}(\tau (x_0,y) \ge \uplambda )&\le (4 \mathbf {C}_6 \text {diam}(S) {{\mathrm{Var}}}\,t_e)^{2d} \sum _{\uplambda =M \text {diam}(S)}^{\infty } {\uplambda }{}^{1-4d} \\&= \mathbf {C}_7 (\text {diam }S)^{2-2d}. \end{aligned}$$

Last, by subadditivity,

$$\begin{aligned} \mathbb {E}\left[ \max _{x,y \in S} \tau (x,y) \right] ^2&\le 4 \mathbb {E} \left[ \max _{y \in S} \tau (x_0,y) \right] ^2 \\&\le 4(M \text {diam }S)^2\\&\quad + 8 (\text {diam }S)\sum _{\uplambda =M\text {diam}(S)}^\infty \uplambda \max _{y \in S} \mathbb {P}(\tau (x_0,y) \ge \uplambda )\\&\le \mathbf {C}_8(\text {diam } S)^2. \end{aligned}$$

\(\square \)

Using the lemma, we find \(\Vert M_x\Vert _2 \le \mathbf {C}_9\text {diam}(B_m) \le \mathbf {C}_{10}\Vert x\Vert _1^{1/4}\). This means

$$\begin{aligned} |{{\mathrm{Var}}}\,\tau (0,x) - {{\mathrm{Var}}}\,F_m| \le 4\mathbf {C}_4\mathbf {C}_{10} \Vert x\Vert _1^{1/2} \Vert x\Vert _1^{1/4} = \mathbf {C}_{11}\Vert x\Vert _1^{3/4}. \end{aligned}$$

\(\square \)

3.2 Bounding the variance by the entropy

Enumerate the edges of \(\mathcal {E}^d\) as \(e_1, e_2, \ldots \). We will bound the variance of \(F_m\) using the martingale decomposition

$$\begin{aligned}F_m-\mathbb {E}F_m = \sum _{k=1}^\infty V_k,\end{aligned}$$

where

$$\begin{aligned} V_k = \mathbb {E}[F_m \mid \mathcal {F}_k] - \mathbb {E}[F_m \mid \mathcal {F}_{k-1}], \end{aligned}$$
(3.2)

and we have written \(\mathcal {F}_k\) for the sigma-algebra generated by the weights \(t_{e_1}, \ldots , t_{e_k}\) (with \(\mathcal {F}_0\) trivial). In particular if \(X\in L^1(\varOmega , \mathbb {P})\), we have

$$\begin{aligned} \mathbb {E}[X\mid \mathcal {F}_k] = \int X\left( (t_e)_{e\in \mathcal {E}^d}\right) \, \prod _{i\ge k+1}\mu \left( \mathrm {d}t_{e_i}\right) . \end{aligned}$$
(3.3)

The idea now is to compare the variance of \(F_m\) to \(\sum _{k=1}^\infty Ent(V_k^2)\). The lower bound comes from the proof of [13, Theorem 2.2].

Lemma 3.3

(Falik–Samorodnitsky) We have the lower bound

$$\begin{aligned} \sum _{k=1}^\infty Ent(V_k^2) \ge {{\mathrm{Var}}}\,F_m ~ \log \left[ \frac{{{\mathrm{Var}}}F_m}{\sum _{k=1}^\infty (\mathbb {E} |V_k|)^2}\right] . \end{aligned}$$
(3.4)

Proof

For \(M \in \mathbb {N}\), define \(\tilde{F}_m = \mathbb {E} [F_m \mid \mathcal {F}_M]\). We first use Proposition 2.4 and the fact that \(\sum _{k=1}^M \mathbb {E} V_k^2 = {{\mathrm{Var}}}\,\tilde{F}_m\):

$$\begin{aligned} \sum _{k=1}^M Ent(V_k^2) \ge \sum _{k=1}^M \mathbb {E} V_k^2 \log \left[ \frac{\mathbb {E}V_k^2}{(\mathbb {E} |V_k|)^2} \right] = - {{\mathrm{Var}}}\tilde{F}_m ~ \sum _{k=1}^M \frac{\mathbb {E}V_k^2}{{{\mathrm{Var}}}\,\tilde{F}_m} \log \left[ \frac{(\mathbb {E} |V_k|)^2}{\mathbb {E}V_k^2} \right] . \end{aligned}$$

Next use Jensen’s inequality with the function \({-}\log \) and sum \(\sum _{k=1}^M \frac{\mathbb {E} V_k^2}{{{\mathrm{Var}}}\,\tilde{F}_m} (\cdot )\) to get the lower bound

$$\begin{aligned} -{{\mathrm{Var}}}\,\tilde{F}_m~ \log \left[ \sum _{k=1}^M \frac{\mathbb {E} V_k^2}{{{\mathrm{Var}}}\,\tilde{F}_m} \cdot \frac{(\mathbb {E} |V_k|)^2}{\mathbb {E} V_k^2} \right] , \end{aligned}$$

which gives the lemma, after a limiting argument to pass to a countable sum.\(\square \)

4 Benaim and Rossignol’s approach

In this section, we explain how the argument developed in [6] can be extended, isolating a more general condition than the “nearly gamma” condition. It includes, for example, all power law distributions with \(2+\epsilon \) moments. We emphasize that the content of this section is independent of the derivation of our main result. In [6], the authors assume that the distribution

$$\begin{aligned} \mu = h(x)\,\mathrm {d}x,~ h \text { continuous} \end{aligned}$$

is an absolutely continuous measure such that

$$\begin{aligned} (\mathrm{supp } h)^\circ :\!{=}\, \{x:h(x)>0\} \subset (0,\infty ) \end{aligned}$$

is an interval. Denoting by \(G(x)\) the distribution function of the standard normal distribution, \(H(x) =\int _{-\infty }^xh(t)\,\mathrm {d}t\), and \(X\) an \(N(0,1)\) variable, the random variable

$$\begin{aligned} Y=T(X), \end{aligned}$$
(4.1)

with \(T=H^{-1}\circ G\), has distribution \(\mu \). Recall the Gaussian logarithmic Sobolev inequality [14, 16, 30]: for any smooth \(f:\mathbb {R}\rightarrow \mathbb {R}\)

$$\begin{aligned} \mathbb {E} f^2(X)\log \frac{f^2(X)}{\mathbb {E}f^2(X)} \le 2\mathbb {E}(f'(X))^2. \end{aligned}$$
(4.2)

Combining (4.1) and (4.2), a calculation yields

$$\begin{aligned} Ent_\mu (f(Y))^2 \le 2\mathbb {E}_\mu ((\psi \cdot f')(Y))^2, \end{aligned}$$
(4.3)

where

$$\begin{aligned} \psi (Y) = \frac{(g\circ G^{-1} \circ H)(Y)}{h(Y)} \end{aligned}$$

for any \(f\) in a suitable Sobolev space.

Benaim and Rossignol apply this inequality to the passage time, using inequality (3.4). It is shown in [6], along the same lines as the proof of Lemma 6.3 that (4.3) implies

$$\begin{aligned} \sum _{k=1}^\infty Ent_\mu (V_k^2)\le 2\sum _{j=1}^\infty \mathbb {E}\left[ \left( \psi \left( t_{e_j}\right) \partial _{t_{e_j}}F_m\right) ^2\right] , \end{aligned}$$
(4.4)

with \(F_m\) as in (3.1). The derivative with respect to the edge weight can be expressed as

$$\begin{aligned} \partial _{t_{e_j}}F_m = \frac{1}{\sharp B_m}\sum _{z\in B_m}\mathbf {1}_{\{e_j\in Geo(z,z+x)\}}. \end{aligned}$$
(4.5)

Observe that the right side of (4.5) is a decreasing function of the edge weight \(t_{e_j}\).

The following simple asymptotics appear in [6, Lemma 5.2]:

Lemma 4.1

$$\begin{aligned} g\circ G^{-1}(y)&\sim y \sqrt{-2\log y}, \quad y\rightarrow 0, \end{aligned}$$
(4.6)
$$\begin{aligned} g\circ G^{-1}(y)&\sim (1-y)\sqrt{-2\log (1-y)}, \quad y\rightarrow 1. \end{aligned}$$
(4.7)

That is, in each case the ratio of the left to the right side tends to 1.

Suppose that there is a constant \(\mathbf {C}_{12}>0\) such that

$$\begin{aligned} \frac{H(t)\sqrt{-\log t}}{h(t)}\le \mathbf {C}_{12} \end{aligned}$$
(4.8)

for all \(t\) with \(I\le t \le I+ \delta \), with \(\delta >0\) and \(I\) the left endpoint of the interval \((\mathrm{supp } h)^\circ \) [as in (1.9)]. The condition (4.8) holds, for example, if the density \(h\) is monotone near \(I\) or if \(h(t)\asymp (t-I)^\alpha \) for some (integrable) power \(\alpha \). The latter condition appears in [6, Lemma 5.3].

For \(M>0\) such that \(F(M)<1\), the expectation in (4.4) can be computed as

$$\begin{aligned} \mathbb {E}\left[ \left( \psi \left( t_{e_j}\right) \partial _{t_{e_j}}F_m\right) ^2\right]&= \mathbb {E}\mathbb {E}_{\mu _j}\left[ \left( \psi \left( t_{e_j}\right) \partial _{t_{e_j}}F_m\right) ^2\right] \nonumber \\&=\mathbb {E}\mathbb {E}_{\mu _j}\left[ \left( \psi \left( t_{e_j}\right) \partial _{t_{e_j}}F_m\right) ^2; t_{e_j} \le M\right] \nonumber \\&\quad +\mathbb {E}\mathbb {E}_{\mu _j}\left[ \left( \psi \left( t_{e_j}\right) \partial _{t_{e_j}}F_m\right) ^2; t_{e_j} > M\right] . \end{aligned}$$
(4.9)

For \(t_{e_j}\le M\), (4.6) implies that the first term in (4.9) is bounded by

$$\begin{aligned} \left( \max \left\{ \mathbf {C}_{12},\sup _{\delta \le t\le M} h(t)^{-1}\right\} \right) ^2\cdot \mathbb {E}_{\mu _j}\left( \partial _{t_{e_j}}F_m\right) ^2. \end{aligned}$$

The maximum is finite by assumption, and we have, by Cauchy–Schwarz,

$$\begin{aligned} \mathbb {E}\left( \partial _{t_{e_j}}F_m\right) ^2 \le \frac{1}{\sharp B_m}\sum _{z\in B_m}\mathbb {E}\left( \mathbf {1}_{\{e_j\in Geo(z,z+x)\}}\right) . \end{aligned}$$

From there, one can conclude the argument as in Sects. 6.2.4 and 6.3.

As for the second term in (4.9), assume first that

$$\begin{aligned} \psi \left( t_{e_j}\right) \le \mathbf {C}_{13}\sqrt{t_{e_j}}. \end{aligned}$$
(4.10)

This is the “nearly gamma” condition of Benaim and Rossignol. The right side of (4.10) is increasing in \(t_{e_j}\). Using this in (4.9) together with the Chebyshev association inequality [8, Theorem 2.14], we find

$$\begin{aligned} \mathbb {E}\mathbb {E}_{\mu _j}\left[ \left( \psi \left( t_{e_j}\right) \partial _{t_{e_j}}F_m\right) ^2; t_{e_j} > M\right]&\le \mathbf {C}^2_{13}\mathbb {E}\mathbb {E}_{\mu _j}\left( \sqrt{t_{e_j}}\cdot \partial _{t_{e_j}}F_m\right) ^2\nonumber \\&\le \mathbf {C}_{13}^2\mathbb {E}\left( t_{e_j}\right) \cdot \mathbb {E}\left( \partial _{t_{e_j}}F_m\right) ^2. \end{aligned}$$
(4.11)

The previous argument shows that the condition (4.10) is not necessary: it is sufficient that \(\psi \) be bounded by some increasing, square integrable function of \(t_{e_j}\). Suppose for example that \(t\mapsto h(t)\) is decreasing for \(t>M\). In this case, by (4.6), we have

$$\begin{aligned} \psi \left( t_{e_j}\right) \mathbf {1}_{\left\{ t_{e_j} > M\right\} }&= \frac{(g\circ G^{-1} \circ H)\left( t_{e_j}\right) }{h\left( t_{e_j}\right) }\mathbf {1}_{\left\{ t_{e_j} > M\right\} } \nonumber \\&\le \mathbf {C}_{14} \frac{\left( 1-H\left( t_{e_j}\right) \right) \cdot \sqrt{-2\log \left( 1-H\left( t_{e_j}\right) \right) }}{h\left( t_{e_j}\right) }\mathbf {1}_{\left\{ t_{e_j} > M\right\} }. \end{aligned}$$
(4.12)

Let us denote by \(K\left( t_{e_j}\right) \) the expression in (4.12). For \(t>M\), we have

$$\begin{aligned} 1-H(t) = \int _t^\infty h(s)\,\mathrm {d}s&= \int _t^\infty s^{2/3+\epsilon } s^{-2/3-\epsilon } h(s)\,\mathrm {d}s \\&\le \left( \int _t^\infty s^{2+3 \epsilon } \, h(s)\mathrm {d}s\right) ^{1/3}\left( \int _t^\infty s^{-1-3\epsilon /2}h(s)\,\mathrm {d}s\right) ^{2/3} \\&\le \mathbf {C}_{15} h(t)^{2/3}, \end{aligned}$$

assuming \(h(s)\) is decreasing for \(s> M\) and that the distribution posesses \(2+3\epsilon \) moments. We have used the \(L^3-L^{3/2}\) Hölder inequality. This gives

$$\begin{aligned} K(t) \le \mathbf {C}_{14}\mathbf {C}_{15} h(t)^{-1/3}\sqrt{-2\log (1-H(t))} \cdot \mathbf {1}_{\{t>M\}}. \end{aligned}$$

Thus \(K(t)\) is bounded by a quantity which is increasing in \(t\). Using the Chebyshev association inequality as in (4.11), we find

$$\begin{aligned}&\mathbb {E}\mathbb {E}_{\mu _j}\left[ \left( \psi \left( t_{e_j}\right) \partial _{t_{e_j}}F_m\right) ^2; t_{e_j} > M\right] \\&\quad \le (\mathbf {C}_{14}\mathbf {C}_{15})^2\mathbb {E}\left( h^{-1/3}\left( t_{e_j}\right) \sqrt{-2\log \left( 1-H\left( t_{e_j}\right) \right) } \right) ^2\\&\quad \qquad \times \mathbb {E}\left( \partial _{t_{e_j}}F_m\right) ^2. \end{aligned}$$

We are left with the task of estimating the first expectation, which is

$$\begin{aligned} \int h(s)^{-2/3}(-2\log (1-H(s)) h(s)\,\mathrm {d}s=\int h(s)^{1/3}(-2\log (1-H(s))\,\mathrm {d}s. \end{aligned}$$

We again use polynomial weights and \(L^3-L^{3/2}\):

$$\begin{aligned} \int h(s)^{1/3}(-2\log (1-H(s)))\,\mathrm {d}s&= \int s^{-2/3-\epsilon } s^{2/3+\epsilon }h(s)^{1/3}(-2\log (1-H(s)))\,\mathrm {d}s\\&\le \left( \int s^{-1-3\epsilon /2}\,\mathrm {d}s\right) ^{2/3}\\&\quad \quad \times \left( \int s^{2+3\epsilon }(-2\log (1-H(s)))^3h(s)\,\mathrm {d}s\right) ^{1/3}. \end{aligned}$$

A further application of Hölder’s inequality allows one to control the logarithm, at the cost of an arbitrarily small increase in the moment assumption. It follows that

$$\begin{aligned} \mathbb {E}\left( \psi \left( t_{e_j}\right) \partial _{t_{e_j}}F_m\right) ^2\le \mathbf {C}_{16}\mathbb {E}\left( \partial _{t_{e_j}}F_m\right) ^2 \end{aligned}$$

if the distribution \(\mu \) has \(2+\epsilon '\) moments. In conclusion, Benaim and Rossignol’s argument extends to the case of distributions with \(2+\epsilon \) moments whose densities are positive and eventually decreasing.

One can derive many variants of the above, the key point being the application of positive association in (4.11).

5 The lower bound

In this section we derive the first generalization of Kesten’s geodesic length estimate and show how it is used to bound the sum \(\sum _{k=1}^\infty (\mathbb {E}|V_k|)^2\) appearing in (3.4). Let \(\mathcal {G}\) be the set of all finite self-avoiding geodesics.

Lemma 5.1

Assuming (1.6) and (1.7), there exists \(\mathbf {C}_{17}>0\) such that for all \(x\) and all finite \(E \subset \mathcal {E}^d\),

$$\begin{aligned} \mathbb {E} \max _{\gamma \in \mathcal {G}} \# (E \cap \gamma ) \le \mathbf {C}_{17} \text {diam}(E). \end{aligned}$$

Proof

Choose \(a, \mathbf {C}_2>0\) from Theorem 1.5. If \(\# (E \cap \gamma ) \ge \uplambda \) for some \(\gamma \in \mathcal {G}\), then we may find the first and last intersections (say \(y\) and \(z\) respectively) of \(\gamma \) with \(V\), the set of endpoints of edges in \(E\). The portion of \(\gamma \) from \(y\) to \(z\) is then a geodesic with at least \(\uplambda \) edges. This means

$$\begin{aligned} \mathbb {P}(\# (E \cap \gamma ) \ge \uplambda \quad \text { for some }\gamma \in \mathcal {G}) \le (\# V) \exp (-\mathbf {C}_2 \uplambda ) + \mathbb {P}\left( \max _{y,z \in V}\tau (y,z) \ge a \uplambda \right) . \end{aligned}$$

Therefore

$$\begin{aligned} \mathbb {E} \max _{\gamma \in \mathcal {G}} \#(E \cap \gamma )&\le \text {diam}(E) + \sum _{\uplambda =\text {diam}(E)}^\infty (\#V) \exp (-\mathbf {C}_2 \uplambda ) \\&\quad + \sum _{\uplambda = \text {diam}(E) }^\infty \mathbb {P}\left( \max _{y,z \in V}\tau (y,z) \ge a \uplambda \right) . \end{aligned}$$

By the inequality \(\text {diam}(E) \ge \mathbf {C}_{18}(\#V)^{1/d}\) for some universal \(\mathbf {C}_{18}\), the middle term is bounded uniformly in \(E\), so we get the upper bound

$$\begin{aligned} \mathbf {C}_{19}\text {diam}(E) + \frac{1}{a} \mathbb {E} \max _{y,z \in V} \tau (y,z). \end{aligned}$$

By Lemma 3.2, this is bounded by \(\mathbf {C}_{20} \text {diam}(E)\).\(\square \)

We will now apply Lemma 5.1 to get an upper bound on \(\sum _{k=1}^\infty (\mathbb {E}|V_k|)^2\). To do so, we use a simple lemma, a variant of which is already found in various places, including the work of Benaim–Rossignol [6, Lemma 5.9]. For its statement, we write an arbitrary element \(\omega \in \varOmega \) as \((t_{e^c},t_e)\), where \(t_{e^c} = (t_f : f \ne e)\). Further, set

$$\begin{aligned} S :=\, \sup \text {supp}(\mu ) = \sup \{x : F(x)<1\} \in \mathbb {R} \cup \{\infty \}. \end{aligned}$$

We use the following short-hand:

$$\begin{aligned} \tau _z = \tau (z,z+x). \end{aligned}$$

Lemma 5.2

For \(e \in \mathcal {E}^d\) and \(z \in \mathbb {Z}^d\), the random variable

$$\begin{aligned} D_{z,e} :=\, \sup \{r < S : e \text { is in a geodesic from }z \text { to } z+x \text { in } (t_{e^c},r)\}\ \end{aligned}$$

has the following properties almost surely.

  1. 1.

    \(D_{z,e} < \infty \).

  2. 2.

    For \(s \le t < S\),

    $$\begin{aligned} \tau _z(t_{e^c}, t) - \tau _z(t_{e^c},s) = \min \{t-s, (D_{z,e}-s)_+\}. \end{aligned}$$
  3. 3.

    For \(s<D_{z,e}\), \(e\in Geo(z,z+x)\) in \((t_{e^c},s)\).

Proof

Part 1 is clear if \(S < \infty \). Otherwise choose any path \(\gamma \) not including \(e\). Then for \(r\) larger than the passage time of this path, \(e\) cannot be in a geodesic in \((t_{e^c},r)\), giving \(D_{z,e} < \infty \).

If \(e\) is in a geodesic \(\gamma \) in \((t_{e^c},t)\) and \(t \ge s\) then the passage time of \(\gamma \) decreases by \(t-s\) in \((t_{e^c},s)\). Since the passage time of no other path decreases by more than \(t-s\), \(\gamma \) is still a geodesic in \((t_{e^c},s)\). This shows that

$$\begin{aligned} \mathbf {1}_{\{e \text { is in a geodesic from } z \text { to } z+x\}} \text { is a non-increasing function of }t_e. \end{aligned}$$
(5.1)

Therefore if \(t < D_{z,e}\), \(e\) is in a geodesic in \((t_{e^c},t)\) and by the above argument, for any \(s \le t\), part 2 holds. We can then extend to \(s \le t \le D_{z,e}\) by continuity.

If \(D_{z,e} < s \le t\) then \(e\) is not in a geodesic from \(z\) to \(z+x\) in \((t_{e^c},s)\). By (1.10), we can almost surely find a geodesic \(\gamma \) in \((t_{e^c},s)\) not containing \(e\) and this path has the same passage time in \((t_{e^c},t)\). However all other paths have no smaller passage time, so \(\tau _z(t_{e^c},t) - \tau _z(t_{e^c},s) = (D_{z,e}-s)_+\) almost surely, proving part 2 in this case. We can then extend the result to \(D_{z,e} \le s \le t\) by continuity and for \(s \le D_{z,e} \le t\) write

$$\begin{aligned} \tau _z(t_{e^c},t) - \tau _z(t_{e^c},s) = \tau _z(t_{e^c},t) - \tau _z(t_{e^c},D_{z,e}) + \tau _z(t_{e^c},D_{z,e}) - \tau _z(t_{e^c},s), \end{aligned}$$

and use the other cases to complete the proof.

For part 3, let \(s<D_{z,e}\), so that by (5.1), \(e\) is in a geodesic \(\gamma _1\) in \((t_{e^c},\frac{s+D_{z,e}}{2})\) from \(z\) to \(x+z\). Assume for a contradiction that \(e\) is not in every geodesic from \(z\) to \(x+z\) in \((t_{e^c},s)\), and choose \(\gamma _2\) as one that does not contain \(e\). Because \(\frac{s + D_{z,e}}{2} \ge s\), \(\gamma _1\) is still a geodesic in \((t_{e^c},s)\) and therefore has the same passage time in this configuration as \(\gamma _2\). But then in \((t_{e^c},\frac{s+D_{z,e}}{2})\) it has strictly larger passage time, contradicting the fact that it is a geodesic.\(\square \)

Proposition 5.3

Assuming (1.6) and (1.7), there exists \(\mathbf {C}_{21}\) such that

$$\begin{aligned} \sum _{k=1}^\infty (\mathbb {E}|V_k|)^2 \le \mathbf {C}_{21} \Vert x\Vert _1^{\frac{5-d}{4}}\quad \text { for all } x. \end{aligned}$$

Proof

Using the definition of \(V_k\),

$$\begin{aligned} \mathbb {E}|V_k|&= \frac{1}{\# B_m} \mathbb {E} \left| \mathbb {E} \left[ \sum _{z \in B_ m} \tau _z \mid \mathcal {F}_k \right] - \mathbb {E} \left[ \sum _{z \in B_m} \tau _z \mid \mathcal {F}_{k-1} \right] \right| \nonumber \\&\le \frac{1}{\# B_m} \sum _{z \in B_m} \mathbb {E} \left| \mathbb {E} \left[ \tau _z \mid \mathcal {F}_k \right] - \mathbb {E} \left[ \tau _z \mid \mathcal {F}_{k-1} \right] \right| . \end{aligned}$$
(5.2)

Write a configuration \(\omega \) as \(\left( t_{<k},t_{e_k},t_{>k}\right) \), where

$$\begin{aligned} t_{<k} = \left( t_{e_j}:j < k\right) \quad \text { and }\quad t_{>k} = \left( t_{e_j} : j > k\right) . \end{aligned}$$

The summand in (5.2) becomes

$$\begin{aligned}&~\int \left| \int \tau _z(t_{<k},t,t_{>k}) \mathbb {P}(\text {d}t_{>k}) - \int \tau _z(t_{<k},s,t_{>k}) \mu (\text {d}s)\mathbb {P}(\text {d}t_{>k}) \right| \mu (\text {d}t) \mathbb {P}(\text {d}t_{<k}) \\&\quad ~\le 2 \mathbb {E} \int \int _{t \ge s} \left| \tau _z(t_{<k},t,t_{>k}) - \tau _z(t_{<k},s,t_{>k}) \right| \mu (\text {d}s) \mu (\text {d}t). \end{aligned}$$

By Lemma 5.2 and \(\mathbb {E}t_e<\infty \), this equals

$$\begin{aligned}&\mathbb {E} \int \int _{t \ge s} \min \{ t-s, \left( D_{z,e_k}-s\right) _+\} \mu (\text {d}s) \mu (\text {d}t) \\&\qquad \le 2 \int t\int _{s < D_{z,e_k}} \mu (\text {d}s) \mu (\text {d}t) = \mathbf {C}_{22} F\left( D_{z,e_k}^-\right) . \end{aligned}$$

Using part 3 of the same lemma, \(\mathbb {E} F\left( D_{z,e_k}^-\right) \le \mathbb {P}(e_k \in Geo(z,z+x))\). Therefore

$$\begin{aligned} \mathbb {E}|V_k| \le \mathbf {C}_{22} \frac{1}{\#B_m} \sum _{z \in B_m} \mathbb {P}(e_k \in Geo(z,z+x)). \end{aligned}$$

By translation invariance, the above probability equals \(\mathbb {P}(e_k+z \in Geo(0,x))\), so we get the bound \(\frac{\mathbf {C}_{22}}{\#B_m} \mathbb {E} \# \left[ Geo(0,x) \cap \{e_k+z : z \in B_m\} \right] \). Lemma 5.1 provides \(\mathbf {C}_{23}\) such that this is no bigger than \(\mathbf {C}_{23}\frac{\text { diam }B_m}{ \#B_m}\). Hence

$$\begin{aligned} \mathbb {E} |V_k| \le \mathbf {C}_{24} \Vert x\Vert _1^{\frac{1-d}{4}}. \end{aligned}$$

This leads to

$$\begin{aligned} \sum _{k=1}^\infty (\mathbb {E} |V_k|)^2&\le \mathbf {C}_{24} \Vert x\Vert _1^{\frac{1-d}{4}} \sum _{k=1}^\infty \mathbb {E}|V_k| \\&\le \mathbf {C}_{25} \Vert x\Vert _1^{\frac{1-d}{4}} \frac{1}{\# B_m}\sum _{z\in B_m} \sum _{k=1}^\infty \mathbb {P}(e_k \in Geo(z,z+x))\\&\le \mathbf {C}_{26} \Vert x\Vert _1^{\frac{5-d}{4}}. \end{aligned}$$

In the last inequality we have used Theorem 1.4.\(\square \)

6 Sublinear variance for general distributions

Combining the results from the previous sections, we have shown so far that if (1.6) and (1.7) hold then

$$\begin{aligned} {{\mathrm{Var}}}\, \tau (0,x) \le {{\mathrm{Var}}}\,F_m + \mathbf {C}_3\Vert x\Vert _1^{3/4} \le \mathbf {C}_{3}\Vert x\Vert _1^{3/4} + \left[ \log \left[ \frac{{{\mathrm{Var}}}F_m}{\Vert x\Vert _1^{\frac{5-d}{4}}} \right] \right] ^{-1} \sum _{k=1}^\infty Ent(V_k^2). \end{aligned}$$
(6.1)

Our goal now is to bound the sum by \(C\Vert x\Vert _1\). We will do this using a Bernoulli encoding.

6.1 Bernoulli encoding

We will now view our edge variables as the push-forward of a Bernoulli sequence. Specifically, for each edge \(e\), let \(\varOmega _e\) be a copy of \(\{0,1\}^\mathbb {N}\) with the product sigma-algebra. We will construct a measurable map \(T_e:\varOmega _e \rightarrow \mathbb {R}\) using the distribution function \(F\). To do this, we create a sequence of partitions of the support of \(\mu \). Recalling \(I :{=} \inf \text {supp}(\mu ) = \inf \{x : F(x) > 0\}\), set

$$\begin{aligned} a_{0,j} = I\quad \text { and }\quad a_{i,j} = \min \left\{ x : F(x) \ge \frac{i}{2^j}\right\} \quad \text { for } j \ge 1 \text { and } 1 \le i \le 2^j-1. \end{aligned}$$

Note that by right continuity of \(F\), the minimum above is attained; that is,

$$\begin{aligned} F(a_{i,j}) \ge \frac{i}{2^j}\quad \text { for } j \ge 1\quad \text { and }\quad 0 \le i \le 2^j -1. \end{aligned}$$
(6.2)

Let us note two properties of the sequence.

$$\begin{aligned}&\text {For }j \ge 1,~ a_{0,j} \le a_{1,j} \le \cdots \le a_{2^j-1,j}. \end{aligned}$$
(6.3)
$$\begin{aligned}&\text {For }i=0, \ldots , 2^j-1, ~ x \ge a_{i,j} \text { if and only if }F(x) \ge \frac{i}{2^j} \text { and } x \ge a_{0,j}.\quad \quad \end{aligned}$$
(6.4)

Each \(\omega \in \varOmega _e\) gives us an “address” for a point in the support of \(\mu \). Given \(\omega = (\omega _1, \omega _2, \ldots )\) and \(j \ge 1\), we associate a number \(T_j(\omega )\) by

$$\begin{aligned} T_j(\omega ) = a_{i(\omega ,j),j},\quad \text { where } i(\omega ,j) = \sum _{l=1}^j 2^{j-l} \omega _l. \end{aligned}$$

\(i(\omega ,j)\) is just the number between \(0\) and \(2^j - 1\) that corresponds to the binary number \(\omega _1 \cdots \omega _j\). It will be important to note that if \(\omega _i \le \hat{\omega }_i\) for all \(i \ge 1\) (written \(\omega \le \hat{\omega }\)), then \(i(\omega ,j) \le i(\hat{\omega },j)\) for all \(j \ge 1\). This, combined with the monotonicity statement (6.3), implies

$$\begin{aligned} \omega \le \hat{\omega }\Rightarrow T_j(\omega ) \le T_j(\hat{\omega })\quad \text { for all } j \ge 1. \end{aligned}$$
(6.5)

It is well-known that one can represent Lebesgue measure on \([0,1]\) using binary expansions and Bernoulli sequences. One way to view the encoding \(T\) in Lemma 6.1 is a composition of this representation with the right-continuous inverse of the distribution function \(F\). The function \(T_j\) instead uses an inverse approximated by simple functions taking dyadic values.

Lemma 6.1

For each \(\omega \), the numbers \((T_j(\omega ))\) form a non-decreasing sequence and have a limit \(T(\omega )\). This map \(T:\varOmega _e \rightarrow \mathbb {R} \cup \{\infty \}\) is measurable and has the following properties.

  1. 1.

    \((\)Monotonicity\()\) If \(\omega \le \hat{\omega }\) then \(T(\omega ) \le T(\hat{\omega })\).

  2. 2.

    \((\)Nesting\()\) For any \(\omega \in \varOmega _e\) and \(j \ge 1\), if \(i(\omega ,j) < 2^j-1\) then

    $$\begin{aligned} a_{i(\omega ,j),j} \le T(\omega ) \le a_{i(\omega ,j)+1,j}. \end{aligned}$$
  3. 3.

    If \(\omega _k = 0\) for some \(k \ge 1\) then \(T(\omega ) < \infty \).

  4. 4.

    Letting \(\pi \) be the product measure \(\prod _{l \in \mathbb {N}} \pi _l\), with each \(\pi _l\) uniform on \(\{0,1\}\), we have

    $$\begin{aligned} \pi \circ T^{-1} = \mu . \end{aligned}$$

    By part 3, \(T\) is \(\pi \)-almost surely finite.

Proof

The functions \(T_j\) are each measurable since their ranges are finite and the pre-image of each point is a cylinder in \(\varOmega _e\). If we show that \(T_j \rightarrow T\) pointwise then \(T\) will also be measurable. Given \(\omega \in \varOmega _e\), we have

$$\begin{aligned} \frac{i(\omega ,j)}{2^j} = \frac{1}{2^j} \sum _{l=1}^j 2^{j-l}\omega _l = \sum _{l=1}^j 2^{-l} \omega _l \le \sum _{l=1}^{j+1} 2^{-l} \omega _l = \frac{i(\omega ,j+1)}{2^{j+1}}. \end{aligned}$$

Therefore if \(x\) is such that \(F(x) \ge \frac{i(\omega ,j+1)}{2^{j+1}}\) then also \(F(x) \ge \frac{i(\omega ,j)}{2^j}\). This means that if \(i(\omega ,j) > 0\),

$$\begin{aligned} T_j(\omega ) = \min \left\{ x : F(x) \ge \frac{i(\omega ,j)}{2^j} \right\} \le \min \left\{ x : F(x) \ge \frac{i(\omega ,j+1)}{2^{j+1}} \right\} = T_{j+1}(\omega ). \end{aligned}$$

Otherwise if \(i(\omega ,j)=0\) then \(T_{j+1}(\omega ) \ge a_{0,j+1}=a_{0,j}=T_j(\omega )\). In either case, \((T_j(\omega ))\) is monotone and has a limit \(T(\omega )\).

For part 1, we simply take limits in (6.5). To prove part 2, we note the lower bound follows from monotonicity. For the upper bound, take \(\omega \in \varOmega _e\) and let \(k \ge j\). Then

$$\begin{aligned} \frac{i(\omega ,k)}{2^k} = \sum _{l=1}^k 2^{-l} \omega _l \le \sum _{l=1}^j 2^{-l} \omega _l + \sum _{l=j+1}^\infty 2^{-l} = \frac{i(\omega ,j)+1}{2^j} \le \frac{2^j-1}{2^j}. \end{aligned}$$

If \(\omega \) is the zero sequence then \(T(\omega ) = I\) and \(T(\omega ) \le a_{i(\omega ,j)+1,j}\). Otherwise we can find \(k \ge j\) such that \(i(\omega ,k) \ne 0\). For this \(k\), \(F(x) \ge \frac{i(\omega ,j)+1}{2^j}\) implies \(F(x) \ge \frac{i(\omega ,k)}{2^k}\), giving

$$\begin{aligned} T_k(\omega ) = a_{i(\omega ,k),k} \le a_{i(\omega ,j)+1,j}. \end{aligned}$$

Taking the limit in \(k\) gives the result.

In part 3, we assume that \(\omega _k=0\) for some \(k \ge 1\). Then \(i(\omega ,k+1) < 2^{k+1}-1\) and therefore by part 2,

$$\begin{aligned} T(\omega ) \le a_{i(\omega ,k+1)+1,j} \le a_{2^{k+1}-1,j} <\infty . \end{aligned}$$

Last we must show that \(\pi \circ T^{-1} = \mu \). The first step is to show that for each \(x \in \mathbb {R}\),

$$\begin{aligned} \pi \circ T_j^{-1} ((-\infty ,x]) \rightarrow \pi \circ T^{-1}((-\infty ,x]). \end{aligned}$$

Consider the sets

$$\begin{aligned} S_j(x) = \{\omega \in \varOmega _e : T_j(\omega ) \le x\}. \end{aligned}$$

If \(T_{j+1}(\omega ) \le x\) then \(T_j(\omega ) \le T_{j+1}(\omega ) \le x\), so these sets are decreasing. If \(\omega \) is in their intersection then \(T_j(\omega ) \le x\) for all \(j\). Since \(T_j(\omega ) \rightarrow T(\omega )\) this means \(T(\omega ) \le x\) and thus \(\omega \in S(x) := \{\omega \in \varOmega _e : T(\omega ) \le x\}\). Conversely, if \(\omega \in S(x)\) then \(T(\omega ) \le x\) and so \(T_j(\omega ) \le T(\omega ) \le x\) for all \(j\), meaning \(\omega \in \cap _j S_j(x)\). Therefore \(\pi \circ T_j^{-1} ((-\infty ,x])\) converges to \(\pi \circ T^{-1}((-\infty ,x])\).

Next we claim that

$$\begin{aligned} x \ge a_{0,j} \Rightarrow \pi \circ T_j^{-1} ((-\infty ,x]) = 2^{-j} \max \left\{ i+1 : F(x) \ge \frac{i}{2^j} \right\} . \end{aligned}$$
(6.6)

The left side of the equality is \(\pi (\{\omega : T_j(\omega ) \le x\})\). The function \(T_j\) is constant on sets of \(\omega \) with the same first \(j\) entries. By definition, if \(\omega \) has first \(j\) entries \(\omega _1 \cdots \omega _j\) then \(T_j(\omega ) = T_j(\omega _1 \cdots \omega _j) = a_{i(\omega ,j),j}\). So

$$\begin{aligned} \pi \circ T_j^{-1}((-\infty ,x]) = 2^{-j} \# \left\{ (\omega _1, \ldots , \omega _j) : a_{i(\omega ,j),j} \le x\right\} . \end{aligned}$$

Also, since \(x \ge a_{0,j}\), (6.4) gives

$$\begin{aligned} \pi \circ T_j^{-1} ((-\infty ,x]) = 2^{-j} \# \left\{ (\omega _1, \ldots , \omega _j) : F(x) \ge \sum _{l=1}^j 2^{-l} \omega _l \right\} \ . \end{aligned}$$

This is exactly the right side of (6.6).

By (6.6), \(\left| \pi \circ T_j^{-1}((-\infty ,x]) - F(x) \right| \le 2^{-j}\) and so \(\pi \circ T_j^{-1} ((-\infty ,x]) \rightarrow F(x)\), completing the proof of part 4.\(\square \)

6.2 Bound on discrete derivatives

In this section we prove the result:

Theorem 6.2

Assume (1.6) and (1.7). There exists \(\mathbf {C}_{27}\) such that

$$\begin{aligned} \sum _{k=1}^\infty Ent(V_k^2) \le \mathbf {C}_{27}\Vert x\Vert _1. \end{aligned}$$

The proof will be broken into subsections. In the first we apply Bonami’s inequality to the Bernoulli encoding of \(F_m\) to get a sum involving discrete derivatives. The next subsection uses the quantities \(D_{z,e}\) from Lemma 5.2 to control the sum of derivatives. In the third subsection, we give a lemma based on the theory of greedy lattice animals and in the final subsection, we use this lemma to achieve the bound \(\mathbf {C}_{27}\Vert x\Vert _1\).

6.2.1 Application of Bonami’s inequality

We will view \(F_m\) as a function of sequences of Bernoulli variables, so define

$$\begin{aligned} \varOmega _B = \prod _e \varOmega _e \end{aligned}$$

where \(\varOmega _e\) is, as in the last section, a copy of \(\{0,1\}^\mathbb {N}\). The measure on \(\varOmega _e\) is \(\pi _e\), a product of the form \(\prod _{j \ge 1} \pi _{e,j}\) with \(\pi _{e,j}\) uniform on \(\{0,1\}\) and the measure on \(\varOmega _B\) is \(\pi : = \prod _e \pi _e\). Here as usual we use the product sigma-algebra. A typical element of \(\varOmega _B\) is denoted \(\omega _B\) and we list the collection of individual Bernoulli variables as

$$\begin{aligned} \omega _B = \left\{ \omega _{e,j} {:}\, e \in \mathcal {E}^d,~ j \ge 1\right\} . \end{aligned}$$

Last, calling \(T_e\) the map from Lemma 6.1 on \(\varOmega _e\), the product map \(T :=\prod _e T_e {:}\, \varOmega _B \rightarrow \varOmega \) is defined

$$\begin{aligned} T (\omega _B) = \left( T_e(\omega _e) {:} \,e \in \mathcal {E}^d\right) . \end{aligned}$$

It is measurable and, by Lemma 6.1, pushes the measure \(\pi \) forward to \(\mathbb {P}\), our original product measure on \(\varOmega \).

We consider \(F_m\) as a function on \(\varOmega _B\); that is, we set \(G = F_m \circ T\). The goal is to estimate the derivative of \(G\), so define the derivative relative to \(\omega _{e,j}\) of a function \(f:\varOmega _B \rightarrow \mathbb {R}\) as

$$\begin{aligned} \left( \Delta _{e,j} f\right) (\omega ) = f\left( \omega ^{e,j,+}\right) - f\left( \omega ^{e,j,-}\right) , \end{aligned}$$

where \(\omega ^{e,j,+}\) agrees with \(\omega \) except possibly at \(\omega _{e,j}\), where it is 1, and \(\omega ^{e,j,-}\) agrees with \(\omega \) except possibly at \(\omega _{e,j}\), where it is 0. Then the following analogue of [6, Eq. (3)] holds.

Lemma 6.3

We have the following inequality:

$$\begin{aligned} \sum _{k=1}^\infty Ent(V_k^2) \le \sum _e \sum _{j=1}^\infty \mathbb {E}_\pi (\Delta _{e,j} G)^2. \end{aligned}$$

Proof

Define a filtration of \(\varOmega _B\) by enumerating the edges of \(\mathcal {E}^d\) as \(\{e_1, e_2, \ldots \}\) as before and setting \(\mathcal {G}_k\) as the sigma-algebra generated by \(\left\{ \omega _{e_r,j} {:}\, r \le k, j \in \mathbb {N}\right\} \). Also define \(W_k = \mathbb {E}_\pi \left[ G \mid \mathcal {G}_k\right] -\mathbb {E}_\pi \left[ G \mid \mathcal {G}_{k-1}\right] \). It is straightforward to verify that, because \(\mathbb {P} = \pi \circ T^{-1}\),

$$\begin{aligned} \mathbb {E}[F_m \mid \mathcal {F}_k] (T(\omega _B)) = \mathbb {E}_\pi [G \mid \mathcal {G}_k](\omega _B)\quad \text { for } \pi \text {-almost every }\omega _B \in \varOmega _B. \end{aligned}$$

Therefore \(Ent(V_k^2) = Ent_\pi (W_k^2)\) for each \(k\). Using tensorization of entropy (Theorem 2.3),

$$\begin{aligned} \sum _{k=1}^\infty Ent_\pi (W_k^2) \le \sum _{k=1}^\infty \mathbb {E}_\pi \sum _e \sum _{j=1}^\infty Ent_{\pi _{e,j}}W_k^2. \end{aligned}$$

For this to be true, we need to check the condition \(W_k^2 \in L^2\), or that \(V_k \in L^4\). Since \(V_k\) is a difference of martingale sequence terms, it suffices to show that \(\tau (0,x)\) is in \(L^4\). But this follows from [10, Lemma 3.1]: if \(Y = \min \{t_1, \ldots , t_{2d}\}\) is a minimum of \(2d\) i.i.d. variables distributed as \(t_e\), then \(\tau (0,x) \in L^4\) if and only if \(\mathbb {E}Y^4<\infty \). By Chebyshev’s inequality,

$$\begin{aligned} \mathbb {E}Y^\alpha = \alpha \int _0^\infty y^{\alpha -1} \mathbb {P}(t_e > y)^{2d}~\text {d}y \le C\int _0^\infty y^{\alpha -1-4d}~\text {d}y, \end{aligned}$$

which is finite if \(\alpha < 4d\). In particular, since \(d \ge 2\), one has \(W_k^2 \in L^2\).

Recall the Bonami–Gross inequality [7, 16], which says that if \(f:\{0,1\} \rightarrow \mathbb {R}\) and \(\nu \) is uniform on \(\{0,1\}\) then

$$\begin{aligned} Ent_\nu f^2 \le (1/2)(f(1)-f(0))^2. \end{aligned}$$

Therefore we get the upper bound \(\sum _{j=1}^\infty \sum _e \sum _{k=1}^\infty \mathbb {E}_\pi (\Delta _{e,j}W_k)^2\). For fixed \(e=e_i \text { and }j\),

$$\begin{aligned} \Delta _{e,j}W_k = {\left\{ \begin{array}{ll} 0 &{} \text { if } k<i \\ \mathbb {E}_\pi [\Delta _{e,j}G \mid \mathcal {G}_k] &{} \text { if } k=i \\ \mathbb {E}_\pi [\Delta _{e,j}G \mid \mathcal {G}_k] - \mathbb {E}_\pi [\Delta _{e,j}G \mid \mathcal {G}_{k-1}] &{} \text { if } k > i \end{array}\right. }. \end{aligned}$$

The first follows because when \(k<i\) then \(W_k\) does not depend on \(\omega _{e,j}\), as this variable is integrated out. A similar idea works for the second, noting that \(\Delta _{e,j} \mathbb {E}_\pi [G \mid \mathcal {G}_{k-1}] = 0\). The third is straightforward. Using orthogonality of martingale differences, \(\sum _{k=1}^\infty \mathbb {E}_\pi (\Delta _{e,j}W_k)^2 = \mathbb {E}_\pi (\Delta _{e,j} G)^2\) and this completes the proof.\(\square \)

6.2.2 Control by edges in geodesics

The first major step is to bound the sum of discrete derivatives by a weighted average of edge-weights in geodesics. The bound we give is analogous to what would appear if we had a LSI for \(\mu \) (see the approach in Benaim–Rossignol [6]); however, we get a logarithmic singularity as \(t_e \downarrow I\).

Proposition 6.4

There exists \(\mathbf {C}_{28}\) such that for all \(x\),

$$\begin{aligned} \sum _e \sum _{j=1}^\infty \mathbb {E}_\pi (\Delta _{e,j} G)^2 \le \mathbf {C}_{28} \mathbb {E} \sum _e (1-\log F(t_e)) \mathbf {1}_{\{e \in Geo(0,x)\}}. \end{aligned}$$

Proof

We begin by using convexity of the square function to get

$$\begin{aligned} \sum _e \sum _{j=1}^\infty \mathbb {E}_\pi \left( \Delta _{e,j} G \right) ^2 \le \frac{1}{\#B_m} \sum _{z \in B_m} \left[ \sum _e \sum _{j=1}^\infty \mathbb {E}_\pi \left( \Delta _{e,j} \tau _z \right) ^2 \right] , \end{aligned}$$
(6.7)

where \(\tau _z=\tau (z,z+x)\). Write \(\mathbb {E}_{e^c}\) for expectation relative to \(\prod _{f \ne e} \pi _f\) and for any \(i\ge 1\), let \(\pi _{e,\ge i}\) be the measure \(\prod _{k \ge i} \pi _{e,k}\). Further, for \(j \ge 1\) write

$$\begin{aligned} \omega _B = (\omega _{e^c}, \omega _{e,< j}, \omega _{e,j}, \omega _{e,> j}), \end{aligned}$$

where \(\omega _{e^c}\) is the configuration \(\omega _B\) projected on the coordinates \((\omega _{f,k} : f \ne e,~ k \ge 1)\), \(\omega _{e,<j}\) is \(\omega _B\) projected on the coordinates \((\omega _{e,k} : k < j)\) and \(\omega _{e, > j}\) is \(\omega _B\) projected on the coordinates \((\omega _{e,k} : k > j)\).

The expectation in (6.7) is now

$$\begin{aligned}&~\mathbb {E}_{e^c} \mathbb {E}_{\pi _{e,1}} \cdots \mathbb {E}_{\pi _{e,j-1}} \left[ \mathbb {E}_{\pi _{e,\ge j}} \left( \Delta _{e,j} \tau _z \right) ^2 \right] \nonumber \\&\quad = ~\mathbb {E}_{e^c} \left[ \frac{1}{2^{j-1}} \sum _{\sigma \in \{0,1\}^{j-1}} \left[ \mathbb {E}_{\pi _{e,\ge j}} \left( \Delta _{e,j} \tau _z (\omega _{e^c}, \sigma , \omega _{e,j}, \omega _{e, > j}) \right) ^2 \right] \right] , \end{aligned}$$
(6.8)

and the innermost term is

$$\begin{aligned} \mathbb {E}_{\pi _{e, \ge j}} \left( \tau _z(\omega _{e^c},\sigma , 1, \omega _{e,>j}) - \tau _z(\omega _{e^c}, \sigma ,0, \omega _{e,>j}) \right) ^2 . \end{aligned}$$
(6.9)

Because of Lemma 5.2, we can rewrite (6.9) as

$$\begin{aligned} \mathbb {E}_{\pi _{e,\ge j}} \min \left\{ (T_e(\sigma ,1,\omega _{e,>j})-T_e(\sigma ,0,\omega _{e,>j}))^2, (D_{z,e}-T_e(\sigma ,0,\omega _{e,>j}))_+^2\right\} . \end{aligned}$$
(6.10)

Note that this allows us to assume \(D_{z,e} > I\):

$$\begin{aligned} \mathbb {E}_\pi (\Delta _{e,j} \tau _z)^2 = \mathbb {E}_{e^c} \left[ \frac{1}{2^{j-1}} \sum _{\sigma \in \{0,1\}^{j-1}} \left[ \mathbb {E}_{\pi _{e,\ge j}} \left( \Delta _{e,j} \tau _z (\omega _{e^c}, \sigma , \omega _{e,j}, \omega _{e, > j}) \right) ^2 \right] \mathbf {1}_{\{I<D_{z,e}\}}\right] . \end{aligned}$$
(6.11)

To simplify notation in the case \(j \ge 2\), we write the values \(a_{1,j-1}, \ldots , a_{2^{j-1}-1,j-1}\) as \(a_1, \ldots , a_{2^{j-1}-1}\) and for a fixed \(\sigma \in \{0,1\}^{j-1}\), \(a_\sigma \) for \(a_{i((\sigma ,0,\omega _{e,>j}),j-1),j-1}\) (note that this does not depend on the configuration outside of \(\sigma \)). Also we write \(a'_\sigma \) for the element of the partition that follows \(a_\sigma \) (when there is one; that is, when \(\sigma \) is not \((1, \ldots , 1)\)). Last, we abbreviate \(T_e(\sigma ,c,\omega _{e,>j})\) by \(T_{e,j}(\sigma ,c)\) for \(c=0,1\). With this notation, we claim the inequalities

$$\begin{aligned} a_\sigma \le T_{e,j}(\sigma ,0) \le T_{e,j}(\sigma ,1) \le a'_\sigma \quad \text { when } \sigma \ne (1, \ldots , 1) \quad \text { and }\quad j \ge 2. \end{aligned}$$

The first and third inequalities follow from the nesting part of Lemma 6.1. The second holds because of the monotonicity part. Therefore we can give an upper bound for (6.10) when \(j \ge 2\) of

$$\begin{aligned} {\left\{ \begin{array}{ll} 0 &{} \text { if } \quad D_{z,e} \le a_\sigma \\ \mathbb {E}_{\pi _{e,\ge j}} \min \{D_{z,e}-a_\sigma , T_{e,j}(\sigma ,1) - a_\sigma \}^2 \mathbf {1}_{\{T_{e,j}(\sigma ,0) < D_{z,e}\}} &{} \text { if } \quad \sigma \ne (1, \ldots , 1) \text { and } a_\sigma < D_{z,e} \le a'_\sigma \\ &{}\text { or } \quad \sigma = (1, \ldots , 1) \\ (a'_\sigma - a_\sigma )^2 &{} \text { if }\quad a'_\sigma \le D_{z,e} \end{array}\right. }. \end{aligned}$$

[Here and above we have strict inequality in the condition of the indicator function since when \(T_e(\sigma ,0,\omega _{e,>j})=D_{z,e}\), (6.10) is zero.] With this, when \(j \ge 2\), the integrand of \(\mathbb {E}_{e^c}\) in (6.11) is no bigger than

$$\begin{aligned} \frac{1}{2^{j-1}}&\bigg [ (a_1-a_0)^2 + \cdots + (a_s-a_{s-1})^2 \nonumber \\&+\! \mathbb {E}_{\pi _{e,\ge j}} \min \{D_{z,e}-a_s,T_{e,j}(\sigma (D_{z,e}),1)-a_s\}^2 \mathbf {1}_{\{T_{e,j}(\sigma (D_{z,e}),0) < D_{z,e}\}} \bigg ] \mathbf {1}_{\{I<D_{z,e}\}} . \end{aligned}$$
(6.12)

Here we have written \(s\) for the largest index \(i\) such that \(a_i < D_{z,e}\) and \(\sigma (D_{z,e})\) for the configuration such that \(a_{\sigma (D_{z,e})} = a_s\). In the case \(j=1\), we have the similar upper bound

$$\begin{aligned} \mathbb {E}_{\pi _{e,\ge j}} \min \{ D_{z,e}-I, T_{e,1}(1)-I\}^2 \mathbf {1}_{\{T_{e,1}(0) < D_{z,e}\}} \mathbf {1}_{\{I<D_{z,e}\}}. \end{aligned}$$
(6.13)

Either way, writing \({1}_j\) (respectively \({0}_j\)) for the configuration \((1, \ldots , 1)\) (respectively \((0, \ldots , 0)\)) of length \(j\),

$$\begin{aligned} \mathbb {E}_{\pi _e} (\Delta _{e,j} \tau _z)^2 \le \frac{1}{2^{j-1}} \mathbb {E}_{\pi _{e,\ge j}} \left[ \min \{D_{z,e},T_{e,j}(1_{j-1},1)\}^2 \mathbf {1}_{\{T_{e,j}(0_{j-1},0) < D_{z,e}\}}\right] \mathbf {1}_{\{I<D_{z,e}\}}. \end{aligned}$$
(6.14)

Note that \(\min \{D_{z,e},T_{e,j}(\varvec{1}_{j-1},1)^2\}\) is an increasing function of \(\omega _{e,\ge j}\) (with all other variables fixed), whereas \(\mathbf {1}_{\{T_{e,j}(\varvec{0}_{j-1},0)< D_{z,e}\}}\) is decreasing (here we use monotonicity of \(T_e\)). Therefore we can apply the Harris-FKG inequality [8, Theorem 2.15] and sum over \(j\) for the upper bound

$$\begin{aligned} \mathbb {E}_{\pi _e} \sum _{j=1}^\infty (\Delta _{e,j} \tau _z)^2&\le \sum _{j=1}^\infty \frac{1}{2^{j-1}} \left[ \mathbb {E}_{\pi _{e,\ge j}} \min \{D_{z,e},T_{e,j}(\varvec{1}_{j-1},1)\}^2~ \pi _{e,\ge j}(T_{e,j}(\varvec{0}_{j-1},0)\right. \nonumber \\&\qquad \quad \qquad \left. < D_{z,e}) \right] \mathbf {1}_{\{I<D_{z,e}\}}. \end{aligned}$$
(6.15)

The goal is now to give a useful bound for this sum. To do this, we consider two types of values of \(j\). Note that \(F(D_{z,e}^-)>0\) and therefore for some \(j\), \(F(D_{z,e}^-) \ge 2^{-j}\). So define

$$\begin{aligned} J(D_{z,e}) = \min \{j \ge 2 : F(D_{z,e}^-) \ge 2^{-(j-1)}\}. \end{aligned}$$

Note that

$$\begin{aligned} 1-\log _2 F(D_{z,e}^-) \le J(D_{z,e}) \le 2-\log _2 F(D_{z,e}^-). \end{aligned}$$
(6.16)

We will estimate the term \(\pi _{e,\ge j}(T_{e,j}(\varvec{0}_{j-1},0) < D_{z,e})\) only when \(j < J(D_{z,e})\). By definition, it is

$$\begin{aligned}&\left( \prod _{k \ge j} \pi _{e,k}\right) \left( \left\{ \omega _e : T_e(0,\ldots , 0,\omega _{e,j+1}, \ldots ) < D_{z,e}\right\} \right) \\&\qquad = \pi _e (\{\omega _e : T_e(0, \ldots , 0, \omega _{e,j+1}, \ldots ) < D_{z,e}\}). \end{aligned}$$

The event in \(\varOmega _e\) listed on the right depends only on \(\omega _{e,k}\) for \(k > j\), so it is independent (under \(\pi _e\)) of the state of the first \(j\) coordinates. Thus the above equals

$$\begin{aligned}&2^j \pi _e (T_e(0, \ldots , 0, \omega _{e,j+1}, \ldots ) < D_{z,e},~ \omega _{e,1}, \ldots , \omega _{e,j} = 0)\\&\qquad \qquad \qquad \le 2^j \pi (T_e(\omega _e) < D_{z,e}) = 2^j F(D_{z,e}^-). \end{aligned}$$

Using this inequality for \(j <J(D_{z,e})\), (6.15) becomes

$$\begin{aligned} \mathbb {E}_{\pi _e} \sum _{j=1}^\infty (\Delta _{e,j} \tau _z)^2&\le 2F(D_{z,e}^-) \mathbb {E}_{\pi _{e,\ge 1}}T_{e,1}(1)^2 \mathbf {1}_{\{I<D_{e,z}\}} + 2F(D_{z,e}^-) \sum _{j=2}^{J(D_{z,e})-1} D_{z,e}^2 \mathbf {1}_{\{I<D_{z,e}\}} \end{aligned}$$
(6.17)
$$\begin{aligned}&\quad + \sum _{j=J(D_{z,e})}^\infty \frac{1}{2^{j-1}} \left[ \mathbb {E}_{\pi _{e,\ge j}} \min \{D_{z,e},T_{e,j}(\varvec{1}_{j-1},1)\}^2 \right] \mathbf {1}_{\{I<D_{z,e}\}} . \end{aligned}$$
(6.18)

The second term on the right of (6.17) is bounded by noting that when this sum is nonempty (that is, \(J(D_{z,e})>2\)), it follows that \(F(D_{z,e}^-) <1/2\) and so \(D_{z,e} \le a_{1,1}\). Using this with (6.16) we obtain

$$\begin{aligned} 2F(D_{z,e}^-) \sum _{j=2}^{J(D_{z,e})-1} D_{z,e}^2 \mathbf {1}_{\{I<D_{z,e}\}} \le 2F(D_{z,e}^-)(1-\log _2 F(D_{z,e}^-)) a_{1,1}^2 \mathbf {1}_{\{I<D_{z,e}\}}. \end{aligned}$$
(6.19)

We next bound \(\mathbb {E}_{\pi _{e,\ge j}} T_{e,j}(\varvec{1}_{j-1},1)^2\). Because \(T_{e,j}(\varvec{1}_{j-1},1)\) only depends on \(\omega _e\) through \(\omega _{e,>j}\),

$$\begin{aligned} \mathbb {E}_{\pi _e} T_{e,j}(\varvec{1}_{j-1},1)^2 = 2^j \mathbb {E}_{\pi _e} T_{e,j}(\varvec{1}_{j-1},1)^2 \mathbf {1}_{\{\omega _{e,\le j} = \varvec{1}_j\}} = 2^j \mathbb {E}_{\pi _e} T_e^2 \mathbf {1}_{\{\omega _{e,\le j} = \varvec{1}_j\}}. \end{aligned}$$

Thus in (6.17),

$$\begin{aligned} 2F(D_{z,e}^-) \mathbb {E}_{\pi _{e,\ge 1}} T_{e,1}(1)^2 \mathbf {1}_{\{I<D_{z,e}\}} \le 4F(D_{z,e}^-) \mathbb {E}_\mu t_e^2 \mathbf {1}_{\{I<D_{z,e}\}} \end{aligned}$$
(6.20)

and

$$\begin{aligned} (6.18) \le 2\sum _{j=J(D_{z,e})}^\infty \left[ \mathbb {E}_{\pi _e} \min \{D_{z,e}, T_e\}^2 \mathbf {1}_{\{\omega _{e,\le j} = \varvec{1}_j\}}\right] \mathbf {1}_{\{I < D_{z,e}\}} \end{aligned}$$

We now consider two cases. If \(D_{z,e} \le a_{1,1}\) then we use (6.16) to obtain the upper bound

$$\begin{aligned} (6.18) \le 2a_{1,1}^2 \sum _{j=J(D_{z,e})}^\infty \pi _e(\omega _{e,\le j} = \varvec{1}_j) \mathbf {1}_{\{I < D_{z,e}\}}&= 2a_{1,1}^2 \sum _{j=J(D_{z,e})}^\infty 2^{-j} \mathbf {1}_{\{I<D_{z,e}\}} \\&\le 4a_{1,1}^2 2^{-J(D_{z,e})} \mathbf {1}_{\{I<D_{z,e}\}} \\&\le 2 a_{1,1}^2 F(D_{z,e}^-) \mathbf {1}_{\{I<D_{z,e}\}}. \end{aligned}$$

On the other hand, if \(D_{z,e}>a_{1,1}\) then we use the bound

$$\begin{aligned} (6.18) \le 2 \left[ \mathbb {E}_{\pi _e} T_e^2 N\right] \mathbf {1}_{\{I<D_{z,e}\}}, \text { where } N = \max \{j \ge 1 : \omega _{e,\le j} = \varvec{1}_j\}. \end{aligned}$$

This is bounded by the variational characterization of entropy, Proposition 2.2. The expectation is no larger than

$$\begin{aligned} 2~Ent_\mu t_e^2 + 2 \mathbb {E}_\mu t_e^2 \log \mathbb {E}_{\pi _e} e^{N/2} . \end{aligned}$$

Because \(N\) has a geometric distribution, this is bounded by \(\mathbf {C}_{29}\) independently of \(e\). As \(D_{z,e}>a_{1,1}\), one has \(F(D_{z,e}^-) \ge 1/2\) and so we obtain

$$\begin{aligned} (6.18) \le 4\mathbf {C}_{29}F(D_{z,e}^-) \mathbf {1}_{\{I<D_{z,e}\}} . \end{aligned}$$

Combined with the case \(D_{z,e} \le a_{1,1}\), our final bound is

$$\begin{aligned} (6.18) \le (4\mathbf {C}_{29}+2a_{1,1}^2) F(D_{z,e}^-) \mathbf {1}_{\{I<D_{z,e}\}} . \end{aligned}$$
(6.21)

Putting together the pieces, (6.20) with (6.19) and (6.21),

$$\begin{aligned} \mathbb {E}_{\pi _e}\sum _{j=1}^\infty (\Delta _{e,j} \tau _z)^2 \le \mathbf {C}_{30} F(D_{z,e}^-)\mathbf {1}_{\{I<D_{z,e}\}} - \mathbf {C}_{31} F(D_{z,e}^-) \log F(D_{z,e}^-) \mathbf {1}_{\{I<D_{z,e}\}} .\nonumber \\ \end{aligned}$$
(6.22)

To bound terms of the second form we use a lemma.

Lemma 6.5

For any \(y>I\), we have

$$\begin{aligned} -F(y^-)\log F(y^-)\le -\int _{[I,y)}\log F(a)\,\mu (\mathrm {d}a). \end{aligned}$$
(6.23)

Proof

Let \(\epsilon >0\). The function \(\log F(x)\) is increasing on \((I, \infty )\). The usual Lebesgue construction gives a measure \(\nu \) on \((I,\infty )\) such that

$$\begin{aligned} \nu (a,b] = \log F(b)-\log F(a)\ge 0 \end{aligned}$$

for \(a,b\in (I,\infty )\). Fix \(x\in (I+\epsilon ,\infty )\), and consider the square

$$\begin{aligned} \square = (I+\epsilon , x]\times (I+\epsilon , x]. \end{aligned}$$

It has two parts:

$$\begin{aligned} \{(a,b):I+\epsilon < a<b \le x\}, \end{aligned}$$
(6.24)
$$\begin{aligned} \{(a,b):I+\epsilon < b\le a \le x\} . \end{aligned}$$
(6.25)

Thus,

$$\begin{aligned} (\mu \times \nu )(\square ) = \int \int _{(6.24)} (\mu \times \nu )(\mathrm {d}a\mathrm {d}b)+ \int \int _{(6.25)} (\mu \times \nu )(\mathrm {d}a\mathrm {d}b). \end{aligned}$$

By Fubini’s theorem, the double integrals may be computed as iterated integrals

$$\begin{aligned} \int \int _{(6.24)} (\mu \times \nu )(\mathrm {d}a\mathrm {d}b)&=\int _{(I+\epsilon ,x]} \mu ((I+\epsilon ,b))\nu (\mathrm {d}b)\nonumber \\&=\int _{(I+\epsilon ,x]}(F(b^-)-F(I+\epsilon ))\log F(\mathrm {d}b) \end{aligned}$$
(6.26)
$$\begin{aligned} \int \int _{(6.25)} (\mu \times \nu )(\mathrm {d}a\mathrm {d}b)&=\int _{(I+\epsilon ,x]} \nu ((I+\epsilon ,a])\mu (\mathrm {d}a)\nonumber \\&= \int _{(I+\epsilon ,x]}(\log F(a)-\log F(I+\epsilon )) F(\mathrm {d}a) . \end{aligned}$$
(6.27)

By definition of the product measure,

$$\begin{aligned} (\mu \times \nu )(\square ) = (F(x)-F(I+\epsilon ))\cdot (\log F(x)-\log F(I+\epsilon )). \end{aligned}$$

This gives the equality:

$$\begin{aligned} (F(x)&-F(I+\epsilon ))\cdot (\log F(x)-\log F(I+\epsilon )) = \int _{(I+\epsilon ,x]}F(b^-)\log F(\mathrm {d}b)\\&\quad +\int _{(I+\epsilon ,x]}\log F(a)F(\mathrm {d}a)-F(I+\epsilon )(\log F(x)-\log F(I+\epsilon )) \\&\quad -\log F(I+\epsilon ) (F(x)-F(I+\epsilon )). \end{aligned}$$

After performing cancellations, we obtain

$$\begin{aligned} F(x)\log F(x)-F(I+\epsilon )\log F(I+\epsilon )&= \int _{(I+\epsilon ,x]}F(b^-)\log F(\mathrm {d}b)\nonumber \\&\quad +\int _{(I+\epsilon ,x]}\log F(a)F(\mathrm {d}a).\qquad \end{aligned}$$
(6.28)

Since \(F(b^-)\ge 0\), this implies the estimate

$$\begin{aligned} -\int _{(I+\epsilon ,x]} \log F(a)\,\mu (\mathrm {d}a)-F(I+\epsilon )\log F(I+\epsilon ) \ge -F(x)\log F(x). \end{aligned}$$

Taking \(\epsilon \downarrow 0\) and using the right continuity of \(F\),

$$\begin{aligned} -\int _{(I,x]}\log F(a)\,\mu (\mathrm {d}a) -F(I)\log F(I) \ge -F(x)\log F(x), \end{aligned}$$

where the second term is interpreted as \(0\) if \(F(I)=0\). Since \(F(I)=\mu (\{I\})\),

$$\begin{aligned} -F(x)\log F(x) \le - \int _{[I,x]} \log F(a) \mu (\text {d}a). \end{aligned}$$

Taking \(x \uparrow y\), (6.23) is proved. \(\square \)

Apply the last lemma in (6.22) with \(y=D_{z,e}\):

$$\begin{aligned} \sum _e \mathbb {E}_\pi \sum _{j=1}^\infty (\Delta _{e,j} \tau _z)^2&\le \mathbf {C}_{32} \sum _e \mathbb {E}_{e^c} \int _{[I,D_{z,e})} (1-\log F(a))~\mu (\text {d}a) \\&= \mathbf {C}_{32} \mathbb {E} \sum _e (1-\log F(t_e)) \mathbf {1}_{\{I\le t_e < D_{z,e}\}}. \end{aligned}$$

By Lemma 5.2, if \(t_e < D_{z,e}\) then \(e\) is in \(Geo(z,z+x)\), so this is bounded above by

$$\begin{aligned} \mathbf {C}_{32} \mathbb {E} \sum _e (1- \log F(t_e)) \mathbf {1}_{\{e \in Geo(z,z+x)\}}. \end{aligned}$$

Translating back from \(z\) to \(0\) and putting this in (6.7) proves the proposition.\(\square \)

6.2.3 Lattice animals

To bound the right side of the inequality in Proposition 6.4 we need finer control than what is given by Kesten’s geodesic length estimates, due to the possible singularity of \(\log F(t_e)\) as \(t_e \downarrow I\). The idea will be that very few edges \(e\) on a geodesic have \(t_e\) close to \(I\). To bound the precise number, we give the main result of this section:

Theorem 6.6

Assume (1.7) and \(\mathbb {E} Y^\alpha <\infty \) for some \(\alpha >1\), where \(Y\) is the minimum of \(2d\) i.i.d. variables distributed as \(t_e\). There exists \(\mathbf {C}_{33}\) such that for all \(x \in \mathbb {Z}^d\) and any Borel set \(B \subset \mathbb {R}\),

$$\begin{aligned} \mathbb {E} \#\{e \in Geo(0,x) : t_e \in B\} \le \mathbf {C}_{33}\Vert x\Vert _1 \mu (B)^{\frac{\alpha -1}{\alpha d}}. \end{aligned}$$

The proof will require an excursion into the theory of greedy lattice animals. We say that a finite set of vertices \(\alpha \subseteq \mathbb {Z}^d\) is a lattice animal if it is connected (under graph connectedness on \(\mathbb {Z}^d\)). One fundamental result on lattice animals is the following, taken from [11, Lemma 1], which describes how a lattice animal may be covered by boxes. We set the notation \(B(l) = [-l,l]^d.\)

Lemma 6.7

(Cox–Gandolfi–Griffin–Kesten) Let \(\alpha \) be a lattice animal with \(0 \in \alpha \) and \(\# \alpha = n,\) and let \(1 \le l \le n.\) There exists a sequence \(x_0, x_1, \ldots , x_r \in \mathbb {Z}^d\), where \(r = \lfloor 2n / l \rfloor \), such that \(x_0 = 0\),

$$\begin{aligned} \alpha \subseteq \bigcup _{i=0}^r (lx_i + B(2l)), \end{aligned}$$
(6.29)

and

$$\begin{aligned} \Vert x_{i+1} - x_i\Vert _{\infty } \le 1, \quad 0\le i \le r-1. \end{aligned}$$

We will use the above theorem in a similar setting to the original model for which it is proved. Let \(\Xi _n\) denote the set of all self-avoiding paths

$$\begin{aligned} \gamma = (0 = v_0, e_1, v_1, \ldots , e_n, v_n) \end{aligned}$$

which begin at the origin and contain \(n\) edges. Denote by \(V(\gamma )\) the vertex set of \(\gamma .\) Assume that we have an edge-indexed set of i.i.d. variables \(\{X_e\}_e\), where \(X_e = 1\) with probability \(p\) and \(0\) otherwise, and denote the joint distribution of \(\{X_e\}\) by \(\mathbb {P}_p\); we denote expectation under this measure by \(\mathbb {E}_p.\) If \(\gamma \in \Xi _ n\) for some \(n\), we define \(X(\gamma ) = \sum _{e \in \gamma } X_e.\) Last, let

$$\begin{aligned} N_n := \max _{\gamma \in \Xi _n} X(\gamma ). \end{aligned}$$
(6.30)

The following lemma (and the proof thereof) is an adaptation of Martin’s [24, Prop. 2.2] extension of a theorem of Lee [22].

Lemma 6.8

There is a constant \(C_d < \infty \) depending only on the dimension \(d\) such that, for all \(p \in (0,1]\) and all \(n \in \mathbb {N},\)

$$\begin{aligned} \frac{\mathbb {E}_p N_n}{n p^{1/d}} < C_d. \end{aligned}$$
(6.31)

Proof

Let \(p\in (0,1]\) be arbitrary. We first consider the case that \(np^{1/d} \le 1.\) In this case, we have

$$\begin{aligned} \frac{\mathbb {E}_p N_n}{n p^{1/d}} \le \frac{1}{n p^{1/d}}\sum _{e \in [-n,n]^d} \mathbb {E}_pX_e \le \frac{2d(2 n + 1)^d p}{n p^{1/d}} \le 2d(3^d) (p^{1/d}n)^{d-1} \le 3^{d+1}d. \end{aligned}$$

In the case that \(n p^{1/d} > 1,\) we set \( l = \lceil p^{-1/d} \rceil \). Note that for any \(\gamma \in \Xi _n,\) \(V(\gamma )\) is a lattice animal with \(n + 1\) vertices. In particular, it can by covered using the results of Theorem 6.7. So for any \(s \ge 0,\)

$$\begin{aligned} \mathbb {P}_p \left( \frac{N_n}{n p^{1/d}} \ge s\right)&= \mathbb {P}_p \left( \max _{\gamma \in \Xi _n} X(\gamma ) \ge n p^{1/d} s \right) \nonumber \\&\le \mathbb {P}_p \left( \max _{x_0, \ldots , x_r} \sum _{\begin{array}{c} e = \{ x, y \} \\ x,y \in \cup _{i=0}^r (lx_i + B(2l)) \end{array}} X_e \ge n p^{1/d} s \right) \nonumber \\&\le \sum _{x_0, \ldots , x_r} \mathbb {P}_p \left( \sum _{\begin{array}{c} e = \{ x, y \} \\ x,y \in \cup _{i=0}^r (lx_i + B(2l)) \end{array}} X_e \ge n p^{1/d} s \right) , \end{aligned}$$
(6.32)

where the outer sum is over all connected subsets of \(\mathbb {Z}^d\) of cardinality \(r+1 = 1 + \lfloor 2(n+1)/l \rfloor \le 5 n p^{1/d}\) which contain the origin.

The expression in (6.32) is bounded above by

$$\begin{aligned}&\sum _{x_0, \ldots , x_r} \exp (-np^{1/d} s) \mathbb {E}_p \exp \left( \sum _{\begin{array}{c} e = \{ x, y \} \\ x,y \in \cup _{i=0}^r (lx_i + B(2l)) \end{array}} X_e \right) \nonumber \\&\quad \le \sum _{x_0, \ldots , x_r} \exp (-np^{1/d} s) \left[ \mathbb {E}_p \exp (X_e)\right] ^{\# \{e = \{ x, y \}: x,y \in \cup _{i=0}^r (lx_i + B(2l))\}}. \end{aligned}$$
(6.33)

Now, note that

  • \(\mathbb {E}_p \exp (X_e) = 1 - p + p\mathrm {e};\)

  • The number of vertices in \(B(2l)\) is \((4l + 1)^d,\) so

    $$\begin{aligned} \# \{e = \{ x, y \} : x,y \in \cup _0^r (lx_i + B(2l))\} \le (r+1) (2d)(4l+1)^d \le \mathbf {C}_{34}(d) n p^{1/d-1}. \end{aligned}$$
  • The number of terms in the sum (6.33) is at most \(3^{d(r+1)} \le 3^{5dnp^{1/d}}.\)

Putting the above into (6.33), we have

$$\begin{aligned} \mathbb {P}_p\left( \frac{N_n}{n p^{1/d}} \ge s\right)&\le \exp (-n p^{1/d} s) 3^{5 d n p^{1/d}} \left[ 1 - p + p \mathrm {e}\right] ^{\mathbf {C}_{34}(d) n p^{1/d - 1}}\nonumber \\&= \exp (-n p^{1/d} s) 3^{5 d n p^{1/d}} \left( \left[ 1 - p + p \mathrm {e}\right] ^{1/p}\right) ^{\mathbf {C}_{34}(d) n p^{1/d}}\nonumber \\&\le \exp (-n p^{1/d} s) 3^{5 d n p^{1/d}} \left[ \mathrm {e}^{\mathrm {e}-1}\right] ^{\mathbf {C}_{34}(d) n p^{1/d }}\nonumber \\&{=:} \exp (-np^{1/d} s + \mathbf {C}_{35} n p^{1/d}) , \end{aligned}$$
(6.34)

where \(\mathbf {C}_{35} = \mathbf {C}_{35}(d)\) again does not depend on \(p\) or \(n\). Then we have, for \(np^{1/d} > 1,\)

$$\begin{aligned} \mathbb {E}_p\left( \frac{N_n}{n p^{1/d}}\right) \le \mathbf {C}_{35} + \mathbb {E}_p \left[ \frac{N_n}{n p^{1/d}} - \mathbf {C}_{35} \right] _+&= \mathbf {C}_{35} + \int _{\mathbf {C}_{35}}^{\infty } \mathbb {P}_p\left( \frac{N_n}{np^{1/d}} \ge s \right) \mathrm {d} s\\&\le \mathbf {C}_{35} + \int _{\mathbf {C}_{35}}^{\infty } \exp \left( - np^{1/d}(s-\mathbf {C}_{35})\right) \mathrm {d} s\\&\le \mathbf {C}_{35} + \int _{\mathbf {C}_{35}}^{\infty } \exp \left( -(s-\mathbf {C}_{35})\right) \mathrm {d} s \le \mathbf {C}_{36} \end{aligned}$$

for some \(\mathbf {C}_{36} = \mathbf {C}_{36}(d)\). \(\square \)

We are now ready to prove the theorem.

Proof of Theorem 6.6

Consider any deterministic ordering of all finite self-avoiding lattice paths and denote by \(\pi (x,y)\) the first geodesic from \(x\) to \(y\) in this ordering. Writing \(Y_B(0,x)\) for the number of edges in \(\pi (x,y)\) with weight in \(B\), note that it suffices to give the bound for \(\mathbb {E} Y_B(0,x)\). Define a set of edge weights \(X_e\) as a function of \(t_e\):

$$\begin{aligned} X_e = {\left\{ \begin{array}{ll} 1 &{} \text { if } t_{e} \in B \\ 0 &{} \text { otherwise} \end{array}\right. }\end{aligned}$$

and build the random variables \(N_n\) for these weights as in (6.30).

On the event \(\{\# \pi (0,x) \le i \},\) we have \(Y_B(0,x) \le N_i\). Therefore, for all \(x \in \mathbb {Z}^d\) and \(\kappa \in \mathbb {N},\)

$$\begin{aligned} \mathbb {E} Y_B(0,x)&\le \mathbb {E} N_{\kappa \Vert x\Vert _1} + \mathbb {E}\left[ \# \pi (0,x) \mathbf {1}_{\{\#\pi (0,x)>\kappa \Vert x\Vert _1\}}\right] \\&= \mathbb {E} N_{\kappa \Vert x\Vert _1} + \int _{\kappa \Vert x\Vert _1}^\infty \mathbb {P}\left( \#\pi (0,x) > s\right) \mathrm {d} s\\&\le C_d \kappa \Vert x\Vert _1 \mu (B)^{1/d} + \int _{\kappa \Vert x\Vert _1}^\infty \mathbb {P}\left( \#\pi (0,x) > s\right) \mathrm {d} s. \end{aligned}$$

To bound the integral above, we use the technique of Kesten (see Eq. (2.26)–(2.27) in [21]). For \(b, j > 0,\) denote by \(D(j,b,x)\) the event that there exists a self-avoiding path \(r\) starting at the origin of at least \(j \Vert x\Vert _1\) steps but \(\tau (r) < j b \Vert x\Vert _1.\) Then for any \(b > 0,\)

$$\begin{aligned} \mathbb {P}\left( \#\pi (0,x) > j\Vert x\Vert _1\right)&\le \mathbb {P}\left( \tau (0,x) \ge bj\Vert x\Vert _1\right) + \mathbb {P}(D(j,b,x)). \end{aligned}$$
(6.35)

By our assumption \(\mathbb {E} Y^\alpha <\infty \), [10, Lemma 3.1] implies that there exists \(\mathbf {C}_{37}\) such that for all \(x\), \(\mathbb {E} \tau (0,x)^\alpha \le \mathbf {C}_{37} \Vert x\Vert _1^\alpha \). Thus for arbitrary \(x \in \mathbb {Z}^d,\)

$$\begin{aligned} \mathbb {P}\left( \tau (0,x) \ge b j \Vert x\Vert _1\right) \le \mathbf {C}_{37} / (b j)^\alpha . \end{aligned}$$

Due to assumption (1.7), we may use Theorem 1.5 to see that, for \(b\) smaller than some \(b_0>0\) (which depends on \(d\) and \(\mu \)), the probability of \(D(j,b,x)\) is bounded above uniformly in \(j\) and \(x\) by \(\exp (-\mathbf {C}_{38} j \Vert x\Vert _1)\). Inserting this bound into (6.35), we see that for \(b\) small enough,

$$\begin{aligned} \mathbb {P}\left( \#\pi (0,x) > j\Vert x\Vert _1\right) \le \frac{\mathbf {C}_{37}}{(b j)^\alpha } + \exp (- \mathbf {C}_{38} j\Vert x\Vert _1). \end{aligned}$$

In particular, setting \(r = s/\Vert x\Vert _1,\)

$$\begin{aligned} \mathbb {E} Y_B(0,x)&\le C_d \kappa \Vert x\Vert _1 \mu (B)^{1/d} + \Vert x\Vert _1\int _{\kappa }^{\infty }\left( \frac{\mathbf {C}_{37}}{(br)^\alpha } + \exp (- \mathbf {C}_{38} r \Vert x\Vert _1)\right) \mathrm {d} r\\&\le C_d \kappa \Vert x\Vert _1 \mu (B)^{1/d} + \frac{\mathbf {C}_{39} \Vert x\Vert _1}{\kappa ^{\alpha -1}} \end{aligned}$$

for some constant \(\mathbf {C}_{39}.\) Choosing \(\kappa = \lceil \mu (B)^{-1/(\alpha d)}\rceil \) completes the proof.\(\square \)

6.2.4 Finishing the proof of Theorem 6.2

We use Theorem 6.6, with a dyadic partition of \([I,\infty )\): let

$$\begin{aligned} x_0 = \infty \quad \text { and }\quad x_n = \min \{x : F(x) \ge 2^{-n}\}\quad \text { for } n \in \mathbb {N}. \end{aligned}$$

Note that for any edge \(e\), \(t_e\) almost surely lies in one of the intervals \([x_i,x_{i-1})\) for \(i \ge 1\). This is clear if \(I<t_e\). Otherwise we must have \(\mu (\{I\})>0\) and we simply take \(i\) to be minimal such that \(2^{-i} \le \mu (\{I\})\).

Now the right side of the inequality in Proposition 6.4 can be rewritten as

$$\begin{aligned} \mathbf {C}_{28} \sum _{i=1}^\infty \sum _e \mathbb {E}&\left[ (1-\log F(t_e)) \mathbf {1}_{\{e \in Geo(z,z+x)\}} \mathbf {1}_{\{t_e \in [x_i,x_{i-1})\}}\right] \\&\le \mathbf {C}_{28} \sum _{i=1}^\infty (1-\log F(x_i)) \mathbb {E} \#\{e \in Geo(z,z+x) : t_e \in [I,x_{i-1})\}. \end{aligned}$$

By Theorem 6.6 with \(\alpha =2\), this is bounded by

$$\begin{aligned}&\mathbf {C}_{28}\mathbf {C}_{33} \Vert x\Vert _1 \sum _{i=1}^\infty (1-\log F(x_i)) F(x_{i-1}^-)^{1/(2d)}\\&\quad \le \mathbf {C}_{28}\mathbf {C}_{33} \Vert x\Vert _1 \sum _{i=1}^\infty \frac{1+i}{2^{(i-1)/(2d)}} \le \mathbf {C}_{27} \Vert x\Vert _1. \end{aligned}$$

6.3 Proof of Theorem 1.1

For \(x \in \mathbb {Z}^d\), if \({{\mathrm{Var}}}\,F_m \le \Vert x\Vert _1^{7/8}\) then by Proposition 3.1, we are done. Otherwise, under assumptions (1.6) and (1.7) we can use (6.1) to find for some \(\mathbf {C}_3\)

$$\begin{aligned} {{\mathrm{Var}}}~\tau (0,x) \le \mathbf {C}_3\Vert x\Vert _1^{3/4} + \left[ \log \left[ \Vert x\Vert _1^{1/8}\right] \right] ^{-1} \sum _{k=1}^\infty Ent(V_k^2). \end{aligned}$$

By Theorem 6.2, \(\sum _{k=1}^\infty Ent(V_k^2) \le \mathbf {C}_{27}\Vert x\Vert _1\), so

$$\begin{aligned} {{\mathrm{Var}}}~\tau (0,x) \le \mathbf {C}_3\Vert x\Vert _1^{3/4} + \frac{8\mathbf {C}_{27} \Vert x\Vert _1}{\log \Vert x\Vert _1} \le \frac{\mathbf {C}_{40} \Vert x\Vert _1}{\log \Vert x\Vert _1}. \end{aligned}$$