This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

footnotetext: Submitted August 19, 2021. Revised January 23, 2023. Accepted February 9, 2023.

Geodesic length and shifted weights
in first-passage percolation

Arjun Krishnan Arjun Krishnan
University of Rochester
Mathematics Department
Hylan 817
Rochester, NY 14627
USA.
[email protected] https://people.math.rochester.edu/faculty/akrish11/
Firas Rassoul-Agha Firas Rassoul-Agha
University of Utah
Mathematics Department
155S 1400E
Salt Lake City, UT 84112
USA.
[email protected] https://www.math.utah.edu/ firas
 and  Timo Seppäläinen Timo Seppäläinen
University of Wisconsin-Madison
Mathematics Department
Van Vleck Hall
480 Lincoln Dr.
Madison WI 53706-1388
USA.
[email protected] https://people.math.wisc.edu/ seppalai/
Abstract.

We study first-passage percolation through related optimization problems over paths of restricted length. The path length variable is in duality with a shift of the weights. This puts into a convex duality framework old observations about the convergence of the normalized Euclidean length of geodesics due to Hammersley and Welsh, Smythe and Wierman, and Kesten, and leads to new results about geodesic length and the regularity of the shape function as a function of the weight shift. For points far enough away from the origin, the ratio of the geodesic length and the 1\ell^{1} distance to the endpoint is uniformly bounded away from one. The shape function is a strictly concave function of the weight shift. Atoms of the weight distribution generate singularities, that is, points of nondifferentiability, in this function. We generalize to all distributions, directions and dimensions an old singularity result of Steele and Zhang for the planar Bernoulli case. When the weight distribution has two or more atoms, a dense set of shifts produce singularities. The results come from a combination of the convex duality, the shape theorems of the different first-passage optimization problems, and modification arguments.

Key words and phrases:
Approximate geodesic, convex duality, first-passage percolation, geodesic, path length, shape function, weight shift
2020 Mathematics Subject Classification:
60K35, 60K37
A. Krishnan was partially supported by a Wiley Assistant Professorship at University of Utah, an AMS-Simons Travel Grant, and Simons Collaboration grant 638966.
F. Rassoul-Agha was partially supported by National Science Foundation grants DMS-1811090 and DMS-2054630.
T. Seppäläinen was partially supported by National Science Foundation grants DMS-1854619 and DMS-2152362 and by the Wisconsin Alumni Research Foundation.

1. Introduction

1.1. Stochastic growth models

Irregular and stochastic growth surrounds us, for example in tumors, bacterial colonies, infections, spread of fluid in a porous medium, and propagating flame fronts. These phenomena attract the attention of mathematicians, scientists and engineers in various disciplines. Simplified mathematical models of stochastic growth have been studied in probability theory for over half a century. This work has inspired some of the central innovations of modern probability, such as the subadditive ergodic theorem, and created new connections between probability and other parts of mathematics, such as representation theory, integrable systems, and partial differential equations.

A class of much-studied stochastic growth models possess a metric-like structure where growth progresses along paths that optimize an energy functional defined in terms of a random environment. Depending on whether the optimal path is chosen through minimization or maximization, these models are called first-passage percolation and last-passage percolation.

A variety of settings for first- and last-passage percolation are studied. The admissible paths can be general or they can be restricted to be directed along some spatial directions. The underlying space can be a graph, the continuum, or a mixture of the two. In the graph case, the environment is given by random weights attached to the vertices or the edges. The most typical choice of graph is the dd-dimensional integer lattice d\mathbb{Z}^{d}. The one-dimensional case usually reduces to classical probability so the real work begins from the planar case d=2d=2.

Much progress in the planar case has taken place over the past 25 years under the rubric Kardar-Parisi-Zhang universality. A universal planar continuum limit, the directed landscape, has recently been constructed [8]. It is expected to be the scaling limit of a wide class of planar first- and last-passage percolation models, but this remains conjectural at present. Evidence for the universality comes from proofs that certain special exactly solvable directed models converge to the directed landscape [9]. We refer the reader to articles [4, 6] and the monograph [2] for general introductions to the field.

Our paper studies first-passage percolation with undirected paths on the integer lattice in arbitrary dimension. This has proved to be, in a sense, the most challenging model, as no exactly solvable version has been discovered. A proof that this model lies in the KPZ class, while universally expected, appears well beyond reach in the current state of the field. Our results concern properties of the geodesics and the regularity of the limiting norm as we perturb the random weights by a common additive constant. We turn to discuss the background.

1.2. First-passage percolation and its limit shape

In first-passage percolation (FPP) a random pseudometric is defined on d\mathbb{Z}^{d} by Tx,y=infπeπt(e)T_{x,y}=\inf_{\pi}\sum_{e\hskip 0.55pt\in\hskip 0.55pt\pi}t(e) where the {t(e)}\{t(e)\} are nonnegative, independent and identically distributed (i.i.d.) random weights on the nearest-neighbor edges between vertices of d\mathbb{Z}^{d} and the infimum is over self-avoiding paths π\pi between the two points xx and yy. A minimizing path is called a geodesic between xx and yy. FPP was introduced by Hammersley and Welsh [12] in 1965 as a simplified model of fluid flow in an inhomogeneous medium. A precise technical definition of the model comes in Section 2.

The fundamental questions of FPP concern the behavior of the passage times Tx,yT_{x,y} and the geodesics as the distance between xx and yy grows. At the level of the law of large numbers, under suitable hypotheses, normalized passage times converge with probability one: n1T𝟎,xnμ(ξ)n^{-1}T_{\mathbf{0},x_{n}}\to\mu(\xi) as nn\to\infty, whenever n1xnξdn^{-1}x_{n}\to\xi\in\mathbb{R}^{d}. The special case μ(𝐞1)=limnn1T𝟎,n𝐞1\mu(\mathbf{e}_{1})=\lim_{n\to\infty}n^{-1}T_{\mathbf{0},n\mathbf{e}_{1}} of the limit is also called the time constant.

The limiting shape function μ\mu is a norm that characterizes the asymptotic shape of a large ball. Define the randomly growing ball in d\mathbb{R}^{d} for t0t\geq 0 by B(t)={xd:T𝟎,xt}B(t)=\{x\in\mathbb{R}^{d}:T_{\mathbf{0},\lfloor{x}\rfloor}\leq t\} where xd\lfloor{x}\rfloor\in\mathbb{Z}^{d} is obtained from xdx\in\mathbb{R}^{d} by taking integer parts coordinatewise. Under the right assumptions, as tt\to\infty the normalized ball t1B(t)t^{-1}B(t) converges to the unit ball ={ξd:μ(ξ)1}\mathcal{B}=\{\xi\in\mathbb{R}^{d}:\mu(\xi)\leq 1\} defined by the norm μ\mu.

The shape function μ\mu is not explicitly known in any nontrivial example. Soft properties such as convexity, continuity, positive homogeneity, and μ(ξ)>0\mu(\xi)>0 for ξ𝟎\xi\neq\mathbf{0} when zero-weight edges are subcritical, are readily established. But anything beyond that, such as strict convexity or differentiability, remain conjectural. The only counterexample to this state of affairs is the classic Durrett-Liggett [10] planar flat edge result, sharpened by Marchand [15], and then extended by Auffinger and Damron [1] to include differentiability at the boundary of the flat edge.

The FPP shape theorem occupies a venerable position as one of the fundamental results of the subject of random growth models and as an early motivator of subadditive ergodic theory. The reader is referred to the monograph [2] for a recent overview of the known results and open problems.

1.3. Differentiability and length of geodesics

The success of the shape theorem contrasts sharply with the situation of another natural limit question, namely the behavior of the normalized Euclidean length (number of edges) of a geodesic as one endpoint is taken to infinity. No useful subadditivity or other related property has been found. This issue has been addressed only a few times over the 55 years of FPP study and the results remain incomplete.

The fundamental observation due to Hammersley and Welsh is the connection between (i) the limit of the normalized length of the geodesic and (ii) the derivative of the shape function as a function of a weight shift. For hh\in\mathbb{R} let μ(h)(ξ)\mu^{(h)}(\xi) denote the shape function for the shifted weights {t(e)+h}\{t(e)+h\}. Let L¯𝟎,x(h)\underline{L}_{\hskip 0.9pt\mathbf{0},x}^{(h)} be the minimal Euclidean length of a geodesic from the origin to the point xx for the shifted weights {t(e)+h}\{t(e)+h\}. Then the important fact is that when n1xnξn^{-1}x_{n}\to\xi,

(1.1) limnn1L¯𝟎,xn(h)=μ(s)(ξ)s|s=h\lim_{n\to\infty}n^{-1}\underline{L}_{\hskip 0.9pt\mathbf{0},x_{n}}^{(h)}\;=\;\frac{\partial\mu^{(s)}(\xi)}{\partial s}\bigg{|}_{s=h}

provided the derivative at hh on the right-hand side exists.

The shape function μ(h)(ξ)\mu^{(h)}(\xi) is a concave function of hh and hence the derivative in (1.1) exists and the limit holds for all but countably many shifts hh. But since the time constant itself remains a mystery, not a single specific nontrivial case where this identity holds has been identified. The first results on the size of the set of exceptional hh at which the derivative on the right fails are proved in the present paper and summarized in Sections 1.5 and 1.6 below.

Here is a brief accounting of the history of (1.1).

Hammersley and Welsh (Theorem 8.2.3 in [12]) gave the first version of (1.1). It was proved for the time constant of planar FPP, so for d=2d=2 and ξ=𝐞1\xi=\mathbf{e}_{1}, and for the particular sequence xn=(n,0)x_{n}=(n,0). Their result applied to the geodesic of the so-called cylinder passage time from (0,0)(0,0) to (n,0)(n,0), and the mode of convergence in (1.1) was convergence in probability.

The limit (1.1) was improved in 1978 by Smythe and Wierman (Theorem 8.2 in [18]) and in 1980 by Kesten [13], in particular from convergence in probability to almost sure convergence. The ultimate version has recently been established by Bates (Theorem 1.25 in [3]): almost sure convergence in (1.1) without any moment assumptions on the weights, in all directions ξ\xi, provided the derivative on the right exists.

A handful of precise results related to (1.1) exist in specific situations defined by criticality in percolation. Let pcp_{c} denote the critical probability of Bernoulli bond percolation on d\mathbb{Z}^{d}. When (t(e)=0)pc\mathbb{P}(t(e)=0)\geq p_{c} the FPP problem becomes in a sense degenerate. Geodesics to far-away points can take advantage of long paths of zero-weight edges and the shape function μ\mu becomes identically zero.

Zhang [21] proved in 1995 that in the supercritical case defined by (t(e)=0)>pc\mathbb{P}(t(e)=0)>p_{c}, for ξ=𝐞1\xi=\mathbf{e}_{1} and h=0h=0, the limit on the left in (1.1) exists and equals a nonrandom constant. In the planar critical case, that is, d=2d=2, (t(e)=0)=1/2=pc\mathbb{P}(t(e)=0)=1/2=p_{c} and h=0h=0, Damron and Tang [7] proved that the left-hand side in (1.1) blows up in all directions ξ\xi.

In 2003 Steele and Zhang [19] proved the first, and before the present paper the only, precise result about the derivative in (1.1), valid for subcritical planar FPP with Bernoulli weights. When the distribution is (t(e)=0)=p=1(t(e)=1)\mathbb{P}(t(e)=0)=p=1-\mathbb{P}(t(e)=1), there exists δ>0\delta>0 such that, if 12δp<12\tfrac{1}{2}-\delta\leq p<\tfrac{1}{2}, d=2d=2 and ξ=𝐞1\xi=\mathbf{e}_{1}, then the derivative in (1.1) fails to exist at h=0h=0. Thus the Hammersley-Welsh differentiability criterion for the convergence of normalized geodesic length faces a limitation.

1.4. Duality of path length and weight shift

We move on to describe the contents of our paper. To investigate (1.1) and more broadly properties of geodesic length, we develop a convex duality between the weight shift hh and a parameter that captures the asymptotic length of a path. This puts the limit (1.1) into a convex-analytic framework. To account for the possibility of nondifferentiability in (1.1), we enlarge the class of paths considered from genuine geodesics to o(n)o(n)-approximate geodesics. These are paths whose endpoints are order nn apart and whose passage times are within o(n)o(n) of the optimal passage time. Through these we can capture the entire superdifferential of the shape function as a function of the shift hh.

To be able to work explicitly with the path-length parameter, we introduce a version of FPP that minimizes over paths with a given number of steps but drops the requirement that paths be self-avoiding (Section 2.3). A further useful variant of the restricted path length FPP process allows zero-length steps that do not increase the passage time. The shape functions gg and gog^{o} of these altered models are no longer positively homogeneous, but they turn out to be continuously differentiable along rays from the origin (Theorem 2.16).

The restricted path length shape functions gg and gog^{o} are connected with the FPP shape function μ\mu in several ways. A key fact is that gg and gog^{o} agree with μ\mu on certain subsets of d\mathbb{R}^{d} described by positively homogeneous functions that are connected with geodesic length (Theorems 2.11 and 2.16). Second, gg and gog^{o} generate μ\mu as the maximal positively homogeneous convex function dominated by gg and gog^{o} (Remark 2.15). Third, gg and gog^{o} contain the information for generating all the shifts μ(h)\mu^{(h)} through convex duality (Theorem 2.17 and Remark 2.18).

From this setting we derive two types of main results for FPP: results on the Euclidean length of geodesics and on the regularity of the shape function as a function of the weight shift, briefly summarized in the next two paragraphs. The proofs come through a combination of

  1. (i)

    versions of the van den Berg-Kesten modification arguments [20],

  2. (ii)

    the convex duality (Theorem 2.17), and

  3. (iii)

    a shape theorem for the altered FPP models (Theorem 2.9 and Theorem B.1 in Appendix B).

Our results are valid on d\mathbb{Z}^{d} in all dimensions d2d\geq 2, under the standard moment bound needed for the shape theorem and the assumption that the minimum of the edge weight t(e)t(e) has probability strictly below pcp_{c}.

1.5. Euclidean length of geodesics

One of our fundamental results is that with probability one, all geodesics from the origin to far enough lattice points xx have length at least (1+δ)|x|1(1+\delta)|x|_{1} for a fixed constant δ>0\delta>0 (Theorem 2.5). The equality in (1.1) between the limiting normalized length of the geodesic and the derivative of the shape function, which is conditional on the existence of these quantities, is generalized to an unconditional identity between the entire interval of the asymptotic normalized lengths of the o(n)o(n)-approximate geodesics and the superdifferential of the shape function as a function of the weight shift (Theorem 2.17). When the random weight t(e)t(e) has an atom at zero or at least two atoms that satisfy suitable linear relations with integer coefficients, there are multiple geodesics whose lengths vary on the same scale as the distance between the endpoints (Theorem 2.6). For any weight distribution with at least two atoms, this happens on a countable dense set of shifts (Theorem 2.7).

1.6. Regularity of the shape function as a function of the weight shift

A second suite of main results concerns the regularity of the shape function μ(h)(ξ)\mu^{(h)}(\xi) as a function of the weight shift hh, in a fixed spatial direction ξd{𝟎}\xi\in\mathbb{R}^{d}\hskip 0.7pt\setminus\{\mathbf{0}\}. This function is strictly concave in hh (Theorem 2.2). In the situations where the atoms of t(e)t(e) bring about geodesics whose asymptotic normalized lengths vary, the concave function hμ(h)(ξ)h\mapsto\mu^{(h)}(\xi) acquires points of nondifferentiability. In particular, there is a countable dense set of these singularities whenever the edge weight has two atoms (Theorems 2.6 and 2.7). We extend the Steele-Zhang nondifferentiability result [19] mentioned above to all dimensions, all directions ξ\xi, and all distributions with an atom at the origin. Furthermore, we disprove their conjecture that h=0h=0 is the only nondifferentiability point in the Bernoulli case (Remark 2.8).

1.7. Organization of the paper

Section 2 describes the models and the main results. Section 3 describes open problems that arise from this work.

The proofs are divided into four sections. Section 4 develops soft results about the relationships between the different shape functions and the Euclidean lengths of optimal paths. The main technical Sections 5 and 6 contain the modification arguments. The final Section 7 combines the results from Sections 4, 5 and 6 to prove the main theorems.

Four appendixes contain auxiliary results that rely on standard material. Appendix A extends the FPP shape function to weights that are allowed small negative values. Appendix B proves a shape theorem for the restricted path length versions of FPP. Appendix C contains the Peierls argument that sets the stage for the modification proofs. Appendix D presents a lemma about the subdifferentials of convex functions.

1.8. Further literature: convergence of empirical measures

We close this introduction with a mention of a significant recent extension to the differentiability approach to limits along geodesics, due to Bates [3]. By representing the weights as functions t(e)=τ(Ue)t(e)=\tau(U_{e}) of uniform random variables, one can consider perturbations t~(e)=τ(Ue)+ψ(Ue)\widetilde{t}(e)=\tau(U_{e})+\psi(U_{e}) of the weights and differentiate the shape function in directions ψ\psi in infinite dimensions. This way the limit in (1.1) can be upgraded to convergence of the empirical distribution of weights along a geodesic, again whenever the required derivative exists. This holds for various uncountable dense collections of weight distributions, exactly as (1.1) holds for an uncountable dense set of shifts hh.

These more general limit results continue to share the fundamental shortcoming of the limit in (1.1), namely, that no particular nontrivial case can be identified where the limit holds. If (t(e)=0)pc\mathbb{P}(t(e)=0)\geq p_{c} the empirical measure along a geodesic converges trivially to a pointmass at zero.

Finding extensions of our results to the general perturbations of [3] presents an interesting open problem.

1.9. Notation and conventions

Here is notation that the reader may wish quick access to. +={0,1,2,3,}\mathbb{Z}_{+}=\{0,1,2,3,\dotsc\}, ={1,2,3,}\mathbb{N}=\{1,2,3,\dotsc\}, and +=[0,)\mathbb{R}_{+}=[0,\infty). For nn\in\mathbb{N}, [n]={1,2,,n}[n]=\{1,2,\dotsc,n\}. Standard basis vectors in d\mathbb{R}^{d} are 𝐞1=(1,0,,0)\mathbf{e}_{1}=(1,0,\dotsc,0), 𝐞2=(0,1,0,,0),,𝐞d=(0,,0,1)\mathbf{e}_{2}=(0,1,0,\dotsc,0),\dotsc,\mathbf{e}_{d}=(0,\dotsc,0,1) and 𝟎\mathbf{0} is the origin of d\mathbb{R}^{d}. The 1\ell^{1} norm of x=(x1,,xd)dx=(x_{1},\dotsc,x_{d})\in\mathbb{R}^{d} is |x|1=i=1d|xi||x|_{1}=\sum_{i=1}^{d}|x_{i}|. Particular subsets of d\mathbb{R}^{d} that recur are ={±𝐞1,,±𝐞d}\mathcal{R}=\{\pm\mathbf{e}_{1},\dotsc,\pm\mathbf{e}_{d}\}, o={𝟎}\mathcal{R}^{o}=\mathcal{R}\cup\{\mathbf{0}\}, 𝒰=co={ξd:|ξ|11}\mathcal{U}=\operatorname{co}\mathcal{R}=\{\xi\in\mathbb{R}^{d}:|\xi|_{1}\leq 1\}, and the topological interior int𝒰\operatorname{int}\mathcal{U}.

A finite or infinite path or sequence is denoted by xm:n=(xm,,xn)x_{m\hskip 0.6pt{\boldsymbol{:}}\hskip 0.3ptn}=(x_{m},\dotsc,x_{n}) for mn-\infty\leq m\leq n\leq\infty. Other notation for lattice paths are x\scaleobj0.5∙x_{\raisebox{0.5pt}{\scaleobj{0.5}{\bullet}}} and π\pi. The steps of a path are the nearest-neighbor edges ei={xi1,xi}e_{i}=\{x_{i-1},x_{i}\}. A finite path xm:nx_{m\hskip 0.6pt{\boldsymbol{:}}\hskip 0.3ptn} that satisfies |xnxm|1=nm|x_{n}-x_{m}|_{1}=n-m is called an 1\ell^{1}-path.

A positively homogeneous function ff satisfies f(cx)=cf(x)f(cx)=cf(x) for c>0c>0 whenever both cxcx and xx are in the domain of ff [17, p. 30]. One-sided derivatives of a function defined around ss\in\mathbb{R} are defined by f(s+)=limh0h1[f(s+h)f(s)]f^{\prime}(s+)=\lim_{h\searrow 0}h^{-1}[f(s+h)-f(s)] and f(s)=limh0h1[f(s)f(sh)]f^{\prime}(s-)=\lim_{h\searrow 0}h^{-1}[f(s)-f(s-h)].

The diamond \diamond is a wild card for three superscripts 𝚎𝚖𝚙𝚝𝚢\langle{\tt empty}\rangle (no superscript at all), oo (zero steps allowed), and sa (self-avoiding) that distinguish different FPP processes with restricted path length.

A real number rr is an atom of the random edge weight t(e)t(e) if {t(e)=r}>0\mathbb{P}\{t(e)=r\}>0. M0=esssupt(e)M_{0}=\mathop{\mathrm{ess\,sup}}t(e) and r0=essinft(e)r_{0}=\mathop{\mathrm{ess\,inf}}t(e). Superscript (b)(b) on any quantity means that it is computed with weights shifted by bb: t(b)(e)=t(e)+bt^{(b)}(e)=t(e)+b.

The symbol \triangle marks the end of a numbered remark.

1.10. Acknowledgements

The authors would like to thank Michael Damron for sketching a route to prove the lower bound in Theorem 2.11(ii), and an anonymous referee for spotting a mistake in the proof of Theorem 2.17.

2. The models and the main results

2.1. Setting

Fix an arbitrary dimension d2d\geq 2. Let d={{x,y}:x,yd,|xy|1=1}{\mathcal{E}}_{d}=\{\{x,y\}:x,y\in\mathbb{Z}^{d},|x-y|_{1}=1\} denote the set of undirected nearest-neighbor edges between vertices of d\mathbb{Z}^{d}. (Ω,𝔖,)(\Omega,\mathfrak{S},\mathbb{P}) is the probability space of an environment ω=(t(e):ed)\omega=(t(e):e\in{\mathcal{E}}_{d}) such that the edge weights {t(e):ed}\{t(e):e\in{\mathcal{E}}_{d}\} are independent and identically distributed (i.i.d.) real-valued random variables. Translations {θx}xd\{\theta_{x}\}_{x\hskip 0.55pt\in\hskip 0.55pt\mathbb{Z}^{d}} act on Ω\Omega by (θxω){u,v}=t({x+u,x+v})(\theta_{x}\omega){\{u,v\}}=t(\{x+u,x+v\}) for a nearest-neighbor edge {u,v}\{u,v\}.

A nearest-neighbor path π=x0:n=(xi)i=0n\pi=x_{0\hskip 0.6pt{\boldsymbol{:}}\hskip 0.3ptn}=(x_{i})_{i=0}^{n} is any finite sequence of vertices x0,x1,,xndx_{0},x_{1},\dotsc,x_{n}\in\mathbb{Z}^{d} that satisfy |xi+1xi|1=1|x_{i+1}-x_{i}|_{1}=1 for each ii. The steps of π\pi are the nearest-neighbor edges ei={xi1,xi}e_{i}=\{x_{i-1},x_{i}\}. The Euclidean length |π||\pi| of π\pi is the number of edges, so in this case |π|=n|\pi|=n. Then we call π\pi an nn-path. The passage time of π\pi is the sum of the weights of its edges:

(2.1) T(π)=i=1nt(ei).T(\pi)=\sum_{i=1}^{n}t(e_{i}).

These definitions apply even if the path repeats vertices or edges, as will be allowed at times in the sequel. For notational consistency we also admit the zero-length path π=x0:0=(x0)\pi=x_{0\hskip 0.6pt{\boldsymbol{:}}\hskip 0.3pt0}=(x_{0}) that has no edges and has zero passage time and length: T(π)=|π|=0T(\pi)=|\pi|=0.

The main results are described next in three parts: results for standard FPP in Section 2.2, results for restricted path-length FPP in Section 2.3, including the connections between the two types of FPP, and finally in Section 2.4 the duality between weight shift and geodesic length.

2.2. Standard first-passage percolation

In standard first-passage percolation (FPP) the passage time between two points is defined as the minimal passage time over all self-avoiding paths. A path π=x0:n=(xi)i=0n\pi=x_{0\hskip 0.6pt{\boldsymbol{:}}\hskip 0.3ptn}=(x_{i})_{i=0}^{n} is self-avoiding if xixjx_{i}\neq x_{j} for all pairs iji\neq j. Let Πx,y sa\Pi^{\text{\rm{\,sa}}}_{x,y} denote the collection of all self-avoiding paths from xx to yy, of arbitrary but finite length. Define the passage time between xx and yy as

(2.2) Tx,y=infπΠx,y saT(π).T_{x,y}=\inf_{\pi\,\in\,\Pi^{\text{\rm{\,sa}}}_{x,y}}T(\pi).

This definition gives Tx,x=0T_{x,x}=0 because the only self-avoiding path from xx to xx is the zero-length path. A geodesic is a self-avoiding path π\pi that minimizes in (2.2).

When t(e)0t(e)\geq 0 the restriction to self-avoiding paths is superfluous in the definition of Tx,yT_{x,y}. Let pcp_{c} denote the critical probability of Bernoulli bond percolation on d\mathbb{Z}^{d}. A frequently used assumption in FPP is that zero-weight edges are subcritical:

(2.3) {t(e)=0}<pc.\mathbb{P}\{t(e)=0\}<p_{c}.

For nonnegative weights, the assumption (2.3) guarantees the existence of a geodesic (Prop. 4.4 in [2]).

For bb\in\mathbb{R}, define bb-shifted weights by

(2.4) ω(b)=(t(b)(e):ed)witht(b)(e)=t(e)+bfored.\omega^{(b)}=(t^{(b)}(e):e\in{\mathcal{E}}_{d})\quad\text{with}\quad t^{(b)}(e)=t(e)+b\quad\text{for}\ \ e\in{\mathcal{E}}_{d}.

All the quantities associated with weights ω(b)\omega^{(b)} acquire the superscript. For example, Tx,y(b)T^{(b)}_{x,y} is the passage time in (2.2) under weights ω(b)\omega^{(b)}. Let

(2.5) r0=-essinfωt(e)r_{0}=\mathbb{P}\text{-}\mathop{\mathrm{ess\,inf}}_{\omega}t(e)

denote the (essential) lower bound of the weights. So in particular, ω(r0)\omega^{(-r_{0})} is the weight configuration shifted so that the lower bound is at zero. Since we shift weights, most of the time we have to replace (2.3) with this assumption:

(2.6) {t(e)=r0}<pc.\displaystyle\mathbb{P}\{t(e)=r_{0}\}<p_{c}.

Let {ti}\{t_{i}\} denote i.i.d. copies of the edge weight t(e)t(e). The following moment assumption will be employed for various values of pp.

(2.7) 𝔼[(min{t1,,t2d})p]<.\mathbb{E}[\,(\min\{t_{1},\dotsc,t_{2d}\})^{p}\,]<\infty.

We record the Cox-Durrett shape theorem ([5], Thm. 2.17 in [2]), with a small extension to weights that can take negative values. This theorem is proved as Theorem A.1 in Appendix A.

Theorem 2.1.

Assume r00r_{0}\geq 0, (2.6), and the moment bound (2.7) with p=dp=d. Then there exists a constant ε0>0\varepsilon_{0}>0, determined by the dimension dd and the distribution of the shifted weights ω(r0)\omega^{(-r_{0})}, and a full-probability event Ω0\Omega_{0} such that the following statements hold. For each real b>r0ε0b>-r_{0}-\varepsilon_{0} there exists a continuous, convex, positively homogeneous shape function μ(b):d+\mu^{(b)}:\mathbb{R}^{d}\to\mathbb{R}_{+} such that the limit

(2.8) μ(b)(ξ)=limnn1T𝟎,xn(b)\mu^{(b)}(\xi)=\lim_{n\to\infty}n^{-1}T^{(b)}_{\mathbf{0},x_{n}}

holds for each ωΩ0\omega\in\Omega_{0}, whenever {xn}d\{x_{n}\}\subset\mathbb{Z}^{d} satisfies xn/nξx_{n}/n\to\xi. We have μ(b)(𝟎)=0\mu^{(b)}(\mathbf{0})=0 and μ(b)(ξ)>0\mu^{(b)}(\xi)>0 for ξ𝟎\xi\neq\mathbf{0}.

If we require the shape function only for a single nonnegative weight distribution without the shifts, then (2.6) can be replaced with the weaker assumption (2.3), and we will occasionally do so. The shape function of unshifted weights is denoted by μ=μ(0)\mu=\mu^{(0)}.

To emphasize dependence on bb with ξ𝟎\xi\neq\mathbf{0} fixed, we write

(2.9) μξ(b)=μ(b)(ξ)for b>r0ε0.\mu_{\xi}(b)=\mu^{(b)}(\xi)\qquad\text{for }\ b>-r_{0}-\varepsilon_{0}.

Several of our main results concern the regularity of μξ\mu_{\xi} and its connections with geodesic length. The reason for allowing negative weights by extending the shift bb below r0-r_{0} is to enable us to talk about the regularity of μξ(b)\mu_{\xi}(b) at b=r0b=-r_{0}. Throughout this paper, ε0\varepsilon_{0} is the constant specified in Theorem 2.1.

Theorem 2.2.

Assume r00r_{0}\geq 0, (2.6), and the moment bound (2.7) with p=dp=d. Fix ξd{𝟎}\xi\in\mathbb{R}^{d}\hskip 0.7pt\setminus\{\mathbf{0}\}.

  1. (i)

    The function μξ\mu_{\xi} of (2.9) is a continuous, strictly increasing, concave function on the open interval (r0ε0,)(-r_{0}-\varepsilon_{0},\infty).

  2. (ii)

    Strict concavity holds on [r0,)[-r_{0},\infty)\hskip 0.7pt: μξ(a+)>μξ(b)\mu_{\xi}^{\prime}(a+)>\mu_{\xi}^{\prime}(b-) for r0a<b<-r_{0}\leq a<b<\infty. Furthermore, μξ(b+)>μξ((r0)+)\mu_{\xi}^{\prime}(b+)>\mu_{\xi}^{\prime}((-r_{0})+) for b(r0ε0,r0)b\in(-r_{0}-\varepsilon_{0},-r_{0}).

Concavity implies that one-sided derivatives μξ(b±)\mu_{\xi}^{\prime}(b\pm) for b>r0ε0b>-r_{0}-\varepsilon_{0} exist, μξ(b)μξ(b+)\mu_{\xi}^{\prime}(b-)\geq\mu_{\xi}^{\prime}(b+), and as functions of bb, they are nonincreasing, μξ(b)\mu_{\xi}^{\prime}(b-) is left-continuous, and μξ(b+)\mu_{\xi}^{\prime}(b+) is right-continuous. Strict concavity is the novel part of the theorem. This property is proved in Section 7, based on the modification argument of Section 5.2.

Introduce the notation

(2.10) L¯𝟎,x\displaystyle\underline{L}_{\hskip 0.9pt\mathbf{0},x} =minimal Euclidean length of a geodesic for T𝟎,x\displaystyle=\text{minimal Euclidean length of a geodesic for $T_{\mathbf{0},x}$}
andL¯𝟎,x\displaystyle\text{and}\qquad\overline{L}_{\hskip 0.9pt\mathbf{0},x} =maximal Euclidean length of a geodesic for T𝟎,x,\displaystyle=\text{maximal Euclidean length of a geodesic for $T_{\mathbf{0},x}$,}

with the superscripted variants L¯𝟎,x(b)=L¯𝟎,x(ω(b))\underline{L}^{(b)}_{\hskip 0.9pt\mathbf{0},x}=\underline{L}_{\hskip 0.9pt\mathbf{0},x}(\omega^{(b)}) and L¯𝟎,x(b)=L¯𝟎,x(ω(b))\overline{L}^{\hskip 0.7pt(b)}_{\hskip 0.9pt\mathbf{0},x}=\overline{L}_{\hskip 0.9pt\mathbf{0},x}(\omega^{(b)}) for shifted weights ω(b)\omega^{(b)}. For a continuous weight distribution L¯𝟎,x=L¯𝟎,x\underline{L}_{\hskip 0.9pt\mathbf{0},x}=\overline{L}_{\hskip 0.9pt\mathbf{0},x} almost surely because in that case geodesics are unique almost surely. This is not the case for all shifts because as bb increases the geodesic jumps occasionally and at the jump locations there are two geodesics.

Recall that a geodesic for standard FPP is by definition self-avoiding. Under the assumptions of Theorem 2.1, Theorem A.1 in Appendix A proves that the following holds on an event Ω0\Omega_{0} of full probability: L¯𝟎,x(b)<\overline{L}^{\hskip 0.7pt(b)}_{\hskip 0.9pt\mathbf{0},x}<\infty for all xdx\in\mathbb{Z}^{d} and b>r0ε0b>-r_{0}-\varepsilon_{0}, and there exist a finite deterministic constant cc and a finite random constant KK such that

(2.11) L¯𝟎,x(b)c|x|1(b+r0)0+ε0b>r0ε0whenever |x|1K.\overline{L}^{\hskip 0.7pt(b)}_{\mathbf{0},x}\leq\frac{c|x|_{1}}{(b+r_{0})\wedge 0+\varepsilon_{0}}\qquad\forall b>-r_{0}-\varepsilon_{0}\ \ \text{whenever }\ \ |x|_{1}\geq K.

We justify part (i) of Theorem 2.2. This sets the stage for further discussion. Let b>r0ε0b>-r_{0}-\varepsilon_{0}. Take 0<δb+r0+ε00<\delta\leq b+r_{0}+\varepsilon_{0} and η>0\eta>0. Considering the shifted weights on the minimal and maximal length geodesics of T𝟎,x(b)T^{(b)}_{\mathbf{0},x} leads to

(2.12) T𝟎,x(bδ)T𝟎,x(b)δL¯𝟎,x(b)andT𝟎,x(b+η)T𝟎,x(b)+ηL¯𝟎,x(b).T^{(b-\delta)}_{\mathbf{0},x}\leq T^{(b)}_{\mathbf{0},x}-\delta\overline{L}^{\hskip 0.7pt(b)}_{\hskip 0.9pt\mathbf{0},x}\quad\text{and}\quad T^{(b+\eta)}_{\mathbf{0},x}\leq T^{(b)}_{\mathbf{0},x}+\eta\underline{L}^{(b)}_{\hskip 0.9pt\mathbf{0},x}.

Rearrange to

(2.13) T𝟎,x(b+η)T𝟎,x(b)ηL¯𝟎,x(b)L¯𝟎,x(b)T𝟎,x(b)T𝟎,x(bδ)δ.\frac{T^{(b+\eta)}_{\mathbf{0},x}-T^{(b)}_{\mathbf{0},x}}{\eta}\leq\underline{L}^{(b)}_{\hskip 0.9pt\mathbf{0},x}\leq\overline{L}^{\hskip 0.7pt(b)}_{\hskip 0.9pt\mathbf{0},x}\leq\frac{T^{(b)}_{\mathbf{0},x}-T^{(b-\delta)}_{\mathbf{0},x}}{\delta}.

Here are the arguments for the properties of μξ\mu_{\xi} claimed in part (i) of Theorem 2.2.

  1. (i.a)

    Strict increasingness. In (2.12) take x=xnx=x_{n} such that xn/nξx_{n}/n\to\xi. Since L¯𝟎,x(b)|x|1\overline{L}^{\hskip 0.7pt(b)}_{\hskip 0.9pt\mathbf{0},x}\geq|x|_{1}, the inequality μξ(bδ)μξ(b)δ|ξ|1\mu_{\xi}(b-\delta)\leq\mu_{\xi}(b)-\delta|\xi|_{1} follows by taking the limit (2.8) in (2.12).

  2. (i.b)

    Concavity follows by taking the same limit in (2.13).

  3. (i.c)

    Continuity of μξ\mu_{\xi} on the open interval (r0ε0,)(-r_{0}-\varepsilon_{0},\infty) follows from concavity.

Since L¯𝟎,x(b)|x|1\underline{L}^{(b)}_{\hskip 0.9pt\mathbf{0},x}\geq|x|_{1}, (2.13) and the monotonicity of the derivatives give the easy bound

(2.14) μξ(b±)|ξ|1.\mu_{\xi}^{\prime}(b\pm)\geq|\xi|_{1}.

A corollary of the strict concavity given in Theorem 2.2(ii) is the strict inequality μξ(b±)>|ξ|1\mu_{\xi}^{\prime}(b\pm)>|\xi|_{1}. The next theorem records a slight strengthening of this and consequences of (2.11) and (2.13). A precise proof is given in Section 7.

Theorem 2.3.

Assume r00r_{0}\geq 0, (2.6), and the moment bound (2.7) with p=dp=d. Let ε0\varepsilon_{0} be the constant specified in Theorem 2.1 and let cc be the constant in (2.11). Then there exists a full-probability event Ω0\Omega_{0} such that the following holds: for all shifts b>r0ε0b>-r_{0}-\varepsilon_{0}, directions ξd{𝟎}\xi\in\mathbb{R}^{d}\hskip 0.7pt\setminus\{\mathbf{0}\}, weight configurations ωΩ0\omega\in\Omega_{0}, and sequences xn/nξx_{n}/n\to\xi, we have the bounds

(2.15) (1+D(b))|ξ|1μξ(b+)\displaystyle(1+D(b))|\xi|_{1}\leq\mu_{\xi}^{\prime}(b+) lim¯nL¯𝟎,xn(b)(ω)n\displaystyle\leq\varliminf_{n\to\infty}\frac{\underline{L}^{(b)}_{\hskip 0.9pt\mathbf{0},\hskip 0.9ptx_{n}}(\omega)}{n}
lim¯nL¯𝟎,xn(b)(ω)nμξ(b)2c(b+r0)0+ε0|ξ|1.\displaystyle\leq\varlimsup_{n\to\infty}\frac{\overline{L}^{\hskip 0.7pt(b)}_{\hskip 0.9pt\mathbf{0},\hskip 0.9ptx_{n}}(\omega)}{n}\leq\mu_{\xi}^{\prime}(b-)\leq\frac{2c}{(b+r_{0})\wedge 0+\varepsilon_{0}}\,|\xi|_{1}.

D(b)D(b) is a nonincreasing function of bb such that D(b)>0D(b)>0 for all b>r0ε0b>-r_{0}-\varepsilon_{0}.

The first inequality in (2.15) says that the strict concavity gap μξ(b+)>|ξ|1\mu_{\xi}^{\prime}(b+)>|\xi|_{1} is uniform across all directions |ξ|1=1|\xi|_{1}=1. This point is further strengthened to a uniformity for fixed weight configurations ω\omega in Theorem 2.5.

Remark 2.4.

Here are points that follow Theorems 2.2 and 2.3. Let ξ𝟎\xi\neq\mathbf{0}.

(i) The inequalities in (2.15) imply the limit of Hammersley-Welsh, Smythe-Wierman and Kesten simultaneously for all sequences. Under the assumptions of Theorem 2.3, suppose μξ\mu_{\xi} is differentiable at b(r0ε0,)b\in(-r_{0}-\varepsilon_{0},\infty). Then (2.15) implies that for all ωΩ0\omega\in\Omega_{0} and sequences xn/nξx_{n}/n\to\xi,

(2.16) limnL¯𝟎,xn(b)(ω)n=limnL¯𝟎,xn(b)(ω)n=μξ(b).\lim_{n\to\infty}\frac{\underline{L}^{(b)}_{\hskip 0.9pt\mathbf{0},\hskip 0.9ptx_{n}}(\omega)}{n}=\lim_{n\to\infty}\frac{\overline{L}^{\hskip 0.7pt(b)}_{\hskip 0.9pt\mathbf{0},\hskip 0.9ptx_{n}}(\omega)}{n}=\mu_{\xi}^{\prime}(b).

By concavity, this happens at all but countably many bb. In particular, if μξ\mu_{\xi} is a differentiable function then geodesic lengths converge with probability one, simultaneously in all directions and at all weight shifts. Presently there is no proof of differentiability under any hypotheses. Further below we show failures of differentiability under assumptions on the atoms of the weight distribution.

Suppose μξ(b+)<μξ(b)\mu_{\xi}^{\prime}(b+)<\mu_{\xi}^{\prime}(b-). Then (2.15) tells us that all the possible asymptotic normalized lengths of geodesics that go in direction ξ\xi form a subset of the interval [μξ(b+),μξ(b)][\hskip 0.9pt\mu_{\xi}^{\prime}(b+),\mu_{\xi}^{\prime}(b-)\hskip 0.9pt]. Presently there is no description of this subset.

For a characterization of [μξ(b+),μξ(b)][\hskip 0.9pt\mu_{\xi}^{\prime}(b+),\mu_{\xi}^{\prime}(b-)\hskip 0.9pt] in terms of path length, given below in Theorem 2.17, we expand the class of paths considered to allow o(n)o(n)-approximate geodesics. These are paths from the origin to nξ+o(n)n\xi+o(n) whose passage times are in the range nμξ(b)+o(n)n\mu_{\xi}(b)+o(n), without necessarily being geodesics between their endpoints.

(ii) The strict concavity of μξ\mu_{\xi} given in Theorem 2.2 and the inequalities in (2.15) imply that, for all ω,ω~Ω0\omega,\widetilde{\omega}\in\Omega_{0} and sequences xn/nξx_{n}/n\to\xi and x~n/nξ\widetilde{x}_{n}/n\to\xi,

(2.17) lim¯nL¯𝟎,xn(b)(ω)nμξ(b)<μξ(a+)lim¯nL¯𝟎,x~n(a)(ω~)nfor all b>a>r0ε0.\varlimsup_{n\to\infty}\frac{\overline{L}^{\hskip 0.7pt(b)}_{\hskip 0.9pt\mathbf{0},\hskip 0.9ptx_{n}}(\omega)}{n}\leq\mu_{\xi}^{\prime}(b-)<\mu_{\xi}^{\prime}(a+)\leq\varliminf_{n\to\infty}\frac{\underline{L}^{(a)}_{\hskip 0.9pt\mathbf{0},\hskip 0.9pt\widetilde{x}_{n}}(\widetilde{\omega})}{n}\quad\text{for all }b>a>-r_{0}-\varepsilon_{0}.

In other words, distinct shifts of a given weight distribution cannot share any possible asymptotic geodesic lengths, even under distinct but typical environments ω\omega and ω~\widetilde{\omega}.

(iii) There is a corresponding monotonicity for geodesics at fixed ω\omega. Namely, when all the weights increase by a common constant, geodesics can only shrink in length. Let π(a)\pi^{(a)} and π(b)\pi^{(b)} be arbitrary geodesics for T𝟎,x(a)T^{(a)}_{\mathbf{0},x} and T𝟎,x(b)T^{(b)}_{\mathbf{0},x}, respectively. Then

(2.18) |π(b)||π(a)||\pi^{(b)}|\leq|\pi^{(a)}| for fixed a<ba<b and ω\omega.

This follows from

(2.19) T(b)(π(a))(ba)|π(a)|\displaystyle T^{(b)}(\pi^{(a)})-(b-a)|\pi^{(a)}| =T(a)(π(a))T(a)(π(b))=T(b)(π(b))(ba)|π(b)|\displaystyle=T^{(a)}(\pi^{(a)})\leq T^{(a)}(\pi^{(b)})=T^{(b)}(\pi^{(b)})-(b-a)|\pi^{(b)}|
T(b)(π(a))(ba)|π(b)|.\displaystyle\leq T^{(b)}(\pi^{(a)})-(b-a)|\pi^{(b)}|.

Furthermore, suppose a unique geodesic is chosen, for example by taking the minimal one according to some ordering of geodesics. Then as aa increases to bb, the geodesic cannot change without its length strictly shrinking:

(2.20) for fixed a<ba<b and ω\omega, |π(b)|=|π(a)||\pi^{(b)}|=|\pi^{(a)}| implies π(b)=π(a)\pi^{(b)}=\pi^{(a)}.

This follows because the string of inequalities (2.19) together with |π(b)|=|π(a)||\pi^{(b)}|=|\pi^{(a)}| implies that T(b)(π(a))T(b)(π(b))T^{(b)}(\pi^{(a)})\leq T^{(b)}(\pi^{(b)}), so π(a)\pi^{(a)} is still at least as good as π(b)\pi^{(b)} for weights {t(b)(e)}\{t^{(b)}(e)\}.

(iv) We establish below in Theorem 2.17 that μξ(b±)|ξ|1\mu_{\xi}^{\prime}(b\pm)\to|\xi|_{1} as bb\to\infty. Naturally, as the weight shift grows very large, it pays less to search for smaller weights at the expense of a longer geodesic. pointlesstext \triangle

The first inequality in (2.15) implies that asymptotically the lengths of geodesics in a particular direction ξ\xi exceed the 1\ell^{1}-distance. The next theorem strengthens this to a uniformity across all sufficiently faraway lattice endpoints. Its proof in Section 7 relies on the convex duality described in Section 2.4, the restricted path length shape theorem of Appendix B, and the modification arguments of Section 5.

Theorem 2.5.

Assume r00r_{0}\geq 0, (2.6), and the moment bound (2.7) with p=dp=d. There exist a deterministic constant δ>0\delta>0 and an almost surely finite random constant KK such that L¯𝟎,x(1+δ)|x|1\underline{L}_{\hskip 0.9pt\mathbf{0},x}\geq(1+\delta)|x|_{1} whenever xdx\in\mathbb{Z}^{d} satisfies |x|1K|x|_{1}\geq K.

We turn to nondifferentiability results for μξ\mu_{\xi}. An atom of the weight distribution is a value rr\in\mathbb{R} such that {t(e)=r}>0\mathbb{P}\{t(e)=r\}>0.

Theorem 2.6.

Assume r00r_{0}\geq 0, (2.6), and the moment bound (2.7) with p=dp=d. Additionally, assume that the weight distribution satisfies at least one of the assumptions (a) and (b) below:

  1. (a)

    zero is an atom;

  2. (b)

    there are two strictly positive atoms r<sr<s such that s/rs/r is rational.

Then there exist constants 0<D,δ,M<0<D,\delta,M<\infty such that

(2.21) (L¯𝟎,xL¯𝟎,xD|x|1)δfor |x|1M\mathbb{P}\bigl{(}\,\overline{L}_{\hskip 0.9pt\mathbf{0},x}-\underline{L}_{\hskip 0.9pt\mathbf{0},x}\geq D|x|_{1}\bigr{)}\geq\delta\qquad\text{for $|x|_{1}\geq M$. }

Furthermore, for all ξd{𝟎}\xi\in\mathbb{R}^{d}\hskip 0.7pt\setminus\{\mathbf{0}\}, μξ(0)μξ(0+)D|ξ|1\mu_{\xi}^{\prime}(0-)-\mu_{\xi}^{\prime}(0+)\geq D|\xi|_{1} and so the function μξ(a)=μ(a)(ξ)\mu_{\xi}(a)=\mu^{(a)}(\xi) is not differentiable at a=0a=0.

For unbounded weights the result above can be proved under more general assumptions on the atoms (see Theorem 6.2 in Section 6).

Theorem 2.7.

Assume r00r_{0}\geq 0, (2.6), and the moment bound (2.7) with p=dp=d. Additionally, assume that the weight distribution has at least two atoms. Then there exists a countably infinite set B[r0,)B\subset[-r_{0},\infty) with these properties.

  1. (i)

    BB is dense in [r0,)[-r_{0},\infty).

  2. (ii)

    For each bBb\in B, conclusion (2.21) of Theorem 2.6 holds for the shifted weights ω(b)\omega^{(b)} with constants D(b),δ(b),M(b)D^{(b)},\delta^{(b)},M^{(b)} that depend on bb.

  3. (iii)

    For each ξd{𝟎}\xi\in\mathbb{R}^{d}\hskip 0.7pt\setminus\{\mathbf{0}\} and bBb\in B, μξ(a)=μ(a)(ξ)\mu_{\xi}(a)=\mu^{(a)}(\xi) is not differentiable at a=ba=b.

The proof of Theorem 2.7 in Section 7.2 constructs the singularity set BB explicitly from two atoms of t(e)t(e) as a countably infinite union of arithmetic sequences.

Remark 2.8.

Standard Bernoulli weights satisfy {t(e)=0}+{t(e)=1}=1\mathbb{P}\{t(e)=0\}+\mathbb{P}\{t(e)=1\}=1. In the subcritical planar Bernoulli case (that is, d=2d=2, t(e){0,1}t(e)\in\{0,1\} and {t(e)=0}<12\mathbb{P}\{t(e)=0\}<\tfrac{1}{2}), Steele and Zhang [19] proved that μ𝐞1(a)\mu_{\mathbf{e}_{1}}(a) is not differentiable at a=0a=0, as long as {t(e)=0}\mathbb{P}\{t(e)=0\} is close enough to 12\tfrac{1}{2}. Furthermore, they conjectured that μ𝐞1(a)\mu_{\mathbf{e}_{1}}(a) is differentiable at all aa such that μ𝐞1(a)>\mu_{\mathbf{e}_{1}}(a)>-\infty except at a=0a=0 (page 1050 in [19]).

Theorem 2.6 above extends the nondifferentiability at a=0a=0 to all directions ξ\xi, all dimensions, and all weight distributions that have an atom at zero. Theorem 2.7 above disproves the Steele-Zhang conjecture by showing that, in all dimensions, in the subcritical Bernoulli case the nondifferentiability points form a countably infinite dense subset of (0,)(0,\infty). \triangle

2.3. Restricted path-length first-passage percolation

Next we discuss FPP models that restrict the length of the paths over which the optimization takes place but give up the self-avoidance requirement. Remark 2.15 below characterizes the FPP shape function μ\mu as the positively homogeneous convex function generated by the restricted path shape functions. In the next Section 2.4 this leads to the convex duality of μξ\mu_{\xi} and a sharpening of Theorem 2.3, and further conceptual understanding of the previous results.

It turns out convenient to consider also a version whose paths are allowed zero steps. In this case the set ={±𝐞1,,±𝐞d}\mathcal{R}=\{\pm\mathbf{e}_{1},\dotsc,\pm\mathbf{e}_{d}\} of admissible steps is augmented to o={𝟎}\mathcal{R}^{o}=\mathcal{R}\cup\{\mathbf{0}\}. For x,ydx,y\in\mathbb{Z}^{d} and nn\in\mathbb{N} define three classes of paths x0:n=(xi)i=0nx_{0\hskip 0.6pt{\boldsymbol{:}}\hskip 0.3ptn}=(x_{i})_{i=0}^{n} from xx to yy of length nn, presented here from largest to smallest:

(2.22) Πx,(n),yo\displaystyle\Pi^{o}_{\hskip 0.55ptx,(n),y} ={x0:n(d)n+1:x0=x,xn=y, each xixi1o},\displaystyle=\{x_{0\hskip 0.6pt{\boldsymbol{:}}\hskip 0.3ptn}\in(\mathbb{Z}^{d})^{n+1}:x_{0}=x,x_{n}=y,\text{ each }x_{i}-x_{i-1}\in\mathcal{R}^{o}\},
Πx,(n),y\displaystyle\Pi_{\hskip 0.55ptx,(n),y} ={x0:n(d)n+1:x0=x,xn=y, each xixi1},\displaystyle=\{x_{0\hskip 0.6pt{\boldsymbol{:}}\hskip 0.3ptn}\in(\mathbb{Z}^{d})^{n+1}:x_{0}=x,x_{n}=y,\text{ each }x_{i}-x_{i-1}\in\mathcal{R}\},
andΠx,(n),y sa\displaystyle\text{and}\qquad\Pi^{\text{\rm{\,sa}}}_{\hskip 0.55ptx,(n),y} ={x0:nΠx,(n),y: points x0,x1,,xn are distinct}.\displaystyle=\{x_{0\hskip 0.6pt{\boldsymbol{:}}\hskip 0.3ptn}\in\Pi_{\hskip 0.55ptx,(n),y}:\text{ points $x_{0},x_{1},\dotsc,x_{n}$ are distinct}\}.

The superscript in Π sa\Pi^{\text{\rm{\,sa}}} is for self-avoiding. Paths in Πx,(n),y\Pi_{\hskip 0.55ptx,(n),y} and Πx,(n),yo\Pi^{o}_{\hskip 0.55ptx,(n),y} are allowed to repeat both vertices and edges. Paths in Πx,(n),y\Pi_{\hskip 0.55ptx,(n),y} are called \mathcal{R}-admissible, and those in Πx,(n),yo\Pi^{o}_{\hskip 0.55ptx,(n),y} o\mathcal{R}^{o}-admissible. An nn-path x0:nx_{0\hskip 0.6pt{\boldsymbol{:}}\hskip 0.3ptn} from x0=xx_{0}=x to xn=yx_{n}=y is an 1\ell^{1}-path if n=|yx|1n=|y-x|_{1}. For n=0n=0 and {𝚎𝚖𝚙𝚝𝚢,o,sa}\diamond\in\{\langle{\tt empty}\rangle,o,\text{sa}\} we define each collection Πx,(0),x\Pi^{\,\diamond}_{\hskip 0.55ptx,(0),x} as consisting only of the zero-length path (x)(x). For xyx\neq y, Πx,(n),y\Pi_{\hskip 0.55ptx,(n),y} and Πx,(n),y sa\Pi^{\text{\rm{\,sa}}}_{\hskip 0.55ptx,(n),y} are nonempty if and only if n|yx|1n-|y-x|_{1} is a nonnegative even integer, while Πx,(n),yo\Pi^{o}_{\hskip 0.55ptx,(n),y} is nonempty if and only if n|yx|1n\geq|y-x|_{1}.

With the three classes of paths go three collections of points reachable by an admissible path of length nn from the origin: for the three superscripts {𝚎𝚖𝚙𝚝𝚢,o,sa}\diamond\in\{\langle{\tt empty}\rangle,o,\text{sa}\}, define

(2.23) 𝒟n={xd:Π𝟎,(n),x}.\mathcal{D}^{\,\diamond}_{n}=\{x\in\mathbb{Z}^{d}:\Pi^{\,\diamond}_{\hskip 0.55pt\mathbf{0},(n),x}\neq\varnothing\}.

If 0k<n0\leq k<n, any kk-path can be augmented to an nn-path by adding nkn-k zero steps, and hence we have 𝒟no=0kn𝒟k\mathcal{D}^{o}_{n}=\cup_{0\leq k\leq n}\mathcal{D}_{k}.

The environment ω=(t(e):ed)\omega=(t(e):e\in{\mathcal{E}}_{d}) is extended to zero steps by stipulating that zero steps always have zero weight, even when weights are shifted: t(b)({x,x})=0t^{(b)}(\{x,x\})=0 xd\forall x\in\mathbb{Z}^{d} and bb\in\mathbb{R}.

Define three point-to-point first-passage times between two points x,ydx,y\in\mathbb{Z}^{d} with restricted path lengths: for {𝚎𝚖𝚙𝚝𝚢,o,sa}\diamond\in\{\langle{\tt empty}\rangle,o,\text{sa}\},

(2.24) Gx,(n),y\displaystyle G^{\,\diamond}_{x,(n),y} =minx0:nΠx,(n),yk=0n1t({xk,xk+1})for yx𝒟n.\displaystyle=\min_{x_{0\hskip 0.6pt{\boldsymbol{:}}\hskip 0.3ptn}\,\in\,\Pi^{\,\diamond}_{\hskip 0.55ptx,(n),y}}\;\sum_{k=0}^{n-1}t(\{x_{k},x_{k+1}\})\quad\text{for }\ y-x\in\mathcal{D}^{\,\diamond}_{n}.

If Πx,(n),y=\Pi^{\,\diamond}_{\hskip 0.55ptx,(n),y}=\varnothing, set Gx,(n),y=G^{\,\diamond}_{x,(n),y}=\infty. Obvious relations hold between these passage times and the standard FPP from (2.2):

(2.25) Gx,(n),yo=mink:|yx|1knGx,(k),y,G^{o}_{x,(n),y}=\min_{k:\,|y-x|_{1}\leq k\leq n}G_{x,(k),y},
Πx,y sa=n|yx|1Πx,(n),y sa\Pi^{\text{\rm{\,sa}}}_{x,y}=\bigcup_{n\geq|y-x|_{1}}\Pi^{\text{\rm{\,sa}}}_{\hskip 0.55ptx,(n),y}

and

(2.26) Tx,y=infπΠx,y saT(π)=infn:n|yx|1Gx,(n),y sa.T_{x,y}=\inf_{\pi\hskip 0.55pt\in\hskip 0.55pt\Pi^{\text{\rm{\,sa}}}_{x,y}}T(\pi)=\inf_{n:\,n\geq|y-x|_{1}}G^{\text{\,\rm{sa}}}_{x,(n),y}.

For nonnegative weights the restriction to self-avoiding paths is superfluous for Tx,yT_{x,y} and hence

(2.27) if r00 then Tx,y=infn:n|yx|1Gx,(n),yo=infn:n|yx|1Gx,(n),y.\text{if \ $r_{0}\geq 0$ \ then }\ \ T_{x,y}=\inf_{n:\,n\geq|y-x|_{1}}G^{o}_{x,(n),y}=\inf_{n:\,n\geq|y-x|_{1}}G_{x,(n),y}.

These identities point to the usefulness of GG and GoG^{o}. Namely, they capture the FPP passage time when the path length parameter nn coincides with a geodesic length. After taking this connection to the limit, the discrepancies between the shape function of GG and the FPP shape function μ\mu reveal which asymptotic path lengths are too short and which are too long to be asymptotic geodesic lengths.

The reader may wonder about the purpose of GoG^{o} and the zero-weight zero step. We shall see that GoG^{o} is a convenient link between standard FPP and restricted path length FPP because it is monotone:

(2.28) if r00 and mn then Gx,(m),yoGx,(n),yoTx,y.\text{if }r_{0}\geq 0\text{ and }m\leq n\text{ then }G^{o}_{x,(m),y}\geq G^{o}_{x,(n),y}\geq T_{x,y}.

The monotonicity is simply a consequence of the fact that any mm-path can be augmented to an nn-path by adding zero steps.

The self-avoiding version Gx,(n),y saG^{\text{\,\rm{sa}}}_{x,(n),y} is mentioned here to complete the overall picture but will not be used in the sequel. Open problem 3.3 points the way to an extension of this work that requires a study of Gx,(n),y saG^{\text{\,\rm{sa}}}_{x,(n),y}.

We state a shape theorem for restricted path length FPP, but only on the open set int𝒰={ξd:|ξ|1<1}\operatorname{int}\mathcal{U}=\{\xi\in\mathbb{R}^{d}:|\xi|_{1}<1\}. Its closure, the compact 1\ell^{1} ball 𝒰\mathcal{U}, is the convex hull of both \mathcal{R} and o\mathcal{R}^{o} and the set of possible asymptotic velocities of admissible paths in Π𝟎,(n),\scaleobj0.5∙\Pi^{\,\diamond}_{\hskip 0.55pt\mathbf{0},(n),\hskip 0.9pt{\raisebox{0.5pt}{\scaleobj{0.5}{\bullet}}}} as nn\to\infty. In this theorem we introduce the parameter α\alpha as a variable that controls asymptotic path length.

Theorem 2.9.

Assume r0>r_{0}>-\infty and that the moment bound (2.7) holds with p=dp=d for the nonnegative weights ti+=ti0t_{i}^{+}=t_{i}\vee 0. Then there exist

  1. (a)

    nonrandom continuous convex functions g:int𝒰[r0,)g:\operatorname{int}\mathcal{U}\to[r_{0},\infty) and go:int𝒰[r00,)g^{o}:\operatorname{int}\mathcal{U}\to[r_{0}\wedge 0,\infty) and

  2. (b)

    an event Ω0\Omega_{0} of \mathbb{P}-probability one

such that the following statement holds for any fixed ωΩ0\omega\in\Omega_{0}: for any ξd\xi\in\mathbb{R}^{d}, any real α>|ξ|1\alpha>|\xi|_{1}, and any sequences knk_{n}\to\infty in \mathbb{N}, xn𝒟knx_{n}\in\mathcal{D}_{k_{n}} and yn𝒟knoy_{n}\in\mathcal{D}^{o}_{k_{n}} such that kn/nαk_{n}/n\to\alpha, xn/nξx_{n}/n\to\xi and yn/nξy_{n}/n\to\xi, we have the laws of large numbers

(2.29) αg(ξα)=limnG𝟎,(kn),xnnandαgo(ξα)=limnG𝟎,(kn),ynon.\alpha\hskip 0.55ptg\biggl{(}\frac{\xi}{\alpha}\biggr{)}=\lim_{n\to\infty}\frac{G_{\mathbf{0},(k_{n}),x_{n}}}{n}\quad\text{and}\quad\alpha\hskip 0.55ptg^{o}\biggl{(}\frac{\xi}{\alpha}\biggr{)}=\lim_{n\to\infty}\frac{G^{o}_{\hskip 0.9pt\mathbf{0},(k_{n}),y_{n}}}{n}.

Furthermore, g(𝟎)=r0g(\mathbf{0})=r_{0} and go(𝟎)=r00g^{o}(\mathbf{0})=r_{0}\wedge 0. In general gogg^{o}\leq g on int𝒰\operatorname{int}\mathcal{U}. If r00r_{0}\leq 0 then g=gog=g^{o} on all of int𝒰\operatorname{int}\mathcal{U}. If r0>0r_{0}>0 then g>gog>g^{o} in a neighborhood of the origin.

The laws of large numbers (2.29) come from Theorem B.1 in Appendix B. The soft properties of gg and gog^{o} stated in the last paragraph of Theorem 2.9 are proved in Lemma 4.1 in Section 4. Figure 2.2 illustrates the limit functions in (2.29).

It is convenient to have gg^{\diamond} defined on the whole of 𝒰\mathcal{U}. An attempt to do this through the laws of large numbers (2.29) would divert attention from the main points of this paper. Furthermore, without stronger moment assumptions there cannot be a finite limit, as can be observed by considering ξ=𝐞1\xi=\mathbf{e}_{1}. Since there is a unique nn-path from 𝟎\mathbf{0} to n𝐞1n\mathbf{e}_{1}, we see that a finite limit is possible only if t(e)L1()t(e)\in L^{1}(\mathbb{P}):

(2.30) limnn1G𝟎,(n),n𝐞1=limnn1k=1nt({(k1)𝐞1,k𝐞1})=𝔼[t(e)].\lim_{n\to\infty}n^{-1}G^{\,\diamond}_{\mathbf{0},(n),n\mathbf{e}_{1}}=\lim_{n\to\infty}n^{-1}\sum_{k=1}^{n}t(\{(k-1)\mathbf{e}_{1},k\mathbf{e}_{1}\})=\mathbb{E}[t(e)].

Instead of limiting passage times, we take radial limits of the shape functions from the interior as stated in the next theorem. The proof of this theorem comes in Lemma 4.1(iv).

Theorem 2.10.

Under the assumptions of Theorem 2.9 we can extend both shape functions to all of 𝒰\mathcal{U} via limits along rays: for {𝚎𝚖𝚙𝚝𝚢,o}\diamond\in\{\langle{\tt empty}\rangle,o\} and |ξ|1=1|\xi|_{1}=1 define g(ξ)=limt1g(tξ)g^{\diamond}(\xi)=\lim_{t\nearrow 1}g^{\diamond}(t\xi). The resulting functions g:𝒰[r0,]g:\mathcal{U}\to[r_{0},\infty] and go:𝒰[r00,]g^{o}:\mathcal{U}\to[r_{0}\wedge 0,\infty] are both convex and lower semicontinuous.

With this theorem we can extend the functions gg^{\diamond} to lower semicontinuous proper convex functions on all of d\mathbb{R}^{d} by setting

(2.31) g(ξ)=+for ξ𝒰.g^{\diamond}(\xi)=+\infty\qquad\text{for }\ \xi\notin\mathcal{U}.

If gg^{\diamond} is finite on 𝒰\mathcal{U}, then gg^{\diamond} is automatically upper semicontinuous on 𝒰\mathcal{U} [17, Theorem 10.2], and hence continuous on 𝒰\mathcal{U}.

The next theorem clarifies the relationship of gg and gog^{o} with μ\mu, beyond the obvious μgog\mu\leq g^{o}\leq g, and links their connection with the asymptotic geodesic lengths from Theorem 2.3. In particular, we introduce here two functions λ¯λ¯\underline{\lambda}\leq\overline{\lambda} that play several roles in our asymptotic results. In the theorem below they are first introduced as the boundaries of the regions where μ\mu coincides with gg and gog^{o}. Part (ii) indicates that λ¯\underline{\lambda} and λ¯\overline{\lambda} are also related to the derivatives of μξ\mu_{\xi} and geodesic length.

These properties are then elaborated on as we proceed. The interval [λ¯(ξ),λ¯(ξ)][\underline{\lambda}(\xi),\overline{\lambda}(\xi)] captures all the asymptotic lengths of geodesics in direction ξ\xi, while the full interval is exactly the set of all asymptotic lengths of approximate geodesics (Remark 2.13). In Theorem 2.16 we see that λ¯\underline{\lambda} and λ¯\overline{\lambda} describe ranges where gg and gog^{o} are affine and where these two functions disagree. The macroscopic description is completed in Theorem 2.17: as the weight shift bb increases, the interval [λ¯(b)(ξ),λ¯(b)(ξ)][\,\underline{\lambda}^{(b)}(\xi),\overline{\lambda}^{\hskip 0.9pt(b)}(\xi)\,] shifts to the left and always equals the superdifferential μξ(b)\partial\mu_{\xi}(b) of the concave function μξ\mu_{\xi}. Then we have reached the desired generalization of the Hammersley-Welsh connection (1.1): the assumptions of differentiability and existence of limiting geodesic length have been dropped, and the correct identity equates the superdifferential with the set of asymptotic lengths of approximate geodesics.

Set

(2.32) μ=sup|ξ|1=1μ(ξ).\mu^{*}=\sup_{|\xi|_{1}=1}\mu(\xi).

In part (ii) of the theorem, on both lines of (2.35) the first inequality depends on the modification arguments and hence the subcriticality assumption is strengthened to (2.6). To capture the complete picture we include in (2.35) the inequalities from (2.15).

Theorem 2.11.

Assume r00r_{0}\geq 0, (2.3), and the moment bound (2.7) with p=dp=d.

  1. (i)

    There exist two positively homogeneous functions λ¯:d+\underline{\lambda}:\mathbb{R}^{d}\to\mathbb{R}_{+} and λ¯:d[0,]\overline{\lambda}:\mathbb{R}^{d}\to[0,\infty] such that λ¯λ¯\underline{\lambda}\leq\overline{\lambda}, and for all ξ𝒰\xi\in\mathcal{U},

    (2.33) go(ξ)=μ(ξ)λ¯(ξ)1g^{o}(\xi)=\mu(\xi)\;\Longleftrightarrow\;\underline{\lambda}(\xi)\leq 1

    and

    (2.34) g(ξ)=μ(ξ)λ¯(ξ)1λ¯(ξ).g(\xi)=\mu(\xi)\;\Longleftrightarrow\;\underline{\lambda}(\xi)\leq 1\leq\overline{\lambda}(\xi).

    Furthermore, λ¯\underline{\lambda} is lower semicontinuous and λ¯\overline{\lambda} is upper semicontinuous. If r0=0r_{0}=0 then λ¯(ξ)\overline{\lambda}(\xi)\equiv\infty, while λ¯\overline{\lambda} is finite in the case r0>0r_{0}>0.

  2. (ii)

    Strengthen the subcriticality assumption to (2.6). There exists a nonrandom constant D>0D>0 and a full-probability event Ω0\Omega_{0} such that, for all ξd{𝟎}\xi\in\mathbb{R}^{d}\hskip 0.7pt\setminus\{\mathbf{0}\}, sequences xn/nξx_{n}/n\to\xi, and ωΩ0\omega\in\Omega_{0},

    (2.35) (1+D)|ξ|1λ¯(ξ)\displaystyle(1+D)|\xi|_{1}\leq\underline{\lambda}(\xi) =μξ(0+)lim¯nn1L¯𝟎,xn(ω)\displaystyle=\mu_{\xi}^{\prime}(0+)\leq\varliminf_{n\to\infty}n^{-1}{\underline{L}_{\hskip 0.9pt\mathbf{0},\hskip 0.9ptx_{n}}(\omega)}
    lim¯nn1L¯𝟎,xn(ω)μξ(0)<λ¯(ξ)=if r0=0\displaystyle\leq\varlimsup_{n\to\infty}n^{-1}{\overline{L}_{\hskip 0.9pt\mathbf{0},\hskip 0.9ptx_{n}}(\omega)}\leq\mu_{\xi}^{\prime}(0-)<\overline{\lambda}(\xi)=\infty\quad\text{if }r_{0}=0
    and(1+D)|ξ|1λ¯(ξ)\displaystyle\text{and}\quad(1+D)|\xi|_{1}\leq\underline{\lambda}(\xi) =μξ(0+)lim¯nn1L¯𝟎,xn(ω)\displaystyle=\mu_{\xi}^{\prime}(0+)\leq\varliminf_{n\to\infty}n^{-1}{\underline{L}_{\hskip 0.9pt\mathbf{0},\hskip 0.9ptx_{n}}(\omega)}
    lim¯nn1L¯𝟎,xn(ω)μξ(0)=λ¯(ξ)(μ/r0)|ξ|1if r0>0.\displaystyle\leq\varlimsup_{n\to\infty}n^{-1}{\overline{L}_{\hskip 0.9pt\mathbf{0},\hskip 0.9ptx_{n}}(\omega)}\leq\mu_{\xi}^{\prime}(0-)=\overline{\lambda}(\xi)\leq(\mu^{*}/r_{0})|\xi|_{1}\quad\text{if }r_{0}>0.

We spell out some of the consequences of the theorems.

Remark 2.12 (Coincidence of shape functions).

There exists a finite constant κ\kappa such that λ¯(ξ)κ|ξ|1\underline{\lambda}(\xi)\leq\kappa|\xi|_{1} ξd\forall\xi\in\mathbb{R}^{d}. This follows from lower semicontinuity and homogeneity, but is also proved directly from Kesten’s fundamental bound in Lemma 4.2 below. Hence the set {μ=go}={λ¯1}\{\mu=g^{o}\}=\{\underline{\lambda}\leq 1\} contains the nondegenerate neighborhood {ξd:|ξ|1κ1}\{\xi\in\mathbb{R}^{d}:|\xi|_{1}\leq\kappa^{-1}\} of the origin.

If r0=0r_{0}=0 then {μ=g}={μ=go}\{\mu=g\}=\{\mu=g^{o}\} because g=gog=g^{o}. If r0>0r_{0}>0 the equality μ(ξ)=g(ξ)\mu(\xi)=g(\xi) holds for at least one nonzero point ξ\xi along each ray from the origin. With all of the above, the first inequality of (2.35) implies that {μ=g}\{\mu=g\} and {μ=go}\{\mu=g^{o}\} are both nonempty closed subsets of int𝒰\operatorname{int}\mathcal{U}. \triangle

Remark 2.13 (o(n)o(n)-approximate geodesics).

For α>|ξ|1>0\alpha>|\xi|_{1}>0, (2.34) gives the equivalence μ(ξ)=αg(ξ/α)\mu(\xi)=\alpha\hskip 0.55ptg({\xi}/{\alpha}) if and only if α[λ¯(ξ),λ¯(ξ)]\alpha\in[\hskip 0.9pt\underline{\lambda}(\xi),\overline{\lambda}(\xi)\hskip 0.9pt]. (This is illustrated in Figure 2.2 below.) By the law of large numbers (2.29), this happens if and only if, with probability one, there are lattice points xnx_{n} and paths πn\pi^{n} from 𝟎\mathbf{0} to xnx_{n} such that xn/nξx_{n}/n\to\xi, |πn|/nα|\pi^{n}|/n\to\alpha and T(πn)/nμ(ξ)T(\pi^{n})/n\to\mu(\xi). These paths πn\pi^{n} do not have to be self-avoiding or geodesics between their endpoints. But T(πn)/nμ(ξ)T(\pi^{n})/n\to\mu(\xi) does imply that T(πn)T(\pi^{n}) is within o(n)o(n) of the passage time of the geodesic between 𝟎\mathbf{0} and xnx_{n}. The asymptotic normalized lengths of true self-avoiding geodesics for μ(ξ)\mu(\xi) are a subset of the interval [λ¯(ξ),λ¯(ξ)][\hskip 0.9pt\underline{\lambda}(\xi),\overline{\lambda}(\xi)\hskip 0.9pt] of asymptotic normalized lengths of o(n)o(n)-approximate geodesics, as indicated in (2.35). \triangle

Remark 2.14 (Convergence of geodesic length).

We now see the connection between the convergence of the normalized geodesic length and the coincidence of shape functions. In the case r0>0r_{0}>0, (2.35) shows that convergence in direction ξ𝟎\xi\neq\mathbf{0} follows from λ¯(ξ)=λ¯(ξ)\underline{\lambda}(\xi)=\overline{\lambda}(\xi), which is equivalent to the condition that the set {μ=g}\{\mu=g\} has empty relative interior on the ξ\xi-directed ray. \triangle

Remark 2.15 (Convexity).

Fix ξd{𝟎}\xi\in\mathbb{R}^{d}\hskip 0.7pt\setminus\{\mathbf{0}\}. For {𝚎𝚖𝚙𝚝𝚢,o}\diamond\in\{\langle{\tt empty}\rangle,o\}, the convexity and continuity of gg^{\diamond} on int𝒰\operatorname{int}\mathcal{U} imply the convexity and continuity of the function ααg(ξ/α)\alpha\mapsto\alpha g^{\diamond}({\xi}/{\alpha}) defined for α(|ξ|1,)\alpha\in(\hskip 0.9pt|\xi|_{1},\infty). By Theorem 2.10, αg(ξ/α)\alpha g^{\diamond}({\xi}/{\alpha}) extends to α=|ξ|1\alpha=|\xi|_{1} by letting α|ξ|1\alpha\searrow|\xi|_{1}. By (2.31), we extend αg(ξ/α)\alpha g^{\diamond}({\xi}/{\alpha}) to α[0,|ξ|1)\alpha\in[0,|\xi|_{1}) by setting its value equal to ++\infty. Thereby ααg(ξ/α)\alpha\mapsto\alpha g^{\diamond}({\xi}/{\alpha}) is a lower semicontinuous proper convex function on +\mathbb{R}_{+}.

For gog^{o}, monotonicity (2.28) implies further that

(2.36) ααgo(ξ/α)\alpha\mapsto\alpha\hskip 0.55ptg^{o}({\xi}/{\alpha}) is nonincreasing for α(|ξ|1,)\alpha\in(\hskip 0.9pt|\xi|_{1},\infty).

A consequence of Theorem 2.11 is that for ξd{𝟎}\xi\in\mathbb{R}^{d}\hskip 0.7pt\setminus\{\mathbf{0}\},

(2.37) μ(ξ)=infα|ξ|1αg(ξα)=infα0αg(ξα)={αgo(ξ/α)α[λ¯(ξ),),αg(ξ/α)α[λ¯(ξ),λ¯(ξ)][λ¯(ξ),).\mu(\xi)=\inf_{\alpha\geq|\xi|_{1}}\alpha\hskip 0.55ptg^{\diamond}\biggl{(}\frac{\xi}{\alpha}\biggr{)}=\inf_{\alpha\geq 0}\alpha\hskip 0.55ptg^{\diamond}\biggl{(}\frac{\xi}{\alpha}\biggr{)}=\begin{cases}\alpha\hskip 0.55ptg^{o}({\xi}/{\alpha})&\forall\alpha\in[\hskip 0.9pt\underline{\lambda}(\xi),\infty),\\[3.0pt] \alpha\hskip 0.55ptg({\xi}/{\alpha})&\forall\alpha\in[\hskip 0.9pt\underline{\lambda}(\xi),\overline{\lambda}(\xi)\hskip 0.9pt]\cap[\hskip 0.9pt\underline{\lambda}(\xi),\infty).\end{cases}

In the language of convex analysis [17, p. 35], the identity above characterizes the standard FPP shape function μ\mu as the positively homogeneous convex function generated by gg^{\diamond}. This means that μ\mu is the greatest positively homogeneous convex function such that μ(𝟎)0\mu(\mathbf{0})\leq 0 and μg\mu\leq g^{\diamond}. Figure 2.2 illustrates (2.37). \triangle

0ttμ(r0)(tξ)=tμ(r0)(ξ)\mu^{(-r_{0})}(t\xi)=t\mu^{(-r_{0})}(\xi)μ(tξ)=tμ(ξ)\mu(t\xi)=t\mu(\xi)1λ¯(r0)(ξ)\displaystyle\frac{1}{\underline{\lambda}^{(-r_{0})}(\xi)}1λ¯(ξ)\displaystyle\frac{1}{\overline{\lambda}(\xi)}1λ¯(ξ)\displaystyle\frac{1}{\underline{\lambda}(\xi)}1|ξ|1\displaystyle\frac{1}{|\xi|_{1}}r0r_{0}thin: tgo(tξ)t\mapsto g^{o}(t\xi)thick: tg(tξ)t\mapsto g(t\xi)thick affineslope μ(r0)(ξ)\mu^{(-r_{0})}(\xi)thick = thin = μ(tξ)\mu(t\xi)affinethick = thinslope ==\inftythin affineslope μ(ξ)\mu(\xi)
Figure 2.1. Illustration of Theorem 2.16 in the case r0>0r_{0}>0. On the tt-axis it is possible that the two middle points 1λ¯(ξ)\frac{1}{\overline{\lambda}(\xi)} and 1λ¯(ξ)\frac{1}{\underline{\lambda}(\xi)} coincide. The separation illustrated here is the case where λ¯(ξ)=μξ(0+)<μξ(0)=λ¯(ξ)\underline{\lambda}(\xi)=\mu_{\xi}^{\prime}(0+)<\mu_{\xi}^{\prime}(0-)=\overline{\lambda}(\xi), which can happen when r0>0r_{0}>0 for example in the situation described in Theorem 2.6. Strict concavity of μξ\mu_{\xi} implies that the middle points are necessarily separated from 1λ¯(r0)(ξ)\frac{1}{\underline{\lambda}^{(-r_{0})}(\xi)} and 1|ξ|1\frac{1}{|\xi|_{1}} (see (2.50) below).

The last theorem of this section records further properties of gg^{\diamond}, illustrated in Figure 2.1. Part (iii) can be proved only in Section 7 after the modification results and hence requires the stronger subcriticality assumption (2.6).

Theorem 2.16.

Assume r00r_{0}\geq 0, (2.3), and the moment bound (2.7) with p=dp=d. Fix ξd{𝟎}\xi\in\mathbb{R}^{d}\hskip 0.7pt\setminus\{\mathbf{0}\}. For {𝚎𝚖𝚙𝚝𝚢,o}\diamond\in\{\langle{\tt empty}\rangle,o\}, the shape functions gg^{\diamond} of Theorem 2.9 have the following properties along the ξ\xi-directed ray from the origin.

  1. (i)

    The function tg(tξ)t\mapsto g^{\diamond}(t\xi) is continuous, convex and strictly increasing for t[0,|ξ|11)t\in[0,|\xi|_{1}^{-1}). Both functions are affine at least in one nondegenerate interval with one endpoint at the origin: for t[0,|ξ|11]t\in[0,|\xi|_{1}^{-1}],

    (2.38) t[0,(λ¯(r0)(ξ))1]\displaystyle t\in[\hskip 0.7pt0,(\underline{\lambda}^{(-r_{0})}(\xi))^{-1}\hskip 0.7pt] g(tξ)=r0+tμ(r0)(ξ)\displaystyle\iff\ g(t\xi)=r_{0}+t\mu^{(-r_{0})}(\xi)
    t[0,(λ¯(ξ))1]\displaystyle t\in[\hskip 0.7pt0,(\underline{\lambda}(\xi))^{-1}\hskip 0.7pt] go(tξ)=tμ(ξ).\displaystyle\iff\ g^{o}(t\xi)=t\mu(\xi).
  2. (ii)

    For t[0,|ξ|11]t\in[0,|\xi|_{1}^{-1}],

    (2.39) t[0,(λ¯(ξ))1)\displaystyle t\in\bigl{[}\hskip 0.7pt0,(\hskip 0.7pt\overline{\lambda}(\xi))^{-1}\hskip 0.7pt\bigr{)} g(tξ)>go(tξ)\displaystyle\iff\ g(t\xi)>g^{o}(t\xi)
    t[(λ¯(ξ))1,|ξ|11]\displaystyle t\in[\hskip 0.7pt(\overline{\lambda}(\xi))^{-1},|\xi|_{1}^{-1}] g(tξ)=go(tξ).\displaystyle\iff\ g(t\xi)=g^{o}(t\xi).
  3. (iii)

    Strengthen the subcriticality assumption to (2.6). The function tg(tξ)t\mapsto g^{\diamond}(t\xi) is continuously differentiable on the open interval (0,|ξ|11)(0,|\xi|_{1}^{-1}) and limt|ξ|11(g)(tξ)=+\lim_{t\nearrow|\xi|_{1}^{-1}}(g^{\diamond})^{\prime}(t\xi)=+\infty. If g(ξ/|ξ|1)<g^{\diamond}(\xi/|\xi|_{1})<\infty then the left derivative of tg(tξ)t\mapsto g^{\diamond}(t\xi) at t=|ξ|11t=|\xi|_{1}^{-1} exists and equals ++\infty.

Notice that the right-hand sides in (2.38) agree if and only if r0=0r_{0}=0, as is consistent with the agreement g=gog=g^{o} when r0=0r_{0}=0. From (2.39) and (2.35) we read that if r0>0r_{0}>0, the set {g>go}\{g>g^{o}\} is an open neighborhood of 𝟎\mathbf{0} that consists of finite rays from the origin, while its complement {g=go}\{g=g^{o}\} contains the nonempty annulus {ζ𝒰:(1+D)1|ζ|11}\{\zeta\in\mathcal{U}:(1+D)^{-1}\leq|\zeta|_{1}\leq 1\}, where DD is the constant in (2.35). Another consequence of (2.38) and (2.39) is that gog^{o} is never strictly between μ\mu and gg but always agrees with at least one of them.

By Lemma D.1 in Appendix D, the differentiability property in part (iii) can be equivalently stated in geometric terms as follows: for ξ(int𝒰){𝟎}\xi\in(\operatorname{int}\mathcal{U})\setminus\{\mathbf{0}\}, the subdifferential g(ξ)\partial g^{\diamond}(\xi) lies on a hyperplane perpendicular to ξ\xi.

2.4. Duality of the weight shift and geodesic length

This section develops the duality between the weight shift variable bb in ωω(b)\omega\mapsto\omega^{(b)} and the path-length variable α\alpha in the limit shapes (2.29). Nonnegative weights (r00r_{0}\geq 0) are assumed throughout.

Fix ξd{𝟎}\xi\in\mathbb{R}^{d}\hskip 0.7pt\setminus\{\mathbf{0}\} for the duration of this section. We restrict the shape function μξ(b)\mu_{\xi}(b) of (2.9) to shifts br0b\geq-r_{0} that preserve the nonnegativity of the weights and then extend it to an upper semicontinuous concave function on all of \mathbb{R} by setting

(2.40) μ¯ξ(b)={μξ(b)=μ(b)(ξ),br0,b<r0.\overline{\mu}_{\xi}(b)=\begin{cases}\mu_{\xi}(b)=\mu^{(b)}(\xi),&b\geq-r_{0}\\ -\infty,&b<-r_{0}.\end{cases}

To emphasize, the function μ¯ξ(b)\overline{\mu}_{\xi}(b) drops the extension to b(r0ε0,r0)b\in(-r_{0}-\varepsilon_{0},-r_{0}) done in Theorem 2.1. The reason for this choice is that developing the duality for shifts b<r0b<-r_{0} requires a study of the shape function of the self-avoiding version G𝟎,(n),x saG^{\text{\,\rm{sa}}}_{\mathbf{0},(n),x} of restricted path length FPP. This is not undertaken in the present paper and is left as open problem 3.3.

By definition, the concave dual μ¯ξ:[,)\overline{\mu}_{\xi}^{*}:\mathbb{R}\to[-\infty,\infty) is another upper semicontinuous concave function, and together μ¯ξ\overline{\mu}_{\xi} and μ¯ξ\overline{\mu}_{\xi}^{*} satisfy

(2.41) μ¯ξ(α)=infb{αbμ¯ξ(b)}andμ¯ξ(b)=infα{αbμ¯ξ(α)}.\overline{\mu}_{\xi}^{*}(\alpha)=\inf_{b\hskip 0.55pt\in\hskip 0.55pt\mathbb{R}}\{\alpha b-\overline{\mu}_{\xi}(b)\}\quad\text{and}\quad\overline{\mu}_{\xi}(b)=\inf_{\alpha\hskip 0.55pt\in\hskip 0.55pt\mathbb{R}}\{\alpha b-\overline{\mu}_{\xi}^{*}(\alpha)\}.

The superdifferential of the concave function μ¯ξ\overline{\mu}_{\xi} at bb is by definition the set

μ¯ξ(b)={α:μ¯ξ(b)μ¯ξ(b)+α(bb)b}.\partial\overline{\mu}_{\xi}(b)=\{\alpha\in\mathbb{R}:\overline{\mu}_{\xi}(b^{\prime})\leq\overline{\mu}_{\xi}(b)+\alpha(b^{\prime}-b)\ \forall b^{\prime}\in\mathbb{R}\}.

By the definition μ¯ξ(b)=\partial\overline{\mu}_{\xi}(b)=\varnothing for b<r0b<-r_{0}. For b>r0b>-r_{0}, μ¯ξ(b)\partial\overline{\mu}_{\xi}(b) is the bounded closed interval [μ¯ξ(b+),μ¯ξ(b)][\hskip 0.9pt\overline{\mu}_{\xi}^{\prime}(b+),\overline{\mu}_{\xi}^{\prime}(b-)\hskip 0.9pt] and so μ¯ξ(b)={α}\partial\overline{\mu}_{\xi}(b)=\{\alpha\} if and only if μ¯ξ(b)=α\overline{\mu}_{\xi}^{\prime}(b)=\alpha. These general equivalences hold:

α,b:αμ¯ξ(b)μ¯ξ(α)+μ¯ξ(b)=αbbμ¯ξ(α).\forall\alpha,b\in\mathbb{R}:\quad\alpha\in\partial\overline{\mu}_{\xi}(b)\ \Longleftrightarrow\ \overline{\mu}_{\xi}^{*}(\alpha)+\overline{\mu}_{\xi}(b)=\alpha b\ \Longleftrightarrow\ b\in\partial\overline{\mu}_{\xi}^{*}(\alpha).

Theorem 2.17 below establishes the convex duality. The qualitative nature of the (negative of the) dual function in (2.43) is illustrated in Figure 2.2, on the left in the case r0=0r_{0}=0 and on the right in the case r0>0r_{0}>0. In particular, on the left the affine portion of ααg(ξ/α)\alpha\mapsto\alpha g(\xi/\alpha) on the interval [λ¯(b)(ξ),λ¯(b)(ξ)][\hskip 0.9pt\underline{\lambda}^{(b)}(\xi),\overline{\lambda}^{\hskip 0.9pt(b)}(\xi)\hskip 0.7pt] is the dual of the superdifferential μξ(b)\partial\mu_{\xi}(b) in (2.46). The infinite slope at the left edge |ξ|1+|\xi|_{1}+ is the dual of the limit (2.48).

A convenient feature of the restricted path length shape function without zero steps is that it transforms trivially under the weight shift:

(2.42) g(b)(ξ)=g(ξ)+b.g^{(b)}(\xi)=g(\xi)+b.

This and (2.37) applied to μ(b)(ξ)\mu^{(b)}(\xi) give (2.44) below for b>r0b>-r_{0}, which is the basis for the duality.

Original weights with r0=0r_{0}=0|ξ|1|\xi|_{1}α\alphaμ(ξ)\mu(\xi)λ¯(b)(ξ)\underline{\lambda}^{(b)}\!(\xi)λ¯(b)(ξ)\ \ \ \ \overline{\lambda}^{(b)}\!(\xi)λ¯(ξ)\ \ \ \underline{\lambda}(\xi)affine, slope b-bconstant=μ(ξ)\mu(\xi)ααgo(ξ/α)=αg(ξ/α)\alpha\mapsto\alpha g^{o}(\xi/\alpha)=\alpha g(\xi/\alpha)Shifted weights ω(b)\omega^{(b)} with b>r0=0b>-r_{0}=0|ξ|1|\xi|_{1}α\alphaμ(b)(ξ)\mu^{(b)}(\xi)λ¯(b)(ξ)\underline{\lambda}^{(b)}\!(\xi)λ¯(b)(ξ)\ \ \ \ \overline{\lambda}^{\hskip 0.9pt(b)}\!(\xi)λ¯(ξ)\ \ \ \underline{\lambda}(\xi)thick=thinthick=thin=μ(b)(ξ)\mu^{(b)}(\xi)thick affine, slope bbthin=μ(b)(ξ)\mu^{(b)}(\xi)thin: αα(go)(b)(ξ/α)\alpha\mapsto\alpha(g^{o})^{(b)}(\xi/\alpha)thick: ααg(b)(ξ/α)\alpha\mapsto\alpha g^{(b)}(\xi/\alpha)
Figure 2.2. Fix ξd{𝟎}\xi\in\mathbb{R}^{d}\hskip 0.55pt\setminus\{\mathbf{0}\}. Left: Graphs of the functions ααμ(ξ/α)=μ(ξ)\alpha\mapsto\alpha\mu(\xi/\alpha)=\mu(\xi) and ααgo(ξ/α)=αg(ξ/α)\alpha\mapsto\alpha g^{o}(\xi/\alpha)=\alpha g(\xi/\alpha) in the case r0=0r_{0}=0. All three agree from λ¯(ξ)\underline{\lambda}(\xi) onwards to λ¯(ξ)=\overline{\lambda}(\xi)=\infty. Right: Graphs of αμ(b)(ξ/α)=μ(b)(ξ)\alpha\mu^{(b)}(\xi/\alpha)=\mu^{(b)}(\xi), α(go)(b)(ξ/α)\alpha(g^{o})^{(b)}(\xi/\alpha) and αg(b)(ξ/α)\alpha g^{(b)}(\xi/\alpha) for the weights shifted by b>r0=0b>-r_{0}=0. The labeling of the α\alpha-axis is the same in both figures. As the weights shift to higher values, the shape functions move up. In particular, the thick graph ααg(b)(ξ/α)\alpha\mapsto\alpha g^{(b)}(\xi/\alpha) on the right is obtained by adding the function αbα\alpha\mapsto b\alpha to the graph on the left. On the possibly degenerate interval [λ¯(b)(ξ),λ¯(b)(ξ)][\hskip 0.9pt\underline{\lambda}^{(b)}(\xi),\overline{\lambda}^{\hskip 0.9pt(b)}(\xi)\hskip 0.7pt] we have the triple coincidence α(go)(b)(ξ/α)=αg(b)(ξ/α)=μ(b)(ξ)\alpha(g^{o})^{(b)}(\xi/\alpha)=\alpha g^{(b)}(\xi/\alpha)=\mu^{(b)}(\xi) and after that αg(b)(ξ/α)\alpha g^{(b)}(\xi/\alpha) separates from the other two. As bb increases, the interval [λ¯(b)(ξ),λ¯(b)(ξ)][\hskip 0.9pt\underline{\lambda}^{(b)}(\xi),\overline{\lambda}^{\hskip 0.9pt(b)}(\xi)\hskip 0.7pt] moves to the left, without overlaps, approaching |ξ|1|\xi|_{1} as bb\nearrow\infty. In both pictures, at the left endpoint |ξ|1+|\xi|_{1}+ the graphs coming from gog^{o} and gg have slope -\infty. The three regions [|ξ|1,λ¯(b)(ξ))[\hskip 0.9pt|\xi|_{1},\hskip 0.9pt\underline{\lambda}^{(b)}(\xi)), [λ¯(b)(ξ),λ¯(b)(ξ)][\hskip 0.9pt\underline{\lambda}^{(b)}(\xi),\overline{\lambda}^{\hskip 0.9pt(b)}(\xi)\hskip 0.7pt] and (λ¯(b)(ξ),)(\,\overline{\lambda}^{\hskip 0.9pt(b)}(\xi),\infty) of qualitatively distinct behavior in the diagram on the right are described in Proposition 4.4.
Theorem 2.17.

Assume r00r_{0}\geq 0, (2.6), and the moment bound (2.7) with p=dp=d. Fix ξd{𝟎}\xi\in\mathbb{R}^{d}\hskip 0.7pt\setminus\{\mathbf{0}\}.

  1. (i)

    The concave dual of μ¯ξ\overline{\mu}_{\xi} is

    (2.43) μ¯ξ(α)={αg(ξ/α),α|ξ|1,α<|ξ|1.\overline{\mu}_{\xi}^{*}(\alpha)=\begin{cases}-\alpha\hskip 0.55ptg(\xi/\alpha),&\alpha\geq|\xi|_{1}\\ -\infty,&\alpha<|\xi|_{1}.\end{cases}

    In particular, we have the identities

    (2.44) μ¯ξ(b)=infα|ξ|1αg(b)(ξ/α)=infα|ξ|1{αg(ξ/α)+αb} for b,\overline{\mu}_{\xi}(b)=\inf_{\alpha\geq|\xi|_{1}}\alpha g^{(b)}(\xi/\alpha)=\inf_{\alpha\geq|\xi|_{1}}\{\alpha\hskip 0.55ptg(\xi/\alpha)+\alpha b\}\quad\text{ for }\ b\in\mathbb{R},

    and

    (2.45) αg(ξ/α)=supbr0{μ¯ξ(b)αb} for α|ξ|1.\alpha\hskip 0.55ptg(\xi/\alpha)=\sup_{b\geq-r_{0}}\{\overline{\mu}_{\xi}(b)-\alpha b\}\quad\text{ for }\ \alpha\geq|\xi|_{1}.
  2. (ii)

    For b>r0b>-r_{0}, the superdifferential μ¯ξ(b)\partial\overline{\mu}_{\xi}(b) is the compact interval

    (2.46) μ¯ξ(b)=[μ¯ξ(b+),μ¯ξ(b)]=[λ¯(b)(ξ),λ¯(b)(ξ)]\partial\overline{\mu}_{\xi}(b)=[\hskip 0.7pt\overline{\mu}_{\xi}^{\prime}(b+),\overline{\mu}_{\xi}^{\prime}(b-)]=[\,\underline{\lambda}^{(b)}(\xi),\overline{\lambda}^{\hskip 0.9pt(b)}(\xi)\,]

    while

    (2.47) μ¯ξ(r0)=[μ¯ξ((r0)+),)=[λ¯(r0)(ξ),λ¯(r0)(ξ)).\partial\overline{\mu}_{\xi}(-r_{0})=[\,\overline{\mu}_{\xi}^{\prime}((-r_{0})+),\infty)=\bigl{[}\,\underline{\lambda}^{(-r_{0})}(\xi),\overline{\lambda}^{\hskip 0.9pt(-r_{0})}(\xi)\bigr{)}.

    Furthermore,

    (2.48) limbμ¯ξ(b±)=|ξ|1.\lim_{b\to\infty}\overline{\mu}_{\xi}^{\prime}(b\pm)=|\xi|_{1}.
Remark 2.18.

(a) Let us make explicit the conversion back to the original FPP shape function μξ(b)=μ(b)(ξ)\mu_{\xi}(b)=\mu^{(b)}(\xi) in Theorem 2.17. In (2.44) μ¯ξ(b)\overline{\mu}_{\xi}(b) can be replaced by μξ(b)\mu_{\xi}(b) for br0b\geq-r_{0}. In each of (2.45), (2.46) and (2.48), μ¯ξ\overline{\mu}_{\xi} can be replaced by μξ\mu_{\xi}. (2.47) cannot be valid for μξ(r0)\partial\mu_{\xi}(-r_{0}) because μξ(b)>\mu_{\xi}(b)>-\infty for some b<r0b<-r_{0}. We do have

(2.49) μξ((r0)+)=μ¯ξ((r0)+)=λ¯(r0)(ξ) but μξ((r0))<=λ¯(r0)(ξ).\mu_{\xi}^{\prime}((-r_{0})+)=\overline{\mu}_{\xi}^{\prime}((-r_{0})+)=\underline{\lambda}^{(-r_{0})}(\xi)\quad\text{ but }\quad\mu_{\xi}^{\prime}((-r_{0})-)<\infty=\overline{\lambda}^{\hskip 0.9pt(-r_{0})}(\xi).

(b) The strict concavity of μ¯ξ\overline{\mu}_{\xi} that was stated in Theorem 2.2 was purposely left out of Theorem 2.17 so that this latter theorem can be proved easily at the end of Section 4, before we turn to the modification arguments. Combining Theorem 2.17 with Theorems 2.2 and 2.3 and (2.35) gives the following. There exists a constant κ<\kappa<\infty that depends on the dimension and the weight distribution such that, for all b>a>r0b>a>-r_{0},

(2.50) |ξ|1<λ¯(b)(ξ)λ¯(b)(ξ)<λ¯(a)(ξ)λ¯(a)(ξ)<λ¯(r0)(ξ)κ|ξ|1<=λ¯(r0)(ξ).|\xi|_{1}<\underline{\lambda}^{(b)}(\xi)\leq\overline{\lambda}^{\hskip 0.9pt(b)}(\xi)<\underline{\lambda}^{(a)}(\xi)\leq\overline{\lambda}^{\hskip 0.9pt(a)}(\xi)<\underline{\lambda}^{(-r_{0})}(\xi)\leq\kappa|\xi|_{1}<\infty=\overline{\lambda}^{\hskip 0.9pt(-r_{0})}(\xi).

The strict inequalities above are due to the strict concavity of μξ\mu_{\xi}.

(c) When the infimum r0r_{0} of the support of the weights is zero, gg and gog^{o} coincide (Theorem 2.9 and Lemma 4.1(ii)). Through (2.42) we get an alternative representation of the concave dual in (2.43) in terms of the restricted path FPP shape that admits zero steps:

(2.51) αg(ξ/α)=αg(r0)(ξ/α)+αr0=α(go)(r0)(ξ/α)+αr0.\alpha\hskip 0.55ptg(\xi/\alpha)=\alpha g^{(-r_{0})}(\xi/\alpha)+\alpha r_{0}=\alpha(g^{o})^{(-r_{0})}(\xi/\alpha)+\alpha r_{0}.

We can combine (2.44) and (2.51) into a statement that shows that both gg and (go)(r0)(g^{o})^{(-r_{0})} contain full information for retrieving all the shifts of μ\mu among nonnegative weights:

(2.52) μξ(b)=infα|ξ|1{αg(ξ/α)+αb}=infα|ξ|1{α(go)(r0)(ξ/α)+α(r0+b)} for br0.\mu_{\xi}(b)=\inf_{\alpha\geq|\xi|_{1}}\{\alpha\hskip 0.55ptg(\xi/\alpha)+\alpha b\}=\inf_{\alpha\geq|\xi|_{1}}\{\alpha\hskip 0.7pt(g^{o})^{(-r_{0})}(\xi/\alpha)+\alpha(r_{0}+b)\}\quad\text{ for }b\geq-r_{0}.

(d) Equations (2.33), (2.51), and positive homogeneity of λ¯\underline{\lambda} and μ\mu show that ααg(ξ/α)\alpha\mapsto\alpha\hskip 0.55ptg(\xi/\alpha) is affine for large α\alpha:

(2.53) αg(ξ/α)=μ(r0)(ξ)+αr0for αλ¯(r0)(ξ).\alpha\hskip 0.55ptg(\xi/\alpha)=\mu^{(-r_{0})}(\xi)+\alpha r_{0}\quad\text{for $\alpha\geq\underline{\lambda}^{(-r_{0})}(\xi)$.}

The reader can recognize this statement as the dual version of μ¯ξ(r0)=[λ¯(r0)(ξ),)\partial\overline{\mu}_{\xi}(-r_{0})=[\,\underline{\lambda}^{(-r_{0})}(\xi),\infty) from the theorem above, and an immediate consequence of (2.38). This affine portion of αg(ξ/α)\alpha\hskip 0.55ptg(\xi/\alpha) is visible in both diagrams of Figure 2.2.

Identities (2.51) and (2.53) suggest that, for αλ¯(r0)(ξ)\alpha\geq\underline{\lambda}^{(-r_{0})}(\xi), the recipe for an optimal path of length approximately nαn\alpha from 𝟎\mathbf{0} to a point close to nξn\xi is this: shift the weights so that their infimum is zero and take the optimal path for the shifted weights ω(r0)\omega^{(-r_{0})}. In particular, once α\alpha is above the FPP geodesic length, we can follow the FPP geodesic of the shifted weights ω(r0)\omega^{(-r_{0})} and extend the path to length nαn\alpha by finding and repeating an edge whose weight is close to the minimum r0r_{0}. \triangle

3. Open problems

We list here open problems raised by the results.

3.1. Asymptotic length of geodesics

Does the Hammersley-Welsh limit generalize in some natural way when μξ(b+)<μξ(b)\mu_{\xi}^{\prime}(b+)<\mu_{\xi}^{\prime}(b-)? For example, are there weight configurations ω\omega and ω~\widetilde{\omega} and sequences xn/nξx_{n}/n\to\xi and x~n/nξ\widetilde{x}_{n}/n\to\xi such that

(3.1) lim¯nL¯𝟎,xn(b)(ω)n=μξ(b+)andlim¯nL¯𝟎,x~n(b)(ω~)n=μξ(b)?\varliminf_{n\to\infty}\frac{\underline{L}^{(b)}_{\hskip 0.9pt\mathbf{0},\hskip 0.9ptx_{n}}(\omega)}{n}=\mu_{\xi}^{\prime}(b+)\qquad\text{and}\qquad\varlimsup_{n\to\infty}\frac{\overline{L}^{\hskip 0.7pt(b)}_{\hskip 0.9pt\mathbf{0},\hskip 0.9pt\widetilde{x}_{n}}(\widetilde{\omega})}{n}=\mu_{\xi}^{\prime}(b-)\,?

If so, can these statements be strengthened to limits, and are they valid for all sequences and almost surely? Even if one cannot know the limits, are the random variables lim¯nn1L¯𝟎,xn\varliminf_{n\to\infty}n^{-1}{\underline{L}_{\hskip 0.9pt\mathbf{0},\hskip 0.9ptx_{n}}} and lim¯nn1L¯𝟎,xn\varlimsup_{n\to\infty}n^{-1}{\overline{L}_{\hskip 0.9pt\mathbf{0},\hskip 0.9ptx_{n}}} almost surely constant?

3.2. Properties of the shape functions

Is μξ\mu_{\xi} differentiable when the weight distribution is continuous? What about the case of a single positive atom which is not covered by Theorems 2.62.7? Is any comparison between μξ\mu_{\xi} and μξ~\mu_{\widetilde{\xi}} possible for two distinct directions ξ\xi and ξ~\widetilde{\xi}? Is the function λ¯\underline{\lambda} defined in (2.33) a norm on d\mathbb{R}^{d}? Do λ¯\underline{\lambda} and λ¯\overline{\lambda} possess more regularity than given in Theorem 2.11(ii)?

3.3. Duality of the weight shift and geodesic length for real-valued weights

The duality described in Section 2.4 restricted the shape function μξ(b)\mu_{\xi}(b) to nonnegative weights through definition (2.40). This leaves open the duality of μξ(b)\mu_{\xi}(b) for b<r0b<-r_{0}. To capture the full convex duality over all shifts bb requires a study of the process G𝟎,(n),x saG^{\text{\,\rm{sa}}}_{\mathbf{0},(n),x}, restricted path length FPP that optimizes over self-avoiding paths, in a manner analogous to our study of G𝟎,(n),xG_{\mathbf{0},(n),x} and its shape function.

The present shortcoming can be seen for example in the case r0=0r_{0}=0 of (2.35) where λ¯(ξ)\overline{\lambda}(\xi) blows up and cannot capture the left derivative μξ(0)\mu_{\xi}^{\prime}(0-). Graphically this same phenomenon appears in the left diagram of Figure 2.2 where the graph of αg(ξ/α)\alpha g(\xi/\alpha) never separates from μ(ξ)\mu(\xi) after λ¯(ξ)\underline{\lambda}(\xi). The graph of the function αg sa(ξ/α)\alpha g^{\text{\rm{\,sa}}}(\xi/\alpha) of the self-avoiding version will separate from μ(ξ)\mu(\xi) for large enough α\alpha and capture μξ(0)\mu_{\xi}^{\prime}(0-).

3.4. Modification arguments for real weights

Do the van den Berg-Kesten modification arguments [20] extend to weights that can take negative values? Such an extension would permit the extension of the strict concavity of μξ(b)\mu_{\xi}(b) to b<r0b<-r_{0}.

3.5. General perturbations of weights

Develop versions of our results for other perturbations of the weights, besides the simple shift t(h)(e)=t(e)+ht^{(h)}(e)=t(e)+h, such as the perturbations considered in [3].

4. The shape functions and lengths of optimal paths

This section develops soft auxiliary results required for the main results of Section 2. Along the way we prove Theorem 2.10, part (i) of Theorem 2.11, parts (i)(ii) of Theorem 2.16, and Theorem 2.17. To begin, assume r0>r_{0}>-\infty and the moment bound (2.7) with p=dp=d for the nonnegative weights t+(e)=t(e)0t^{+}(e)=t(e)\vee 0. Take the existence of the continuous, convex functions g,go:int𝒰[r00,)g,g^{o}:\operatorname{int}\mathcal{U}\to[r_{0}\wedge 0,\infty) that satisfy the laws of large numbers (2.29) from Theorem B.1 in Appendix B. The limit implies ggog\geq g^{o}. Extend the shape functions gg and gog^{o} to all of 𝒰\mathcal{U} through radial limits: for {𝚎𝚖𝚙𝚝𝚢,o}\diamond\in\{\langle{\tt empty}\rangle,o\}, define

(4.1) g(ξ)=limt1g(tξ)[r00,]for |ξ|1=1.g^{\diamond}(\xi)=\lim_{t\nearrow 1}g^{\diamond}(t\xi)\;\in\;[r_{0}\wedge 0,\infty]\qquad\text{for }\ |\xi|_{1}=1.

The limit exists because tg(tξ)t\mapsto g^{\diamond}(t\xi) is a convex function on the interval [0,1)[0,1). Monotonicity (2.36), ggog\geq g^{o}, and the limit combine to give, for |ξ|1τα|\xi|_{1}\leq\tau\leq\alpha,

(4.2) αgo(ξ/α)τgo(ξ/τ)τg(ξ/τ).\alpha\hskip 0.55ptg^{o}({\xi}/{\alpha})\leq\tau g^{o}({\xi}/{\tau})\leq\tau g({\xi}/{\tau}).

Part (iv) of the next lemma proves Theorem 2.10.

Lemma 4.1.

Assume r0>r_{0}>-\infty and the moment bound (2.7) with p=dp=d for the nonnegative weights t+(e)=t(e)0t^{+}(e)=t(e)\vee 0. The restricted path shape functions have the following properties.

  1. (i)

    g(𝟎)=r0g(\mathbf{0})=r_{0} and go(𝟎)=r00g^{o}(\mathbf{0})=r_{0}\wedge 0.

  2. (ii)

    If r00r_{0}\leq 0 then g=gog=g^{o} on all of 𝒰\mathcal{U}. If r0>0r_{0}>0 then g>gog>g^{o} in an open neighborhood of the origin.

  3. (iii)

    For all ξd{𝟎}\xi\in\mathbb{R}^{d}\hskip 0.7pt\setminus\{\mathbf{0}\} and α|ξ|1\alpha\geq|\xi|_{1},

    (4.3) αgo(ξα)=infτ:|ξ|1τατg(ξτ)\alpha\hskip 0.55ptg^{o}\biggl{(}\frac{\xi}{\alpha}\biggr{)}=\inf_{\tau:\,|\xi|_{1}\leq\tau\leq\alpha}\tau g\biggl{(}\frac{\xi}{\tau}\biggr{)}

    and the infimum on the right is attained at some τ[|ξ|1,α]\tau\in[\,|\xi|_{1},\alpha]. In particular, |ξ|1=1|\xi|_{1}=1 implies go(ξ)=g(ξ)g^{o}(\xi)=g(\xi).

  4. (iv)

    For {𝚎𝚖𝚙𝚝𝚢,o}\diamond\in\{\langle{\tt empty}\rangle,o\}, the extended function gg^{\diamond} is convex and lower semicontinuous on 𝒰\mathcal{U}.

Proof.

Part (i). The lower bounds gr0g\geq r_{0} and gor00g^{o}\geq r_{0}\wedge 0 on all of int𝒰\operatorname{int}\mathcal{U} follow from t(e)r0t(e)\geq r_{0} and t({x,x})=0t(\{x,x\})=0. Also immediate is go(𝟎)0g^{o}(\mathbf{0})\leq 0. Given ε>0\varepsilon>0, we can fix as measurable functions of almost every ω\omega,

(4.4) an edge e={x,y}e=\{x,y\} such that t(e)<r0+εt(e)<r_{0}+\varepsilon, and a path π\pi from 𝟎\mathbf{0} to xx.

For large enough nn consider paths x0:nx_{0\hskip 0.6pt{\boldsymbol{:}}\hskip 0.3ptn} that follow π\pi and then repeat edge ee n|π|n-|\pi| times. Then xn/n0x_{n}/n\to 0 and in the limit go(𝟎)g(𝟎)r0+εg^{o}(\mathbf{0})\leq g(\mathbf{0})\leq r_{0}+\varepsilon.

Part (ii). The claim for r00r_{0}\leq 0 is true because the zero steps of a path πnΠ𝟎,(n),xo\pi_{n}\in\Pi^{o}_{\hskip 0.55pt\mathbf{0},(n),x} can be replaced by repetitions of an edge with weight close to r0r_{0}. Here is a detailed proof.

Fix ξint𝒰\xi\in\operatorname{int}\mathcal{U} and a sequence xn𝒟nx_{n}\in\mathcal{D}_{n} such that xn/nξx_{n}/n\to\xi. Let πn\pi_{n} be an optimal path for G𝟎,(n),xnoG^{o}_{\mathbf{0},(n),x_{n}} and let knk_{n} be the number of zero steps in πn\pi_{n}. Let ee and π\pi be as in (4.4). We construct an \mathcal{R}-admissible path πn\pi_{n}^{\prime} of length nn from 𝟎\mathbf{0} to xnx_{n} or xn+𝐞1x_{n}+\mathbf{e}_{1} that repeats edge ee as many times as possible, as follows.

  • First, if knk_{n} is even, let yn=xny_{n}=x_{n}, and if knk_{n} is odd, let yn=xn+𝐞1y_{n}=x_{n}+\mathbf{e}_{1}. Then πn\pi_{n} (plus the ynxny_{n}-x_{n} step if necessary) goes from 𝟎\mathbf{0} to yny_{n} in n2kn/2n-2\lfloor{k_{n}/2}\rfloor nonzero steps.

  • The remaining 2kn/22\lfloor{k_{n}/2}\rfloor steps are spent in an initial segment from 𝟎\mathbf{0} back to 𝟎\mathbf{0} by first following π\pi to xx, then back and forth across ee altogether 2(kn/2|π|)+2(\lfloor{k_{n}/2}\rfloor-|\pi|)^{+} times, and then from xx back to 𝟎\mathbf{0} along π\pi (in reverse direction). If kn/2|π|\lfloor{k_{n}/2}\rfloor\leq|\pi| then the initial segment does not go all the way to xx but turns back towards 𝟎\mathbf{0} after kn/2\lfloor{k_{n}/2}\rfloor steps along π\pi.

Let e1,,eme_{1},\dotsc,e_{m} denote the edges of π\pi. We get the following bound:

(4.5) G𝟎,(n),yn\displaystyle G_{\mathbf{0},(n),y_{n}} T(πn)=2i=1mkn/2t(ei)+2(kn/2|π|)+t(e)+T(πn)+t({xn,yn})\displaystyle\leq T(\pi_{n}^{\prime})=2\sum_{i=1}^{m\wedge\lfloor{k_{n}/2}\rfloor}t(e_{i})+2\bigl{(}\lfloor{k_{n}/2}\rfloor-|\pi|\bigr{)}^{+}t(e)+T(\pi_{n})+t(\{x_{n},y_{n}\})
2i=1mt+(ei)+2(kn/2|π|)+(r0+ε)+G𝟎,(n),xno+t+({xn,xn+𝐞1}).\displaystyle\leq 2\sum_{i=1}^{m}t^{+}(e_{i})+2\bigl{(}\lfloor{k_{n}/2}\rfloor-|\pi|\bigr{)}^{+}(r_{0}+\varepsilon)+G^{o}_{\mathbf{0},(n),x_{n}}+t^{+}(\{x_{n},x_{n}+\mathbf{e}_{1}\}).

Divide through by nn and let nn\to\infty along a suitable subsequence, utilizing r00r_{0}\leq 0 and yn/nξy_{n}/n\to\xi. We obtain

g(ξ)ε+go(ξ)+lim¯nn1t+({xn,xn+𝐞1}).g(\xi)\leq\varepsilon+g^{o}(\xi)+\varliminf_{n\to\infty}n^{-1}t^{+}(\{x_{n},x_{n}+\mathbf{e}_{1}\}).

The last term vanishes almost surely because n1t+({xn,xn+𝐞1})0n^{-1}t^{+}(\{x_{n},x_{n}+\mathbf{e}_{1}\})\to 0 in probability. Since ggog\geq g^{o} always, letting ε0\varepsilon\searrow 0 establishes the equality g=gog=g^{o} under r00r_{0}\leq 0.

The statement for r0>0r_{0}>0 in Part (ii) follows from Part (i) and continuity.

Part (iii). For r00r_{0}\leq 0 (4.3) follows from go=gg^{o}=g and (4.2).

Assume r0>0r_{0}>0. The inequalities in (4.2) imply that \leq holds in (4.3). To prove the opposite inequality \geq in (4.3), consider first α>|ξ|1\alpha>|\xi|_{1} so that we can take advantage of the laws of large numbers. Choose knk_{n}\to\infty and xn𝒟knox_{n}\in\mathcal{D}^{o}_{k_{n}} so that kn/nαk_{n}/n\to\alpha, |xn|1|x_{n}|_{1}\to\infty and xn/nξx_{n}/n\to\xi. Begin with

G𝟎,(kn),xno=minj:|xn|1jknG𝟎,(j),xn.G^{o}_{\mathbf{0},(k_{n}),x_{n}}=\min_{j:\,|x_{n}|_{1}\leq j\leq k_{n}}G_{\mathbf{0},(j),x_{n}}.

Let ε>0\varepsilon>0 and choose a partition |ξ|1=τ0<τ1<<τm=α|\xi|_{1}=\tau_{0}<\tau_{1}<\dotsm<\tau_{m}=\alpha such that τiτi1<ε\tau_{i}-\tau_{i-1}<\varepsilon. Choose integers n,i\ell_{n,i} such that |xn|1=n,0<n,1<<n,m|x_{n}|_{1}=\ell_{n,0}<\ell_{n,1}<\dotsm<\ell_{n,m}, n,mkn\ell_{n,m}\geq k_{n}, n,i/nτi\ell_{n,i}/n\to\tau_{i} and xn𝒟n,ix_{n}\in\mathcal{D}_{\ell_{n,i}}. (When n,i>|xn|1\ell_{n,i}>|x_{n}|_{1}, xn𝒟n,ix_{n}\in\mathcal{D}_{\ell_{n,i}} only requires n,i\ell_{n,i} to have the right parity.) Then

G𝟎,(kn),xnomin1imminn,i1jn,iG𝟎,(j),xnmin1imG𝟎,(n,i),xn2T(π)2nε(r0+ε)\displaystyle G^{o}_{\mathbf{0},(k_{n}),x_{n}}\;\geq\;\min_{1\leq i\leq m}\;\min_{\ell_{n,i-1}\leq j\leq\ell_{n,i}}G_{\mathbf{0},(j),x_{n}}\;\geq\;\min_{1\leq i\leq m}G_{\mathbf{0},(\ell_{n,i}),x_{n}}-2T(\pi)-2n\varepsilon(r_{0}+\varepsilon)

where we again utilize (4.4): for n,i1jn,i\ell_{n,i-1}\leq j\leq\ell_{n,i} whenever xn𝒟jx_{n}\in\mathcal{D}_{j}, construct an n,i\ell_{n,i}-path from 𝟎\mathbf{0} to xnx_{n} by first going from 𝟎\mathbf{0} to one endpoint of ee, repeating ee as many times as needed, returning to 𝟎\mathbf{0}, and then following an optimal jj-path from 𝟎\mathbf{0} to xnx_{n}. (If n,ij\ell_{n,i}-j is too small to allow travel all the way to ee, then proceed part of the way and return to 𝟎\mathbf{0}. n,ij\ell_{n,i}-j is even because xn𝒟n,i𝒟jx_{n}\in\mathcal{D}_{\ell_{n,i}}\cap\mathcal{D}_{j}.) The number of repetitions of ee is at most 2nε2n\varepsilon when nn is large enough.

In the limit

αgo(ξ/α)min1imτig(ξ/τi)2ε(r0+ε)infτ:|ξ|1τατg(ξ/τ)2ε(r0+ε).\alpha\hskip 0.55ptg^{o}({\xi}/{\alpha})\geq\min_{1\leq i\leq m}\tau_{i}\hskip 0.55ptg(\xi/\tau_{i})-2\varepsilon(r_{0}+\varepsilon)\geq\inf_{\tau:\,|\xi|_{1}\leq\tau\leq\alpha}\tau g({\xi}/{\tau})-2\varepsilon(r_{0}+\varepsilon).

Let ε0\varepsilon\searrow 0 to complete the proof of (4.3) in the case α>|ξ|1\alpha>|\xi|_{1}. The infimum in (4.3) is attained because on the right either the extended function is continuous down to τ=|ξ|1\tau=|\xi|_{1} or then it blows up to \infty.

To complete the proof of (4.3) we show that go(ξ)=g(ξ)g^{o}(\xi)=g(\xi) when |ξ|1=1|\xi|_{1}=1. Only go(ξ)g(ξ)g^{o}(\xi)\geq g(\xi) needs proof. Let c<g(ξ)c<g(\xi). Since gr0>0g\geq r_{0}>0 we can assume c>0c>0. Pick u<1u<1 so that g(sξ)>cg(s\xi)>c for s[u,1]s\in[u,1]. Then by (4.3) applied to the case α>1\alpha>1, for t[u,1)t\in[u,1) we have

go(tξ)=tinfs[t,1]g(sξ)stc.g^{o}(t\xi)=t\cdot\inf_{s\hskip 0.7pt\in\hskip 0.7pt[\hskip 0.9ptt\hskip 0.7pt,\hskip 0.9pt1]}\frac{g(s\xi)}{s}\geq t\hskip 0.7ptc.

Letting t1t\nearrow 1 gives go(ξ)cg^{o}(\xi)\geq c.

Part (iv). Convexity extends readily to all of 𝒰\mathcal{U}. If ξ=αξ+(1α)ξ′′\xi=\alpha\xi^{\prime}+(1-\alpha)\xi^{\prime\prime} in 𝒰\mathcal{U}, then for 0<t<10<t<1 convexity on int𝒰\operatorname{int}\mathcal{U} gives g(tξ)αg(tξ)+(1α)g(tξ′′)g^{\diamond}(t\xi)\leq\alpha g^{\diamond}(t\xi^{\prime})+(1-\alpha)g^{\diamond}(t\xi^{\prime\prime}) and we can let t1t\nearrow 1.

We check the lower semicontinuity of the extension gog^{o} on 𝒰\mathcal{U}. Since gog^{o} is continuous in the interior, we need to consider only limits to the boundary. Let |ξ|1=1|\xi|_{1}=1, go(ξ)>cg^{o}(\xi)>c and ξjξ\xi_{j}\to\xi in 𝒰\mathcal{U}. By the limit in (4.1) we can pick t<1t<1 so that t1go(tξ)>ct^{-1}g^{o}(t\xi)>c. By the continuity of gog^{o} on int𝒰\operatorname{int}\mathcal{U}, go(tξj)go(tξ)g^{o}(t\xi_{j})\to g^{o}(t\xi). Pick j0j_{0} so that t1go(tξj)>ct^{-1}g^{o}(t\xi_{j})>c for jj0j\geq j_{0}. Apply (4.2) to ξj\xi_{j} with α=t1\alpha=t^{-1} and τ=1\tau=1 to get go(ξj)t1go(tξj)>cg^{o}(\xi_{j})\geq t^{-1}g^{o}(t\xi_{j})>c, again for all jj0j\geq j_{0}. Lower semicontinuity of gog^{o} has been established.

Lower semicontinuity of gg follows from ggog\geq g^{o} and the equality g=gog=g^{o} on the boundary: when |ξ|1=1|\xi|_{1}=1 and ξjξ\xi_{j}\to\xi in 𝒰\mathcal{U}, lim¯jg(ξj)lim¯jgo(ξj)go(ξ)=g(ξ).\varliminf_{j\to\infty}g(\xi_{j})\geq\varliminf_{j\to\infty}g^{o}(\xi_{j})\geq g^{o}(\xi)=g(\xi).


In the remainder of this section we investigate the connections of gog^{o} and gg with standard FPP and assume r00r_{0}\geq 0 and either (2.3) or (2.6). We begin with the fact that μ\mu and gog^{o} coincide in a neighborhood of the origin. Since the next lemma considers nonnegative weights without any shifts, the weaker subcriticality assumption (2.3) is sufficient.

Lemma 4.2.

Assume r00r_{0}\geq 0, (2.3), and the moment bound (2.7) with p=dp=d. Then there exists a constant κ(1,)\kappa\in(1,\infty) and a positively homogeneous function λ¯:d+\underline{\lambda}:\mathbb{R}^{d}\to\mathbb{R}_{+} such that |ξ|1λ¯(ξ)κ|ξ|1|\xi|_{1}\leq\underline{\lambda}(\xi)\leq\kappa|\xi|_{1} ξd\forall\xi\in\mathbb{R}^{d} and

(4.6) for ξ𝒰,μ(ξ)=go(ξ)λ¯(ξ)1.\displaystyle\text{for $\xi\in\mathcal{U}$,}\quad\mu(\xi)=g^{o}(\xi)\;\Longleftrightarrow\;\underline{\lambda}(\xi)\leq 1.

In particular, μ(ξ)=go(ξ)\mu(\xi)=g^{o}(\xi) in the neighborhood {ξd:|ξ|1κ1}\{\xi\in\mathbb{R}^{d}:|\xi|_{1}\leq\kappa^{-1}\} of the origin.

Proof.

We claim that there exists a constant κ(1,)\kappa\in(1,\infty) such that

(4.7) ξd{𝟎}:μ(ξ/α)=go(ξ/α) for ακ|ξ|1.\forall\xi\in\mathbb{R}^{d}\hskip 0.7pt\setminus\{\mathbf{0}\}:\ \mu(\xi/\alpha)=g^{o}(\xi/\alpha)\text{ for }\alpha\geq\kappa|\xi|_{1}.

It suffices to prove that a constant κ\kappa works for all |ξ|1=1|\xi|_{1}=1. Towards this end we show the existence of a deterministic constant κ\kappa and a random constant M1M_{1} such that

(4.8) L¯𝟎,x12κ|x|1 for all |x|1M1.\overline{L}_{\hskip 0.9pt\mathbf{0},x}\leq\tfrac{1}{2}\kappa|x|_{1}\quad\text{ for all }\quad|x|_{1}\geq M_{1}.

By Kesten’s foundational estimate (Proposition 5.8 in [14], also Lemma 4.5 in [2]), valid under the subcriticality assumption (2.3), there are positive constants δ,c1\delta,c_{1} such that, for all kk\in\mathbb{N},

(4.9) ( self-avoiding path γ such that 𝟎γ|γ|k, and T(γ)kδ)ec1k.\mathbb{P}\bigl{(}\exists\text{ self-avoiding path $\gamma$ such that $\mathbf{0}\in\gamma$, $|\gamma|\geq k$, and $T(\gamma)\leq k\delta$}\bigr{)}\leq e^{-c_{1}k}.

By adding these probabilities over the cases |γ|=kn|\gamma|=k\geq n we get

(4.10) { self-avoiding path γ from the origin with |γ|n and T(γ)δ|γ|}Cec1n.\mathbb{P}\bigl{\{}\hskip 0.7pt\text{$\exists$ self-avoiding path $\gamma$ from the origin with $|\gamma|\geq n$ and $T(\gamma)\leq\delta|\gamma|$}\hskip 0.7pt\bigr{\}}\leq Ce^{-c_{1}n}.

Thus there exists a random constant M1M_{1} such that any self-avoiding path γ\gamma from the origin of length |γ|M1|\gamma|\geq M_{1} satisfies T(γ)δ|γ|T(\gamma)\geq\delta|\gamma|.

Since the FPP shape function μ\mu is positively homogeneous, by the FPP shape theorem ([2, p. 11] or (A.8) in Appendix A) we can increase M1M_{1} if necessary so that, for a deterministic constant c2c_{2},

(4.11) T𝟎,xc2|x|1for all |x|1M1.T_{\mathbf{0},x}\leq c_{2}|x|_{1}\quad\text{for all $|x|_{1}\geq M_{1}$.}

Let |x|1M1|x|_{1}\geq M_{1} and let π\pi be a geodesic for T𝟎,xT_{\mathbf{0},x}. Then

δ|π|T(π)=T𝟎,xc2|x|1\delta|\pi|\leq T(\pi)=T_{\mathbf{0},x}\leq c_{2}|x|_{1}

from which |π|(c2/δ)|x|1|\pi|\leq(c_{2}/\delta)|x|_{1}. (4.8) has been verified.


Given ξ\xi such that |ξ|1=1|\xi|_{1}=1, let xn/nξx_{n}/n\to\xi. Then for all large enough nn, L¯𝟎,xnnκ\overline{L}_{\hskip 0.9pt\mathbf{0},\hskip 0.9ptx_{n}}\leq n\kappa. Hence, recalling (2.25),

T𝟎,xn=min|xn|1knκG𝟎,(k),xn=G𝟎,(nκ),xno.T_{\mathbf{0},x_{n}}=\min_{|x_{n}|_{1}\leq k\leq n\kappa}G_{\hskip 0.9pt\mathbf{0},(k),x_{n}}=G^{o}_{\hskip 0.9pt\mathbf{0},(\lfloor{n\kappa}\rfloor),x_{n}}.

In the limit μ(ξ)=κgo(ξ/κ)\mu(\xi)=\kappa g^{o}(\xi/\kappa). (The requirement κ>1\kappa>1 was imposed precisely to justify the limit n1G𝟎,(nκ),xnoκgo(ξ/κ)n^{-1}G^{o}_{\hskip 0.9pt\mathbf{0},(\lfloor{n\kappa}\rfloor),x_{n}}\to\kappa g^{o}(\xi/\kappa).) By the lower bound goμg^{o}\geq\mu and the monotonicity in (4.2), μ(ξ)=αgo(ξ/α)\mu(\xi)=\alpha\hskip 0.55ptg^{o}(\xi/\alpha) for ακ\alpha\geq\kappa. (4.7) has been verified.

Define

(4.12) λ¯(ξ)=inf{α|ξ|1:μ(ξ/α)=go(ξ/α)}.\underline{\lambda}(\xi)=\inf\{\alpha\geq|\xi|_{1}:\mu(\xi/\alpha)=g^{o}(\xi/\alpha)\}.

The claimed properties of the function λ¯\underline{\lambda} follow. ∎

Later in the paper (Corollary 7.2) after much more work we can show that λ¯(ξ)(1+D)|ξ|1\underline{\lambda}(\xi)\geq(1+D)|\xi|_{1}.

In the next lemma we strengthen the subcriticality assumption to (2.6) so that we can apply Lemma 4.2 to the shifted weights ω(r0)\omega^{(-r_{0})} and μ(r0)(ξ)>0\mu^{(-r_{0})}(\xi)>0.

Lemma 4.3.

Assume r00r_{0}\geq 0, (2.6), and the moment bound (2.7) with p=dp=d. For {𝚎𝚖𝚙𝚝𝚢,o}\diamond\in\{\langle{\tt empty}\rangle,o\}, the shape functions gg^{\diamond} have the following properties for a fixed ξd{𝟎}\xi\in\mathbb{R}^{d}\hskip 0.7pt\setminus\{\mathbf{0}\}.

  1. (i)

    On the ξ\xi-directed ray these functions are affine in a nondegenerate interval started from zero: for 0t|ξ|110\leq t\leq|\xi|_{1}^{-1},

    (4.13) t[0,(λ¯(ξ))1]\displaystyle t\in[\hskip 0.7pt0,(\underline{\lambda}(\xi))^{-1}\hskip 0.7pt] go(tξ)=tμ(ξ)\displaystyle\iff\ g^{o}(t\xi)=t\mu(\xi)
    and t[0,(λ¯(r0)(ξ))1]\displaystyle\text{and }\quad t\in[\hskip 0.7pt0,(\underline{\lambda}^{(-r_{0})}(\xi))^{-1}\hskip 0.7pt] g(tξ)=r0+tμ(r0)(ξ).\displaystyle\iff\ g(t\xi)=r_{0}+t\mu^{(-r_{0})}(\xi).
  2. (ii)

    The function tg(tξ)t\mapsto g^{\diamond}(t\xi) is continuous, convex, and strictly increasing for t[0,|ξ|11)t\in[0,|\xi|_{1}^{-1}).

Proof.

Part (i). The first line of (4.13) is exactly (4.6). Shifting weights gives g(ζ)=r0+g(r0)(ζ)g(\zeta)=r_{0}+g^{(-r_{0})}(\zeta) and Lemma 4.1(ii) gives g(r0)(ζ)=(go)(r0)(ζ)g^{(-r_{0})}(\zeta)=(g^{o})^{(-r_{0})}(\zeta). Then the first line of (4.13) applied to ω(r0)\omega^{(-r_{0})} gives the second line.

Part (ii). Continuity and convexity on int𝒰\operatorname{int}\mathcal{U} are already in the construction of the functions gg^{\diamond}. Since μ(ξ)μ(r0)(ξ)>0\mu(\xi)\geq\mu^{(-r_{0})}(\xi)>0 (Theorem 2.1), tg(tξ)t\mapsto g^{\diamond}(t\xi) is strictly increasing on a nondegenerate interval from 0. By convexity, it has to be strictly increasing on the entire interval [0,|ξ|11)[0,|\xi|_{1}^{-1}). ∎

Since the functions ααg(ξ/α)\alpha\mapsto\alpha g^{\diamond}(\xi/\alpha) are central to our treatment, we rewrite (4.6) in this form:

(4.14) for ξd{𝟎} and α|ξ|1,αgo(ξα)=μ(ξ)αλ¯(ξ).\displaystyle\text{for $\xi\in\mathbb{R}^{d}\hskip 0.7pt\setminus\{\mathbf{0}\}$ and $\alpha\geq|\xi|_{1}$,}\quad\alpha\hskip 0.55ptg^{o}\biggl{(}\frac{\xi}{\alpha}\biggr{)}=\mu(\xi)\;\Longleftrightarrow\;\alpha\geq\underline{\lambda}(\xi).

Together with (4.3) the above implies that some τ|ξ|1\tau\geq|\xi|_{1} satisfies τg(ξ/τ)=μ(ξ)\tau g(\xi/\tau)=\mu(\xi). By the μgog\mu\leq g^{o}\leq g inequalities, any such τ\tau must satisfy τλ¯(ξ)\tau\geq\underline{\lambda}(\xi). Now we have

(4.15)  for ξd{𝟎},μ(ξ)=infα:α|ξ|1αg(ξα).\text{ for }\xi\in\mathbb{R}^{d}\hskip 0.7pt\setminus\{\mathbf{0}\},\ \ \mu(\xi)=\inf_{\alpha:\,\alpha\geq|\xi|_{1}}\alpha\hskip 0.55ptg\biggl{(}\frac{\xi}{\alpha}\biggr{)}.

Furthermore, for ξd{𝟎}\xi\in\mathbb{R}^{d}\hskip 0.7pt\setminus\{\mathbf{0}\},

(4.16) λ¯(ξ)=sup{α|ξ|1:αg(ξα)=μ(ξ)}[λ¯(ξ),]\overline{\lambda}(\xi)=\sup\Bigl{\{}\alpha\geq|\xi|_{1}:\alpha\hskip 0.55ptg\biggl{(}\frac{\xi}{\alpha}\biggr{)}=\mu(\xi)\Bigr{\}}\;\in\;[\hskip 0.7pt\underline{\lambda}(\xi),\infty]

is a meaningful definition as the supremum of a nonempty set. Positive homogeneity of λ¯\overline{\lambda} on d{𝟎}\mathbb{R}^{d}\hskip 0.7pt\setminus\{\mathbf{0}\} follows from the positive homogeneity of μ\mu. By Lemma 4.1(ii) and (4.14),

(4.17) r0=0impliesλ¯(ξ)=.r_{0}=0\quad\text{implies}\quad\overline{\lambda}(\xi)=\infty.

Recall μ=sup|ξ|1=1μ(ξ)\mu^{*}=\sup_{|\xi|_{1}=1}\mu(\xi). Let α\alpha be such that αg(ξ/α)=μ(ξ)\alpha\hskip 0.55ptg(\xi/\alpha)=\mu(\xi). Then

αr0αg(ξ/α)=μ(ξ)μ|ξ|1.\alpha r_{0}\leq\alpha\hskip 0.55ptg(\xi/\alpha)=\mu(\xi)\leq\mu^{*}|\xi|_{1}.

Thus

(4.18) r0>0impliesλ¯(ξ)(μ/r0)|ξ|1.r_{0}>0\quad\text{implies}\quad\overline{\lambda}(\xi)\leq(\mu^{*}/r_{0})|\xi|_{1}.

Since r0>0r_{0}>0 implies that g(𝟎)=r0>0=μ(𝟎)g(\mathbf{0})=r_{0}>0=\mu(\mathbf{0}), (4.16) is not a meaningful definition of λ¯(𝟎)\overline{\lambda}(\mathbf{0}). Cued by (4.17) and (4.18), we can retain positive homogeneity by defining

(4.19) λ¯(𝟎)={0,r0>0,r0=0.\overline{\lambda}(\mathbf{0})=\begin{cases}0,&r_{0}>0\\ \infty,&r_{0}=0.\end{cases}

The next proposition collects properties of the functions ααg(ξ/α)\alpha\mapsto\alpha g^{\diamond}(\xi/\alpha) for {𝚎𝚖𝚙𝚝𝚢,o}\diamond\in\{\langle{\tt empty}\rangle,o\}. These properties are implicit in the definitions and previously established facts. Note that part (i) below is still conditional for we have not yet proved that |ξ|1<λ¯(ξ)|\xi|_{1}<\underline{\lambda}(\xi). The trichotomy in the proposition is illustrated in Figure 2.2.

Proposition 4.4.

Assume r00r_{0}\geq 0, (2.3), and the moment bound (2.7) with p=dp=d. Fix ξd{𝟎}\xi\in\mathbb{R}^{d}\hskip 0.7pt\setminus\{\mathbf{0}\}. Then for α[|ξ|1,)\alpha\in[\,|\xi|_{1},\infty), the functions ααg(ξ/α)\alpha\mapsto\alpha\hskip 0.55ptg(\xi/\alpha) and ααgo(ξ/α)\alpha\mapsto\alpha\hskip 0.55ptg^{o}(\xi/\alpha) have the following properties.

  1. (i)

    For |ξ|1α<λ¯(ξ)|\xi|_{1}\leq\alpha<\underline{\lambda}(\xi),   αgo(ξ/α)=αg(ξ/α)\alpha\hskip 0.55ptg^{o}(\xi/\alpha)=\alpha\hskip 0.55ptg(\xi/\alpha) are strictly decreasing, convex, and strictly above μ(ξ)\mu(\xi).

  2. (ii)

    For λ¯(ξ)αλ¯(ξ)\underline{\lambda}(\xi)\leq\alpha\leq\overline{\lambda}(\xi),   αgo(ξ/α)=αg(ξ/α)=μ(ξ)\alpha\hskip 0.55ptg^{o}(\xi/\alpha)=\alpha\hskip 0.55ptg(\xi/\alpha)=\mu(\xi).

  3. (iii)

    For α>λ¯(ξ)\alpha>\overline{\lambda}(\xi), αgo(ξ/α)=μ(ξ)\alpha\hskip 0.55ptg^{o}(\xi/\alpha)=\mu(\xi), while αg(ξ/α)>μ(ξ)\alpha\hskip 0.55ptg(\xi/\alpha)>\mu(\xi) and αg(ξ/α)\alpha\hskip 0.55ptg(\xi/\alpha) is convex and strictly increasing. This case is nonempty if and only if r0>0r_{0}>0.

Proof.

The inequalities

(4.20) λ¯(ξ)<and|ξ|1λ¯(ξ)λ¯(ξ).\underline{\lambda}(\xi)<\infty\quad\text{and}\quad|\xi|_{1}\leq\underline{\lambda}(\xi)\leq\overline{\lambda}(\xi)\leq\infty.

are built into the definitions and Lemma 4.2.

Part (i). Assume |ξ|1<λ¯(ξ)|\xi|_{1}<\underline{\lambda}(\xi) so there is something to check. Since ααgo(ξ/α)\alpha\mapsto\alpha\hskip 0.55ptg^{o}(\xi/\alpha) is nonincreasing, convex, and reaches its minimum μ(ξ)\mu(\xi) at α=λ¯(ξ)\alpha=\underline{\lambda}(\xi) but not before, it must be strictly decreasing for |ξ|1α<λ¯(ξ)|\xi|_{1}\leq\alpha<\underline{\lambda}(\xi).

Suppose α0go(ξ/α0)<α0g(ξ/α0)\alpha_{0}g^{o}(\xi/\alpha_{0})<\alpha_{0}g(\xi/\alpha_{0}) for some α0>|ξ|1\alpha_{0}>|\xi|_{1}. (Equality holds at α0=|ξ|1\alpha_{0}=|\xi|_{1} by Lemma 4.1(iii).) We show that λ¯(ξ)<α0\underline{\lambda}(\xi)<\alpha_{0}. By (4.3), for some τ0[|ξ|1,α0)\tau_{0}\in[\hskip 0.9pt|\xi|_{1},\alpha_{0}) and all α[τ0,α0]\alpha\in[\tau_{0},\alpha_{0}]

α0go(ξ/α0)=τ0g(ξ/τ0)=inf|ξ|1τα0τg(ξ/τ)=inf|ξ|1τατg(ξ/τ)=αgo(ξ/α).\alpha_{0}g^{o}(\xi/\alpha_{0})=\tau_{0}g(\xi/\tau_{0})=\inf_{|\xi|_{1}\leq\tau\leq\alpha_{0}}\tau g(\xi/\tau)=\inf_{|\xi|_{1}\leq\tau\leq\alpha}\tau g(\xi/\tau)=\alpha\hskip 0.55ptg^{o}(\xi/\alpha).

Thus ααgo(ξ/α)\alpha\mapsto\alpha\hskip 0.55ptg^{o}(\xi/\alpha) is constant on [τ0,α0][\tau_{0},\alpha_{0}] with τ0<α0\tau_{0}<\alpha_{0}. It must be that λ¯(ξ)τ0<α0\underline{\lambda}(\xi)\leq\tau_{0}<\alpha_{0}.

Part (ii). From Part (i) and (4.14), the behavior of αgo(ξ/α)\alpha\hskip 0.55ptg^{o}(\xi/\alpha) is completely determined. Furthermore, αg(ξ/α)\alpha\hskip 0.55ptg(\xi/\alpha) achieves its minimum μ(ξ)\mu(\xi) at α=λ¯(ξ)\alpha=\underline{\lambda}(\xi) by a combination of (4.3) with Part (i) and (4.14). Then αg(ξ/α)\alpha\hskip 0.55ptg(\xi/\alpha) must be nondecreasing for αλ¯(ξ)\alpha\geq\underline{\lambda}(\xi), and definition (4.16) forces αg(ξ/α)=μ(ξ)\alpha\hskip 0.55ptg(\xi/\alpha)=\mu(\xi) for λ¯(ξ)αλ¯(ξ)\underline{\lambda}(\xi)\leq\alpha\leq\overline{\lambda}(\xi).

Part (iii) follows from convexity and the definitions. ∎

The next lemma shows that λ¯\underline{\lambda} is lower semicontinuous and λ¯\overline{\lambda} upper semicontinuous.

Lemma 4.5.

Let ξiξ\xi_{i}\to\xi in d{𝟎}\mathbb{R}^{d}\hskip 0.7pt\setminus\{\mathbf{0}\}. Then

(4.21) λ¯(ξ)lim¯iλ¯(ξi)lim¯iλ¯(ξi)λ¯(ξ).\underline{\lambda}(\xi)\leq\varliminf_{i\to\infty}\underline{\lambda}(\xi_{i})\leq\varlimsup_{i\to\infty}\overline{\lambda}(\xi_{i})\leq\overline{\lambda}(\xi).
Proof.

If λ¯(ξ)=|ξ|1\underline{\lambda}(\xi)=|\xi|_{1}, the first inequality of (4.21) is trivial. Suppose |ξ|1<α<λ¯(ξ)|\xi|_{1}<\alpha<\underline{\lambda}(\xi). Then αgo(ξ/α)>μ(ξ)\alpha\hskip 0.55ptg^{o}(\xi/\alpha)>\mu(\xi). By continuity on int𝒰\operatorname{int}\mathcal{U}, αgo(ξi/α)>μ(ξi)\alpha\hskip 0.55ptg^{o}(\xi_{i}/\alpha)>\mu(\xi_{i}) for large ii, which implies λ¯(ξi)>α\underline{\lambda}(\xi_{i})>\alpha.

If λ¯(ξ)=\overline{\lambda}(\xi)=\infty, the last inequality of (4.21) is trivial. By (4.17) and (4.18), the complementary case has r0>0r_{0}>0 and therefore λ¯(ξi)(μ/r0)|ξi|1\overline{\lambda}(\xi_{i})\leq(\mu^{*}/r_{0})|\xi_{i}|_{1}. Then

λ¯(ξi)g(ξiλ¯(ξi))=μ(ξi).\overline{\lambda}(\xi_{i})\hskip 0.9ptg\biggl{(}\frac{\xi_{i}}{\overline{\lambda}(\xi_{i})}\biggr{)}=\mu(\xi_{i}).

Suppose a subsequence satisfies λ¯(ξi)τ>λ¯(ξ)|ξ|1\overline{\lambda}(\xi_{i})\to\tau>\overline{\lambda}(\xi)\geq|\xi|_{1}. Then for all large enough ii, λ¯(ξi)(1+δ)|ξi|1\overline{\lambda}(\xi_{i})\geq(1+\delta)|\xi_{i}|_{1} for some δ>0\delta>0. Continuity of gg on int𝒰\operatorname{int}\mathcal{U} and of μ\mu on d\mathbb{R}^{d} then leads to τg(ξ/τ)=μ(ξ)\tau g(\xi/\tau)=\mu(\xi), a contradiction. ∎

At this point we have covered everything needed to prove part (i) of Theorem 2.11 and parts (i)(ii) of Theorem 2.16. The proofs of these theorems will be completed in Section 7.1 after the modification arguments. As the last item of this section we prove the claims about the convex duality.

Lemma 4.6.

Assume (2.7) with p=1p=1. For all ξd\xi\in\mathbb{R}^{d}, we have

limbμ(b)(ξ)b=|ξ|1.\lim_{b\to\infty}\frac{\mu^{(b)}(\xi)}{b}=|\xi|_{1}.
Proof.

We may assume that ξd{𝟎}\xi\in\mathbb{Z}^{d}\setminus\{\mathbf{0}\} for the following. The extension to ξd\xi\in\mathbb{R}^{d} follows from the homogeneity and convexity of μ(b)(ξ)\mu^{(b)}(\xi) in ξ\xi.

Claim 4.7.

For each ξd\xi\in\mathbb{Z}^{d}, there exist 2d2d edge-disjoint paths {πi}i=12d\{\pi_{i}\}_{i=1}^{2d} from 𝟎\mathbf{0} to ξ\xi such that their Euclidean lengths satisfy |π1|=|ξ|1|\pi_{1}|=|\xi|_{1} and |πi||ξ|1+8|\pi_{i}|\leq|\xi|_{1}+8 for i2i\neq 2.

The proof of this claim comes after the proof of the lemma, but it is intuitively clear that there exist 2d2d edge-disjoint paths from 𝟎\mathbf{0} to ξ\xi such that at least one path has length |ξ|1|\xi|_{1} and the length of each path is at most |ξ|1+Cξ|\xi|_{1}+C_{\xi} for some constant CξC_{\xi} (see [14, Fig. 2.1] for the case ξ=k𝐞i\xi=k\mathbf{e}_{i}). Then,

(4.22) (1+r0b)|ξ|1μ(b)(ξ)b𝔼[T𝟎,ξ(b)]b𝔼[b1mini=1,,2dT(b)(πi)],\left(1+\frac{r_{0}}{b}\right)|\xi|_{1}\leq\frac{\mu^{(b)}(\xi)}{b}\leq\frac{\mathbb{E}[T^{(b)}_{\mathbf{0},\xi}]}{b}\leq\mathbb{E}\Bigl{[}b^{-1}\min_{i=1,\ldots,2d}T^{(b)}(\pi_{i})\Bigr{]},

where the first inequality follows from T𝟎,ξ(b)|ξ|1(b+r0)T^{(b)}_{\mathbf{0},\xi}\geq|\xi|_{1}(b+r_{0}), the second from subadditivity, and the third from the fact that T𝟎,ξ(b)T^{(b)}_{\mathbf{0},\xi} is an infimum over all paths from 𝟎\mathbf{0} to ξ\xi. Denote the integrand on the right-hand side of (4.22) by ZbZ_{b}.

Since T(b)(πi)(Cξ+|ξ|1)b+T(πi)T^{(b)}(\pi_{i})\leq(C_{\xi}+|\xi|_{1})b+T(\pi_{i}) for i=1,,2di=1,\ldots,2d, we have

Zb(Cξ+|ξ|1)+mini=1,,2dT(πi)for all b1.Z_{b}\leq(C_{\xi}+|\xi|_{1})+\min_{i=1,\ldots,2d}T(\pi_{i})\quad\text{for all }b\geq 1.

Next, we show that mini=1,,2dT(πi)\min_{i=1,\ldots,2d}T(\pi_{i}) is integrable (see [2, Theorem 2.2]) in preparation for the dominated convergence theorem. A union bound over the edges of each path πi\pi_{i} and independence of the edge weights in the paths implies

{mini=1,,2dT(πi)s}(maxi|πi|{tes|πi|})2d.\mathbb{P}\bigl{\{}\min_{i=1,\ldots,2d}T(\pi_{i})\geq s\bigr{\}}\leq\Bigl{(}\max_{i}|\pi_{i}|\,\mathbb{P}\Bigl{\{}t_{e}\geq\frac{s}{|\pi_{i}|}\Bigr{\}}\Bigr{)}^{2d}.

Integrating over s0s\geq 0 shows that for some constant CξC_{\xi},

𝔼[mini=1,,2dT(πi)]Cξ𝔼[min{t1,,t2d}]<.\mathbb{E}\bigl{[}\min_{i=1,\ldots,2d}T(\pi_{i})\bigr{]}\leq C_{\xi}\,\mathbb{E}[\,\min\{t_{1},\dotsc,t_{2d}\}\,]<\infty.

Since ZbZ_{b} can be written as

Zb=mini=1,,2d(|πi|+T(πi)b),Z_{b}=\min_{i=1,\ldots,2d}\Bigl{(}|\pi_{i}|+\frac{T(\pi_{i})}{b}\Big{)},

we see that limbZb=|π1|=|ξ|1\lim_{b\to\infty}Z_{b}=|\pi_{1}|=|\xi|_{1}. Therefore, by the dominated convergence theorem, we have

|ξ|1limbμ(b)(ξ)blimb𝔼[Zb]=|ξ|1.|\xi|_{1}\leq\lim_{b\to\infty}\frac{\mu^{(b)}(\xi)}{b}\leq\lim_{b\to\infty}\mathbb{E}[Z_{b}]=|\xi|_{1}.\qed
Proof of Claim 4.7.

For general ξd\xi\in\mathbb{Z}^{d}, let kk be the number of non-zero coordinates of ξ\xi and suppose k>1k>1. This is the effective dimension of the rectangle formed with the origin and ξ\xi as extreme opposing corners. We may assume without loss of generality that the first kk coordinates of ξ\xi are non-zero and the rest are 0. So let ξ=(a1,a2,,ak,0,,0)\xi=(a_{1},a_{2},\dotsc,a_{k},0,\dotsc,0).

The first kk disjoint paths run along the edges of the rectangle. Such a path is encoded by a permutation σ𝕊k\sigma\in\mathbb{S}_{k}. For example, σ=(1,2,,k)\sigma=(1,2,\ldots,k) corresponds to the path 𝟎a1𝐞1a1𝐞1+a2𝐞2\mathbf{0}\to a_{1}\mathbf{e}_{1}\to a_{1}\mathbf{e}_{1}+a_{2}\mathbf{e}_{2}\to\cdots. Two paths encoded by permutations σ=(σ1,,σk)\sigma=(\sigma_{1},\ldots,\sigma_{k}) and μ=(μ1,,μk)\mu=(\mu_{1},\ldots,\mu_{k}) meet (share a vertex) before ξ\xi if and only if for some j<kj<k, {σ1,,σj}={μ1,,μj}\{\sigma_{1},\ldots,\sigma_{j}\}=\{\mu_{1},\ldots,\mu_{j}\}. Consider the kk paths corresponding to the cyclic permutations:

π1=(1,2,,k),π2=(2,3,,1),,πk=(k,1,,k1).\displaystyle\pi_{1}=(1,2,\ldots,k),\,\pi_{2}=(2,3,\ldots,1),\dotsc,\pi_{k}=(k,1,\ldots,k-1).

These kk paths are vertex disjoint, except for their first and last vertices, and have length |ξ|1|\xi|_{1}.

The next dkd-k paths are formed as follows. For each j{k+1,,d}j\in\{k+1,\dotsc,d\}, start with an 𝐞j\mathbf{e}_{j} step, follow the path π1\pi_{1} to ξ+𝐞j\xi+\mathbf{e}_{j}, and conclude with a 𝐞j-\mathbf{e}_{j} step to ξ\xi. Get another dkd-k paths by starting with 𝐞j-\mathbf{e}_{j} and finishing with 𝐞j\mathbf{e}_{j}. These paths have length |ξ|1+2|\xi|_{1}+2.

Now we have altogether k+2(dk)k+2(d-k) paths. The final kk paths are a little trickier.

For each i={1,,k}i=\{1,\ldots,k\}, pair direction 𝐞i\mathbf{e}_{i} with path pi+1modkp_{i+1\!\!\!\mod\!k}. We construct the path for i=1i=1, and the rest are similar. The first step is 𝐞1-\mathbf{e}_{1}. Then follow π2\pi_{2} until π2\pi_{2} is about to step in the 𝐞k\mathbf{e}_{k} direction (the last step before it steps in the 𝐞1\mathbf{e}_{1} direction). On the 𝐞k\mathbf{e}_{k} segment take ak+1a_{k}+1 steps and then take a1+1a_{1}+1 steps in the 𝐞1\mathbf{e}_{1} direction (this avoids the π2\pi_{2} path), ending up at ξ+𝐞k\xi+\mathbf{e}_{k}. Finish at ξ\xi by taking a final 𝐞k-\mathbf{e}_{k} step. Replacing 𝐞1\mathbf{e}_{1} and π2\pi_{2} by 𝐞j\mathbf{e}_{j} and pj+1modkp_{j+1\!\!\!\mod\!k} for j=2,,kj=2,\ldots,k gives us kk such paths that are disjoint from each other and all the previous paths (except for their first and last vertices). All these have length |ξ|1+4|\xi|_{1}+4. Notice the crucial assumption of k>1k>1 for this construction.

The k=1k=1 case is covered in [14, Fig 2.1], as mentioned earlier. One can verify that this gives the worst case of |ξ|1+8|\xi|_{1}+8. ∎

Proof of Theorem 2.17.

Step 1. Identity (2.44). For br0b\geq-r_{0}, (2.44) is a combination of (4.15) and (2.42). For large α\alpha

(4.23) αg(ξ/α)μ(ξ)+αr0\alpha\hskip 0.55ptg(\xi/\alpha)\leq\mu(\xi)+\alpha r_{0}

because an nα\lfloor{n\alpha}\rfloor-path from 𝟎\mathbf{0} to a point close to nξn\xi can be created by following the strategy in the proof of Lemma 4.1(ii): repeat an edge close to the origin with weight close to r0r_{0} as many times as needed, and then follow a geodesic to a point close to nξn\xi. Bound (4.23) implies that the right-hand side of (2.44) equals -\infty for b<r0b<-r_{0}. Identity (2.44) has been verified for all bb\in\mathbb{R}.

Step 2. The duality. The convexity and lower semicontinuity of ααg(ξ/α)\alpha\mapsto\alpha\hskip 0.55ptg(\xi/\alpha) for α|ξ|1\alpha\geq|\xi|_{1} imply that the function defined by the right-hand side of (2.43) is concave and upper semicontinuous. Thus (2.44) implies that μ¯ξ\overline{\mu}_{\xi} is the concave dual of this function. Then we can identify the dual μ¯ξ\overline{\mu}_{\xi}^{*} of μ¯ξ\overline{\mu}_{\xi} as (2.43), which gives (2.45).

Step 3. The superdifferentials. Let b>r0b>-r_{0}. Then λ¯(b)(ξ)<\overline{\lambda}^{\hskip 0.9pt(b)}(\xi)<\infty by (4.18). By Proposition 4.4 and the duality,

(4.24) [λ¯(b)(ξ),λ¯(b)(ξ)]\displaystyle{[\,\underline{\lambda}^{(b)}(\xi),\overline{\lambda}^{\hskip 0.9pt(b)}(\xi)\,]} ={α|ξ|1:μ¯ξ(b)=αg(b)(ξ/α)}\displaystyle=\{\alpha\geq|\xi|_{1}:\overline{\mu}_{\xi}(b)=\alpha g^{(b)}(\xi/\alpha)\}
={α|ξ|1:μ¯ξ(b)=αg(ξ/α)+αb}\displaystyle=\{\alpha\geq|\xi|_{1}:\overline{\mu}_{\xi}(b)=\alpha\hskip 0.55ptg(\xi/\alpha)+\alpha b\}
={α:μ¯ξ(b)=αbμ¯ξ(α)}=μ¯ξ(b).\displaystyle=\{\alpha\in\mathbb{R}:\overline{\mu}_{\xi}(b)=\alpha b-\overline{\mu}_{\xi}^{*}(\alpha)\}=\partial\overline{\mu}_{\xi}(b).

Similarly

(4.25) [λ¯(r0)(ξ),)\displaystyle{[\,\underline{\lambda}^{(-r_{0})}(\xi),\infty)} ={α|ξ|1:μ¯ξ(r0)=αg(r0)(ξ/α)}\displaystyle=\{\alpha\geq|\xi|_{1}:\overline{\mu}_{\xi}(-r_{0})=\alpha g^{(-r_{0})}(\xi/\alpha)\}
={α:μ¯ξ(r0)=αr0μ¯ξ(α)}=μ¯ξ(r0).\displaystyle=\{\alpha\in\mathbb{R}:\overline{\mu}_{\xi}(-r_{0})=-\alpha r_{0}-\overline{\mu}_{\xi}^{*}(\alpha)\}=\partial\overline{\mu}_{\xi}(-r_{0}).

Fix a>r0a>-r_{0} and let b>ab>a. From |ξ|1μξ(b±)|\xi|_{1}\leq\mu_{\xi}^{\prime}(b\pm) given in (2.14), concavity, and Lemma 4.6,

|ξ|1μ¯ξ(b±)μ¯ξ(b)μ¯ξ(a)ba|ξ|1as b.|\xi|_{1}\leq\overline{\mu}_{\xi}^{\prime}(b\pm)\leq\frac{\overline{\mu}_{\xi}(b)-\overline{\mu}_{\xi}(a)}{b-a}\;\to\;|\xi|_{1}\quad\text{as }b\to\infty.

5. Modification proofs for strict concavity

The modification arguments provide the power to go beyond soft results. In particular, these give us the strict concavity of the shape function in the shift variable (Theorem 2.2(ii)), the strict separation of λ¯(ξ)\underline{\lambda}(\xi) from |ξ|1|\xi|_{1} (Theorem 2.11(ii)), and the strict exceedance of 1\ell^{1} distance by the geodesic length (Theorem 2.5).

5.1. Preparation for the modification arguments

We adapt to our goals the modification argument of van den Berg and Kesten [20]. Throughout this section r0=essinft(e)0r_{0}=\mathop{\mathrm{ess\,inf}}t(e)\geq 0.

An NN\!-box BB is by definition a rectangular subset of d\mathbb{Z}^{d} of the form

(5.1) B={x=(x1,,xd)d:aixiai+3N for i[d]k,akxkak+N}B=\{x=(x_{1},\dotsc,x_{d})\in\mathbb{Z}^{d}:a_{i}\leq x_{i}\leq a_{i}+3N\text{ for }i\in[d]\setminus k,\,a_{k}\leq x_{k}\leq a_{k}+N\}

for some a=(a1,,ad)da=(a_{1},\dotsc,a_{d})\in\mathbb{Z}^{d} and k[d]k\in[d]. In other words, one of the dimensions of BB has size NN and the other d1d-1 dimensions are of size 3N3N. The two large 3N××3N3N\times\dotsm\times 3N faces of BB in (5.1) are the subsets

{xB:xk=ak}and{xB:xk=ak+N}.\{x\in B:x_{k}=a_{k}\}\quad\text{and}\quad\{x\in B:x_{k}=a_{k}+N\}.

The interior of BB is defined by requiring ai<xi<ai+3Na_{i}<x_{i}<a_{i}+3N and ak<xk<ak+Na_{k}<x_{k}<a_{k}+N in (5.1). The boundary B\partial B of BB is the set of points of BB that have a nearest-neighbor vertex in the complement of BB. Our convention will be that an edge ee lies in BB if both its endpoints lie in BB, otherwise eBce\in B^{c}. A suitable 1\ell^{1}-neighborhood around BB is defined by

(5.2) B¯={xd:yB:|xy|13N(d1)+N}.\overline{B}=\{x\in\mathbb{Z}^{d}:\exists y\in B:\,|x-y|_{1}\leq 3N(d-1)+N\}.

The significance of the choice 3N(d1)+N3N(d-1)+N is that the 1\ell^{1}-distance from any point in BB to the boundary of B¯\overline{B} is at least as large as the distance between any two points in BB.

Introduce two parameters 0<s0,δ0<0<s_{0},\delta_{0}<\infty whose choices are made precise later. Consider these conditions on the edge weights in BB and B¯\overline{B}:

(5.3) maxeBt(e)s0,\max_{e\in B}t(e)\leq s_{0}\,,
(5.4) eBt(e)s0,\sum_{e\in B}t(e)\leq s_{0}\,,

and

(5.5) T(π)>(r0+δ0)|yx|1T(\pi)>(r_{0}+\delta_{0})|y-x|_{1} for every self-avoiding path π\pi that stays entirely in B¯\overline{B}
and whose endpoints xx and yy satisfy |yx|1N|y-x|_{1}\geq N.

The properties of a black box stated in the next definition depend on whether the weights are bounded or unbounded. We let M0=-esssupt(e)M_{0}=\mathbb{P}\text{-}\mathop{\mathrm{ess\,sup}}t(e).

Definition 5.1 (Black box).

  1. (i)

    In the case of bounded weights (M0<M_{0}<\infty), color a box BB black if conditions (5.3) and (5.5) are satisfied.

  2. (ii)

    In the case of unbounded weights (M0=M_{0}=\infty), color a box BB black if conditions (5.4) and (5.5) are satisfied.

By choosing s0s_{0} and NN large enough and δ0\delta_{0} small enough, the probability of a given BB being black can be made as close to 11 as desired. This is evident for conditions (5.3) and (5.4). For condition (5.5) it follows from Lemma 5.5 in [20] that we quote here:

Lemma 5.2.

[20, Lemma 5.5] Assume (2.6), that is, that the infimum of the passage time is subcritical. Then there exist constants δ0>0\delta_{0}>0 and D0>0D_{0}>0 such that for all x,ydx,y\in\mathbb{Z}^{d},

(5.6) {Tx,y(r0+δ0)|yx|1}eD0|yx|1.\mathbb{P}\{T_{x,y}\leq(r_{0}+\delta_{0})|y-x|_{1}\}\leq e^{-D_{0}|y-x|_{1}}.

When r0>0r_{0}>0, Lemma 5.5 of [20] requires the weaker assumption P{t(e)=r0}<pcP\{t(e)=r_{0}\}<\vec{p}_{c} where pc\vec{p}_{c} is the critical probability of oriented bond percolation on d\mathbb{Z}^{d}. However, since we consider shifts of weights that can turn r0r_{0} into zero, it is simpler to assume (2.6) for all r00r_{0}\geq 0 instead of keeping track when we might get by with the weaker assumption.

The probability of the complement of (5.5) is then bounded by

{(5.5) fails}x,yB¯:|yx|1N{Tx,y(r0+δ0)|yx|1}CdN2deD0N.\displaystyle\mathbb{P}\{\text{\eqref{1.bl2} fails}\}\leq\sum_{x,\,y\,\in\overline{B}:\,|y-x|_{1}\geq N}\mathbb{P}\{T_{x,y}\leq(r_{0}+\delta_{0})|y-x|_{1}\}\leq C_{d}N^{2d}e^{-D_{0}N}.

The bound above decreases for large enough NN and hence gives us this conclusion:

(5.7) There exists a fixed δ0>0\delta_{0}>0 such that for any ε>0\varepsilon>0 there exist NN and s0s_{0}
such that {box B is black}1ε\mathbb{P}\{\text{box $B$ is black}\}\geq 1-\varepsilon while {t(e)s0}>0\mathbb{P}\{t(e)\geq s_{0}\}>0. Increasing NN and s0s_{0}
while keeping δ0\delta_{0} fixed cannot violate this condition as long as {t(e)s0}>0\mathbb{P}\{t(e)\geq s_{0}\}>0.

Condition {t(e)s0}>0\mathbb{P}\{t(e)\geq s_{0}\}>0 is included above simply to point out that s0s_{0} is not chosen so large that property (5.3) becomes trivial for bounded weights.

A nearest-neighbor path π=(xi)i=0n\pi=(x_{i})_{i=0}^{n} that lies in BB is a short crossing of BB if x0x_{0} and xnx_{n} lie on opposite large faces of BB. More generally, we say that

(5.8) a path π\pi crosses BB if some segment πxk,xm=(xi)i=km\pi_{x_{k},x_{m}}=(x_{i})_{i=k}^{m} of π\pi is a short crossing of BB
and neither endpoint of π\pi lies in BB.

The second part of the definition ensures that π\pi genuinely “goes through” BB.

Let \mathcal{B} be the countable set of all triples (B,v,w)(B,v,w) where BB is an NN-box and vv and ww are two distinct points on the boundary of BB. A path π\pi has a (B,v,w)(B,v,w)-crossing if (5.8) holds and vv is the point where π\pi first enters BB and ww is the point through which π\pi last exits BB. (Then the short crossing of BB is some segment πv,wπv,w\pi_{v^{\prime},w^{\prime}}\subset\pi_{v,w}.) If π\pi crosses BB, then π\pi has a (B,v,w)(B,v,w)-crossing for some (B,v,w)(B,v,w)\in\mathcal{B} with (v,w)(v,w) uniquely determined by π\pi and BB.

Partition the set \mathcal{B} of all elements (B,v,w)(B,v,w) into KK subcollections 1,,K\mathcal{B}_{1},\dotsc,\mathcal{B}_{K} such that within each j\mathcal{B}_{j} all boxes BB are separated by distance NN. Any particular box BB appears at most once in any particular j\mathcal{B}_{j}. The number KK of subcollections depends only on the dimension dd and the size parameter NN. The particular size NN of the separation of boxes in j\mathcal{B}_{j} is taken for convenience only. In the end what matters is that the boxes are separated and that once NN is fixed, KK is a constant.

Let 𝔹(0,r)={xd:|x|1r}{\mathbb{B}}(0,r)=\{x\in\mathbb{Z}^{d}:|x|_{1}\leq r\} denote the 1\ell^{1}-ball (diamond) of radius r\lfloor{r}\rfloor in d\mathbb{Z}^{d}, with (inner) boundary 𝔹(0,r)={xd:|x|1=r}\partial{\mathbb{B}}(0,r)=\{x\in\mathbb{Z}^{d}:|x|_{1}=\lfloor{r}\rfloor\}. The lemma below is proved in Appendix C.

Lemma 5.3.

By fixing s0s_{0} and NN large enough and δ0\delta_{0} small enough as in (5.7), we can ensure the existence of constants 0<δ1,D1,n1<0<\delta_{1},D_{1},n_{1}<\infty such that, for all nn1n\geq n_{1},

(5.9) {\displaystyle\mathbb{P}\bigl{\{} every lattice path π\pi from the origin to 𝔹(0,n)\partial{\mathbb{B}}(0,n) has an index
     j(π)[K]j(\pi)\in[K] such that π\pi has at least nδ1\lfloor{n\delta_{1}}\rfloor (B,v,w)(B,v,w)-crossings
 of black boxes B such that (B,v,w)j(π)}1eD1n.\displaystyle\quad\;\text{ of black boxes $B$ such that $(B,v,w)\in\mathcal{B}_{j(\pi)}$}\bigr{\}}\geq 1-e^{-D_{1}n}.

We turn to the modification argument for the strict concavity of μξ\mu_{\xi} claimed in Theorem 2.2.

5.2. Strict concavity

Let δ0>0\delta_{0}>0 be the quantity in (5.5) in the definition of a black box. In addition to t(e)0t(e)\geq 0 we consider two complementary assumptions on the weight distribution. Either the weights are unbounded:

(5.10) M0=M_{0}=\infty

and satisfy a moment bound, or the weights are bounded and have a strictly positive support point close enough to the lower bound:

(5.11) the support of t(e)t(e) contains a point r1r_{1} that satisfies
0<r1<r0+δ0<M0<.\displaystyle\qquad\qquad\qquad\qquad 0<r_{1}<r_{0}+\delta_{0}<M_{0}<\infty.

If r0>0r_{0}>0 we can choose r1=r0r_{1}=r_{0}. Let ε0>0\varepsilon_{0}>0 be the constant that appears in Theorem 2.1 and in Theorem A.1, also equal to the constant δ\delta in (4.10) for the shifted weights ω(r0)\omega^{(-r_{0})}.

Theorem 5.4.

Assume r00r_{0}\geq 0 and (2.6), in other words, that weights are nonnegative and the infimum is subcritical. Furthermore, assume that one of these two cases holds:

  1. (a)

    Unbounded case: the weights satisfy (5.10) and the moment bound (2.7) with p=1p=1.

  2. (b)

    Bounded case: the weights satisfy (5.11).

Then there exists a finite positive constant MM and a function D(b)>0D(b)>0 of b>0b>0 such that the following bounds hold for all b(0,r0+ε0)b\in(0,r_{0}+\varepsilon_{0}) and all |x|1M|x|_{1}\geq M:

  1. (i)

    In the unbounded case (a),

    (5.12) 𝔼[T𝟎,x(b)]𝔼[T𝟎,x]b𝔼[L¯𝟎,x]D(b)b|x|1.\mathbb{E}[T^{(-b)}_{\mathbf{0},x}]\leq\mathbb{E}[T_{\mathbf{0},x}]-b\,\mathbb{E}[\,\overline{L}_{\hskip 0.9pt\mathbf{0},x}]-D(b)b|x|_{1}.
  2. (ii)

    In the bounded case (b),

    (5.13) 𝔼[T𝟎,x(b)]𝔼[T𝟎,x]b𝔼[L¯𝟎,x]D(b)b|x|1.\mathbb{E}[T^{(-b)}_{\mathbf{0},x}]\leq\mathbb{E}[T_{\mathbf{0},x}]-b\,\mathbb{E}[\,\underline{L}_{\hskip 0.9pt\mathbf{0},x}]-D(b)b|x|_{1}.

Condition (2.7) with p=1p=1 guarantees that the expectation 𝔼[T𝟎,x]\mathbb{E}[T_{\mathbf{0},x}] above is finite (Lemma 2.3 in [2]). This together with Lemma A.3 then implies that 𝔼[T𝟎,x(b)]\mathbb{E}[T^{(-b)}_{\mathbf{0},x}] is finite for b(0,r0+ε0)b\in(0,r_{0}+\varepsilon_{0}). Since 𝔼[L¯𝟎,x]𝔼[L¯𝟎,x]\mathbb{E}[\,\overline{L}_{\hskip 0.9pt\mathbf{0},x}]\geq\mathbb{E}[\,\underline{L}_{\hskip 0.9pt\mathbf{0},x}], (5.12) provides a better bound than (5.13). This is due to the fact that the modification argument gives sharper control of the geodesic under unbounded weights.

Our modification proofs force the geodesic to follow explicitly constructed paths. These paths are parametrized by two integers kk and \ell whose choice is governed by the support of t(e)t(e) through the lemma below.

Lemma 5.5.

Fix reals 0<r<s0<r<s and b>0b>0. Then there exist arbitrarily large positive integers k,k,\ell such that

(5.14) k(s+δ)<(k+2)(rδ)<(k+2)(r+δ)<k(sδ)+(21)bk(s+\delta)<(k+2\ell)(r-\delta)<(k+2\ell)(r+\delta)<k(s-\delta)+(2\ell-1)b

holds for sufficiently small real δ>0\delta>0.

Proof.

It suffices to show the existence of arbitrarily large positive integers k,k,\ell that satisfy the strict inequalities

(5.15) ks<(k+2)r<ks+(21)bks<(k+2\ell)r<ks+(2\ell-1)b

and then choose δ>0\delta>0 small enough. Let 0<ε<b/r0<\varepsilon<b/r and choose an integer m>2/εm>2/\varepsilon. Then for each kk\in\mathbb{N} there exists \ell\in\mathbb{N} such that

(5.16) k(sr1)<2<k(sr1)+mε,k\Bigl{(}\,\frac{s}{r}-1\Bigr{)}<2\ell<k\Bigl{(}\,\frac{s}{r}-1\Bigr{)}+m\varepsilon,

and kk and \ell can be taken arbitrarily large. Rearranging (5.16) and remembering the choice of ε\varepsilon gives

ks<(k+2)r<ks+mεr<ks+mb.ks<(k+2\ell)r<ks+m\varepsilon r<ks+mb.

To get (5.15), take kk and \ell large enough to have m<21m<2\ell-1. ∎

Proof of Theorem 5.4.

The proof has three stages. The first and the last are common to bounded and unbounded weights. The most technical middle stage has to be tailored separately to the two cases. We present the stages in their logical order, with separate cases for the middle stage.


Stage 1 for both bounded and unbounded weights.

Let π(x)\pi(x) be a geodesic for T𝟎,xT_{\mathbf{0},x}. When geodesics are not unique, π(x)\pi(x) will be chosen in particular measurable ways that are made precise later in the proofs. Assume that |x|1n1|x|_{1}\geq n_{1} so that Lemma 5.3 applies with n=|x|1n=|x|_{1}. The event in (5.9) lies in the union

j=1K{π(x) crosses at least |x|1δ1 black boxes from j}.\bigcup_{j=1}^{K}\{\text{$\pi(x)$ crosses at least $\lfloor{|x|_{1}\delta_{1}}\rfloor$ black boxes from $\mathcal{B}_{j}$}\}.

By (5.9), there is a nonrandom index j(x)[K]j(x)\in[K] such that

(5.17) {π(x) crosses at least |x|1δ1 black boxes from j(x)}1eD1|x|1K.\mathbb{P}\bigl{\{}\text{$\pi(x)$ crosses at least $\lfloor{|x|_{1}\delta_{1}}\rfloor$ black boxes from $\mathcal{B}_{j(x)}$}\bigr{\}}\geq\frac{1-e^{-D_{1}|x|_{1}}}{K}.

Define the event

(5.18) ΛB,v,w,x={B is black and π(x) has a (B,v,w)-crossing}.\Lambda_{B,v,w,x}=\{\text{$B$ is black and $\pi(x)$ has a $(B,v,w)$-crossing}\}.

Consequently

(5.19) {ΛB,v,w,x occurs for at least |x|1δ1 elements (B,v,w)j(x)}1eD1|x|1K.\displaystyle\mathbb{P}\{\text{$\Lambda_{B,v,w,x}$ occurs for at least $\lfloor{\hskip 0.55pt|x|_{1}\delta_{1}}\rfloor$ elements $(B,v,w)\in\mathcal{B}_{j(x)}$}\}\geq\frac{1-e^{-D_{1}|x|_{1}}}{K}.

Turn this into a lower bound on the expected number of events, with a new constant D1>0D_{1}>0:

(5.20) (B,v,w)j(x)(ΛB,v,w,x)=𝔼[#{(B,v,w)j(x):ΛB,v,w,x occurs}]D1|x|1.\displaystyle\sum_{(B,v,w)\in\mathcal{B}_{j(x)}}\mathbb{P}(\Lambda_{B,v,w,x})=\mathbb{E}\bigl{[}\,\#\{(B,v,w)\in\mathcal{B}_{j(x)}:\text{$\Lambda_{B,v,w,x}$ occurs}\}\bigr{]}\geq D_{1}|x|_{1}.

The next Stage 2 of the proof shows that, after a modification of the environment on a black box, the geodesic encounters a k+2k+2\ell detour whose weights are determined by the modification. By this we mean that the geodesic runs through a straight-line kk-step path segment of the form π+=(π0++i𝐮)0ik\pi^{+}=(\pi^{+}_{0}+i\mathbf{u})_{0\leq i\leq k} parallel to an integer unit vector 𝐮{±𝐞i}i=1d\mathbf{u}\in\{\pm\mathbf{e}_{i}\}_{i=1}^{d}, with some initial vertex π0+\pi^{+}_{0}. A k+2k+2\ell detour associated to π+\pi^{+} is a path π++=(πi++)0ik+2\pi^{++}=(\pi^{++}_{i})_{0\leq i\leq k+2\ell} that shares both endpoints with π+\pi^{+} and translates the kk-segment by \ell steps in a direction perpendicular to 𝐮\mathbf{u}: so for some integer unit vector 𝐮𝐮\mathbf{u}^{\prime}\perp\mathbf{u},

(5.21) πi++={π0++i𝐮,0iπ0++𝐮+(i)𝐮,+1ik+π0++𝐮+k𝐮(ik)𝐮,k++1ik+2.\pi^{++}_{i}=\begin{cases}\pi^{+}_{0}+i\mathbf{u}^{\prime},&0\leq i\leq\ell\\ \pi^{+}_{0}+\ell\mathbf{u}^{\prime}+(i-\ell)\mathbf{u},&\ell+1\leq i\leq k+\ell\\ \pi^{+}_{0}+\ell\mathbf{u}^{\prime}+k\mathbf{u}-(i-k-\ell)\mathbf{u}^{\prime},&k+\ell+1\leq i\leq k+2\ell.\end{cases}

In particular, π+\pi^{+} and π++\pi^{++} are edge-disjoint while they share their endpoints.

The k×k\times\ell rectangle G=[π0+,π0++k𝐮]×[π0+,π0++𝐮]G=[\pi^{+}_{0},\pi^{+}_{0}+k\mathbf{u}\hskip 0.55pt]\times[\pi^{+}_{0},\pi^{+}_{0}+\ell\mathbf{u}^{\prime}\hskip 0.7pt] enclosed by π+\pi^{+} and π++\pi^{++} will be called a detour rectangle. Its relative boundary on the plane spanned by {𝐮,𝐮}\{\mathbf{u},\mathbf{u}^{\prime}\} is G=π+π++\partial G=\pi^{+}\cup\pi^{++}. Throughout we use superscripts ++ and ++++ to indicate objects associated with the two portions of the boundaries of detour rectangles GG. Figure 5.1 illustrates.

π+\pi^{+}π0+\pi_{0}^{+}π++\pi_{\ell}^{++}πk+++\pi_{k+\ell}^{++}πk+=πk+2++\pi_{k}^{+}=\pi_{k+2\ell}^{++}kk\ellGG𝐮\mathbf{u}^{\prime}𝐮\mathbf{u}
Figure 5.1. Illustration of (5.21): 𝐮\mathbf{u} and 𝐮\mathbf{u}^{\prime} are two perpendicular unit vectors in d\mathbb{Z}^{d}, π+\pi^{+} is a path that takes kk 𝐮\mathbf{u}-steps, while the detour π++\pi^{++} first takes \ell 𝐮\mathbf{u}^{\prime}-steps, followed by kk 𝐮\mathbf{u}-steps, and last \ell (𝐮)(-\mathbf{u}^{\prime})-steps. The detour rectangle GG is bounded by these paths.

Stage 2 is undertaken separately for bounded and unbounded weights.


Stage 2 for bounded weights.

Lemma 5.6.

Assume (5.11). For i{0,1,2}i\in\{0,1,2\} there exist nondecreasing sequences {si(q)}q\{s_{i}(q)\}_{q\in\mathbb{N}} with the following properties:

(5.22) r0+δ0<s0(q)s1(q)s2(q)=M0,r_{0}+\delta_{0}<s_{0}(q)\leq s_{1}(q)\leq s_{2}(q)=M_{0},
(5.23) limqs0(q)=M0andlimq{t(e)s0(q)}=1,\lim_{q\to\infty}s_{0}(q)=M_{0}\quad\text{and}\quad\lim_{q\to\infty}\mathbb{P}\{t(e)\leq s_{0}(q)\}=1,
(5.24)  for ε>0 and q,{s0(q)εt(e)s0(q)}>0,\text{ for }\varepsilon>0\text{ and }q\in\mathbb{N},\ \ \mathbb{P}\{s_{0}(q)-\varepsilon\leq t(e)\leq s_{0}(q)\}>0,
(5.25) and for i{0,1} and q,{si(q)t(e)si+1(q)}>0.\text{and for }i\in\{0,1\}\text{ and }q\in\mathbb{N},\ \ \mathbb{P}\bigl{\{}s_{i}(q)\leq t(e)\leq s_{i+1}(q)\bigr{\}}>0.
Proof.

If {t(e)=M0}>0\mathbb{P}\{t(e)=M_{0}\}>0 then let si(q)=M0s_{i}(q)=M_{0} for all ii and qq. So suppose {t(e)=M0}=0\mathbb{P}\{t(e)=M_{0}\}=0.

Let s0(0)=r0+δ0s_{0}(0)=r_{0}+\delta_{0}. For q1q\geq 1 define inductively s0(q)s_{0}(q) in the interval (s0(q1)(M0q1),M0)(s_{0}(q-1)\vee(M_{0}-q^{-1}),M_{0}) so that {s0(q)εt(e)s0(q)}>0\mathbb{P}\{s_{0}(q)-\varepsilon\leq t(e)\leq s_{0}(q)\}>0 for all ε>0\varepsilon>0. This can be done as follows. Let s0(q)s_{0}(q) be an atom of t(e)t(e) in (s0(q1)(M0q1),M0)(s_{0}(q-1)\vee(M_{0}-q^{-1}),M_{0}) if one exists. If not, the c.d.f. of t(e)t(e) is continuous in this interval and we take s0(q)s_{0}(q) to be a point of strict increase which must exist.

Then s0(q)M0s_{0}(q)\to M_{0} and thus {t(e)s0(q)}1\mathbb{P}\{t(e)\leq s_{0}(q)\}\to 1. Furthermore, {t(e)>s0(q)}>0\mathbb{P}\{t(e)>s_{0}(q)\}>0 for all qq because s0(q)<M0s_{0}(q)<M_{0}. Pick s(q)[s0(q),M0)s^{\prime}(q)\in[s_{0}(q),M_{0}) so that {s0(q)t(e)s(q)}>0\mathbb{P}\{s_{0}(q)\leq t(e)\leq s^{\prime}(q)\}>0. Define a nondecreasing sequence by s1(q)=maxjqs(j)s_{1}(q)=\max_{j\leq q}s^{\prime}(j). Since s1(q)<M0s_{1}(q)<M_{0} we have {t(e)>s1(q)}>0\mathbb{P}\{t(e)>s_{1}(q)\}>0. ∎

We fix various parameters for this stage of the proof. Fix b(0,r1)b\in(0,r_{1}) and determine k,,δk,\ell,\delta^{\prime} by applying Lemma 5.5 to 0<b<r1<M00<b<r_{1}<M_{0} to have

(5.26) k(M0+δ)<(k+2)(r1δ)<(k+2)(r1+δ)<k(M0δ)+(21)b.k(M_{0}+\delta^{\prime})<(k+2\ell)(r_{1}-\delta^{\prime})<(k+2\ell)(r_{1}+\delta^{\prime})<k(M_{0}-\delta^{\prime})+(2\ell-1)b.

Since s0(q)M0s_{0}(q)\to M_{0} from below, we can fix qq large enough and δ(0,δ)\delta\in(0,\delta^{\prime}) small enough so that

(5.27) k(s0+δ)<(k+2)(r1δ)<(k+2)(r1+δ)<k(s0δ)+(21)bk(s_{0}+\delta)<(k+2\ell)(r_{1}-\delta)<(k+2\ell)(r_{1}+\delta)<k(s_{0}-\delta)+(2\ell-1)b

holds for s0=s0(q)s_{0}=s_{0}(q). Note that this continues to hold if we increase qq to take s0s_{0} closer to M0M_{0} or decrease δ\delta.

Take NN large enough, δ0>0\delta_{0}>0 small enough, and qq large enough so that the crossing bound (5.9) of Lemma 5.3 is satisfied for the choice s0=s0(q)s_{0}=s_{0}(q). Drop qq from the notation and henceforth write si=si(q)s_{i}=s_{i}(q).

Shrink δ>0\delta>0 further so that

(5.28) r1+δ<r0+δ0\displaystyle r_{1}+\delta<r_{0}+\delta_{0}

and

(5.29) (+1)s0>(+1)(r1+δ)+kδ.\displaystyle(\ell+1)s_{0}>(\ell+1)(r_{1}+\delta)+k\delta.

The construction to come will attach k+2k+2\ell detours to edges of cubes. The number of such attachments per edge is given by the parameter

k0=30dM0r0+δ0r1δ+2.\displaystyle k_{0}=\Bigl{\lceil}\frac{30dM_{0}}{r_{0}+\delta_{0}-r_{1}-\delta}\Bigr{\rceil}+2.

Let m1m_{1} be an even positive integer and define two constants

(5.30) c1=2ks0+2m1(r1+δ)c_{1}=2ks_{0}+2m_{1}(r_{1}+\delta)

and

(5.31) c2=r0+δ0((r1+δ)m1m1+k+s0km1+k).c_{2}=r_{0}+\delta_{0}-\Bigl{(}(r_{1}+\delta)\frac{m_{1}}{m_{1}+k}+s_{0}\frac{k}{m_{1}+k}\Bigr{)}.

We have the lower bound

c2c2=r0+δ0((r1+δ)m1m1+k+M0km1+k).c_{2}\geq c_{2}^{\prime}=r_{0}+\delta_{0}-\Bigl{(}(r_{1}+\delta)\frac{m_{1}}{m_{1}+k}+M_{0}\frac{k}{m_{1}+k}\Bigr{)}.

Fix m1m_{1} large enough so that

(5.32) m116M0r0+δ0r1δ,\displaystyle m_{1}\geq\frac{16\ell M_{0}}{r_{0}+\delta_{0}-r_{1}-\delta}\,,
(5.33) m1(r1δ)>(k+2)(r1+δ),\displaystyle m_{1}(r_{1}-\delta)>(k+2\ell)(r_{1}+\delta),
(5.34) c2>0andc2(k0(m1+k)2)6dM04m1+3(k+1)(+1).\displaystyle c_{2}^{\prime}>0\quad\text{and}\quad\frac{c_{2}^{\prime}\bigl{(}k_{0}(m_{1}+k)-2\ell\bigr{)}}{6dM_{0}}\geq 4m_{1}+3(k+1)(\ell+1).

Note that after fixing m1m_{1}, (5.32) and (5.33) remain true as we shrink δ\delta and (5.34) remains true with c2c_{2} in place of c2c_{2}^{\prime} as we increase s0s_{0} towards M0M_{0}.

Set three size-determining integer parameters as

(5.35) 1=k0(m1+k),2=12,and2′′=32.\ell_{1}=k_{0}(m_{1}+k),\quad\ell_{2}^{\prime}=\ell_{1}-2\ell,\quad\text{and}\quad\ell_{2}^{\prime\prime}=3\ell_{2}^{\prime}.

Set

(5.36) m2=c226dM0=c2(k0(m1+k)2)6dM04m1+3(k+1)(+1)m_{2}=\biggl{\lfloor}\frac{c_{2}\ell_{2}^{\prime}}{6dM_{0}}\biggr{\rfloor}=\biggl{\lfloor}\frac{c_{2}\bigl{(}k_{0}(m_{1}+k)-2\ell\bigr{)}}{6dM_{0}}\biggr{\rfloor}\geq 4m_{1}+3(k+1)(\ell+1)

where we appealed to (5.34).

As the last step fix NN so that N22N-2\ell_{2}^{\prime} is a multiple of 1\ell_{1} and large enough so that

(5.37) Q=c2N4d(2′′+1)M0c1c2N/2.Q=c_{2}N-4d(\ell_{2}^{\prime\prime}+\ell_{1})M_{0}-c_{1}\geq c_{2}N/2.

Increasing NN may force us to take s0s_{0} closer to M0M_{0} to maintain the crossing bound (5.9). As observed above, this can be done while maintaining all the inequalities above.

We perform a construction within each NN-box BB. Let VV be a box inside BB that is tiled with cubes ViV_{i} of the form j=1d[uj,uj+1]\prod_{j=1}^{d}[u_{j},u_{j}+\ell_{1}] where (u1,,ud)d(u_{1},\dotsc,u_{d})\in\mathbb{Z}^{d} is the lower left corner of the cube and the side-length 1\ell_{1} comes from (5.35). The cubes ViV_{i} are nonoverlapping but neighboring cubes share a (d1)(d-1)-dimensional face. Then, V=i=1αViV=\bigcup_{i=1}^{\alpha}V_{i} where α=3d11d(N22)d\alpha=3^{d-1}\ell_{1}^{-d}({N-2\ell_{2}^{\prime}})^{d} is the number of cubes required to tile VV. Inside box BB, VV is surrounded by an annular region BVB\setminus V whose thickness (perpendicular distance from a face of VV to BcB^{c}) is 2\ell_{2}^{\prime} in the direction where BB has width NN and 2′′\ell_{2}^{\prime\prime} in the other directions.

A boundary edge of a cube VjV_{j} is one of the 2d1d2^{d-1}d line segments (one-dimensional faces) of length 1\ell_{1} that lie on the boundary Vj\partial V_{j}.

m1/2m_{1}/2kkm1m_{1}m1/2m_{1}/2kk\ell1\ell_{1}
Figure 5.2. k+2k+2\ell-detours attached to the south and west boundaries of 1×1\ell_{1}\times\ell_{1} 22-faces. In this illustration each edge has k0=2k_{0}=2 detours attached to it, spaced m1m_{1} apart.

Attach (k+2)(k+2\ell)-detours along each of the boundary edges of the tiling so that the kk-path π+\pi^{+} is on the boundary edge and the detour π++\pi^{++} is in the interior of one of the two-dimensional faces adjacent to this boundary edge. Adopt the convention that if the boundary edge is [v,v+1𝐞i][v,v+\ell_{1}\mathbf{e}_{i}] then the detour lies on the 2-dimensional face [v,v+1𝐞i]×[v,v+1𝐞j][v,v+\ell_{1}\mathbf{e}_{i}]\times[v,v+\ell_{1}\mathbf{e}_{j}] for some jij\neq i (in other words, the detour points into a positive coordinate direction). See Figure 5.2.

Place k0k_{0} detours on each boundary edge of the tiling so that the detours are exactly distance m1m_{1} apart from each other and a detour that is right next to a corner vertex of the tiling is exactly distance m1/2m_{1}/2 from that vertex. This is consistent with the definition of 1\ell_{1} in (5.35).

Since m1/2>m_{1}/2>\ell by (5.11) and (5.32), distinct detour rectangles that happen to lie on the same two-dimensional face do not intersect and the points on a detour are closer to the boundary edge of the detour than to any other boundary edge.

Inside a particular NN-box BB, for j{0,1,2}j\in\{0,1,2\} let WjW_{j} denote the union of the jj-dimensional faces of the cubes {Vi}\{V_{i}\} tiling VV. Let W1W_{1}^{\prime} be the union of W1W_{1} (the boundary edges) and the detours π++\pi^{++} attached to the boundary edges.

We describe in more detail the structure of the detours on the two-dimensional faces inside a particular BB. Let HW2H\subset W_{2} be a two-dimensional 1×1\ell_{1}\times\ell_{1} face. For simplicity of notation suppose H=[0,1𝐞1]×[0,1𝐞2]H=[0,\ell_{1}\mathbf{e}_{1}]\times[0,\ell_{1}\mathbf{e}_{2}]. Assume without loss of generality that the boundary edge [0,1𝐞1][0,\ell_{1}\mathbf{e}_{1}] has its detours contained in HH. For i[k0]i\in[k_{0}] define the iith detour rectangle:

Gi,S=[(m1/2+(i1)(k+m1))𝐞1,(m1/2+(i1)(k+m1)+k)𝐞1]×[0,𝐞2].G_{i,S}=\bigl{[}\bigl{(}m_{1}/2+(i-1)(k+m_{1})\bigr{)}\mathbf{e}_{1},\bigl{(}m_{1}/2+(i-1)(k+m_{1})+k\bigr{)}\mathbf{e}_{1}]\times[0,\ell\mathbf{e}_{2}].

The subscript SS identifies these detour rectangles as attached to the southern boundary of HH. Similarly, if the detour rectangles attached to the western boundary of HH lie in HH, we denote these by {Gi,W:1ik0}\{G_{i,W}:1\leq i\leq k_{0}\}.

For a label U{S,W}U\in\{S,W\}, let πi,U++=Gi,UH\pi^{++}_{i,U}=\partial G_{i,U}\setminus\partial H be the portion of the boundary of Gi,UG_{i,U} in the interior of HH. πi,U++\pi^{++}_{i,U} is the detour path of k+2k+2\ell edges. Let πi,U+=Gi,UH\pi^{+}_{i,U}=\partial G_{i,U}\cap\partial H be the portion of the boundary of Gi,UG_{i,U} that lies on the boundary of HH. πi,U+\pi^{+}_{i,U} is a straight path of kk edges, the path bypassed by the detour. Let

(5.38) H=HU{S,W}1ik0Gi,U,H¯=HU{S,W}1ik0Gi,U,andH+=U{S,W}1ik0πi,U+.H^{\prime}=\partial H\cup\!\!\!\!\!\bigcup_{\stackrel{{\scriptstyle 1\leq i\leq k_{0}}}{{U\in\{S,W\}}}}\!\!\!\!\partial G_{i,U},\quad\overline{H}=\partial H\cup\!\!\!\!\!\bigcup_{\stackrel{{\scriptstyle 1\leq i\leq k_{0}}}{{U\in\{S,W\}}}}\!\!\!\!G_{i,U},\quad\text{and}\quad H^{+}=\!\!\!\!\!\bigcup_{\stackrel{{\scriptstyle 1\leq i\leq k_{0}}}{{U\in\{S,W\}}}}\!\!\!\!\!\pi^{+}_{i,U}.

See Figure 5.3. Let W¯1\overline{W}_{\!1} (resp. W1+W^{+}_{1}) be the union of all H¯\overline{H} (resp. H+H^{+}) as HH ranges over all the two-dimensional faces that lie in W2W_{2}. The union of all HH^{\prime} equals W1W^{\prime}_{1} as already defined above.

HHHH^{\prime}H¯\overline{H}H+H^{+}
Figure 5.3. From left to right: a two-dimensional face HH (shaded), HH^{\prime} that consists of the boundary H\partial H of HH and the boundaries of the detour rectangles in HH, H¯\overline{H} that consists of H\partial H and the full (shaded) detour rectangles in HH, and finally H+H^{+} that consists of the π+\pi^{+}-parts of the boundaries of the detour rectangles in HH.

Since multiple geodesics are possible, we have to make a particular measurable choice of a geodesic to work on and one that relates suitably to the structure defined above. For this purpose order the admissible steps for example as in

(5.39) 𝐞1𝐞1𝐞2𝐞2𝐞d𝐞d\varnothing\prec\mathbf{e}_{1}\prec-\mathbf{e}_{1}\prec\mathbf{e}_{2}\prec-\mathbf{e}_{2}\prec\dotsm\prec\mathbf{e}_{d}\prec-\mathbf{e}_{d}

and then order the paths lexicographically. Here \varnothing stands for a missing step. So if π\pi^{\prime} extends π\pi with one or more steps, then ππ\pi\prec\pi^{\prime} in lexicographic ordering. Recall the choice of index j(x)j(x) in (5.17).

Lemma 5.7.

Fix xd{𝟎}x\in\mathbb{Z}^{d}\setminus\{\mathbf{0}\}. There exists a unique geodesic π\pi for T𝟎,xT_{\mathbf{0},x} that satisfies the following two conditions.

  1. (i)

    For every NN-box Bj(x)B\in\mathcal{B}_{j(x)} and points u,vπB=πBu,v\in\pi_{B}=\pi\cap B the following holds: if both u,vW¯1u,v\in\overline{W}_{\!1} or uW¯1u\in\overline{W}_{\!1} and vBv\in\partial B (or vice versa), and if every edge of πu,v\pi_{u,v} lies in BB but not in W¯1\overline{W}_{\!1}, then there is no geodesic between uu and vv that remains in BB, uses only edges with strictly positive weights, and uses at least one edge in W¯1\overline{W}_{\!1}.

  2. (ii)

    π\pi is lexicographically first among all geodesics of T𝟎,xT_{\mathbf{0},x} that satisfy point (i).

Proof.

It suffices to show the existence of a geodesic that satisfies point (i). Point (ii) then picks a unique one.

Start with any T𝟎,xT_{\mathbf{0},x}-geodesic π\pi of maximal Euclidean length. For the purpose of this proof consider π\pi as an ordered sequence of vertices and the edges connecting them.

Consider in order each segment πu,v\pi_{u,v} that violates point (i). When this violation happens, there is a particular NN-box Bj(x)B\in\mathcal{B}_{j(x)} such that πu,vBW¯1\pi_{u,v}\subset B\setminus\overline{W}_{\!1} and there is an alternative geodesic πu,vB\pi^{\prime}_{u,v}\subset B that uses only edges with strictly positive weights and uses at least one edge in W¯1\overline{W}_{\!1}. Replace the original segment πu,v\pi_{u,v} with πu,v\pi^{\prime}_{u,v}.

Since we replaced one geodesic segment with another, T(πu,v)=T(πu,v)T(\pi^{\prime}_{u,v})=T(\pi_{u,v}). Suppose that after the replacement, the full path is no longer self-avoiding. Then a portion of it can be removed and this portion contains part of πu,v\pi^{\prime}_{u,v}. Since πu,v\pi^{\prime}_{u,v} uses only edges with strictly positive weights, this removal reduces the passage time by a strictly positive amount, contradicting the assumption that the original passage time was optimal. Consequently the new path is still a self-avoiding geodesic.

Since the original path was a geodesic of maximal Euclidean length, it follows that |πu,v||πu,v||\pi^{\prime}_{u,v}|\leq|\pi_{u,v}|. Since the replacement inserted into the geodesic at least one new edge from W¯1\overline{W}_{\!1}, πu,v\pi^{\prime}_{u,v} has strictly fewer edges in BW¯1B\setminus\overline{W}_{\!1} than πu,v\pi_{u,v}.

The new segment πu,v\pi^{\prime}_{u,v} may in turn contain smaller segments πu1,v1,,πum,vm\pi^{\prime}_{u_{1},v_{1}},\dotsc,\pi^{\prime}_{u_{m},v_{m}} that violate point (i). Replace each of these with alternative segments πu1,v1′′,,πum,vm′′\pi^{\prime\prime}_{u_{1},v_{1}},\dotsc,\pi^{\prime\prime}_{u_{m},v_{m}}. Continue like this until the entire path segment between uu and vv has been cleaned up, in the sense that no smaller segment of it violates (i). This process must end because each replacement leaves strictly shorter segments that can potentially violate point (i).

Observe that the clean-up of the segment πu,v\pi_{u,v} happens entirely inside the particular NN-box BB, does not alter the endpoints u,vu,v of the original segment, and does not alter the other portions π𝟎,u\pi_{\mathbf{0},u} and πv,x\pi_{v,x} of the geodesic because each replacement step produced a self-avoiding geodesic.

Proceed in this manner through all the path segments that are in violation of point (i). There are only finitely many. At the conclusion of this process we have a geodesic that satisfies point (i). ∎

Define the event

(5.40) ΓB={ω:\displaystyle\Gamma_{B}=\Big{\{}\,\omega: r1δ<t(e)<r1+δeW1W1+,\displaystyle r_{1}-\delta<t(e)<r_{1}+\delta\quad\forall e\in W_{1}^{\prime}\,\setminus W^{+}_{1},
s0δ<t(e)s0eW1+,\displaystyle s_{0}-\delta<t(e)\leq s_{0}\quad\forall e\in W^{+}_{1},
s0t(e)s1eW¯1W1,\displaystyle s_{0}\leq t(e)\leq s_{1}\quad\forall e\in\overline{W}_{\!1}\,\setminus W_{1}^{\prime},
s1t(e)M0eBW¯1}.\displaystyle s_{1}\leq t(e)\leq M_{0}\quad\forall e\in B\setminus\overline{W}_{\!1}\,\Big{\}}.

A key consequence of the definition of the event ΓB\Gamma_{B} is that, by (5.27), the boundary paths π+\pi^{+} and π++\pi^{++} of all detour rectangles GG in W1W_{1}^{\prime} satisfy

(5.41) T(π+)<T(π++)<T(π+)+(21)b.T(\pi^{+})<T(\pi^{++})<T(\pi^{+})+(2\ell-1)b.

Once the parameters have been fixed, then up to translations and rotations there are only finitely many ways to choose the constructions above. Thus

(5.42) D2>0 such that (ΓB)D2 for all B.\exists D_{2}>0\text{ such that }\mathbb{P}(\Gamma_{B})\geq D_{2}\text{ for all }B.

D2D_{2} depends on NN and the probabilities of the events on t(e)t(e) that appear in ΓB\Gamma_{B}. In particular, D2D_{2} does not depend on xx.

Our point of view shifts now to the implications of the event ΓB\Gamma_{B} for a particular Bj(x)B\in\mathcal{B}_{j(x)}.

Let γ\gamma be a self-avoiding path in W1W_{1}. Then if ωΓB\omega\in\Gamma_{B},

(5.43) T(γ)|γ|1(s0km1+k+(r1+δ)m1m1+k)+c1T(\gamma)\leq\left|{\gamma}\right|_{1}\left(s_{0}\frac{k}{m_{1}+k}+(r_{1}+\delta)\frac{m_{1}}{m_{1}+k}\right)+c_{1}

where c1c_{1} came from (5.30). The main term on the right of (5.43) contains the weights of the kk-paths of detours and m1m_{1}-gaps completely covered by γ\gamma, and c1c_{1} accounts for the partially covered pieces at either end of γ\gamma.

We say that a point yW¯1y\in\overline{W}_{\!1} is associated with a boundary edge II of a cube Vi0V_{i_{0}} if either yIy\in I or yy lies in one of the detour rectangles Gi,UG_{i,U} attached to the edge II. We say that points y,zW¯1y,z\in\overline{W}_{\!1} are (1,W1)(\ell^{1},W_{1})-related if they are each associated to boundary edges IVi0I\subset V_{i_{0}} and JVj0J\subset V_{j_{0}} such that every point on II can be connected to every point on JJ by an 1\ell^{1}-path that remains entirely within W1W_{1}. Recall that an 1\ell^{1}-path xm:nx_{m\hskip 0.6pt{\boldsymbol{:}}\hskip 0.3ptn} satisfies |xnxm|1=nm|x_{n}-x_{m}|_{1}=n-m.

Lemma 5.8.

Let ωΓB\omega\in\Gamma_{B}. Let y,zW¯1y,z\in\overline{W}_{\!1} be two (1,W1)(\ell^{1},W_{1})-related points. Suppose a geodesic between yy and zz lies within BB. Then there exists a geodesic between yy and zz that stays within BB and uses at least one edge in W¯1\overline{W}_{\!1}.

Proof.

There are two cases:

  1. (A)

    y,zy,z are connected by an 1\ell^{1}-path inside W¯1\overline{W}_{\!1}.

  2. (B)

    y,zy,z cannot be connected by an 1\ell^{1}-path that remains entirely inside W¯1\overline{W}_{\!1}.

In case (A), any 1\ell^{1}-path inside W¯1\overline{W}_{\!1} takes weights that are at most s1s_{1} and any path inside BW¯1B\setminus\overline{W}_{\!1} takes weights that are at least s1s_{1}. Since we assume the existence of a geodesic between yy and zz that lies entirely inside BB, we see that there must exist a geodesic that remains entirely within W¯1\overline{W}_{\!1}.

In case (B), suppose π^B\hat{\pi}\subset B is a self-avoiding path between yy and zz that lies outside W¯1\overline{W}_{\!1}. Construct a path πW¯1\pi^{\prime}\subset\overline{W}_{\!1} from yy to zz by concatenating the following path segments: using at most \ell steps, connect yy to the closest point yy^{\prime} on the boundary edge II that yy is associated with; using at most \ell steps, connect zz to the closest point zz^{\prime} on the boundary edge JJ that zz is associated with; connect yy^{\prime} to zz^{\prime} with an 1\ell^{1}-path π′′\pi^{\prime\prime} in W1W_{1}. We show that T(π)T(π^)T(\pi^{\prime})\leq T(\hat{\pi}), thus proving the lemma.

We argue that

(5.44) π′′\pi^{\prime\prime} uses at least m1/2m_{1}/2 edges in W1W1+W_{1}^{\prime}\setminus W^{+}_{1}.

Indeed, observe that yy and zz cannot both be on W1W_{1} nor both in the same detour rectangle Gi,UG_{i,U}, for otherwise we would be in case (A). On the other hand, if yy is in a detour rectangle and zz is on W1W_{1}, then π′′\pi^{\prime\prime} is an 1\ell^{1}-path that connects yy^{\prime} to z=zz^{\prime}=z. If in this case |π′′(W1W1+)|1<m1/2\left|{\pi^{\prime\prime}\cap(W_{1}^{\prime}\setminus W^{+}_{1})}\right|_{1}<m_{1}/2, then it must be the case that π′′I=J\pi^{\prime\prime}\subset I=J. But then in this case π\pi^{\prime} is an 1\ell^{1}-path from yy to zz and we are again in case (A). The symmetric case of yW1y\in W_{1} and zz in a detour rectangle is similar. Lastly, if yy and zz belong to different detour rectangles, then the segment of π′′\pi^{\prime\prime} that connects the two rectangles must be of length at least m1m_{1}, the distance between two neighboring detours.

We have verified (5.44). From (5.44) and m18m_{1}\geq 8\ell comes the lower bound

|zy|1m1/22m1/4.\left|{z-y}\right|_{1}\geq m_{1}/2-2\ell\geq m_{1}/4.

The m1/2m_{1}/2 edges in π′′(W1W1+)\pi^{\prime\prime}\cap(W_{1}^{\prime}\setminus W^{+}_{1}) all have weight at most r1+δr_{1}+\delta. Furthermore, |π′′|1|zy|1+2|\pi^{\prime\prime}|_{1}\leq\left|{z-y}\right|_{1}+2\ell and all the edges along π′′\pi^{\prime\prime} have weight no larger than s0s_{0}. This gives the bound

T(π′′)m1(r1+δ)/4+(|zy|1m1/4+2)s0.T(\pi^{\prime\prime})\leq m_{1}(r_{1}+\delta)/4+(\left|{z-y}\right|_{1}-m_{1}/4+2\ell)s_{0}.

Since π^\hat{\pi} connects yy to zz and the weights along π^\hat{\pi} are at least s1s_{1},

T(π^)m1s1/4+(|zy|1m1/4)s1.T(\hat{\pi})\geq m_{1}s_{1}/4+(\left|{z-y}\right|_{1}-m_{1}/4)s_{1}.

Together these observations give the lower bound

T(π^)T(π′′)m1(s1(r1+δ))/42s0.T(\hat{\pi})-T(\pi^{\prime\prime})\geq m_{1}(s_{1}-(r_{1}+\delta))/4-2\ell s_{0}.

From this,

T(π)T(π′′)+2s1T(π^)m1(s1(r1+δ))/4+4s1<T(π^).\displaystyle T(\pi^{\prime})\leq T(\pi^{\prime\prime})+2\ell s_{1}\leq T(\hat{\pi})-m_{1}(s_{1}-(r_{1}+\delta))/4+4\ell s_{1}<T(\hat{\pi}).

The last inequality used (5.32) and r1+δ<r0+δ0<s1M0r_{1}+\delta<r_{0}+\delta_{0}<s_{1}\leq M_{0}. ∎

\ellkk1\ell_{1}1\ell_{1}22\ell22\ell\ellIIGSG_{S}
Figure 5.4. The proof of Lemma 5.9. The light grid is W1W_{1}. The thicker square and its two detours are part of W¯1\overline{W}_{\!1}. The thickest edge of the square is denoted by II. The hashed box, denoted by GSG_{S}, is a detour rectangle attached to II. The points that are within distance 12\ell_{1}-2\ell from a point on IGSI\cup G_{S} are all inside the dashed rectangle. All these points that are also on W1W_{1} can be reached from any point on II via an 1\ell^{1} path that stays on W1W_{1}.
Lemma 5.9.

Let ωΓB\omega\in\Gamma_{B}. Suppose y,zW¯1y,z\in\overline{W}_{\!1} are not (1,W1)(\ell^{1},W_{1})-related and that they are connected by a path π^\hat{\pi} that remains entirely in BW¯1B\setminus\overline{W}_{\!1}. Then

T(π^)s1(12).T(\hat{\pi})\geq s_{1}(\ell_{1}-2\ell).
Proof.

Inspection of Figure 5.4 convinces that any two points y,zW¯1y,z\in\overline{W}_{\!1} such that |zy|1<12|z-y|_{1}<\ell_{1}-2\ell must be (1,W1)(\ell^{1},W_{1})-related. Thus |π^|112|\hat{\pi}|_{1}\geq\ell_{1}-2\ell and by assumption it uses only weights s1\geq s_{1}. ∎

Lemma 5.10.

Let ωΓB\omega\in\Gamma_{B} and y,zBy,z\in B. Assume that either both y,zW¯1y,z\in\overline{W}_{\!1} or that yW¯1y\in\overline{W}_{\!1} and zBz\in\partial B. Let π\pi be a geodesic between yy and zz. Assume that the edges of π\pi lie entirely outside W¯1\overline{W}_{\!1}. Then either there is a geodesic between yy and zz inside BB that uses at least one edge in W¯1\overline{W}_{\!1} or

(5.45) T(π)min{s1(12),s12}=s12.T(\pi)\geq\min\bigl{\{}s_{1}(\ell_{1}-2\ell),\,s_{1}\ell_{2}^{\prime}\bigr{\}}=s_{1}\ell_{2}^{\prime}.
Proof.

If π\pi reaches the boundary B\partial B (in either case of y,zy,z) then π\pi must travel through BVB\setminus V and consequently T(π)s12T(\pi)\geq s_{1}\ell_{2}^{\prime}. The other possibility is that π\pi stays inside BW¯1B\setminus\overline{W}_{\!1}. If yy and zz are (1,W1)(\ell^{1},W_{1})-related then Lemma 5.8 gives a geodesic in BB that uses an edge in W¯1\overline{W}_{\!1}. If yy and zz are not (1,W1)(\ell^{1},W_{1})-related, Lemma 5.9 gives T(π)s1(12)T(\pi)\geq s_{1}(\ell_{1}-2\ell). The last equality of (5.45) is from (5.35). ∎

Henceforth we often work with two coupled environments ω\omega and ω\omega^{*}. Quantities calculated in the ω\omega^{*} environment will be marked with a star if ω\omega^{*} is not explicitly present. For example, T𝟎,x=T𝟎,x(ω)T^{*}_{\mathbf{0},x}=T_{\mathbf{0},x}(\omega^{*}) denotes the passage time between 𝟎\mathbf{0} and xx in the environment ω\omega^{*}.

Recall the event ΛB,v,w,x\Lambda_{B,v,w,x} defined in (5.18).

Lemma 5.11.

Let ω\omega and ω\omega^{*} be two environments that agree outside BB and satisfy ωΛB,v,w,x\omega\in\Lambda_{B,v,w,x} and ωΓB\omega^{*}\in\Gamma_{B}. Then there exists a self-avoiding path π~\widetilde{\pi} from 𝟎\mathbf{0} to xx such that

T(π~)T(π(x))Q.T^{*}(\widetilde{\pi})\leq T(\pi(x))-Q.
Proof.

Since box BB is black on the event ΛB,v,w,x\Lambda_{B,v,w,x},

T(πv,w(x))>(r0+δ0)(|wv|1N).\displaystyle T(\pi_{v,w}(x))>(r_{0}+\delta_{0})(|w-v|_{1}\vee N).

The bound above comes from (5.5), on account of these observations: regardless of whether πv,w(x)\pi_{v,w}(x) exits B¯\overline{B}, there is a segment inside B¯\overline{B} of length |wv|1|w-v|_{1}, and furthermore πv,w(x)\pi_{v,w}(x) contains a short crossing of BB that has length at least NN.

Define a path π\pi^{\prime} from vv to ww in BB as follows. Let λ1\lambda_{1} be an 1\ell^{1}-path from vv to some point aW1a\in W_{1}. Similarly, let λ3\lambda_{3} be an 1\ell^{1}-path from ww to some bW1b\in W_{1}. These paths satisfy |λ1|1|λ3|1d2′′+(d2)1\left|{\lambda_{1}}\right|_{1}\vee\left|{\lambda_{3}}\right|_{1}\leq d\ell_{2}^{\prime\prime}+(d-2)\ell_{1}. Let λ2\lambda_{2} be a shortest path from aa to bb that remains in W1W_{1}. Since |ab|1|vw|1+2d2′′+2(d2)1\left|{a-b}\right|_{1}\leq\left|{v-w}\right|_{1}+2d\ell_{2}^{\prime\prime}+2(d-2)\ell_{1}, |λ2|1|vw|1+2d2′′+2d1\left|{\lambda_{2}}\right|_{1}\leq\left|{v-w}\right|_{1}+2d\ell_{2}^{\prime\prime}+2d\ell_{1}. (To go from aa to bb along W1W_{1} use 212\ell_{1} steps to go from aa and bb to the nearest vertices aa^{\prime} and bb^{\prime} in W0W_{0}, respectively, and an 1\ell^{1}-path along W1W_{1} will take |ab|1|ab|1+21\left|{a^{\prime}-b^{\prime}}\right|_{1}\leq\left|{a-b}\right|_{1}+2\ell_{1} steps.)

Let π\pi^{\prime} be the concatenation of λ1\lambda_{1}, λ2\lambda_{2} and λ3\lambda_{3}. Define π~\widetilde{\pi} as the concatenation of π𝟎,v(x)\pi_{\mathbf{0},v}(x), π\pi^{\prime}, and πw,x(x)\pi_{w,x}(x). The next calculation uses (5.43), (5.37) and (5.31), and the facts that ω=ω\omega=\omega^{*} outside BB, ωΛB,v,w,x\omega\in\Lambda_{B,v,w,x} and ωΓB\omega^{*}\in\Gamma_{B}.

T(π(x))T(π~)=T(πv,w(x))T(π)\displaystyle T(\pi(x))-T^{*}(\widetilde{\pi})=T(\pi_{v,w}(x))-T^{*}(\pi^{\prime})
(r0+δ0)max(|vw|1,N)2(d2′′+(d2)1)M0\displaystyle\quad\geq(r_{0}+\delta_{0})\max(\left|{v-w}\right|_{1},N)-2(d\ell_{2}^{\prime\prime}+(d-2)\ell_{1})M_{0}
(|vw|1+2d2′′+2d1)(s0km1+k+(r1+δ)m1m1+k)c1\displaystyle\quad\quad-(\left|{v-w}\right|_{1}+2d\ell_{2}^{\prime\prime}+2d\ell_{1})\left(s_{0}\frac{k}{m_{1}+k}+(r_{1}+\delta)\frac{m_{1}}{m_{1}+k}\right)-c_{1}
(r0+δ0)max(|vw|1,N)2(d2′′+(d2)1)M0\displaystyle\quad\geq(r_{0}+\delta_{0})\max(\left|{v-w}\right|_{1},N)-2(d\ell_{2}^{\prime\prime}+(d-2)\ell_{1})M_{0}
(max(|vw|1,N)+2d2′′+2d1)(s0km1+k+(r1+δ)m1m1+k)c1\displaystyle\quad\quad-\bigl{(}\max(\left|{v-w}\right|_{1},N)+2d\ell_{2}^{\prime\prime}+2d\ell_{1}\bigr{)}\left(s_{0}\frac{k}{m_{1}+k}+(r_{1}+\delta)\frac{m_{1}}{m_{1}+k}\right)-c_{1}
=(r0+δ0)max(|vw|1,N)2(d2′′+(d2)1)M0\displaystyle\quad=(r_{0}+\delta_{0})\max(\left|{v-w}\right|_{1},N)-2(d\ell_{2}^{\prime\prime}+(d-2)\ell_{1})M_{0}
(max(|vw|1,N)+2d2′′+2d1)(r0+δ0c2)c1\displaystyle\quad\quad-\bigl{(}\max(\left|{v-w}\right|_{1},N)+2d\ell_{2}^{\prime\prime}+2d\ell_{1}\bigr{)}(r_{0}+\delta_{0}-c_{2})-c_{1}
=c2max(|vw|1,N)2(d2′′+(d2)1)M0(2d2′′+2d1)(r0+δ0)c1\displaystyle\quad=c_{2}\max(\left|{v-w}\right|_{1},N)-2(d\ell_{2}^{\prime\prime}+(d-2)\ell_{1})M_{0}-(2d\ell_{2}^{\prime\prime}+2d\ell_{1})(r_{0}+\delta_{0})-c_{1}
c2N4d(2′′+1)M0c1=Q.\displaystyle\quad\geq c_{2}N-4d(\ell_{2}^{\prime\prime}+\ell_{1})M_{0}-c_{1}=Q.

In the first inequality, 2(d2′′+(d2)1)M02(d\ell_{2}^{\prime\prime}+(d-2)\ell_{1})M_{0} bounds the time spent on λ1\lambda_{1} and λ3\lambda_{3} and the remaining negative terms bound the passage time of λ2\lambda_{2}. The lemma is proved. ∎

Henceforth we assume that ωΛB,v,w,x\omega\in\Lambda_{B,v,w,x} and ωΓB\omega^{*}\in\Gamma_{B}. Let π(x)\pi^{*}(x) be the geodesic from 𝟎\mathbf{0} to xx in the ω\omega^{*}-environment specified in Lemma 5.7. By Lemma 5.11,

(5.46) T(π(x))T(π~)T(π(x))QT(π(x))Q.T^{*}(\pi^{*}(x))\leq T^{*}(\widetilde{\pi})\leq T(\pi(x))-Q\leq T(\pi^{*}(x))-Q.

This implies that π(x)\pi^{*}(x) must use edges in W¯1\overline{W}_{\!1} because ω\omega and ω\omega^{*} agree outside BB, while t(e)s0s1t(e)t(e)\leq s_{0}\leq s_{1}\leq t^{*}(e) on edges in BW¯1B\setminus\overline{W}_{\!1}.

BBVVW¯1\overline{W}_{1}a0a_{0}a1=x1a_{1}=x_{1}y1y_{1}z1=x2z_{1}=x_{2}y2y_{2}z2=x3z_{2}=x_{3}y3y_{3}z3=x4z_{3}=x_{4}y4y_{4}z4=b1z_{4}=b_{1}
Figure 5.5. The path segment πa0,b1(x)\pi^{*}_{a_{0},b_{1}}(x) with four excursions π1,,π4\pi^{1},\dotsc,\pi^{4}. The segment πi,1\pi^{i,1} inside W¯1\overline{W}_{1} goes from xix_{i} to yiy_{i} and the segment πi,2\pi^{i,2} outside W¯1\overline{W}_{1} goes from yiy_{i} to ziz_{i}. Note that W¯1\overline{W}_{1} is not actually a box but is represented as one above for the purpose of illustration.

Let a0a_{0} be the first vertex of π(x)\pi^{*}(x) in BB, a1a_{1} the first vertex of π(x)\pi^{*}(x) in W¯1\overline{W}_{\!1}, and b1b_{1} the last vertex of π(x)\pi^{*}(x) in BB. Decompose the path segment πa1,b1(x)\pi^{*}_{a_{1},b_{1}}(x) between a1a_{1} and b1b_{1} into excursions π1,,πσ\pi^{1},\ldots,\pi^{\sigma} (σ\sigma\in\mathbb{N}) as follows: each excursion πi\pi^{i} begins with a nonempty segment πi,1\pi^{i,1} of edges inside W¯1\overline{W}_{\!1}, followed by a nonempty segment πi,2\pi^{i,2} of edges outside W¯1\overline{W}_{\!1}. The excursions π1,,πσ1\pi^{1},\ldots,\pi^{\sigma-1} begin and end at a vertex in W¯1\overline{W}_{\!1}, while the last excursion πσ\pi^{\sigma} begins in W¯1\overline{W}_{\!1} and ends at the vertex b1b_{1} where π(x)\pi^{*}(x) exits BB. Figure 5.5 illustrates.

By ωΓB\omega^{*}\in\Gamma_{B}, (5.27) and (5.28), r1δ>0r_{1}-\delta>0 and hence t(e)>0t^{*}(e)>0 for all edges eBe\in B. Then condition (i) of Lemma 5.7 ensures that those portions of the segments π1,2,π2,2,,πσ,2\pi^{1,2},\pi^{2,2},\ldots,\pi^{\sigma,2} that connect W¯1\overline{W}_{\!1} to itself or to B\partial B inside BB cannot be replaced by segments that use edges in W¯1\overline{W}_{\!1}. Therefore these segments obey bound (5.45). This gives the last inequality below:

T(πa0,b1(x))T(πa1,b1(x))i=1σT(πi,2)σs12.T^{*}(\pi^{*}_{a_{0},b_{1}}(x))\geq T^{*}(\pi^{*}_{a_{1},b_{1}}(x))\geq\sum_{i=1}^{\sigma}T^{*}(\pi^{i,2})\geq\sigma s_{1}\ell_{2}^{\prime}.

Since the maximal side length of BB is 3N3N, a0a_{0} and b1b_{1} can be connected with a path πo\pi^{o} such that T(πo)3dNM0.T^{*}(\pi^{o})\leq 3dNM_{0}. Since π(x)\pi^{*}(x) is optimal, σs123dNM0\sigma s_{1}\ell_{2}^{\prime}\leq 3dNM_{0}, and therefore

(5.47) σ3dNM0s12.\sigma\leq\frac{3dNM_{0}}{s_{1}\ell_{2}^{\prime}}\,.

Using (5.46), and that ω=ω\omega=\omega^{*} outside BB while ωω\omega\leq\omega^{*} on BW¯1B\setminus\overline{W}_{\!1},

Q\displaystyle Q T(π(x))T(π(x))\displaystyle\leq T(\pi^{*}(x))-T^{*}(\pi^{*}(x))
=T(πa0,a1(x))T(πa0,a1(x))+T(πa1,b1(x))T(πa1,b1(x))\displaystyle=T(\pi^{*}_{a_{0},a_{1}}(x))-T^{*}(\pi^{*}_{a_{0},a_{1}}(x))+T(\pi^{*}_{a_{1},b_{1}}(x))-T^{*}(\pi^{*}_{a_{1},b_{1}}(x))
i=1σ[T(πi)T(πi)].\displaystyle\leq\sum_{i=1}^{\sigma}\bigl{[}T(\pi^{i})-T^{*}(\pi^{i})\bigr{]}.

Then some excursion π¯{π1,,πσ}\bar{\pi}\in\{\pi^{1},\ldots,\pi^{\sigma}\} must satisfy

(5.48) T(π¯)T(π¯)\displaystyle T(\bar{\pi})-T^{*}(\bar{\pi}) Qσc2Ns126dNM0=c2s126dM0.\displaystyle\geq\frac{Q}{\sigma}\geq\frac{c_{2}Ns_{1}\ell_{2}^{\prime}}{6dNM_{0}}=\frac{c_{2}s_{1}\ell_{2}^{\prime}}{6dM_{0}}.

The second inequality comes from (5.37) and (5.47). The only positive contributions to T(π¯)T(π¯)T(\bar{\pi})-T^{*}(\bar{\pi}) can come from π¯1\bar{\pi}^{1}, the segment of π¯\bar{\pi} in W¯1\overline{W}_{\!1}. Since BB is black, t(e)t(e)t(e)s0s1t(e)-t^{*}(e)\leq t(e)\leq s_{0}\leq s_{1} for all edges eBe\in B. Therefore the number of edges |π¯1|1|\bar{\pi}^{1}|_{1} satisfies s1|π¯1|1T(π¯)T(π¯)s_{1}|\bar{\pi}^{1}|_{1}\geq T(\bar{\pi})-T^{*}(\bar{\pi}). From this and (5.36)

(5.49) |π¯1|1T(π¯)T(π¯)s1c226dM0m24m1+3(k+1)(+1).|\bar{\pi}^{1}|_{1}\geq\frac{T(\bar{\pi})-T^{*}(\bar{\pi})}{s_{1}}\geq\frac{c_{2}\ell_{2}^{\prime}}{6dM_{0}}\geq m_{2}\geq 4m_{1}+3(k+1)(\ell+1).

The next lemma ensures that the path segment π¯1\bar{\pi}^{1} goes through the kk-path of at least one k+2k+2\ell-detour.

Lemma 5.12.

Let ω\omega and ω\omega^{*} be two environments that agree outside BB and satisfy ωΛB,v,w,x\omega\in\Lambda_{B,v,w,x} and ωΓB\omega^{*}\in\Gamma_{B}. Let π(x)\pi^{*}(x) be the geodesic for T𝟎,x(ω)T_{\mathbf{0},x}(\omega^{*}) chosen in Lemma 5.7. Then there exists a detour rectangle GG in BB with boundary paths (π+,π++)(\pi^{+},\pi^{++}) such that π(x)\pi^{*}(x) follows π+\pi^{+} and does not touch π++\pi^{++}, except at the endpoints shared by π+\pi^{+} and π++\pi^{++}.

Proof.

By construction, the portion π¯1\bar{\pi}^{1} of π(x)\pi^{*}(x) has a continuous path segment of length m24m1+3(k+1)(+1)m_{2}\geq 4m_{1}+3(k+1)(\ell+1) in W¯1\overline{W}_{\!1}. This forces π¯1\bar{\pi}^{1} to enter at least three k×k\times\ell detour rectangles, because these rectangles are m1m_{1} apart along W¯1\overline{W}_{\!1} and the path can use at most (k+1)(+1)(k+1)(\ell+1) edges in a gives detour rectangle. Let GG be a middle rectangle along this path segment, in other words, one that is both entered and exited, and such that π¯1\bar{\pi}^{1} covers two m1m_{1}-segments on W1W_{1} that connect GG to some neighboring detour rectangles. (Recall Figure 5.2.) We can assume without loss of generality that GG lies in the (𝐞1,𝐞2)(\mathbf{e}_{1},\mathbf{e}_{2})-plane and that it is attached to a boundary edge of some VjV_{j} that lies along 𝐞1\mathbf{e}_{1}.

Let π+\pi^{+} and π++\pi^{++} be the boundary paths of GG and aa and bb their common endpoints. Let π^=π¯a,b1\hat{\pi}=\bar{\pi}^{1}_{a,b} denote the segment of π¯1\bar{\pi}^{1} between aa and bb. For ease of language assume that π^\hat{\pi} visits aa first and then bb. We show that π^=π+\hat{\pi}=\pi^{+} by showing that all other cases are strictly worse.

Definition (5.40) of ΓB\Gamma_{B} and inequality (5.27) imply T(π++)(k+2)(r1δ)>s0kT(π+)T^{*}(\pi^{++})\geq(k+2\ell)(r_{1}-\delta)>s_{0}k\geq T^{*}(\pi^{+}) and rule out the case π^=π++\hat{\pi}=\pi^{++}. If π^\hat{\pi} coincides with neither π+\pi^{+} nor π++\pi^{++}, there are points aa^{\prime} and bb^{\prime} on G\partial G such that π^\hat{\pi} visits a,a,b,ba,a^{\prime},b^{\prime},b in this order and π=π^a,b\pi^{\prime}=\hat{\pi}_{a^{\prime},b^{\prime}} lies in the interior GGG\setminus\partial G.

If aa^{\prime} and bb^{\prime} lie on the same or on adjacent sides of G\partial G, the 1\ell^{1}-path from aa^{\prime} to bb^{\prime} along G\partial G has smaller weight than π\pi^{\prime}.

Suppose aa^{\prime} and bb^{\prime} lie on opposite \ell-sides of π++\pi^{++}. Then

T(π^)T(π^a,a)+s0k+T(π^b,b)>s0kT(π+).T^{*}(\hat{\pi})\geq T^{*}(\hat{\pi}_{a,a^{\prime}})+s_{0}k+T^{*}(\hat{\pi}_{b^{\prime},b})>s_{0}k\geq T^{*}(\pi^{+}).

The term s0ks_{0}k is a lower bound on T(π)T^{*}(\pi^{\prime}). The strict inequality comes from r1δ>0r_{1}-\delta>0 (from (5.27)) and because the segments π^a,a\hat{\pi}_{a,a^{\prime}} and π^b,b\hat{\pi}_{b^{\prime},b} are not degenerate paths. This is the case because no edge connects the interior GGG\setminus\partial G to either aa or bb.

The remaining option is that aa^{\prime} and bb^{\prime} lie on opposite kk-sides of G\partial G. Let aa^{\prime} be the first point at which π^\hat{\pi} leaves G\partial G, and let bb^{\prime} be the point of first return to G\partial G.

aabbkk\ellπ^a,b\ \widehat{\pi}_{a^{\prime},b^{\prime}}bb^{\prime}aa^{\prime}
Figure 5.6. Case 1: aa^{\prime} lies on the kk-side of π++\pi^{++} and bπ+b^{\prime}\in\pi^{+}.

Case 1. Suppose aa^{\prime} lies on the kk-side of π++\pi^{++} and bπ+b^{\prime}\in\pi^{+} (Figure 5.6). Fix coordinates as follows: aa is at the origin, a=a1𝐞1+𝐞2a^{\prime}=a_{1}^{\prime}\mathbf{e}_{1}+\ell\mathbf{e}_{2}, and b=b1𝐞1b^{\prime}=b_{1}^{\prime}\mathbf{e}_{1}. Then,

T(π^a,b)\displaystyle T^{*}(\hat{\pi}_{a,b^{\prime}}) =T(π^a,a)+T(π^a,b)\displaystyle=T^{*}(\hat{\pi}_{a,a^{\prime}})+T^{*}(\hat{\pi}_{a^{\prime},b^{\prime}})
(r1δ)|aa|1+|ab|1s0\displaystyle\geq(r_{1}-\delta)|a-a^{\prime}|_{1}+|a^{\prime}-b^{\prime}|_{1}s_{0}
=(+a1)(r1δ)+(+|b1a1|)s0\displaystyle=(\ell+a_{1}^{\prime})(r_{1}-\delta)+(\ell+|b_{1}^{\prime}-a_{1}^{\prime}|)s_{0}
2r1δ+(a1(r1δ)+|b1a1|s0).\displaystyle\geq 2\ell r_{1}-\ell\delta+\bigl{(}a_{1}^{\prime}(r_{1}-\delta)+|b_{1}^{\prime}-a_{1}^{\prime}|s_{0}\bigr{)}.

Combine the above with T(πa,b+)s0b1T^{*}(\pi^{+}_{a,b^{\prime}})\leq s_{0}b_{1}^{\prime} and develop further:

T(π^a,b)T(πa,b+)\displaystyle T^{*}(\hat{\pi}_{a,b^{\prime}})-T^{*}(\pi^{+}_{a,b^{\prime}}) 2r1δ+a1(r1δ)+|b1a1|s0s0b1\displaystyle\geq 2\ell r_{1}-\ell\delta+a_{1}^{\prime}(r_{1}-\delta)+|b_{1}^{\prime}-a_{1}^{\prime}|s_{0}-s_{0}b_{1}^{\prime}
2r1δa1(s0r1+δ)\displaystyle\geq 2\ell r_{1}-\ell\delta-a_{1}^{\prime}(s_{0}-r_{1}+\delta)
>2(r1δ)k(s0r1+δ)>0.\displaystyle>2\ell(r_{1}-\delta)-k(s_{0}-r_{1}+\delta)>0.

The last inequality is from (5.27). Thus π^\hat{\pi} cannot cross the interior of GG from π++\pi^{++} to π+\pi^{+}.

Case 2. Suppose aπ+a^{\prime}\in\pi^{+} and bb^{\prime} lies on the kk-side of π++\pi^{++}, so that a=a1𝐞1a^{\prime}=a_{1}^{\prime}\mathbf{e}_{1}, and b=b1𝐞1+𝐞2b^{\prime}=b_{1}^{\prime}\mathbf{e}_{1}+\ell\mathbf{e}_{2}. Then

T(π^a,b)\displaystyle T^{*}(\hat{\pi}_{a,b^{\prime}}) =T(π^a,a)+T(π^a,b)\displaystyle=T^{*}(\hat{\pi}_{a,a^{\prime}})+T^{*}(\hat{\pi}_{a^{\prime},b^{\prime}})
a1(s0δ)+(+|b1a1|)s0\displaystyle\geq a^{\prime}_{1}(s_{0}-\delta)+(\ell+|b_{1}^{\prime}-a_{1}^{\prime}|)s_{0}
(+b1)s0a1δ\displaystyle\geq(\ell+b_{1}^{\prime})s_{0}-a^{\prime}_{1}\delta
>(+b1)(r1+δ)+(+1)(s0r1δ)kδ\displaystyle>(\ell+b_{1}^{\prime})(r_{1}+\delta)+(\ell+1)(s_{0}-r_{1}-\delta)-k\delta
>(+b1)(r1+δ)T(πa,b++).\displaystyle>(\ell+b_{1}^{\prime})(r_{1}+\delta)\geq T^{*}(\pi^{++}_{a,b^{\prime}}).

The last strict inequality is from (5.29). Thus it is strictly better to take π++\pi^{++} from aa to bb^{\prime}.

In conclusion, π^\hat{\pi} does not coincide with π++\pi^{++}, nor does π^\hat{\pi} visit the interior of the detour rectangle. The only possibility is that π^=π+\hat{\pi}=\pi^{+}.

It remains to argue that π(x)\pi^{*}(x) does not touch π++\pi^{++} except at the endpoints aa and bb when it goes through π+\pi^{+}. Suppose on the contrary that π(x)\pi^{*}(x) visits a vertex z^\widehat{z} on π++\pi^{++}. This has to happen either before vertex aa or after vertex bb. The two cases are similar so suppose z^\widehat{z} is visited before aa. Then, by the choice of the detour rectangle GG, the segment πz^,a(x)\pi^{*}_{\widehat{z},a}(x) contains an m1m_{1}-segment on W1W_{1} that ends at aa. Hence by the definition of ΓB\Gamma_{B} and (5.33),

T(πz^,a(x))m1(r1δ)>(k+2)(r1+δ).T^{*}(\pi^{*}_{\widehat{z},a}(x))\geq m_{1}(r_{1}-\delta)>(k+2\ell)(r_{1}+\delta).

However, (k+2)(r1+δ)(k+2\ell)(r_{1}+\delta) is an upper bound on the passage time of the path from aa to z^\widehat{z} along π++\pi^{++}, which is then strictly faster than πz^,a(x)\pi^{*}_{\widehat{z},a}(x). Since πz^,a(x)\pi^{*}_{\widehat{z},a}(x) must be a geodesic, the supposed visit to z^\widehat{z} cannot happen. ∎

Stage 2 for unbounded weights.

In the unbounded weights case we construct first the k+2k+2\ell detour for a given triple (B,v,w)(B,v,w) and then the good event ΓB,v,w\Gamma_{B,v,w}. Given any k,k,\ell\in\mathbb{N}, the construction below can be carried out for all large enough NN. We label the construction below so that we can refer to it again. Figure 5.7 gives an illustration.

σ\sigma^{\prime}σ\sigmav+(k++2)𝐮v+(k+\ell+2)\mathbf{u}x0x_{0}kkπ+\pi^{+}π++\pi^{++}\ellwwv+𝐮v+\mathbf{u}^{\prime}vvv𝐮v-\mathbf{u}^{\prime}v+𝐮v+\ell\mathbf{u}v+(k++1)𝐮v+(k+\ell+1)\mathbf{u}v+(+1)𝐮v+(\ell+1)\mathbf{u}𝐮\mathbf{u}^{\prime}𝐮\mathbf{u}
Figure 5.7. Illustration of Construction 5.13. The set AA of (5.50) consists of the straight line from vv to v+(+1)𝐮v+(\ell+1)\mathbf{u} and the k×k\times\ell detour rectangle bounded by the union of the kk-path π+\pi^{+} and the k+2k+2\ell-detour π++\pi^{++}. The figure shows case (ii) of the AA-avoiding self-avoiding path from v+(k++2)𝐮v+(k+\ell+2)\mathbf{u} to ww via the point x0x_{0} on σσ\sigma\cup\sigma^{\prime}.
Construction 5.13 (The k+2k+2\ell detour for the unbounded weights case).

Fix two unit vectors 𝐮\mathbf{u} and 𝐮\mathbf{u}^{\prime} among {±𝐞i}i=1d\{\pm\mathbf{e}_{i}\}_{i=1}^{d} perpendicular to each other so that the point v+(k++2)𝐮+𝐮v+(k+\ell+2)\mathbf{u}+\ell\mathbf{u}^{\prime} lies in BB. Hence also the rectangle of size (k++2)×(k+\ell+2)\times\ell with corners vv and v+(k++2)𝐮+𝐮v+(k+\ell+2)\mathbf{u}+\ell\mathbf{u}^{\prime} lies in BB. Switch the labels 𝐮\mathbf{u} and 𝐮\mathbf{u}^{\prime} if necessary to guarantee that ww does not lie in the set

(5.50) A={v+h𝐮:0h}{v+i𝐮+j𝐮:+1ik++1, 0j}.A=\{v+h\mathbf{u}:0\leq h\leq\ell\}\cup\{v+i\mathbf{u}+j\mathbf{u}^{\prime}:\ell+1\leq i\leq k+\ell+1,\,0\leq j\leq\ell\}.

The two versions of AA obtained by interchanging 𝐮\mathbf{u} and 𝐮\mathbf{u}^{\prime} have only vv in common, so at least one of them does not contain ww.

From v+(k++2)𝐮v+(k+\ell+2)\mathbf{u} there is a self-avoiding path to ww that stays inside BB and does not intersect AA. The existence of such a path and an upper bound on the minimal length of such a path can be seen as follows.

  1. (i)

    If ww does not lie on the plane through vv spanned by {𝐮,𝐮}\{\mathbf{u},\mathbf{u}^{\prime}\}, take a minimal length path from v+(k++2)𝐮v+(k+\ell+2)\mathbf{u} to ww that begins with a step 𝐳\mathbf{z} perpendicular to this plane. Unit vector 𝐳\mathbf{z} is chosen so that (wv)𝐳>0(w-v)\cdot\mathbf{z}>0. This path will not return to the {𝐮,𝐮}\{\mathbf{u},\mathbf{u}^{\prime}\} plane and hence avoids AA. The length of this path is at most k++2+|wv|1k+\ell+2+|w-v|_{1}. This is because a possible AA-avoiding route to ww takes first the 𝐳\mathbf{z}-step from v+(k++2)𝐮v+(k+\ell+2)\mathbf{u}, then k++2k+\ell+2 (𝐮)(-\mathbf{u})-steps to v+𝐳v+\mathbf{z}, and from v+𝐳v+\mathbf{z} a minimal length path to ww. A path from vv to ww includes a 𝐳\mathbf{z}-step, hence the distance from v+𝐳v+\mathbf{z} to ww is |wv|11|w-v|_{1}-1.

  2. (ii)

    Suppose ww lies on the plane through vv spanned by {𝐮,𝐮}\{\mathbf{u},\mathbf{u}^{\prime}\}. Then we move on this plane from v+(k++2)𝐮v+(k+\ell+2)\mathbf{u} to ww and take care to avoid AA. First define the minimal AA-avoiding path σ\sigma from v+(k++2)𝐮v+(k+\ell+2)\mathbf{u} to v𝐮v-\mathbf{u}^{\prime} in k++3k+\ell+3 steps, and a minimal AA-avoiding path σ\sigma^{\prime} from v+(k++2)𝐮v+(k+\ell+2)\mathbf{u} to v+𝐮v+\mathbf{u}^{\prime} in k+3+3k+3\ell+3 steps. (We may be forced to pick between v±𝐮v\pm\mathbf{u}^{\prime} depending on which side of AA the point ww lies.) Let x0x_{0} be a closest point to ww on σσ\sigma\cup\sigma^{\prime} (possibly x0=wx_{0}=w). The AA-avoiding self-avoiding path from v+(k++2)𝐮v+(k+\ell+2)\mathbf{u} to ww then goes first to x0x_{0} along σ\sigma or σ\sigma^{\prime} and from there takes a minimal length path to ww. The length of this path is at most k+3+4+|wv|1k+3\ell+4+|w-v|_{1}.

Using the construction above, fix a self-avoiding path π\pi^{\prime} in BB from vv to ww that begins with k++2k+\ell+2 𝐮\mathbf{u}-steps from vv to v+(k++2)𝐮v+(k+\ell+2)\mathbf{u}, avoids AA after that, and has

(5.51) |π||wv|1+2k+4+6.|\pi^{\prime}|\leq|w-v|_{1}+2k+4\ell+6.

Let π+π\pi^{+}\subset\pi^{\prime} be the 𝐮\mathbf{u}-directed straight line segment of length kk from π0+=π+1=v+(+1)𝐮\pi^{+}_{0}=\pi^{\prime}_{\ell+1}=v+(\ell+1)\mathbf{u} to πk+=πk++1=v+(k++1)𝐮\pi^{+}_{k}=\pi^{\prime}_{k+\ell+1}=v+(k+\ell+1)\mathbf{u}. Let π++A\pi^{++}\subset A be the detour of length k+2k+2\ell between the endpoints π0++=π0+\pi^{++}_{0}=\pi^{+}_{0} and πk+2++=πk+\pi^{++}_{k+2\ell}=\pi^{+}_{k} defined as in (5.21). The two endpoints of π++\pi^{++} lie on π\pi^{\prime} but π++\pi^{++} is edge-disjoint from π\pi^{\prime}. This completes the construction of the k+2k+2\ell detour. \triangle

Let b>0b>0 be given. By assumption (5.10) we can choose r<sr<s in the support of t(e)t(e) so that b<r<sb<r<s. Choose k,,δk,\ell,\delta to satisfy (5.14).

Fix an element (B,v,w)(B,v,w) for a while. Define the following event ΓB,v,w\Gamma_{B,v,w} that depends only on the weights t(e)t(e) in BB. Constants s0s_{0} and δ0\delta_{0} are from definition (5.4)–(5.5) of a black NN-box BB.

(5.52) ΓB,v,w={\displaystyle\Gamma_{B,v,w}=\bigl{\{} t(e)[r0,r0+δ0/2) for eππ+,\displaystyle t(e)\in[r_{0},r_{0}+\delta_{0}/2)\,\text{ for }\,e\in\pi^{\prime}\setminus\pi^{+},
t(e)(sδ,s+δ) for eπ+,\displaystyle t(e)\in(s-\delta,s+\delta)\,\text{ for }\,e\in\pi^{+},
t(e)(rδ,r+δ) for eπ++,and\displaystyle t(e)\in(r-\delta,r+\delta)\,\text{ for }\,e\in\pi^{++},\ \ \text{and}
t(e)>s0 for eB(ππ++)}.\displaystyle t(e)>s_{0}\,\text{ for }\,e\in B\setminus(\pi^{\prime}\cup\pi^{++})\,\bigr{\}}.

By (5.14), on the event ωΓB,v,w\omega\in\Gamma_{B,v,w},

(5.53) T(π+)<T(π++)<T(π+)+(21)b.T(\pi^{+})<T(\pi^{++})<T(\pi^{+})+(2\ell-1)b.

Once NN has been fixed, then up to translations and rotations there are only finitely many ways to choose the points vv and ww on the boundary of BB and the paths π,π+,π++\pi^{\prime},\pi^{+},\pi^{++} constructed above. Thus

(5.54) D2>0 such that (ΓB,v,w)D2 for all triples (B,v,w).\exists D_{2}>0\text{ such that }\mathbb{P}(\Gamma_{B,v,w})\geq D_{2}\text{ for all triples }(B,v,w).

D2D_{2} depends on NN and the probabilities of the events on t(e)t(e) that appear in ΓB,v,w\Gamma_{B,v,w}. In particular, D2D_{2} does not depend on xx.

On the event ΛB,v,w,x\Lambda_{B,v,w,x} of (5.18), π(x)\pi(x) crosses BB, vv is the point of first entry into BB and ww the point of last exit from BB. Hence on this event we can define π¯\overline{\pi} as the self-avoiding path from 𝟎\mathbf{0} to xx obtained by concatenating the segments π𝟎,v(x)\pi_{\mathbf{0},v}(x), π\pi^{\prime}, and πw,x(x)\pi_{w,\,x}(x). For future reference at (5.57), note that π¯\overline{\pi} is edge-disjoint from π++\pi^{++}.

Lemma 5.14.

Let ω\omega and ω\omega^{*} be two environments that agree outside BB and satisfy ωΛB,v,w,x\omega\in\Lambda_{B,v,w,x} and ωΓB,v,w\omega^{*}\in\Gamma_{B,v,w}. Then π¯\overline{\pi} is a geodesic for T𝟎,x(ω)T_{\mathbf{0},x}(\omega^{*}). Furthermore, if π(x)\pi(x) was chosen to be a geodesic of maximal Euclidean length for T𝟎,x(ω)T_{\mathbf{0},x}(\omega), then π¯\overline{\pi} is a geodesic of maximal length for T𝟎,x(ω)T_{\mathbf{0},x}(\omega^{*}). The same works for minimal length.

Proof.

Since box BB is black on the event ΛB,v,w,x\Lambda_{B,v,w,x},

(5.55) T(πv,w(x))>(r0+δ0)(|wv|1N).T(\pi_{v,w}(x))>(r_{0}+\delta_{0})(|w-v|_{1}\vee N).

The bound above comes from (5.5), on account of these observations: regardless of whether πv,w(x)\pi_{v,w}(x) exits B¯\overline{B}, there is a segment inside B¯\overline{B} of length |wv|1|w-v|_{1}, and furthermore πv,w(x)\pi_{v,w}(x) contains a short crossing of BB that has length at least NN.

From ωΓB,v,w\omega^{*}\in\Gamma_{B,v,w},

T(π¯v,w)=T(π)\displaystyle T^{*}(\overline{\pi}_{v,w})=T^{*}(\pi^{\prime}) <k(s+δ)+(|wv|1+k+4+6)(r0+12δ0)\displaystyle<k(s+\delta)+(|w-v|_{1}+k+4\ell+6)(r_{0}+\tfrac{1}{2}\delta_{0})
|wv|1(r0+12δ0)+k(s+r0+δ+12δ0)+(4+6)(r0+12δ0)\displaystyle\leq|w-v|_{1}(r_{0}+\tfrac{1}{2}\delta_{0})+k(s+r_{0}+\delta+\tfrac{1}{2}\delta_{0})+(4\ell+6)(r_{0}+\tfrac{1}{2}\delta_{0})
T(πv,w(x))12(|wv|1N)δ0+C1δ0+C2\displaystyle\leq T(\pi_{v,w}(x))-\tfrac{1}{2}(|w-v|_{1}\vee N)\delta_{0}+C_{1}\delta_{0}+C_{2}
<T(πv,w(x)).\displaystyle<T(\pi_{v,w}(x)).

Before the last inequality above, Ci=Ci(k,,δ,s,r0)C_{i}=C_{i}(k,\ell,\delta,s,r_{0}) are constants determined by the quantities in parentheses. The last inequality is then guaranteed by fixing NN large enough relative to δ0\delta_{0} and these other constants. Observation (5.7) is used here.

Outside BB the weights ω\omega^{*} and ω\omega agree, and the segments π¯𝟎,v=π𝟎,v(x)\overline{\pi}_{\mathbf{0},v}=\pi_{\mathbf{0},v}(x) and π¯w,x=πw,x(x)\overline{\pi}_{w,x}=\pi_{w,x}(x) agree and lie outside BB. Hence the inequality above gives T(π¯)<T(π(x))T^{*}(\overline{\pi})<T(\pi(x)) and thereby, for any geodesic π(x)\pi^{*}(x) from 𝟎\mathbf{0} to xx in environment ω\omega^{*},

(5.56) T(π(x))T(π¯)<T(π(x)).T^{*}(\pi^{*}(x))\leq T^{*}(\overline{\pi})<T(\pi(x)).

This implies that every geodesic π(x)\pi^{*}(x) must enter BB since otherwise

T(π(x))=T(π(x))T(π(x))>T(π¯),T^{*}(\pi^{*}(x))=T(\pi^{*}(x))\geq T(\pi(x))>T^{*}(\overline{\pi}),

contradicting the optimality of π(x)\pi^{*}(x) under ω\omega^{*}.

If πB(x)ππ++\pi^{*}_{B}(x)\not\subset\pi^{\prime}\cup\pi^{++}, then π(x)\pi^{*}(x) must use an edge ee in BB with weight >s0>s_{0}. Then by property (5.4) of a black box BB, T(πB(x))s0<T(πB(x))T(\pi^{*}_{B}(x))\leq s_{0}<T^{*}(\pi^{*}_{B}(x)). Since ω\omega and ω\omega^{*} agree on BcB^{c}, we get

T(π(x))T(π(x))\displaystyle T(\pi(x))\leq T(\pi^{*}(x)) =T(πBc(x))+T(πB(x))\displaystyle=T(\pi^{*}_{B^{c}}(x))+T(\pi^{*}_{B}(x))
<T(πBc(x))+T(πB(x))=T(π(x)),\displaystyle<T^{*}(\pi^{*}_{B^{c}}(x))+T^{*}(\pi^{*}_{B}(x))=T^{*}(\pi^{*}(x)),

contradicting (5.56). Consequently πB(x)ππ++\pi^{*}_{B}(x)\subset\pi^{\prime}\cup\pi^{++}. Part of event ΛB,v,w,x\Lambda_{B,v,w,x} is that {𝟎,x}B=\{\mathbf{0},x\}\cap B=\varnothing. Thus π(x)\pi^{*}(x) must both enter and exit BB. As a geodesic π(x)\pi^{*}(x) does not backtrack on itself. Hence it must traverse the route between vv and ww. By (5.53) π+\pi^{+} is better under ω\omega^{*} than π++\pi^{++}, and hence πB(x)=π=π¯B\pi^{*}_{B}(x)=\pi^{\prime}=\overline{\pi}_{B}.

Outside BB, under both ω\omega and ω\omega^{*} since they agree on BcB^{c}, π¯Bc\overline{\pi}_{B^{c}} is an optimal union of two paths that connect the origin to one of vv and ww, and the other one of vv and ww to xx. This concludes the proof that π¯\overline{\pi} is a geodesic for T𝟎,x(ω)T_{\mathbf{0},x}(\omega^{*}).

Suppose π(x)\pi(x) is a geodesic of maximal Euclidean length under ω\omega but under ω\omega^{*} there is a geodesic π\pi^{*} strictly longer than π¯\overline{\pi}. The argument above showed πB=π¯B\pi^{*}_{B}=\overline{\pi}_{B}. Hence outside BB, πBc\pi^{*}_{B^{c}} must provide an ω\omega^{*}-geodesic from 𝟎\mathbf{0} or xx to one of vv or ww that is longer than that given by π¯Bc=πBc(x)\overline{\pi}_{B^{c}}=\pi_{B^{c}}(x). This contradicts the choice of π(x)\pi(x) as a maximal length geodesic, again because ω\omega and ω\omega^{*} agree on BcB^{c}. Same works for minimal. This completes the proof of Lemma 5.14. ∎

Stage 3 for both bounded and unbounded weights.

We choose a particular geodesic π(x)\pi(x) for T𝟎,xT_{\mathbf{0},x}. In the bounded weights case, let π(x)\pi(x) be the geodesic specified in Lemma 5.7. In the unbounded weights case, let π(x)\pi(x) be the unique lexicographically first geodesic among the geodesics of maximal Euclidean length. Let b>0b>0. For NN-boxes Bj(x)B\in\mathcal{B}_{j(x)} define the event

(5.57) ΨB,x={\displaystyle\Psi_{B,x}=\bigl{\{} inside BB \exists edge-disjoint path segments π+\pi^{+} and π++\pi^{++} that share both endpoints
and satisfy π+π(x)\pi^{+}\subset\pi(x), (π(x)π+)π++(\pi(x)\setminus\pi^{+})\cup\pi^{++} is a self-avoiding path,
|π++|=|π+|+2, and T(π+)<T(π++)<T(π+)+(21)b }.\displaystyle\text{$|\pi^{++}|=|\pi^{+}|+2\ell$, and $T(\pi^{+})<T(\pi^{++})<T(\pi^{+})+(2\ell-1)b$ }\bigr{\}}.

Couple two i.i.d. edge weight configurations ω={t(e)}ed\omega=\{t(e)\}_{e\,\in\,{\mathcal{E}}_{d}} and ω={t(e)}ed\omega^{*}=\{t^{*}(e)\}_{e\,\in\,{\mathcal{E}}_{d}} so that t(e)=t(e)t^{*}(e)=t(e) for eBe\notin B (that is, at least one endpoint of ee lies outside BB) and so that the weights {t(e)}ed\{t(e)\}_{e\,\in\,{\mathcal{E}}_{d}} and {t(e)}eB\{t^{*}(e)\}_{e\,\in\,B} are independent.

Lemma 5.12 for bounded weights (with ΓB,v,w=ΓB\Gamma_{B,v,w}=\Gamma_{B}) and Lemma 5.14 for unbounded weights imply that

{ωΛB,v,w,x}{ωΓB,v,w}{ωΨB,x}.\{\omega\in\Lambda_{B,v,w,x}\}\cap\{\omega^{*}\in\Gamma_{B,v,w}\}\,\subset\,\{\omega^{*}\in\Psi_{B,x}\}.

In particular, by inequalities (5.41) and (5.53), ωΓB,v,w\omega^{*}\in\Gamma_{B,v,w} implies T(π+)<T(π++)<T(π+)+(21)bT^{*}(\pi^{+})<T^{*}(\pi^{++})<T^{*}(\pi^{+})+(2\ell-1)b required for ωΨB,x\omega^{*}\in\Psi_{B,x}, where TT^{*} denotes passage time in the environment ω\omega^{*}.

By the independence of {ωΛB,v,w,x}\{\omega\in\Lambda_{B,v,w,x}\} and {ωΓB,v,w}\{\omega^{*}\in\Gamma_{B,v,w}\}, and then by (5.42) for bounded weights and by (5.54) for unbounded weights,

(5.58) (ΨB,x)={ωΨB,x}\displaystyle\mathbb{P}(\Psi_{B,x})=\mathbb{P}\{\omega^{*}\in\Psi_{B,x}\} {ωΛB,v,w,x}{ωΓB,v,w}D2(ΛB,v,w,x).\displaystyle\geq\mathbb{P}\{\omega\in\Lambda_{B,v,w,x}\}\mathbb{P}\{\omega^{*}\in\Gamma_{B,v,w}\}\,\geq\,D_{2}\mathbb{P}(\Lambda_{B,v,w,x}).

Let YY be the number of (B,v,w)j(x)(B,v,w)\in\mathcal{B}_{j(x)} for which ΨB,x\Psi_{B,x} occurs. By the above and (5.20),

(5.59) 𝔼[Y]\displaystyle\mathbb{E}[Y] (B,v,w)j(x)(ΨB,x)\displaystyle\geq\sum_{(B,v,w)\,\in\,\mathcal{B}_{j(x)}}\mathbb{P}(\Psi_{B,x})
(B,v,w)j(x)D2(ΛB,v,w,x)D2D1|x|1D3|x|1\displaystyle\geq\sum_{(B,v,w)\,\in\,\mathcal{B}_{j(x)}}D_{2}\mathbb{P}(\Lambda_{B,v,w,x})\geq D_{2}D_{1}|x|_{1}\equiv D_{3}|x|_{1}

for a new constant D3D_{3}.

Since we have arranged the boxes in the elements (B,v,w)j(x)(B,v,w)\in\mathcal{B}_{j(x)} separated, we can define a self-avoiding path π^\widehat{\pi} from 𝟎\mathbf{0} to xx by replacing each π+\pi^{+} segment with the π++\pi^{++} segment in each box Bj(x)B\in\mathcal{B}_{j(x)} for which event ΨB,x\Psi_{B,x} happens.

Reduce the weights on each edge ee from t(e)t(e) to t(b)(e)=t(e)bt^{(-b)}(e)=t(e)-b. By the definition of ΨB,x\Psi_{B,x}, the t(b)t^{(-b)}-passage times of the segments π+\pi^{+} and π++\pi^{++} obey this inequality:

T(b)(π++)=T(π++)b|π++|<T(π+)+(21)bb|π++|=T(b)(π+)b.\displaystyle T^{(-b)}(\pi^{++})=T(\pi^{++})-b|\pi^{++}|<T(\pi^{+})+(2\ell-1)b-b|\pi^{++}|=T^{(-b)}(\pi^{+})-b.

Consequently, along the entire path π(x)\pi(x), the replacements of π+\pi^{+} with π++\pi^{++} reduce the t(b)t^{(-b)}-passage time by at least bYbY. We get the following bound:

(5.60) T𝟎,x(b)\displaystyle T^{(-b)}_{\mathbf{0},x} T(b)(π^)<T(b)(π(x))bY=T(π(x))b|π(x)|bY\displaystyle\leq T^{(-b)}(\widehat{\pi})<T^{(-b)}(\pi(x))-bY=T(\pi(x))-b\,|\pi(x)|-bY
{T𝟎,xbL¯𝟎,xbYin the bounded weights case,=T𝟎,xbL¯𝟎,xbYin the unbounded weights case.\displaystyle\begin{cases}\leq T_{\mathbf{0},x}-b\underline{L}_{\hskip 0.9pt\mathbf{0},x}-bY&\text{in the bounded weights case,}\\[2.0pt] =T_{\mathbf{0},x}-b\overline{L}_{\hskip 0.9pt\mathbf{0},x}-bY&\text{in the unbounded weights case.}\end{cases}

The case distinction above comes because in the unbounded case |π(x)|=L¯𝟎,x|\pi(x)|=\overline{L}_{\hskip 0.9pt\mathbf{0},x} by our choice of π(x)\pi(x), while in the bounded case our choice is different, but any geodesic satisfies |π(x)|L¯𝟎,x|\pi(x)|\geq\underline{L}_{\hskip 0.9pt\mathbf{0},x}. Note that the inequality above does not require that π^\widehat{\pi} be a geodesic for T𝟎,x(b)T^{(-b)}_{\mathbf{0},x}, as long as π^\widehat{\pi} is self-avoiding.

In order to take expectations in (5.60) we restrict to b(0,r0+ε0)b\in(0,r_{0}+\varepsilon_{0}) which guarantees that 𝔼[T𝟎,x(b)]\mathbb{E}[T^{(-b)}_{\mathbf{0},x}] is finite, even if b<r0-b<-r_{0} so that weights ω(b)\omega^{(-b)} can be negative (Theorem A.1 in Appendix A). By Lemma 2.3 in [2], moment bound (2.7) with p=1p=1 is equivalent to the finite expectation 𝔼[T𝟎,x]<\mathbb{E}[T_{\mathbf{0},x}]<\infty for all xx. The inequalities above then force L¯𝟎,x\underline{L}_{\hskip 0.9pt\mathbf{0},x} and L¯𝟎,x\overline{L}_{\hskip 0.9pt\mathbf{0},x} to have finite expectations. Apply (5.59): in the bounded weights case

𝔼[T𝟎,x(b)]\displaystyle\mathbb{E}[T^{(-b)}_{\mathbf{0},x}] 𝔼[T𝟎,x]b𝔼(L¯𝟎,x)b𝔼(Y)𝔼[T𝟎,x]b𝔼(L¯𝟎,x)D3b|x|1,\displaystyle\leq\mathbb{E}[T_{\mathbf{0},x}]-b\,\mathbb{E}(\underline{L}_{\hskip 0.9pt\mathbf{0},x})-b\,\mathbb{E}(Y)\leq\mathbb{E}[T_{\mathbf{0},x}]-b\mathbb{E}(\underline{L}_{\hskip 0.9pt\mathbf{0},x})-D_{3}b|x|_{1},

while in the unbounded weights case 𝔼(L¯𝟎,x)\mathbb{E}(\underline{L}_{\hskip 0.9pt\mathbf{0},x}) is replaced by 𝔼(L¯𝟎,x)\mathbb{E}(\overline{L}_{\hskip 0.9pt\mathbf{0},x}). This completes the proof of Theorem 5.4. ∎

6. Modification proofs for nondifferentiability

In this section we consider three scenarios under which we prove that, with probability bounded away from zero, there are geodesics between two points whose lengths differ on the scale of the distance between the endpoints. The setting and modification proofs in this section borrow heavily from Section 5.

Assumption 6.1.

We assume one of these three situations for nonnegative weights.

  1. (i)

    Zero is an atom: r0=0r_{0}=0 and 0<{t(e)=0}<pc0<\mathbb{P}\{t(e)=0\}<p_{c}.

  2. (ii)

    The weights are unbounded (M0=)(M_{0}=\infty) and there exist strictly positive integers kk and \ell and atoms r1,,rk+2,s1,,skr^{\prime}_{1},\dotsc,r^{\prime}_{k+2\ell},s^{\prime}_{1},\dotsc,s^{\prime}_{k} (not necessarily all distinct) such that

    (6.1) i=1k+2ri=j=1ksj.\ \sum_{i=1}^{k+2\ell}r^{\prime}_{i}=\sum_{j=1}^{k}s^{\prime}_{j}.
  3. (iii)

    The weights are bounded (M0<)(M_{0}<\infty) and there exist strictly positive integers kk and \ell and atoms r<sr<s such that (k+2)r=ks(k+2\ell)r=ks.

Theorem 6.2.

Assume r00r_{0}\geq 0, (2.6), and the moment bound (2.7) with p>1p>1. Furthermore, assume one of the three scenarios (i)(iii) of Assumption 6.1. Then there exist constants 0<D,δ,M<0<D,\delta,M<\infty such that

(6.2) (L¯𝟎,xL¯𝟎,xD|x|1)δfor |x|1M\mathbb{P}\bigl{(}\,\overline{L}_{\hskip 0.9pt\mathbf{0},x}-\underline{L}_{\hskip 0.9pt\mathbf{0},x}\geq D|x|_{1}\bigr{)}\geq\delta\qquad\text{for $|x|_{1}\geq M$. }

Before the proof some observations about the assumptions are in order.

Remark 6.3.

Condition (6.1) of case (ii) is trivially true if zero is an atom for t(e)t(e). Since this situation is taken care of by case (i) of Assumption 6.1, let us suppose zero is not an atom. Then a necessary condition for (6.1) is that t(e)t(e) has at least two strictly positive atoms.

A sufficient condition for (6.1) is the existence of two atoms r<sr<s in (0,)(0,\infty) such that s/rs/r is rational. This is exactly the assumption on the atoms in case (iii) of Assumption 6.1. If t(e)t(e) has exactly two atoms r<sr<s in (0,)(0,\infty) and no others, then (6.1) holds if and only if s/rs/r is rational.

With more than two atoms, rational ratios are not necessary for (6.1). For example, if θ>0\theta>0 is irrational and {1,θ,1+2θ}\{1,\theta,1+2\theta\} are atoms, then (6.1) is satisfied and the ratios θ,1+2θ,θ1+2\theta,1+2\theta,\theta^{-1}+2 are irrational.

We can prove a more general result for unbounded weights because arbitrarily large weights can be used to force the geodesic to follow a specific path. With bounded weights the control of the geodesic is less precise. Hence the assumption in case (iii) is more restrictive on the atoms. \triangle

Proof of Theorem 6.2.

We prove the theorem by considering each case of Assumption 6.1 in turn.

Proof of Theorem 6.2 in Case (i) of Assumption 6.1.

We assume that zero is an atom. In this case conditions (5.3) or (5.4) are not needed for a black box, so color a box BB black if (5.5) holds. Fix NN large enough and δ0\delta_{0} small enough. Consider points xx with |x|1|x|_{1} large enough so that the Peierls estimate (5.9) is valid for n=|x|1n=|x|_{1}.

Let π(x)\pi(x) be the unique geodesic for T𝟎,xT_{\mathbf{0},x} that is lexicographically first among the geodesics of minimal Euclidean length. For this purpose order ={±𝐞1,,±𝐞d}\mathcal{R}=\{\pm\mathbf{e}_{1},\dotsc,\pm\mathbf{e}_{d}\} in some way, for example as in (5.39). The index j(x)j(x) and the event ΛB,v,w,x\Lambda_{B,v,w,x} are defined as before in (5.17) and (5.18), and estimate (5.20) holds. Let ΓB={ω:t(e)=0eB}\Gamma_{B}=\{\omega:t(e)=0\;\forall e\in B\} be the event that all edge weights in BB are zero and D2=(ΓB)>0.D_{2}=\mathbb{P}(\Gamma_{B})>0.

Given an NN-box BB, define edge weight configuration ω={t(e)}ed\omega^{*}=\{t^{*}(e)\}_{e\,\in\,{\mathcal{E}}_{d}} by setting t(e)=t(e)t^{*}(e)=t(e) for eBe\notin B (that is, at least one endpoint of ee lies outside BB) and by resampling {t(e)}eB\{t^{*}(e)\}_{e\,\in\,B} independently. Then ω\omega^{*} has the same i.i.d. distribution as the original weights ω={t(e)}ed\omega=\{t(e)\}_{e\,\in\,{\mathcal{E}}_{d}}.

Lemma 6.4.

On the event {ωΛB,v,w,x}{ωΓB}\{\omega\in\Lambda_{B,v,w,x}\}\cap\{\omega^{*}\in\Gamma_{B}\}, every geodesic from 𝟎\mathbf{0} to xx in the ω\omega^{*} environment uses at least one edge in BB.

Proof.

On the event {ωΛB,v,w,x}\{\omega\in\Lambda_{B,v,w,x}\}, π(x)\pi(x) goes through vv and ww. Let π\pi^{\prime} be an arbitrary path from vv to ww that remains inside BB and define π¯\overline{\pi} as the path from 𝟎\mathbf{0} to xx obtained by concatenating the segments π𝟎,v(x)\pi_{\mathbf{0},v}(x), π\pi^{\prime}, and πw,x(x)\pi_{w,\,x}(x). Then on the event {ωΛB,v,w,x}{ωΓB}\{\omega\in\Lambda_{B,v,w,x}\}\cap\{\omega^{*}\in\Gamma_{B}\},

T(π¯v,w)=T(π)=0<δ0(|wv|1N)<T(πv,w(x)).\displaystyle T^{*}(\overline{\pi}_{v,w})=T^{*}(\pi^{\prime})=0<\delta_{0}(|w-v|_{1}\vee N)<T(\pi_{v,w}(x)).

The justification for the last inequality was given below (5.55).

Outside BB weights ω\omega^{*} and ω\omega agree, and the segments π¯𝟎,v=π𝟎,v(x)\overline{\pi}_{\mathbf{0},v}=\pi_{\mathbf{0},v}(x) and π¯w,x=πw,x(x)\overline{\pi}_{w,x}=\pi_{w,x}(x) agree and lie outside BB. Hence the inequality above gives T(π¯)<T(π(x))T^{*}(\overline{\pi})<T(\pi(x)) and thereby, for any geodesic π(x)\pi^{*}(x) from 𝟎\mathbf{0} to xx in environment ω\omega^{*},

(6.3) T(π(x))T(π¯)<T(π(x)).T^{*}(\pi^{*}(x))\leq T^{*}(\overline{\pi})<T(\pi(x)).

This implies that every geodesic π(x)\pi^{*}(x) must use at least one edge in BB. For otherwise

T(π(x))=T(π(x))T(π(x))>T(π¯),T^{*}(\pi^{*}(x))=T(\pi^{*}(x))\geq T(\pi(x))>T^{*}(\overline{\pi}),

contradicting the optimality of π(x)\pi^{*}(x) for ω\omega^{*}. ∎

For NN-boxes BB such that 𝟎,xB\mathbf{0},x\notin B define the event

(6.4) ΨB,x={\displaystyle\Psi_{B,x}=\{ inside BB \exists path segments π+\pi^{+} and π++\pi^{++} that share both endpoints
and satisfy π+π(x)\pi^{+}\subset\pi(x), (π(x)π+)π++(\pi(x)\setminus\pi^{+})\cup\pi^{++} is a self-avoiding path,
|π++||π+|+2, and T(π+)=T(π++) }.\displaystyle\text{$|\pi^{++}|\geq|\pi^{+}|+2$, and $T(\pi^{+})=T(\pi^{++})$ }\}.

In particular, on the event ΨB,x\Psi_{B,x}, replacing π+\pi^{+} with π++\pi^{++} creates an alternative geodesic.

By Lemma 6.4, ωΨB,x\omega^{*}\in\Psi_{B,x} holds on the event {ωΛB,v,w,x}{ωΓB}\{\omega\in\Lambda_{B,v,w,x}\}\cap\{\omega^{*}\in\Gamma_{B}\}. This is seen as follows. Let π(x)\pi^{*}(x) be the lexicographically first geodesic of minimal Euclidean length in environment ω\omega^{*}. By Lemma 6.4, π(x)\pi^{*}(x) uses at least one edge in BB. Let u1u_{1} be the first and u2u_{2} the last point of π(x)\pi^{*}(x) in BB. Since {ωΓB}\{\omega^{*}\in\Gamma_{B}\} ensures that all edges in BB have zero weight and π(x)\pi^{*}(x) is a minimal length geodesic, the segment πu1,u2(x)\pi^{*}_{u_{1},u_{2}}(x) must be a path of length |u2u1|1|u_{2}-u_{1}|_{1} from u1u_{1} to u2u_{2} inside BB. Now take π+=πu1,u2(x)\pi^{+}=\pi^{*}_{u_{1},u_{2}}(x) and let π++\pi^{++} be any other path inside BB from u1u_{1} to u2u_{2} that takes more than the minimal number |u2u1|1|u_{2}-u_{1}|_{1} of steps. By the choice of u1u_{1} and u2u_{2}, the other portions π𝟎,u1(x)\pi^{*}_{\mathbf{0},u_{1}}(x) and πu2,x(x)\pi_{u_{2},x}^{*}(x) of the geodesic lie outside BB, and consequently π++\pi^{++} does not touch these paths except at the points u1u_{1} and u2u_{2}.

By the independence of {ωΛB,v,w,x}\{\omega\in\Lambda_{B,v,w,x}\} and {ωΓB}\{\omega^{*}\in\Gamma_{B}\},

(6.5) (ΨB,x)={ΨB,x occurs for ω}\displaystyle\mathbb{P}(\Psi_{B,x})=\mathbb{P}\{\text{$\Psi_{B,x}$ occurs for $\omega^{*}$}\} ({ωΛB,v,w,x}{ωΓB})\displaystyle\geq\mathbb{P}(\{\omega\in\Lambda_{B,v,w,x}\}\cap\{\omega^{*}\in\Gamma_{B}\})
={ωΛB,v,w,x}{ωΓB}D2(ΛB,v,w,x).\displaystyle=\mathbb{P}\{\omega\in\Lambda_{B,v,w,x}\}\mathbb{P}\{\omega^{*}\in\Gamma_{B}\}\,\geq\,D_{2}\mathbb{P}(\Lambda_{B,v,w,x}).

Let YY be the number of (B,v,w)j(x)(B,v,w)\in\mathcal{B}_{j(x)} for which ΨB,x\Psi_{B,x} occurs. By (5.20), for another constant D3>0D_{3}>0,

(6.6) 𝔼[Y](B,v,w)j(x)D2(ΛB,v,w,x)D3|x|1.\displaystyle\mathbb{E}[Y]\geq\sum_{(B,v,w)\in\mathcal{B}_{j(x)}}D_{2}\mathbb{P}(\Lambda_{B,v,w,x})\geq D_{3}|x|_{1}.

By Proposition 4.7(1) of [2], under the assumption {t(e)=0}<pc\mathbb{P}\{t(e)=0\}<p_{c}, for any p>0p>0 there exists a finite constant CpC_{p} such that for all xdx\in\mathbb{Z}^{d},

(6.7) 𝔼[(L¯𝟎,x)p]Cp𝔼[(T𝟎,x)p].\mathbb{E}[\,(\overline{L}_{\hskip 0.9pt\mathbf{0},x})^{p}\,]\leq C_{p}\hskip 0.7pt\mathbb{E}[\,(T_{\mathbf{0},x})^{p}\,].

By Lemma 2.3 in [2], under assumption (2.7) there exists a finite constant CC^{\prime} such that for all xdx\in\mathbb{Z}^{d}

(6.8) 𝔼[(T𝟎,x)p]C|x|1p.\mathbb{E}[\,(T_{\mathbf{0},x})^{p}\,]\leq C^{\prime}|x|_{1}^{p}.

An obvious upper bound on YY is the number of edges on the geodesic π(x)\pi(x). Let p>1p>1 be the power for which (2.7) is assumed to hold and q=pp1q=\frac{p}{p-1} its conjugate exponent. Then, by a combination of (6.6), (6.7) and (6.8),

D3|x|1𝔼(Y)\displaystyle D_{3}|x|_{1}\leq\mathbb{E}(Y) =𝔼(Y,Y<D3|x|1/2)+𝔼(Y,YD3|x|1/2)\displaystyle=\mathbb{E}(Y,\,Y<D_{3}|x|_{1}/2)+\mathbb{E}(Y,\,Y\geq D_{3}|x|_{1}/2)
D3|x|1/2+𝔼(|π(x)|,YD3|x|1/2)\displaystyle\leq D_{3}|x|_{1}/2+\mathbb{E}(|\pi(x)|,Y\geq D_{3}|x|_{1}/2)
D3|x|1/2+(𝔼[|π(x)|p])1p(YD3|x|1/2)1q\displaystyle\leq D_{3}|x|_{1}/2+\bigl{(}\mathbb{E}[\,|\pi(x)|^{p}\,]\bigr{)}^{\frac{1}{p}}\,\mathbb{P}(Y\geq D_{3}|x|_{1}/2)^{\frac{1}{q}}
D3|x|1/2+C|x|1(YD3|x|1/2)1q.\displaystyle\leq D_{3}|x|_{1}/2+C|x|_{1}\mathbb{P}(Y\geq D_{3}|x|_{1}/2)^{\frac{1}{q}}.

From this we get the bound

(Y12D3|x|1)δ3>0for large enough |x|1.\mathbb{P}\bigl{(}Y\geq\tfrac{1}{2}D_{3}|x|_{1}\bigr{)}\geq\delta_{3}>0\quad\text{for large enough $|x|_{1}$.}

Since we have arranged the boxes BB in the elements (B,v,w)j(x)(B,v,w)\in\mathcal{B}_{j(x)} separated, we can define a self-avoiding path π^(x)\widehat{\pi}(x) from 𝟎\mathbf{0} to xx by replacing each π+\pi^{+} segment of π(x)\pi(x) with the π++\pi^{++} segment in each box BB for which event ΨB,x\Psi_{B,x} happens. This path π^(x)\widehat{\pi}(x) has the same passage time T(π^(x))=T(π(x))T(\widehat{\pi}(x))=T(\pi(x)) and hence both π(x)\pi(x) and π^(x)\widehat{\pi}(x) are geodesics. By the construction, the numbers of edges on these paths satisfy |π^(x)||π(x)|+2Y|\widehat{\pi}(x)|\geq|\pi(x)|+2Y. Thus we get these inequalities between the maximal and minimal geodesic length:

L¯𝟎,x|π^(x)||π(x)|+2YL¯𝟎,x+2Y\displaystyle\overline{L}_{\hskip 0.9pt\mathbf{0},x}\geq|\widehat{\pi}(x)|\geq|\pi(x)|+2Y\geq\underline{L}_{\hskip 0.9pt\mathbf{0},x}+2Y

and then

(L¯𝟎,xL¯𝟎,xD3|x|1)(Y12D3|x|1)δ3.\mathbb{P}\bigl{(}\,\overline{L}_{\hskip 0.9pt\mathbf{0},x}-\underline{L}_{\hskip 0.9pt\mathbf{0},x}\geq D_{3}|x|_{1}\bigr{)}\geq\mathbb{P}\bigl{(}Y\geq\tfrac{1}{2}D_{3}|x|_{1}\bigr{)}\geq\delta_{3}.

(6.2) has been proved.


Proof of Theorem 6.2 in Case (ii) of Assumption 6.1.

By assumption (6.1) we can fix s1<s_{1}<\infty large enough so that, for i.i.d. copies ti,tjt_{i},t_{j}^{\prime} of the edge weight t(e)t(e),

(6.9) {tis1i[k+2],tjs1j[k], and i=1k+2ti=j=1ktj}>0.\mathbb{P}\biggl{\{}\,t_{i}\leq s_{1}\;\forall i\in[k+2\ell],\;t_{j}^{\prime}\leq s_{1}\;\forall j\in[k],\;\text{ and }\;\sum_{i=1}^{k+2\ell}t_{i}=\sum_{j=1}^{k}t_{j}^{\prime}\biggr{\}}>0.

Apply Construction 5.13 of the k+2k+2\ell detour in an NN-box BB with given boundary points vv and ww, to define paths π\pi^{\prime}, π+\pi^{+} and π++\pi^{++} in BB with |π+|=k|\pi^{+}|=k and |π++|=k+2|\pi^{++}|=k+2\ell. Define the event ΓB,v,w\Gamma_{B,v,w} that depends only on the weights t(e)t(e) in BB:

(6.10) ΓB,v,w={\displaystyle\Gamma_{B,v,w}=\Bigl{\{} t(e)[r0,r0+δ0/2) for eππ+,\displaystyle t(e)\in[r_{0},r_{0}+\delta_{0}/2)\,\text{ for }\,e\in\pi^{\prime}\setminus\pi^{+},
t(e)s1 for eπ+π++,\displaystyle t(e)\leq s_{1}\,\text{ for }\,e\in\pi^{+}\cup\pi^{++},
eπ++t(e)=eπ+t(e)and\displaystyle\sum_{e\,\in\,\pi^{++}}t(e)=\sum_{e^{\prime}\in\pi^{+}}t(e^{\prime})\quad\ \ \text{and}
t(e)>s0 for eB(ππ++)}.\displaystyle t(e)>s_{0}\,\text{ for }\,e\in B\setminus(\pi^{\prime}\cup\pi^{++})\,\Bigr{\}}.

By (6.9), unbounded weights, and the detour construction, there exists a constant D2D_{2} such that (ΓB,v,w)D2>0\mathbb{P}(\Gamma_{B,v,w})\geq D_{2}>0 for all triples (B,v,w)(B,v,w).

The steps follow those of the proof of Theorem 5.4 and the proof of Case (i) of Theorem 6.2. First sample ω\omega, and then define ω={t(e)}ed\omega^{*}=\{t^{*}(e)\}_{e\,\in\,{\mathcal{E}}_{d}} by setting t(e)=t(e)t^{*}(e)=t(e) for eBe\notin B and by resampling {t(e)}eB\{t^{*}(e)\}_{e\,\in\,B} independently. Let π(x)\pi(x) be a self-avoiding geodesic of minimal Euclidean length for T𝟎,x(ω)T_{\mathbf{0},x}(\omega). On the event {ωΛB,v,w,x}{ωΓB,v,w}\{\omega\in\Lambda_{B,v,w,x}\}\cap\{\omega^{*}\in\Gamma_{B,v,w}\} define the path π¯\overline{\pi} from 𝟎\mathbf{0} to xx by concatenating the segments π𝟎,v(x)\pi_{\mathbf{0},v}(x), π\pi^{\prime}, and πw,x(x)\pi_{w,x}(x).

Lemma 6.5.

When NN is fixed large enough, on the event {ωΛB,v,w,x}{ωΓB,v,w}\{\omega\in\Lambda_{B,v,w,x}\}\cap\{\omega^{*}\in\Gamma_{B,v,w}\} the path π¯\overline{\pi} is a self-avoiding geodesic of minimal Euclidean length for T𝟎,x(ω)T_{\mathbf{0},x}(\omega^{*}).

Proof.

As before, since box BB is black on the event ΛB,v,w,x\Lambda_{B,v,w,x},

T(πv,w(x))>(r0+δ0)(|wv|1N).\displaystyle T(\pi_{v,w}(x))>(r_{0}+\delta_{0})(|w-v|_{1}\vee N).

Then by ωΓB,v,w\omega^{*}\in\Gamma_{B,v,w},

T(π¯v,w)=T(π)\displaystyle T^{*}(\overline{\pi}_{v,w})=T^{*}(\pi^{\prime}) <ks1+(|wv|1+k+4+6)(r0+12δ0)\displaystyle<ks_{1}+(|w-v|_{1}+k+4\ell+6)(r_{0}+\tfrac{1}{2}\delta_{0})
(|wv|1N)(r0+12δ0)+k(s1+r0+12δ0)+(4+6)(r0+12δ0)\displaystyle\leq\bigl{(}|w-v|_{1}\vee N\bigr{)}(r_{0}+\tfrac{1}{2}\delta_{0})+k(s_{1}+r_{0}+\tfrac{1}{2}\delta_{0})+(4\ell+6)(r_{0}+\tfrac{1}{2}\delta_{0})
T(πv,w(x))12(|wv|1N)δ0+C\displaystyle\leq T(\pi_{v,w}(x))-\tfrac{1}{2}(|w-v|_{1}\vee N)\delta_{0}+C
<T(πv,w(x)).\displaystyle<T(\pi_{v,w}(x)).

Before the last inequality above, C=C(k,,s1,r0,δ0)C=C(k,\ell,s_{1},r_{0},\delta_{0}) is a constant determined by the quantities fixed thus far in the proof. The last inequality is then guaranteed by fixing NN large enough relative to these other constants. Outside BB weights ω\omega^{*} and ω\omega agree, and the segments π¯𝟎,v=π𝟎,v(x)\overline{\pi}_{\mathbf{0},v}=\pi_{\mathbf{0},v}(x) and π¯w,x=πw,x(x)\overline{\pi}_{w,x}=\pi_{w,x}(x) agree and lie outside BB. Hence the inequality above gives T(π¯)<T(π(x))T^{*}(\overline{\pi})<T(\pi(x)) and thereby, for any geodesic π(x)\pi^{*}(x) from 𝟎\mathbf{0} to xx in environment ω\omega^{*},

(6.11) T(π(x))T(π¯)<T(π(x)).T^{*}(\pi^{*}(x))\leq T^{*}(\overline{\pi})<T(\pi(x)).

As explained below (6.3), this implies that every ω\omega^{*} geodesic π(x)\pi^{*}(x) must enter BB.

If πB(x)ππ++\pi^{*}_{B}(x)\not\subset\pi^{\prime}\cup\pi^{++}, then π(x)\pi^{*}(x) must use an edge ee in BB with weight >s0>s_{0}. Then by property (5.4) of a black box BB, T(πB(x))s0<T(πB(x))T(\pi^{*}_{B}(x))\leq s_{0}<T^{*}(\pi^{*}_{B}(x)). Since tt and tt^{*} agree on BcB^{c}, we get

T(π(x))T(π(x))\displaystyle T(\pi(x))\leq T(\pi^{*}(x)) =T(πBc(x))+T(πB(x))\displaystyle=T(\pi^{*}_{B^{c}}(x))+T(\pi^{*}_{B}(x))
=T(πBc(x))+T(πB(x))<T(πBc(x))+T(πB(x))=T(π(x)),\displaystyle=T^{*}(\pi^{*}_{B^{c}}(x))+T(\pi^{*}_{B}(x))<T^{*}(\pi^{*}_{B^{c}}(x))+T^{*}(\pi^{*}_{B}(x))=T^{*}(\pi^{*}(x)),

contradicting (6.11). Consequently πB(x)ππ++\pi^{*}_{B}(x)\subset\pi^{\prime}\cup\pi^{++}. As a geodesic π(x)\pi^{*}(x) does not backtrack on itself. Hence it must traverse the route between vv to ww, either via π\pi^{\prime} or via π\pi^{\prime} with π+\pi^{+} replaced by π++\pi^{++}. By (6.10) T(π+)=T(π++)T^{*}(\pi^{+})=T^{*}(\pi^{++}) so there is no travel time distinction between the two routes between vv and ww.

Since ω\omega and ω\omega^{*} agree on BcB^{c}, π¯Bc\overline{\pi}_{B^{c}} is an optimal union of two paths that connect 𝟎\mathbf{0} to one of vv and ww, and xx to the other one of vv and ww. Thus π¯\overline{\pi} is a geodesic for T𝟎,x(ω)T_{\mathbf{0},x}(\omega^{*}).

The argument above showed that every geodesic of T𝟎,x(ω)T_{\mathbf{0},x}(\omega^{*}) goes from vv to ww utilizing edges in ππ++\pi^{\prime}\cup\pi^{++} and otherwise remains outside BB. If there were a geodesic πo\pi^{o} strictly shorter than π¯\overline{\pi}, πo\pi^{o} would have to use an alternative shorter geodesic path between 𝟎\mathbf{0} and vv or between ww and xx. This contradicts the choice of π(x)\pi(x) as the shortest geodesic. ∎

Define ΨB,x\Psi_{B,x} as in (6.4) above. By Lemma 6.5, ωΨB,x\omega^{*}\in\Psi_{B,x} holds on the event {ωΛB,v,w,x}{ωΓB,v,w}\{\omega\in\Lambda_{B,v,w,x}\}\cap\{\omega^{*}\in\Gamma_{B,v,w}\}. The proof of this case is completed exactly as was done in the previous case from equation (6.5) onwards.


Proof of Theorem 6.2 in Case (iii) of Assumption 6.1.

The weights are now assumed bounded. We work under assumption (6.1) until the last stage of the proof where we have to invoke the more stringent assumption of Case (iii) under which (6.1) is restricted to the case where all ri=rr^{\prime}_{i}=r and all sj=ss_{j}^{\prime}=s. Since the case of a zero atom has been taken care of, we can assume that these atoms {ri,sj}\{r^{\prime}_{i},s^{\prime}_{j}\} are strictly positive and that zero is not an atom. Since zero is not an atom, condition (5.11) holds.

As in the cases above, all that is needed for the conclusion is that the geodesic encounters (π+,π++)(\pi^{+},\pi^{++})-pairs whose passage times coincide. This proof follows closely the bounded weight case of Stage 2 of the proof of Theorem 5.4, which required condition (5.11). Lemma 5.6 can be enhanced to include the additional conclusion

(6.12) maxi,j{ri,sj}s0(q).\max_{i,j}\{r^{\prime}_{i},s^{\prime}_{j}\}\leq s_{0}(q).

The only change required in the proof of Lemma 5.6 is that induction begins with s0(0)=(r0+δ0)maxi,j{ri,sj}s_{0}(0)=(r_{0}+\delta_{0})\vee\max_{i,j}\{r^{\prime}_{i},s^{\prime}_{j}\}, after the case {t(e)=M0}>0\mathbb{P}\{t(e)=M_{0}\}>0 has been taken care of.

The construction of W1W_{1}, W1+W^{+}_{1}, W1W^{\prime}_{1}, W¯1\overline{W}_{\!1} and W2W_{2} in each black box BB goes exactly as before around (5.38). Let {πB,j+,πB,j++}1jj1(B)\{\pi^{+}_{B,\hskip 0.55ptj}\,,\pi^{++}_{B,\hskip 0.55ptj}\}_{1\leq j\leq j_{1}(B)} be the π+\pi^{+} and π++\pi^{++} boundary path segments of the detour rectangles {GB,j}1jj1(B)\{G_{B,\hskip 0.55ptj}\}_{1\leq j\leq j_{1}(B)} constructed in the box BB. In particular,

W1+=jπB,j+W1 and W1=(W1jπB,j++)(W1jGB,j)=W¯1.W_{1}^{+}=\bigcup_{j}\pi^{+}_{B,\hskip 0.55ptj}\subset W_{1}\quad\text{ and }\quad W^{\prime}_{1}=\biggl{(}W_{1}\cup\bigcup_{j}\pi^{++}_{B,\hskip 0.55ptj}\biggr{)}\subset\biggl{(}W_{1}\cup\bigcup_{j}G_{B,\hskip 0.55ptj}\biggr{)}=\overline{W}_{\!1}.

Define the event

(6.13) ΓB={ω:\displaystyle\Gamma_{B}=\Big{\{}\,\omega: r1δ<t(e)<r1+δeW1W1+,\displaystyle r_{1}-\delta<t(e)<r_{1}+\delta\quad\forall e\in W_{1}\,\setminus W^{+}_{1},
eπB,j++t(e)=eπB,j+t(e)j,\displaystyle\sum_{e\,\in\,\pi^{++}_{B,j}}t(e)=\sum_{e^{\prime}\,\in\,\pi^{+}_{B,j}}t(e^{\prime})\quad\forall j,
0<t(e)s0eW1,\displaystyle 0<t(e)\leq s_{0}\quad\forall e\in W_{1}^{\prime},
s0t(e)s1eW¯1W1,\displaystyle s_{0}\leq t(e)\leq s_{1}\quad\forall e\in\overline{W}_{\!1}\,\setminus W_{1}^{\prime},
and s1t(e)M0eBW¯1}.\displaystyle s_{1}\leq t(e)\leq M_{0}\quad\forall e\in B\setminus\overline{W}_{\!1}\,\Big{\}}.

The condition t(e)s0eW1t(e)\leq s_{0}\;\forall e\in W_{1}^{\prime} is implied by the conditions before it. It is stated explicitly merely for clarity. The condition t(e)>0eW1t(e)>0\;\forall e\in W_{1}^{\prime} can be imposed because (i) for eW1W1+e\in W_{1}\,\setminus W^{+}_{1} it follows from t(e)>r1δt(e)>r_{1}-\delta (recall from (5.27) that r1δ>0r_{1}-\delta>0), and (ii) for edges e1jj1(B)(πB,j+πB,j++)e\in\bigcup_{1\leq j\leq j_{1}(B)}(\pi^{+}_{B,\hskip 0.55ptj}\cup\pi^{++}_{B,\hskip 0.55ptj}) we can use the strictly positive atoms {ri,sj}\{r^{\prime}_{i},s^{\prime}_{j}\}. Again (ΓB)D2\mathbb{P}(\Gamma_{B})\geq D_{2} for a constant D2D_{2}.

As before, given an NN-box BB we work with two environments ω\omega and ω\omega^{*} that agree outside BB. Let π(x)\pi^{*}(x) be the T𝟎,x(ω)T_{\mathbf{0},x}(\omega^{*}) geodesic specified in Lemma 5.7. Starting from inequality (5.43), Stage 2 for bounded weights in the proof of Theorem 5.4 can be followed down to inequality (5.49), to get the existence of an excursion π¯\bar{\pi} in π(x)\pi^{*}(x) whose segment π¯1\bar{\pi}^{1} in W¯1\overline{W}_{\!1} satisfies (5.49). The previous Lemma 5.12 is then replaced by the next lemma.

Lemma 6.6.

Assume ωΛB,v,w,x\omega\in\Lambda_{B,v,w,x} and ωΓB\omega^{*}\in\Gamma_{B}. Then there exist three path segments π^,π+,π++\hat{\pi},\pi^{+},\pi^{++} in BB with the same endpoints and such that the following holds:

  1. (i)

    the pair (π+,π++)(\pi^{+},\pi^{++}) forms the boundaries of a detour rectangle,

  2. (ii)

    π^π(x)\hat{\pi}\subset\pi^{*}(x), and

  3. (iii)

    replacing π^\hat{\pi} in π(x)\pi^{*}(x) with either π+\pi^{+} or π++\pi^{++} produces two self-avoiding geodesics for T𝟎,x(ω)T_{\mathbf{0},x}(\omega^{*}).

Proof.

As in the proof of Lemma 5.12, π¯1\bar{\pi}^{1} has a segment π^=π¯a,b1\hat{\pi}=\bar{\pi}^{1}_{a,b} between the common endpoints aa and bb of the boundary paths π+\pi^{+} and π++\pi^{++} of some detour rectangle GG in BB. We show that π(x)\pi^{*}(x) can be redirected to take either π+\pi^{+} or π++\pi^{++}, by showing that (i) π^\hat{\pi} cannot be strictly better than π+\pi^{+} or π++\pi^{++} and (ii) replacing π^\hat{\pi} with π+\pi^{+} or π++\pi^{++} does not violate the requirement that a geodesic be self-avoiding.

Suppose T(π^)<T(π+)=T(π++)T^{*}(\hat{\pi})<T^{*}(\pi^{+})=T^{*}(\pi^{++}). Then there are points aa^{\prime} and bb^{\prime} on G\partial G such that π^\hat{\pi} visits a,a,b,ba,a^{\prime},b^{\prime},b in this order and the edges of π=π^a,b\pi^{\prime}=\hat{\pi}_{a^{\prime},b^{\prime}} lie in the interior GGG\setminus\partial G. Recall that on the event ΓB\Gamma_{B}, the weights on G\partial G are at most s0s_{0} while the weights in the interior GGG\setminus\partial G are at least s0s_{0}.

The points aa^{\prime} and bb^{\prime} cannot lie on the same or on adjacent sides of G\partial G since the 1\ell^{1}-path from aa^{\prime} to bb^{\prime} along G\partial G has no larger weight than π\pi^{\prime}.

Suppose aa^{\prime} and bb^{\prime} lie on opposite \ell-sides of GG. Then

T(π^)T(π)s0kT(π+)=T(π++).T^{*}(\hat{\pi})\geq T^{*}(\pi^{\prime})\geq s_{0}k\geq T^{*}(\pi^{+})=T^{*}(\pi^{++}).

So we can do at least as well by picking π+\pi^{+} or π++\pi^{++}.

The remaining option is that aa^{\prime} and bb^{\prime} lie on opposite kk-sides of GG. Let us suppose that aa^{\prime} is the first point at which π^\hat{\pi} leaves G\partial G and bb^{\prime} the first return to G\partial G.

For this argument we use the most restrictive assumption that there are two atoms r<sr<s such that (k+2)r=ks(k+2\ell)r=ks, with weights t(e)=st(e)=s on edges eπ+e\in\pi^{+} and t(e)=rt(e)=r on edges eπ++e\in\pi^{++}.

Case 1. Suppose aa^{\prime} lies on the kk-segment of π++\pi^{++} and bπ+b^{\prime}\in\pi^{+}. (See again Figure 5.6.) We can assume that aa is at the origin, a=a1𝐞1+𝐞2a^{\prime}=a_{1}^{\prime}\mathbf{e}_{1}+\ell\mathbf{e}_{2}, and b=b1𝐞1b^{\prime}=b_{1}^{\prime}\mathbf{e}_{1}. Then,

T(π^a,b)\displaystyle T^{*}(\hat{\pi}_{a,b^{\prime}}) =T(π^a,a)+T(π^a,b)\displaystyle=T^{*}(\hat{\pi}_{a,a^{\prime}})+T^{*}(\hat{\pi}_{a^{\prime},b^{\prime}})
|aa|1r+|ab|1s0\displaystyle\geq|a-a^{\prime}|_{1}r+|a^{\prime}-b^{\prime}|_{1}s_{0}
=(+a1)r+(+|b1a1|)s0.\displaystyle=(\ell+a_{1}^{\prime})r+(\ell+|b_{1}^{\prime}-a_{1}^{\prime}|)s_{0}.

From a1k1a_{1}^{\prime}\leq k-1 and the assumptions s0s>rs_{0}\geq s>r and ks=(k+2)rks=(k+2\ell)r we deduce:

(r+s)2r=k(sr)>a1(sr)(b1|b1a1|)sa1r\displaystyle\ell(r+s)\geq 2\ell r=k(s-r)>a_{1}^{\prime}(s-r)\geq\bigl{(}b_{1}^{\prime}-|b_{1}^{\prime}-a_{1}^{\prime}|\bigr{)}s-a_{1}^{\prime}r
(+a1)r+(+|b1a1|)s0>b1s\displaystyle\quad\implies\ \ (\ell+a_{1}^{\prime})r+(\ell+|b_{1}^{\prime}-a_{1}^{\prime}|)s_{0}>b_{1}^{\prime}s
T(π^a,b)>T(πa,b+).\displaystyle\quad\implies\ \ T^{*}(\hat{\pi}_{a,b^{\prime}})>T^{*}(\pi^{+}_{a,b^{\prime}}).

In other words, we can do better by taking π+\pi^{+} from aa to bb^{\prime}.

Case 2. Suppose aπ+a^{\prime}\in\pi^{+} and bb^{\prime} lies on the kk-segment of π++\pi^{++} so that a=a1𝐞1a^{\prime}=a_{1}^{\prime}\mathbf{e}_{1} and b=b1𝐞1+𝐞2b^{\prime}=b_{1}^{\prime}\mathbf{e}_{1}+\ell\mathbf{e}_{2}. Then,

T(π^a,b)\displaystyle T^{*}(\hat{\pi}_{a,b^{\prime}}) a1s+(+|b1a1|)s0>(+b1)r=T(πa,b++).\displaystyle\geq a^{\prime}_{1}s+(\ell+|b_{1}^{\prime}-a_{1}^{\prime}|)s_{0}>(\ell+b_{1}^{\prime})r=T^{*}(\pi^{++}_{a,b^{\prime}}).

This time it is better to take π++\pi^{++} from aa to bb^{\prime}.

We have shown that the passage time is not made worse by forcing π^\hat{\pi} to take π+\pi^{+} or π++\pi^{++}. Suppose doing so violates self-avoidance of the overall path from 𝟎\mathbf{0} to xx. Then we can cut out part of the path, and the removed piece includes at least one edge of either π+\pi^{+} or π++\pi^{++}. The assumption ωΓB\omega^{*}\in\Gamma_{B} implies that t(e)>0t^{*}(e)>0 for these edges. Consequently the original passage time could not have been optimal. ∎

The event ΨB,x\Psi_{B,x} earlier defined in (6.4) has to be reworded slightly for the present case. Let π(x)\pi(x) be the T𝟎,x(ω)T_{\mathbf{0},x}(\omega) geodesic chosen in Lemma 5.7.

(6.14) ΨB,x={\displaystyle\Psi_{B,x}=\{ inside BB \exists path segments π^\hat{\pi}, π+\pi^{+} and π++\pi^{++} that share both endpoints
and satisfy π^π(x)\hat{\pi}\subset\pi(x), both (π(x)π^)π+(\pi(x)\setminus\hat{\pi})\cup\pi^{+} and (π(x)π^)π++(\pi(x)\setminus\hat{\pi})\cup\pi^{++}
are self-avoiding paths from 𝟎\mathbf{0} to xx,
|π++||π+|+2, and T(π^)=T(π+)=T(π++) }.\displaystyle\text{$|\pi^{++}|\geq|\pi^{+}|+2$, and $T(\hat{\pi})=T(\pi^{+})=T(\pi^{++})$ }\}.

It is of course possible that π^\hat{\pi} agrees with either π+\pi^{+} or π++\pi^{++}. By Lemma 6.6, ωΨB,x\omega^{*}\in\Psi_{B,x} holds on the event {ωΛB,v,w,x}{ωΓB,v,w}\{\omega\in\Lambda_{B,v,w,x}\}\cap\{\omega^{*}\in\Gamma_{B,v,w}\}.

Now follow the proof of the previous case from equation (6.5) onwards. Again, since the boxes BB in the elements (B,v,w)j(x)(B,v,w)\in\mathcal{B}_{j(x)} are separated, we can define two self-avoiding paths π+(x)\pi^{+}(x) and π++(x)\pi^{++}(x) from 𝟎\mathbf{0} to xx by replacing each π^\hat{\pi} segment of π(x)\pi(x) with the π+\pi^{+} (respectively, π++\pi^{++}) segment in each box BB that appears among (B,v,w)j(x)(B,v,w)\in\mathcal{B}_{j(x)} and for which event ΨB,x\Psi_{B,x} happens. Then both π+(x)\pi^{+}(x) and π++(x)\pi^{++}(x) are self-avoiding geodesics for T𝟎,x(ω)T_{\mathbf{0},x}(\omega).

By the construction, the Euclidean lengths of these paths satisfy |π++(x)||π+(x)|+2Y|\pi^{++}(x)|\geq|\pi^{+}(x)|+2Y where YY is again the number of (B,v,w)j(x)(B,v,w)\in\mathcal{B}_{j(x)} for which ΨB,x\Psi_{B,x} occurs. Hence

L¯𝟎,x|π++(x)||π+(x)|+2YL¯𝟎,x+2Y.\displaystyle\overline{L}_{\mathbf{0},x}\geq|\pi^{++}(x)|\geq|\pi^{+}(x)|+2Y\geq\underline{L}_{\hskip 0.9pt\mathbf{0},x}+2Y.

This completes the proof of the third case and thereby the proof of Theorem 6.2. ∎


7. Proofs of the main theorems

This section proves the remaining claims of Section 2 by appeal to the preparatory work of Section 4 and the modification results of Sections 5 and 6.

7.1. Strict concavity, derivatives, and geodesic length

The next theorem gives part (ii) of Theorem 2.2 and thereby completes the proof of Theorem 2.2. Recall that r0=essinft(e)r_{0}=\mathop{\mathrm{ess\,inf}}t(e) and ε0>0\varepsilon_{0}>0 is the constant specified in Theorems 2.1 and A.1.

Theorem 7.1.

Assume r00r_{0}\geq 0, (2.6), and moment bound (2.7) with p=dp=d. Then there exist strictly positive constants D(a,h)D(a,h) such that the following holds for all ξd{𝟎}\xi\in\mathbb{R}^{d}\hskip 0.7pt\setminus\{\mathbf{0}\}: whenever ar0a\geq-r_{0} and r0ε0<ah<a-r_{0}-\varepsilon_{0}<a-h<a,

(7.1) μξ(ah)\displaystyle\mu_{\xi}(a-h) μξ(a)hμξ(a+)D(a,h)h|ξ|1.\displaystyle\leq\mu_{\xi}(a)-h\hskip 0.7pt\mu_{\xi}^{\prime}(a+)-D(a,h)\hskip 0.55pth\hskip 0.7pt|\xi|_{1}.

As a consequence, μξ(a0+)>μξ(a1)\mu_{\xi}^{\prime}(a_{0}+)>\mu_{\xi}^{\prime}(a_{1}-) whenever r0a0<a1<-r_{0}\leq a_{0}<a_{1}<\infty and μξ(b±)>μξ((r0)+)\mu_{\xi}^{\prime}(b\pm)>\mu_{\xi}^{\prime}((-r_{0})+) for all b(r0ε0,r0)b\in(-r_{0}-\varepsilon_{0},-r_{0}).

Note that the theorem does not rule out a linear segment of μξ\mu_{\xi} immediately to the left of r0-r_{0} which happens if μξ(b+)=μξ((r0))\mu_{\xi}^{\prime}(b+)=\mu_{\xi}^{\prime}((-r_{0})-) for some b(r0ε0,r0)b\in(-r_{0}-\varepsilon_{0},-r_{0}). But this does force μξ((r0))>μξ((r0)+)\mu_{\xi}^{\prime}((-r_{0})-)>\mu_{\xi}^{\prime}((-r_{0})+) and thereby a singularity at r0-r_{0}.

Proof.

We start by deriving the last statement of strict concavity from (7.1). Suppose that μξ(a0+)=μξ(a1)=τ\mu_{\xi}^{\prime}(a_{0}+)=\mu_{\xi}^{\prime}(a_{1}-)=\tau for some r0a0<a1<-r_{0}\leq a_{0}<a_{1}<\infty. Then by concavity μξ\mu_{\xi} must be affine on the open interval (a0,a1)(a_{0},a_{1}): μξ(a)=μξ(a0)+τ(aa0)\mu_{\xi}(a)=\mu_{\xi}(a_{0})+\tau(a-a_{0}) and μξ(a)=τ\mu_{\xi}^{\prime}(a)=\tau for a(a0,a1)a\in(a_{0},a_{1}). This violates (7.1). The second claim of the last statement follows similarly.

For this and a later proof, we check here the validity of the middle portion of (2.15). Let b>r0ε0b>-r_{0}-\varepsilon_{0}, ξd{𝟎}\xi\in\mathbb{R}^{d}\hskip 0.7pt\setminus\{\mathbf{0}\}, ωΩ0=\omega\in\Omega_{0}= the full measure event specified in Theorem A.1, and xn/nξx_{n}/n\to\xi. Take limits (2.8) in the extremes of (2.13), limits lim¯n1L¯𝟎,xn(b)(ω)\varliminf n^{-1}\underline{L}^{(b)}_{\hskip 0.9pt\mathbf{0},\hskip 0.9ptx_{n}}(\omega) and lim¯n1L¯𝟎,xn(b)(ω)\varlimsup n^{-1}\overline{L}^{\hskip 0.7pt(b)}_{\hskip 0.9pt\mathbf{0},\hskip 0.9ptx_{n}}(\omega) in the middle of (2.13), and then let δ,η0\delta,\eta\searrow 0. This gives

(7.2) μξ(b+)lim¯nL¯𝟎,xn(b)(ω)nlim¯nL¯𝟎,xn(b)(ω)nμξ(b).\mu_{\xi}^{\prime}(b+)\leq\varliminf_{n\to\infty}\frac{\underline{L}^{(b)}_{\hskip 0.9pt\mathbf{0},\hskip 0.9ptx_{n}}(\omega)}{n}\leq\varlimsup_{n\to\infty}\frac{\overline{L}^{\hskip 0.7pt(b)}_{\hskip 0.9pt\mathbf{0},\hskip 0.9ptx_{n}}(\omega)}{n}\leq\mu_{\xi}^{\prime}(b-).

To prove (7.1), consider first the case where a>r0a>-r_{0} or a=r0a=-r_{0} but {t(e)=r0}=0\mathbb{P}\{t(e)=r_{0}\}=0. The hypotheses of Theorem 5.4 are satisfied for the shifted weights ω(a)\omega^{(a)}. In particular, the extra assumption (5.11) of the bounded weights case that requires the existence of a positive support point r1r_{1} close enough to the lower bound is valid because either essinft(a)(e)>0\mathop{\mathrm{ess\,inf}}t^{(a)}(e)>0 or essinft(a)(e)=0\mathop{\mathrm{ess\,inf}}t^{(a)}(e)=0 but 0 is not an atom.

From Theorem 5.4 applied to the shifted weights ω(a)\omega^{(a)} we take the conclusion (5.13) which is valid in both cases of the theorem:

(7.3) 𝔼[T𝟎,x(ah)]𝔼[T𝟎,x(a)]h𝔼[L¯𝟎,x(a)]D(a,h)h|x|1.\mathbb{E}[T^{(a-h)}_{\mathbf{0},x}]\leq\mathbb{E}[T^{(a)}_{\mathbf{0},x}]-h\,\mathbb{E}[\,\underline{L}^{(a)}_{\hskip 0.9pt\mathbf{0},x}]-D(a,h)\hskip 0.55pth\hskip 0.55pt|x|_{1}.

The constant D(a,h)D(a,h) given by the theorem depends now also on aa.

In (7.3) take x=xnx=x_{n}, divide through by nn, and let nn\to\infty along a suitable subsequence. The expectations of normalized passage times converge by Theorem A.1. We obtain

(7.4) μξ(ah)μξ(a)hlim¯nn1𝔼[L¯𝟎,xn(a)]D(a,h)h|ξ|1.\mu_{\xi}(a-h)\leq\mu_{\xi}(a)-h\varlimsup_{n\to\infty}n^{-1}\mathbb{E}[\,\underline{L}^{(a)}_{\hskip 0.9pt\mathbf{0},\hskip 0.9ptx_{n}}\,]-D(a,h)\hskip 0.55pth\hskip 0.55pt|\xi|_{1}.

By Fatou’s lemma and (7.2),

(7.5) lim¯nn1𝔼[L¯𝟎,xn(a)]lim¯nn1𝔼[L¯𝟎,xn(a)]𝔼[lim¯nn1L¯𝟎,xn(a)]μξ(a+).\displaystyle\varlimsup_{n\to\infty}n^{-1}\mathbb{E}[\,\underline{L}^{(a)}_{\hskip 0.9pt\mathbf{0},\hskip 0.9ptx_{n}}\,]\geq\varliminf_{n\to\infty}n^{-1}\mathbb{E}[\,\underline{L}^{(a)}_{\hskip 0.9pt\mathbf{0},\hskip 0.9ptx_{n}}\,]\geq\mathbb{E}\bigl{[}\,\varliminf_{n\to\infty}n^{-1}\underline{L}^{(a)}_{\hskip 0.9pt\mathbf{0},\hskip 0.9ptx_{n}}\,\bigr{]}\geq\mu_{\xi}^{\prime}(a+).

This substituted into (7.4) gives (7.1).

Last we take up the case a=r0a=-r_{0} and 0<{t(e)=r0}<pc0<\mathbb{P}\{t(e)=r_{0}\}<p_{c}. The shifted weights ω(r0)\omega^{(-r_{0})} satisfy 0<{t(e)=0}<pc0<\mathbb{P}\{t(e)=0\}<p_{c}. This puts us in case (i) of Theorem 6.2. Its conclusion (6.2) implies the existence of a constant D>0D>0 such that

(L¯𝟎,xn(r0)L¯𝟎,xn(r0)D|xn|1 for infinitely many n)δ.\mathbb{P}\bigl{(}\,\overline{L}^{\hskip 0.7pt(-r_{0})}_{\hskip 0.9pt\mathbf{0},x_{n}}-\underline{L}^{(-r_{0})}_{\hskip 0.9pt\mathbf{0},x_{n}}\geq D|x_{n}|_{1}\text{ for infinitely many $n$}\hskip 0.9pt\bigr{)}\geq\delta.

Hence (7.2) implies μξ((r0))μξ((r0)+)D|ξ|1\mu_{\xi}^{\prime}((-r_{0})-)-\mu_{\xi}^{\prime}((-r_{0})+)\geq D|\xi|_{1}. Note that DD does not depend on the sequence {xn}\{x_{n}\} or ξ\xi. (7.1) comes from concavity:

μξ(r0h)μξ(r0)μξ((r0))hμξ(r0)μξ((r0)+)hDh|ξ|1.\mu_{\xi}(-r_{0}-h)\leq\mu_{\xi}(-r_{0})-\mu_{\xi}^{\prime}((-r_{0})-)h\leq\mu_{\xi}(-r_{0})-\mu_{\xi}^{\prime}((-r_{0})+)h-D\hskip 0.55pth\hskip 0.55pt|\xi|_{1}.\qed
Corollary 7.2.

Assume r00r_{0}\geq 0, (2.6), and moment bound (2.7) with p=dp=d. There exists a constant D>0D>0 such that λ¯(ξ)(1+D)|ξ|1\underline{\lambda}(\xi)\geq(1+D)|\xi|_{1} for all ξd{𝟎}\xi\in\mathbb{R}^{d}\setminus\{\mathbf{0}\}.

Proof.

Fix 0<ab<a0<a-b<a and let D=D(a,b)D=D(a,b) from (7.1). Then, for ξ𝟎\xi\neq\mathbf{0},

(7.6) λ¯(ξ)=μξ(0+)μξ(a)μξ(ab)bμξ(a+)+D(a,b)|ξ|1(1+D)|ξ|1.\underline{\lambda}(\xi)=\mu_{\xi}^{\prime}(0+)\geq\frac{\mu_{\xi}(a)-\mu_{\xi}(a-b)}{b}\geq\mu_{\xi}^{\prime}(a+)+D(a,b)|\xi|_{1}\geq(1+D)|\xi|_{1}.

The first equality is from the characterization of the superdifferential in (2.46) if r0>0r_{0}>0 and in (2.49) if r0=0r_{0}=0. The first inequality is concavity and the second one is (7.1). The last inequality is the easy bound from (2.14). ∎

Proof of Theorem 2.3.

We prove the first inequality of (2.15). For br0b\geq-r_{0} the characterizations of the superdifferentials in (2.46) and (2.49) give μξ(b+)=λ¯(b)(ξ)\mu_{\xi}^{\prime}(b+)=\underline{\lambda}^{(b)}(\xi). Corollary 7.2 gives constants D(b)>0D(b)>0 such that λ¯(b)(ξ)(1+D(b))|ξ|1\underline{\lambda}^{(b)}(\xi)\geq(1+D(b))|\xi|_{1}. By the monotonicity of the derivatives, D(b)=D(r0)D(b)=D(-r_{0}) works for b<r0b<-r_{0}. To produce a nonincreasing function, replace D(b)D(b) with infr0abD(a)\inf_{-r_{0}\leq a\leq b}D(a).

The three middle inequalities of (2.15) are in (7.2) above.

To prove the rightmost bound of (2.15), consider first b(r0ε0,r0]b\in(-r_{0}-\varepsilon_{0},-r_{0}]. Take a=(br0ε0)/2(r0ε0,b)a=(b-r_{0}-\varepsilon_{0})/2\in(-r_{0}-\varepsilon_{0},b). Let ωΩ0\omega\in\Omega_{0} and xn/nξx_{n}/n\to\xi. Concavity, (7.2), and (A.2) give

μξ(b)μξ(a+)lim¯nL¯𝟎,xn(a)(ω)nc(a+r0)0+ε0|ξ|1=2c(b+r0)0+ε0|ξ|1.\mu_{\xi}^{\prime}(b-)\leq\mu_{\xi}^{\prime}(a+)\leq\varliminf_{n\to\infty}\frac{\underline{L}^{(a)}_{\hskip 0.9pt\mathbf{0},\hskip 0.9ptx_{n}}(\omega)}{n}\leq\frac{c}{(a+r_{0})\wedge 0+\varepsilon_{0}}\,|\xi|_{1}=\frac{2c}{(b+r_{0})\wedge 0+\varepsilon_{0}}\,|\xi|_{1}.

The rightmost bound of (2.15) extends to all br0b\geq-r_{0} because μξ(b)\mu_{\xi}^{\prime}(b-) is nonincreasing in bb. ∎

Proof of Theorem 2.5.

Using Proposition 4.4(i), the continuity of the shape functions μ\mu and gog^{o} on int𝒰\operatorname{int}\mathcal{U}, and λ¯(ξ)(1+D)|ξ|1\underline{\lambda}(\xi)\geq(1+D)|\xi|_{1} from Corollary 7.2 above, choose constants η,δ>0\eta,\delta>0 small enough so that for any |ξ|1=1|\xi|_{1}=1,

(7.7) |μ(ξ)τgo(ξ/τ)|ητ1+δ.|\,\mu(\xi)-\tau g^{o}(\xi/\tau)\,|\leq\eta\ \ \implies\ \ \tau\geq 1+\delta.

From (4.8) or (A.2) pick finite deterministic κ\kappa and random KK such that

(7.8) L¯𝟎,xκ|x|1 for all |x|1K.\overline{L}_{\hskip 0.9pt\mathbf{0},x}\leq\kappa|x|_{1}\quad\text{ for all }\quad|x|_{1}\geq K.

Let α=δ/4\alpha=\delta/4. Increase κ\kappa if necessary so that κ>2+α\kappa>2+\alpha. Let 0<ε<η/(1+κ)0<\varepsilon<\eta/(1+\kappa). Increase KK if necessary so that (i) K4/δK\geq 4/\delta, (ii) KK works in (B.1) for α,ε\alpha,\varepsilon, and (iii) KK satisfies the FPP shape theorem ([2, p. 11], also (A.3))

(7.9) |T𝟎,xμ(x)|ε|x|1for |x|1K.|\,T_{\mathbf{0},x}-\mu(x)\,|\leq\varepsilon|x|_{1}\qquad\text{for }|x|_{1}\geq K.

Let |x|1K|x|_{1}\geq K and let π\pi be a geodesic for T𝟎,xT_{\mathbf{0},x}. Let k=|π|(1+α)|x|1k=|\pi|\vee\lceil{\hskip 0.55pt(1+\alpha)|x|_{1}\hskip 0.55pt}\rceil. Then

T𝟎,x=G𝟎,(|π|),xo=G𝟎,(k),xo.T_{\mathbf{0},x}=G^{o}_{\mathbf{0},(|\pi|),x}=G^{o}_{\mathbf{0},(k),x}.

A combination of (B.1) and (7.9), the homogeneity of μ\mu, and kκ|x|1k\leq\kappa|x|_{1} give

|μ(x)kgo(x/k)|ε|x|1+εkε(1+κ)|x|1\displaystyle|\,\mu(x)\,-\,k\hskip 0.7ptg^{o}(x/k)\,|\leq\varepsilon|x|_{1}+\varepsilon k\leq\varepsilon(1+\kappa)|x|_{1}
|μ(x|x|1)k|x|1go(x/|x|1k/|x|1)|ε(1+κ)<η.\displaystyle\implies\quad\biggl{\lvert}\,\mu\biggl{(}\frac{x}{|x|_{1}}\biggr{)}\,-\,\frac{k}{|x|_{1}}\hskip 0.7ptg^{o}\biggl{(}\frac{x/|x|_{1}}{k/|x|_{1}}\biggr{)}\,\biggr{\rvert}\,\leq\,\varepsilon(1+\kappa)<\eta.

Now (7.7) implies k(1+δ)|x|1k\geq(1+\delta)|x|_{1}. On the other hand, |x|1K>4/δ|x|_{1}\geq K>4/\delta implies that

k=|π|(1+α)|x|1|π|(1+δ/2)|x|1.k=|\pi|\vee\lceil{\hskip 0.55pt(1+\alpha)|x|_{1}\hskip 0.55pt}\rceil\leq|\pi|\vee(1+\delta/2)|x|_{1}.

Together these force |π|(1+δ)|x|1|\pi|\geq(1+\delta)|x|_{1}. ∎

Proof of Theorem 2.11.

Part (i). The statements about λ¯(ξ)\underline{\lambda}(\xi) come from Lemma 4.2. The statements about λ¯(ξ)\overline{\lambda}(\xi) come from the definition (4.16) and Proposition 4.4(ii). The semicontinuity claims are in Lemma 4.5. The finite-infinite dichotomy of λ¯(ξ)\overline{\lambda}(\xi) is in (4.17) and (4.18).

Part (ii). To derive (2.35), combine Corollary 7.2, (4.17), (4.18), (7.2), and the characterizations of the derivatives μξ(0±)\mu_{\xi}^{\prime}(0\pm) from (2.46) when r0>0r_{0}>0 and from (2.49) when r0=0r_{0}=0. ∎

Proof of Theorem 2.16.

Part (i) was proved in Lemma 4.3. Part (ii) comes from Proposition 4.4.

Part (iii). Begin by noting that differentiability of tg(tξ)t\mapsto g^{\diamond}(t\xi) is equivalent to differentiability of ττg(ξ/τ)\tau\mapsto\tau g^{\diamond}(\xi/\tau) and on an open interval a differentiable convex function is continuously differentiable.

Since λ¯(ξ)>|ξ|1\underline{\lambda}(\xi)>|\xi|_{1} and by the limit (2.48), the union of the superdifferentials on the right-hand sides of (2.46) and (2.47) is equal to the interval (|ξ|1,)(|\xi|_{1},\infty). General convex analysis gives the equivalence

bτ[τg(ξ/τ)]τμξ(b).-b\in\partial_{\tau}[\tau\hskip 0.55ptg(\xi/\tau)]\ \iff\ \tau\in\partial\mu_{\xi}(b).

By the strict concavity of μξ\mu_{\xi}, a given τ\tau lies in μξ(b)\partial\mu_{\xi}(b) for a unique bb, and hence the subdifferential τ[τg(ξ/τ)]\partial_{\tau}[\tau\hskip 0.55ptg(\xi/\tau)] consists of a unique value b(,r0]-b\in(-\infty,r_{0}]. This implies that ττg(ξ/τ)\tau\mapsto\tau\hskip 0.55ptg(\xi/\tau) is differentiable at τ(|ξ|1,)\tau\in(|\xi|_{1},\infty).

Continuous differentiability of ττgo(ξ/τ)\tau\mapsto\tau g^{o}(\xi/\tau) for τ>|ξ|1\tau>|\xi|_{1} now follows from Proposition 4.4. Namely, τgo(ξ/τ)=τg(ξ/τ)\tau g^{o}(\xi/\tau)=\tau g(\xi/\tau) for τ[|ξ|1,λ¯(ξ)]\tau\in[\,|\xi|_{1},\underline{\lambda}(\xi)\hskip 0.7pt], which we now know to be a nondegenerate interval, and their common left τ\tau-derivative vanishes at the minimum τ=λ¯(ξ)\tau=\underline{\lambda}(\xi). On [λ¯(ξ),)[\hskip 0.7pt\underline{\lambda}(\xi),\infty), τgo(ξ/τ)=μ(ξ)\tau g^{o}(\xi/\tau)=\mu(\xi) is constant and hence connects in a C1C^{1} fashion to the part on [|ξ|1,λ¯(ξ)][\hskip 0.9pt|\xi|_{1},\underline{\lambda}(\xi)\hskip 0.7pt].

If g(ξ/|ξ|1)=g^{\diamond}(\xi/|\xi|_{1})=\infty then necessarily limt|ξ|11(g)(tξ)=+\lim_{t\nearrow|\xi|_{1}^{-1}}(g^{\diamond})^{\prime}(t\xi)=+\infty.

The remaining claims follow if we assume g(ξ/|ξ|1)<g^{\diamond}(\xi/|\xi|_{1})<\infty and show that

(7.10) limt|ξ|11g(|ξ|11ξ)g(tξ)|ξ|11t=+.\lim_{t\nearrow|\xi|_{1}^{-1}}\frac{g^{\diamond}(|\xi|_{1}^{-1}\xi)-g^{\diamond}(t\xi)}{|\xi|_{1}^{-1}-t}=+\infty.

It suffices to treat gg since go=gg^{o}=g close enough to the boundary of 𝒰\mathcal{U} by part (ii).

Take α=1/t>|ξ|1\alpha=1/t>|\xi|_{1} and rewrite the ratio above as

|ξ|1g(|ξ|11ξ)+|ξ|1|ξ|1g(ξ/|ξ|1)αg(ξ/α)α|ξ|1.\displaystyle|\xi|_{1}g(|\xi|_{1}^{-1}\xi)+|\xi|_{1}\,\frac{|\xi|_{1}g(\xi/|\xi|_{1})-\alpha\hskip 0.55ptg(\xi/\alpha)}{\alpha-|\xi|_{1}}.

Thus by the duality in Theorem 2.17, (7.10) is equivalent to

(7.11) limα|ξ|1μ¯ξ(α)μ¯ξ(|ξ|1)α|ξ|1=.\lim_{\alpha\searrow|\xi|_{1}}\frac{\overline{\mu}_{\xi}^{*}(\alpha)-\overline{\mu}_{\xi}^{*}(|\xi|_{1})}{\alpha-|\xi|_{1}}=\infty.

By concavity, the ratio in (7.11) above is a nonincreasing function of α>|ξ|1\alpha>|\xi|_{1}. Hence if (7.11) fails, there exists b0<b_{0}<\infty such that, α>|ξ|1\forall\alpha>|\xi|_{1} and bb0\forall b\geq b_{0},

|ξ|1bμ¯ξ(|ξ|1)αbμ¯ξ(α).|\xi|_{1}b-\overline{\mu}_{\xi}^{*}(|\xi|_{1})\leq\alpha b-\overline{\mu}_{\xi}^{*}(\alpha).

It then follows from the duality ((2.41) or (2.44)) that

μ¯ξ(b)=|ξ|1bμ¯ξ(|ξ|1)for bb0.\overline{\mu}_{\xi}(b)=|\xi|_{1}b-\overline{\mu}_{\xi}^{*}(|\xi|_{1})\qquad\text{for }b\geq b_{0}.

This contradicts the strict concavity of μ¯ξ\overline{\mu}_{\xi}. (7.10) has been verified. ∎

7.2. Nondifferentiability

Proof of Theorem 2.6.

Bound (2.21) is contained in Theorem 6.2. (2.21) implies that, along any subsequence {ni}\{n_{i}\},

(L¯𝟎,xniL¯𝟎,xniD|xni|1 for infinitely many i)δ.\mathbb{P}\bigl{(}\,\overline{L}_{\hskip 0.9pt\mathbf{0},x_{n_{i}}}-\underline{L}_{\hskip 0.9pt\mathbf{0},x_{n_{i}}}\geq D|x_{n_{i}}|_{1}\text{ for infinitely many $i$}\hskip 0.9pt\bigr{)}\geq\delta.

Now (7.2) implies μξ(0)μξ(0+)D|ξ|1\mu_{\xi}^{\prime}(0-)-\mu_{\xi}^{\prime}(0+)\geq D|\xi|_{1}. ∎

Proof of Theorem 2.7.

Let r<sr<s be two atoms of t(e)t(e) in [r0,)[r_{0},\infty). Fix an arbitrary \ell\in\mathbb{N} and then pick kk\in\mathbb{N} so that

(k1)(sr)2rr0<k(sr)2.\frac{(k-1)(s-r)}{2\ell}\leq r-r_{0}<\frac{k(s-r)}{2\ell}.

For m+m\in\mathbb{Z}_{+} let

bm=(k+m)(sr)2r(r0,).b_{m}=\frac{(k+m)(s-r)}{2\ell}-r\;\in\;(-r_{0},\infty).

Then bm+rb_{m}+r and bm+sb_{m}+s are atoms of t(bm)(e)t^{(b_{m})}(e) such that

(k+m)(s+bm)=(k+m+2)(r+bm)for all m+.(k+m)(s+b_{m})=(k+m+2\ell)(r+b_{m})\qquad\text{for all $m\in\mathbb{Z}_{+}$.}

The other hypotheses of Theorem 2.6 are inherited by ω(bm)\omega^{(b_{m})} and so the conclusions of Theorem 2.6 hold for all ω(bm)\omega^{(b_{m})}. In particular, since μξ(bm)(a)=μξ(a+bm)\mu_{\xi}^{(b_{m})}(a)=\mu_{\xi}(a+b_{m}), μξ(bm)\mu_{\xi}^{(b_{m})} has a corner at 0 if and only if μξ\mu_{\xi} has a corner at bmb_{m}.

No point of [r0,)[-r_{0},\infty) is farther than sr2\frac{s-r}{2\ell} from the nearest bmb_{m}. We get the dense set BB by combining the collections {bm}\{b_{m}\} for all \ell\in\mathbb{N}. ∎

Appendix A First-passage percolation with slightly negative weights

This appendix extends the shape theorem of standard FPP to real-valued weights {t(e)}\{t(e)\} under certain hypotheses. The setting is the same as in Section 2.1. As before, {ti}\{t_{i}\} denotes i.i.d. copies of the edge weight t(e)t(e). Assumption (2.7) is reformulated for positive parts as

(A.1) 𝔼[(min{t1+,,t2d+})p]<.\mathbb{E}[\,(\min\{t^{+}_{1},\dotsc,t^{+}_{2d}\})^{p}\,]<\infty.

Passage times Tx,yT_{x,y} are defined as in (2.2) and now the restriction to self-avoiding paths is essential.

Theorem A.1.

Assume r0=essinft(e)0r_{0}=\mathop{\mathrm{ess\,inf}}t(e)\geq 0, (2.6), and (A.1) (equivalently, (2.7)) with p=dp=d. Then there exist

  1. (a)

    a constant ε0>0\varepsilon_{0}>0 determined by the distribution of the shifted weights ω(r0)\omega^{(-r_{0})},

  2. (b)

    for each real b>r0ε0b>-r_{0}-\varepsilon_{0}, a positively homogeneous continuous convex function μ(b):d+\mu^{(b)}:\mathbb{R}^{d}\to\mathbb{R}_{+}, and

  3. (c)

    an event Ω0\Omega_{0} of full probability,

such that the properties listed below in points (i)–(iii) are satisfied.

  1. (i)

    For each ωΩ0\omega\in\Omega_{0} and b>r0ε0b>-r_{0}-\varepsilon_{0} the following pointwise statements hold. For each xdx\in\mathbb{Z}^{d}, T𝟎,x(b)T^{(b)}_{\mathbf{0},x} is finite and has a geodesic, that is, a self-avoiding path π\pi from 𝟎\mathbf{0} to xx such that T𝟎,x(b)=T(b)(π)T^{(b)}_{\mathbf{0},x}=T^{(b)}(\pi). There exist a deterministic finite constant cc and an ω\omega-dependent finite constant K=K(ω)K=K(\omega) such that

    (A.2) L¯𝟎,x(b)cε0+(r0+b)0|x|1whenever |x|1K.\overline{L}^{(b)}_{\mathbf{0},x}\leq\frac{c}{\varepsilon_{0}+(r_{0}+b)\wedge 0}\,|x|_{1}\qquad\text{whenever }\ \ |x|_{1}\geq K.

    The shape theorem holds, locally uniformly in the shift bb: for any a0<a1a_{0}<a_{1} in (r0ε0,)(-r_{0}-\varepsilon_{0},\infty),

    (A.3) limnsup|x|1nsupb[a0,a1]|T𝟎,x(b)μ(b)(x)||x|1=0.\displaystyle\lim_{n\to\infty}\;\sup_{|x|_{1}\geq n}\,\sup_{b\hskip 0.9pt\in\hskip 0.9pt[a_{0},a_{1}]}\,\frac{|\hskip 0.9ptT^{(b)}_{\mathbf{0},x}-\mu^{(b)}(x)\hskip 0.9pt|}{|x|_{1}}=0.
  2. (ii)

    For each b>r0ε0b>-r_{0}-\varepsilon_{0} the following statements hold. T𝟎,x(b)L1()T^{(b)}_{\mathbf{0},x}\in L^{1}(\mathbb{P}) for all xdx\in\mathbb{Z}^{d}. For any sequence xndx_{n}\in\mathbb{Z}^{d} with xn/nξdx_{n}/n\to\xi\in\mathbb{R}^{d}, the convergence n1T𝟎,xn(b)μ(b)(ξ)n^{-1}T^{(b)}_{\mathbf{0},x_{n}}\to\mu^{(b)}(\xi) holds almost surely and in L1()L^{1}(\mathbb{P}).

  3. (iii)

    The shape function satisfies these Lipschitz bounds for shifts b2>b1>r0ε0b_{2}>b_{1}>-r_{0}-\varepsilon_{0} and all ξd\xi\in\mathbb{R}^{d}:

    (A.4) μ(b1)(ξ)μ(b2)(ξ)μ(b1)(ξ)+c|ξ|1ε0+(r0+b1)0(b2b1).\mu^{(b_{1})}(\xi)\leq\mu^{(b_{2})}(\xi)\leq\mu^{(b_{1})}(\xi)+\frac{c\hskip 0.55pt|\xi|_{1}}{\varepsilon_{0}+(r_{0}+b_{1})\wedge 0}\,(b_{2}-b_{1}).

    For b>r0ε0b>-r_{0}-\varepsilon_{0}, μ(b)(𝟎)=0\mu^{(b)}(\mathbf{0})=0 and μ(b)(ξ)>0\mu^{(b)}(\xi)>0 for all ξ0\xi\neq 0.

We prove this theorem at the end of the section after proving a more general shape result in Theorem A.4 below.

Lemma A.2.

Let \mathbb{P} be a probability measure invariant under a group {θx}xd\{\theta_{x}\}_{x\hskip 0.7pt\in\hskip 0.7pt\mathbb{Z}^{d}} of measurable bijections. Let AA be a nonnegative random variable such that 𝔼[Ad]<\mathbb{E}[A^{d}]<\infty. Then

(A.5) limmm1max|x|1mAθx=0 with probability one.\displaystyle\lim_{m\to\infty}m^{-1}\max_{|x|_{1}\leq m}A\circ\theta_{x}=0\quad\text{ with probability one.}
Proof.

The conclusion is equivalent to |x|11Aθx0|x|_{1}^{-1}A\circ\theta_{x}\to 0 as |x|1|x|_{1}\to\infty. Apply Borel-Cantelli with the estimate below for ε>0\varepsilon>0:

x{Aθxε|x|1}\displaystyle\sum_{x}\mathbb{P}\{A\circ\theta_{x}\geq\varepsilon|x|_{1}\} =k=0|x|1=k{Aθxkε}1+C(d)k=1kd1{Akε}\displaystyle=\sum_{k=0}^{\infty}\sum_{|x|_{1}=k}\mathbb{P}\{A\circ\theta_{x}\geq k\varepsilon\}\leq 1+C(d)\sum_{k=1}^{\infty}k^{d-1}\mathbb{P}\{A\geq k\varepsilon\}
1+C(d,ε)𝔼[Ad]<.\displaystyle\leq 1+C(d,\varepsilon)\,\mathbb{E}[A^{d}]<\infty.\qed

Because the inequalities in the proof can be reversed with different constants, an i.i.d. example shows that p<dp<d moments does not suffice for the conclusion.

Let x=(x)0x^{-}=(-x)\vee 0 denote the negative part of a real number. Following [18], define the random variable

(A.6) A=2supxdT𝟎,x\displaystyle A=2\sup_{x\hskip 0.7pt\in\hskip 0.7pt\mathbb{Z}^{d}}T_{\mathbf{0},x}^{\hskip 0.7pt-}

We first prove a moment bound for the shifts of AA that was used in the concavity result of Section 5.2.

Lemma A.3.

Assume r00r_{0}\geq 0 and the subcriticality assumption (2.6). Let δ>0\delta>0 be the constant in the bound (4.10) for the shifted weights ω(r0)\omega^{(-r_{0})}. Then there exists s>0s>0 such that 𝔼[esA(b)]<\mathbb{E}[e^{sA^{(b)}}]<\infty for all shifts A(b)=2supxd(T𝟎,x(b))A^{(b)}=2\sup_{x\hskip 0.7pt\in\hskip 0.7pt\mathbb{Z}^{d}}\bigl{(}T_{\mathbf{0},x}^{(b)}\bigr{)}^{-} such that br0δb\geq-r_{0}-\delta.

Proof.

By monotonicity it is enough to consider the case b=r0δb=-r_{0}-\delta. The proof is the same as that of the corollary of Theorem 3 in [13]. A(r0δ)a>0A^{(-r_{0}-\delta)}\geq a>0 implies the existence of a self-avoiding path γ\gamma from 𝟎\mathbf{0} such that T(r0δ)(γ)<a/4T^{(-r_{0}-\delta)}(\gamma)<-a/4. Turn this into

δ|γ|T(r0)(γ)δ|γ|=T(r0δ)(γ)<a/4<0.-\delta|\gamma|\leq T^{(-r_{0})}(\gamma)-\delta|\gamma|=T^{(-r_{0}-\delta)}(\gamma)<-a/4<0.

Then |γ|>a/(4δ)|\gamma|>a/(4\delta) and (4.10) gives the bound

{A(r0δ)a}\displaystyle\mathbb{P}\{A^{(-r_{0}-\delta)}\geq a\} { self-avoiding path γ from the origin\displaystyle\leq\mathbb{P}\bigl{\{}\text{$\exists$ self-avoiding path $\gamma$ from the origin}
such that |γ|a/(4δ) and T(r0)(γ)δ|γ|}Cec1a/(4δ).\displaystyle\qquad\qquad\text{such that $|\gamma|\geq{a}/{(4\delta)}$ and $T^{(-r_{0})}(\gamma)\leq\delta|\gamma|$}\bigr{\}}\leq Ce^{-c_{1}a/(4\delta)}.\qed

The next item is a shape theorem whose hypotheses are stated in terms of the random variable AA of (A.6).

Theorem A.4.

Let ω=(t(e):ed)\omega=(t(e):e\in{\mathcal{E}}_{d}) be i.i.d. real-valued weights.

(i) Assume (A.1) with p=1p=1 and that the random variable from (A.6) satisfies AL1A\in L^{1}. Then Tx,yT_{x,y} is a finite integrable random variable for all x,ydx,y\in\mathbb{Z}^{d}. There exists a non-random positively homogeneous continuous convex function μ:d+\mu:\mathbb{R}^{d}\to\mathbb{R}_{+} such that for any sequence {xn}d\{x_{n}\}\subset\mathbb{Z}^{d} with xn/nξdx_{n}/n\to\xi\in\mathbb{R}^{d},

(A.7) limn𝔼[|n1T𝟎,xnμ(ξ)|]=0.\displaystyle\lim_{n\to\infty}\mathbb{E}[\hskip 0.9pt|n^{-1}\hskip 0.7ptT_{\mathbf{0},x_{n}}-\mu(\xi)|\hskip 0.9pt]=0.

(ii) Assume furthermore (A.1) with p=dp=d and 𝔼[Ad]<\mathbb{E}[A^{d}]<\infty. Then the following hold with probability one:

(A.8) limnsup|x|1n|T𝟎,xμ(x)||x|1=0\displaystyle\lim_{n\to\infty}\sup_{|x|_{1}\geq n}\frac{|T_{\mathbf{0},x}-\mu(x)|}{|x|_{1}}=0

and for all ξd\xi\in\mathbb{R}^{d} and any sequence xndx_{n}\in\mathbb{Z}^{d} such that xn/nξx_{n}/n\to\xi

(A.9) μ(ξ)=limnT𝟎,xnn.\displaystyle\mu(\xi)=\lim_{n\to\infty}\frac{T_{\mathbf{0},x_{n}}}{n}\,.
Proof.

Let Ax=AθxA_{x}=A\circ\theta_{x}. Consider two paths πx,yΠx,y sa\pi_{x,y}\in\Pi^{\text{\rm{\,sa}}}_{x,y} and πy,zΠy,z sa\pi_{y,z}\in\Pi^{\text{\rm{\,sa}}}_{y,z}. Their concatenation may fail to be self-avoiding. Choose a point uu belonging to both paths such that erasing the portion of πx,y\pi_{x,y} from uu to yy (denoted by πu,y\pi^{\prime}_{u,y}) and erasing the portion of πy,z\pi_{y,z} from yy to uu (denoted by πy,u′′\pi^{\prime\prime}_{y,u}) leaves a self-avoiding path πx,z\pi_{x,z} from xx to zz. (If the concatenation was self-avoiding to begin with, then u=yu=y.) Note that πu,y\pi^{\prime}_{u,y} and πy,u′′\pi^{\prime\prime}_{y,u} are self-avoiding paths. This implies that

T(πx,y)+T(πy,z)\displaystyle T(\pi_{x,y})+T(\pi_{y,z}) =T(πx,z)+T(πu,y)+T(πy,u′′)Tx,z+Tu,y+Ty,u\displaystyle=T(\pi_{x,z})+T(\pi^{\prime}_{u,y})+T(\pi^{\prime\prime}_{y,u})\geq T_{x,z}+T_{u,y}+T_{y,u}
Tx,zTu,yTy,uTx,zAy.\displaystyle\geq T_{x,z}-T^{-}_{u,y}-T^{-}_{y,u}\geq T_{x,z}-A_{y}.

Taking infimum over πx,y\pi_{x,y} and πy,z\pi_{y,z} gives Tx,y+Ty,zTx,zAy.T_{x,y}+T_{y,z}\geq T_{x,z}-A_{y}. Rearranging, we get

(A.10) 0Tx,z+AzTx,y+Ay+Ty,z+Az.\displaystyle 0\leq T_{x,z}+A_{z}\leq T_{x,y}+A_{y}+T_{y,z}+A_{z}.

To apply the subadditive ergodic theorem, we derive a moment bound.

Let ω+=(t(e)+:ed)\omega^{+}=(t(e)^{+}:e\in{\mathcal{E}}_{d}). Take any 1\ell^{1}-path x0:kx_{0\hskip 0.6pt{\boldsymbol{:}}\hskip 0.3ptk} from 𝟎\mathbf{0} to xx (where k=|x|1k=|x|_{1}) and use the subadditivity of the passage times in weights ω+\omega^{+} to write

T𝟎,x(ω+)i=0k1Txi,xi+1(ω+).T_{\mathbf{0},x}(\omega^{+})\leq\sum_{i=0}^{k-1}T_{x_{i},x_{i+1}}(\omega^{+}).

Since 𝔼[T𝟎,±𝐞i(ω+)]\mathbb{E}[T_{\mathbf{0},\pm\mathbf{e}_{i}}(\omega^{+})] are all identical,

(A.11) 𝔼[T𝟎,x(ω)]𝔼[T𝟎,x(ω+)]𝔼[T𝟎,𝐞1(ω+)]|x|1.\displaystyle\mathbb{E}[T_{\mathbf{0},x}(\omega)]\leq\mathbb{E}[T_{\mathbf{0},x}(\omega^{+})]\leq\mathbb{E}[T_{\mathbf{0},\mathbf{e}_{1}}(\omega^{+})]\,|x|_{1}.

Assumption (A.1) with p=1p=1 implies that 𝔼[T𝟎,𝐞1(ω+)]<\mathbb{E}[T_{\mathbf{0},\mathbf{e}_{1}}(\omega^{+})]<\infty (Lemma 2.3 in [2]). By the assumption AL1A\in L^{1},

𝔼[T𝟎,x+Ax]C|x|1+𝔼[A]<.\mathbb{E}[T_{\mathbf{0},x}+A_{x}]\leq C|x|_{1}+\mathbb{E}[A]<\infty.

Standard subadditivity arguments give the existence of a positively homogeneous convex function μ¯:d+\overline{\mu}:\mathbb{Q}^{d}\to\mathbb{R}_{+} such that for all ζd\zeta\in\mathbb{Q}^{d} and \ell\in\mathbb{N} with ζd\ell\zeta\in\mathbb{Z}^{d}, almost surely and in L1L^{1},

(A.12) μ¯(ζ)=limnT𝟎,nζ+Anζn=limnT𝟎,nζn[0,C|ζ|1],\displaystyle\overline{\mu}(\zeta)=\lim_{n\to\infty}\frac{T_{\mathbf{0},n\ell\zeta}+A_{n\ell\zeta}}{n\ell}=\lim_{n\to\infty}\frac{T_{\mathbf{0},n\ell\zeta}}{n\ell}\in[0,C|\zeta|_{1}],

and μ¯(ζ)\overline{\mu}(\zeta) does not depend on the choice of \ell. The assumption AL1A\in L^{1} allows us to drop the term AnζA_{n\ell\zeta} from above. The first inequality of (A.10) gives μ¯(ζ)0\overline{\mu}(\zeta)\geq 0.

Fix x,ydx,y\in\mathbb{Z}^{d}. Use subadditivity (A.10) to write

T𝟎,xT𝟎,y\displaystyle T_{\mathbf{0},x}-T_{\mathbf{0},y} =T𝟎,x+Ax+Tx,y+AyT𝟎,yAyTx,yAx\displaystyle=T_{\mathbf{0},x}+A_{x}+T_{x,y}+A_{y}-T_{\mathbf{0},y}-A_{y}-T_{x,y}-A_{x}
Tx,yAxTx,y(ω+)Ax.\displaystyle\geq-T_{x,y}-A_{x}\geq-T_{x,y}(\omega^{+})-A_{x}.

Switching xx and yy gives a complementary bound and so

(A.13) |T𝟎,xT𝟎,y|Tx,y(ω+)+Ax+Ay.\displaystyle|T_{\mathbf{0},x}-T_{\mathbf{0},y}|\leq T_{x,y}(\omega^{+})+A_{x}+A_{y}.

By (A.11)

(A.14) 𝔼[|T𝟎,xT𝟎,y|]C|xy|1+2𝔼[A].\displaystyle\mathbb{E}[\hskip 0.9pt|T_{\mathbf{0},x}-T_{\mathbf{0},y}|\hskip 0.9pt]\leq C|x-y|_{1}+2\mathbb{E}[A].

Now take ζ,ηd\zeta,\eta\in\mathbb{Q}^{d} and \ell\in\mathbb{N} such that ζ\ell\zeta and η\ell\eta are both in d\mathbb{Z}^{d} and apply the above to get

|𝔼[T𝟎,nζ]𝔼[T𝟎,nη]|Cn|ζη|1+2𝔼[A].|\mathbb{E}[T_{\mathbf{0},n\ell\zeta}]-\mathbb{E}[T_{\mathbf{0},n\ell\eta}]|\leq Cn\ell|\zeta-\eta|_{1}+2\mathbb{E}[A].

Divide by nn\ell and take nn to \infty to get

(A.15) |μ¯(ζ)μ¯(η)|C|ζη|1.\displaystyle|\overline{\mu}(\zeta)-\overline{\mu}(\eta)|\leq C|\zeta-\eta|_{1}.

As a Lipschitz function μ¯\overline{\mu} extends uniquely to a continuous positively homogenous convex function μ:d\mu:\mathbb{R}^{d}\to\mathbb{R}.

Fix ξd{𝟎}\xi\in\mathbb{R}^{d}\setminus\{\mathbf{0}\} and a sequence xnx_{n} in d\mathbb{Z}^{d} such that xn/nξx_{n}/n\to\xi. Fix ζd\zeta\in\mathbb{Q}^{d}. Take \ell\in\mathbb{N} such that ζd\ell\zeta\in\mathbb{Z}^{d}. For nn\in\mathbb{N} let mn=n/m_{n}=\lfloor{n/\ell}\rfloor. By (A.14),

𝔼[|n1T𝟎,xnμ(ξ)|]\displaystyle\mathbb{E}[|n^{-1}T_{\mathbf{0},x_{n}}-\mu(\xi)|] n1𝔼[|T𝟎,xnT𝟎,mnζ|]+𝔼[|n1T𝟎,mnζμ¯(ζ)|]+|μ¯(ζ)μ(ξ)|\displaystyle\leq n^{-1}\mathbb{E}[\hskip 0.9pt|T_{\mathbf{0},x_{n}}-T_{\mathbf{0},m_{n}\ell\zeta}|\hskip 0.9pt]+\mathbb{E}[|n^{-1}T_{\mathbf{0},m_{n}\ell\zeta}-\overline{\mu}(\zeta)|]+|\overline{\mu}(\zeta)-\mu(\xi)|
n1C|xnmnζ|1+2n1𝔼[A]+𝔼[|n1T𝟎,mnζμ¯(ζ)|]+|μ¯(ζ)μ(ξ)|.\displaystyle\leq n^{-1}C|x_{n}-m_{n}\ell\zeta|_{1}+2n^{-1}\mathbb{E}[A]+\mathbb{E}[|n^{-1}T_{\mathbf{0},m_{n}\ell\zeta}-\overline{\mu}(\zeta)|]+|\overline{\mu}(\zeta)-\mu(\xi)|.

Take nn\to\infty to get

lim¯nn1𝔼[|T𝟎,xnμ(ξ)|]C|ξζ|1+|μ¯(ζ)μ(ξ)|.\varlimsup_{n\to\infty}n^{-1}\mathbb{E}[|T_{\mathbf{0},x_{n}}-\mu(\xi)|]\leq C|\xi-\zeta|_{1}+|\overline{\mu}(\zeta)-\mu(\xi)|.

Let ζξ\zeta\to\xi to get (A.7). This completes the proof of part (i).

Now strengthen the assumptions to 𝔼[Ad]<\mathbb{E}[A^{d}]<\infty and (A.1) with p=dp=d. For (A.8) we follow the proof of the Cox-Durrett shape theorem presented in [2, Section 2.3].

Let Ω0\Omega_{0} be the full probability event on which (A.12) holds for all ζd\zeta\in\mathbb{Q}^{d}. By Lemma 2.22 and Claim 1 on p. 22 of [2], under (A.1) there exists a finite positive constant κ\kappa and a full probability event Ω1\Omega_{1} such that for any ωΩ1\omega\in\Omega_{1} and yd{𝟎}y\in\mathbb{Z}^{d}\setminus\{\mathbf{0}\}, there exists a strictly increasing random sequence m(n)m(n)\in\mathbb{N} such that m(n+1)/m(n)1m(n+1)/m(n)\to 1 as nn\to\infty and

(A.16) Tm(n)y,z(1+ω+)κ|m(n)yz|1for all zd and n.\displaystyle T_{m(n)y,z}(1+\omega^{+})\leq\kappa\hskip 0.55pt|m(n)y-z|_{1}\quad\text{for all }z\in\mathbb{Z}^{d}\text{ and }n\in\mathbb{N}.

Here, Tx,y(1+ω+)T_{x,y}(1+\omega^{+}) is the first-passage time from xx to yy under the weights 1+ω+=(1+t(e)+:ed)1+\omega^{+}=(1+t(e)^{+}:e\in{\mathcal{E}}_{d}). The results from [2] apply because these weights are strictly positive and satisfy (A.1) with p=dp=d.

Let Ω2\Omega_{2} be the full probability event on which (A.5) holds for the random variable AA of (A.6). We show that (A.8) holds for each fixed ωΩ0Ω1Ω2\omega\in\Omega_{0}\cap\Omega_{1}\cap\Omega_{2}. Let xkdx_{k}\in\mathbb{Z}^{d} be an ω\omega-dependent sequence such that |xk|1|x_{k}|_{1}\to\infty and

(A.17) limk|T𝟎,xkμ(xk)||xk|1=lim¯nsup|x|1n|T𝟎,xμ(x)||x|1.\displaystyle\lim_{k\to\infty}\frac{|T_{\mathbf{0},x_{k}}-\mu(x_{k})|}{|x_{k}|_{1}}=\varlimsup_{n\to\infty}\sup_{|x|_{1}\geq n}\frac{|T_{\mathbf{0},x}-\mu(x)|}{|x|_{1}}\,.

By passing to a subsequence we can assume xk/|xk|1ξdx_{k}/|x_{k}|_{1}\to\xi\in\mathbb{R}^{d} with |ξ|1=1|\xi|_{1}=1. Let ζd\zeta\in\mathbb{Q}^{d} satisfy |ζ|1=1|\zeta|_{1}=1 and pick \ell\in\mathbb{N} such that ζd\ell\zeta\in\mathbb{Z}^{d}. Choose m(n)m(n) in (A.16) for y=ζy=\ell\zeta. For each kk\in\mathbb{N} take nkn_{k}\in\mathbb{N} such that

(A.18) m(nk)|xk|1m(nk+1).\displaystyle m(n_{k})\ell\leq|x_{k}|_{1}\leq m(n_{k}+1)\ell.

Abbreviate mk=m(nk)m_{k}=m(n_{k}). There exists ω\omega-dependent k0k_{0} such that for all kk0k\geq k_{0}, m(nk+1)2mkm(n_{k}+1)\leq 2m_{k}. Triangle inequality:

|T𝟎,xk|xk|1μ(ξ)|\displaystyle\Bigl{|}\frac{T_{\mathbf{0},x_{k}}}{|x_{k}|_{1}}-\mu(\xi)\Bigr{|} |T𝟎,xkT𝟎,mkζ||xk|1+mk|xk|1|T𝟎,mkζmkμ(ζ)|\displaystyle\leq\frac{|T_{\mathbf{0},x_{k}}-T_{\mathbf{0},m_{k}\ell\zeta}|}{|x_{k}|_{1}}+\frac{m_{k}\ell}{|x_{k}|_{1}}\cdot\Bigl{|}\frac{T_{\mathbf{0},m_{k}\ell\zeta}}{m_{k}\ell}-\mu(\zeta)\Bigr{|}
+|mk|xk|11||μ(ζ)|+|μ(ζ)μ(ξ)|.\displaystyle\qquad+\Bigl{|}\frac{m_{k}\ell}{|x_{k}|_{1}}-1\Bigr{|}\cdot|\mu(\zeta)|+|\mu(\zeta)-\mu(\xi)|.

Use (A.13), (A.16) applied to y=ζy=\ell\zeta, and take kk0k\geq k_{0}:

|T𝟎,xk|xk|1μ(ξ)|\displaystyle\Bigl{|}\frac{T_{\mathbf{0},x_{k}}}{|x_{k}|_{1}}-\mu(\xi)\Bigr{|} Tmkζ,xk(1+ω+)+Amkζ+Axk|xk|1+mk|xk|1|T𝟎,mkζmkμ(ζ)|\displaystyle\leq\frac{T_{m_{k}\ell\zeta,x_{k}}(1+\omega^{+})+A_{m_{k}\ell\zeta}+A_{x_{k}}}{|x_{k}|_{1}}+\frac{m_{k}\ell}{|x_{k}|_{1}}\cdot\Bigl{|}\frac{T_{\mathbf{0},m_{k}\ell\zeta}}{m_{k}\ell}-\mu(\zeta)\Bigr{|}
+|mk|xk|11||μ(ζ)|+|μ(ζ)μ(ξ)|\displaystyle\qquad+\Bigl{|}\frac{m_{k}\ell}{|x_{k}|_{1}}-1\Bigr{|}\cdot|\mu(\zeta)|+|\mu(\zeta)-\mu(\xi)|
κ|mkζxk|1|xk|1+2max|x|12mkAxmk+mk|xk|1|T𝟎,mkζmkμ(ζ)|\displaystyle\leq\frac{\kappa|m_{k}\ell\zeta-x_{k}|_{1}}{|x_{k}|_{1}}+\frac{2\max_{|x|_{1}\leq 2m_{k}\ell}A_{x}}{m_{k}\ell}+\frac{m_{k}\ell}{|x_{k}|_{1}}\cdot\Bigl{|}\frac{T_{\mathbf{0},m_{k}\ell\zeta}}{m_{k}\ell}-\mu(\zeta)\Bigr{|}
+|mk|xk|11||μ(ζ)|+|μ(ζ)μ(ξ)|.\displaystyle\qquad+\Bigl{|}\frac{m_{k}\ell}{|x_{k}|_{1}}-1\Bigr{|}\cdot|\mu(\zeta)|+|\mu(\zeta)-\mu(\xi)|.

As kk\to\infty the right-hand side converges to κ|ζξ|+|μ(ζ)μ(ξ)|\kappa|\zeta-\xi|+|\mu(\zeta)-\mu(\xi)|. Letting ζξ\zeta\to\xi then proves that T𝟎,xk/|xk|1μ(ξ)T_{\mathbf{0},x_{k}}/|x_{k}|_{1}\to\mu(\xi) as kk\to\infty. Since μ\mu is continuous and homogeneous, we also have μ(xk)/|xk|1=μ(xk/|xk|1)μ(ξ)\mu(x_{k})/|x_{k}|_{1}=\mu(x_{k}/|x_{k}|_{1})\to\mu(\xi). Now (A.8) follows from (A.17).

(A.9) follows from (A.8) and the continuity and homogeneity of μ\mu. ∎

Remark A.5.

In the last inequality of the proof above (mk)1max|x|12mkAx(m_{k}\ell)^{-1}\max_{|x|_{1}\leq 2m_{k}\ell}A_{x} can be replaced by a smaller term as follows. First, fix a rational ε>0\varepsilon>0. Take k0k_{0} to be large enough so that for all kk0k\geq k_{0}, m(nk+1)2mkm(n_{k}+1)\leq 2m_{k}, as before, but also

|mk|xk|11|ε3and|xk|xk|ξ|1ε3.\Bigl{|}\frac{m_{k}\ell}{|x_{k}|_{1}}-1\Bigr{|}\leq\frac{\varepsilon}{3\ell}\quad\text{and}\quad\Bigl{|}\frac{x_{k}}{|x_{k}|}-\xi\Bigr{|}_{1}\leq\frac{\varepsilon}{3\ell}\,.

Take ζd\zeta\in\mathbb{Q}^{d} (still with |ζ|1=1|\zeta|_{1}=1) so that |ζξ|1ε/(3)|\zeta-\xi|_{1}\leq\varepsilon/(3\ell). Now we have

|xkmkζ|1|xk|1|mk|xk|11|+|ζξ|1+|xk|xk|ξ|1ε.\frac{|x_{k}-m_{k}\ell\zeta|_{1}}{|x_{k}|_{1}}\leq\Bigl{|}\frac{m_{k}\ell}{|x_{k}|_{1}}-1\Bigr{|}+|\zeta-\xi|_{1}+\Bigl{|}\frac{x_{k}}{|x_{k}|}-\xi\Bigr{|}_{1}\leq\frac{\varepsilon}{\ell}\,.

Consequently,

|xkmkζ|1ε1|xk|1εm(nk+1)2εmk.|x_{k}-m_{k}\ell\zeta|_{1}\leq\varepsilon\ell^{-1}|x_{k}|_{1}\leq\varepsilon m(n_{k}+1)\leq 2\varepsilon m_{k}.

Thus, instead of (A.5) one now needs the hypothesis that for any fixed zdz\in\mathbb{Z}^{d},

limε0lim¯mm1maxx:|xmz|1εmAθx=0almost surely.\lim_{\varepsilon\searrow 0}\;\varlimsup_{m\to\infty}\;m^{-1}\max_{x:\,|x-mz|_{1}\leq\varepsilon m}A\circ\theta_{x}=0\quad\text{almost surely}.

In our application to the proof of Theorem A.1 the random variable AA has all moments, so we do not pursue sharper assumptions on AA than those stated in Theorem A.4. \triangle

Proof of Theorem A.1.

The constant in point (a) is taken to be ε0=δ=\varepsilon_{0}=\delta= the constant in the bound (4.10) for the shifted weights ω(r0)\omega^{(-r_{0})}. Then by Lemma A.3, A(b)A^{(b)} has all moments for all b>r0ε0b>-r_{0}-\varepsilon_{0}. Hence we can apply Theorem A.4 to the shifted weights ω(b)\omega^{(b)} to define the shape functions μ(b)\mu^{(b)} whose existence is asserted in point (b). We specify the full-probability event Ω0\Omega_{0} for point (c) in the course of the proof.

Proof of part (i). Start with the obvious point that T𝟎,x(b)T(b)(γ~)<T^{(b)}_{\mathbf{0},x}\leq T^{(b)}(\widetilde{\gamma})<\infty for any particular self-avoiding path γ~\widetilde{\gamma} from 𝟎\mathbf{0} to xx. By bound (4.10) there exists a full probability event Ω0\Omega_{0} and a finite random variable K(ω)K(\omega) such that on the event Ω0\Omega_{0}, every self-avoiding path γ\gamma from the origin such that |γ|K|\gamma|\geq K satisfies the bound T(r0)(γ)>ε0|γ|T^{(-r_{0})}(\gamma)>\varepsilon_{0}|\gamma|. Then for any shift bb these paths satisfy

(A.19) T(b)(γ)=T(r0)(γ)+(r0+b)|γ|>(ε0+r0+b)|γ|.T^{(b)}(\gamma)=T^{(-r_{0})}(\gamma)+(r_{0}+b)|\gamma|>(\varepsilon_{0}+r_{0}+b)|\gamma|.

From this we conclude that, for any xdx\in\mathbb{Z}^{d}, b>r0ε0b>-r_{0}-\varepsilon_{0}, and any path γ\gamma,

(A.20) |γ|KT𝟎,x(b)ε0+r0+bimpliesT(b)(γ)>T𝟎,x(b).|\gamma|\;\geq\;K\vee\frac{T^{(b)}_{\mathbf{0},x}}{\varepsilon_{0}+r_{0}+b}\qquad\text{implies}\qquad T^{(b)}(\gamma)>T^{(b)}_{\mathbf{0},x}.

Thus the infimum that defines T𝟎,x(b)T^{(b)}_{\mathbf{0},x} in (2.2) cannot be taken outside a certain ω\omega-dependent finite set of paths. Consequently on the event Ω0\Omega_{0} a minimizing path exists and both T𝟎,x(b)T^{(b)}_{\mathbf{0},x} and L¯𝟎,x(b)\overline{L}^{(b)}_{\mathbf{0},x} are finite for all xdx\in\mathbb{Z}^{d} and b>r0ε0b>-r_{0}-\varepsilon_{0}.

Next, shrink the event Ω0\Omega_{0} (if needed) so that for ωΩ0\omega\in\Omega_{0} the shape theorem (A.8) is valid for the weights ω(r0)\omega^{(-r_{0})}. Then we can increase KK and pick a deterministic positive constant cc so that T𝟎,x(r0)c|x|1T^{(-r_{0})}_{\mathbf{0},x}\leq c|x|_{1} whenever |x|1K|x|_{1}\geq K. By monotonicity T𝟎,x(b)c|x|1T^{(b)}_{\mathbf{0},x}\leq c|x|_{1} for all br0b\leq-r_{0} whenever |x|1K|x|_{1}\geq K. If necessary increase cc so that cε0c\geq\varepsilon_{0}. Then by (A.20), when |x|1K|x|_{1}\geq K and b(r0ε0,r0]b\in(-r_{0}-\varepsilon_{0},-r_{0}], a self-avoiding path γ\gamma between 𝟎\mathbf{0} and xx that satisfies

|γ|c|x|1ε0+r0+b|\gamma|\geq\frac{c|x|_{1}}{\varepsilon_{0}+r_{0}+b}

cannot be a geodesic for T𝟎,x(b)T^{(b)}_{\mathbf{0},x}. We conclude that for ωΩ0\omega\in\Omega_{0},

L¯𝟎,x(b)cε0+r0+b|x|1whenever b(r0ε0,r0] and |x|1K.\overline{L}^{(b)}_{\mathbf{0},x}\leq\frac{c}{\varepsilon_{0}+r_{0}+b}|x|_{1}\qquad\text{whenever }\ b\in(-r_{0}-\varepsilon_{0},-r_{0}]\ \text{ and }\ |x|_{1}\geq K.

Since L¯𝟎,x(b)\overline{L}^{(b)}_{\mathbf{0},x} is nonincreasing in bb (Remark 2.4(iii)), we can extend the bound above to all br0b\geq-r_{0} in the form (A.2).

By taking advantage of (A.2) now proved, we get these Lipschitz bounds: for all ωΩ0\omega\in\Omega_{0}, b2>b1>r0ε0b_{2}>b_{1}>-r_{0}-\varepsilon_{0} and |x|1K|x|_{1}\geq K, and with π𝟎,x(b)\pi^{(b)}_{\mathbf{0},x} denoting a geodesic of T𝟎,x(b)T^{(b)}_{\mathbf{0},x},

(A.21) T𝟎,x(b1)\displaystyle T^{(b_{1})}_{\mathbf{0},x} T𝟎,x(b2)T(b2)(π𝟎,x(b1))=T(b1)(π𝟎,x(b1))+(b2b1)|π𝟎,x(b1)|\displaystyle\leq T^{(b_{2})}_{\mathbf{0},x}\leq T^{(b_{2})}(\pi^{(b_{1})}_{\mathbf{0},x})=T^{(b_{1})}(\pi^{(b_{1})}_{\mathbf{0},x})+(b_{2}-b_{1})|\pi^{(b_{1})}_{\mathbf{0},x}|
T𝟎,x(b1)+c(b2b1)|x|1ε0+(r0+b1)0T𝟎,x(b1)+κ(b1)(b2b1)|x|1.\displaystyle\leq T^{(b_{1})}_{\mathbf{0},x}+\frac{c\hskip 0.9pt(b_{2}-b_{1})|x|_{1}}{\varepsilon_{0}+(r_{0}+b_{1})\wedge 0}\equiv T^{(b_{1})}_{\mathbf{0},x}+\kappa(b_{1})(b_{2}-b_{1})|x|_{1}.

The last equality defines the constant κ(b)\kappa(b) which is nonincreasing in bb.

We establish the locally uniform shape theorem (A.3). Let BB be a countable dense subset of (r0ε0,)(-r_{0}-\varepsilon_{0},\infty). Shrink the event Ω0\Omega_{0} further so that for ωΩ0\omega\in\Omega_{0} the shape theorem (A.8) holds for the shifted weights ω(b)\omega^{(b)} for all bBb\in B. By passing to the limit, (A.21) gives the macroscopic Lipschitz bounds (A.4) for shifts b1<b2b_{1}<b_{2} in this countable dense set BB.

Let ωΩ0\omega\in\Omega_{0}, ε>0\varepsilon>0, and a0<a1a_{0}<a_{1} in BB. Pick a partition a0=b0<b1<<bm=a1a_{0}=b_{0}<b_{1}<\dotsm<b_{m}=a_{1} so that each biBb_{i}\in B and κ(a0)(bibi1)<ε/2\kappa(a_{0})(b_{i}-b_{i-1})<\varepsilon/2. Fix a constant K0=K0b0,b1,,bm(ω)K_{0}=K_{0}^{b_{0},b_{1},\dotsc,b_{m}}(\omega) such that

|T𝟎,x(bi)μ(bi)(x)|ε|x|1/2for i=0,1,,m whenever |x|1K0.|\,T^{(b_{i})}_{\mathbf{0},x}-\mu^{(b_{i})}(x)\,|\leq\varepsilon|x|_{1}/2\qquad\text{for $i=0,1,\dotsc,m$ whenever }\ |x|_{1}\geq K_{0}.

Now for i[m]i\in[m], b[bi1,bi]b\in[b_{i-1},b_{i}], and |x|1K0|x|_{1}\geq K_{0}, utilizing the monotonicity in bb of T𝟎,x(b)T^{(b)}_{\mathbf{0},x} and μ(b)(x)\mu^{(b)}(x) and the Lipschitz bounds (A.21) and (A.4),

|T𝟎,x(b)μ(b)(x)||T𝟎,x(bi)μ(bi)(x)|+κ(a0)(bibi1)|x|1ε|x|1.\displaystyle|\,T^{(b)}_{\mathbf{0},x}-\mu^{(b)}(x)|\leq|\,T^{(b_{i})}_{\mathbf{0},x}-\mu^{(b_{i})}(x)\hskip 0.9pt|+\kappa(a_{0})(b_{i}-b_{i-1})|x|_{1}\leq\varepsilon|x|_{1}.

The shape theorem (A.3) has been proved.

Proof of part (ii). The integrability and L1L^{1} convergence follow from Theorem A.4(i). The almost sure convergence comes from the homogeneity and continuity of μ(b)\mu^{(b)} and the shape theorem (A.3).

Proof of part (iii). We already established (A.4) for a dense set of shifts b1<b2b_{1}<b_{2}. Monotonicity of bμ(b)(ξ)b\mapsto\mu^{(b)}(\xi) extends (A.4) to all shifts bb.

That μ(b)(𝟎)=0\mu^{(b)}(\mathbf{0})=0 follows from homogeneity. The final claim that μ(b)(ξ)>0\mu^{(b)}(\xi)>0 for b>r0ε0b>-r_{0}-\varepsilon_{0} and ξ𝟎\xi\neq\mathbf{0} follows from (A.19), which implies T𝟎,x(b)(ε0+r0+b)|x|1T^{(b)}_{\mathbf{0},x}\geq(\varepsilon_{0}+r_{0}+b)|x|_{1} whenever |x|1K|x|_{1}\geq K. ∎

Appendix B Restricted path length shape theorem

This section proves the next shape theorem in the interior of 𝒰\mathcal{U} for the restricted path length FPP processes defined in (2.24). As throughout, the edge weights {t(e):ed}\{t(e):e\in{\mathcal{E}}_{d}\} are independent and identically distributed (i.i.d.) real-valued random variables, r0=essinft(e)r_{0}=\mathop{\mathrm{ess\,inf}}t(e), the set 𝒟\mathcal{D}^{\,\diamond}_{\ell} of points reachable by \ell-paths is defined by (2.23), and 𝒰={ξd:|ξ|11}\mathcal{U}=\{\xi\in\mathbb{R}^{d}:|\xi|_{1}\leq 1\} is the 1\ell^{1} unit ball. We also write {ti}\{t_{i}\} for i.i.d. copies of the edge weight t(e)t(e).

Theorem B.1.

Assume r0>r_{0}>-\infty and moment assumption (A.1) with p=dp=d. Fix {𝚎𝚖𝚙𝚝𝚢,o}\diamond\in\{\langle{\tt empty}\rangle,o\}. There exists a deterministic, continuous, convex shape function g:int𝒰[r00,)g^{\diamond}:\operatorname{int}\mathcal{U}\to[r_{0}\wedge 0,\infty) that satisfies the following: for each α,ε>0\alpha,\varepsilon>0 there exists an almost-surely finite random constant K(α,ε)K(\alpha,\varepsilon) such that

(B.1) |G𝟎,(k),xkg(x/k)|εk|\,G^{\,\diamond}_{\mathbf{0},(k),x}-k\hskip 0.7ptg^{\diamond}(x/k)\,|\leq\varepsilon k

whenever kK(α,ε)k\geq K(\alpha,\varepsilon), k(1+α)|x|1k\geq(1+\alpha)|x|_{1}, and x𝒟kx\in\mathcal{D}^{\,\diamond}_{k}.

The shape theorem can be proved all the way to the boundary of 𝒰\mathcal{U}. This requires (i) stronger moment bounds that vary with the dimension of each boundary face and (ii) further technical constructions beyond what is done in the proof below, because there are fewer paths to the boundary than to interior points. We have no need for the shape theorem on all of 𝒰\mathcal{U} in the present paper. Our purposes are met by extending the shape function from the interior to the boundary through radial limits (Theorem 2.10 and Lemma 4.1).

We begin with a basic tail bound on G𝟎,(),xG^{\,\diamond}_{\mathbf{0},(\ell),x}.

Lemma B.2.

Assume the weights are arbitrary real-valued i.i.d. random variables. Let \ell\in\mathbb{N}, {𝚎𝚖𝚙𝚝𝚢,o}\diamond\in\{\langle{\tt empty}\rangle,o\} and x𝒟x\in\mathcal{D}^{\,\diamond}_{\ell}. Assume |x|18\ell-|x|_{1}\geq 8. Then for any real s0s\geq 0,

(B.2) {G𝟎,(),xs}2d{min(t1,,t2d)s/}.\displaystyle\mathbb{P}\{G^{\,\diamond}_{\mathbf{0},(\ell),x}\geq s\}\leq\ell^{\hskip 0.55pt2d}\hskip 0.9pt\mathbb{P}\{\min(t_{1},\dotsc,t_{2d})\geq s/\ell\}\,.
Proof.

It is enough to prove the lemma for =𝚎𝚖𝚙𝚝𝚢\diamond=\langle{\tt empty}\rangle. Then we can assume that |x|1\ell-|x|_{1} is an even integer because otherwise Π𝟎,(),x=\Pi_{\mathbf{0},(\ell),x}=\varnothing. The reason that the case =o\diamond=o is also covered is that G𝟎,(),xoG𝟎,(1),xG𝟎,(),xG^{o}_{\mathbf{0},(\ell),x}\leq G_{\mathbf{0},(\ell-1),x}\wedge G_{\mathbf{0},(\ell),x}.

To prove (B.2) we construct a total of 2d2d edge-disjoint paths in Π𝟎,(),x\Pi_{\hskip 0.55pt\mathbf{0},(\ell),x}. Let A={z:zx>0}A=\{z\in\mathcal{R}:z\cdot x>0\} and B={z:zx=0}B=\{z\in\mathcal{R}:z\cdot x=0\} with cardinalities ν10\nu_{1}\geq 0 and ν2=2d2ν1\nu_{2}=2d-2\nu_{1}, respectively. Enumerate these sets as A={z1,,zν1}A=\{z_{1},\dotsc,z_{\nu_{1}}\} and B={zν1+1,,zν1+ν2}B=\{z_{\nu_{1}+1},\dotsc,z_{\nu_{1}+\nu_{2}}\}.

Suppose x𝟎x\neq\mathbf{0}, in which case ν11\nu_{1}\geq 1. For each i[ν1]i\in[\nu_{1}] let πiΠ𝟎,(|x|1),x\pi_{i}^{\prime}\in\Pi_{\hskip 0.55pt\mathbf{0},(|x|_{1}),x} be the 1\ell^{1}-path from 𝟎\mathbf{0} to xx that takes the necessary steps in the order zi,zi+1,,zν1,z1,,zi1z_{i},z_{i+1},\dotsc,z_{\nu_{1}},z_{1},\dotsc,z_{i-1}. Then for each i[ν1]i\in[\nu_{1}] let πiΠ𝟎,(),x\pi_{i}\in\Pi_{\hskip 0.55pt\mathbf{0},(\ell),x} be the path that starts with (|x|1)/2(\ell-|x|_{1})/2 repetitions of the (zi,zi)(z_{i},-z_{i}) pair and then follows πi\pi^{\prime}_{i}. For i[ν2]i\in[\nu_{2}] let πν1+iΠ𝟎,(),x\pi_{\nu_{1}+i}\in\Pi_{\hskip 0.55pt\mathbf{0},(\ell),x} be the path that starts with a zν1+iz_{\nu_{1}+i} step, then repeats the (z1,z1)(z_{1},-z_{1}) pair (|x|12)/2(\ell-|x|_{1}-2)/2 times, then follows the steps of π1\pi^{\prime}_{1}, and finishes with a zν1+i-z_{\nu_{1}+i} step. Thus far we have constructed ν1+ν2=2dν1\nu_{1}+\nu_{2}=2d-\nu_{1} paths. For the remaining ν1\nu_{1} paths we distinguish two cases.

If ν1=1\nu_{1}=1 we need only one more path π2dΠ𝟎,(),x\pi_{2d}\in\Pi_{\hskip 0.55pt\mathbf{0},(\ell),x}. Take this to be the path that starts with a z1-z_{1} step, repeats the (z1,z1)(z_{1},-z_{1}) pair (|x|18)/2(\ell-|x|_{1}-8)/2 times, takes two z2z_{2} steps, one z1z_{1} step, follows the steps of π1\pi^{\prime}_{1}, takes one z1z_{1} step, two z2-z_{2} steps, and finishes with a z1-z_{1} step.

If ν1>1\nu_{1}>1, then for i[ν11]i\in[\nu_{1}-1], let πν1+ν2+iΠ𝟎,(),x\pi_{\nu_{1}+\nu_{2}+i}\in\Pi_{\hskip 0.55pt\mathbf{0},(\ell),x} be the path that starts with a zi-z_{i} step, repeats the (zi,zi)(z_{i},-z_{i}) pair (|x|14)/2(\ell-|x|_{1}-4)/2 times, takes a zi+1z_{i+1} step, follows the steps of πi+1\pi^{\prime}_{i+1}, and ends with a ziz_{i} step followed by a zi+1-z_{i+1} step. For i=ν1i=\nu_{1} the path π2d\pi_{2d} is defined similarly, except that zi+1z_{i+1} and πi+1\pi^{\prime}_{i+1} are replaced by z1z_{1} and π1\pi^{\prime}_{1}, respectively.

One can check that the paths πiΠ𝟎,(),x\pi_{i}\in\Pi_{\hskip 0.55pt\mathbf{0},(\ell),x}, i[2d]i\in[2d], are edge-disjoint. From

(B.3) G𝟎,(),xmini[2d]T(πi)\displaystyle G_{\mathbf{0},(\ell),x}\leq\min_{i\in[2d]}T(\pi_{i})

follows

(B.4) {G𝟎,(),xs}\displaystyle\mathbb{P}\{G_{\mathbf{0},(\ell),x}\geq s\} i=12d{T(πi)s}({t(e)s/})2d\displaystyle\leq\prod_{i=1}^{2d}\mathbb{P}\{T(\pi_{i})\geq s\}\leq\bigl{(}\ell\hskip 0.7pt\mathbb{P}\{t(e)\geq s/\ell\}\bigr{)}^{2d}
=2d{min(t1,,t2d)s/}.\displaystyle=\ell^{2d}\mathbb{P}\{\min(t_{1},\dotsc,t_{2d})\geq s/\ell\}\,.

If x=𝟎x=\mathbf{0} (and hence ν1=0\nu_{1}=0 and ν2=2d\nu_{2}=2d) then redo the last computation with the edge-disjoint paths πi\pi_{i}, i[2d]i\in[2d], that just repeat the pair (zi,zi)(z_{i},-z_{i}). ∎

Below we use the condition that a rational point ζ𝒰\zeta\in\mathcal{U} satisfies ζ𝒟\ell\zeta\in\mathcal{D}^{\,\diamond}_{\ell} for a positive integer \ell such that ζd\ell\zeta\in\mathbb{Z}^{d}. When zero steps are admissible (=o\diamond=o) this is of course trivial, and without zero steps (=𝚎𝚖𝚙𝚝𝚢\diamond=\langle{\tt empty}\rangle) this can be achieved if (1|ζ|1)\ell(1-|\zeta|_{1}) is even. Therefore, one can take for example =2\ell=2\ell^{\prime} for \ell^{\prime}\in\mathbb{N} such that ζd\ell^{\prime}\zeta\in\mathbb{Z}^{d}. Properties of convex sets used below can be found in Chapter 18 of [17].

The following theorem comes by a standard application of the subadditive ergodic theorem.

Theorem B.3.

Assume r0>r_{0}>-\infty. Fix ζd𝒰\zeta\in\mathbb{Q}^{d}\cap\mathcal{U} and {𝚎𝚖𝚙𝚝𝚢,o}\diamond\in\{\langle{\tt empty}\rangle,o\}. Let \ell\in\mathbb{N} be such that ζ𝒟\ell\zeta\in\mathcal{D}^{\,\diamond}_{\ell}. Assume 𝔼[G0,(),ζ]<\mathbb{E}[G^{\,\diamond}_{0,(\ell),\ell\zeta}]<\infty. Then the limit

(B.5) g(ζ)=infn𝔼[G𝟎,(n),nζ]n=limnG𝟎,(n),nζn[r00,1𝔼[G0,(),ζ]]\displaystyle g^{\diamond}(\zeta)=\inf_{n\hskip 0.55pt\in\hskip 0.55pt\mathbb{N}}\frac{\mathbb{E}[G^{\,\diamond}_{\mathbf{0},(n\ell),n\ell\zeta}]}{n\ell}=\lim_{n\to\infty}\frac{G^{\,\diamond}_{\mathbf{0},(n\ell),n\ell\zeta}}{n\ell}\in\bigl{[}r_{0}\wedge 0,\,\ell^{-1}\mathbb{E}[G^{\,\diamond}_{0,(\ell),\ell\zeta}]\bigr{]}

exists almost surely and in L1L^{1} and does not depend on the choice of \ell. As a function of ζd𝒰\zeta\in\mathbb{Q}^{d}\cap\mathcal{U}, gg^{\diamond} is convex. Precisely, if ζ,ηd𝒰\zeta,\eta\in\mathbb{Q}^{d}\cap\mathcal{U} are such that 𝔼[G0,(),ζ]<\mathbb{E}[G^{\,\diamond}_{0,(\ell),\ell\zeta}]<\infty and 𝔼[G0,(),η]<\mathbb{E}[G^{\,\diamond}_{0,(\ell),\ell\eta}]<\infty for some \ell\in\mathbb{N}, then for any t(0,1)t\in(0,1)\cap\mathbb{Q}, 𝔼[G0,(),(tζ+(1t)η)]<\mathbb{E}[G^{\,\diamond}_{0,(\ell^{\prime}),\ell^{\prime}(t\zeta+(1-t)\eta)}]<\infty for some \ell^{\prime}\in\mathbb{N} and

(B.6) g(tζ+(1t)η)tg(ζ)+(1t)g(η).\displaystyle g^{\diamond}\bigl{(}t\zeta+(1-t)\eta\bigr{)}\leq tg^{\diamond}(\zeta)+(1-t)g^{\diamond}(\eta).
Remark B.4 (Conditions for finiteness).

By Lemma B.2, assumption (A.1) with p=1p=1 implies that 𝔼[G0,(),ζ]<\mathbb{E}[G^{\,\diamond}_{0,(\ell),\ell\zeta}]<\infty for any ζd𝒰\zeta\in\mathbb{Q}^{d}\cap\mathcal{U} and any large enough \ell\in\mathbb{N} that satisfies ζ𝒟\ell\zeta\in\mathcal{D}^{\,\diamond}_{\ell}.

Next, from convexity we deduce local boundedness and then a local Lipschitz property.

Lemma B.5.

Assume r0>r_{0}>-\infty and (A.1) with p=1p=1. Fix ζdint𝒰\zeta\in\mathbb{Q}^{d}\cap\operatorname{int}\mathcal{U} and {𝚎𝚖𝚙𝚝𝚢,o}\diamond\in\{\langle{\tt empty}\rangle,o\}. There exist ε>0\varepsilon>0 and a finite constant CC such that

(B.7) g(η)Cfor all ηd𝒰 such that |ηζ|1ε.\displaystyle g^{\diamond}(\eta)\leq C\quad\text{for all }\eta\in\mathbb{Q}^{d}\cap\mathcal{U}\text{ such that }|\eta-\zeta|_{1}\leq\varepsilon.
Proof.

Take ε>0\varepsilon>0 rational and small enough so that 𝒜={η𝒰:|ηζ|1ε}int𝒰\mathcal{A}=\{\eta\in\mathcal{U}:|\eta-\zeta|_{1}\leq\varepsilon\}\subset\operatorname{int}\mathcal{U}. Let {ηi:i[2d]}dint𝒰\{\eta_{i}:i\in[2d]\}\subset\mathbb{Q}^{d}\cap\operatorname{int}\mathcal{U} be the extreme points of the convex set 𝒜\mathcal{A}. For ηd𝒜\eta\in\mathbb{Q}^{d}\cap\mathcal{A} write η=i=12dαiηi\eta=\sum_{i=1}^{2d}\alpha_{i}\eta_{i} with rational αi[0,1]\alpha_{i}\in[0,1] such that i[2d]αi=1\sum_{i\in[2d]}\alpha_{i}=1. By bound (B.5) and Remark B.4, g(ηi)<g^{\diamond}(\eta_{i})<\infty for i[2d]i\in[2d]. Convexity (B.6) implies

g(η)i[2d]αig(ηi)maxi[2d]g(ηi)g^{\diamond}(\eta)\leq\sum_{i\in[2d]}\alpha_{i}g^{\diamond}(\eta_{i})\leq\max_{i\in[2d]}g^{\diamond}(\eta_{i})

and Lemma B.5 is proved. ∎

Lemma B.6.

Assume r0>r_{0}>-\infty and (A.1) with p=1p=1. Fix ζdint𝒰\zeta\in\mathbb{Q}^{d}\cap\operatorname{int}\mathcal{U}. There exist ε>0\varepsilon>0 and a finite positive constant C=C(ζ,ε,r0)C=C(\zeta,\varepsilon,r_{0}) such that for both {𝚎𝚖𝚙𝚝𝚢,o}\diamond\in\{\langle{\tt empty}\rangle,o\}

|g(η)g(η)|C|ηη|1η,ηdint𝒰 with |ηζ|1ε and |ηζ|1ε.|g^{\diamond}(\eta)-g^{\diamond}(\eta^{\prime})|\leq C|\eta-\eta^{\prime}|_{1}\quad\forall\eta,\eta^{\prime}\in\mathbb{Q}^{d}\cap\operatorname{int}\mathcal{U}\text{ with }|\eta-\zeta|_{1}\leq\varepsilon\text{ and }|\eta^{\prime}-\zeta|_{1}\leq\varepsilon.
Proof.

The assumptions of Lemma B.5 are satisfied and therefore there exists a rational ε>0\varepsilon>0 and a finite constant CC such that (B.7) holds. By taking ε>0\varepsilon>0 smaller, if necessary, we can also guarantee that for any ηd\eta\in\mathbb{R}^{d}, |ηζ|1ε|\eta-\zeta|_{1}\leq\varepsilon implies ηint𝒰\eta\in\operatorname{int}\mathcal{U}.

Take ηη\eta\neq\eta^{\prime} in int𝒰\operatorname{int}\mathcal{U} with |ηζ|1ε/2|\eta-\zeta|_{1}\leq\varepsilon/2 and |ηζ|1ε/2|\eta^{\prime}-\zeta|_{1}\leq\varepsilon/2. Abbreviate δ=2ε1|ηη|1\delta=2\varepsilon^{-1}|\eta-\eta^{\prime}|_{1} and write

η=11+δη+δ1+δ(η+δ1(ηη)).\eta=\frac{1}{1+\delta}\cdot\eta^{\prime}+\frac{\delta}{1+\delta}\cdot\bigl{(}\eta+\delta^{-1}(\eta-\eta^{\prime})\bigr{)}.

Note that

|η+δ1(ηη)ζ|1ε2+δ1|ηη|1=ε.|\eta+\delta^{-1}(\eta-\eta^{\prime})-\zeta|_{1}\leq\frac{\varepsilon}{2}+\delta^{-1}|\eta-\eta^{\prime}|_{1}=\varepsilon.

Therefore, η+δ1(ηη)int𝒰\eta+\delta^{-1}(\eta-\eta^{\prime})\in\operatorname{int}\mathcal{U}. By convexity (B.6) and boundedness (B.7) we have

g(η)11+δg(η)+δ1+δg(η+δ1(ηη))11+δg(η)+Cδ1+δ.g^{\diamond}(\eta)\leq\frac{1}{1+\delta}\cdot g^{\diamond}(\eta^{\prime})+\frac{\delta}{1+\delta}\cdot g^{\diamond}\bigl{(}\eta+\delta^{-1}(\eta-\eta^{\prime})\bigr{)}\leq\frac{1}{1+\delta}\cdot g^{\diamond}(\eta^{\prime})+\frac{C\delta}{1+\delta}\,.

From Cg(η)r00C\geq g^{\diamond}(\eta^{\prime})\geq r_{0}\wedge 0,

g(η)g(η)δ1+δ(g(η)+C)δ(|r00|+C)=2ε1(|r00|+C)|ηη|1.\displaystyle g^{\diamond}(\eta)-g^{\diamond}(\eta^{\prime})\leq\frac{\delta}{1+\delta}(-g^{\diamond}(\eta^{\prime})+C)\leq\delta(|r_{0}\wedge 0|+C)=2\varepsilon^{-1}(|r_{0}\wedge 0|+C)\hskip 0.7pt|\eta-\eta^{\prime}|_{1}.

The other bound comes by switching around η\eta and η\eta^{\prime}. ∎

The next lemma is an immediate consequence of the local Lipschitz property proved in the previous lemma.

Lemma B.7.

Assume r0>r_{0}>-\infty and (A.1) with p=1p=1. Then gg and gog^{o} extend to locally Lipschitz, continuous, convex functions on int𝒰\operatorname{int}\mathcal{U}.

Before we prove the shape theorem we need two more auxiliary lemmas.

Lemma B.8.

Assume (A.1) with p=dp=d. Then there exists a finite constant κ\kappa such that

(B.8) {pair ϵ<ρ in (0,1)0=0(ϵ,ρ,ω) such that 0,{𝚎𝚖𝚙𝚝𝚢,o}:supy𝒟ϵ|y|1ρ1G𝟎,(),yκ}=1.\displaystyle\begin{split}&\mathbb{P}\Biggl{\{}\forall\hskip 0.7pt\text{pair }\epsilon<\rho\text{ in }(0,1)\ \exists\ell_{0}=\ell_{0}(\epsilon,\rho,\omega)\text{ such that }\\ &\qquad\qquad\forall\ell\geq\ell_{0},\,\forall\diamond\in\{\langle{\tt empty}\rangle,o\}:\sup_{\begin{subarray}{c}y\hskip 0.55pt\in\hskip 0.55pt\mathcal{D}^{\,\diamond}_{\ell}\\ \epsilon\ell\leq|y|_{1}\leq\rho\ell\end{subarray}}\ell^{-1}G^{\,\diamond}_{\mathbf{0},(\ell),y}\leq\kappa\Biggr{\}}=1.\end{split}
Proof.

It is enough to work with =𝚎𝚖𝚙𝚝𝚢\diamond=\langle{\tt empty}\rangle since G𝟎,(),yoG𝟎,(),yG^{o}_{\mathbf{0},(\ell),y}\leq G_{\mathbf{0},(\ell),y}. It is also enough to work with fixed ϵ<ρ\epsilon<\rho since the suprema in question increase as we increase ρ\rho and decrease ϵ\epsilon.

Fix an integer r5r\geq 5 such that

ρ(1+8/r)<1.\rho(1+8/r)<1.

The strategy of the proof will be to bound G𝟎,(),yG_{\mathbf{0},(\ell),y} by constructing edge-disjoint paths on the coarse-grained lattice rdr\mathbb{Z}^{d} to a point y¯\underline{y} that approximates yy. An approach to finding such paths was developed in the proof of Lemma B.2.

Take \ell large enough so that

(B.9) 2(d+8)rϵ1andρ(1+8/r)+(d+8)(r+8)+dr+8.\displaystyle\ell\geq 2(d+8)r\epsilon^{-1}\quad\text{and}\quad\rho(1+8/r)\ell+(d+8)(r+8)+dr+8\leq\ell.

For each y𝒟y\in\mathcal{D}_{\ell} with ϵ|y|1ρ\epsilon\ell\leq|y|_{1}\leq\rho\ell pick y¯rd\underline{y}\in r\mathbb{Z}^{d} so that |yy¯|1dr|y-\underline{y}|_{1}\leq dr. As in Lemma B.2, let ν1\nu_{1} be the number of zz\in\mathcal{R} such that y¯z>0\underline{y}\cdot z>0 and let ν2=2d2ν1\nu_{2}=2d-2\nu_{1}. Following the construction in the proof of Lemma B.2 we can produce edge-disjoint nearest-neighbor paths πi\pi^{\prime}_{i}, i[2d]i\in[2d], on the coarse-grained lattice rdr\mathbb{Z}^{d} from 𝟎\mathbf{0} to y¯\underline{y} such that, in terms of the number steps taken on rdr\mathbb{Z}^{d}, πi\pi^{\prime}_{i} has length ¯i=¯=|y¯|1/r\underline{\ell}_{i}=\underline{\ell}=|\underline{y}|_{1}/r for i[ν1]i\in[\nu_{1}], πi\pi^{\prime}_{i} has length ¯i=¯+2\underline{\ell}_{i}=\underline{\ell}+2 for i{ν1+1,,ν1+ν2}i\in\{\nu_{1}+1,\dotsc,\nu_{1}+\nu_{2}\}, and for i>ν1+ν2i>\nu_{1}+\nu_{2}, πi\pi^{\prime}_{i} has length ¯i=¯+8\underline{\ell}_{i}=\underline{\ell}+8 if ν1=1\nu_{1}=1 and ¯i=¯+4\underline{\ell}_{i}=\underline{\ell}+4 if ν1>1\nu_{1}>1.

From |y|1ρ|y|_{1}\leq\rho\ell and |yy¯|1dr|y-\underline{y}|_{1}\leq dr follows ¯=|y¯|1/r(ρ+dr)/r\underline{\ell}=|\underline{y}|_{1}/r\leq(\rho\ell+dr)/r, and then from (B.9)

(¯+8)(r+8)+|yy¯|1+8((ρ+dr)/r+8)(r+8)+dr+8.(\underline{\ell}+8)(r+8)+|y-\underline{y}|_{1}+8\leq\bigl{(}(\rho\ell+dr)/r+8\bigr{)}(r+8)+dr+8\leq\ell.

Define

q=12(8(r+8)|yy¯|18¯r8).q=\Bigl{\lfloor}\frac{1}{2}\Bigl{(}\frac{\ell-8(r+8)-|y-\underline{y}|_{1}-8}{\underline{\ell}}-r-8\Bigr{)}\Bigr{\rfloor}.

Then

0q2¯r2(|y|1dr)rϵ1.0\leq q\leq\frac{\ell}{2\underline{\ell}}\leq\frac{r\ell}{2(|y|_{1}-dr)}\leq r\epsilon^{-1}.

Define

(B.10) m=8(r+8)|yy¯|18¯(r+8+2q)2.m=\Bigl{\lfloor}\frac{\ell-8(r+8)-|y-\underline{y}|_{1}-8-\underline{\ell}(r+8+2q)}{2}\Bigr{\rfloor}.

Then

0m¯.0\leq m\leq\underline{\ell}.

Let πi,s\pi^{\prime}_{i,s} denote the position (on the original lattice) of the path πi\pi^{\prime}_{i} after ss steps (of size rr). Let

(B.11) i=(r+8)¯i+2q¯+2m.\ell_{i}=(r+8)\underline{\ell}_{i}+2q\underline{\ell}+2m.

For each i[2d]i\in[2d] we have this identity:

(r+10+2q)m+(r+8+2q)(¯m)+(r+8)(¯i¯)+i=.(r+10+2q)m+(r+8+2q)(\underline{\ell}-m)+(r+8)(\underline{\ell}_{i}-\underline{\ell})+\ell-\ell_{i}=\ell.

This equation gives a way to decompose the \ell steps from 𝟎\mathbf{0} to yy so that we first go through the vertices {πi,s}0s¯i\{\pi^{\prime}_{i,s}\}_{0\leq s\leq\underline{\ell}_{i}} and then use the last i\ell-\ell_{i} steps to go from y¯\underline{y} to yy. We continue with this next bound:

(B.12) G𝟎,(),y\displaystyle G_{\mathbf{0},(\ell),y} mini[2d]{s=0m1Gπi,s,(r+10+2q),πi,s+1+s=m¯1Gπi,s,(r+8+2q),πi,s+1\displaystyle\leq\min_{i\in[2d]}\biggl{\{}\;\sum_{s=0}^{m-1}G_{\pi^{\prime}_{i,s},(r+10+2q),\pi^{\prime}_{i,s+1}}+\sum_{s=m}^{\underline{\ell}-1}G_{\pi^{\prime}_{i,s},(r+8+2q),\pi^{\prime}_{i,s+1}}
+s=¯¯i1Gπi,s,(r+8),πi,s+1+Gy¯,(i),y}.\displaystyle\qquad\qquad\qquad\qquad\qquad\qquad+\sum_{s=\underline{\ell}}^{\underline{\ell}_{i}-1}G_{\pi^{\prime}_{i,s},(r+8),\pi^{\prime}_{i,s+1}}+G_{\underline{y},(\ell-\ell_{i}),y}\biggr{\}}.

Bound mm in (B.10) by dropping \lfloor{\ }\rfloor to turn (B.11) into this inequality (note that terms 2q¯2q\underline{\ell} cancel):

i\displaystyle\ell_{i} (r+8)(¯+8)+8(r+8)|yy¯|18(r+8)¯=|yy¯|18.\displaystyle\leq(r+8)(\underline{\ell}+8)+\ell-8(r+8)-|y-\underline{y}|_{1}-8-(r+8)\underline{\ell}=\ell-|y-\underline{y}|_{1}-8.

Similarly,

i|yy¯|110dr10.\ell_{i}\geq\ell-|y-\underline{y}|_{1}-10\geq\ell-dr-10.

Fix κ>0\kappa>0. For j=d+1,,2dj=d+1,\dotsc,2d let 𝐞j=𝐞jd\mathbf{e}_{j}=-\mathbf{e}_{j-d}. Define the events

1={j[2d]:G𝟎,(r+8+2q),r𝐞jκ/14 or G𝟎,(r+10+2q),r𝐞jκ/14},\displaystyle{\mathcal{E}}^{1}_{\ell}=\bigl{\{}\exists j\in[2d]:G_{\mathbf{0},(r+8+2q),r\mathbf{e}_{j}}\geq\kappa\ell/14\text{ or }G_{\mathbf{0},(r+10+2q),r\mathbf{e}_{j}}\geq\kappa\ell/14\bigr{\}},
2={j[2d]:Gr𝐞j,(r+8),𝟎κ/14 or Gr𝐞j,(r+8+2q),𝟎κ/14 or Gr𝐞j,(r+10+2q),𝟎κ/14},\displaystyle{\mathcal{E}}^{2}_{\ell}=\bigl{\{}\exists j\in[2d]:G_{-r\mathbf{e}_{j},(r+8),\mathbf{0}}\geq\kappa\ell/14\text{ or }G_{-r\mathbf{e}_{j},(r+8+2q),\mathbf{0}}\geq\kappa\ell/14\text{ or }G_{-r\mathbf{e}_{j},(r+10+2q),\mathbf{0}}\geq\kappa\ell/14\bigr{\}},
3={k,z𝒟k:|z|1dr,|z|1+8kdr+10,Gz,(k),𝟎κ/14},\displaystyle{\mathcal{E}}^{3}_{\ell}=\bigl{\{}\exists k\in\mathbb{N},z\in\mathcal{D}_{k}:|z|_{1}\leq dr,\ |z|_{1}+8\leq k\leq dr+10,\ G_{z,(k),\mathbf{0}}\geq\kappa\ell/14\bigr{\}},

and the event ,y4{\mathcal{E}}^{4}_{\ell,y} on which for all i[2d]i\in[2d]

(B.13) s=1m(¯1)1Gπi,s,(r+10+2q),πi,s+1+s=m1¯2Gπi,s,(r+8+2q),πi,s+1κ/14.\displaystyle\sum_{s=1}^{m\wedge(\underline{\ell}-1)-1}G_{\pi^{\prime}_{i,s},(r+10+2q),\pi^{\prime}_{i,s+1}}+\sum_{s=m\vee 1}^{\underline{\ell}-2}G_{\pi^{\prime}_{i,s},(r+8+2q),\pi^{\prime}_{i,s+1}}\geq\kappa\ell/14.

Then for 0\ell_{0} large enough to satisfy (B.9) and κ>0\kappa>0,

(B.14) {sup0supy𝒟ϵ|y|1ρ1G𝟎,(),y>κ}\displaystyle\Biggl{\{}\sup_{\ell\geq\ell_{0}}\sup_{\begin{subarray}{c}y\hskip 0.55pt\in\hskip 0.55pt\mathcal{D}_{\ell}\\ \epsilon\ell\leq|y|_{1}\leq\rho\ell\end{subarray}}\ell^{-1}G_{\mathbf{0},(\ell),y}>\kappa\Biggr{\}} (01)(0y¯rdϵ/2|y¯|12θy¯)\displaystyle\;\subset\;\biggl{(}\;\bigcup_{\ell\geq\ell_{0}}{\mathcal{E}}^{1}_{\ell}\biggr{)}\cup\biggl{(}\;\bigcup_{\ell\geq\ell_{0}}\bigcup_{\begin{subarray}{c}\underline{y}\hskip 0.55pt\in\hskip 0.55ptr\mathbb{Z}^{d}\\ \epsilon\ell/2\leq|\underline{y}|_{1}\leq\ell\end{subarray}}{\mathcal{E}}^{2}_{\ell}\circ\theta_{\underline{y}}\biggr{)}
(0ydϵ|y|13θy)(0y𝒟ϵ|y|1,y4).\displaystyle\qquad\cup\biggl{(}\;\bigcup_{\ell\geq\ell_{0}}\bigcup_{\begin{subarray}{c}y\hskip 0.55pt\in\hskip 0.55pt\mathbb{Z}^{d}\\ \epsilon\ell\leq|y|_{1}\leq\ell\end{subarray}}{\mathcal{E}}^{3}_{\ell}\circ\theta_{y}\biggr{)}\cup\biggl{(}\;\bigcup_{\ell\geq\ell_{0}}\bigcup_{\begin{subarray}{c}y\hskip 0.55pt\in\hskip 0.55pt\mathcal{D}_{\ell}\\ \epsilon\ell\leq|y|_{1}\leq\ell\end{subarray}}{\mathcal{E}}^{4}_{\ell,y}\biggr{)}.

Here is the explanation for the inclusion above.

  1. (i)

    Further down the proof we add auxiliary paths around the rr-steps of the path πi\pi^{\prime}_{i}. Because the first rr-steps share their initial point 𝟎\mathbf{0}, their auxiliary paths would intersect and independence would be lost. The same is true for the last rr-steps that share the endpoint y¯\underline{y}. Hence these special steps are handled separately.

    The event 1{\mathcal{E}}^{1}_{\ell} takes care of the first step of the path πi\pi^{\prime}_{i} which is either in the first sum on the right in (B.12), or in the second sum in case m=0m=0 and the first sum is empty.

    The event 2{\mathcal{E}}^{2}_{\ell} takes care of the last step to y¯\underline{y} which can come from any one of the three sums on the right in (B.12). We have to check that the possible endpoints fall within the range ϵ/2|πi,s|1\epsilon\ell/2\leq|\pi_{i,s}^{\prime}|_{1}\leq\ell of the union of shifts of 2{\mathcal{E}}^{2}_{\ell}: for i[2d]i\in[2d] and ¯s¯i\underline{\ell}\leq s\leq\underline{\ell}_{i},

    |πi,s|1|y|1(d+8)rϵ/2|\pi_{i,s}^{\prime}|_{1}\geq|y|_{1}-(d+8)r\geq\epsilon\ell/2

    and since πi,s\pi_{i,s}^{\prime} is on an admissible path of length \ell from 𝟎\mathbf{0} to yy, it must be that |πi,s|1|\pi_{i,s}^{\prime}|_{1}\leq\ell.

  2. (ii)

    The event 3{\mathcal{E}}^{3}_{\ell} takes care of the path segment from y¯\underline{y} to yy.

  3. (iii)

    On the complement of the first three unions on the right-hand side of (B.14) we have for each i[2d]i\in[2d],

    Gπi,0,(r+10+2q),πi,11{m1}+Gπi,0,(r+8+2q),πi,11{m=0}+Gπi,¯1,(r+10+2q),πi,¯1{m=¯}\displaystyle G_{\pi^{\prime}_{i,0},(r+10+2q),\pi^{\prime}_{i,1}}\mbox{\mymathbb{1}}\{m\geq 1\}+G_{\pi^{\prime}_{i,0},(r+8+2q),\pi^{\prime}_{i,1}}\mbox{\mymathbb{1}}\{m=0\}+G_{\pi^{\prime}_{i,\underline{\ell}-1},(r+10+2q),\pi^{\prime}_{i,\underline{\ell}}}\mbox{\mymathbb{1}}\{m=\underline{\ell}\}
    +Gπi,¯1,(r+8+2q),πi,¯1{m<¯}+s=¯¯i1Gπi,s,(r+8),πi,s+1+Gy¯,(i),y<13κ/14.\displaystyle\qquad\qquad+G_{\pi^{\prime}_{i,\underline{\ell}-1},(r+8+2q),\pi^{\prime}_{i,\underline{\ell}}}\mbox{\mymathbb{1}}\{m<\underline{\ell}\}+\sum_{s=\underline{\ell}}^{\underline{\ell}_{i}-1}G_{\pi^{\prime}_{i,s},(r+8),\pi^{\prime}_{i,s+1}}+G_{\underline{y},(\ell-\ell_{i}),y}<13\kappa\ell/14.

    Since ¯i¯8\underline{\ell}_{i}-\underline{\ell}\leq 8, the left-hand side has at most 13 terms, which explains the bound on the right. Thus, if in addition G𝟎,(),y>κG_{\mathbf{0},(\ell),y}>\kappa\ell, then event ,y4{\mathcal{E}}^{4}_{\ell,y} must occur.

By bounding the probabilities of the unions on the right of (B.14), we show next that for some fixed κ\kappa that does not depend on 0<ϵ<ρ<10<\epsilon<\rho<1,

(B.15) lim0{sup0supy𝒟ϵ|y|1ρ1G𝟎,(),y>κ}=0.\lim_{\ell_{0}\to\infty}\mathbb{P}\Biggl{\{}\sup_{\ell\geq\ell_{0}}\sup_{\begin{subarray}{c}y\hskip 0.55pt\in\hskip 0.55pt\mathcal{D}_{\ell}\\ \epsilon\ell\leq|y|_{1}\leq\rho\ell\end{subarray}}\ell^{-1}G_{\mathbf{0},(\ell),y}>\kappa\Biggr{\}}=0.

This will imply the conclusion (B.8) as we point out at the end of the proof.

By (B.2), (1)\mathbb{P}({\mathcal{E}}_{\ell}^{1}) is summable if (A.1) is satisfied with p=1p=1. Then (01)0\mathbb{P}\bigl{(}\,\bigcup_{\ell\geq\ell_{0}}{\mathcal{E}}_{\ell}^{1}\,\bigr{)}\to 0 as 0\ell_{0}\to\infty. Next, observe that

0ydϵ|y|13θyϵ0yd|y|1=3θy\bigcup_{\ell\geq\ell_{0}}\bigcup_{\begin{subarray}{c}y\hskip 0.55pt\in\hskip 0.55pt\mathbb{Z}^{d}\\ \epsilon\ell\leq|y|_{1}\leq\ell\end{subarray}}{\mathcal{E}}^{3}_{\ell}\circ\theta_{y}\subset\bigcup_{\ell\geq\epsilon\ell_{0}}\bigcup_{\begin{subarray}{c}y\hskip 0.55pt\in\hskip 0.55pt\mathbb{Z}^{d}\\ |y|_{1}=\ell\end{subarray}}{\mathcal{E}}^{3}_{\ell}\circ\theta_{y}

and hence

(0ydϵ|y|13θy)ϵ0(yd|y|1=3θy),\mathbb{P}\Bigl{(}\;\bigcup_{\ell\geq\ell_{0}}\bigcup_{\begin{subarray}{c}y\hskip 0.55pt\in\hskip 0.55pt\mathbb{Z}^{d}\\ \epsilon\ell\leq|y|_{1}\leq\ell\end{subarray}}{\mathcal{E}}^{3}_{\ell}\circ\theta_{y}\Bigr{)}\leq\sum_{\ell\geq\epsilon\ell_{0}}\mathbb{P}\Bigl{(}\bigcup_{\begin{subarray}{c}y\hskip 0.55pt\in\hskip 0.55pt\mathbb{Z}^{d}\\ |y|_{1}=\ell\end{subarray}}{\mathcal{E}}^{3}_{\ell}\circ\theta_{y}\Bigr{)},

which goes to 0 when 0\ell_{0}\to\infty if d1(3)\ell^{d-1}\mathbb{P}({\mathcal{E}}^{3}_{\ell}) is summable. This is the case if (A.1) is satisfied with p=dp=d. The union over 2θy¯{\mathcal{E}}^{2}_{\ell}\circ\theta_{\underline{y}} is controlled similarly.

𝟎\mathbf{0}y¯\underline{y}
Figure B.1. The light dashed grid is the coarse-grained lattice rdr\mathbb{Z}^{d}. The thin lines along this grid represent four πi\pi^{\prime}_{i}-paths from 𝟎\mathbf{0} to y¯\underline{y}. Three rr-steps on two πi\pi^{\prime}_{i}-paths are decorated with auxiliary paths represented by thick lines. The auxiliary paths are edge-disjoint as long as they associate (i) with different πi\pi^{\prime}_{i}-paths, (ii) with non-consecutive rr-steps on the same path πi\pi^{\prime}_{i}, or (iii) with rr-steps that are neither the first nor the last one of a πi\pi^{\prime}_{i}-path.

It remains to control the probability of the union of the events ,y4{\mathcal{E}}^{4}_{\ell,y} in (B.14). For i[2d]i\in[2d] and s[¯2]s\in[\underline{\ell}-2], for each segment [πi,s,πi,s+1][\pi^{\prime}_{i,s},\pi^{\prime}_{i,s+1}], bound both passage times Gπi,s,(r+10+2q),πi,s+1G_{\pi^{\prime}_{i,s},(r+10+2q),\pi^{\prime}_{i,s+1}} and Gπi,s,(r+8+2q),πi,s+1G_{\pi^{\prime}_{i,s},(r+8+2q),\pi^{\prime}_{i,s+1}} as was done in (B.3) by using 2d2d independent auxiliary paths of the appropriate lengths. For each segment [πi,s,πi,s+1][\pi^{\prime}_{i,s},\pi^{\prime}_{i,s+1}] add the two upper bounds and denote the result by Aπi,s,πi,s+1A_{\pi^{\prime}_{i,s},\pi^{\prime}_{i,s+1}}.

The terms for s=0s=0 and s¯1s\geq\underline{\ell}-1 were excluded from the events ,y4{\mathcal{E}}^{4}_{\ell,y} so that for distinct indices i[2d]i\in[2d] the 2d2d auxiliary paths constructed around the segments {[πi,s,πi,s+1]}s[¯2]\{[\pi^{\prime}_{i,s},\pi^{\prime}_{i,s+1}]\}_{s\in[\hskip 0.55pt\underline{\ell}-2]} stay separated. (We chose r5r\geq 5 at the outset to guarantee this separation.) Replace the edge weights t(e)t(e) with t+(e)=max(t(e),0)t^{+}(e)=\max(t(e),0) to ensure that the upper bounds are nonnegative. After these steps, the left-hand side of (B.13) is bounded above by s=1¯2Aπi,s,πi,s+1\sum_{s=1}^{\underline{\ell}-2}A_{\pi^{\prime}_{i,s},\pi^{\prime}_{i,s+1}}.

All the AA-terms have the same distribution as A𝟎,r𝐞1A_{\mathbf{0},r\mathbf{e}_{1}}. As explained above, over distinct indices i[2d]i\in[2d] the random vectors {Aπi,s,πi,s+1:s[¯2]}\{A_{\pi^{\prime}_{i,s},\pi^{\prime}_{i,s+1}}:s\in[\underline{\ell}-2]\} are independent. For any particular i[2d]i\in[2d], {Aπi,s,πi,s+1:s[¯2] even}\{A_{\pi^{\prime}_{i,s},\pi^{\prime}_{i,s+1}}:s\in[\underline{\ell}-2]\text{ even}\} are i.i.d. and {Aπi,s,πi,s+1:s[¯2] odd}\{A_{\pi^{\prime}_{i,s},\pi^{\prime}_{i,s+1}}:s\in[\underline{\ell}-2]\text{ odd}\} are i.i.d. because now we skip every other rr-step. See Figure B.1.

We derive the concluding estimate. Recall that

¯(ρ+dr)/r(ρr1+1).\underline{\ell}\leq(\rho\ell+dr)/r\leq(\rho r^{-1}+1)\ell.

Let c=(ρr1+1)/2c=\lceil{(\rho r^{-1}+1)/2}\rceil. Let SnS_{n} denote the sum of nn independent copies of A𝟎,r𝐞1A_{\mathbf{0},r\mathbf{e}_{1}}. Since the AA-terms are nonnegative we have

(s[¯2] evenAπi,s,πi,s+1κ/28)(Scκ/28).\mathbb{P}\Bigl{(}\sum_{s\in[\underline{\ell}-2]\text{ even}}A_{\pi^{\prime}_{i,s},\pi^{\prime}_{i,s+1}}\geq\kappa\ell/28\Bigr{)}\leq\mathbb{P}(S_{c\hskip 0.55pt\ell}\geq\kappa\ell/28).

The same holds for the sum over odd ss. Thus we have

(,y4)22d(Scκ/28)2d,\mathbb{P}({\mathcal{E}}_{\ell,y}^{4})\leq 2^{2d}\hskip 0.7pt\mathbb{P}(S_{c\hskip 0.55pt\ell}\geq\kappa\ell/28)^{2d},

Take κ>28c𝔼[A𝟎,r𝐞1]\kappa>28\hskip 0.55ptc\hskip 0.55pt\mathbb{E}[A_{\mathbf{0},r\mathbf{e}_{1}}] and use the fact that there are no more than (2+1)d(2\ell+1)^{d} points y𝒟y\in\mathcal{D}_{\ell} to get

(0y𝒟ϵ|y|1ρ,y4)\displaystyle\mathbb{P}\biggl{(}\;\bigcup_{\ell\geq\ell_{0}}\bigcup_{\begin{subarray}{c}y\hskip 0.55pt\in\hskip 0.55pt\mathcal{D}_{\ell}\\ \epsilon\ell\leq|y|_{1}\leq\rho\ell\end{subarray}}{\mathcal{E}}^{4}_{\ell,y}\biggr{)} 0(2+1)d(,y4)0(8+4)d(Scκ/28)2d\displaystyle\leq\sum_{\ell\geq\ell_{0}}(2\ell+1)^{d}\mathbb{P}({\mathcal{E}}^{4}_{\ell,y})\leq\sum_{\ell\geq\ell_{0}}(8\ell+4)^{d}\mathbb{P}(S_{c\hskip 0.55pt\ell}\geq\kappa\ell/28)^{2d}
0(8+4)dc2dVar(A𝟎,r𝐞1)2d(κ/28c𝔼[A𝟎,r𝐞1])4d2d.\displaystyle\leq\sum_{\ell\geq\ell_{0}}\frac{(8\ell+4)^{d}c^{2d}\operatorname{Var}(A_{\mathbf{0},r\mathbf{e}_{1}})^{2d}}{(\kappa/28-c\mathbb{E}[A_{\mathbf{0},r\mathbf{e}_{1}}])^{4d}\ell^{2d}}\,.

The bound (B.4) can be utilized to show that each G𝟎,(),xG_{\mathbf{0},(\ell),x} and thereby A𝟎,r𝐞1A_{\mathbf{0},r\mathbf{e}_{1}} is square-integrable if (A.1) holds with p=2p=2. The above then converges to 0 as 0\ell_{0}\to\infty. We have verified (B.15). The claim of the lemma follows:

{00:supy𝒟ϵ|y|1ρ1G𝟎,(),y>κ}=lim0{0:supy𝒟ϵ|y|1ρ1G𝟎,(),y>κ}=0.\mathbb{P}\Biggl{\{}\forall\ell_{0}\ \exists\ell\geq\ell_{0}:\sup_{\begin{subarray}{c}y\hskip 0.55pt\in\hskip 0.55pt\mathcal{D}_{\ell}\\ \epsilon\ell\leq|y|_{1}\leq\rho\ell\end{subarray}}\ell^{-1}G_{\mathbf{0},(\ell),y}>\kappa\Biggr{\}}=\lim_{\ell_{0}\to\infty}\mathbb{P}\Biggl{\{}\exists\ell\geq\ell_{0}:\sup_{\begin{subarray}{c}y\hskip 0.55pt\in\hskip 0.55pt\mathcal{D}_{\ell}\\ \epsilon\ell\leq|y|_{1}\leq\rho\ell\end{subarray}}\ell^{-1}G_{\mathbf{0},(\ell),y}>\kappa\Biggr{\}}=0.\qed
Lemma B.9.

Assume (A.1) with p=dp=d. Then for any 0<ϵ<ρ<10<\epsilon<\rho<1 there exists a deterministic constant κ(0,)\kappa\in(0,\infty) such that, with probability one for each xdx\in\mathbb{Z}^{d}, there exists a strictly increasing random sequence {m(n)}n\{m(n)\}_{n\hskip 0.55pt\in\hskip 0.55pt\mathbb{N}}\subset\mathbb{N} such that m(n+1)/m(n)1m(n+1)/m(n)\to 1 and for {𝚎𝚖𝚙𝚝𝚢,o}\diamond\in\{\langle{\tt empty}\rangle,o\} and \ell\in\mathbb{N}

(B.16) Gm(n)x,(),zκzm(n)x+𝒟 such that ϵ|zm(n)x|1ρ.\displaystyle\begin{split}&G^{\,\diamond}_{m(n)x,(\ell),z}\leq\kappa\ell\quad\forall z\in m(n)x+\mathcal{D}^{\,\diamond}_{\ell}\text{ such that }\epsilon\ell\leq|z-m(n)x|_{1}\leq\rho\ell.\end{split}
Proof.

If x=𝟎x=\mathbf{0} take m(n)=0+nm(n)=\ell_{0}+n from Lemma B.8. Next suppose x𝟎x\neq\mathbf{0}. Fix ϵ<ρ\epsilon<\rho in (0,1)(0,1). Apply Lemma B.8 to choose a finite constant κ\kappa such that

(){{𝚎𝚖𝚙𝚝𝚢,o}:sup,y𝒟ϵ|y|1ρ1G𝟎,(),yκ}>0.\mathbb{P}({\mathcal{E}})\equiv\mathbb{P}\Biggl{\{}\forall\diamond\in\{\langle{\tt empty}\rangle,o\}:\ \sup_{\begin{subarray}{c}\ell\in\mathbb{N},y\hskip 0.55pt\in\hskip 0.55pt\mathcal{D}^{\,\diamond}_{\ell}\\ \epsilon\ell\leq|y|_{1}\leq\rho\ell\end{subarray}}\ell^{-1}G^{\,\diamond}_{\mathbf{0},(\ell),y}\leq\kappa\Biggr{\}}>0.

The ergodic theorem implies that with probability one, for each xd{𝟎}x\in\mathbb{Z}^{d}\setminus\{\mathbf{0}\} there exist infinitely many mm\in\mathbb{N} such that

{𝚎𝚖𝚙𝚝𝚢,o}:sup,ymx+𝒟ϵ|ymx|1ρ1Gmx,(),yκ.\forall\diamond\in\{\langle{\tt empty}\rangle,o\}:\sup_{\begin{subarray}{c}\ell\in\mathbb{N},y\hskip 0.55pt\in\hskip 0.55ptmx+\mathcal{D}^{\,\diamond}_{\ell}\\ \epsilon\ell\leq|y-mx|_{1}\leq\rho\ell\end{subarray}}\ell^{-1}G^{\,\diamond}_{mx,(\ell),y}\leq\kappa.

Enumerate these mm’s as a strictly increasing sequence {m(n):n}\{m(n):n\in\mathbb{N}\}. Then for \mathbb{P}-almost every ω\omega

limnnm(n)=limn1m(n)k=1m(n)1{θkxω}=()>0.\lim_{n\to\infty}\frac{n}{m(n)}=\lim_{n\to\infty}\frac{1}{m(n)}\sum_{k=1}^{m(n)}\mbox{\mymathbb{1}}\{\theta_{kx}\omega\in{\mathcal{E}}\}=\mathbb{P}({\mathcal{E}})>0.

Consequently, m(n+1)/m(n)m(n+1)/m(n) converges to 11. ∎

We are ready for the shape theorem.

Theorem B.10.

Assume r0>r_{0}>-\infty and (A.1) with p=dp=d. Fix {𝚎𝚖𝚙𝚝𝚢,o}\diamond\in\{\langle{\tt empty}\rangle,o\}. Let 𝒱\mathcal{V} be a closed subset of int𝒰\operatorname{int}\mathcal{U}. The following holds with probability one:

(B.17) limmaxx𝒟:x/𝒱1|G𝟎,(),xg(x/)|=0.\displaystyle\lim_{\ell\to\infty}\;\max_{x\hskip 0.55pt\in\hskip 0.55pt\mathcal{D}^{\,\diamond}_{\ell}:\,x/\ell\hskip 0.55pt\in\hskip 0.55pt\mathcal{V}}\;\ell^{\hskip 0.55pt-1}|G^{\,\diamond}_{\mathbf{0},(\ell),x}-\ell g^{\diamond}(x/\ell)|=0.
Proof.

The proof follows steps similar to those of (A.3). We treat the case =o\diamond=o, the other case being a simpler version. Let Ω0\Omega_{0} be the full probability event that consists of intersecting the event on which (B.5) holds for all ζdint𝒰\zeta\in\mathbb{Q}^{d}\cap\operatorname{int}\mathcal{U} with the event in (B.8) and the events in Lemma B.9 for all rational ϵ<ρ\epsilon<\rho in (0,1)(0,1). Fix ωΩ0\omega\in\Omega_{0}. We show that for this ω\omega

(B.18) lim¯minx𝒟:x/𝒱1(G𝟎,(),xogo(x/))0and\displaystyle\varliminf_{\ell\to\infty}\;\min_{x\hskip 0.55pt\in\hskip 0.55pt\mathcal{D}^{\,\diamond}_{\ell}:\,x/\ell\hskip 0.55pt\in\hskip 0.55pt\mathcal{V}}\;\ell^{-1}\bigl{(}G^{o}_{\mathbf{0},(\ell),x}-\ell g^{o}(x/\ell)\bigr{)}\geq 0\quad\text{and}
(B.19) lim¯maxx𝒟:x/𝒱1(G𝟎,(),xogo(x/))0.\displaystyle\varlimsup_{\ell\to\infty}\;\max_{x\hskip 0.55pt\in\hskip 0.55pt\mathcal{D}^{\,\diamond}_{\ell}:\,x/\ell\hskip 0.55pt\in\hskip 0.55pt\mathcal{V}}\;\ell^{-1}\bigl{(}G^{o}_{\mathbf{0},(\ell),x}-\ell g^{o}(x/\ell)\bigr{)}\leq 0.

Proof of (B.18). Fix (ω\omega-dependent) sequences k\ell_{k}\to\infty and xk𝒟kox_{k}\in\mathcal{D}^{o}_{\ell_{k}} that realize the lim¯\varliminf on the left-hand side of (B.18). Since xk𝒟kox_{k}\in\mathcal{D}^{o}_{\ell_{k}} there are coefficients ai,k±+a^{\pm}_{i,k}\in\mathbb{Z}_{+} such that

(B.20) xk=i=1d(ai,k+ai,k)𝐞iandi=1d(ai,k++ai,k)k.\displaystyle x_{k}=\sum_{i=1}^{d}(a_{i,k}^{+}-a^{-}_{i,k})\mathbf{e}_{i}\quad\text{and}\quad\sum_{i=1}^{d}(a_{i,k}^{+}+a_{i,k}^{-})\leq\ell_{k}.

Pass to subsequences, still denoted by k\ell_{k} and xkx_{k}, such that

(B.21) ai,k±/kkαi±[0,1]withi=1d(αi++αi)1.\displaystyle a_{i,k}^{\pm}/\ell_{k}\mathop{\longrightarrow}_{k\to\infty}\alpha_{i}^{\pm}\in[0,1]\quad\text{with}\quad\sum_{i=1}^{d}(\alpha_{i}^{+}+\alpha_{i}^{-})\leq 1.

Let ξ=i=1d(αi+αi)𝐞i=limkxk/k𝒱int𝒰\xi=\sum_{i=1}^{d}(\alpha_{i}^{+}-\alpha_{i}^{-})\mathbf{e}_{i}=\lim_{k\to\infty}x_{k}/\ell_{k}\in\mathcal{V}\subset\operatorname{int}\mathcal{U}. We approximate ξ\xi with a rational point ζ\zeta to which we can apply (B.5). Bound (B.18) comes by building a path from xkx_{k} to a multiple of ζ\zeta and by the subadditivity of passage times. Here are the details.

First, we dispose of the case where there are infinitely many kk for which ai,k+=ai,k=0a_{i,k}^{+}=a_{i,k}^{-}=0 for all i[d]i\in[d]. If this is the case, then going along a further subsequence we can assume that xk=𝟎x_{k}=\mathbf{0} for all kk. Applying (B.5) with ζ=𝟎\zeta=\mathbf{0} gives k1G𝟎,(k),xkogo(𝟎)\ell_{k}^{-1}G^{o}_{\mathbf{0},(\ell_{k}),x_{k}}\to g^{o}(\mathbf{0}) and since go(xk/k)=go(𝟎)g^{o}(x_{k}/\ell_{k})=g^{o}(\mathbf{0}) for all kk we see that the lim¯\varliminf on the left-hand side of (B.18) is 0. We can therefore assume that for each kk there exists some i[d]i\in[d] such that ai,k+1a_{i,k}^{+}\geq 1 or ai,k1a_{i,k}^{-}\geq 1. Consequently, if we let \mathcal{I} denote the set of indices i[d]i\in[d] for which ai,k+1a_{i,k}^{+}\geq 1 or ai,k1a_{i,k}^{-}\geq 1 for infinitely many kk, then \mathcal{I}\neq\varnothing.

Let

(B.22) γ=min{αi:αi>0,i[d]}min{αi+:αi+>0,i[d]}>0,\displaystyle\gamma=\min\{\alpha_{i}^{-}:\alpha_{i}^{-}>0,i\in[d]\}\wedge\min\{\alpha_{i}^{+}:\alpha_{i}^{+}>0,i\in[d]\}>0,

with the convention that min=\min\varnothing=\infty, which takes care of the case αi±=0\alpha_{i}^{\pm}=0 for all i[d]i\in[d]. Let δ\delta be a rational in (0,(γ1)/(4d))\bigl{(}0,(\gamma\wedge 1)/(4d)\bigr{)}. For i[d]i\in[d]\setminus\mathcal{I} let βi+=βi=0\beta_{i}^{+}=\beta_{i}^{-}=0 and note that we also have αi+=αi=0\alpha_{i}^{+}=\alpha_{i}^{-}=0. For ii\in\mathcal{I} take βi±[δ,1]\beta_{i}^{\pm}\in[\delta,1]\cap\mathbb{Q} such that |αi±βi±|2dδ|\alpha_{i}^{\pm}-\beta_{i}^{\pm}|\leq 2d\delta,

i=1d(βi++βi)1,andj:(1+5dγ1)(βj+βj)αj+αj.\sum_{i=1}^{d}(\beta_{i}^{+}+\beta_{i}^{-})\leq 1,\quad\text{and}\quad\forall j\in\mathcal{I}:(1+5d\gamma^{-1})(\beta_{j}^{+}-\beta_{j}^{-})\neq\alpha_{j}^{+}-\alpha_{j}^{-}.

Let ζ=i=1d(βi+βi)𝐞i\zeta=\sum_{i=1}^{d}(\beta_{i}^{+}-\beta_{i}^{-})\mathbf{e}_{i} and take δ>0\delta>0 small enough so that ζint𝒰\zeta\in\operatorname{int}\mathcal{U}. We will eventually take δ0\delta\to 0, which sends ζξ\zeta\to\xi.

We have for all ii\in\mathcal{I}

(B.23) (1+5dδγ1)βi+αi+δand(1+5dδγ1)βiαiδ.\displaystyle(1+5d\delta\gamma^{-1})\beta_{i}^{+}-\alpha_{i}^{+}\geq\delta\quad\text{and}\quad(1+5d\delta\gamma^{-1})\beta_{i}^{-}-\alpha_{i}^{-}\geq\delta.

To see this, note that when αi+>0\alpha_{i}^{+}>0 we have

(1+5dδγ1)βi+αi+(1+5dδγ1)(αi+2dδ)αi+dδ/2δ(1+5d\delta\gamma^{-1})\beta_{i}^{+}-\alpha_{i}^{+}\geq(1+5d\delta\gamma^{-1})(\alpha_{i}^{+}-2d\delta)-\alpha_{i}^{+}\geq d\delta/2\geq\delta

and when αi+=0\alpha_{i}^{+}=0 (but ii\in\mathcal{I}) we have

(1+5dδγ1)βi+αi+=(1+5dδγ1)βi+βi+δ.(1+5d\delta\gamma^{-1})\beta_{i}^{+}-\alpha_{i}^{+}=(1+5d\delta\gamma^{-1})\beta_{i}^{+}\geq\beta^{+}_{i}\geq\delta.

The same holds with superscript -.

Let

ζ=i((1+5dδγ1)(βi+βi)(αi+αi))𝐞ii((1+5dδγ1)(βi++βi)(αi++αi)).\zeta^{\prime}=\frac{\sum_{i\in\mathcal{I}}\Bigl{(}(1+5d\delta\gamma^{-1})(\beta_{i}^{+}-\beta_{i}^{-})-(\alpha_{i}^{+}-\alpha_{i}^{-})\Bigr{)}\mathbf{e}_{i}}{\sum_{i\in\mathcal{I}}\Bigl{(}(1+5d\delta\gamma^{-1})(\beta_{i}^{+}+\beta_{i}^{-})-(\alpha_{i}^{+}+\alpha_{i}^{-})\Bigr{)}}.

The choice of βi±\beta_{i}^{\pm} guarantees that ζ𝟎\zeta^{\prime}\neq\mathbf{0}. Furthermore, (B.23) shows that ζ\zeta^{\prime} is a convex combination of the vectors {±𝐞i:i}\{\pm\mathbf{e}_{i}:i\in\mathcal{I}\} with all strictly positive coefficients. Consequently, ζint𝒰\zeta^{\prime}\in\operatorname{int}\mathcal{U}.

Take rational ϵ<ρ\epsilon<\rho in (0,1)(0,1) such that ϵ<|ζ|1<ρ\epsilon<|\zeta^{\prime}|_{1}<\rho. Let \ell\in\mathbb{N} be such that βi+,βi\ell\beta_{i}^{+},\ell\beta_{i}^{-}\in\mathbb{N} for ii\in\mathcal{I} and take n¯k\bar{n}_{k} such that

m(n¯k1)(1+5dδγ1)k/m(n¯k),m(\bar{n}_{k}-1)\leq(1+5d\delta\gamma^{-1})\ell_{k}/\ell\leq m(\bar{n}_{k}),

for the sequence m(n)m(n) in Lemma B.9 corresponding to the above choice of ϵ\epsilon and ρ\rho and to x=ζx=\ell\zeta. Abbreviate m¯k=m(n¯k)\overline{m}_{k}=m(\bar{n}_{k}). Using (B.23) we have for ii\in\mathcal{I}

(B.24) limkk1(m¯kβi+ai,k+)=(1+5dδγ1)βi+αi+δ.\displaystyle\lim_{k\to\infty}\ell_{k}^{-1}(\overline{m}_{k}\ell\beta_{i}^{+}-a_{i,k}^{+})=(1+5d\delta\gamma^{-1})\beta_{i}^{+}-\alpha_{i}^{+}\geq\delta.

The same holds with superscript -. Thus, for all ii\in\mathcal{I} and for large kk

(B.25) m¯kβi±ai,k±+δk/2.\displaystyle\overline{m}_{k}\ell\beta_{i}^{\pm}\geq a_{i,k}^{\pm}+\delta\ell_{k}/2.

This implies that when kk is large, m¯kζ\overline{m}_{k}\ell\zeta (which belongs to d\mathbb{Z}^{d}) is accessible from xkx_{k} by an \mathcal{R}-admissible path of length

(B.26) j¯k=i=1d(m¯kβi+ai,k+)+i=1d(m¯kβiai,k).\displaystyle\overline{j}_{k}=\sum_{i=1}^{d}(\overline{m}_{k}\ell\beta_{i}^{+}-a_{i,k}^{+})+\sum_{i=1}^{d}(\overline{m}_{k}\ell\beta_{i}^{-}-a_{i,k}^{-}).

Note that

(B.27) limkj¯k/k=i=1d((1+5dδγ1)βi+αi+)+i=1d((1+5dδγ1)βiαi)(4d+5γ1)dδ.\displaystyle\lim_{k\to\infty}\overline{j}_{k}/\ell_{k}=\sum_{i=1}^{d}((1+5d\delta\gamma^{-1})\beta_{i}^{+}-\alpha_{i}^{+})+\sum_{i=1}^{d}((1+5d\delta\gamma^{-1})\beta_{i}^{-}-\alpha_{i}^{-})\leq(4d+5\gamma^{-1})d\delta.

The first equality and (B.24) imply that

limkm¯kζxkj¯k=ζ\lim_{k\to\infty}\frac{\overline{m}_{k}\ell\zeta-x_{k}}{\overline{j}_{k}}=\zeta^{\prime}

and therefore ϵj¯k|m¯kζxk|1ρj¯k\epsilon\overline{j}_{k}\leq|\overline{m}_{k}\ell\zeta-x_{k}|_{1}\leq\rho\overline{j}_{k} for kk large enough. This will allow us to apply (B.16).

Since xkx_{k} is accessible from 𝟎\mathbf{0} by an \mathcal{R}-admissible path of length i=1d(ai,k++ai,k)k\sum_{i=1}^{d}(a_{i,k}^{+}+a_{i,k}^{-})\leq\ell_{k}, concatenating this path and the one from xkx_{k} to m¯kζ\overline{m}_{k}\ell\zeta gives an \mathcal{R}-admissible path from 𝟎\mathbf{0} to m¯kζ\overline{m}_{k}\ell\zeta of length

i=1dm¯k(βi++βi)m¯k.\sum_{i=1}^{d}\overline{m}_{k}\ell(\beta_{i}^{+}+\beta_{i}^{-})\leq\overline{m}_{k}\ell.

Hence m¯kζ𝒟m¯ko\overline{m}_{k}\ell\zeta\in\mathcal{D}^{o}_{\overline{m}_{k}\ell}. Subadditivity now gives

G𝟎,(m¯k),m¯kζoG𝟎,(k),xko+Gxk,(j¯k),m¯kζo.\displaystyle G^{o}_{\mathbf{0},(\overline{m}_{k}\ell),\overline{m}_{k}\ell\zeta}\leq G^{o}_{\mathbf{0},(\ell_{k}),x_{k}}+G^{o}_{x_{k},(\overline{j}_{k}),\overline{m}_{k}\ell\zeta}.

Using this, (B.16), and (B.27), we get

(1+5dδγ1)go(ζ)=limkG𝟎,(m¯k),m¯kζolim¯kG𝟎,(k),xkok+κ(4d+5γ1)dδ.(1+5d\delta\gamma^{-1})g^{o}(\zeta)=\lim_{k\to\infty}G^{o}_{\mathbf{0},(\overline{m}_{k}\ell),\overline{m}_{k}\ell\zeta}\leq\varliminf_{k\to\infty}\frac{G^{o}_{\mathbf{0},(\ell_{k}),x_{k}}}{\ell_{k}}+\kappa(4d+5\gamma^{-1})d\delta.

Taking δ0\delta\to 0 and the continuity of gog^{o} on int𝒰\operatorname{int}\mathcal{U} gives

go(ξ)lim¯kG𝟎,(k),xkok.g^{o}(\xi)\leq\varliminf_{k\to\infty}\frac{G^{o}_{\mathbf{0},(\ell_{k}),x_{k}}}{\ell_{k}}.

Since xk/k𝒱int𝒰x_{k}/\ell_{k}\in\mathcal{V}\subset\operatorname{int}\mathcal{U} and xk/kξx_{k}/\ell_{k}\to\xi, using again the continuity of gog^{o} on int𝒰\operatorname{int}\mathcal{U} completes the proof of (B.18):

lim¯kk1(G𝟎,(k),xkokgo(xk/k))0.\displaystyle\varliminf_{k\to\infty}\ell_{k}^{\hskip 0.55pt-1}\bigl{(}G^{o}_{\mathbf{0},(\ell_{k}),x_{k}}-\ell_{k}g^{o}(x_{k}/\ell_{k})\bigr{)}\geq 0.

Proof of (B.19). Proceed similarly to the proof of (B.18), but with the sequences k\ell_{k}\to\infty and xk𝒟kox_{k}\in\mathcal{D}^{o}_{\ell_{k}} realizing the lim¯\varlimsup on the left-hand side of (B.19). Again, we have the representation (B.20), the limits (B.21), and ξ=i[d](αi+αi)𝐞i𝒱\xi=\sum_{i\in[d]}(\alpha_{i}^{+}-\alpha_{i}^{-})\mathbf{e}_{i}\in\mathcal{V}.

We start by treating the case when ξ=𝟎\xi=\mathbf{0}. In this case let jk=2|xk|1j_{k}=2|x_{k}|_{1} or jk=2|xk|1+1j_{k}=2|x_{k}|_{1}+1, so that kjk\ell_{k}-j_{k} is even. Observe that jk/k0j_{k}/\ell_{k}\to 0 and hence kjk\ell_{k}\geq j_{k} for kk large. Thus, one can make an admissible loop of length kjk\ell_{k}-j_{k} from 𝟎\mathbf{0} back to 𝟎\mathbf{0} and then take a path of length jkj_{k} from 𝟎\mathbf{0} to xkx_{k}. From (B.5) we have k1G𝟎,(kjk),𝟎ogo(𝟎)\ell_{k}^{-1}G^{o}_{\mathbf{0},(\ell_{k}-j_{k}),\mathbf{0}}\to g^{o}(\mathbf{0}). If jkj_{k} is bounded then so is |xk|1|x_{k}|_{1} and we have k1G𝟎,(jk),xko0\ell_{k}^{-1}G^{o}_{\mathbf{0},(j_{k}),x_{k}}\to 0. On the other hand, if jkj_{k}\to\infty along some subsequence, then along this subsequence, and for kk large, we have jk/3|xk|12jk/3j_{k}/3\leq|x_{k}|_{1}\leq 2j_{k}/3 and, applying (B.8), we then get

G𝟎,(k),xkoG𝟎,(kjk),𝟎o+G𝟎,(jk),xkoG𝟎,(kjk),𝟎o+κjk,G^{o}_{\mathbf{0},(\ell_{k}),x_{k}}\leq G^{o}_{\mathbf{0},(\ell_{k}-j_{k}),\mathbf{0}}+G^{o}_{\mathbf{0},(j_{k}),x_{k}}\leq G^{o}_{\mathbf{0},(\ell_{k}-j_{k}),\mathbf{0}}+\kappa j_{k},

for kk large enough. Dividing by k\ell_{k} and taking kk\to\infty we deduce that

lim¯kk1G𝟎,(k),xkogo(𝟎).\varlimsup_{k\to\infty}\ell_{k}^{-1}G^{o}_{\mathbf{0},(\ell_{k}),x_{k}}\leq g^{o}(\mathbf{0}).

The continuity of gog^{o} at 𝟎\mathbf{0} implies then that the lim¯\varlimsup on the left-hand side of (B.19) is 0. For the rest of the proof we can and will assume that ξ𝟎\xi\neq\mathbf{0}.

Define γ(0,)\gamma\in(0,\infty) as in (B.22). Let δ\delta be a rational in (0,γ/2)(0,\gamma/2). Choose βi±\beta_{i}^{\pm}, i[d]i\in[d], so that for \scaleobj0.9{,+}{\scaleobj{0.9}{\boxempty}}\in\{-,+\}, when αi\scaleobj0.8=0\alpha_{i}^{\scaleobj{0.8}{\boxempty}}=0 we have βi\scaleobj0.8=0\beta_{i}^{\scaleobj{0.8}{\boxempty}}=0 and when αi\scaleobj0.8>0\alpha_{i}^{\scaleobj{0.8}{\boxempty}}>0 we have βi\scaleobj0.8[δ,1]\beta_{i}^{\scaleobj{0.8}{\boxempty}}\in[\delta,1]\cap\mathbb{Q} such that |αi\scaleobj0.8βi\scaleobj0.8|δ|\alpha_{i}^{\scaleobj{0.8}{\boxempty}}-\beta_{i}^{\scaleobj{0.8}{\boxempty}}|\leq\delta and overall we have

i=1d(βi++βi)1and(12δγ1)i[d](βi+βi)𝐞ii[d](αi+αi)𝐞i.\sum_{i=1}^{d}(\beta_{i}^{+}+\beta_{i}^{-})\leq 1\quad\text{and}\quad(1-2\delta\gamma^{-1})\sum_{i\in[d]}(\beta_{i}^{+}-\beta_{i}^{-})\mathbf{e}_{i}\neq\sum_{i\in[d]}(\alpha_{i}^{+}-\alpha_{i}^{-})\mathbf{e}_{i}.

This is possible since ξ𝟎\xi\neq\mathbf{0} and therefore αi\scaleobj0.8>0\alpha_{i}^{\scaleobj{0.8}{\boxempty}}>0 for some i[d]i\in[d] and \scaleobj0.9{,+}{\scaleobj{0.9}{\boxempty}}\in\{-,+\}. Let ζ=i=1d(βi+βi)𝐞i\zeta=\sum_{i=1}^{d}(\beta_{i}^{+}-\beta_{i}^{-})\mathbf{e}_{i} and choose δ\delta small enough so that ζint𝒰\zeta\in\operatorname{int}\mathcal{U}. Note that

αi\scaleobj0.8(12δγ1)βi\scaleobj0.80for all i[d] and \scaleobj0.8{,+}.\alpha_{i}^{\scaleobj{0.8}{\boxempty}}-(1-2\delta\gamma^{-1})\beta_{i}^{\scaleobj{0.8}{\boxempty}}\geq 0\quad\text{for all }i\in[d]\text{ and }{\scaleobj{0.8}{\boxempty}}\in\{-,+\}.

Indeed, this clearly holds when αi\scaleobj0.8=0\alpha_{i}^{\scaleobj{0.8}{\boxempty}}=0 and when αi\scaleobj0.8>0\alpha_{i}^{\scaleobj{0.8}{\boxempty}}>0 we have

αi\scaleobj0.8(12δγ1)βi\scaleobj0.8αi\scaleobj0.8(12δγ1)(αi\scaleobj0.8+δ)δ.\alpha_{i}^{\scaleobj{0.8}{\boxempty}}-(1-2\delta\gamma^{-1})\beta_{i}^{\scaleobj{0.8}{\boxempty}}\geq\alpha_{i}^{\scaleobj{0.8}{\boxempty}}-(1-2\delta\gamma^{-1})(\alpha_{i}^{\scaleobj{0.8}{\boxempty}}+\delta)\geq\delta.

The above two observations imply that

ζ=i[d]((αi+αi)(12δγ1)(βi+βi))𝐞iδ+i[d]((αi++αi)(12δγ1)(βi++βi))int𝒰{𝟎}.\zeta^{\prime}=\frac{\sum_{i\in[d]}\Bigl{(}(\alpha_{i}^{+}-\alpha_{i}^{-})-(1-2\delta\gamma^{-1})(\beta_{i}^{+}-\beta_{i}^{-})\Bigr{)}\mathbf{e}_{i}}{\delta+\sum_{i\in[d]}\Bigl{(}(\alpha_{i}^{+}+\alpha_{i}^{-})-(1-2\delta\gamma^{-1})(\beta_{i}^{+}+\beta_{i}^{-})\Bigr{)}}\in\operatorname{int}\mathcal{U}\setminus\{\mathbf{0}\}.

We can then find rational ϵ<ρ\epsilon<\rho in (0,1)(0,1) such that ϵ<|ζ|1<ρ\epsilon<|\zeta^{\prime}|_{1}<\rho.

Let \ell\in\mathbb{N} be such that βi+,βi+\ell\beta_{i}^{+},\ell\beta_{i}^{-}\in\mathbb{Z}_{+} for i[d]i\in[d] and take n¯k\underline{n}_{k} such that

m(n¯k)(12δγ1)k/m(n¯k+1),m(\underline{n}_{k})\leq(1-2\delta\gamma^{-1})\ell_{k}/\ell\leq m(\underline{n}_{k}+1),

for the sequence m(n)m(n) in Lemma B.9 corresponding to x=ζx=\ell\zeta and to the above choice of ϵ\epsilon and ρ\rho. Abbreviate m¯k=m(n¯k)\underline{m}_{k}=m(\underline{n}_{k}) and observe that if αi+>0\alpha_{i}^{+}>0 then

(B.28) limkk1(ai,k+m¯kβi+)=αi+(12δγ1)βi+δ.\displaystyle\lim_{k\to\infty}\ell_{k}^{-1}(a_{i,k}^{+}-\underline{m}_{k}\ell\beta_{i}^{+})=\alpha_{i}^{+}-(1-2\delta\gamma^{-1})\beta_{i}^{+}\geq\delta.

Then for large kk

(B.29) ai,k+m¯kβi+0\displaystyle a_{i,k}^{+}-\underline{m}_{k}\ell\beta_{i}^{+}\geq 0

This inequality is trivial if αi+=βi+=0\alpha_{i}^{+}=\beta_{i}^{+}=0. The same computation works with minus sign superscripts. This implies that xkx_{k} is accessible from m¯kζ\underline{m}_{k}\ell\zeta in

j¯k=i=1d(ai,k+m¯kβi+)+i=1d(ai,km¯kβi)\underline{j}_{k}=\sum_{i=1}^{d}(a_{i,k}^{+}-\underline{m}_{k}\ell\beta_{i}^{+})+\sum_{i=1}^{d}(a_{i,k}^{-}-\underline{m}_{k}\ell\beta_{i}^{-})

\mathcal{R}-steps and δk\lfloor{\delta\ell_{k}}\rfloor 𝟎\mathbf{0}-steps. Note that

limkj¯k/k=i=1d(αi+(12δγ1)βi+)+i=1d(αi(12δγ1)βi)(2d+2γ1)δ.\lim_{k\to\infty}\underline{j}_{k}/\ell_{k}=\sum_{i=1}^{d}(\alpha_{i}^{+}-(1-2\delta\gamma^{-1})\beta_{i}^{+})+\sum_{i=1}^{d}(\alpha_{i}^{-}-(1-2\delta\gamma^{-1})\beta_{i}^{-})\leq(2d+2\gamma^{-1})\delta.

As a consequence,

limkxkm¯kζδk+j¯k=ζ\lim_{k\to\infty}\frac{x_{k}-\underline{m}_{k}\ell\zeta}{\lfloor{\delta\ell_{k}}\rfloor+\underline{j}_{k}}=\zeta^{\prime}

and one can then apply (B.16). Then, as in the proof of (B.18), using subadditivity then taking kk\to\infty and then δ0\delta\to 0 and using the continuity of gog^{o} on int𝒰\operatorname{int}\mathcal{U} gives

lim¯kG𝟎,(k),xkokgo(ξ).\varlimsup_{k\to\infty}\frac{G^{o}_{\mathbf{0},(\ell_{k}),x_{k}}}{\ell_{k}}\leq g^{o}(\xi).

Another use of the continuity of gog^{o} completes the proof of (B.19). ∎

Proof of Theorem B.1.

Apply Theorem B.10 with 𝒱={ξ𝒰:|ξ|11/(1+α)}\mathcal{V}=\{\xi\in\mathcal{U}:|\xi|_{1}\leq 1/(1+\alpha)\}. ∎

Appendix C Peierls argument

This appendix follows the ideas of [11, 20]. First we prove a general estimate and then specialize it to prove Lemma 5.3. Let dd\in\mathbb{N}. Tile d\mathbb{Z}^{d} by NN-cubes S(𝐤)=N𝐤+[0,N)dS(\mathbf{k})=N\mathbf{k}+[0,N)^{d} indexed by 𝐤d\mathbf{k}\in\mathbb{Z}^{d}. Each NN-cube S(𝐤)S(\mathbf{k}) is colored randomly black or white in a shift-stationary manner. Let p=p(N)p=p(N) be the marginal probability that a particular cube is black and assume that

(C.1) p(N)1asN.p(N)\to 1\quad\text{as}\quad N\to\infty.

Assume finite range dependence: there is a strictly positive integer constant a0a_{0} such that

(C.2) 𝐮d\forall\mathbf{u}\in\mathbb{Z}^{d}, the colors of the cubes {S(𝐤):𝐤𝐮+a0d}\{S(\mathbf{k}):\mathbf{k}\in\mathbf{u}+a_{0}\mathbb{Z}^{d}\} are i.i.d.

There are K0=a0dK_{0}=a_{0}^{d} distinct i.i.d. collections, indexed by 𝐮{0,1,,a01}d\mathbf{u}\in\{0,1,\dotsc,a_{0}-1\}^{d}.

It may be desirable to let the separation of the cubes be a parameter. For a positive integer a1a_{1} and 𝐮{0,1,,a11}d\mathbf{u}\in\{0,1,\dotsc,a_{1}-1\}^{d}, define the collection 𝒮a1,𝐮={S(𝐤):𝐤𝐮+a1d}\mathcal{S}_{a_{1},\mathbf{u}}=\{S(\mathbf{k}):\mathbf{k}\in\mathbf{u}+a_{1}\mathbb{Z}^{d}\} of cubes with lower left corners on the grid 𝐮+a1d\mathbf{u}+a_{1}\mathbb{Z}^{d}. For a given a1a_{1}, K1=a1dK_{1}=a_{1}^{d} is the number of distinct collections 𝒮a1,𝐮\mathcal{S}_{a_{1},\mathbf{u}} indexed by 𝐮{0,1,,a11}d\mathbf{u}\in\{0,1,\dotsc,a_{1}-1\}^{d}. We always consider a1a0a_{1}\geq a_{0} where a0a_{0} is the fixed constant of the independence assumption (C.2).

Let 𝔹(0,r)={xd:|x|1r}{\mathbb{B}}(0,r)=\{x\in\mathbb{Z}^{d}:|x|_{1}\leq r\} denote the 1\ell^{1}-ball (diamond) of radius r\lfloor{r}\rfloor in d\mathbb{Z}^{d}, with (inner) boundary 𝔹(0,r)={xd:|x|1=r}\partial{\mathbb{B}}(0,r)=\{x\in\mathbb{Z}^{d}:|x|_{1}=\lfloor{r}\rfloor\}.

Lemma C.1.

Assume (C.1) and (C.2). Let a1a0a_{1}\in\mathbb{Z}_{\geq a_{0}} and K1=a1dK_{1}=a_{1}^{d}. Then there exists a constant N0=N0(d)N_{0}=N_{0}(d) such that for NN0N\geq N_{0} and n2(d+1)Nn\geq 2(d+1)N,

(C.3) {\displaystyle\mathbb{P}\bigl{\{} \forall ​lattice path π\pi from the origin to 𝔹(0,n)\partial{\mathbb{B}}(0,n) \exists ​​𝐮([0,a11])d\mathbf{u}\in([0,a_{1}-1]\cap\mathbb{Z})^{d} such that
π intersects at least n4NK1 black cubes from 𝒮a1,𝐮}1exp(n2N).\displaystyle\qquad\text{$\pi$ intersects at least $\frac{n}{4NK_{1}}$ black cubes from $\mathcal{S}_{a_{1},\mathbf{u}}$}\,\bigr{\}}\geq 1\,-\,\exp\Bigl{(}-\,\frac{n}{2N}\Bigr{)}.

To prove Lemma C.1 we record a Bernoulli large deviation bound.

Lemma C.2.

Assume (C.2) and let p(0,1)p\in(0,1) be the marginal probability of a black cube. Then there exist constants A(p,K,δ)>0A(p,K,\delta)>0 such that, for all integers a1a0a_{1}\geq a_{0}, mm\in\mathbb{N}, and δ(0,p/K1)\delta\in(0,p/K_{1}), with K1=a1dK_{1}=a_{1}^{d}, and for any particular sequence S(𝐤1),,S(𝐤m)S(\mathbf{k}_{1}),\dotsc,S(\mathbf{k}_{m}) of distinct NN-cubes, the following estimate holds for some 𝐮\mathbf{u} determined by {S(𝐤i)}i=1m\{S(\mathbf{k}_{i})\}_{i=1}^{m}:

{S(𝐤1),,S(𝐤m) contains at least mδ black cubes from 𝒮a1,𝐮}1eA(p,K1,δ)m.\displaystyle\mathbb{P}\bigl{\{}\text{$S(\mathbf{k}_{1}),\dotsc,S(\mathbf{k}_{m})$ contains at least $m\delta$ black cubes from $\mathcal{S}_{a_{1},\mathbf{u}}$}\bigr{\}}\geq 1-e^{-A(p,K_{1},\delta)m}.

Furthermore, limp1A(p,K,δ)=\lim_{p\nearrow 1}A(p,K,\delta)=\infty for all KK\in\mathbb{N} and δ(0,p/K)\delta\in(0,p/K).

Proof.

Pick 𝐮\mathbf{u} so that 𝒮a1,𝐮\mathcal{S}_{a_{1},\mathbf{u}} contains at least m/K1\lceil{m/K_{1}}\rceil of the cubes S(𝐤1),,S(𝐤m)S(\mathbf{k}_{1}),\dotsc,S(\mathbf{k}_{m}). Since these are colored independently and δ<p/K1\delta<p/K_{1}, basic large deviations gives

(at most mδ cubes among {S(𝐤i)}i=1m𝒮a1,𝐮 are black)\displaystyle\mathbb{P}(\text{at most $m\delta$ cubes among $\{S(\mathbf{k}_{i})\}_{i=1}^{m}\cap\mathcal{S}_{a_{1},\mathbf{u}}$ are black})
(at most mδ cubes among m/K1 independently colored cubes are black)\displaystyle\leq\mathbb{P}(\text{at most $m\delta$ cubes among $\lceil{m/K_{1}}\rceil$ independently colored cubes are black})
exp{mK1Ip(K1δ)}=eA(p,K1,δ)m\displaystyle\leq\exp\Bigl{\{}-\,\frac{m}{K_{1}}I_{p}(K_{1}\delta)\Bigr{\}}=e^{-A(p,K_{1},\delta)m}

where the last equality defines AA and the well-known Cramér rate function [16] of the Bernoulli(p)(p) distribution is

Ip(s)=slogsp+(1s)log1s1pfor s[0,1].I_{p}(s)=s\log\frac{s}{p}+(1-s)\log\frac{1-s}{1-p}\qquad\text{for }\ s\in[0,1].

To complete the proof, observe that

limp1A(p,K,δ)=limp11KIp(Kδ)=limp1(δlogKδp+1KδKlog1Kδ1p)=.\lim_{p\nearrow 1}A(p,K,\delta)=\lim_{p\nearrow 1}\frac{1}{K}I_{p}(K\delta)=\lim_{p\nearrow 1}\Bigl{(}\delta\log\frac{K\delta}{p}+\frac{1-K\delta}{K}\log\frac{1-K\delta}{1-p}\Bigr{)}=\infty.\qed
Proof of Lemma C.1.

Consider for the moment a fixed path π\pi from 0 to a point yy such that |y|1=n|y|_{1}=n. Assume n>dNn>dN so that yS(𝟎)y\notin S(\mathbf{0}).

For j+j\in\mathbb{Z}_{+} let level jj of NN-cubes refer to the collection j={S(𝐤):|𝐤|1=j}\mathcal{L}_{j}=\{S(\mathbf{k}):|\mathbf{k}|_{1}=j\}. Since points x=(x1,,xd)S(𝐤)x=(x_{1},\dotsc,x_{d})\in S(\mathbf{k}) satisfy

kiNxikiN+N1 for i[d],k_{i}N\leq x_{i}\leq k_{i}N+N-1\quad\text{ for $i\in[d]$,}

level jj cubes are subsets of {x:Njd(N1)|x|1Nj+d(N1)}\{x:Nj-d(N-1)\leq|x|_{1}\leq Nj+d(N-1)\}.

To reach the point yy, path π\pi must have entered and exited at least one NN-cube at levels 0,1,,m00,1,\dotsc,m_{0} where m0m_{0} satisfies

Nm0+d(N1)<|y|1N(m0+1)+d(N1).Nm_{0}+d(N-1)<|y|_{1}\leq N(m_{0}+1)+d(N-1).

This calculation excludes the cube that contains the endpoint yy. From this

(C.4) m0|y|1d(N1)N1nN(d+1).m_{0}\geq\frac{|y|_{1}-d(N-1)}{N}-1\geq\frac{n}{N}-(d+1).

Consider the sequence of NN-cubes that path π\pi intersects: S(𝟎)=S(𝐤0),S(𝐤1),,S(𝐤m1)S(\mathbf{0})=S(\mathbf{k}_{0}),S(\mathbf{k}_{1}),\dotsc,S(\mathbf{k}_{m_{1}}), with the initial point 0S(𝟎)=S(𝐤0)0\in S(\mathbf{0})=S(\mathbf{k}_{0}) and the final point yS(𝐤m1)y\in S(\mathbf{k}_{m_{1}}). Remove loops from this sequence (if any), for example by the following procedure:

  1. (1)

    Let i0i_{0} be the minimal index such that 𝐤i0=𝐤j\mathbf{k}_{i_{0}}=\mathbf{k}_{j} for some j>i0j>i_{0}. Let j0j_{0} be the maximal jj for i0i_{0}. Then remove S(𝐤i0+1),,S(𝐤j0)S(\mathbf{k}_{i_{0}+1}),\dotsc,S(\mathbf{k}_{j_{0}}).

  2. (2)

    Repeat the same step on the remaining sequence S(𝐤0),,S(𝐤i0),S(𝐤j0+1),,S(𝐤m1)S(\mathbf{k}_{0}),\dotsc,S(\mathbf{k}_{i_{0}}),S(\mathbf{k}_{j_{0}+1}),\dotsc,S(\mathbf{k}_{m_{1}}), as long as loops remain.

After loop removal relabel the sequence of remaining cubes consecutively to arrive at a new sequence S(𝐤0),S(𝐤1),S(𝐤m2)S(\mathbf{k}_{0}),S(\mathbf{k}_{1})\dotsc,S(\mathbf{k}_{m_{2}}) of distinct NN-cubes with m2m1m_{2}\leq m_{1} and still 0S(𝟎)=S(𝐤0)0\in S(\mathbf{0})=S(\mathbf{k}_{0}) and yS(𝐤m2)y\in S(\mathbf{k}_{m_{2}}). This sequence takes nearest-neighbor steps on the coarse-grained lattice of NN-cubes, in the sense that |𝐤i𝐤i1|1=1|\mathbf{k}_{i}-\mathbf{k}_{i-1}|_{1}=1, because this property is preserved by the loop removal. Since π\pi enters and leaves behind at least one NN-cube on each level 0,,m00,\dotsc,m_{0}, we have the bound m21m0m_{2}-1\geq m_{0}.

We have now associated to each path π\pi a sequence of m0m_{0} distinct NN-cubes that π\pi both enters from the outside and exits again. We apply Lemma C.2 to these sequences of cubes.

Take a1a01a_{1}\geq a_{0}\geq 1 and K1=a1dK_{1}=a_{1}^{d} as in the statement of Lemma C.1. Let δ0=(2K1)1\delta_{0}=(2K_{1})^{-1}. Fix NN large enough so that p=p(N)>12=δ0K1p=p(N)>\tfrac{1}{2}=\delta_{0}K_{1} and the constant given by Lemma C.1 satisfies

A(p,K1,δ0)>log2d+1.A(p,K_{1},\delta_{0})>\log 2d+1.

Consider n2(d+1)Nn\geq 2(d+1)N to guarantee that the rightmost expression in (C.4) and thereby also m0m_{0} is larger than n/(2N)n/(2N). Then also m0δ0n/(4NK1)m_{0}\delta_{0}\geq n/(4NK_{1}). By Lemma C.2,

{path π:0𝔹(0,n) 𝐮 such that π enters and exits\displaystyle\mathbb{P}\bigl{\{}\text{$\forall\,$path $\pi:0\to\partial{\mathbb{B}}(0,n)$ $\exists\mathbf{u}$ such that $\pi$ enters and exits}
at least n4NK1 distinct black cubes from 𝒮a1,𝐮}\displaystyle\qquad\qquad\qquad\qquad\text{at least $\frac{n}{4NK_{1}}$ distinct black cubes from $\mathcal{S}_{a_{1},\mathbf{u}}$}\bigr{\}}
{every nearest-neighbor sequence of m0 N-cubes starting at S(𝟎)\displaystyle\geq\mathbb{P}\bigl{\{}\text{every nearest-neighbor sequence of $m_{0}$ $N$-cubes starting at $S(\mathbf{0})$ }
 contains at least m0δ0 black cubes from some 𝒮a1,𝐮}\displaystyle\qquad\qquad\qquad\qquad\text{ contains at least $m_{0}\delta_{0}$ black cubes from some $\mathcal{S}_{a_{1},\mathbf{u}}$}\bigr{\}}
1(2d)m0eA(p,K1,δ0)m01em01en/(2N).\displaystyle\geq 1-(2d)^{m_{0}}e^{-A(p,K_{1},\delta_{0})m_{0}}\geq 1-e^{-m_{0}}\geq 1-e^{-n/(2N)}.

This completes the proof of Lemma C.1. ∎

Proof of Lemma 5.3.

Surround each NN-cube S(𝐤)S(\mathbf{k}) with 2d2d NN-boxes so that each d1d-1 dimensional face of S(𝐤)S(\mathbf{k}) is directly opposite a large face of one of the NN-boxes. Precisely, first put S(𝐤)S(\mathbf{k}) at the center of the 3N3N-cube T(𝐤)=N𝐤+[N,2N]dT(\mathbf{k})=N\mathbf{k}+[-N,2N]^{d} on d\mathbb{Z}^{d}, and then define 2d2d NN-boxes B±j(𝐤)=T(𝐤)T(𝐤±2𝐞j)B^{\pm j}(\mathbf{k})=T(\mathbf{k})\cap T(\mathbf{k}\pm 2\mathbf{e}_{j}) for j[d]j\in[d]. Any lattice path that enters S(𝐤)S(\mathbf{k}) and exits T(𝐤)T(\mathbf{k}) must cross in the sense of (5.8) one of the NN-boxes that surround S(𝐤)S(\mathbf{k}).

Color S(𝐤)S(\mathbf{k}) black if all 2d2d NN-boxes surrounding it are black. The probability that S(𝐤)S(\mathbf{k}) is black can be made arbitrarily close to 1 by choosing s0s_{0} and NN large enough and δ0>0\delta_{0}>0 small enough in the definition (5.4)–(5.5) of a black NN-box. The color of S(𝐤)S(\mathbf{k}) depends only on the edge variables in the union T¯(𝐤)\overline{T}(\mathbf{k}) of the 2d2d boxes B¯±j(𝐤)\overline{B}^{\,\pm j}(\mathbf{k}) enlarged as in (5.2). The separation of a0a_{0} in (C.2) can be fixed large enough to guarantee that over 𝐤𝐮+a0d\mathbf{k}\in\mathbf{u}+a_{0}\mathbb{Z}^{d} the cubes T¯(𝐤)\overline{T}(\mathbf{k}) are pairwise disjoint.

Apply Lemma C.1 with K1=a1d=a0dK_{1}=a_{1}^{d}=a_{0}^{d}. Tighten the requirement n2(d+1)Nn\geq 2(d+1)N of Lemma C.1 to n4dNn\geq 4dN to guarantee that if a path π\pi intersects S(𝐤)S(\mathbf{k}) then it also intersects the complement of T(𝐤)T(\mathbf{k}). (If π\pi remains inside T(𝐤)T(\mathbf{k}) then the 1\ell^{1}-distance between the endpoints of π\pi is at most 3dN3dN and π\pi cannot connect the origin to 𝔹(0,n)\partial{\mathbb{B}}(0,n).) Thus for every S(𝐤)S(\mathbf{k}) intersected by π\pi, at least one of the NN-boxes surrounding S(𝐤)S(\mathbf{k}) is crossed by π\pi in the sense of (5.8). In conclusion, on the event in (C.3) each path from the origin to 𝔹(0,n)\partial{\mathbb{B}}(0,n) crosses at least n/(4NK1)=na0d/(4N)\lceil{n/(4NK_{1})}\rceil=\lceil{na_{0}^{-d}/(4N)}\rceil disjoint NN-boxes. Of these, at least na0d/(4N)/K\lceil{na_{0}^{-d}/(4N)}\rceil/K must come from some particular collection j\mathcal{B}_{j}. Thus in Lemma 5.3 we can take δ1=1/(4a0dNK)\delta_{1}=1/(4a_{0}^{d}NK), n1=4dNn_{1}=4dN and D1=1/(2N)D_{1}=1/(2N). ∎

Appendix D Convex analysis

Lemma D.1.

Let ff be a proper convex function on d\mathbb{R}^{d} (<f-\infty<f\leq\infty and ff is not identically \infty) and ξri(domf)\xi\in\operatorname{ri}(\operatorname{dom}f). Then the following statements are equivalent.

  1. (a)

    For some bb\in\mathbb{R}, f(ξ){hd:hξ=b}\partial f(\xi)\subset\{h\in\mathbb{R}^{d}:h\cdot\xi=b\}.

  2. (b)

    ff^{*} is constant over f(ξ)\partial f(\xi).

  3. (c)

    tf(tξ)t\mapsto f(t\xi) is differentiable at t=1t=1.

Proof.

(a)\implies(b). For all hf(ξ)h\in\partial f(\xi), f(h)=hξf(ξ)=bf(ξ)f^{*}(h)=h\cdot\xi-f(\xi)=b-f(\xi).

(b)\implies(a). Suppose f(h)=sf^{*}(h)=s for all hf(ξ)h\in\partial f(\xi). Then for all hf(ξ)h\in\partial f(\xi), hξ=f(h)+f(ξ)=s+f(ξ)h\cdot\xi=f^{*}(h)+f(\xi)=s+f(\xi).

(c)\implies(a). Let b=(d/dt)f(tξ)|t=1b=(d/dt)f(t\xi)|_{t=1} and hf(ξ)h\in\partial f(\xi). Then for all |s|ε|s|\leq\varepsilon, by convexity, f(ξ+sξ)f(ξ)shξf(\xi+s\xi)-f(\xi)\geq s\hskip 0.55pth\cdot\xi. This says that hξh\cdot\xi lies in the subdifferential of the function tf(tξ)t\mapsto f(t\xi) at t=1t=1, but by assumption this latter equals the singleton {b}\{b\}.

(a)\implies(c). The directional derivatives satisfy the following, where in both equations the second equality comes from [17, Thm. 23.4].

f(ξ;ξ)=lims0f(ξ+sξ)f(ξ)s=sup{ξh:hf(ξ)}=b\displaystyle f^{\prime}(\xi;\xi)=\lim_{s\searrow 0}\frac{f(\xi+s\xi)-f(\xi)}{s}=\sup\{\xi\cdot h:h\in\partial f(\xi)\}=b

and

f(ξ;ξ)=lims0f(ξsξ)f(ξ)s=sup{ξh:hf(ξ)}=b.\displaystyle f^{\prime}(\xi;-\xi)=\lim_{s\searrow 0}\frac{f(\xi-s\xi)-f(\xi)}{s}=\sup\{-\xi\cdot h:h\in\partial f(\xi)\}=-b.

From this we see the equality of the left and right derivatives of φ(t)=f(tξ)\varphi(t)=f(t\xi) at t=1t=1:

φ(t)=limt0f(ξ+tξ)f(ξ)t=f(ξ;ξ)=b\varphi^{\prime}(t-)=\lim_{t\nearrow 0}\frac{f(\xi+t\xi)-f(\xi)}{t}=-f^{\prime}(\xi;-\xi)=b

and

φ(t+)=limt0f(ξ+tξ)f(ξ)t=f(ξ;ξ)=b.\varphi^{\prime}(t+)=\lim_{t\searrow 0}\frac{f(\xi+t\xi)-f(\xi)}{t}=f^{\prime}(\xi;\xi)=b.\qed

References

  • [1] Antonio Auffinger and Michael Damron. Differentiability at the edge of the percolation cone and related results in first-passage percolation. Probab. Theory Related Fields, 156(1-2):193–227, 2013.
  • [2] Antonio Auffinger, Jack Hanson, and Michael Damron. 50 years of first passage percolation, volume 68 of University Lecture Series. American Mathematical Society, Providence, RI, 2017.
  • [3] Erik Bates. Empirical distributions, geodesic lengths, and a variational formula in first-passage percolation. 2020. Preprint (arXiv:2006.12580).
  • [4] Ivan Corwin. Kardar-Parisi-Zhang universality. Notices Amer. Math. Soc., 63(3):230--239, 2016.
  • [5] J. Theodore Cox and Richard Durrett. Some limit theorems for percolation processes with necessary and sufficient conditions. Ann. Probab., 9(4):583--603, 1981.
  • [6] Michael Damron, Firas Rassoul-Agha, and Timo Seppäläinen. Random growth models. Notices Amer. Math. Soc., 63(9):1004--1008, 2016.
  • [7] Michael Damron and Pengfei Tang. Superlinearity of geodesic length in 2d critical first-passage percolation. In Sojourns in Probability Theory and Statistical Physics - II, pages 101--122. Springer Singapore, 2019.
  • [8] Duncan Dauvergne, Janosch Ortmann, and Bálint Virág. The directed landscape. Acta Math., 2022. To appear (arXiv 1812.00309).
  • [9] Duncan Dauvergne and Bálint Virág. The scaling limit of the longest increasing subsequence. 04 2021. Preprint (arXiv 2104.08210).
  • [10] Richard Durrett and Thomas M. Liggett. The shape of the limit set in Richardson’s growth model. Ann. Probab., 9(2):186--193, 1981.
  • [11] Geoffrey Grimmett and Harry Kesten. First-passage percolation, network flows and electrical resistances. Z. Wahrsch. Verw. Gebiete, 66(3):335--366, 1984.
  • [12] John M. Hammersley and J. A. Dominic Welsh. First-passage percolation, subadditive processes, stochastic networks, and generalized renewal theory. In Proc. Internat. Res. Semin., Statist. Lab., Univ. California, Berkeley, Calif, pages 61--110. Springer-Verlag, New York, 1965.
  • [13] Harry Kesten. On the time constant and path length of first-passage percolation. Adv. in Appl. Probab., 12(4):848--863, 1980.
  • [14] Harry Kesten. Aspects of first passage percolation. In École d’été de probabilités de Saint-Flour, XIV---1984, volume 1180 of Lecture Notes in Math., pages 125--264. Springer, Berlin, 1986.
  • [15] R. Marchand. Strict inequalities for the time constant in first passage percolation. Ann. Appl. Probab., 12(3):1001--1038, 2002.
  • [16] Firas Rassoul-Agha and Timo Seppäläinen. A course on large deviations with an introduction to Gibbs measures, volume 162 of Graduate Studies in Mathematics. American Mathematical Society, Providence, RI, 2015.
  • [17] R. Tyrrell Rockafellar. Convex analysis. Princeton Mathematical Series, No. 28. Princeton University Press, Princeton, N.J., 1970.
  • [18] R. T. Smythe and John C. Wierman. First-passage percolation on the square lattice, volume 671 of Lecture Notes in Mathematics. Springer, Berlin, 1978.
  • [19] J. Michael Steele and Yu Zhang. Nondifferentiability of the time constants of first-passage percolation. Ann. Probab., 31(2):1028--1051, 2003.
  • [20] J. van den Berg and H. Kesten. Inequalities for the time constant in first-passage percolation. Ann. Appl. Probab., 3(1):56--80, 1993.
  • [21] Yu Zhang. Supercritical behaviors in first-passage percolation. Stochastic Process. Appl., 59(2):251--266, 1995.