Geodesic length and shifted weights
in first-passage percolation
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 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 shift2020 Mathematics Subject Classification:
60K35, 60K371. 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 -dimensional integer lattice . The one-dimensional case usually reduces to classical probability so the real work begins from the planar case .
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 by where the are nonnegative, independent and identically distributed (i.i.d.) random weights on the nearest-neighbor edges between vertices of and the infimum is over self-avoiding paths between the two points and . A minimizing path is called a geodesic between and . 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 and the geodesics as the distance between and grows. At the level of the law of large numbers, under suitable hypotheses, normalized passage times converge with probability one: as , whenever . The special case of the limit is also called the time constant.
The limiting shape function is a norm that characterizes the asymptotic shape of a large ball. Define the randomly growing ball in for by where is obtained from by taking integer parts coordinatewise. Under the right assumptions, as the normalized ball converges to the unit ball defined by the norm .
The shape function is not explicitly known in any nontrivial example. Soft properties such as convexity, continuity, positive homogeneity, and for 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 let denote the shape function for the shifted weights . Let be the minimal Euclidean length of a geodesic from the origin to the point for the shifted weights . Then the important fact is that when ,
(1.1) |
provided the derivative at on the right-hand side exists.
The shape function is a concave function of and hence the derivative in (1.1) exists and the limit holds for all but countably many shifts . 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 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 and , and for the particular sequence . Their result applied to the geodesic of the so-called cylinder passage time from to , 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 , 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 denote the critical probability of Bernoulli bond percolation on . When 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 becomes identically zero.
Zhang [21] proved in 1995 that in the supercritical case defined by , for and , the limit on the left in (1.1) exists and equals a nonrandom constant. In the planar critical case, that is, , and , Damron and Tang [7] proved that the left-hand side in (1.1) blows up in all directions .
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 , there exists such that, if , and , then the derivative in (1.1) fails to exist at . 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 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 -approximate geodesics. These are paths whose endpoints are order apart and whose passage times are within of the optimal passage time. Through these we can capture the entire superdifferential of the shape function as a function of the shift .
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 and 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 and are connected with the FPP shape function in several ways. A key fact is that and agree with on certain subsets of described by positively homogeneous functions that are connected with geodesic length (Theorems 2.11 and 2.16). Second, and generate as the maximal positively homogeneous convex function dominated by and (Remark 2.15). Third, and contain the information for generating all the shifts 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
-
(i)
versions of the van den Berg-Kesten modification arguments [20],
-
(ii)
the convex duality (Theorem 2.17), and
- (iii)
Our results are valid on in all dimensions , under the standard moment bound needed for the shape theorem and the assumption that the minimum of the edge weight has probability strictly below .
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 have length at least for a fixed constant (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 -approximate geodesics and the superdifferential of the shape function as a function of the weight shift (Theorem 2.17). When the random weight 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 as a function of the weight shift , in a fixed spatial direction . This function is strictly concave in (Theorem 2.2). In the situations where the atoms of bring about geodesics whose asymptotic normalized lengths vary, the concave function 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 , and all distributions with an atom at the origin. Furthermore, we disprove their conjecture that 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 of uniform random variables, one can consider perturbations of the weights and differentiate the shape function in directions 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 .
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 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. , , and . For , . Standard basis vectors in are , and is the origin of . The norm of is . Particular subsets of that recur are , , , and the topological interior .
A finite or infinite path or sequence is denoted by for . Other notation for lattice paths are and . The steps of a path are the nearest-neighbor edges . A finite path that satisfies is called an -path.
A positively homogeneous function satisfies for whenever both and are in the domain of [17, p. 30]. One-sided derivatives of a function defined around are defined by and .
The diamond is a wild card for three superscripts (no superscript at all), (zero steps allowed), and sa (self-avoiding) that distinguish different FPP processes with restricted path length.
A real number is an atom of the random edge weight if . and . Superscript on any quantity means that it is computed with weights shifted by : .
The symbol marks the end of a numbered remark.
1.10. Acknowledgements
2. The models and the main results
2.1. Setting
Fix an arbitrary dimension . Let denote the set of undirected nearest-neighbor edges between vertices of . is the probability space of an environment such that the edge weights are independent and identically distributed (i.i.d.) real-valued random variables. Translations act on by for a nearest-neighbor edge .
A nearest-neighbor path is any finite sequence of vertices that satisfy for each . The steps of are the nearest-neighbor edges . The Euclidean length of is the number of edges, so in this case . Then we call an -path. The passage time of is the sum of the weights of its edges:
(2.1) |
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 that has no edges and has zero passage time and 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 is self-avoiding if for all pairs . Let denote the collection of all self-avoiding paths from to , of arbitrary but finite length. Define the passage time between and as
(2.2) |
This definition gives because the only self-avoiding path from to is the zero-length path. A geodesic is a self-avoiding path that minimizes in (2.2).
When the restriction to self-avoiding paths is superfluous in the definition of . Let denote the critical probability of Bernoulli bond percolation on . A frequently used assumption in FPP is that zero-weight edges are subcritical:
(2.3) |
For nonnegative weights, the assumption (2.3) guarantees the existence of a geodesic (Prop. 4.4 in [2]).
For , define -shifted weights by
(2.4) |
All the quantities associated with weights acquire the superscript. For example, is the passage time in (2.2) under weights . Let
(2.5) |
denote the (essential) lower bound of the weights. So in particular, 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) |
Let denote i.i.d. copies of the edge weight . The following moment assumption will be employed for various values of .
(2.7) |
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 , (2.6), and the moment bound (2.7) with . Then there exists a constant , determined by the dimension and the distribution of the shifted weights , and a full-probability event such that the following statements hold. For each real there exists a continuous, convex, positively homogeneous shape function such that the limit
(2.8) |
holds for each , whenever satisfies . We have and for .
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 .
To emphasize dependence on with fixed, we write
(2.9) |
Several of our main results concern the regularity of and its connections with geodesic length. The reason for allowing negative weights by extending the shift below is to enable us to talk about the regularity of at . Throughout this paper, is the constant specified in Theorem 2.1.
Theorem 2.2.
Concavity implies that one-sided derivatives for exist, , and as functions of , they are nonincreasing, is left-continuous, and 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) | ||||
with the superscripted variants and for shifted weights . For a continuous weight distribution almost surely because in that case geodesics are unique almost surely. This is not the case for all shifts because as 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 of full probability: for all and , and there exist a finite deterministic constant and a finite random constant such that
(2.11) |
We justify part (i) of Theorem 2.2. This sets the stage for further discussion. Let . Take and . Considering the shifted weights on the minimal and maximal length geodesics of leads to
(2.12) |
Rearrange to
(2.13) |
Here are the arguments for the properties of claimed in part (i) of Theorem 2.2.
- (i.a)
-
(i.b)
Concavity follows by taking the same limit in (2.13).
-
(i.c)
Continuity of on the open interval follows from concavity.
Since , (2.13) and the monotonicity of the derivatives give the easy bound
(2.14) |
A corollary of the strict concavity given in Theorem 2.2(ii) is the strict inequality . 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 , (2.6), and the moment bound (2.7) with . Let be the constant specified in Theorem 2.1 and let be the constant in (2.11). Then there exists a full-probability event such that the following holds: for all shifts , directions , weight configurations , and sequences , we have the bounds
(2.15) | ||||
is a nonincreasing function of such that for all .
The first inequality in (2.15) says that the strict concavity gap is uniform across all directions . This point is further strengthened to a uniformity for fixed weight configurations in Theorem 2.5.
Remark 2.4.
(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 is differentiable at . Then (2.15) implies that for all and sequences ,
(2.16) |
By concavity, this happens at all but countably many . In particular, if 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 . Then (2.15) tells us that all the possible asymptotic normalized lengths of geodesics that go in direction form a subset of the interval . Presently there is no description of this subset.
For a characterization of in terms of path length, given below in Theorem 2.17, we expand the class of paths considered to allow -approximate geodesics. These are paths from the origin to whose passage times are in the range , without necessarily being geodesics between their endpoints.
(ii) The strict concavity of given in Theorem 2.2 and the inequalities in (2.15) imply that, for all and sequences and ,
(2.17) |
In other words, distinct shifts of a given weight distribution cannot share any possible asymptotic geodesic lengths, even under distinct but typical environments and .
(iii) There is a corresponding monotonicity for geodesics at fixed . Namely, when all the weights increase by a common constant, geodesics can only shrink in length. Let and be arbitrary geodesics for and , respectively. Then
(2.18) | for fixed and . |
This follows from
(2.19) | ||||
Furthermore, suppose a unique geodesic is chosen, for example by taking the minimal one according to some ordering of geodesics. Then as increases to , the geodesic cannot change without its length strictly shrinking:
(2.20) | for fixed and , implies . |
This follows because the string of inequalities (2.19) together with implies that , so is still at least as good as for weights .
(iv) We establish below in Theorem 2.17 that as . Naturally, as the weight shift grows very large, it pays less to search for smaller weights at the expense of a longer geodesic.
The first inequality in (2.15) implies that asymptotically the lengths of geodesics in a particular direction exceed the -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.
We turn to nondifferentiability results for . An atom of the weight distribution is a value such that .
Theorem 2.6.
Assume , (2.6), and the moment bound (2.7) with . Additionally, assume that the weight distribution satisfies at least one of the assumptions (a) and (b) below:
-
(a)
zero is an atom;
-
(b)
there are two strictly positive atoms such that is rational.
Then there exist constants such that
(2.21) |
Furthermore, for all , and so the function is not differentiable at .
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.
The proof of Theorem 2.7 in Section 7.2 constructs the singularity set explicitly from two atoms of as a countably infinite union of arithmetic sequences.
Remark 2.8.
Standard Bernoulli weights satisfy . In the subcritical planar Bernoulli case (that is, , and ), Steele and Zhang [19] proved that is not differentiable at , as long as is close enough to . Furthermore, they conjectured that is differentiable at all such that except at (page 1050 in [19]).
Theorem 2.6 above extends the nondifferentiability at to all directions , 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 .
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 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 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 of admissible steps is augmented to . For and define three classes of paths from to of length , presented here from largest to smallest:
(2.22) | ||||
The superscript in is for self-avoiding. Paths in and are allowed to repeat both vertices and edges. Paths in are called -admissible, and those in -admissible. An -path from to is an -path if . For and we define each collection as consisting only of the zero-length path . For , and are nonempty if and only if is a nonnegative even integer, while is nonempty if and only if .
With the three classes of paths go three collections of points reachable by an admissible path of length from the origin: for the three superscripts , define
(2.23) |
If , any -path can be augmented to an -path by adding zero steps, and hence we have .
The environment is extended to zero steps by stipulating that zero steps always have zero weight, even when weights are shifted: and .
Define three point-to-point first-passage times between two points with restricted path lengths: for ,
(2.24) |
If , set . Obvious relations hold between these passage times and the standard FPP from (2.2):
(2.25) |
and
(2.26) |
For nonnegative weights the restriction to self-avoiding paths is superfluous for and hence
(2.27) |
These identities point to the usefulness of and . Namely, they capture the FPP passage time when the path length parameter coincides with a geodesic length. After taking this connection to the limit, the discrepancies between the shape function of and the FPP shape function 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 and the zero-weight zero step. We shall see that is a convenient link between standard FPP and restricted path length FPP because it is monotone:
(2.28) |
The monotonicity is simply a consequence of the fact that any -path can be augmented to an -path by adding zero steps.
The self-avoiding version 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 .
We state a shape theorem for restricted path length FPP, but only on the open set . Its closure, the compact ball , is the convex hull of both and and the set of possible asymptotic velocities of admissible paths in as . In this theorem we introduce the parameter as a variable that controls asymptotic path length.
Theorem 2.9.
Assume and that the moment bound (2.7) holds with for the nonnegative weights . Then there exist
-
(a)
nonrandom continuous convex functions and and
-
(b)
an event of -probability one
such that the following statement holds for any fixed : for any , any real , and any sequences in , and such that , and , we have the laws of large numbers
(2.29) |
Furthermore, and . In general on . If then on all of . If then 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 and 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 defined on the whole of . 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 . Since there is a unique -path from to , we see that a finite limit is possible only if :
(2.30) |
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 via limits along rays: for and define . The resulting functions and are both convex and lower semicontinuous.
With this theorem we can extend the functions to lower semicontinuous proper convex functions on all of by setting
(2.31) |
If is finite on , then is automatically upper semicontinuous on [17, Theorem 10.2], and hence continuous on .
The next theorem clarifies the relationship of and with , beyond the obvious , and links their connection with the asymptotic geodesic lengths from Theorem 2.3. In particular, we introduce here two functions that play several roles in our asymptotic results. In the theorem below they are first introduced as the boundaries of the regions where coincides with and . Part (ii) indicates that and are also related to the derivatives of and geodesic length.
These properties are then elaborated on as we proceed. The interval captures all the asymptotic lengths of geodesics in direction , 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 and describe ranges where and are affine and where these two functions disagree. The macroscopic description is completed in Theorem 2.17: as the weight shift increases, the interval shifts to the left and always equals the superdifferential of the concave function . 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) |
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.
-
(i)
There exist two positively homogeneous functions and such that , and for all ,
(2.33) and
(2.34) Furthermore, is lower semicontinuous and is upper semicontinuous. If then , while is finite in the case .
-
(ii)
Strengthen the subcriticality assumption to (2.6). There exists a nonrandom constant and a full-probability event such that, for all , sequences , and ,
(2.35)
We spell out some of the consequences of the theorems.
Remark 2.12 (Coincidence of shape functions).
There exists a finite constant such that . 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 contains the nondegenerate neighborhood of the origin.
If then because . If the equality holds for at least one nonzero point along each ray from the origin. With all of the above, the first inequality of (2.35) implies that and are both nonempty closed subsets of .
Remark 2.13 (-approximate geodesics).
For , (2.34) gives the equivalence if and only if . (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 and paths from to such that , and . These paths do not have to be self-avoiding or geodesics between their endpoints. But does imply that is within of the passage time of the geodesic between and . The asymptotic normalized lengths of true self-avoiding geodesics for are a subset of the interval of asymptotic normalized lengths of -approximate geodesics, as indicated in (2.35).
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 , (2.35) shows that convergence in direction follows from , which is equivalent to the condition that the set has empty relative interior on the -directed ray.
Remark 2.15 (Convexity).
Fix . For , the convexity and continuity of on imply the convexity and continuity of the function defined for . By Theorem 2.10, extends to by letting . By (2.31), we extend to by setting its value equal to . Thereby is a lower semicontinuous proper convex function on .
For , monotonicity (2.28) implies further that
(2.36) | is nonincreasing for . |
A consequence of Theorem 2.11 is that for ,
(2.37) |
In the language of convex analysis [17, p. 35], the identity above characterizes the standard FPP shape function as the positively homogeneous convex function generated by . This means that is the greatest positively homogeneous convex function such that and . Figure 2.2 illustrates (2.37).
The last theorem of this section records further properties of , 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 , (2.3), and the moment bound (2.7) with . Fix . For , the shape functions of Theorem 2.9 have the following properties along the -directed ray from the origin.
-
(i)
The function is continuous, convex and strictly increasing for . Both functions are affine at least in one nondegenerate interval with one endpoint at the origin: for ,
(2.38) -
(ii)
For ,
(2.39) -
(iii)
Strengthen the subcriticality assumption to (2.6). The function is continuously differentiable on the open interval and . If then the left derivative of at exists and equals .
Notice that the right-hand sides in (2.38) agree if and only if , as is consistent with the agreement when . From (2.39) and (2.35) we read that if , the set is an open neighborhood of that consists of finite rays from the origin, while its complement contains the nonempty annulus , where is the constant in (2.35). Another consequence of (2.38) and (2.39) is that is never strictly between and but always agrees with at least one of them.
2.4. Duality of the weight shift and geodesic length
This section develops the duality between the weight shift variable in and the path-length variable in the limit shapes (2.29). Nonnegative weights () are assumed throughout.
Fix for the duration of this section. We restrict the shape function of (2.9) to shifts that preserve the nonnegativity of the weights and then extend it to an upper semicontinuous concave function on all of by setting
(2.40) |
To emphasize, the function drops the extension to done in Theorem 2.1. The reason for this choice is that developing the duality for shifts requires a study of the shape function of the self-avoiding version 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 is another upper semicontinuous concave function, and together and satisfy
(2.41) |
The superdifferential of the concave function at is by definition the set
By the definition for . For , is the bounded closed interval and so if and only if . These general equivalences hold:
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 and on the right in the case . In particular, on the left the affine portion of on the interval is the dual of the superdifferential in (2.46). The infinite slope at the left edge 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) |
This and (2.37) applied to give (2.44) below for , which is the basis for the duality.
Theorem 2.17.
Remark 2.18.
(a) Let us make explicit the conversion back to the original FPP shape function in Theorem 2.17. In (2.44) can be replaced by for . In each of (2.45), (2.46) and (2.48), can be replaced by . (2.47) cannot be valid for because for some . We do have
(2.49) |
(b) The strict concavity of 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 that depends on the dimension and the weight distribution such that, for all ,
(2.50) |
The strict inequalities above are due to the strict concavity of .
(c) When the infimum of the support of the weights is zero, and 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) |
We can combine (2.44) and (2.51) into a statement that shows that both and contain full information for retrieving all the shifts of among nonnegative weights:
(2.52) |
(d) Equations (2.33), (2.51), and positive homogeneity of and show that is affine for large :
(2.53) |
The reader can recognize this statement as the dual version of from the theorem above, and an immediate consequence of (2.38). This affine portion of is visible in both diagrams of Figure 2.2.
Identities (2.51) and (2.53) suggest that, for , the recipe for an optimal path of length approximately from to a point close to is this: shift the weights so that their infimum is zero and take the optimal path for the shifted weights . In particular, once is above the FPP geodesic length, we can follow the FPP geodesic of the shifted weights and extend the path to length by finding and repeating an edge whose weight is close to the minimum .
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 ? For example, are there weight configurations and and sequences and such that
(3.1) |
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 and almost surely constant?
3.2. Properties of the shape functions
Is differentiable when the weight distribution is continuous? What about the case of a single positive atom which is not covered by Theorems 2.6–2.7? Is any comparison between and possible for two distinct directions and ? Is the function defined in (2.33) a norm on ? Do and 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 to nonnegative weights through definition (2.40). This leaves open the duality of for . To capture the full convex duality over all shifts requires a study of the process , restricted path length FPP that optimizes over self-avoiding paths, in a manner analogous to our study of and its shape function.
The present shortcoming can be seen for example in the case of (2.35) where blows up and cannot capture the left derivative . Graphically this same phenomenon appears in the left diagram of Figure 2.2 where the graph of never separates from after . The graph of the function of the self-avoiding version will separate from for large enough and capture .
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 to .
3.5. General perturbations of weights
Develop versions of our results for other perturbations of the weights, besides the simple shift , 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 and the moment bound (2.7) with for the nonnegative weights . Take the existence of the continuous, convex functions that satisfy the laws of large numbers (2.29) from Theorem B.1 in Appendix B. The limit implies . Extend the shape functions and to all of through radial limits: for , define
(4.1) |
The limit exists because is a convex function on the interval . Monotonicity (2.36), , and the limit combine to give, for ,
(4.2) |
Lemma 4.1.
Assume and the moment bound (2.7) with for the nonnegative weights . The restricted path shape functions have the following properties.
-
(i)
and .
-
(ii)
If then on all of . If then in an open neighborhood of the origin.
-
(iii)
For all and ,
(4.3) and the infimum on the right is attained at some . In particular, implies .
-
(iv)
For , the extended function is convex and lower semicontinuous on .
Proof.
Part (i). The lower bounds and on all of follow from and . Also immediate is . Given , we can fix as measurable functions of almost every ,
(4.4) | an edge such that , and a path from to . |
For large enough consider paths that follow and then repeat edge times. Then and in the limit .
Part (ii). The claim for is true because the zero steps of a path can be replaced by repetitions of an edge with weight close to . Here is a detailed proof.
Fix and a sequence such that . Let be an optimal path for and let be the number of zero steps in . Let and be as in (4.4). We construct an -admissible path of length from to or that repeats edge as many times as possible, as follows.
-
•
First, if is even, let , and if is odd, let . Then (plus the step if necessary) goes from to in nonzero steps.
-
•
The remaining steps are spent in an initial segment from back to by first following to , then back and forth across altogether times, and then from back to along (in reverse direction). If then the initial segment does not go all the way to but turns back towards after steps along .
Let denote the edges of . We get the following bound:
(4.5) | ||||
Divide through by and let along a suitable subsequence, utilizing and . We obtain
The last term vanishes almost surely because in probability. Since always, letting establishes the equality under .
Assume . The inequalities in (4.2) imply that holds in (4.3). To prove the opposite inequality in (4.3), consider first so that we can take advantage of the laws of large numbers. Choose and so that , and . Begin with
Let and choose a partition such that . Choose integers such that , , and . (When , only requires to have the right parity.) Then
where we again utilize (4.4): for whenever , construct an -path from to by first going from to one endpoint of , repeating as many times as needed, returning to , and then following an optimal -path from to . (If is too small to allow travel all the way to , then proceed part of the way and return to . is even because .) The number of repetitions of is at most when is large enough.
In the limit
Let to complete the proof of (4.3) in the case . The infimum in (4.3) is attained because on the right either the extended function is continuous down to or then it blows up to .
To complete the proof of (4.3) we show that when . Only needs proof. Let . Since we can assume . Pick so that for . Then by (4.3) applied to the case , for we have
Letting gives .
Part (iv). Convexity extends readily to all of . If in , then for convexity on gives and we can let .
We check the lower semicontinuity of the extension on . Since is continuous in the interior, we need to consider only limits to the boundary. Let , and in . By the limit in (4.1) we can pick so that . By the continuity of on , . Pick so that for . Apply (4.2) to with and to get , again for all . Lower semicontinuity of has been established.
Lower semicontinuity of follows from and the equality on the boundary: when and in , ∎
In the remainder of this section we investigate the connections of and with standard FPP and assume and either (2.3) or (2.6). We begin with the fact that and 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.
Proof.
We claim that there exists a constant such that
(4.7) |
It suffices to prove that a constant works for all . Towards this end we show the existence of a deterministic constant and a random constant such that
(4.8) |
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 such that, for all ,
(4.9) |
By adding these probabilities over the cases we get
(4.10) |
Thus there exists a random constant such that any self-avoiding path from the origin of length satisfies .
Since the FPP shape function is positively homogeneous, by the FPP shape theorem ([2, p. 11] or (A.8) in Appendix A) we can increase if necessary so that, for a deterministic constant ,
(4.11) |
Given such that , let . Then for all large enough , . Hence, recalling (2.25),
In the limit . (The requirement was imposed precisely to justify the limit .) By the lower bound and the monotonicity in (4.2), for . (4.7) has been verified.
Define
(4.12) |
The claimed properties of the function follow. ∎
Later in the paper (Corollary 7.2) after much more work we can show that .
In the next lemma we strengthen the subcriticality assumption to (2.6) so that we can apply Lemma 4.2 to the shifted weights and .
Lemma 4.3.
Assume , (2.6), and the moment bound (2.7) with . For , the shape functions have the following properties for a fixed .
-
(i)
On the -directed ray these functions are affine in a nondegenerate interval started from zero: for ,
(4.13) -
(ii)
The function is continuous, convex, and strictly increasing for .
Proof.
Since the functions are central to our treatment, we rewrite (4.6) in this form:
(4.14) |
Together with (4.3) the above implies that some satisfies . By the inequalities, any such must satisfy . Now we have
(4.15) |
Furthermore, for ,
(4.16) |
is a meaningful definition as the supremum of a nonempty set. Positive homogeneity of on follows from the positive homogeneity of . By Lemma 4.1(ii) and (4.14),
(4.17) |
Recall . Let be such that . Then
Thus
(4.18) |
Since implies that , (4.16) is not a meaningful definition of . Cued by (4.17) and (4.18), we can retain positive homogeneity by defining
(4.19) |
The next proposition collects properties of the functions for . 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 . The trichotomy in the proposition is illustrated in Figure 2.2.
Proposition 4.4.
Proof.
Part (i). Assume so there is something to check. Since is nonincreasing, convex, and reaches its minimum at but not before, it must be strictly decreasing for .
Suppose for some . (Equality holds at by Lemma 4.1(iii).) We show that . By (4.3), for some and all
Thus is constant on with . It must be that .
Part (ii). From Part (i) and (4.14), the behavior of is completely determined. Furthermore, achieves its minimum at by a combination of (4.3) with Part (i) and (4.14). Then must be nondecreasing for , and definition (4.16) forces for .
Part (iii) follows from convexity and the definitions. ∎
The next lemma shows that is lower semicontinuous and upper semicontinuous.
Lemma 4.5.
Let in . Then
(4.21) |
Proof.
If , the first inequality of (4.21) is trivial. Suppose . Then . By continuity on , for large , which implies .
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 . For all , we have
Proof.
We may assume that for the following. The extension to follows from the homogeneity and convexity of in .
Claim 4.7.
For each , there exist edge-disjoint paths from to such that their Euclidean lengths satisfy and for .
The proof of this claim comes after the proof of the lemma, but it is intuitively clear that there exist edge-disjoint paths from to such that at least one path has length and the length of each path is at most for some constant (see [14, Fig. 2.1] for the case ). Then,
(4.22) |
where the first inequality follows from , the second from subadditivity, and the third from the fact that is an infimum over all paths from to . Denote the integrand on the right-hand side of (4.22) by .
Since for , we have
Next, we show that is integrable (see [2, Theorem 2.2]) in preparation for the dominated convergence theorem. A union bound over the edges of each path and independence of the edge weights in the paths implies
Integrating over shows that for some constant ,
Since can be written as
we see that . Therefore, by the dominated convergence theorem, we have
Proof of Claim 4.7.
For general , let be the number of non-zero coordinates of and suppose . This is the effective dimension of the rectangle formed with the origin and as extreme opposing corners. We may assume without loss of generality that the first coordinates of are non-zero and the rest are . So let .
The first disjoint paths run along the edges of the rectangle. Such a path is encoded by a permutation . For example, corresponds to the path . Two paths encoded by permutations and meet (share a vertex) before if and only if for some , . Consider the paths corresponding to the cyclic permutations:
These paths are vertex disjoint, except for their first and last vertices, and have length .
The next paths are formed as follows. For each , start with an step, follow the path to , and conclude with a step to . Get another paths by starting with and finishing with . These paths have length .
Now we have altogether paths. The final paths are a little trickier.
For each , pair direction with path . We construct the path for , and the rest are similar. The first step is . Then follow until is about to step in the direction (the last step before it steps in the direction). On the segment take steps and then take steps in the direction (this avoids the path), ending up at . Finish at by taking a final step. Replacing and by and for gives us such paths that are disjoint from each other and all the previous paths (except for their first and last vertices). All these have length . Notice the crucial assumption of for this construction.
The case is covered in [14, Fig 2.1], as mentioned earlier. One can verify that this gives the worst case of . ∎
Proof of Theorem 2.17.
Step 1. Identity (2.44). For , (2.44) is a combination of (4.15) and (2.42). For large
(4.23) |
because an -path from to a point close to 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 as many times as needed, and then follow a geodesic to a point close to . Bound (4.23) implies that the right-hand side of (2.44) equals for . Identity (2.44) has been verified for all .
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 from (Theorem 2.11(ii)), and the strict exceedance of 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 .
An -box is by definition a rectangular subset of of the form
(5.1) |
for some and . In other words, one of the dimensions of has size and the other dimensions are of size . The two large faces of in (5.1) are the subsets
The interior of is defined by requiring and in (5.1). The boundary of is the set of points of that have a nearest-neighbor vertex in the complement of . Our convention will be that an edge lies in if both its endpoints lie in , otherwise . A suitable -neighborhood around is defined by
(5.2) |
The significance of the choice is that the -distance from any point in to the boundary of is at least as large as the distance between any two points in .
Introduce two parameters whose choices are made precise later. Consider these conditions on the edge weights in and :
(5.3) |
(5.4) |
and
(5.5) | for every self-avoiding path that stays entirely in | |||
and whose endpoints and satisfy . |
The properties of a black box stated in the next definition depend on whether the weights are bounded or unbounded. We let .
Definition 5.1 (Black box).
By choosing and large enough and small enough, the probability of a given being black can be made as close to 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.
When , Lemma 5.5 of [20] requires the weaker assumption where is the critical probability of oriented bond percolation on . However, since we consider shifts of weights that can turn into zero, it is simpler to assume (2.6) for all 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
The bound above decreases for large enough and hence gives us this conclusion:
(5.7) | There exists a fixed such that for any there exist and | |||
such that while . Increasing and | ||||
while keeping fixed cannot violate this condition as long as . |
Condition is included above simply to point out that is not chosen so large that property (5.3) becomes trivial for bounded weights.
A nearest-neighbor path that lies in is a short crossing of if and lie on opposite large faces of . More generally, we say that
(5.8) | a path crosses if some segment of is a short crossing of | |||
and neither endpoint of lies in . |
The second part of the definition ensures that genuinely “goes through” .
Let be the countable set of all triples where is an -box and and are two distinct points on the boundary of . A path has a -crossing if (5.8) holds and is the point where first enters and is the point through which last exits . (Then the short crossing of is some segment .) If crosses , then has a -crossing for some with uniquely determined by and .
Partition the set of all elements into subcollections such that within each all boxes are separated by distance . Any particular box appears at most once in any particular . The number of subcollections depends only on the dimension and the size parameter . The particular size of the separation of boxes in is taken for convenience only. In the end what matters is that the boxes are separated and that once is fixed, is a constant.
Let denote the -ball (diamond) of radius in , with (inner) boundary . The lemma below is proved in Appendix C.
Lemma 5.3.
By fixing and large enough and small enough as in (5.7), we can ensure the existence of constants such that, for all ,
(5.9) | every lattice path from the origin to has an index | |||
such that has at least -crossings | ||||
We turn to the modification argument for the strict concavity of claimed in Theorem 2.2.
5.2. Strict concavity
Let be the quantity in (5.5) in the definition of a black box. In addition to we consider two complementary assumptions on the weight distribution. Either the weights are unbounded:
(5.10) |
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 contains a point that satisfies | |||
If we can choose . Let be the constant that appears in Theorem 2.1 and in Theorem A.1, also equal to the constant in (4.10) for the shifted weights .
Theorem 5.4.
Assume and (2.6), in other words, that weights are nonnegative and the infimum is subcritical. Furthermore, assume that one of these two cases holds:
- (a)
-
(b)
Bounded case: the weights satisfy (5.11).
Then there exists a finite positive constant and a function of such that the following bounds hold for all and all :
Condition (2.7) with guarantees that the expectation above is finite (Lemma 2.3 in [2]). This together with Lemma A.3 then implies that is finite for . Since , (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 and whose choice is governed by the support of through the lemma below.
Lemma 5.5.
Fix reals and . Then there exist arbitrarily large positive integers such that
(5.14) |
holds for sufficiently small real .
Proof.
It suffices to show the existence of arbitrarily large positive integers that satisfy the strict inequalities
(5.15) |
and then choose small enough. Let and choose an integer . Then for each there exists such that
(5.16) |
and and can be taken arbitrarily large. Rearranging (5.16) and remembering the choice of gives
To get (5.15), take and large enough to have . ∎
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 be a geodesic for . When geodesics are not unique, will be chosen in particular measurable ways that are made precise later in the proofs. Assume that so that Lemma 5.3 applies with . The event in (5.9) lies in the union
By (5.9), there is a nonrandom index such that
(5.17) |
Define the event
(5.18) |
Consequently
(5.19) |
Turn this into a lower bound on the expected number of events, with a new constant :
(5.20) |
The next Stage 2 of the proof shows that, after a modification of the environment on a black box, the geodesic encounters a detour whose weights are determined by the modification. By this we mean that the geodesic runs through a straight-line -step path segment of the form parallel to an integer unit vector , with some initial vertex . A detour associated to is a path that shares both endpoints with and translates the -segment by steps in a direction perpendicular to : so for some integer unit vector ,
(5.21) |
In particular, and are edge-disjoint while they share their endpoints.
The rectangle enclosed by and will be called a detour rectangle. Its relative boundary on the plane spanned by is . Throughout we use superscripts and to indicate objects associated with the two portions of the boundaries of detour rectangles . Figure 5.1 illustrates.
Stage 2 is undertaken separately for bounded and unbounded weights.
Stage 2 for bounded weights.
Lemma 5.6.
Assume (5.11). For there exist nondecreasing sequences with the following properties:
(5.22) |
(5.23) |
(5.24) |
(5.25) |
Proof.
If then let for all and . So suppose .
Let . For define inductively in the interval so that for all . This can be done as follows. Let be an atom of in if one exists. If not, the c.d.f. of is continuous in this interval and we take to be a point of strict increase which must exist.
Then and thus . Furthermore, for all because . Pick so that . Define a nondecreasing sequence by . Since we have . ∎
We fix various parameters for this stage of the proof. Fix and determine by applying Lemma 5.5 to to have
(5.26) |
Since from below, we can fix large enough and small enough so that
(5.27) |
holds for . Note that this continues to hold if we increase to take closer to or decrease .
Take large enough, small enough, and large enough so that the crossing bound (5.9) of Lemma 5.3 is satisfied for the choice . Drop from the notation and henceforth write .
Shrink further so that
(5.28) |
and
(5.29) |
The construction to come will attach detours to edges of cubes. The number of such attachments per edge is given by the parameter
Let be an even positive integer and define two constants
(5.30) |
and
(5.31) |
We have the lower bound
Fix large enough so that
(5.32) | ||||
(5.33) | ||||
(5.34) |
Note that after fixing , (5.32) and (5.33) remain true as we shrink and (5.34) remains true with in place of as we increase towards .
As the last step fix so that is a multiple of and large enough so that
(5.37) |
Increasing may force us to take closer to 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 -box . Let be a box inside that is tiled with cubes of the form where is the lower left corner of the cube and the side-length comes from (5.35). The cubes are nonoverlapping but neighboring cubes share a -dimensional face. Then, where is the number of cubes required to tile . Inside box , is surrounded by an annular region whose thickness (perpendicular distance from a face of to ) is in the direction where has width and in the other directions.
A boundary edge of a cube is one of the line segments (one-dimensional faces) of length that lie on the boundary .
Attach -detours along each of the boundary edges of the tiling so that the -path is on the boundary edge and the detour 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 then the detour lies on the 2-dimensional face for some (in other words, the detour points into a positive coordinate direction). See Figure 5.2.
Place detours on each boundary edge of the tiling so that the detours are exactly distance apart from each other and a detour that is right next to a corner vertex of the tiling is exactly distance from that vertex. This is consistent with the definition of in (5.35).
Since 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 -box , for let denote the union of the -dimensional faces of the cubes tiling . Let be the union of (the boundary edges) and the detours attached to the boundary edges.
We describe in more detail the structure of the detours on the two-dimensional faces inside a particular . Let be a two-dimensional face. For simplicity of notation suppose . Assume without loss of generality that the boundary edge has its detours contained in . For define the th detour rectangle:
The subscript identifies these detour rectangles as attached to the southern boundary of . Similarly, if the detour rectangles attached to the western boundary of lie in , we denote these by .
For a label , let be the portion of the boundary of in the interior of . is the detour path of edges. Let be the portion of the boundary of that lies on the boundary of . is a straight path of edges, the path bypassed by the detour. Let
(5.38) |
See Figure 5.3. Let (resp. ) be the union of all (resp. ) as ranges over all the two-dimensional faces that lie in . The union of all equals as already defined above.
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) |
and then order the paths lexicographically. Here stands for a missing step. So if extends with one or more steps, then in lexicographic ordering. Recall the choice of index in (5.17).
Lemma 5.7.
Fix . There exists a unique geodesic for that satisfies the following two conditions.
-
(i)
For every -box and points the following holds: if both or and (or vice versa), and if every edge of lies in but not in , then there is no geodesic between and that remains in , uses only edges with strictly positive weights, and uses at least one edge in .
-
(ii)
is lexicographically first among all geodesics of 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 -geodesic of maximal Euclidean length. For the purpose of this proof consider as an ordered sequence of vertices and the edges connecting them.
Consider in order each segment that violates point (i). When this violation happens, there is a particular -box such that and there is an alternative geodesic that uses only edges with strictly positive weights and uses at least one edge in . Replace the original segment with .
Since we replaced one geodesic segment with another, . 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 . Since 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 . Since the replacement inserted into the geodesic at least one new edge from , has strictly fewer edges in than .
The new segment may in turn contain smaller segments that violate point (i). Replace each of these with alternative segments . Continue like this until the entire path segment between and 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 happens entirely inside the particular -box , does not alter the endpoints of the original segment, and does not alter the other portions and of the geodesic because each replacement step produced a self-avoiding geodesic.
Define the event
(5.40) | ||||
A key consequence of the definition of the event is that, by (5.27), the boundary paths and of all detour rectangles in satisfy
(5.41) |
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) |
depends on and the probabilities of the events on that appear in . In particular, does not depend on .
Our point of view shifts now to the implications of the event for a particular .
Let be a self-avoiding path in . Then if ,
(5.43) |
where came from (5.30). The main term on the right of (5.43) contains the weights of the -paths of detours and -gaps completely covered by , and accounts for the partially covered pieces at either end of .
We say that a point is associated with a boundary edge of a cube if either or lies in one of the detour rectangles attached to the edge . We say that points are -related if they are each associated to boundary edges and such that every point on can be connected to every point on by an -path that remains entirely within . Recall that an -path satisfies .
Lemma 5.8.
Let . Let be two -related points. Suppose a geodesic between and lies within . Then there exists a geodesic between and that stays within and uses at least one edge in .
Proof.
There are two cases:
-
(A)
are connected by an -path inside .
-
(B)
cannot be connected by an -path that remains entirely inside .
In case (A), any -path inside takes weights that are at most and any path inside takes weights that are at least . Since we assume the existence of a geodesic between and that lies entirely inside , we see that there must exist a geodesic that remains entirely within .
In case (B), suppose is a self-avoiding path between and that lies outside . Construct a path from to by concatenating the following path segments: using at most steps, connect to the closest point on the boundary edge that is associated with; using at most steps, connect to the closest point on the boundary edge that is associated with; connect to with an -path in . We show that , thus proving the lemma.
We argue that
(5.44) | uses at least edges in . |
Indeed, observe that and cannot both be on nor both in the same detour rectangle , for otherwise we would be in case (A). On the other hand, if is in a detour rectangle and is on , then is an -path that connects to . If in this case , then it must be the case that . But then in this case is an -path from to and we are again in case (A). The symmetric case of and in a detour rectangle is similar. Lastly, if and belong to different detour rectangles, then the segment of that connects the two rectangles must be of length at least , the distance between two neighboring detours.
We have verified (5.44). From (5.44) and comes the lower bound
The edges in all have weight at most . Furthermore, and all the edges along have weight no larger than . This gives the bound
Since connects to and the weights along are at least ,
Lemma 5.9.
Let . Suppose are not -related and that they are connected by a path that remains entirely in . Then
Proof.
Inspection of Figure 5.4 convinces that any two points such that must be -related. Thus and by assumption it uses only weights . ∎
Lemma 5.10.
Let and . Assume that either both or that and . Let be a geodesic between and . Assume that the edges of lie entirely outside . Then either there is a geodesic between and inside that uses at least one edge in or
(5.45) |
Proof.
If reaches the boundary (in either case of ) then must travel through and consequently . The other possibility is that stays inside . If and are -related then Lemma 5.8 gives a geodesic in that uses an edge in . If and are not -related, Lemma 5.9 gives . The last equality of (5.45) is from (5.35). ∎
Henceforth we often work with two coupled environments and . Quantities calculated in the environment will be marked with a star if is not explicitly present. For example, denotes the passage time between and in the environment .
Recall the event defined in (5.18).
Lemma 5.11.
Let and be two environments that agree outside and satisfy and . Then there exists a self-avoiding path from to such that
Proof.
Since box is black on the event ,
The bound above comes from (5.5), on account of these observations: regardless of whether exits , there is a segment inside of length , and furthermore contains a short crossing of that has length at least .
Define a path from to in as follows. Let be an -path from to some point . Similarly, let be an -path from to some . These paths satisfy . Let be a shortest path from to that remains in . Since , . (To go from to along use steps to go from and to the nearest vertices and in , respectively, and an -path along will take steps.)
Let be the concatenation of , and . Define as the concatenation of , , and . The next calculation uses (5.43), (5.37) and (5.31), and the facts that outside , and .
In the first inequality, bounds the time spent on and and the remaining negative terms bound the passage time of . The lemma is proved. ∎
Henceforth we assume that and . Let be the geodesic from to in the -environment specified in Lemma 5.7. By Lemma 5.11,
(5.46) |
This implies that must use edges in because and agree outside , while on edges in .
Let be the first vertex of in , the first vertex of in , and the last vertex of in . Decompose the path segment between and into excursions () as follows: each excursion begins with a nonempty segment of edges inside , followed by a nonempty segment of edges outside . The excursions begin and end at a vertex in , while the last excursion begins in and ends at the vertex where exits . Figure 5.5 illustrates.
By , (5.27) and (5.28), and hence for all edges . Then condition (i) of Lemma 5.7 ensures that those portions of the segments that connect to itself or to inside cannot be replaced by segments that use edges in . Therefore these segments obey bound (5.45). This gives the last inequality below:
Since the maximal side length of is , and can be connected with a path such that Since is optimal, , and therefore
(5.47) |
Using (5.46), and that outside while on ,
Then some excursion must satisfy
(5.48) |
The second inequality comes from (5.37) and (5.47). The only positive contributions to can come from , the segment of in . Since is black, for all edges . Therefore the number of edges satisfies . From this and (5.36)
(5.49) |
The next lemma ensures that the path segment goes through the -path of at least one -detour.
Lemma 5.12.
Let and be two environments that agree outside and satisfy and . Let be the geodesic for chosen in Lemma 5.7. Then there exists a detour rectangle in with boundary paths such that follows and does not touch , except at the endpoints shared by and .
Proof.
By construction, the portion of has a continuous path segment of length in . This forces to enter at least three detour rectangles, because these rectangles are apart along and the path can use at most edges in a gives detour rectangle. Let be a middle rectangle along this path segment, in other words, one that is both entered and exited, and such that covers two -segments on that connect to some neighboring detour rectangles. (Recall Figure 5.2.) We can assume without loss of generality that lies in the -plane and that it is attached to a boundary edge of some that lies along .
Let and be the boundary paths of and and their common endpoints. Let denote the segment of between and . For ease of language assume that visits first and then . We show that by showing that all other cases are strictly worse.
Definition (5.40) of and inequality (5.27) imply and rule out the case . If coincides with neither nor , there are points and on such that visits in this order and lies in the interior .
If and lie on the same or on adjacent sides of , the -path from to along has smaller weight than .
Suppose and lie on opposite -sides of . Then
The term is a lower bound on . The strict inequality comes from (from (5.27)) and because the segments and are not degenerate paths. This is the case because no edge connects the interior to either or .
The remaining option is that and lie on opposite -sides of . Let be the first point at which leaves , and let be the point of first return to .
Case 1. Suppose lies on the -side of and (Figure 5.6). Fix coordinates as follows: is at the origin, , and . Then,
Combine the above with and develop further:
The last inequality is from (5.27). Thus cannot cross the interior of from to .
Case 2. Suppose and lies on the -side of , so that , and . Then
The last strict inequality is from (5.29). Thus it is strictly better to take from to .
In conclusion, does not coincide with , nor does visit the interior of the detour rectangle. The only possibility is that .
It remains to argue that does not touch except at the endpoints and when it goes through . Suppose on the contrary that visits a vertex on . This has to happen either before vertex or after vertex . The two cases are similar so suppose is visited before . Then, by the choice of the detour rectangle , the segment contains an -segment on that ends at . Hence by the definition of and (5.33),
However, is an upper bound on the passage time of the path from to along , which is then strictly faster than . Since must be a geodesic, the supposed visit to cannot happen. ∎
Stage 2 for unbounded weights.
In the unbounded weights case we construct first the detour for a given triple and then the good event . Given any , the construction below can be carried out for all large enough . We label the construction below so that we can refer to it again. Figure 5.7 gives an illustration.
Construction 5.13 (The detour for the unbounded weights case).
Fix two unit vectors and among perpendicular to each other so that the point lies in . Hence also the rectangle of size with corners and lies in . Switch the labels and if necessary to guarantee that does not lie in the set
(5.50) |
The two versions of obtained by interchanging and have only in common, so at least one of them does not contain .
From there is a self-avoiding path to that stays inside and does not intersect . The existence of such a path and an upper bound on the minimal length of such a path can be seen as follows.
-
(i)
If does not lie on the plane through spanned by , take a minimal length path from to that begins with a step perpendicular to this plane. Unit vector is chosen so that . This path will not return to the plane and hence avoids . The length of this path is at most . This is because a possible -avoiding route to takes first the -step from , then -steps to , and from a minimal length path to . A path from to includes a -step, hence the distance from to is .
-
(ii)
Suppose lies on the plane through spanned by . Then we move on this plane from to and take care to avoid . First define the minimal -avoiding path from to in steps, and a minimal -avoiding path from to in steps. (We may be forced to pick between depending on which side of the point lies.) Let be a closest point to on (possibly ). The -avoiding self-avoiding path from to then goes first to along or and from there takes a minimal length path to . The length of this path is at most .
Using the construction above, fix a self-avoiding path in from to that begins with -steps from to , avoids after that, and has
(5.51) |
Let be the -directed straight line segment of length from to . Let be the detour of length between the endpoints and defined as in (5.21). The two endpoints of lie on but is edge-disjoint from . This completes the construction of the detour.
Let be given. By assumption (5.10) we can choose in the support of so that . Choose to satisfy (5.14).
Fix an element for a while. Define the following event that depends only on the weights in . Constants and are from definition (5.4)–(5.5) of a black -box .
(5.52) | ||||
By (5.14), on the event ,
(5.53) |
Once has been fixed, then up to translations and rotations there are only finitely many ways to choose the points and on the boundary of and the paths constructed above. Thus
(5.54) |
depends on and the probabilities of the events on that appear in . In particular, does not depend on .
On the event of (5.18), crosses , is the point of first entry into and the point of last exit from . Hence on this event we can define as the self-avoiding path from to obtained by concatenating the segments , , and . For future reference at (5.57), note that is edge-disjoint from .
Lemma 5.14.
Let and be two environments that agree outside and satisfy and . Then is a geodesic for . Furthermore, if was chosen to be a geodesic of maximal Euclidean length for , then is a geodesic of maximal length for . The same works for minimal length.
Proof.
Since box is black on the event ,
(5.55) |
The bound above comes from (5.5), on account of these observations: regardless of whether exits , there is a segment inside of length , and furthermore contains a short crossing of that has length at least .
From ,
Before the last inequality above, are constants determined by the quantities in parentheses. The last inequality is then guaranteed by fixing large enough relative to and these other constants. Observation (5.7) is used here.
Outside the weights and agree, and the segments and agree and lie outside . Hence the inequality above gives and thereby, for any geodesic from to in environment ,
(5.56) |
This implies that every geodesic must enter since otherwise
contradicting the optimality of under .
If , then must use an edge in with weight . Then by property (5.4) of a black box , . Since and agree on , we get
contradicting (5.56). Consequently . Part of event is that . Thus must both enter and exit . As a geodesic does not backtrack on itself. Hence it must traverse the route between and . By (5.53) is better under than , and hence .
Outside , under both and since they agree on , is an optimal union of two paths that connect the origin to one of and , and the other one of and to . This concludes the proof that is a geodesic for .
Suppose is a geodesic of maximal Euclidean length under but under there is a geodesic strictly longer than . The argument above showed . Hence outside , must provide an -geodesic from or to one of or that is longer than that given by . This contradicts the choice of as a maximal length geodesic, again because and agree on . 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 for . In the bounded weights case, let be the geodesic specified in Lemma 5.7. In the unbounded weights case, let be the unique lexicographically first geodesic among the geodesics of maximal Euclidean length. Let . For -boxes define the event
(5.57) | inside edge-disjoint path segments and that share both endpoints | |||
and satisfy , is a self-avoiding path, | ||||
Couple two i.i.d. edge weight configurations and so that for (that is, at least one endpoint of lies outside ) and so that the weights and are independent.
Lemma 5.12 for bounded weights (with ) and Lemma 5.14 for unbounded weights imply that
In particular, by inequalities (5.41) and (5.53), implies required for , where denotes passage time in the environment .
By the independence of and , and then by (5.42) for bounded weights and by (5.54) for unbounded weights,
(5.58) |
Since we have arranged the boxes in the elements separated, we can define a self-avoiding path from to by replacing each segment with the segment in each box for which event happens.
Reduce the weights on each edge from to . By the definition of , the -passage times of the segments and obey this inequality:
Consequently, along the entire path , the replacements of with reduce the -passage time by at least . We get the following bound:
(5.60) | ||||
The case distinction above comes because in the unbounded case by our choice of , while in the bounded case our choice is different, but any geodesic satisfies . Note that the inequality above does not require that be a geodesic for , as long as is self-avoiding.
In order to take expectations in (5.60) we restrict to which guarantees that is finite, even if so that weights can be negative (Theorem A.1 in Appendix A). By Lemma 2.3 in [2], moment bound (2.7) with is equivalent to the finite expectation for all . The inequalities above then force and to have finite expectations. Apply (5.59): in the bounded weights case
while in the unbounded weights case is replaced by . 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.
-
(i)
Zero is an atom: and .
-
(ii)
The weights are unbounded and there exist strictly positive integers and and atoms (not necessarily all distinct) such that
(6.1) -
(iii)
The weights are bounded and there exist strictly positive integers and and atoms such that .
Theorem 6.2.
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 . 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 has at least two strictly positive atoms.
A sufficient condition for (6.1) is the existence of two atoms in such that is rational. This is exactly the assumption on the atoms in case (iii) of Assumption 6.1. If has exactly two atoms in and no others, then (6.1) holds if and only if is rational.
With more than two atoms, rational ratios are not necessary for (6.1). For example, if is irrational and are atoms, then (6.1) is satisfied and the ratios 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.
Proof of Theorem 6.2.
We prove the theorem by considering each case of Assumption 6.1 in turn.
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 black if (5.5) holds. Fix large enough and small enough. Consider points with large enough so that the Peierls estimate (5.9) is valid for .
Let be the unique geodesic for that is lexicographically first among the geodesics of minimal Euclidean length. For this purpose order in some way, for example as in (5.39). The index and the event are defined as before in (5.17) and (5.18), and estimate (5.20) holds. Let be the event that all edge weights in are zero and
Given an -box , define edge weight configuration by setting for (that is, at least one endpoint of lies outside ) and by resampling independently. Then has the same i.i.d. distribution as the original weights .
Lemma 6.4.
On the event , every geodesic from to in the environment uses at least one edge in .
Proof.
On the event , goes through and . Let be an arbitrary path from to that remains inside and define as the path from to obtained by concatenating the segments , , and . Then on the event ,
The justification for the last inequality was given below (5.55).
Outside weights and agree, and the segments and agree and lie outside . Hence the inequality above gives and thereby, for any geodesic from to in environment ,
(6.3) |
This implies that every geodesic must use at least one edge in . For otherwise
contradicting the optimality of for . ∎
For -boxes such that define the event
(6.4) | inside path segments and that share both endpoints | |||
and satisfy , is a self-avoiding path, | ||||
In particular, on the event , replacing with creates an alternative geodesic.
By Lemma 6.4, holds on the event . This is seen as follows. Let be the lexicographically first geodesic of minimal Euclidean length in environment . By Lemma 6.4, uses at least one edge in . Let be the first and the last point of in . Since ensures that all edges in have zero weight and is a minimal length geodesic, the segment must be a path of length from to inside . Now take and let be any other path inside from to that takes more than the minimal number of steps. By the choice of and , the other portions and of the geodesic lie outside , and consequently does not touch these paths except at the points and .
By the independence of and ,
(6.5) | ||||
Let be the number of for which occurs. By (5.20), for another constant ,
(6.6) |
By Proposition 4.7(1) of [2], under the assumption , for any there exists a finite constant such that for all ,
(6.7) |
By Lemma 2.3 in [2], under assumption (2.7) there exists a finite constant such that for all
(6.8) |
An obvious upper bound on is the number of edges on the geodesic . Let be the power for which (2.7) is assumed to hold and its conjugate exponent. Then, by a combination of (6.6), (6.7) and (6.8),
From this we get the bound
Since we have arranged the boxes in the elements separated, we can define a self-avoiding path from to by replacing each segment of with the segment in each box for which event happens. This path has the same passage time and hence both and are geodesics. By the construction, the numbers of edges on these paths satisfy . Thus we get these inequalities between the maximal and minimal geodesic length:
and then
(6.2) has been proved.
By assumption (6.1) we can fix large enough so that, for i.i.d. copies of the edge weight ,
(6.9) |
Apply Construction 5.13 of the detour in an -box with given boundary points and , to define paths , and in with and . Define the event that depends only on the weights in :
(6.10) | ||||
By (6.9), unbounded weights, and the detour construction, there exists a constant such that for all triples .
The steps follow those of the proof of Theorem 5.4 and the proof of Case (i) of Theorem 6.2. First sample , and then define by setting for and by resampling independently. Let be a self-avoiding geodesic of minimal Euclidean length for . On the event define the path from to by concatenating the segments , , and .
Lemma 6.5.
When is fixed large enough, on the event the path is a self-avoiding geodesic of minimal Euclidean length for .
Proof.
As before, since box is black on the event ,
Then by ,
Before the last inequality above, is a constant determined by the quantities fixed thus far in the proof. The last inequality is then guaranteed by fixing large enough relative to these other constants. Outside weights and agree, and the segments and agree and lie outside . Hence the inequality above gives and thereby, for any geodesic from to in environment ,
(6.11) |
As explained below (6.3), this implies that every geodesic must enter .
If , then must use an edge in with weight . Then by property (5.4) of a black box , . Since and agree on , we get
contradicting (6.11). Consequently . As a geodesic does not backtrack on itself. Hence it must traverse the route between to , either via or via with replaced by . By (6.10) so there is no travel time distinction between the two routes between and .
Since and agree on , is an optimal union of two paths that connect to one of and , and to the other one of and . Thus is a geodesic for .
The argument above showed that every geodesic of goes from to utilizing edges in and otherwise remains outside . If there were a geodesic strictly shorter than , would have to use an alternative shorter geodesic path between and or between and . This contradicts the choice of as the shortest geodesic. ∎
Define as in (6.4) above. By Lemma 6.5, holds on the event . The proof of this case is completed exactly as was done in the previous case from equation (6.5) onwards.
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 and all . Since the case of a zero atom has been taken care of, we can assume that these atoms 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 -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) |
The only change required in the proof of Lemma 5.6 is that induction begins with , after the case has been taken care of.
The construction of , , , and in each black box goes exactly as before around (5.38). Let be the and boundary path segments of the detour rectangles constructed in the box . In particular,
Define the event
(6.13) | ||||
and |
The condition is implied by the conditions before it. It is stated explicitly merely for clarity. The condition can be imposed because (i) for it follows from (recall from (5.27) that ), and (ii) for edges we can use the strictly positive atoms . Again for a constant .
As before, given an -box we work with two environments and that agree outside . Let be the 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 in whose segment in satisfies (5.49). The previous Lemma 5.12 is then replaced by the next lemma.
Lemma 6.6.
Assume and . Then there exist three path segments in with the same endpoints and such that the following holds:
-
(i)
the pair forms the boundaries of a detour rectangle,
-
(ii)
, and
-
(iii)
replacing in with either or produces two self-avoiding geodesics for .
Proof.
As in the proof of Lemma 5.12, has a segment between the common endpoints and of the boundary paths and of some detour rectangle in . We show that can be redirected to take either or , by showing that (i) cannot be strictly better than or and (ii) replacing with or does not violate the requirement that a geodesic be self-avoiding.
Suppose . Then there are points and on such that visits in this order and the edges of lie in the interior . Recall that on the event , the weights on are at most while the weights in the interior are at least .
The points and cannot lie on the same or on adjacent sides of since the -path from to along has no larger weight than .
Suppose and lie on opposite -sides of . Then
So we can do at least as well by picking or .
The remaining option is that and lie on opposite -sides of . Let us suppose that is the first point at which leaves and the first return to .
For this argument we use the most restrictive assumption that there are two atoms such that , with weights on edges and on edges .
Case 1. Suppose lies on the -segment of and . (See again Figure 5.6.) We can assume that is at the origin, , and . Then,
From and the assumptions and we deduce:
In other words, we can do better by taking from to .
Case 2. Suppose and lies on the -segment of so that and . Then,
This time it is better to take from to .
We have shown that the passage time is not made worse by forcing to take or . Suppose doing so violates self-avoidance of the overall path from to . Then we can cut out part of the path, and the removed piece includes at least one edge of either or . The assumption implies that for these edges. Consequently the original passage time could not have been optimal. ∎
The event earlier defined in (6.4) has to be reworded slightly for the present case. Let be the geodesic chosen in Lemma 5.7.
(6.14) | inside path segments , and that share both endpoints | |||
and satisfy , both and | ||||
are self-avoiding paths from to , | ||||
It is of course possible that agrees with either or . By Lemma 6.6, holds on the event .
Now follow the proof of the previous case from equation (6.5) onwards. Again, since the boxes in the elements are separated, we can define two self-avoiding paths and from to by replacing each segment of with the (respectively, ) segment in each box that appears among and for which event happens. Then both and are self-avoiding geodesics for .
By the construction, the Euclidean lengths of these paths satisfy where is again the number of for which occurs. Hence
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 and is the constant specified in Theorems 2.1 and A.1.
Theorem 7.1.
Note that the theorem does not rule out a linear segment of immediately to the left of which happens if for some . But this does force and thereby a singularity at .
Proof.
We start by deriving the last statement of strict concavity from (7.1). Suppose that for some . Then by concavity must be affine on the open interval : and for . 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 , , the full measure event specified in Theorem A.1, and . Take limits (2.8) in the extremes of (2.13), limits and in the middle of (2.13), and then let . This gives
(7.2) |
To prove (7.1), consider first the case where or but . The hypotheses of Theorem 5.4 are satisfied for the shifted weights . In particular, the extra assumption (5.11) of the bounded weights case that requires the existence of a positive support point close enough to the lower bound is valid because either or but is not an atom.
From Theorem 5.4 applied to the shifted weights we take the conclusion (5.13) which is valid in both cases of the theorem:
(7.3) |
The constant given by the theorem depends now also on .
Corollary 7.2.
Proof.
Proof of Theorem 2.3.
Proof of Theorem 2.5.
Using Proposition 4.4(i), the continuity of the shape functions and on , and from Corollary 7.2 above, choose constants small enough so that for any ,
(7.7) |
From (4.8) or (A.2) pick finite deterministic and random such that
(7.8) |
Let . Increase if necessary so that . Let . Increase if necessary so that (i) , (ii) works in (B.1) for , and (iii) satisfies the FPP shape theorem ([2, p. 11], also (A.3))
(7.9) |
Proof of Theorem 2.11.
Proof of Theorem 2.16.
Part (iii). Begin by noting that differentiability of is equivalent to differentiability of and on an open interval a differentiable convex function is continuously differentiable.
Since 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 . General convex analysis gives the equivalence
By the strict concavity of , a given lies in for a unique , and hence the subdifferential consists of a unique value . This implies that is differentiable at .
Continuous differentiability of for now follows from Proposition 4.4. Namely, for , which we now know to be a nondegenerate interval, and their common left -derivative vanishes at the minimum . On , is constant and hence connects in a fashion to the part on .
If then necessarily .
The remaining claims follow if we assume and show that
(7.10) |
It suffices to treat since close enough to the boundary of by part (ii).
Take and rewrite the ratio above as
Thus by the duality in Theorem 2.17, (7.10) is equivalent to
(7.11) |
By concavity, the ratio in (7.11) above is a nonincreasing function of . Hence if (7.11) fails, there exists such that, and ,
It then follows from the duality ((2.41) or (2.44)) that
This contradicts the strict concavity of . (7.10) has been verified. ∎
7.2. Nondifferentiability
Proof of Theorem 2.6.
Proof of Theorem 2.7.
Let be two atoms of in . Fix an arbitrary and then pick so that
For let
Then and are atoms of such that
The other hypotheses of Theorem 2.6 are inherited by and so the conclusions of Theorem 2.6 hold for all . In particular, since , has a corner at if and only if has a corner at .
No point of is farther than from the nearest . We get the dense set by combining the collections for all . ∎
Appendix A First-passage percolation with slightly negative weights
This appendix extends the shape theorem of standard FPP to real-valued weights under certain hypotheses. The setting is the same as in Section 2.1. As before, denotes i.i.d. copies of the edge weight . Assumption (2.7) is reformulated for positive parts as
(A.1) |
Passage times are defined as in (2.2) and now the restriction to self-avoiding paths is essential.
Theorem A.1.
Assume , (2.6), and (A.1) (equivalently, (2.7)) with . Then there exist
-
(a)
a constant determined by the distribution of the shifted weights ,
-
(b)
for each real , a positively homogeneous continuous convex function , and
-
(c)
an event of full probability,
such that the properties listed below in points (i)–(iii) are satisfied.
-
(i)
For each and the following pointwise statements hold. For each , is finite and has a geodesic, that is, a self-avoiding path from to such that . There exist a deterministic finite constant and an -dependent finite constant such that
(A.2) The shape theorem holds, locally uniformly in the shift : for any in ,
(A.3) -
(ii)
For each the following statements hold. for all . For any sequence with , the convergence holds almost surely and in .
-
(iii)
The shape function satisfies these Lipschitz bounds for shifts and all :
(A.4) For , and for all .
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 be a probability measure invariant under a group of measurable bijections. Let be a nonnegative random variable such that . Then
(A.5) |
Proof.
The conclusion is equivalent to as . Apply Borel-Cantelli with the estimate below for :
Because the inequalities in the proof can be reversed with different constants, an i.i.d. example shows that moments does not suffice for the conclusion.
Let denote the negative part of a real number. Following [18], define the random variable
(A.6) |
We first prove a moment bound for the shifts of that was used in the concavity result of Section 5.2.
Lemma A.3.
Proof.
The next item is a shape theorem whose hypotheses are stated in terms of the random variable of (A.6).
Theorem A.4.
Let be i.i.d. real-valued weights.
(i) Assume (A.1) with and that the random variable from (A.6) satisfies . Then is a finite integrable random variable for all . There exists a non-random positively homogeneous continuous convex function such that for any sequence with ,
(A.7) |
(ii) Assume furthermore (A.1) with and . Then the following hold with probability one:
(A.8) |
and for all and any sequence such that
(A.9) |
Proof.
Let . Consider two paths and . Their concatenation may fail to be self-avoiding. Choose a point belonging to both paths such that erasing the portion of from to (denoted by ) and erasing the portion of from to (denoted by ) leaves a self-avoiding path from to . (If the concatenation was self-avoiding to begin with, then .) Note that and are self-avoiding paths. This implies that
Taking infimum over and gives Rearranging, we get
(A.10) |
To apply the subadditive ergodic theorem, we derive a moment bound.
Let . Take any -path from to (where ) and use the subadditivity of the passage times in weights to write
Since are all identical,
(A.11) |
Assumption (A.1) with implies that (Lemma 2.3 in [2]). By the assumption ,
Standard subadditivity arguments give the existence of a positively homogeneous convex function such that for all and with , almost surely and in ,
(A.12) |
and does not depend on the choice of . The assumption allows us to drop the term from above. The first inequality of (A.10) gives .
Fix . Use subadditivity (A.10) to write
Switching and gives a complementary bound and so
(A.13) |
By (A.11)
(A.14) |
Now take and such that and are both in and apply the above to get
Divide by and take to to get
(A.15) |
As a Lipschitz function extends uniquely to a continuous positively homogenous convex function .
Fix and a sequence in such that . Fix . Take such that . For let . By (A.14),
Take to get
Let to get (A.7). This completes the proof of part (i).
Now strengthen the assumptions to and (A.1) with . For (A.8) we follow the proof of the Cox-Durrett shape theorem presented in [2, Section 2.3].
Let be the full probability event on which (A.12) holds for all . By Lemma 2.22 and Claim 1 on p. 22 of [2], under (A.1) there exists a finite positive constant and a full probability event such that for any and , there exists a strictly increasing random sequence such that as and
(A.16) |
Here, is the first-passage time from to under the weights . The results from [2] apply because these weights are strictly positive and satisfy (A.1) with .
Let be the full probability event on which (A.5) holds for the random variable of (A.6). We show that (A.8) holds for each fixed . Let be an -dependent sequence such that and
(A.17) |
By passing to a subsequence we can assume with . Let satisfy and pick such that . Choose in (A.16) for . For each take such that
(A.18) |
Abbreviate . There exists -dependent such that for all , . Triangle inequality:
Use (A.13), (A.16) applied to , and take :
As the right-hand side converges to . Letting then proves that as . Since is continuous and homogeneous, we also have . Now (A.8) follows from (A.17).
Remark A.5.
In the last inequality of the proof above can be replaced by a smaller term as follows. First, fix a rational . Take to be large enough so that for all , , as before, but also
Take (still with ) so that . Now we have
Consequently,
Thus, instead of (A.5) one now needs the hypothesis that for any fixed ,
In our application to the proof of Theorem A.1 the random variable has all moments, so we do not pursue sharper assumptions on than those stated in Theorem A.4.
Proof of Theorem A.1.
The constant in point (a) is taken to be the constant in the bound (4.10) for the shifted weights . Then by Lemma A.3, has all moments for all . Hence we can apply Theorem A.4 to the shifted weights to define the shape functions whose existence is asserted in point (b). We specify the full-probability event for point (c) in the course of the proof.
Proof of part (i). Start with the obvious point that for any particular self-avoiding path from to . By bound (4.10) there exists a full probability event and a finite random variable such that on the event , every self-avoiding path from the origin such that satisfies the bound . Then for any shift these paths satisfy
(A.19) |
From this we conclude that, for any , , and any path ,
(A.20) |
Thus the infimum that defines in (2.2) cannot be taken outside a certain -dependent finite set of paths. Consequently on the event a minimizing path exists and both and are finite for all and .
Next, shrink the event (if needed) so that for the shape theorem (A.8) is valid for the weights . Then we can increase and pick a deterministic positive constant so that whenever . By monotonicity for all whenever . If necessary increase so that . Then by (A.20), when and , a self-avoiding path between and that satisfies
cannot be a geodesic for . We conclude that for ,
Since is nonincreasing in (Remark 2.4(iii)), we can extend the bound above to all in the form (A.2).
By taking advantage of (A.2) now proved, we get these Lipschitz bounds: for all , and , and with denoting a geodesic of ,
(A.21) | ||||
The last equality defines the constant which is nonincreasing in .
We establish the locally uniform shape theorem (A.3). Let be a countable dense subset of . Shrink the event further so that for the shape theorem (A.8) holds for the shifted weights for all . By passing to the limit, (A.21) gives the macroscopic Lipschitz bounds (A.4) for shifts in this countable dense set .
Let , , and in . Pick a partition so that each and . Fix a constant such that
Now for , , and , utilizing the monotonicity in of and and the Lipschitz bounds (A.21) and (A.4),
The shape theorem (A.3) has been proved.
Proof of part (ii). The integrability and convergence follow from Theorem A.4(i). The almost sure convergence comes from the homogeneity and continuity of and the shape theorem (A.3).
Proof of part (iii). We already established (A.4) for a dense set of shifts . Monotonicity of extends (A.4) to all shifts .
That follows from homogeneity. The final claim that for and follows from (A.19), which implies whenever . ∎
Appendix B Restricted path length shape theorem
This section proves the next shape theorem in the interior of for the restricted path length FPP processes defined in (2.24). As throughout, the edge weights are independent and identically distributed (i.i.d.) real-valued random variables, , the set of points reachable by -paths is defined by (2.23), and is the unit ball. We also write for i.i.d. copies of the edge weight .
Theorem B.1.
Assume and moment assumption (A.1) with . Fix . There exists a deterministic, continuous, convex shape function that satisfies the following: for each there exists an almost-surely finite random constant such that
(B.1) |
whenever , , and .
The shape theorem can be proved all the way to the boundary of . 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 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 .
Lemma B.2.
Assume the weights are arbitrary real-valued i.i.d. random variables. Let , and . Assume . Then for any real ,
(B.2) |
Proof.
It is enough to prove the lemma for . Then we can assume that is an even integer because otherwise . The reason that the case is also covered is that .
To prove (B.2) we construct a total of edge-disjoint paths in . Let and with cardinalities and , respectively. Enumerate these sets as and .
Suppose , in which case . For each let be the -path from to that takes the necessary steps in the order . Then for each let be the path that starts with repetitions of the pair and then follows . For let be the path that starts with a step, then repeats the pair times, then follows the steps of , and finishes with a step. Thus far we have constructed paths. For the remaining paths we distinguish two cases.
If we need only one more path . Take this to be the path that starts with a step, repeats the pair times, takes two steps, one step, follows the steps of , takes one step, two steps, and finishes with a step.
If , then for , let be the path that starts with a step, repeats the pair times, takes a step, follows the steps of , and ends with a step followed by a step. For the path is defined similarly, except that and are replaced by and , respectively.
One can check that the paths , , are edge-disjoint. From
(B.3) |
follows
(B.4) | ||||
If (and hence and ) then redo the last computation with the edge-disjoint paths , , that just repeat the pair . ∎
Below we use the condition that a rational point satisfies for a positive integer such that . When zero steps are admissible () this is of course trivial, and without zero steps () this can be achieved if is even. Therefore, one can take for example for such that . 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 . Fix and . Let be such that . Assume . Then the limit
(B.5) |
exists almost surely and in and does not depend on the choice of . As a function of , is convex. Precisely, if are such that and for some , then for any , for some and
(B.6) |
Remark B.4 (Conditions for finiteness).
Next, from convexity we deduce local boundedness and then a local Lipschitz property.
Lemma B.5.
Assume and (A.1) with . Fix and . There exist and a finite constant such that
(B.7) |
Proof.
Lemma B.6.
Assume and (A.1) with . Fix . There exist and a finite positive constant such that for both
Proof.
The next lemma is an immediate consequence of the local Lipschitz property proved in the previous lemma.
Lemma B.7.
Assume and (A.1) with . Then and extend to locally Lipschitz, continuous, convex functions on .
Before we prove the shape theorem we need two more auxiliary lemmas.
Lemma B.8.
Assume (A.1) with . Then there exists a finite constant such that
(B.8) |
Proof.
It is enough to work with since . It is also enough to work with fixed since the suprema in question increase as we increase and decrease .
Fix an integer such that
The strategy of the proof will be to bound by constructing edge-disjoint paths on the coarse-grained lattice to a point that approximates . An approach to finding such paths was developed in the proof of Lemma B.2.
Take large enough so that
(B.9) |
For each with pick so that . As in Lemma B.2, let be the number of such that and let . Following the construction in the proof of Lemma B.2 we can produce edge-disjoint nearest-neighbor paths , , on the coarse-grained lattice from to such that, in terms of the number steps taken on , has length for , has length for , and for , has length if and if .
Let denote the position (on the original lattice) of the path after steps (of size ). Let
(B.11) |
For each we have this identity:
This equation gives a way to decompose the steps from to so that we first go through the vertices and then use the last steps to go from to . We continue with this next bound:
(B.12) | ||||
Bound in (B.10) by dropping to turn (B.11) into this inequality (note that terms cancel):
Similarly,
Fix . For let . Define the events
and the event on which for all
(B.13) |
Then for large enough to satisfy (B.9) and ,
(B.14) | ||||
Here is the explanation for the inclusion above.
-
(i)
Further down the proof we add auxiliary paths around the -steps of the path . Because the first -steps share their initial point , their auxiliary paths would intersect and independence would be lost. The same is true for the last -steps that share the endpoint . Hence these special steps are handled separately.
The event takes care of the first step of the path which is either in the first sum on the right in (B.12), or in the second sum in case and the first sum is empty.
The event takes care of the last step to 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 of the union of shifts of : for and ,
and since is on an admissible path of length from to , it must be that .
-
(ii)
The event takes care of the path segment from to .
-
(iii)
On the complement of the first three unions on the right-hand side of (B.14) we have for each ,
Since , the left-hand side has at most 13 terms, which explains the bound on the right. Thus, if in addition , then event must occur.
By bounding the probabilities of the unions on the right of (B.14), we show next that for some fixed that does not depend on ,
(B.15) |
This will imply the conclusion (B.8) as we point out at the end of the proof.
By (B.2), is summable if (A.1) is satisfied with . Then as . Next, observe that
and hence
which goes to when if is summable. This is the case if (A.1) is satisfied with . The union over is controlled similarly.
It remains to control the probability of the union of the events in (B.14). For and , for each segment , bound both passage times and as was done in (B.3) by using independent auxiliary paths of the appropriate lengths. For each segment add the two upper bounds and denote the result by .
The terms for and were excluded from the events so that for distinct indices the auxiliary paths constructed around the segments stay separated. (We chose at the outset to guarantee this separation.) Replace the edge weights with to ensure that the upper bounds are nonnegative. After these steps, the left-hand side of (B.13) is bounded above by .
All the -terms have the same distribution as . As explained above, over distinct indices the random vectors are independent. For any particular , are i.i.d. and are i.i.d. because now we skip every other -step. See Figure B.1.
We derive the concluding estimate. Recall that
Let . Let denote the sum of independent copies of . Since the -terms are nonnegative we have
The same holds for the sum over odd . Thus we have
Take and use the fact that there are no more than points to get
The bound (B.4) can be utilized to show that each and thereby is square-integrable if (A.1) holds with . The above then converges to as . We have verified (B.15). The claim of the lemma follows:
Lemma B.9.
Assume (A.1) with . Then for any there exists a deterministic constant such that, with probability one for each , there exists a strictly increasing random sequence such that and for and
(B.16) |
Proof.
If take from Lemma B.8. Next suppose . Fix in . Apply Lemma B.8 to choose a finite constant such that
The ergodic theorem implies that with probability one, for each there exist infinitely many such that
Enumerate these ’s as a strictly increasing sequence . Then for -almost every
Consequently, converges to . ∎
We are ready for the shape theorem.
Theorem B.10.
Assume and (A.1) with . Fix . Let be a closed subset of . The following holds with probability one:
(B.17) |
Proof.
The proof follows steps similar to those of (A.3). We treat the case , the other case being a simpler version. Let be the full probability event that consists of intersecting the event on which (B.5) holds for all with the event in (B.8) and the events in Lemma B.9 for all rational in . Fix . We show that for this
(B.18) | |||
(B.19) |
Proof of (B.18). Fix (-dependent) sequences and that realize the on the left-hand side of (B.18). Since there are coefficients such that
(B.20) |
Pass to subsequences, still denoted by and , such that
(B.21) |
Let . We approximate with a rational point to which we can apply (B.5). Bound (B.18) comes by building a path from to a multiple of and by the subadditivity of passage times. Here are the details.
First, we dispose of the case where there are infinitely many for which for all . If this is the case, then going along a further subsequence we can assume that for all . Applying (B.5) with gives and since for all we see that the on the left-hand side of (B.18) is . We can therefore assume that for each there exists some such that or . Consequently, if we let denote the set of indices for which or for infinitely many , then .
Let
(B.22) |
with the convention that , which takes care of the case for all . Let be a rational in . For let and note that we also have . For take such that ,
Let and take small enough so that . We will eventually take , which sends .
We have for all
(B.23) |
To see this, note that when we have
and when (but ) we have
The same holds with superscript .
Let
The choice of guarantees that . Furthermore, (B.23) shows that is a convex combination of the vectors with all strictly positive coefficients. Consequently, .
Take rational in such that . Let be such that for and take such that
for the sequence in Lemma B.9 corresponding to the above choice of and and to . Abbreviate . Using (B.23) we have for
(B.24) |
The same holds with superscript . Thus, for all and for large
(B.25) |
This implies that when is large, (which belongs to ) is accessible from by an -admissible path of length
(B.26) |
Note that
(B.27) |
The first equality and (B.24) imply that
and therefore for large enough. This will allow us to apply (B.16).
Since is accessible from by an -admissible path of length , concatenating this path and the one from to gives an -admissible path from to of length
Hence . Subadditivity now gives
Using this, (B.16), and (B.27), we get
Taking and the continuity of on gives
Since and , using again the continuity of on completes the proof of (B.18):
Proof of (B.19). Proceed similarly to the proof of (B.18), but with the sequences and realizing the on the left-hand side of (B.19). Again, we have the representation (B.20), the limits (B.21), and .
We start by treating the case when . In this case let or , so that is even. Observe that and hence for large. Thus, one can make an admissible loop of length from back to and then take a path of length from to . From (B.5) we have . If is bounded then so is and we have . On the other hand, if along some subsequence, then along this subsequence, and for large, we have and, applying (B.8), we then get
for large enough. Dividing by and taking we deduce that
The continuity of at implies then that the on the left-hand side of (B.19) is . For the rest of the proof we can and will assume that .
Define as in (B.22). Let be a rational in . Choose , , so that for , when we have and when we have such that and overall we have
This is possible since and therefore for some and . Let and choose small enough so that . Note that
Indeed, this clearly holds when and when we have
The above two observations imply that
We can then find rational in such that .
Let be such that for and take such that
for the sequence in Lemma B.9 corresponding to and to the above choice of and . Abbreviate and observe that if then
(B.28) |
Then for large
(B.29) |
This inequality is trivial if . The same computation works with minus sign superscripts. This implies that is accessible from in
-steps and -steps. Note that
As a consequence,
and one can then apply (B.16). Then, as in the proof of (B.18), using subadditivity then taking and then and using the continuity of on gives
Another use of the continuity of completes the proof of (B.19). ∎
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 . Tile by -cubes indexed by . Each -cube is colored randomly black or white in a shift-stationary manner. Let be the marginal probability that a particular cube is black and assume that
(C.1) |
Assume finite range dependence: there is a strictly positive integer constant such that
(C.2) | , the colors of the cubes are i.i.d. |
There are distinct i.i.d. collections, indexed by .
It may be desirable to let the separation of the cubes be a parameter. For a positive integer and , define the collection of cubes with lower left corners on the grid . For a given , is the number of distinct collections indexed by . We always consider where is the fixed constant of the independence assumption (C.2).
Let denote the -ball (diamond) of radius in , with (inner) boundary .
Lemma C.1.
To prove Lemma C.1 we record a Bernoulli large deviation bound.
Lemma C.2.
Assume (C.2) and let be the marginal probability of a black cube. Then there exist constants such that, for all integers , , and , with , and for any particular sequence of distinct -cubes, the following estimate holds for some determined by :
Furthermore, for all and .
Proof.
Pick so that contains at least of the cubes . Since these are colored independently and , basic large deviations gives
where the last equality defines and the well-known Cramér rate function [16] of the Bernoulli distribution is
To complete the proof, observe that
Proof of Lemma C.1.
Consider for the moment a fixed path from to a point such that . Assume so that .
For let level of -cubes refer to the collection . Since points satisfy
level cubes are subsets of .
To reach the point , path must have entered and exited at least one -cube at levels where satisfies
This calculation excludes the cube that contains the endpoint . From this
(C.4) |
Consider the sequence of -cubes that path intersects: , with the initial point and the final point . Remove loops from this sequence (if any), for example by the following procedure:
-
(1)
Let be the minimal index such that for some . Let be the maximal for . Then remove .
-
(2)
Repeat the same step on the remaining sequence , as long as loops remain.
After loop removal relabel the sequence of remaining cubes consecutively to arrive at a new sequence of distinct -cubes with and still and . This sequence takes nearest-neighbor steps on the coarse-grained lattice of -cubes, in the sense that , because this property is preserved by the loop removal. Since enters and leaves behind at least one -cube on each level , we have the bound .
We have now associated to each path a sequence of distinct -cubes that both enters from the outside and exits again. We apply Lemma C.2 to these sequences of cubes.
Proof of Lemma 5.3.
Surround each -cube with -boxes so that each dimensional face of is directly opposite a large face of one of the -boxes. Precisely, first put at the center of the -cube on , and then define -boxes for . Any lattice path that enters and exits must cross in the sense of (5.8) one of the -boxes that surround .
Color black if all -boxes surrounding it are black. The probability that is black can be made arbitrarily close to 1 by choosing and large enough and small enough in the definition (5.4)–(5.5) of a black -box. The color of depends only on the edge variables in the union of the boxes enlarged as in (5.2). The separation of in (C.2) can be fixed large enough to guarantee that over the cubes are pairwise disjoint.
Apply Lemma C.1 with . Tighten the requirement of Lemma C.1 to to guarantee that if a path intersects then it also intersects the complement of . (If remains inside then the -distance between the endpoints of is at most and cannot connect the origin to .) Thus for every intersected by , at least one of the -boxes surrounding is crossed by in the sense of (5.8). In conclusion, on the event in (C.3) each path from the origin to crosses at least disjoint -boxes. Of these, at least must come from some particular collection . Thus in Lemma 5.3 we can take , and . ∎
Appendix D Convex analysis
Lemma D.1.
Let be a proper convex function on ( and is not identically ) and . Then the following statements are equivalent.
-
(a)
For some , .
-
(b)
is constant over .
-
(c)
is differentiable at .
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.