• Aucun résultat trouvé

In this subsection, we recall a few well-known properties of the DGFF that are valid on any finite graph. Let (V, E) be a finite graph, and letV⊂V be a non-empty set of vertices. Let h denote a sample from the DGFF with boundary values given by some functionh:V!R.

When f is a function on V, the discrete gradient ∇f is the function on the set of ordered pairs (v, u) such that {v, u}∈E defined by ∇f((u, v))=f(v)−f(u). When defining the norm of the gradient we sum over undirected edges, i.e., we write

k∇f(v)k2= X

{u,v}∈E

(f(v)−f(u))2. (2.1)

Thus, the probability density ofhis proportional toe−k∇h(v)k2/2. Therefore, whenh=0 on V, h is a standard Gaussian with respect to the norm k∇hk on Ω. The (discrete) Dirichlet inner product that defines this norm can be written

(f, g)= X

{u,v}∈E

(f(v)−f(u))(g(v)−g(u)). (2.2)

Now write ∆f(v)=P

u∼v(f(u)−f(v)), where the sum is over all neighborsuof v.

By expanding and rearranging the summands in (2.2), we find

(f, g)=−(∆f, g). (2.3)

LetV0⊂VD. We claim that the vector space of functions f:V!Rthat are zero on V\V0, and the vector space of functionsf:V!Rthat are discrete-harmonic on V0 (i.e.,

∆f=0 onV0) are orthogonal to each other with respect to the inner product (·,·), and together they span RV. This basic observation will be used frequently below. Indeed, that they are orthogonal follows immediately from (2.3), and a dimension count now shows that the two spaces together spanRV.

The following consequence of this orthogonality property will be used below. Let V0⊂V satisfyV0⊃V and let h0 denote the function that is discrete-harmonic in V\V0

and equal to h in V0. Then h−h0 and h0 are independent random variables, because h=h0+(h−h0) is the corresponding orthogonal decomposition ofh. It also follows that h−h0is the DGFF inV\V0 with zero boundary values onV0. (2.4) Observe that the Markov property and the effect of boundary conditions that were men-tioned in§1.2 are immediate consequences of (2.4). An additional useful consequence is that the law ofh0is proportional to e−(h0,h0)2/2 times the Lebesgue measure onRV0.

We now derive a useful well-known expression for the expectation ofh(v)h(u):

E[h(v)h(u)]−E[h(v)]E[h(u)] =G(u, v)

deg(v), (2.5)

whereG(u, v) is the expected number of visits tov by a simple random walk started at ubefore it hits V and deg(v) is the degree ofv; that is, the number of edges incident with it. (The function G is known as the Green function.) As we have noted in the introduction, his the sum of the discrete-harmonic extension of h and a DGFF with zero boundary values. It therefore suffices to prove (2.5) in the case whereh=0 onV. In this case, E[h(v)]=0=E[h(u)]. Setting Gv(u)=G(u, v), we observe (or recall) that

∆Gv(u)=−deg(v)1v(u). Thus,

h(v) = (h,1v) =−(h,∆Gv) deg(v)

(2.3)

= (h, Gv)

deg(v) .

If X is a standard Gaussian in Rn, and x, y∈Rn, then E[(X·x)(X·y)]=x·y. Conse-quently, whenh=0, we have

E[h(v)h(u)] =E[(h, Gv)(h, Gu)]

deg(v) deg(u) = (Gv, Gu)

deg(v) deg(u)

= −(Gv,∆Gu)

deg(v) deg(u)=(Gv,1u)

deg(v) =G(u, v) deg(v). This proves (2.5).

2.2. Some assumptions and notation

We will make frequent use of the following notation and assumptions:

(h) A bounded domain (non-empty, open, connected set) D⊂R2 whose boundary

∂D is a subgraph of TG is fixed. The set of vertices of TG inD is denoted byVD and V denotes the set of vertices in ∂D. A constant ¯Λ>0 is fixed, as well as a function h:V!Rsatisfyingkhk6Λ. The DGFF on¯ D with boundary values given byh is denoted byh. Also setV=VD=VD∪V.

We denote by TGthe hexagonal grid which is dual to the triangular lattice TG—so that each hexagonal face of TGis centered at a vertex of TG. Generally, a TG-hexagon will mean a closed hexagonal face of TG. Denote byBRthe union of all TG-hexagons that intersect the ballB(0, R).

Sometimes, in addition to (h) we will need to assume:

(D) The domainD is simply connected, and (to avoid minor but annoying triviali-ties)∂D is a simple closed curve. We fix two distinct midpoints of TG-edgesx andy

on ∂D. Let the counterclockwise (respectively, clockwise) arc of ∂D from x to y be denoted by∂+ (respectively,∂).

IfH⊂D is a TG-hexagon, we writeh(H) as a shorthand for the value ofhon the center ofH (which is a vertex of TG). Assuming (h) and (D), letD+denote the union of all TG-hexagons contained in D where h is positive together with the intersection of Dwith TG-hexagons centered at vertices in ∂+. Let D be the closure of D\D+ (which almost surely consists of TG-hexagons inDwhereh<0 and the intersection ofD with TG-hexagons whose center is in∂). Then∂D∩∂D+ necessarily consists of the interface we previously calledγ, and a collection of disjoint simple closed paths. We use the terminterface (orzero-height interface) to describe a simple (or simple closed) path in∂D∩∂D+oriented so thatD+is on its right (that is, oriented clockwise aroundD+).

Throughout, the notationO(s) represents any quantityf such that|f|6Csfor some absolute constantC. We use the notationOΛ¯(s) if the constant also depends on ¯Λ. When introducing a constant c, we often write c=c(a, b) as shorthand to indicate that c may depend onaandb.

2.3. Simple random walk background

We need to recall a very useful property of the discrete-harmonic measure of simple random walk.

Lemma 2.1. (Hit near) Let v be a vertex of the grid TG,and let H be a connected subgraph of TG. Set d=diamH. The probability that a simple random walk on TG started from v exits the ball B(v, d)before hitting H is at most c(dist(v, H)/d)ζ1,where c and ζ1∈(0,1) are absolute constants.

Likewise,the same bound applies to the probability that a simple random walk started at some vertex outside B(v, d)will hit B(v,dist(v, H))before H.

In fact, we may takeζ1=12. The continuous version of this statement is known as the Beurling projection theorem (the extremal case is whenH is a line segment). The above statement can probably be deduced from the discrete Beurling theorem as given

in [L1, Theorem 2.5.2], though the setting there is slightly different. In any case, since we do not require any particular value forζ1, the lemma is rather easily proved directly (see [Sch, Lemma 2.1]).

We will also use the (well-known) discrete Harnack principle, in the following form.

Lemma 2.2. (Harnack principle) Let D, V, V and VD be as in (h), let v, u∈VD, and let f:V!R be discrete-harmonic in VD.

(1) If v and uare neighbors, then

|f(v)−f(u)|6O(1) kfk dist(v, ∂D).

(2) If we assume that f>0 and that there is a path of length ` from v to u whose minimal distance to ∂D is %, then

f(u)>f(v) exp

−O(`+%)

%

.

The proof of statement (1) can be obtained by noting thatf(v) is the expected value off evaluated at the first hitting point onV of a random walk started atv, and observing that a random walk started atv and a random walk started atumay be coupled so that they meet before reaching a distance ofr:=dist(v, ∂D) with probability 1−O(1/r) and walk together after they meet. (See also the more general [LSW4, Lemma 6.2], for example.)

To prove statement (2), letW:={w∈V:f(w)>f(v)}, and observe that the maximum principle implies thatW contains a path fromv to∂D. If we assume that`612%, then a random walk started atuhas probability bounded away from zero to hitW before∂D, which by the optional sampling theorem implies (2) in this case. The case `>12% now follows by induction ond2`/%e.

We also need the following well-known estimate on the Green functionG(u, v).

Lemma 2.3. Let D, VD and V be as in (h), let u, v∈VD, and let ε>0. Set r:=

dist(u, ∂D),and suppose that within distance r/εfrom uthere is a connected component of ∂D of diameter at least εr. If |u−v|623r, say, then GD(v, u) (the expected number of visits to uby a simple random walk starting at v before hitting ∂D)satisfies

GD(v, u) = exp(Oε(1)) log r+1

|u−v|+1.

The probabilityHD(v, u) that a random walk started fromv hits ubefore ∂Dcan be expressed asGD(v, u)/GD(u, u) and hence by the lemma

HD(v, u) = exp(Oε(1))

1−log(|u−v|+1) log(r+1)

. (2.6)

Let pR denote the probability that a simple random walk started at 0 does not return to 0 before exiting the ballB(0, R). We now show that the lemma follows from the well-known estimate

pR=exp(O(1))

log(R+2). (2.7)

In the setting of the lemma, consider some vertexw∈VDsuch that|w−u|>r. It is easy to see that with probability bounded away from zero (by a function ofε), a simple random walk started fromwwill hit∂Dbeforeu. Hence,q:=min{1−HD(w, u):w∈VD\B(u, r)}

is bounded away from 0 by a positive function ofε. Clearly, 1

qpr>GD(u, u)> 1 pr

.

This proves the caseu=v of the lemma from (2.7). Now start a simple random walk at a vertexv6=usatisfying|v−u|623r, and let the walk stop when it hits∂D. It is easy to see that the probability that the walk visitsv after exiting the ballB v,14|v−u|

, say, is within a bounded factor from HD(v, u). Therefore, the expected number of visits tov after exitingB v,14|v−u|

is within a bounded factor ofGD(v, u). But the former is the same asGD(v, v)−GB(v,|v−u|/4)(v, v). The lemma follows.

We have not found a reference proving (2.7) in a way that generalizes to the setting of Theorem 1.4, though the result is well known. In fact, it easily follows from Rayleigh’s method, as explained in [DoSn,§2.2]. In that book, the goal is to show thatpR!0, as R!∞, for lattices in the plane but not in R3. However, the method easily yields the quantitative bounds (2.7).

3. The height gap in the discrete setting