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


11affiliationtext: Tsinghua University, China

Multiple SLEs and Dyson Brownian motion:
transition density and Green’s function

Hao Wu [email protected]. Supported by Beijing Natural Science Foundation (JQ20001). Lu Yang [email protected]
Abstract

1 Introduction

In this article, we consider connections between multiple chordal SLEs and Dyson Brownian motion. The motivation to study multiple SLEs is to describe the scaling limit of multiple interfaces in polygons, see for instance [BBK05, Dub07, KP16]. The study on its connections with Dyson Brownian motion goes back to [Car03] in multiple radial case, and some recent research including [KK21, CM22, CLM23] has also proved properties of multiple chordal SLE driven by Dyson Brownian motion, which is a local formulation of multiple chordal SLE defined through multiple Loewner equations.

There is another global formulation of multiple chordal SLE defined as commuting Loewner chains and we first introduce its notions. We call (Ω;x1,,x2N)(\Omega;x_{1},\ldots,x_{2N}) a polygon if Ω\Omega\subset\mathbb{C} is a simply connected domain such that Ω\partial\Omega is locally connected and x1,,x2Nx_{1},\ldots,x_{2N} are 2N2N marked points lying on Ω\partial\Omega in counterclockwise order. We consider NN non-crossing simple curves in Ω\Omega such that each curve connects two points among {x1,,x2N}\{x_{1},\ldots,x_{2N}\}. These curves can have various planar connectivities. We describe such connectivities by link patterns α={{a1,b1},,{aN,bN}}\alpha=\{\{a_{1},b_{1}\},\ldots,\{a_{N},b_{N}\}\} where {a1,b1,,aN,bN}={1,2,,2N}\{a_{1},b_{1},\ldots,a_{N},b_{N}\}=\{1,2,\ldots,2N\}, and denote by LPN\mathrm{LP}_{N} the set of all link patterns. Note that #LPN\#\mathrm{LP}_{N} is the Catalan number CN=1N+1(2NN)C_{N}=\frac{1}{N+1}\binom{2N}{N}.

When N=1N=1, we denote by Xsimple(Ω;x1,x2)X_{\mathrm{simple}}(\Omega;x_{1},x_{2}) the set of continuous simple unparameterized curves in Ω\Omega connecting x1x_{1} and x2x_{2} such that they only touch the boundary Ω\partial\Omega in {x1,x2}\{x_{1},x_{2}\}. We denote by X(Ω;x1,x2)X(\Omega;x_{1},x_{2}) the closure of Xsimple(Ω;x1,x2)X_{\mathrm{simple}}(\Omega;x_{1},x_{2}) in the metric topology of the set XX of all planar oriented curves (i.e., continuous mappings from [0,1][0,1] to \mathbb{C} modulo reparameterization) equipped with metric

dist(η,η~):=infφ,φ~supt[0,1]|η(φ(t))η~(φ~(t))|,η,η~X,\mathrm{dist}(\eta,\tilde{\eta}):=\inf_{\varphi,\tilde{\varphi}}\sup_{t\in[0,1]}|\eta(\varphi(t))-\tilde{\eta}(\tilde{\varphi}(t))|,\quad\eta,\tilde{\eta}\in X,

where the infimum is taken over all increasing homeomorphisms φ,φ~:[0,1][0,1]\varphi,\tilde{\varphi}:[0,1]\mapsto[0,1]. Chordal SLE in (Ω;x1,x2)(\Omega;x_{1},x_{2}) is a continuous random curve in Ω\Omega connecting x1x_{1} and x2x_{2}. Note that SLEκ\mathrm{SLE}_{\kappa} in (Ω;x1,x2)(\Omega;x_{1},x_{2}) is a random curve in Xsimple(Ω;x1,x2)X_{\mathrm{simple}}(\Omega;x_{1},x_{2}) when κ4\kappa\leq 4 and that SLEκ\mathrm{SLE}_{\kappa} in (Ω;x1,x2)(\Omega;x_{1},x_{2}) is a random curve in X(Ω;x1,x2)X(\Omega;x_{1},x_{2}) when κ(4,8)\kappa\in(4,8). In general, for N2N\geq 2 and α={{a1,b1},,{aN,bN}}LPN\alpha=\{\{a_{1},b_{1}\},\ldots,\{a_{N},b_{N}\}\}\in\mathrm{LP}_{N}, we denote by Xα(Ω;x1,,x2N)X_{\alpha}(\Omega;x_{1},\ldots,x_{2N}) the set of families (γ1,,γN)(\gamma_{1},\ldots,\gamma_{N}) of curves such that, for each s{1,,N}s\in\{1,\ldots,N\}, we have γsX(Ω;xas,xbs)\gamma_{s}\in X(\Omega;x_{a_{s}},x_{b_{s}}) and γs\gamma_{s} does not disconnect any two points xa,xbx_{a},x_{b} with {a,b}α\{a,b\}\in\alpha from each other.

Fix κ(0,8)\kappa\in(0,8). For N1N\geq 1 and αLPN\alpha\in\mathrm{LP}_{N}, there exists a unique probability measure on (γ1,,γN)Xα(Ω;x1,,x2N)(\gamma_{1},\ldots,\gamma_{N})\in X_{\alpha}(\Omega;x_{1},\ldots,x_{2N}) such that, for each j{1,,N}j\in\{1,\ldots,N\}, the conditional law of γj\gamma_{j} given {γ1,,γN}{γj}\{\gamma_{1},\ldots,\gamma_{N}\}\setminus\{\gamma_{j}\} is the chordal SLEκ\mathrm{SLE}_{\kappa} connecting xajx_{a_{j}} and xbjx_{b_{j}} in the connected component of the domain Ωkjγk\Omega\setminus\bigcup_{k\neq j}\gamma_{k} having xajx_{a_{j}} and xbjx_{b_{j}} on its boundary. For the existence and uniqueness of such measures, see [KL07], [PW19, Theorem 1.3] and [BPW21, Theorem 1.2] for κ(0,4]\kappa\in(0,4], and [AHSY23, Theorem 1.2], [Zha23, Theorem 4.2] and [FLPW24, Theorem 1.21] for κ(4,8)\kappa\in(4,8). We call this measure chordal NN-SLEκ\mathrm{SLE}_{\kappa} in polygon (Ω;x1,,x2N)(\Omega;x_{1},\ldots,x_{2N}) associated to link pattern α\alpha.

In the companion paper [FWY24], we derived the connection between chordal NN-SLEκ\mathrm{SLE}_{\kappa} in the unit disc and Dyson circular ensemble for κ(0,4]\kappa\in(0,4]. In this article, we will prove corresponding conclusions for chordal NN-SLEκ\mathrm{SLE}_{\kappa} and Dyson Brownian motion, and extend the conclusions from κ(0,4]\kappa\in(0,4] to κ(0,8)\kappa\in(0,8). The main conclusions are summarized as follows. We will parameterize multiple chordal SLEs by a “common time” and derive the transition density of the driving function in Theorem 1.1: it can be given by the transition density of Dyson Brownian motion and Green’s function (Definition 1.2). Using such connection and the asymptotic of the transition density of Dyson Brownian motion, we derive the asymptotic of the probability of a rare event and show that, conditional on this rare event, the driving function of multiple chordal SLEs converges to Dyson Brownian motion (Proposition 1.3). In particular, we take κ=4\kappa=4 as an example and connect our result to Gaussian free field level lines (Proposition 4.1). In this case, the driving function has an explicit SDE (4.2), Green’s function has an explicit formula (4.4), and the Dyson Brownian motion is of parameter β=2\beta=2. Finally, we give one application of our main result: we derive the asymptotic of the probability for chordal NN-SLE to hit a small neighborhood of a boundary point (Theorem 1.4).

Let us elaborate on how we extend our results from κ(0,4]\kappa\in(0,4] to κ(0,8)\kappa\in(0,8). To prove the first main conclusion Theorem 1.1, we first relate global multiple SLEs to multiple SLE partition functions “pure partition functions” (see Section 2.1); second, we use these partition functions to construct 2N2N-time local martingales (see Section 3.2); and finally, we take the common time in these local martingales and derive the desired transition density. This proof works for all κ(0,8)\kappa\in(0,8). To prove the second main conclusion Proposition 1.3 and Theorem 1.4, we need further analysis on pure partition functions in Section 2.2. Such analysis relies on the relation between Coulomb gas integrals and pure partition functions which is only available due to recent development in [FPW24, FLPW24]. Using these tools, we are also able to extend previous result [FWY24, Theorem 1.2] from κ(0,4]\kappa\in(0,4] to κ(0,8)\kappa\in(0,8), see Remark 1.5.

1.1 Multiple SLEs and Dyson Brownian motion

We focus on the polygon (;x1,,x2N)(\mathbb{H};x_{1},\ldots,x_{2N}) where \mathbb{H} is the upper-half plane and x1<<x2Nx_{1}<\cdots<x_{2N}. We denote

𝔛2N:={(x1,,x2N)2N:x1<<x2N}.\mathfrak{X}_{2N}:=\{(x_{1},\ldots,x_{2N})\in\mathbb{R}^{2N}:x_{1}<\cdots<x_{2N}\}.

We denote by α(𝒙)\mathbb{P}_{\alpha}^{(\boldsymbol{x})} the law of chordal NN-SLEκ\mathrm{SLE}_{\kappa} in polygon (;x1,,x2N)(\mathbb{H};x_{1},\ldots,x_{2N}) associated to link pattern α={{a1,b1},,{aN,bN}}LPN\alpha=\{\{a_{1},b_{1}\},\ldots,\{a_{N},b_{N}\}\}\in\mathrm{LP}_{N}. Suppose (γ1,,γN)α(𝒙)(\gamma_{1},\ldots,\gamma_{N})\sim\mathbb{P}_{\alpha}^{(\boldsymbol{x})}. Recall that γj\gamma_{j} is a curve in \mathbb{H} from xajx_{a_{j}} to xbjx_{b_{j}}. We view (γ1,,γN)(\gamma_{1},\ldots,\gamma_{N}) as a 2N2N-tuple of continuous non-self-crossing curves 𝜼=(η1,,η2N)\boldsymbol{\eta}=(\eta_{1},\ldots,\eta_{2N}): for j{1,,N}j\in\{1,\ldots,N\}, we define ηaj\eta_{a_{j}} to be γj\gamma_{j} and ηbj\eta_{b_{j}} to be the time-reversal of γj\gamma_{j}. In this way, ηj\eta_{j} is a continuous curve in \mathbb{H} starting from xjx_{j} for j{1,,2N}j\in\{1,\ldots,2N\}. We introduce the common parameterization under which the 2N2N curves (η1,,η2N)(\eta_{1},\ldots,\eta_{2N}) grow simultaneously and are parameterized by a single time parameter tt. For 𝒕=(t1,,t2N)+2N\boldsymbol{t}=(t_{1},\ldots,t_{2N})\in\mathbb{R}_{+}^{2N}, let H𝒕H_{\boldsymbol{t}} be the unbounded connected component of (jηj([0,tj]))\mathbb{H}\setminus(\cup_{j}\eta_{j}([0,t_{j}])). Let g𝒕:H𝒕g_{\boldsymbol{t}}:H_{\boldsymbol{t}}\to\mathbb{H} be the unique conformal transformation with

g𝒕(z)=z+(𝒕)z+o(1/|z|),z.g_{\boldsymbol{t}}(z)=z+\frac{\aleph(\boldsymbol{t})}{z}+o(1/|z|),\quad z\to\infty.

For a>0a>0, we say that the collection of curves (ηj([0,tj]),1j2N)(\eta_{j}([0,t_{j}]),1\leq j\leq 2N) have aa-common parameterization if for each t<τ:=min1j2Ntjt<\tau:=\min_{1\leq j\leq 2N}t_{j},

j(t,t,,t)=a,j{1,2,,2N}.\displaystyle\partial_{j}\aleph(t,t,\ldots,t)=a,\quad j\in\{1,2,\ldots,2N\}.

In particular, under aa-common parameterization, we denote gt:=g(t,,t)g_{t}:=g_{(t,\ldots,t)} and then

gt(z)=z+2aNtz+o(1/|z|),z.\displaystyle g_{t}(z)=z+\frac{2aNt}{z}+o(1/|z|),\quad z\to\infty.

Moreover, under aa-common parameterization, gtg_{t} satisfies the following chordal Loewner equation

tgt(z)=j=12Nagt(z)Xtj,z(jηj([0,t])),t<τ.\displaystyle\partial_{t}g_{t}(z)=\sum_{j=1}^{2N}\frac{a}{g_{t}(z)-X_{t}^{j}},\quad\forall z\in\mathbb{H}\setminus\left(\cup_{j}\eta_{j}([0,t])\right),\quad\forall t<\tau.

We say that {(ηj(t),0t<τ)}1j2N\{(\eta_{j}(t),0\leq t<\tau)\}_{1\leq j\leq 2N} is the chordal Loewner chain with aa-common parameterization and with driving functions {(Xtj,0t<τ)}1j2N\{(X_{t}^{j},0\leq t<\tau)\}_{1\leq j\leq 2N}.

For chordal NN-SLEκ\mathrm{SLE}_{\kappa}, we view it as a chordal Loewner chain (𝜼(t)=(η1(t),,η2N(t)),t0)(\boldsymbol{\eta}(t)=(\eta_{1}(t),\ldots,\eta_{2N}(t)),t\geq 0) with 2/κ2/\kappa-common parameterization and with driving functions (𝑿t=(Xt1,,Xt2N),t0)(\boldsymbol{X}_{t}=(X^{1}_{t},\ldots,X^{2N}_{t}),t\geq 0). Denote by (t,t0)(\mathcal{F}_{t},t\geq 0) the filtration generated by (𝜼(t),t0)(\boldsymbol{\eta}(t),t\geq 0), and define the lifetime TT as the first time tt that Xtj=XtkX_{t}^{j}=X_{t}^{k} for some jkj\neq k, then we have α(𝒙)[T<]=1\mathbb{P}_{\alpha}^{(\boldsymbol{x})}[T<\infty]=1. We will prove in Lemma 3.2 that, under α(𝒙)\mathbb{P}_{\alpha}^{(\boldsymbol{x})}, the driving functions (𝑿t,0t<T)(\boldsymbol{X}_{t},0\leq t<T) satisfies the SDE

dXtj=dBtj+(jlog𝒵α)(Xt1,,Xt2N)dt+kj2/κXtjXtkdt,1j2N,\mathrm{d}X_{t}^{j}=\mathrm{d}B_{t}^{j}+(\partial_{j}\log\mathcal{Z}_{\alpha})(X_{t}^{1},\ldots,X_{t}^{2N})\mathrm{d}t+\sum_{k\neq j}\frac{2/\kappa}{X_{t}^{j}-X_{t}^{k}}\mathrm{d}t,\quad 1\leq j\leq 2N, (1.1)

where {Bj}1j2N\{B^{j}\}_{1\leq j\leq 2N} are independent standard Brownian motions and 𝒵α\mathcal{Z}_{\alpha} is the pure partition function for chordal NN-SLEκ\mathrm{SLE}_{\kappa} defined in Section 2.3. The main goal of this article is to compare the solution to (1.1) with Dyson Brownian motion.

Theorem 1.1.

Fix κ(0,8)\kappa\in(0,8) and αLPN\alpha\in\mathrm{LP}_{N}.

  • Denote by pα(t;,)p_{\alpha}(t;\cdot,\cdot) the transition density for the solution to (1.1).111By saying that pα(t;,)p_{\alpha}(t;\cdot,\cdot) is the transition density, we mean that for any t>0t>0 and 𝒙𝔛2N\boldsymbol{x}\in\mathfrak{X}_{2N}, if (𝑿t,0t<T)(\boldsymbol{X}_{t},0\leq t<T) starts from 𝒙\boldsymbol{x}, then for any bounded measurable function ff on 𝔛2N\mathfrak{X}_{2N}, 𝔼α(𝒙)[𝟙{T>t}f(𝑿t)]=𝔛2Npα(t;𝒙,𝒚)f(𝒚)d𝒚.\mathbb{E}_{\alpha}^{(\boldsymbol{x})}\left[\mathbb{1}_{\{T>t\}}f(\boldsymbol{X}_{t})\right]=\int_{\mathfrak{X}_{2N}}p_{\alpha}(t;\boldsymbol{x},\boldsymbol{y})f(\boldsymbol{y})\mathrm{d}\boldsymbol{y}.

  • Denote by p(t;,)p_{*}(t;\cdot,\cdot) the transition density for Dyson Brownian motion with parameter β=8/κ\beta=8/\kappa:

    dXtj=dBtj+kj4/κXtjXtkdt,1j2N,\mathrm{d}X_{t}^{j}=\mathrm{d}B_{t}^{j}+\sum_{k\neq j}\frac{4/\kappa}{X_{t}^{j}-X_{t}^{k}}\mathrm{d}t,\quad 1\leq j\leq 2N, (1.2)

    where {Bj}1j2N\{B^{j}\}_{1\leq j\leq 2N} are independent standard Brownian motions.

Then we have

pα(t;𝒙,𝒚)=p(t;𝒙,𝒚)Gα(𝒙)Gα(𝒚),for all t0, and 𝒙,𝒚𝔛2N,p_{\alpha}(t;\boldsymbol{x},\boldsymbol{y})=p_{*}(t;\boldsymbol{x},\boldsymbol{y})\frac{G_{\alpha}(\boldsymbol{x})}{G_{\alpha}(\boldsymbol{y})},\quad\text{for all }t\geq 0,\text{ and }\boldsymbol{x},\boldsymbol{y}\in\mathfrak{X}_{2N}, (1.3)

where GαG_{\alpha} is the Green’s function in Definition 1.2.

Definition 1.2.

Fix κ(0,8)\kappa\in(0,8) and αLPN\alpha\in\mathrm{LP}_{N}. For (x1,,x2N)𝔛2N(x_{1},\ldots,x_{2N})\in\mathfrak{X}_{2N}, we define Green’s function in the boundary for chordal NN-SLEκ\mathrm{SLE}_{\kappa} as

Gα(x1,,x2N):=𝒵(x1,,x2N)𝒵α(x1,,x2N),G_{\alpha}(x_{1},\ldots,x_{2N}):=\frac{\mathcal{Z}_{*}(x_{1},\ldots,x_{2N})}{\mathcal{Z}_{\alpha}(x_{1},\ldots,x_{2N})},

where 𝒵\mathcal{Z}_{*} is the partition function for SLEκ(2,,2)\mathrm{SLE}_{\kappa}(2,\ldots,2):

𝒵(x1,,x2N)=1j<k2N(xkxj)2/κ,\mathcal{Z}_{*}(x_{1},\ldots,x_{2N})=\prod_{1\leq j<k\leq 2N}(x_{k}-x_{j})^{2/\kappa}, (1.4)

and 𝒵α\mathcal{Z}_{\alpha} is the pure partition function for chordal NN-SLEκ\mathrm{SLE}_{\kappa} defined in Section 2.3.

The relation (1.3) provides a tool to analyze multiple SLEs using Dyson Brownian motion. For instance, one may analyze multiple SLEs as NN\to\infty using (1.3) as the asymptotic of Dyson Brownian motion as NN\to\infty is known, see [dMS16, HK18, HS21]. One may analyze multiple SLEs as κ0+\kappa\to 0+ using (1.3) as the asymptotic of Dyson Brownian motion as β\beta\to\infty is known, this is related to large deviation principle for multiple SLEs, see [PW24, AHP24]. In the following, we analyze multiple SLEs as tt\to\infty using (1.3) as the asymptotic of Dyson Brownian motion as tt\to\infty is known [Rös98] (see Lemma 2.6) and we derive the asymptotic of the probability for multiple SLEs to hit a small neighborhood of a boundary point.

1.2 Estimates on multiple SLEs

Proposition 1.3.

Fix κ(0,8)\kappa\in(0,8), αLPN\alpha\in\mathrm{LP}_{N} and 𝐱𝔛2N\boldsymbol{x}\in\mathfrak{X}_{2N}. We have

α(𝒙)[T>t]=1𝒥αGα(𝒙)(1t)A2N+(1+O(|𝒙|t)),as t,\mathbb{P}_{\alpha}^{(\boldsymbol{x})}[T>t]=\mathcal{I}_{*}^{-1}\mathcal{J}_{\alpha}G_{\alpha}(\boldsymbol{x})\left(\frac{1}{\sqrt{t}}\right)^{A_{2N}^{+}}\left(1+O\left(\frac{|\boldsymbol{x}|}{\sqrt{t}}\right)\right),\quad\text{as }t\to\infty, (1.5)

where \mathcal{I}_{*} and 𝒥α\mathcal{J}_{\alpha} are normalization constants defined in (2.13) and (2.14), GαG_{\alpha} is Green’s function in Definition 1.2, and A2N+A_{2N}^{+} is the half-plane arm exponent for SLE:

A2N+=N(4N+4κ)κ.A_{2N}^{+}=\frac{N(4N+4-\kappa)}{\kappa}. (1.6)

Moreover,

  • For t>0t>0, denote by μα|t(𝒙)\mu_{\alpha\,|\,t}^{(\boldsymbol{x})} the law of the solution to (1.1) started from 𝒙\boldsymbol{x} conditional on {T>t}\{T>t\}.

  • Denote by μ(𝒙)\mu^{(\boldsymbol{x})}_{*} the law of Dyson Brownian motion (1.2) with parameter β=8/κ\beta=8/\kappa started from 𝒙\boldsymbol{x}.

Denote by (t,t0)(\mathcal{F}_{t},t\geq 0) the filtration generated by the Brownian motions in the equations (1.1) or (1.2). Then, for any s>0s>0, when both measures μα|t(𝐱)\mu^{(\boldsymbol{x})}_{\alpha\,|\,t} and μ(𝐱)\mu^{(\boldsymbol{x})}_{*} are restricted to s\mathcal{F}_{s}, the total variation distance between the two measures goes to zero as tt\to\infty:

limtdistTV(μα|t(𝒙)[|s],μ(𝒙)[|s])=0.\lim_{t\to\infty}\mathrm{dist}_{\mathrm{TV}}\left(\mu_{\alpha\,|\,t}^{(\boldsymbol{x})}[\cdot\,|\,_{\mathcal{F}_{s}}],\mu^{(\boldsymbol{x})}_{*}[\cdot\,|\,_{\mathcal{F}_{s}}]\right)=0. (1.7)
Theorem 1.4.

Fix κ(0,8)\kappa\in(0,8), 𝐱=(x1,,x2N)𝔛2N\boldsymbol{x}=(x_{1},\ldots,x_{2N})\in\mathfrak{X}_{2N}, αLPN\alpha\in\mathrm{LP}_{N} and suppose (γ1,,γN)α(𝐱)(\gamma_{1},\ldots,\gamma_{N})\sim\mathbb{P}_{\alpha}^{(\boldsymbol{x})}. Then we have222For two functions F1(𝐱;R)F_{1}(\boldsymbol{x};R) and F2(𝐱;R)F_{2}(\boldsymbol{x};R), we write F1(𝐱;R)F2(𝐱;R)F_{1}(\boldsymbol{x};R)\asymp F_{2}(\boldsymbol{x};R) as R/|𝐱|R/|\boldsymbol{x}|\to\infty if there exist constants C(1,)C\in(1,\infty) and R0R_{0} such that C1F1(𝐱;R)F2(𝐱;R)CC^{-1}\leq\frac{F_{1}(\boldsymbol{x};R)}{F_{2}(\boldsymbol{x};R)}\leq C for all RR0|𝐱|R\geq R_{0}|\boldsymbol{x}|.

α(𝒙)[γjB(0,R),1jN]Gα(𝒙)RA2N+,as R/|𝒙|,\mathbb{P}_{\alpha}^{(\boldsymbol{x})}[\gamma_{j}\cap\partial B(0,R)\neq\emptyset,1\leq j\leq N]\asymp G_{\alpha}(\boldsymbol{x})R^{-A_{2N}^{+}},\quad\text{as }R/|\boldsymbol{x}|\to\infty, (1.8)

where A2N+A_{2N}^{+} is half-plane arm exponent (1.6) and GαG_{\alpha} is Green’s function in Definition 1.2, and the implicit constants depend only on κ,N,α\kappa,N,\alpha.

The rest of this article is organized as follows. We give preliminaries on multiple SLEs and Dyson Brownian motion in Section 2. We prove Theorem 1.1 in Section 3. The main ingredients for the proof are 2N2N-time local martingales constructed in Section 3.2. Using Theorem 1.1 and the asymptotic of the transition density of Dyson Brownian motion, we complete the proof of Proposition 1.3 in Section 3.3. We prove Proposition 4.1 in Section 4. Finally, we prove Proposition 1.4 in Section 5 using Proposition 1.3. The conclusion in Proposition 1.4 is proved for κ(0,8),N=1\kappa\in(0,8),N=1 in [Law15] and is proved for κ(0,8),N=2\kappa\in(0,8),N=2 in [Zha21]. We will prove (1.8) in Section 5 for all κ(0,8),N3,αLPN\kappa\in(0,8),N\geq 3,\alpha\in\mathrm{LP}_{N}.

Remark 1.5.

In the companion paper [FWY24], we derived the asymptotic of the probability for chordal NN-SLEκ\mathrm{SLE}_{\kappa} to hit a small neighborhood of an interior point for κ(0,4]\kappa\in(0,4]. Using recent development in [FPW24, FLPW24], we are able to extend it from κ(0,4]\kappa\in(0,4] to κ(0,8)\kappa\in(0,8). More precisely, fix κ(0,8)\kappa\in(0,8), consider the polygon (𝕌;x1,,x2N)(\mathbb{U};x_{1},\ldots,x_{2N}) where 𝕌\mathbb{U} is the unit disc. Let (γ1,,γN)(\gamma_{1},\ldots,\gamma_{N}) be the chordal NN-SLEκ\mathrm{SLE}_{\kappa} in (𝕌;x1,,x2N)(\mathbb{U};x_{1},\ldots,x_{2N}) associated to link pattern αLPN\alpha\in\mathrm{LP}_{N} and denote its law by α(𝕌;𝒙)\mathbb{P}_{\alpha}^{(\mathbb{U};\boldsymbol{x})}. Then333For two functions F1(𝒙;ϵ)F_{1}(\boldsymbol{x};\epsilon) and F2(𝒙;ϵ)F_{2}(\boldsymbol{x};\epsilon), we write F1(𝒙;ϵ)F2(𝒙;ϵ)F_{1}(\boldsymbol{x};\epsilon)\sim F_{2}(\boldsymbol{x};\epsilon) as ϵ0\epsilon\to 0 if the limit limϵ0F1(𝒙;ϵ)F2(𝒙;ϵ)\lim_{\epsilon\to 0}\frac{F_{1}(\boldsymbol{x};\epsilon)}{F_{2}(\boldsymbol{x};\epsilon)} exists and takes value in (0,)(0,\infty).

α(𝕌;𝒙)[γjB(0,ϵ),1jN]G~α(𝒙)ϵA2N,as ϵ0,\mathbb{P}_{\alpha}^{(\mathbb{U};\boldsymbol{x})}\left[\gamma_{j}\cap B(0,\epsilon)\neq\emptyset,1\leq j\leq N\right]\sim\tilde{G}_{\alpha}(\boldsymbol{x})\epsilon^{A_{2N}},\quad\text{as }\epsilon\to 0, (1.9)

where A2NA_{2N} is whole-plane arm exponent

A2N=16N2(κ4)28κ;A_{2N}=\frac{16N^{2}-(\kappa-4)^{2}}{8\kappa};

and G~α\tilde{G}_{\alpha} is Green’s function defined in [FWY24, Definition 1.3]. See discussion at the end of Section 5.

2 Preliminaries

We fix parameters

κ(0,8),a=2κ,h=6κ2κ,c=(6κ)(3κ8)2κ.\kappa\in(0,8),\qquad a=\frac{2}{\kappa},\qquad h=\frac{6-\kappa}{2\kappa},\qquad c=\frac{(6-\kappa)(3\kappa-8)}{2\kappa}. (2.1)

\mathbb{H}-hulls and half-plane capacity

An \mathbb{H}-hull is a relatively closed subset KK of \mathbb{H} such that KK is bounded and K\mathbb{H}\setminus K is a simply connected domain. For an \mathbb{H}-hull KK, there exists a unique conformal map gK:Kg_{K}:\mathbb{H}\setminus K\to\mathbb{H} with

gK(z)=z+cKz+o(1/|z|),zg_{K}(z)=z+\frac{c_{K}}{z}+o(1/|z|),\quad z\to\infty

for some constant cK0c_{K}\geq 0. The constant cKc_{K} is called the half-plane capacity of KK and denoted as hcap(K)\text{hcap}(K). We give some estimates about the relation between half-plane capacity and diameter of \mathbb{H}-hulls, which will be used in the proof of Theorem 1.4.

Lemma 2.1.

Suppose KK\subset\mathbb{H} is an \mathbb{H}-hull with half-plane capacity hcap(K)>0\emph{hcap}(K)>0 as described above.

  • If K{|z|>r}=K\cap\{|z|>r\}=\emptyset, then hcap(K)r2\emph{hcap}(K)\leq r^{2}.

  • If hcap(K)<r2/4\emph{hcap}(K)<r^{2}/4, then K{|z|>r}=K\cap\{|z|>r\}=\emptyset.

Proof.

This is a standard estimate using the Koebe’s 1/41/4 Theorem and Koebe’s distortion Theorem. See [Zha21, Proposition 2.4 and Lemma 6.1]. ∎

2.1 Pure partition functions

The description of the Loewner chain of chordal NN-SLEκ\mathrm{SLE}_{\kappa} involves the notion of pure paritition functions of multiple SLEκ\mathrm{SLE}_{\kappa}. These are the recursive collection {𝒵α:αN0LPN}\{\mathcal{Z}_{\alpha}\colon\alpha\in\bigsqcup_{N\geq 0}\mathrm{LP}_{N}\} of functions 𝒵α:𝔛2N\mathcal{Z}_{\alpha}\colon\mathfrak{X}_{2N}\to\mathbb{R} uniquely determined by the following 4 properties:

  • BPZ equations: for all j{1,,2N}j\in\{1,\ldots,2N\},

    [12j2+kj(axkxjkah(xkxj)2)]𝒵α(x1,,x2N)=0.\displaystyle\left[\frac{1}{2}\partial_{j}^{2}+\sum_{k\neq j}\left(\frac{a}{x_{k}-x_{j}}\partial_{k}-\frac{ah}{(x_{k}-x_{j})^{2}}\right)\right]\mathcal{Z}_{\alpha}(x_{1},\ldots,x_{2N})=0. (2.2)
  • Möbius covariance: for all Möbius maps φ\varphi of the upper half-plane \mathbb{H} such that φ(x1)<<φ(x2N)\varphi(x_{1})<\cdots<\varphi(x_{2N}), we have

    𝒵α(x1,,x2N)=j=12Nφ(xj)h×𝒵α(φ(x1),,φ(x2N)).\displaystyle\mathcal{Z}_{\alpha}(x_{1},\ldots,x_{2N})=\prod_{j=1}^{2N}\varphi^{\prime}(x_{j})^{h}\times\mathcal{Z}_{\alpha}(\varphi(x_{1}),\ldots,\varphi(x_{2N})).
  • Asymptotics: with 𝒵1\mathcal{Z}_{\emptyset}\equiv 1 for the empty link pattern LP0\emptyset\in\mathrm{LP}_{0}, the collection {𝒵α:αLPN}\{\mathcal{Z}_{\alpha}\colon\alpha\in\mathrm{LP}_{N}\} satisfies the following recursive asymptotics property. Fix N1N\geq 1 and j{1,2,,2N1}j\in\{1,2,\ldots,2N-1\}. Then, we have

    limxj,xj+1ξ𝒵α(x1,,x2N)(xj+1xj)2h={𝒵α/{j,j+1}(x1,,xj1,xj+2,,x2N),if {j,j+1}α,0,if {j,j+1}α,\displaystyle\lim_{x_{j},x_{j+1}\to\xi}\frac{\mathcal{Z}_{\alpha}(x_{1},\ldots,x_{2N})}{(x_{j+1}-x_{j})^{-2h}}=\begin{cases}\mathcal{Z}_{\alpha/\{j,j+1\}}(x_{1},\ldots,x_{j-1},x_{j+2},\ldots,x_{2N}),&\quad\text{if }\{j,j+1\}\in\alpha,\\ 0,&\quad\text{if }\{j,j+1\}\not\in\alpha,\end{cases}

    where ξ(xj1,xj+2)\xi\in(x_{j-1},x_{j+2}) (with the convention that x0=x_{0}=-\infty and x2N+1=+x_{2N+1}=+\infty), and α/{k,l}\alpha/\{k,l\} denotes the link pattern in LPN1\mathrm{LP}_{N-1} obtained by removing {k,l}\{k,l\} from α\alpha and then relabeling the remaining indices so that they are the first 2(N1)2(N-1) positive integers.

  • The functions satisfy the following power-law bound: there exist constants C(0,)C\in(0,\infty) and p(0,)p\in(0,\infty) such that for all (x1,,x2N)𝔛2N(x_{1},\ldots,x_{2N})\in\mathfrak{X}_{2N},

    |𝒵α(x1,,x2N)|C1j<k2N|xkxj|μkj(p),where μkj(p)={p,if |xkxj|1,p,if |xkxj|<1.\displaystyle|\mathcal{Z}_{\alpha}(x_{1},\ldots,x_{2N})|\leq C\prod_{1\leq j<k\leq 2N}|x_{k}-x_{j}|^{\mu_{kj}(p)},\quad\text{where }\mu_{kj}(p)=\begin{cases}p,&\text{if }|x_{k}-x_{j}|\geq 1,\\ -p,&\text{if }|x_{k}-x_{j}|<1.\end{cases} (2.3)

The uniqueness when κ(0,8)\kappa\in(0,8) of such collection of functions were proved in [FK15a]. The existence is proved in [FK15b, KP16, PW19, Wu20, FLPW24]. Pure partition functions enjoy a refined power law bound which we summarize in the following two lemmas. They will be used later.

Lemma 2.2.

Fix κ(0,6]\kappa\in(0,6], pure partition functions satisfy the following refined power law bound:

0<𝒵α(x1,,x2N){k,l}α|xkxl|2h,for all x1<<x2N.\displaystyle 0<\mathcal{Z}_{\alpha}(x_{1},\ldots,x_{2N})\leq\prod_{\{k,l\}\in\alpha}|x_{k}-x_{l}|^{-2h},\quad\text{for all }x_{1}<\cdots<x_{2N}. (2.4)
Lemma 2.3.

Fix κ(4,8)\kappa\in(4,8), pure partition functions satisfy the following refined power law bound: there exists constant C(0,)C\in(0,\infty) (depending only on NN and κ\kappa) such that for all (x1,,x2N)𝔛2N(x_{1},\ldots,x_{2N})\in\mathfrak{X}_{2N},

0<𝒵α(x1,,x2N)C1j<k2N|xkxj|μkj,where μkj={10/κ,if |xkxj|1,6/κ,if |xkxj|<1.\displaystyle 0<\mathcal{Z}_{\alpha}(x_{1},\ldots,x_{2N})\leq C\prod_{1\leq j<k\leq 2N}|x_{k}-x_{j}|^{\mu_{kj}},\quad\text{where }\mu_{kj}=\begin{cases}10/\kappa,&\text{if }|x_{k}-x_{j}|\geq 1,\\ -6/\kappa,&\text{if }|x_{k}-x_{j}|<1.\end{cases} (2.5)

The bound (2.4) is only proved for κ(0,6]\kappa\in(0,6] (see [Wu20, Theorem 1.7]) and we will prove the bound (2.5) for κ(4,8)\kappa\in(4,8) in Section 2.2. Note that the bound (2.5) is stronger than (2.3) but is weaker than (2.4). Nevertheless, it suffices for us to complete the proof of main conclusions.

2.2 Proof of Lemma 2.3

The proof of Lemma 2.3 relies on recent development [FPW24, FLPW24] on the relation between Coulomb gas integrals and pure partition functions, both of which are solutions to BPZ equations (2.2). We first introduce some necessary notations. Fix κ(4,8)\kappa\in(4,8). For N1N\geq 1, we define

𝒢(N)(x1,,x2N):=(2|cos(4π/κ)|Γ(28/κ)Γ(14/κ)2)Nx1x2x2N1x2Nf(𝒙;u1,,uN)du1duN,\displaystyle\mathcal{G}^{(N)}(x_{1},\ldots,x_{2N}):=\left(\frac{2|\cos(4\pi/\kappa)|\Gamma(2-8/\kappa)}{\Gamma(1-4/\kappa)^{2}}\right)^{N}\int_{x_{1}}^{x_{2}}\cdots\int_{x_{2N-1}}^{x_{2N}}f(\boldsymbol{x};u_{1},\ldots,u_{N})\mathrm{d}u_{1}\cdots\mathrm{d}u_{N}, (2.6)

where duq\mathrm{d}u_{q} is integrated from x2q1x_{2q-1} to x2qx_{2q} for 1qN1\leq q\leq N, and the integrand is

f(𝒙;u1,,uN):=1j<k2N(xkxj)2/κ1r<sN|usur|8/κ1j2N,1qN|uqxj|4/κ.\displaystyle f(\boldsymbol{x};u_{1},\ldots,u_{N}):=\prod_{1\leq j<k\leq 2N}(x_{k}-x_{j})^{2/\kappa}\prod_{1\leq r<s\leq N}|u_{s}-u_{r}|^{8/\kappa}\prod_{1\leq j\leq 2N,1\leq q\leq N}|u_{q}-x_{j}|^{-4/\kappa}. (2.7)

The function 𝒢(N)\mathcal{G}^{(N)} is the Coulomb gas integral function with link pattern

¯={{1,2},,{2N1,2N}}LPN,\boldsymbol{\underline{\cap\cap}}=\{\{1,2\},\ldots,\{2N-1,2N\}\}\in\mathrm{LP}_{N},

see for instance [FLPW24, Definition 1.4]. We have the following two lemmas for 𝒢(N)\mathcal{G}^{(N)}.

Lemma 2.4.

Fix κ(4,8)\kappa\in(4,8). For all (x1,,x2N)𝔛2N(x_{1},\ldots,x_{2N})\in\mathfrak{X}_{2N},

𝒢(N)(x1,,x2N)=αLPNν(α,¯)𝒵α(x1,,x2N),\mathcal{G}^{(N)}(x_{1},\ldots,x_{2N})=\sum_{\alpha\in\mathrm{LP}_{N}}\mathcal{M}_{\nu}(\alpha,\boldsymbol{\underline{\cap\cap}})\mathcal{Z}_{\alpha}(x_{1},\ldots,x_{2N}), (2.8)

where ν=2|cos(4π/κ)|\nu=2|\cos(4\pi/\kappa)| and ν(α,¯)=ν(α,¯)\mathcal{M}_{\nu}(\alpha,\boldsymbol{\underline{\cap\cap}})=\nu^{\ell(\alpha,\boldsymbol{\underline{\cap\cap}})} and (α,¯)\ell(\alpha,\boldsymbol{\underline{\cap\cap}}) denotes the number of loops in the meander formed from the two link patterns α\alpha and ¯\boldsymbol{\underline{\cap\cap}}.

Proof.

See [FPW24, Proposition 1.10] and [FLPW24, Proposition 1.11]. ∎

Lemma 2.5.

Fix κ(4,8)\kappa\in(4,8). There exists constant C(0,)C\in(0,\infty) (depending only on NN and κ\kappa) such that for all (x1,,x2N)𝔛2N(x_{1},\ldots,x_{2N})\in\mathfrak{X}_{2N},

0<𝒢(N)(x1,,x2N)C1j<k2N|xkxj|μkj,where μkj={10/κ,if |xkxj|1,6/κ,if |xkxj|<1.0<\mathcal{G}^{(N)}(x_{1},\ldots,x_{2N})\leq C\prod_{1\leq j<k\leq 2N}|x_{k}-x_{j}|^{\mu_{kj}},\quad\text{where }\mu_{kj}=\begin{cases}10/\kappa,&\text{if }|x_{k}-x_{j}|\geq 1,\\ -6/\kappa,&\text{if }|x_{k}-x_{j}|<1.\end{cases} (2.9)
Proof.

The second term in the integrand (2.7) can be bounded from above: for 1r<sN1\leq r<s\leq N,

|usur||x2sx2r1|.|u_{s}-u_{r}|\leq|x_{2s}-x_{2r-1}|.

The third term in the integrand (2.7) can be bounded from above: for 1qN1\leq q\leq N,

1j2N|uqxj|4/κ1j<2q1|x2q1xj|4/κ×|uqx2q1|4/κ|uqx2q|4/κ×2q<j2N|x2qxj|4/κ.\prod_{1\leq j\leq 2N}|u_{q}-x_{j}|^{-4/\kappa}\leq\prod_{1\leq j<2q-1}|x_{2q-1}-x_{j}|^{-4/\kappa}\times|u_{q}-x_{2q-1}|^{-4/\kappa}|u_{q}-x_{2q}|^{-4/\kappa}\times\prod_{2q<j\leq 2N}|x_{2q}-x_{j}|^{-4/\kappa}.

Note that, using change of variables v=uqx2q1x2qx2q1v=\frac{u_{q}-x_{2q-1}}{x_{2q}-x_{2q-1}},

x2q1x2q|uqx2q1|4/κ|uqx2q|4/κduq=\displaystyle\int_{x_{2q-1}}^{x_{2q}}|u_{q}-x_{2q-1}|^{-4/\kappa}|u_{q}-x_{2q}|^{-4/\kappa}\mathrm{d}u_{q}= (x2qx2q1)18/κ01v4/κ(1v)4/κdv\displaystyle(x_{2q}-x_{2q-1})^{1-8/\kappa}\int_{0}^{1}v^{-4/\kappa}(1-v)^{-4/\kappa}\mathrm{d}v
=\displaystyle= (x2qx2q1)18/κΓ(14/κ)2Γ(28/κ).\displaystyle(x_{2q}-x_{2q-1})^{1-8/\kappa}\frac{\Gamma(1-4/\kappa)^{2}}{\Gamma(2-8/\kappa)}.

Plugging these three observations into (2.6), there exists constant C(0,)C\in(0,\infty) (depending only on NN and κ\kappa),

𝒢(N)(x1,,x2N)\displaystyle\mathcal{G}^{(N)}(x_{1},\ldots,x_{2N})\leq C1j<k2N(xkxj)2/κ×1r<sN|x2sx2r1|8/κ\displaystyle C\prod_{1\leq j<k\leq 2N}(x_{k}-x_{j})^{2/\kappa}\times\prod_{1\leq r<s\leq N}|x_{2s}-x_{2r-1}|^{8/\kappa} (2.10)
×1qN1j<2q1|x2q1xj|4/κ×|x2qx2q1|18/κ×2q<j2N|x2qxj|4/κ.\displaystyle\times\prod_{1\leq q\leq N}\prod_{1\leq j<2q-1}|x_{2q-1}-x_{j}|^{-4/\kappa}\times|x_{2q}-x_{2q-1}|^{1-8/\kappa}\times\prod_{2q<j\leq 2N}|x_{2q}-x_{j}|^{-4/\kappa}.

In the right-hand side of (2.10),

  • the accumulated exponent for the term |x2qx2q1||x_{2q}-x_{2q-1}| with 1qN1\leq q\leq N is given by

    2/κ+18/κ=16/κ;2/\kappa+1-8/\kappa=1-6/\kappa;
  • the accumulated exponent for the term |x2qx2r1||x_{2q}-x_{2r-1}| with 1r<qN1\leq r<q\leq N is given by

    2/κ+8/κ=10/κ;2/\kappa+8/\kappa=10/\kappa;
  • the accumulated exponent for the term |x2q+1x2q||x_{2q+1}-x_{2q}| with 1qN11\leq q\leq N-1 is given by

    2/κ4/κ=2/κ;2/\kappa-4/\kappa=-2/\kappa;
  • the accumulated exponent for the term |x2j+1x2q||x_{2j+1}-x_{2q}| with 1q<jN11\leq q<j\leq N-1 is given by

    2/κ4/κ4/κ=6/κ;2/\kappa-4/\kappa-4/\kappa=-6/\kappa;
  • the accumulated exponent for the term |x2rx2q||x_{2r}-x_{2q}| with 1q<rN1\leq q<r\leq N is given by

    2/κ4/κ=2/κ;2/\kappa-4/\kappa=-2/\kappa;
  • the accumulated exponent for the term |x2r1x2q1||x_{2r-1}-x_{2q-1}| with 1q<rN1\leq q<r\leq N is given by

    2/κ4/κ=2/κ.2/\kappa-4/\kappa=-2/\kappa.

Combining these observations, we obtain (2.9) as desired. ∎

Proof of Lemma 2.3.

From (2.8), for any αLPN\alpha\in\mathrm{LP}_{N}, we have 𝒢(N)ν(α,¯)𝒵α\mathcal{G}^{(N)}\geq\mathcal{M}_{\nu}(\alpha,\boldsymbol{\underline{\cap\cap}})\mathcal{Z}_{\alpha}. Combining with (2.9), we obtain (2.5) as desired. ∎

2.3 Chordal NN-SLEκ\mathrm{SLE}_{\kappa}

The law of chordal NN-SLE\mathrm{SLE} can be described by the pure partition functions defined in Section 2.1. To this end, we first introduce usual parameterization. We fix N1N\geq 1 and let 𝒙=(x1,,x2N)𝔛2N\boldsymbol{x}=(x_{1},\ldots,x_{2N})\in\mathfrak{X}_{2N}. Let 𝜼=(η1,,η2N)\boldsymbol{\eta}=(\eta_{1},\ldots,\eta_{2N}) be 2N2N-tuple of continuous non-self-crossing curves ηj:[0,Tj]¯\eta_{j}:[0,T_{j}]\to\overline{\mathbb{H}} such that ηj(0)=xj\eta_{j}(0)=x_{j}. For j{1,,2N}j\in\{1,\ldots,2N\}, let HtjH_{t}^{j} be the unbounded connected component of ηj([0,t])\mathbb{H}\setminus\eta_{j}([0,t]). Let gtj:Htjg_{t}^{j}:H_{t}^{j}\to\mathbb{H} be the unique conformal transformation with limz|gtj(z)z|=0\lim_{z\to\infty}|g_{t}^{j}(z)-z|=0. We say that ηj=(ηj(t),0t<Tj)\eta_{j}=(\eta_{j}(t),0\leq t<T_{j}) has aa-usual parameterization if for any t<Tjt<T_{j},

gtj(z)=z+atz+o(1/|z|), as z.\displaystyle g_{t}^{j}(z)=z+\frac{at}{z}+o(1/|z|),\quad\text{ as }z\to\infty.

Under this parameterization, the half-plane capacity of ηj[0,t]\eta_{j}[0,t] is atat and gtjg_{t}^{j} satisfies the following chordal Loewner equation

tgtj(z)=agtj(z)Wtj,zHtj,t<Tj.\displaystyle\partial_{t}{g}_{t}^{j}(z)=\frac{a}{g^{j}_{t}(z)-W_{t}^{j}},\quad\forall z\in H_{t}^{j},\quad\forall t<T_{j}.

We say that (ηj(t),0t<Tj)(\eta_{j}(t),0\leq t<T_{j}) is the chordal Loewner chain with aa-usual parameterization and with driving function (Wtj,0t<Tj)(W_{t}^{j},0\leq t<T_{j}).

Let (γ1,,γN)α(𝒙)(\gamma_{1},\ldots,\gamma_{N})\sim\mathbb{P}_{\alpha}^{(\boldsymbol{x})} be the chordal NN-SLEκ\mathrm{SLE}_{\kappa} in polygon (;x1,,x2N)(\mathbb{H};x_{1},\ldots,x_{2N}) associated to link pattern α={{a1,b1},,{aN,bN}}LPN\alpha=\{\{a_{1},b_{1}\},\ldots,\{a_{N},b_{N}\}\}\in\mathrm{LP}_{N}. We view it as a 2N2N-tuple of continuous non-self-crossing curves 𝜼=(η1,,η2N)\boldsymbol{\eta}=(\eta_{1},\ldots,\eta_{2N}) as described in Section 1.1: for j{1,,N}j\in\{1,\ldots,N\}, we define ηaj\eta_{a_{j}} to be γj\gamma_{j} and ηbj\eta_{b_{j}} to be the time-reversal of γj\gamma_{j}. In this way, ηj\eta_{j} is a continuous curve in \mathbb{H} starting from xjx_{j} for j{1,,2N}j\in\{1,\ldots,2N\}. For j{1,,2N}j\in\{1,\ldots,2N\}, if we parameterize ηj\eta_{j} with aa-usual parameterization, then its driving function (Wtj,t0)(W^{j}_{t},t\geq 0) satisfies (see [PW19, Proposition 4.10] for κ(0,4]\kappa\in(0,4] and [FLPW24, Theorem 1.21] for κ(4,8)\kappa\in(4,8))

{dWtj=dBtj+(jlog𝒵α)(Vt1,,Vtj1,Wtj,Vtj+1,,Vt2N)dt,W0j=xj;dVtk=aVtkWtjdt,V0k=xk,k{1,,j1,j+1,,2N};\begin{cases}\mathrm{d}W_{t}^{j}=\mathrm{d}B_{t}^{j}+(\partial_{j}\log\mathcal{Z}_{\alpha})(V^{1}_{t},\ldots,V^{j-1}_{t},W_{t}^{j},V^{j+1}_{t},\ldots,V^{2N}_{t})\mathrm{d}t,\quad W_{0}^{j}=x_{j};\\ \mathrm{d}V_{t}^{k}=\frac{a}{V_{t}^{k}-W_{t}^{j}}\mathrm{d}t,\quad V_{0}^{k}=x_{k},\quad k\in\{1,\ldots,j-1,j+1,\ldots,2N\};\end{cases} (2.11)

where (Btj,t0)(B^{j}_{t},t\geq 0) is a standard Brownian motion.

2.4 Dyson Brownian motion

In this section, we recall a classical result for Dyson Brownian motion. For each v>0v>0 and 𝒙=(x1,,x2N)𝔛2N\boldsymbol{x}=(x_{1},\ldots,x_{2N})\in\mathfrak{X}_{2N}, define

fv(𝒙)=1j<k2N(xkxj)v.f_{v}(\boldsymbol{x})=\prod_{1\leq j<k\leq 2N}(x_{k}-x_{j})^{v}.

Note that 𝒵(𝒙):=𝒵(x1,,x2N)\mathcal{Z}_{*}(\boldsymbol{x}):=\mathcal{Z}_{*}(x_{1},\ldots,x_{2N}) in (1.4) is exactly fa(𝒙)f_{a}(\boldsymbol{x}).

Lemma 2.6.

Fix κ(0,8)\kappa\in(0,8) and a=2/κa=2/\kappa. Denote by p(t;𝐱,)p_{*}(t;\boldsymbol{x},\cdot) the transition density for Dyson Brownian motion (1.2) with parameter β=8/κ\beta=8/\kappa. Then, for any 𝐱,𝐲𝔛2N\boldsymbol{x},\boldsymbol{y}\in\mathfrak{X}_{2N},

p(t;𝒙,𝒚)=1(1t)BNf4a(𝒚)exp(|𝒚|22t)(1+O(|𝒙|t)),p_{*}(t;\boldsymbol{x},\boldsymbol{y})=\mathcal{I}_{*}^{-1}\left(\frac{1}{\sqrt{t}}\right)^{B_{N}}f_{4a}(\boldsymbol{y})\exp\left(-\frac{|\boldsymbol{y}|^{2}}{2t}\right)\left(1+O\left(\frac{|\boldsymbol{x}|}{\sqrt{t}}\right)\right), (2.12)

where

BN=2N(8N4+κ)κ,B_{N}=\frac{2N(8N-4+\kappa)}{\kappa},

and (0,)\mathcal{I}_{*}\in(0,\infty) is a normalization constant

=𝔛2Nf4a(𝒙)e12|𝒙|2d𝒙.\mathcal{I}_{*}=\int_{\mathfrak{X}_{2N}}f_{4a}(\boldsymbol{x})e^{-\frac{1}{2}|\boldsymbol{x}|^{2}}\mathrm{d}\boldsymbol{x}. (2.13)
Proof.

The explicit formula of transition density p(t;𝒙,𝒚)p_{*}(t;\boldsymbol{x},\boldsymbol{y}) can be obtained from [Rös98, Equation (4.5)]:

p(t;𝒙,𝒚)=1(1t)BNf4a(𝒚)exp(|𝒙|2+|𝒚|22t)J2aA(𝒙t,𝒚t),𝒙,𝒚𝔛2N,p_{*}(t;\boldsymbol{x},\boldsymbol{y})=\mathcal{I}_{*}^{-1}\left(\frac{1}{\sqrt{t}}\right)^{B_{N}}f_{4a}(\boldsymbol{y})\exp\left(-\frac{|\boldsymbol{x}|^{2}+|\boldsymbol{y}|^{2}}{2t}\right)J_{2a}^{A}\left(\frac{\boldsymbol{x}}{\sqrt{t}},\frac{\boldsymbol{y}}{\sqrt{t}}\right),\quad\boldsymbol{x},\boldsymbol{y}\in\mathfrak{X}_{2N},

where J2aAJ_{2a}^{A} is multivariate Bessel function of type AA with multiplicity k=2ak=2a and is analytic on 2N×2N\mathbb{R}^{2N}\times\mathbb{R}^{2N} with J2aA(𝟎,𝒚)=1J_{2a}^{A}(\boldsymbol{0},\boldsymbol{y})=1 (see [Rös98, Section 2] or [Voi19, Section 1]). Letting |𝒙|/t0|\boldsymbol{x}|/\sqrt{t}\to 0, we obtain (2.12) and complete the proof. ∎

The following lemma will be used in the proof of Proposition 1.3.

Lemma 2.7.

The constant

𝒥α=𝔛2Nf4a(𝒙)Gα(𝒙)1e12|𝒙|2d𝒙\mathcal{J}_{\alpha}=\int_{\mathfrak{X}_{2N}}f_{4a}(\boldsymbol{x})G_{\alpha}(\boldsymbol{x})^{-1}e^{-\frac{1}{2}|\boldsymbol{x}|^{2}}\mathrm{d}\boldsymbol{x} (2.14)

is finite, where GαG_{\alpha} is the Green’s function in Definition 1.2.

Proof.

Note that

0<f4a(𝒙)Gα(𝒙)1e12|𝒙|2=f4a(𝒙)𝒵α(𝒙)𝒵(𝒙)1e12|𝒙|2=f3a(𝒙)𝒵α(𝒙)e12|𝒙|2.0<f_{4a}(\boldsymbol{x})G_{\alpha}(\boldsymbol{x})^{-1}e^{-\frac{1}{2}|\boldsymbol{x}|^{2}}=f_{4a}(\boldsymbol{x})\mathcal{Z}_{\alpha}(\boldsymbol{x})\mathcal{Z}_{*}(\boldsymbol{x})^{-1}e^{-\frac{1}{2}|\boldsymbol{x}|^{2}}=f_{3a}(\boldsymbol{x})\mathcal{Z}_{\alpha}(\boldsymbol{x})e^{-\frac{1}{2}|\boldsymbol{x}|^{2}}.

From the power-law bound (2.4) and (2.5) and the relation (2.1), we know that f3a(𝒙)𝒵α(𝒙)e12|𝒙|2f_{3a}(\boldsymbol{x})\mathcal{Z}_{\alpha}(\boldsymbol{x})e^{-\frac{1}{2}|\boldsymbol{x}|^{2}} is integrable on 𝔛2N\mathfrak{X}_{2N}. Thus 𝒥α\mathcal{J}_{\alpha} is a finite constant. ∎

3 Proof of Theorem 1.1 and Proposition 1.3

3.1 Transition density and proof of Theorem 1.1

Fix N1N\geq 1 and 𝒙=(x1,,x2N)𝔛2N\boldsymbol{x}=(x_{1},\ldots,x_{2N})\in\mathfrak{X}_{2N}. To derive transition density pα(t;,)p_{\alpha}(t;\cdot,\cdot) for driving functions of chordal NN-SLEκ\mathrm{SLE}_{\kappa}, we derive Radon-Nikodym derivatives between three measures:

  • Recall that α(𝒙)\mathbb{P}_{\alpha}^{(\boldsymbol{x})} denotes the law of chordal NN-SLEκ\mathrm{SLE}_{\kappa} in polygon (;x1,,x2N)(\mathbb{H};x_{1},\ldots,x_{2N}) associated to link pattern αLPN\alpha\in\mathrm{LP}_{N}. We view it as a measure on 2N2N-tuples of curves 𝜼=(η1,,η2N)\boldsymbol{\eta}=(\eta_{1},\ldots,\eta_{2N}).

  • Denote by (𝒙)\mathbb{P}_{*}^{(\boldsymbol{x})} the law of 2N2N-tuples of curves 𝜼=(η1,,η2N)\boldsymbol{\eta}=(\eta_{1},\ldots,\eta_{2N}) in polygon (;x1,,x2N)(\mathbb{H};x_{1},\ldots,x_{2N}) driven by Dyson Brownian motion (1.2).

  • Denote by \mathbb{P} the law on 𝜼=(η1,,η2N)\boldsymbol{\eta}=(\eta_{1},\ldots,\eta_{2N}) where (η1,,η2N)(\eta_{1},\ldots,\eta_{2N}) are independent chordal SLEκ\mathrm{SLE}_{\kappa} in \mathbb{H} started from (x1,,x2N)(x_{1},\ldots,x_{2N}) respectively.

We fix the same parameters as in (2.1). We parameterize these curves with aa-common parameterization and recall from Section 2.3 that (Wtj,t0)(W_{t}^{j},t\geq 0) is the driving function of (ηj(t),t0)(\eta^{j}(t),t\geq 0) and from Section 1.1 that {(Xtj,t0)}1j2N\{(X_{t}^{j},t\geq 0)\}_{1\leq j\leq 2N} is driving functions of 𝜼=(η1,,η2N)\boldsymbol{\eta}=(\eta_{1},\ldots,\eta_{2N}) as a chordal Loewner chain with aa-common parameterization. Under aa-common parameterization, recall that (t,t0)(\mathcal{F}_{t},t\geq 0) is the filtration generated by (𝜼(t),t0)(\boldsymbol{\eta}(t),t\geq 0) and TT is the lifetime of driving functions (𝑿t,t0)(\boldsymbol{X}_{t},t\geq 0), and we define the following normalized conformal maps for t<Tt<T:

  • gtjg_{t}^{j} is the conformal map from the unbounded connected component of ηj([0,t])\mathbb{H}\setminus\eta_{j}([0,t]) onto \mathbb{H} with limz|gtj(z)z|=0\lim_{z\to\infty}|g_{t}^{j}(z)-z|=0;

  • gtg_{t} is the conformal map from the unbounded connected component of jηj([0,t])\mathbb{H}\setminus\cup_{j}\eta_{j}([0,t]) onto \mathbb{H} with gt(z)=z+2aNt/z+o(1/|z|)g_{t}(z)=z+2aNt/z+o(1/|z|) as zz\to\infty;

  • gt,jg_{t,j} is the conformal map from the unbounded connected component of kjgtj(ηk([0,t]))\mathbb{H}\setminus\cup_{k\neq j}g_{t}^{j}(\eta_{k}([0,t])) onto \mathbb{H} with limz|gt,j(z)z|=0\lim_{z\to\infty}|g_{t,j}(z)-z|=0.

Then we have gt=gt,jgtjg_{t}=g_{t,j}\circ g_{t}^{j} for 1j2N1\leq j\leq 2N. For a conformal map gg, we denote the Schwarzian derivative of gg by

Sg:=(g′′g)12(g′′g)2=g′′′g32(g′′g)2.Sg:=\left(\frac{g^{\prime\prime}}{g^{\prime}}\right)^{\prime}-\frac{1}{2}\left(\frac{g^{\prime\prime}}{g^{\prime}}\right)^{2}=\frac{g^{\prime\prime\prime}}{g^{\prime}}-\frac{3}{2}\left(\frac{g^{\prime\prime}}{g^{\prime}}\right)^{2}.
Lemma 3.1.

The Radon-Nikodym derivative between (𝐱)\mathbb{P}_{*}^{(\boldsymbol{x})} and \mathbb{P} when both measures are restricted to t\mathcal{F}_{t} and {T>t}\{T>t\} is given by

d(𝒙)[|t]d[|t{T>t}]=MtM0,t>0,\frac{\mathrm{d}\mathbb{P}_{*}^{(\boldsymbol{x})}[\cdot|_{\mathcal{F}_{t}}]}{\mathrm{d}\mathbb{P}[\cdot|_{\mathcal{F}_{t}\cap\{T>t\}}]}=\frac{M^{*}_{t}}{M^{*}_{0}},\quad t>0,

where

Mt=j=12Ngt,j(Wtj)hexp(ac120tSgs,j(Wsj)ds)×𝒵(Xt1,,Xt2N),M^{*}_{t}=\prod_{j=1}^{2N}g_{t,j}^{\prime}(W_{t}^{j})^{h}\exp\left(-\frac{ac}{12}\int_{0}^{t}Sg_{s,j}(W_{s}^{j})\mathrm{d}s\right)\times\mathcal{Z}_{*}(X^{1}_{t},\ldots,X^{2N}_{t}), (3.1)

and 𝒵\mathcal{Z}_{*} is the partition function (1.4).

Lemma 3.2.

The Radon-Nikodym derivative between α(𝐱)\mathbb{P}_{\alpha}^{(\boldsymbol{x})} and \mathbb{P} when both measures are restricted to t\mathcal{F}_{t} and {T>t}\{T>t\} is given by

dα(𝒙)[|t{T>t}]d[|t{T>t}]=MtαM0α,t>0,\frac{\mathrm{d}\mathbb{P}_{\alpha}^{(\boldsymbol{x})}[\cdot|_{\mathcal{F}_{t}\cap\{T>t\}}]}{\mathrm{d}\mathbb{P}[\cdot|_{\mathcal{F}_{t}\cap\{T>t\}}]}=\frac{M^{\alpha}_{t}}{M^{\alpha}_{0}},\quad t>0,

where

Mtα=j=12Ngt,j(Wtj)hexp(ac120tSgs,j(Wsj)ds)×𝒵α(Xt1,,Xt2N),M_{t}^{\alpha}=\prod_{j=1}^{2N}g_{t,j}^{\prime}(W_{t}^{j})^{h}\exp\left(-\frac{ac}{12}\int_{0}^{t}Sg_{s,j}(W_{s}^{j})\mathrm{d}s\right)\times\mathcal{Z}_{\alpha}(X^{1}_{t},\ldots,X^{2N}_{t}), (3.2)

and 𝒵α\mathcal{Z}_{\alpha} is the pure partition function in Section 2.3. Moreover, under α(𝐱)\mathbb{P}_{\alpha}^{(\boldsymbol{x})}, the driving functions (𝐗t,0t<T)(\boldsymbol{X}_{t},0\leq t<T) satisfies the SDE (1.1) where {Bj}1j2N\{B^{j}\}_{1\leq j\leq 2N} are independent standard Brownian motions.

We will prove Lemmas 3.1 and 3.2 in Section 3.2 where we introduce local martingales with 2N2N time parameters. Assuming these lemmas, we are able to derive the transition density for driving functions of chordal NN-SLEκ\mathrm{SLE}_{\kappa}.

Proof of Theorem 1.1.

For each t>0t>0, we obtain from Lemmas 3.1 and 3.2 the Radon-Nikodym derivative of α(𝒙)\mathbb{P}_{\alpha}^{(\boldsymbol{x})} against (𝒙)\mathbb{P}_{*}^{(\boldsymbol{x})} when both measures are restricted to t\mathcal{F}_{t} and {T>t}\{T>t\}:

dα(𝒙)[|t{T>t}]d(𝒙)[|t]=Mtα/M0αMt/M0=𝒵α(𝑿t)/𝒵α(𝒙)𝒵(𝑿t)/𝒵(𝒙)=Gα(𝒙)Gα(𝑿t).\frac{\mathrm{d}\mathbb{P}_{\alpha}^{(\boldsymbol{x})}[\cdot|_{\mathcal{F}_{t}\cap\{T>t\}}]}{\mathrm{d}\mathbb{P}_{*}^{(\boldsymbol{x})}[\cdot|_{\mathcal{F}_{t}}]}=\frac{M^{\alpha}_{t}/M^{\alpha}_{0}}{M^{*}_{t}/M^{*}_{0}}=\frac{\mathcal{Z}_{\alpha}(\boldsymbol{X}_{t})/\mathcal{Z}_{\alpha}(\boldsymbol{x})}{\mathcal{Z}_{*}(\boldsymbol{X}_{t})/\mathcal{Z}_{*}(\boldsymbol{x})}=\frac{G_{\alpha}(\boldsymbol{x})}{G_{\alpha}(\boldsymbol{X}_{t})}.

Since (𝑿t,t0)(\boldsymbol{X}_{t},t\geq 0) is a Markov process with transition density p(t;𝒙,𝒚)p_{*}(t;\boldsymbol{x},\boldsymbol{y}) under (𝒙)\mathbb{P}_{*}^{(\boldsymbol{x})}, the Radon-Nikodym derivative gives the transition density pα(t;𝒙,𝒚)p_{\alpha}(t;\boldsymbol{x},\boldsymbol{y}) as desired. ∎

3.2 2N2N-time local martingales

We will prove Lemmas 3.1 and 3.2 in this section. To this end, we introduce local martingales with 2N2N time parameters. Consider 2N2N chordal curves 𝜼=(η1,,η2N)\boldsymbol{\eta}=(\eta_{1},\ldots,\eta_{2N}) in the polygon (;x1,,x2N)(\mathbb{H};x_{1},\ldots,x_{2N}). Suppose each ηj\eta_{j} is parameterized with aa-usual parameterization and we get 2N2N time parameters 𝒕=(t1,,t2N)+2N\boldsymbol{t}=(t_{1},\ldots,t_{2N})\in\mathbb{R}_{+}^{2N}. We define the following normalized conformal maps:

  • gtjjg_{t_{j}}^{j} is the conformal map from the unbounded connected component of ηj([0,tj])\mathbb{H}\setminus\eta_{j}([0,t_{j}]) onto \mathbb{H} with gtjj(z)=z+atj/z+o(1/|z|)g_{t_{j}}^{j}(z)=z+at_{j}/z+o(1/|z|) as zz\to\infty, 1j2N1\leq j\leq 2N.

  • g𝒕g_{\boldsymbol{t}} is the conformal map from the unbounded connected component of j=12Nηj([0,tj])\mathbb{H}\setminus\bigcup_{j=1}^{2N}\eta_{j}([0,t_{j}]) onto \mathbb{H} with g𝒕(z)=z+(𝒕)/z+o(1/|z|)g_{\boldsymbol{t}}(z)=z+\aleph(\boldsymbol{t})/z+o(1/|z|) as zz\to\infty.

  • g𝒕,jg_{\boldsymbol{t},j} is the conformal map from the unbounded connecetd component of kjgtjj(ηk([0,tk]))\mathbb{H}\setminus\bigcup_{k\neq j}g_{t_{j}}^{j}(\eta_{k}([0,t_{k}])) onto \mathbb{H} with limz|g𝒕,j(z)z|=0\lim_{z\to\infty}|g_{\boldsymbol{t},j}(z)-z|=0, 1j2N1\leq j\leq 2N.

Then g𝒕=g𝒕,jgtjjg_{\boldsymbol{t}}=g_{\boldsymbol{t},j}\circ g_{t_{j}}^{j} for 1j2N1\leq j\leq 2N. Denote by (Wtjj,tj0)(W^{j}_{t_{j}},t_{j}\geq 0) the driving function of ηj\eta_{j} as a chordal Loewner chain with aa-usual parameterization. Let 𝑿𝒕=(X𝒕1,,X𝒕2N)\boldsymbol{X}_{\boldsymbol{t}}=(X^{1}_{\boldsymbol{t}},\ldots,X^{2N}_{\boldsymbol{t}}) with X𝒕j=g𝒕,j(Wtjj)X^{j}_{\boldsymbol{t}}=g_{\boldsymbol{t},j}(W_{t_{j}}^{j}), 1j2N1\leq j\leq 2N.

Lemma 3.3.

Recall that \mathbb{P} denotes the law on 𝛈=(η1,,η2N)\boldsymbol{\eta}=(\eta_{1},\ldots,\eta_{2N}) where (η1,,η2N)(\eta_{1},\ldots,\eta_{2N}) are independent chordal SLEκ\mathrm{SLE}_{\kappa} in \mathbb{H} started from (x1,,x2N)(x_{1},\ldots,x_{2N}) respectively. Then the process

M𝒕=j=12Ng𝒕,j(Wtjj)hexp(ac120tjSg𝒕,j(Wtjj)dtj)×𝒵(X𝒕1,,X𝒕2N)M^{*}_{\boldsymbol{t}}=\prod_{j=1}^{2N}g_{\boldsymbol{t},j}^{\prime}(W_{t_{j}}^{j})^{h}\exp\left(-\frac{ac}{12}\int_{0}^{t_{j}}Sg_{\boldsymbol{t},j}(W_{t_{j}}^{j})\mathrm{d}t_{j}\right)\times\mathcal{Z}_{*}(X^{1}_{\boldsymbol{t}},\ldots,X^{2N}_{\boldsymbol{t}}) (3.3)

is a 2N2N-time-parameter local martingale with respect to \mathbb{P}.

Proof.

Note that {Wj}1j2N\{W^{j}\}_{1\leq j\leq 2N} are independent standard Brownian motions under \mathbb{P}. Since X𝒕j=g𝒕,j(Wtjj)X_{\boldsymbol{t}}^{j}=g_{\boldsymbol{t},j}(W_{t_{j}}^{j}) for 1j2N1\leq j\leq 2N, by Itô’s formula and a standard calculation, we have

dX𝒕j=g𝒕,j(Wtjj)dWtjjhg𝒕,j′′(Wtjj)dtj+akj1X𝒕jX𝒕kg𝒕,k(Wtkk)2dtk,1j2N.\mathrm{d}X_{\boldsymbol{t}}^{j}=g_{\boldsymbol{t},j}^{\prime}(W_{t_{j}}^{j})\mathrm{d}W_{t_{j}}^{j}-hg_{\boldsymbol{t},j}^{\prime\prime}(W_{t_{j}}^{j})\mathrm{d}t_{j}+a\sum_{k\neq j}\frac{1}{X_{\boldsymbol{t}}^{j}-X_{\boldsymbol{t}}^{k}}g_{\boldsymbol{t},k}^{\prime}(W_{t_{k}}^{k})^{2}\mathrm{d}t_{k},\quad 1\leq j\leq 2N.

Note that M𝒕M_{\boldsymbol{t}}^{*} defined in (3.3) has a decomposition M𝒕=N𝒕×N𝒕M_{\boldsymbol{t}}^{*}=N_{\boldsymbol{t}}\times N_{\boldsymbol{t}}^{*}, where

N𝒕=j=12Ng𝒕,j(Wtjj)hexp(ac120tjSg𝒕,j(Wtjj)dtj)×exp(ahj=12N0tjkjg𝒕,j(Wtjj)2(X𝒕jX𝒕k)2dtj),N_{\boldsymbol{t}}=\prod_{j=1}^{2N}g_{\boldsymbol{t},j}^{\prime}(W_{t_{j}}^{j})^{h}\exp\left(-\frac{ac}{12}\int_{0}^{t_{j}}Sg_{\boldsymbol{t},j}(W_{t_{j}}^{j})\mathrm{d}t_{j}\right)\times\exp\left(ah\sum_{j=1}^{2N}\int_{0}^{t_{j}}\sum_{k\neq j}\frac{g_{\boldsymbol{t},j}^{\prime}(W_{t_{j}}^{j})^{2}}{(X_{\boldsymbol{t}}^{j}-X_{\boldsymbol{t}}^{k})^{2}}\mathrm{d}t_{j}\right), (3.4)

and

N𝒕=𝒵(X𝒕1,,X𝒕2N)×exp(ahj=12N0tjkjg𝒕,j(Wtjj)2(X𝒕jX𝒕k)2dtj).N_{\boldsymbol{t}}^{*}=\mathcal{Z}_{*}(X^{1}_{\boldsymbol{t}},\ldots,X^{2N}_{\boldsymbol{t}})\times\exp\left(-ah\sum_{j=1}^{2N}\int_{0}^{t_{j}}\sum_{k\neq j}\frac{g_{\boldsymbol{t},j}^{\prime}(W_{t_{j}}^{j})^{2}}{(X_{\boldsymbol{t}}^{j}-X_{\boldsymbol{t}}^{k})^{2}}\mathrm{d}t_{j}\right).

It is proved by direct calculation that N𝒕N_{\boldsymbol{t}} is a local martingale with respect to \mathbb{P} satisfying

dN𝒕=N𝒕j=12Nhg𝒕,j′′(Wtjj)g𝒕,j(Wtjj)dWtjj.\mathrm{d}N_{\boldsymbol{t}}=N_{\boldsymbol{t}}\sum_{j=1}^{2N}h\frac{g_{\boldsymbol{t},j}^{\prime\prime}(W_{t_{j}}^{j})}{g_{\boldsymbol{t},j}^{\prime}(W_{t_{j}}^{j})}\mathrm{d}W_{t_{j}}^{j}.

Denote by \mathbb{Q} the probability measure obtained by tilting \mathbb{P} by N𝒕N_{\boldsymbol{t}}. By Girsanov’s theorem, under \mathbb{Q}, we have

dX𝒕j=g𝒕,j(Wtjj)dBtjj+akj1X𝒕jX𝒕kg𝒕,k(Wtkk)2dtk,1j2N,\mathrm{d}X_{\boldsymbol{t}}^{j}=g_{\boldsymbol{t},j}^{\prime}(W_{t_{j}}^{j})\mathrm{d}B_{t_{j}}^{j}+a\sum_{k\neq j}\frac{1}{X^{j}_{\boldsymbol{t}}-X^{k}_{\boldsymbol{t}}}g_{\boldsymbol{t},k}^{\prime}(W_{t_{k}}^{k})^{2}\mathrm{d}t_{k},\quad 1\leq j\leq 2N,

where {Bj}1j2N\{B^{j}\}_{1\leq j\leq 2N} are independent standard Brownian motions.

It remains to prove that N𝒕N_{\boldsymbol{t}}^{*} is a local martingale with respect to \mathbb{Q}. By Itô’s formula, we have

dN𝒕=N𝒕j=12NkjaX𝒕jX𝒕kg𝒕,j(Wtjj)dBtjj.\mathrm{d}N_{\boldsymbol{t}}^{*}=N_{\boldsymbol{t}}^{*}\sum_{j=1}^{2N}\sum_{k\neq j}\frac{a}{X_{\boldsymbol{t}}^{j}-X_{\boldsymbol{t}}^{k}}g_{\boldsymbol{t},j}^{\prime}(W_{t_{j}}^{j})\mathrm{d}B_{t_{j}}^{j}.

This completes the proof. ∎

Proof of Lemma 3.1.

Denote by ^\hat{\mathbb{P}} the probability measure obtained by tilting \mathbb{P} by M𝒕M_{\boldsymbol{t}}^{*}, then under ^\hat{\mathbb{P}}, for 1j2N1\leq j\leq 2N,

dX𝒕j=g𝒕,j(Wtjj)dBtjj+akj1X𝒕jX𝒕kg𝒕,j(Wtjj)2dtj+akj1X𝒕jX𝒕kg𝒕,k(Wtkk)2dtk,\mathrm{d}X_{\boldsymbol{t}}^{j}=g_{\boldsymbol{t},j}^{\prime}(W_{t_{j}}^{j})\mathrm{d}B_{t_{j}}^{j}+a\sum_{k\neq j}\frac{1}{X^{j}_{\boldsymbol{t}}-X^{k}_{\boldsymbol{t}}}g_{\boldsymbol{t},j}^{\prime}(W_{t_{j}}^{j})^{2}\mathrm{d}t_{j}+a\sum_{k\neq j}\frac{1}{X^{j}_{\boldsymbol{t}}-X^{k}_{\boldsymbol{t}}}g_{\boldsymbol{t},k}^{\prime}(W_{t_{k}}^{k})^{2}\mathrm{d}t_{k}, (3.5)

where {Bj}1j2N\{B^{j}\}_{1\leq j\leq 2N} are independent Brownian motions under ^\hat{\mathbb{P}}. Under aa-common parameterization, we have dtj=g𝒕,j(Wtjj)2dt\mathrm{d}t_{j}=g_{\boldsymbol{t},j}^{\prime}(W_{t_{j}}^{j})^{-2}\mathrm{d}t for 1j2N1\leq j\leq 2N (see [HL21, Lemma 3.2] for similar proof), then the SDE (3.5) becomes SDE (1.2), and the martingale (3.3) becomes (3.1). This implies ^=(𝒙)\hat{\mathbb{P}}=\mathbb{P}_{*}^{(\boldsymbol{x})} and completes the proof. ∎

Lemma 3.4.

Recall that \mathbb{P} denotes the law on 𝛈=(η1,,η2N)\boldsymbol{\eta}=(\eta_{1},\ldots,\eta_{2N}) where (η1,,η2N)(\eta_{1},\ldots,\eta_{2N}) are independent chordal SLEκ\mathrm{SLE}_{\kappa} in \mathbb{H} started from (x1,,x2N)(x_{1},\ldots,x_{2N}) respectively. Then the process

M𝒕α=j=12Ng𝒕,j(Wtjj)hexp(ac120tjSg𝒕,j(Wtjj)dtj)×𝒵α(X𝒕1,,X𝒕2N)M^{\alpha}_{\boldsymbol{t}}=\prod_{j=1}^{2N}g_{\boldsymbol{t},j}^{\prime}(W_{t_{j}}^{j})^{h}\exp\left(-\frac{ac}{12}\int_{0}^{t_{j}}Sg_{\boldsymbol{t},j}(W_{t_{j}}^{j})\mathrm{d}t_{j}\right)\times\mathcal{Z}_{\alpha}(X^{1}_{\boldsymbol{t}},\ldots,X^{2N}_{\boldsymbol{t}}) (3.6)

is a 2N2N-time-parameter local martingale with respect to \mathbb{P}.

Proof.

Note that M𝒕M_{\boldsymbol{t}}^{*} defined in (3.6) has a decomposition M𝒕α=N𝒕×N𝒕αM_{\boldsymbol{t}}^{\alpha}=N_{\boldsymbol{t}}\times N_{\boldsymbol{t}}^{\alpha}, where N𝒕N_{\boldsymbol{t}} is defined in (3.4) and

N𝒕α=𝒵α(X𝒕1,,X𝒕2N)×exp(ahj=12N0tjkjg𝒕,j(Wtjj)2(X𝒕jX𝒕k)2dtj).N_{\boldsymbol{t}}^{\alpha}=\mathcal{Z}_{\alpha}(X^{1}_{\boldsymbol{t}},\ldots,X^{2N}_{\boldsymbol{t}})\times\exp\left(-ah\sum_{j=1}^{2N}\int_{0}^{t_{j}}\sum_{k\neq j}\frac{g_{\boldsymbol{t},j}^{\prime}(W_{t_{j}}^{j})^{2}}{(X_{\boldsymbol{t}}^{j}-X_{\boldsymbol{t}}^{k})^{2}}\mathrm{d}t_{j}\right).

Recall from the proof of Lemma 3.3 that N𝒕N_{\boldsymbol{t}} is a local martingale with respect to \mathbb{P}, and \mathbb{Q} is the probability measure obtained by tilting \mathbb{P} by N𝒕N_{\boldsymbol{t}}. It remains to prove that N𝒕αN_{\boldsymbol{t}}^{\alpha} is a local martingale with respect to \mathbb{Q}. By Itô’s formula and the PDE (2.2) satisfied by 𝒵α\mathcal{Z}_{\alpha}, we have

dN𝒕α=N𝒕αj=12N(jlog𝒵α)(X𝒕1,,X𝒕2N)g𝒕,j(Wtjj)dBtjj.\mathrm{d}N_{\boldsymbol{t}}^{\alpha}=N_{\boldsymbol{t}}^{\alpha}\sum_{j=1}^{2N}(\partial_{j}\log\mathcal{Z}_{\alpha})(X_{\boldsymbol{t}}^{1},\ldots,X_{\boldsymbol{t}}^{2N})g_{\boldsymbol{t},j}^{\prime}(W_{t_{j}}^{j})\mathrm{d}B_{t_{j}}^{j}.

This completes the proof. ∎

Proof of Lemma 3.2.

Denote by ^\hat{\mathbb{P}} the probability measure obtained by tilting \mathbb{P} by M𝒕αM^{\alpha}_{\boldsymbol{t}} in (3.6), then under ^\hat{\mathbb{P}}, for 1j2N1\leq j\leq 2N,

dX𝒕j=g𝒕,j(Wtjj)dBtjj+(jlog𝒵α)(X𝒕1,,X𝒕2N)g𝒕,j(Wtjj)2dtj+akj1X𝒕jX𝒕kg𝒕,k(Wtkk)2dtk,\mathrm{d}X^{j}_{\boldsymbol{t}}=g_{\boldsymbol{t},j}^{\prime}(W_{t_{j}}^{j})\mathrm{d}B_{t_{j}}^{j}+(\partial_{j}\log\mathcal{Z}_{\alpha})(X^{1}_{\boldsymbol{t}},\ldots,X^{2N}_{\boldsymbol{t}})g_{\boldsymbol{t},j}^{\prime}(W_{t_{j}}^{j})^{2}\mathrm{d}t_{j}+a\sum_{k\neq j}\frac{1}{X^{j}_{\boldsymbol{t}}-X^{k}_{\boldsymbol{t}}}g_{\boldsymbol{t},k}^{\prime}(W_{t_{k}}^{k})^{2}\mathrm{d}t_{k}, (3.7)

where {Bj}1j2N\{B^{j}\}_{1\leq j\leq 2N} are independent standard Brownian motions under ^\hat{\mathbb{P}}. Suppose that {ηj([0,Tj])}1j2N\{\eta_{j}([0,T_{j}])\}_{1\leq j\leq 2N} are restrictions of {ηj}1j2N\{\eta_{j}\}_{1\leq j\leq 2N} such that, for each 1j2N1\leq j\leq 2N, the curve ηj([0,Tj])\eta_{j}([0,T_{j}]) does not disconnect any ηk([0,Tk])\eta_{k}([0,T_{k}]) with kjk\neq j from \infty. We will argue that ^\hat{\mathbb{P}} is the same as α(𝒙)\mathbb{P}_{\alpha}^{(\boldsymbol{x})} when restricted to {ηj([0,Tj])}1j2N\{\eta_{j}([0,T_{j}])\}_{1\leq j\leq 2N}.

On the one hand, by the domain Markov property of chordal NN-SLEκ\mathrm{SLE}_{\kappa}, (η1,,η2N)(\eta_{1},\ldots,\eta_{2N}) commute with each other in the following sense. Let τj\tau_{j} be any stopping time before TjT_{j} for 1j2N1\leq j\leq 2N. For each ηj\eta_{j}, conditionally on kjηk([0,τk])\bigcup_{k\neq j}\eta_{k}([0,\tau_{k}]), let gg be the conformal map from the unbounded connected component of kjηk([0,τk])\mathbb{H}\setminus\bigcup_{k\neq j}\eta_{k}([0,\tau_{k}]) onto \mathbb{H} with limz|g(z)z|=0\lim_{z\to\infty}|g(z)-z|=0, then the gg-image of ηj\eta_{j} is a chordal Loewner chain in \mathbb{H} started from g(xj)g(x_{j}) with aa-usual parameterization and with driving function (Wtj,t0)(W^{j}_{t},t\geq 0) satisfying SDE (2.11) up to a time change.

On the other hand, we know from (3.7) that under the measure ^\hat{\mathbb{P}} and conditionally on kjηk([0,τk])\bigcup_{k\neq j}\eta_{k}([0,\tau_{k}]), (g(ηj(tj)),tj0)(g(\eta_{j}(t_{j})),t_{j}\geq 0) is a chordal Loewner chain in \mathbb{H} started from g(xj)g(x_{j}) and driven by (Xtjj,tj0)(X_{t_{j}}^{j},t_{j}\geq 0) satisfying the SDE

dXtjj=g𝒕,j(Wtjj)dBtjj+(jlog𝒵α)(X𝒕1,,X𝒕2N)g𝒕,j(Wtjj)2dtj,\mathrm{d}X_{t_{j}}^{j}=g_{\boldsymbol{t},j}^{\prime}(W_{t_{j}}^{j})\mathrm{d}B_{t_{j}}^{j}+(\partial_{j}\log\mathcal{Z}_{\alpha})(X^{1}_{\boldsymbol{t}},\ldots,X^{2N}_{\boldsymbol{t}})g_{\boldsymbol{t},j}^{\prime}(W_{t_{j}}^{j})^{2}\mathrm{d}t_{j},

where 𝒕=(τ1,,τj1,tj,τj+1,,τ2N)\boldsymbol{t}=(\tau_{1},\ldots,\tau_{j-1},t_{j},\tau_{j+1},\ldots,\tau_{2N}), tj0t_{j}\geq 0. This is the same as SDE (2.11) after a time change, which implies that the chordal Loewner curves (η1,,η2N)(\eta_{1},\ldots,\eta_{2N}) under ^\hat{\mathbb{P}} locally commute with each other in the sense of the last paragraph. Therefore, the joint law α(𝒙)\mathbb{P}_{\alpha}^{(\boldsymbol{x})} of chordal NN-SLEκ\mathrm{SLE}_{\kappa} (η1,,η2N)(\eta_{1},\ldots,\eta_{2N}) coincides with ^\hat{\mathbb{P}} when restricted to {ηj([0,Tj])}1j2N\{\eta_{j}([0,T_{j}])\}_{1\leq j\leq 2N}.

Finally, we parameterize these 2N2N-tuple of curves with aa-common parameterization by making the time change dtj=g𝒕,j(Wtjj)2dt\mathrm{d}t_{j}=g_{\boldsymbol{t},j}^{\prime}(W_{t_{j}}^{j})^{-2}\mathrm{d}t, then the 2N2N-time-parameter local martingale (3.6) becomes MtαM_{t}^{\alpha} in (3.2), and the SDE (3.7) becomes SDE (1.1). Note also that, on the event {T>t}\{T>t\}, the collection {ηj([0,t])}1j2N\{\eta_{j}([0,t])\}_{1\leq j\leq 2N} has the property that the curve ηj([0,t])\eta_{j}([0,t]) does not disconnect any ηk([0,t])\eta_{k}([0,t]) with kjk\neq j from \infty. This completes the proof. ∎

3.3 Proof of Proposition 1.3

Recall that α(𝒙)\mathbb{P}_{\alpha}^{(\boldsymbol{x})} denotes the law of chordal NN-SLEκ\mathrm{SLE}_{\kappa} in polygon (;x1,,x2N)(\mathbb{H};x_{1},\ldots,x_{2N}) associated to link pattern αLPN\alpha\in\mathrm{LP}_{N}. We view it as a 2N2N-tuple of continuous non-self-crossing curves 𝜼=(η1,,η2N)\boldsymbol{\eta}=(\eta_{1},\ldots,\eta_{2N}) as described in Section 1.1: for j{1,,N}j\in\{1,\ldots,N\}, we define ηaj\eta_{a_{j}} to be γj\gamma_{j} and ηbj\eta_{b_{j}} to be the time-reversal of γj\gamma_{j}. In this way, ηj\eta_{j} is a continuous non-self-crossing curve in \mathbb{H} starting from xjx_{j} for j{1,,2N}j\in\{1,\ldots,2N\}. We parameterize 𝜼=(η1,,η2N)\boldsymbol{\eta}=(\eta_{1},\ldots,\eta_{2N}) with aa-common parameterization. We denote by (t,t0)(\mathcal{F}_{t},t\geq 0) the filtration generated by (𝜼(t),t0)(\boldsymbol{\eta}(t),t\geq 0) and by TT the lifetime of its driving function (𝑿t,t0)(\boldsymbol{X}_{t},t\geq 0). Under α(𝒙)\mathbb{P}_{\alpha}^{(\boldsymbol{x})}, the lifetime TT is finite almost surely. We will prove Proposition 1.3 in this section.

Proof of (1.5).

For each 𝒙=(x1,,x2N)𝔛2N\boldsymbol{x}=(x_{1},\ldots,x_{2N})\in\mathfrak{X}_{2N}, we have

α(𝒙)[T>t]\displaystyle\mathbb{P}_{\alpha}^{(\boldsymbol{x})}[T>t] =𝔛2Npα(t;𝒙,𝒚)d𝒚\displaystyle=\int_{\mathfrak{X}_{2N}}p_{\alpha}(t;\boldsymbol{x},\boldsymbol{y})\mathrm{d}\boldsymbol{y}
=𝔛2Np(t;𝒙,𝒚)Gα(𝒙)Gα(𝒚)d𝒚\displaystyle=\int_{\mathfrak{X}_{2N}}p_{*}(t;\boldsymbol{x},\boldsymbol{y})\frac{G_{\alpha}(\boldsymbol{x})}{G_{\alpha}(\boldsymbol{y})}\mathrm{d}\boldsymbol{y} (due to (1.3))
=𝔛2N1(1t)BNf4a(𝒚)exp(|𝒚|22t)(1+O(|𝒙|t))Gα(𝒙)Gα(𝒚)d𝒚\displaystyle=\int_{\mathfrak{X}_{2N}}\mathcal{I}_{*}^{-1}\left(\frac{1}{\sqrt{t}}\right)^{B_{N}}f_{4a}(\boldsymbol{y})\exp\left(-\frac{|\boldsymbol{y}|^{2}}{2t}\right)\left(1+O\left(\frac{|\boldsymbol{x}|}{\sqrt{t}}\right)\right)\frac{G_{\alpha}(\boldsymbol{x})}{G_{\alpha}(\boldsymbol{y})}\mathrm{d}\boldsymbol{y} (due to (2.12))
=1Gα(𝒙)(1t)BN(1+O(|𝒙|t))×𝔛2Nf4a(𝒚)Gα(𝒚)1exp(|𝒚|22t)d𝒚\displaystyle=\mathcal{I}_{*}^{-1}G_{\alpha}(\boldsymbol{x})\left(\frac{1}{\sqrt{t}}\right)^{B_{N}}\left(1+O\left(\frac{|\boldsymbol{x}|}{\sqrt{t}}\right)\right)\times\int_{\mathfrak{X}_{2N}}f_{4a}(\boldsymbol{y})G_{\alpha}(\boldsymbol{y})^{-1}\exp\left(-\frac{|\boldsymbol{y}|^{2}}{2t}\right)\mathrm{d}\boldsymbol{y}
=1Gα(𝒙)(1t)BN(1+O(|𝒙|t))×𝒥α×(1t)BN\displaystyle=\mathcal{I}_{*}^{-1}G_{\alpha}(\boldsymbol{x})\left(\frac{1}{\sqrt{t}}\right)^{B_{N}}\left(1+O\left(\frac{|\boldsymbol{x}|}{\sqrt{t}}\right)\right)\times\mathcal{J}_{\alpha}\times\left(\frac{1}{\sqrt{t}}\right)^{-B_{N}^{\prime}}

where 𝒥α\mathcal{J}_{\alpha} is the constant in (2.14) and the exponent BNB_{N}^{\prime} is obtained by change of variable in integration:

BN=N(12N12+3κ)κ.B_{N}^{\prime}=\frac{N(12N-12+3\kappa)}{\kappa}.

Note that

BNBN=2N(8N4+κ)κN(12N12+3κ)κ=N(4N+4κ)κ=A2N+.B_{N}-B_{N}^{\prime}=\frac{2N(8N-4+\kappa)}{\kappa}-\frac{N(12N-12+3\kappa)}{\kappa}=\frac{N(4N+4-\kappa)}{\kappa}=A_{2N}^{+}.

This completes the proof. ∎

Using the estimates in (1.5), we are able to show that the law α(𝒙)\mathbb{P}_{\alpha}^{(\boldsymbol{x})} conditional on the event {T>t}\{T>t\} will converge to (𝒙)\mathbb{P}_{*}^{(\boldsymbol{x})}. This implies the convergence of driving functions.

Corollary 3.5.

Consider the following two measures:

  • For t>0t>0, denote by α|t(𝒙)\mathbb{P}_{\alpha\,|\,t}^{(\boldsymbol{x})} the measure α(𝒙)\mathbb{P}_{\alpha}^{(\boldsymbol{x})} conditional on the event {T>t}\{T>t\}.

  • Recall that (𝒙)\mathbb{P}_{*}^{(\boldsymbol{x})} denotes the law of 2N2N-tuples of curves 𝜼=(η1,,η2N)\boldsymbol{\eta}=(\eta_{1},\ldots,\eta_{2N}) in polygon (;x1,,x2N)(\mathbb{H};x_{1},\ldots,x_{2N}) driven by Dyson Brownian motion (1.2).

Then for any s>0s>0, when both measures α|t(𝐱)\mathbb{P}_{\alpha\,|\,t}^{(\boldsymbol{x})} and (𝐱)\mathbb{P}_{*}^{(\boldsymbol{x})} are restricted to s\mathcal{F}_{s}, the total variation distance between the two measures goes to zero as tt\to\infty:

limtdistTV(α|t(𝒙)[|s],(𝒙)[|s])=0.\lim_{t\to\infty}\mathrm{dist}_{\mathrm{TV}}\left(\mathbb{P}_{\alpha\,|\,t}^{(\boldsymbol{x})}[\cdot\,|\,_{\mathcal{F}_{s}}],\mathbb{P}_{*}^{(\boldsymbol{x})}[\cdot\,|\,_{\mathcal{F}_{s}}]\right)=0.
Proof.

For any t>s>0t>s>0, from the domain Markov property of chordal NN-SLEκ\mathrm{SLE}_{\kappa} and Proposition 1.3, we have

dα|t(𝒙)[|s]dα(𝒙)[|s{T>s}]=α(𝑿s)[T>ts]α(𝒙)[T>t]=(tst)A2N+/2Gα(Xs1,,Xs2N)Gα(x1,,x2N)(1+O(|𝑿s|t)).\frac{\mathrm{d}\mathbb{P}_{\alpha\,|\,t}^{(\boldsymbol{x})}\left[\cdot|_{\mathcal{F}_{s}}\right]}{\mathrm{d}\mathbb{P}_{\alpha}^{(\boldsymbol{x})}\left[\cdot|_{\mathcal{F}_{s}\cap\{T>s\}}\right]}=\frac{\mathbb{P}_{\alpha}^{(\boldsymbol{X}_{s})}[T>t-s]}{\mathbb{P}_{\alpha}^{(\boldsymbol{x})}[T>t]}=\left(\frac{t-s}{t}\right)^{-A_{2N}^{+}/2}\frac{G_{\alpha}(X^{1}_{s},\ldots,X^{2N}_{s})}{G_{\alpha}(x_{1},\ldots,x_{2N})}\left(1+O\left(\frac{|\boldsymbol{X}_{s}|}{\sqrt{t}}\right)\right).

Combining this with

dα(𝒙)[|s{T>s}]d(𝒙)[|s]=Gα(x1,,x2N)Gα(Xs1,,Xs2N),\frac{\mathrm{d}\mathbb{P}_{\alpha}^{(\boldsymbol{x})}\left[\cdot|_{\mathcal{F}_{s}\cap\{T>s\}}\right]}{\mathrm{d}\mathbb{P}_{*}^{(\boldsymbol{x})}\left[\cdot|_{\mathcal{F}_{s}}\right]}=\frac{G_{\alpha}(x_{1},\ldots,x_{2N})}{G_{\alpha}(X^{1}_{s},\ldots,X^{2N}_{s})},

we conclude that

dα|t(𝒙)[|s]d(𝒙)[|s]=(1st)A2N+/2(1+O(|𝑿s|t)),\frac{\mathrm{d}\mathbb{P}_{\alpha\,|\,t}^{(\boldsymbol{x})}\left[\cdot|_{\mathcal{F}_{s}}\right]}{\mathrm{d}\mathbb{P}_{*}^{(\boldsymbol{x})}\left[\cdot|_{\mathcal{F}_{s}}\right]}=\left(1-\frac{s}{t}\right)^{-A_{2N}^{+}/2}\left(1+O\left(\frac{|\boldsymbol{X}_{s}|}{\sqrt{t}}\right)\right),

which implies that

limt𝔼(𝒙)[(1dα|t(𝒙)[|s]d(𝒙)[|s])+]=0.\lim_{t\to\infty}\mathbb{E}_{*}^{(\boldsymbol{x})}\left[\left(1-\frac{\mathrm{d}\mathbb{P}_{\alpha\,|\,t}^{(\boldsymbol{x})}\left[\cdot|_{\mathcal{F}_{s}}\right]}{\mathrm{d}\mathbb{P}_{*}^{(\boldsymbol{x})}\left[\cdot|_{\mathcal{F}_{s}}\right]}\right)^{+}\right]=0.

This gives the convergence in total-variation distance and completes the proof. ∎

Proof of Proposition 1.3.

The estimate in (1.5) is already proved. The convergence (1.7) follows from Corollary 3.5 because μα|t(𝒙)\mu_{\alpha\,|\,t}^{(\boldsymbol{x})} is the induced measure on the driving function (𝑿t,t0)(\boldsymbol{X}_{t},t\geq 0) under α|t(𝒙)\mathbb{P}_{\alpha\,|\,t}^{(\boldsymbol{x})} and μ(𝒙)\mu_{*}^{(\boldsymbol{x})} is the induced measure on the driving function under (𝒙)\mathbb{P}_{*}^{(\boldsymbol{x})}. ∎

4 Gaussian free field level lines and Dyson Brownian motion β=2\beta=2

We will give a very brief summary on Gaussian free field (GFF) and its level lines, more details can be found in [She07, SS13, WW17]. Suppose DD\subset\mathbb{C} is a domain. The Sobolev space H01(D)H_{0}^{1}(D) is the Hilbert space closure of C0(D)C_{0}^{\infty}(D) with respect to the Dirichlet inner product (f,g)=12πfgdx(f,g)_{\nabla}=\frac{1}{2\pi}\int\nabla f\cdot\nabla g\;\mathrm{d}x. The zero-boundary GFF Γ\Gamma on DD is given by Γ=nXnfn\Gamma=\sum_{n}X_{n}f_{n} where {Xn}\{X_{n}\} is a sequence of i.i.d. normal Gaussian random variables and {fn}\{f_{n}\} is an orthonormal basis for H01(D)H_{0}^{1}(D). Such sum does not converge in H01(D)H_{0}^{1}(D), but it does converge in an appropriate space of distributions. The GFF with boundary data Γ0\Gamma_{0} is the sum of the zero-boundary GFF on DD and the function in DD, still denoted by Γ0\Gamma_{0}, which is harmonic in DD and is equal to Γ0\Gamma_{0} on D\partial D. It is shown in [SS13] that variants of SLE4\mathrm{SLE}_{4} process can be coupled with GFF as “level lines”. In this section, we focus on the coupling between chordal NN-SLE4\mathrm{SLE}_{4} and GFF.

We fix λ=π/2\lambda=\pi/2 and fix 𝒙=(x1,,x2N)𝔛2N\boldsymbol{x}=(x_{1},\ldots,x_{2N})\in\mathfrak{X}_{2N}. We consider GFF Γ\Gamma in \mathbb{H} with the following boundary data:

λ on (x2j1,x2j) for 1jN,λ on (x2j,x2j+1) for 0jN,\lambda\text{ on }(x_{2j-1},x_{2j})\text{ for }1\leq j\leq N,\quad-\lambda\text{ on }(x_{2j},x_{2j+1})\text{ for }0\leq j\leq N, (4.1)

with the convention that x0=x_{0}=-\infty and x2N+1=x_{2N+1}=\infty. For 1jN1\leq j\leq N, let η2j1\eta_{2j-1} be the level line of Γ\Gamma starting from x2j1x_{2j-1} and let η2j\eta_{2j} be the level line of Γ-\Gamma starting from x2jx_{2j}. We denote the law of 𝜼=(η1,,η2N)\boldsymbol{\eta}=(\eta_{1},\ldots,\eta_{2N}) by GFF(𝒙)\mathbb{P}_{\mathrm{GFF}}^{(\boldsymbol{x})}. We view it as a chordal Loewner chain with 1/21/2-common parameterization and with driving functions (𝑿t=(Xt1,,Xt2N),t0)(\boldsymbol{X}_{t}=(X_{t}^{1},\ldots,X_{t}^{2N}),t\geq 0). Denote by (t,t0)(\mathcal{F}_{t},t\geq 0) the filtration generated by (𝜼(t),t0)(\boldsymbol{\eta}(t),t\geq 0), and the lifetime TT as the first time tt that Xtj=XtkX_{t}^{j}=X_{t}^{k} for some jkj\neq k. We will prove in Lemma 4.2 that, under GFF(𝒙)\mathbb{P}_{\mathrm{GFF}}^{(\boldsymbol{x})}, the driving functions (𝑿t,0t<T)(\boldsymbol{X}_{t},0\leq t<T) satisfies the SDE

dXtj=dBtj+kj12((1)jk+1)XtjXtkdt,1j2N,\mathrm{d}X_{t}^{j}=\mathrm{d}B_{t}^{j}+\sum_{k\neq j}\frac{\frac{1}{2}\left((-1)^{j-k}+1\right)}{X_{t}^{j}-X_{t}^{k}}\mathrm{d}t,\quad 1\leq j\leq 2N, (4.2)

where {Bj}1j2N\{B^{j}\}_{1\leq j\leq 2N} are independent standard Brownian motions. The main conclusion of this section is the following.

Proposition 4.1.

Consider level lines 𝛈=(η1,,η2N)\boldsymbol{\eta}=(\eta_{1},\ldots,\eta_{2N}) of the GFF with boundary data (4.1).

  • Denote by q(t;,)q(t;\cdot,\cdot) the transition density for the solution to (4.2).

  • Denote by q(t;,)q_{*}(t;\cdot,\cdot) the transition density for Dyson Brownian motion with parameter β=2\beta=2:

    dXtj=dBtj+kj1XtjXtkdt,1j2N,\mathrm{d}X_{t}^{j}=\mathrm{d}B_{t}^{j}+\sum_{k\neq j}\frac{1}{X_{t}^{j}-X_{t}^{k}}\mathrm{d}t,\quad 1\leq j\leq 2N,

    where {Bj}1j2N\{B^{j}\}_{1\leq j\leq 2N} are independent standard Brownian motions.

Then we have

q(t;𝒙,𝒚)=q(t;𝒙,𝒚)G(𝒙)G(𝒚),for all t0, and 𝒙,𝒚𝔛2N,q(t;\boldsymbol{x},\boldsymbol{y})=q_{*}(t;\boldsymbol{x},\boldsymbol{y})\frac{G(\boldsymbol{x})}{G(\boldsymbol{y})},\quad\text{for all }t\geq 0,\text{ and }\boldsymbol{x},\boldsymbol{y}\in\mathfrak{X}_{2N}, (4.3)

where

G(x1,,x2N)=1j<k2N(xkxj)12(1(1)jk).G(x_{1},\ldots,x_{2N})=\prod_{1\leq j<k\leq 2N}(x_{k}-x_{j})^{\frac{1}{2}(1-(-1)^{j-k})}. (4.4)

Moreover,

GFF(𝒙)[T>t]=N1𝒥NG(𝒙)tN2/2(1+O(|𝒙|t)),as t,\mathbb{P}_{\mathrm{GFF}}^{(\boldsymbol{x})}[T>t]=\mathcal{I}_{N}^{-1}\mathcal{J}_{N}G(\boldsymbol{x})t^{-N^{2}/2}\left(1+O\left(\frac{|\boldsymbol{x}|}{\sqrt{t}}\right)\right),\quad\text{as }t\to\infty, (4.5)

where N\mathcal{I}_{N} and 𝒥N\mathcal{J}_{N} are constants:

N=(2π)Nn=12N1n!,𝒥N=𝔛2N1j<k2N(xkxj)12(3+(1)jk)×e12|𝒙|2d𝒙.\mathcal{I}_{N}=(2\pi)^{N}\prod_{n=1}^{2N-1}n!,\qquad\mathcal{J}_{N}=\int_{\mathfrak{X}_{2N}}\prod_{1\leq j<k\leq 2N}(x_{k}-x_{j})^{\frac{1}{2}(3+(-1)^{j-k})}\times e^{-\frac{1}{2}|\boldsymbol{x}|^{2}}\mathrm{d}\boldsymbol{x}.

The law of the solution to (4.2) conditional on {T>t}\{T>t\} converges to the law of Dyson Brownian motion with parameter β=2\beta=2 in total variation distance as in (1.7).

The rest of this section is devoted to proving Proposition 4.1. We denote the law of 𝜼=(η1,,η2N)\boldsymbol{\eta}=(\eta_{1},\ldots,\eta_{2N}) by GFF(𝒙)\mathbb{P}_{\mathrm{GFF}}^{(\boldsymbol{x})}. For j{1,,2N}j\in\{1,\ldots,2N\}, if we parameterize ηj\eta_{j} with 1/21/2-usual parameterization, then its driving function (Wtj,t0)(W_{t}^{j},t\geq 0) satisfies (see e.g. [MS16, Theorem 1.1])

{dWtj=dBtj+(jlog𝒵GFF(N))(Vt1,,Vtj1,Wtj,Vtj+1,,Vt2N)dt,W0j=xj;dVtk=1/2VtkWtjdt,V0k=xk,k{1,,j1,j+1,,2N};\begin{cases}\mathrm{d}W_{t}^{j}=\mathrm{d}B_{t}^{j}+(\partial_{j}\log\mathcal{Z}_{\mathrm{GFF}}^{(N)})(V^{1}_{t},\ldots,V^{j-1}_{t},W_{t}^{j},V^{j+1}_{t},\ldots,V^{2N}_{t})\mathrm{d}t,\quad W_{0}^{j}=x_{j};\\ \mathrm{d}V_{t}^{k}=\frac{1/2}{V_{t}^{k}-W_{t}^{j}}\mathrm{d}t,\quad V_{0}^{k}=x_{k},\quad k\in\{1,\ldots,j-1,j+1,\ldots,2N\};\end{cases}

where (Btj,t0)(B^{j}_{t},t\geq 0) is a standard Brownian motion and 𝒵GFF(N)\mathcal{Z}_{\mathrm{GFF}}^{(N)} is given by

𝒵GFF(N)(x1,,x2N)=1j<k2N(xkxj)12(1)jk.\mathcal{Z}_{\mathrm{GFF}}^{(N)}(x_{1},\ldots,x_{2N})=\prod_{1\leq j<k\leq 2N}(x_{k}-x_{j})^{\frac{1}{2}(-1)^{j-k}}. (4.6)

If we parameterize 𝜼=(η1,,η2N)\boldsymbol{\eta}=(\eta_{1},\ldots,\eta_{2N}) with 1/21/2-common parameterization and use the same notations as in Section 3.1 for κ=4\kappa=4, we will obtain the following lemma.

Lemma 4.2.

Recall that \mathbb{P} denotes the law on 𝛈=(η1,,η2N)\boldsymbol{\eta}=(\eta_{1},\ldots,\eta_{2N}) where (η1,,η2N)(\eta_{1},\ldots,\eta_{2N}) are independent chordal SLE4\mathrm{SLE}_{4} in \mathbb{H} started from (x1,,x2N)(x_{1},\ldots,x_{2N}) respectively. The Radon-Nikodym derivative between GFF(𝐱)\mathbb{P}_{\mathrm{GFF}}^{(\boldsymbol{x})} and \mathbb{P} when both measures are restricted to t\mathcal{F}_{t} and {T>t}\{T>t\} is given by

dGFF(𝒙)[|t{T>t}]d[|t{T>t}]=MtM0,t>0,\frac{\mathrm{d}\mathbb{P}_{\mathrm{GFF}}^{(\boldsymbol{x})}[\cdot|_{\mathcal{F}_{t}\cap\{T>t\}}]}{\mathrm{d}\mathbb{P}[\cdot|_{\mathcal{F}_{t}\cap\{T>t\}}]}=\frac{M_{t}}{M_{0}},\quad t>0,

where

Mt=j=12Ngt,j(Wtj)1/4exp(1240tSgs,j(Wsj)ds)×𝒵GFF(N)(Xt1,,Xt2N),M_{t}=\prod_{j=1}^{2N}g_{t,j}^{\prime}(W_{t}^{j})^{1/4}\exp\left(-\frac{1}{24}\int_{0}^{t}Sg_{s,j}(W_{s}^{j})\mathrm{d}s\right)\times\mathcal{Z}_{\mathrm{GFF}}^{(N)}(X^{1}_{t},\ldots,X^{2N}_{t}),

and 𝒵GFF(N)\mathcal{Z}_{\mathrm{GFF}}^{(N)} is the partition function (4.6). Moreover, under GFF(𝐱)\mathbb{P}_{\mathrm{GFF}}^{(\boldsymbol{x})}, the driving functions (𝐗t,0t<T)(\boldsymbol{X}_{t},0\leq t<T) satisfies the SDE (4.2) where {Bj}1j2N\{B^{j}\}_{1\leq j\leq 2N} are independent standard Brownian motions.

Proof.

We fix κ=4\kappa=4. From [PW19, Theorem 1.4], we know that

𝒵GFF(N)(𝒙)=αLPN𝒵α(𝒙),\mathcal{Z}_{\mathrm{GFF}}^{(N)}(\boldsymbol{x})=\sum_{\alpha\in\mathrm{LP}_{N}}\mathcal{Z}_{\alpha}(\boldsymbol{x}), (4.7)

and that

GFF(𝒙)=α𝒵α(𝒙)𝒵GFF(N)(𝒙)α(𝒙).\mathbb{P}_{\mathrm{GFF}}^{(\boldsymbol{x})}=\sum_{\alpha}\frac{\mathcal{Z}_{\alpha}(\boldsymbol{x})}{\mathcal{Z}_{\mathrm{GFF}}^{(N)}(\boldsymbol{x})}\mathbb{P}_{\alpha}^{(\boldsymbol{x})}. (4.8)

Therefore,

dGFF(𝒙)[|t{T>t}]d[|t{T>t}]=\displaystyle\frac{\mathrm{d}\mathbb{P}_{\mathrm{GFF}}^{(\boldsymbol{x})}[\cdot\,|\,_{\mathcal{F}_{t}\cap\{T>t\}}]}{\mathrm{d}\mathbb{P}[\cdot\,|\,_{\mathcal{F}_{t}\cap\{T>t\}}]}= αLPN𝒵α(𝒙)𝒵GFF(N)(𝒙)×dα(𝒙)[|t{T>t}]d[|t{T>t}]\displaystyle\sum_{\alpha\in\mathrm{LP}_{N}}\frac{\mathcal{Z}_{\alpha}(\boldsymbol{x})}{\mathcal{Z}_{\mathrm{GFF}}^{(N)}(\boldsymbol{x})}\times\frac{\mathrm{d}\mathbb{P}_{\alpha}^{(\boldsymbol{x})}[\cdot\,|\,_{\mathcal{F}_{t}\cap\{T>t\}}]}{\mathrm{d}\mathbb{P}[\cdot\,|\,_{\mathcal{F}_{t}\cap\{T>t\}}]} (due to (4.8))
=\displaystyle= αLPN𝒵α(𝒙)𝒵GFF(N)(𝒙)×MtαM0α\displaystyle\sum_{\alpha\in\mathrm{LP}_{N}}\frac{\mathcal{Z}_{\alpha}(\boldsymbol{x})}{\mathcal{Z}_{\mathrm{GFF}}^{(N)}(\boldsymbol{x})}\times\frac{M_{t}^{\alpha}}{M_{0}^{\alpha}} (due to Lemma 3.2)
=\displaystyle= MtM0.\displaystyle\frac{M_{t}}{M_{0}}. (due to (4.7))

Note that 𝒵GFF(N)\mathcal{Z}_{\mathrm{GFF}}^{(N)} is a combination of 𝒵α\mathcal{Z}_{\alpha} as in (4.7) and hence it also satisfies the PDE (2.2), and that under α(𝒙)\mathbb{P}_{\alpha}^{(\boldsymbol{x})}, the driving functions (𝑿t,0t<T)(\boldsymbol{X}_{t},0\leq t<T) satisfies the SDE (1.1) with κ=4\kappa=4. Then similar calculation using Itô’s formula implies that under GFF(𝒙)\mathbb{P}_{\mathrm{GFF}}^{(\boldsymbol{x})}, (𝑿t,0t<T)(\boldsymbol{X}_{t},0\leq t<T) satisfies the SDE

dXtj=dBtj+(jlog𝒵GFF(N))(Xt1,,Xt2N)dt+kj1/2XtjXtkdt,1j2N,\mathrm{d}X_{t}^{j}=\mathrm{d}B_{t}^{j}+(\partial_{j}\log\mathcal{Z}_{\mathrm{GFF}}^{(N)})(X_{t}^{1},\ldots,X_{t}^{2N})\mathrm{d}t+\sum_{k\neq j}\frac{1/2}{X_{t}^{j}-X_{t}^{k}}\mathrm{d}t,\quad 1\leq j\leq 2N,

where {Bj}1j2N\{B^{j}\}_{1\leq j\leq 2N} are independent standard Brownian motions. From the explicit expression (4.6), this is the same as SDE (4.2) and we complete the proof. ∎

Proof of Proposition 4.1.

We fix κ=4\kappa=4. For each t>0t>0, we obtain from Lemma 3.1 with κ=4\kappa=4 and Lemma 4.2, the Radon-Nikodym derivative of GFF(𝒙)\mathbb{P}_{\mathrm{GFF}}^{(\boldsymbol{x})} against (𝒙)\mathbb{P}_{*}^{(\boldsymbol{x})} when both measures are restricted to t\mathcal{F}_{t} and {T>t}\{T>t\}:

dGFF(𝒙)[|t{T>t}]d(𝒙)[|t]=Mt/M0Mt/M0=𝒵GFF(N)(𝑿t)/𝒵GFF(N)(𝒙)𝒵(𝑿t)/𝒵(𝒙)=G(𝒙)G(𝑿t),\frac{\mathrm{d}\mathbb{P}_{\mathrm{GFF}}^{(\boldsymbol{x})}[\cdot|_{\mathcal{F}_{t}\cap\{T>t\}}]}{\mathrm{d}\mathbb{P}_{*}^{(\boldsymbol{x})}[\cdot|_{\mathcal{F}_{t}}]}=\frac{M_{t}/M_{0}}{M^{*}_{t}/M^{*}_{0}}=\frac{\mathcal{Z}_{\mathrm{GFF}}^{(N)}(\boldsymbol{X}_{t})/\mathcal{Z}_{\mathrm{GFF}}^{(N)}(\boldsymbol{x})}{\mathcal{Z}_{*}(\boldsymbol{X}_{t})/\mathcal{Z}_{*}(\boldsymbol{x})}=\frac{G(\boldsymbol{x})}{G(\boldsymbol{X}_{t})},

where

G(𝒙)=𝒵(𝒙)/𝒵GFF(N)(𝒙)=1j<k2N(xkxj)12(1(1)jk).G(\boldsymbol{x})=\mathcal{Z}_{*}(\boldsymbol{x})/\mathcal{Z}_{\mathrm{GFF}}^{(N)}(\boldsymbol{x})=\prod_{1\leq j<k\leq 2N}(x_{k}-x_{j})^{\frac{1}{2}(1-(-1)^{j-k})}.

Since (𝑿t,t0)(\boldsymbol{X}_{t},t\geq 0) is a Markov process with transition density q(t;𝒙,𝒚)q_{*}(t;\boldsymbol{x},\boldsymbol{y}) under (𝒙)\mathbb{P}_{*}^{(\boldsymbol{x})}, the Radon-Nikodym derivative gives the transition density q(t;𝒙,𝒚)q(t;\boldsymbol{x},\boldsymbol{y}) as desired in (4.3).

Lemma 2.6 gives the asymptotic of transition density for Dyson Brownian motion with parameter β=2\beta=2:

q(t;𝒙,𝒚)=N1t2N2f2(𝒚)exp(|𝒚|22t)(1+O(|𝒙|t)),as t,\displaystyle q_{*}(t;\boldsymbol{x},\boldsymbol{y})=\mathcal{I}_{N}^{-1}t^{-2N^{2}}f_{2}(\boldsymbol{y})\exp\left(-\frac{|\boldsymbol{y}|^{2}}{2t}\right)\left(1+O\left(\frac{|\boldsymbol{x}|}{\sqrt{t}}\right)\right),\quad\text{as }t\to\infty,

where (see [Meh04, Eq. (17.6.7)])

N=𝔛2Nf2(𝒙)e12|𝒙|2d𝒙=1(2N)!2N|f2(𝒙)|e12|𝒙|2d𝒙=(2π)Nn=12N1n!.\displaystyle\mathcal{I}_{N}=\int_{\mathfrak{X}_{2N}}f_{2}(\boldsymbol{x})e^{-\frac{1}{2}|\boldsymbol{x}|^{2}}\mathrm{d}\boldsymbol{x}=\frac{1}{(2N)!}\int_{\mathbb{R}^{2N}}|f_{2}(\boldsymbol{x})|e^{-\frac{1}{2}|\boldsymbol{x}|^{2}}\mathrm{d}\boldsymbol{x}=(2\pi)^{N}\prod_{n=1}^{2N-1}n!.

Then we obtain

GFF(𝒙)[T>t]\displaystyle\mathbb{P}_{\mathrm{GFF}}^{(\boldsymbol{x})}[T>t] =𝔛2Nq(t;𝒙,𝒚)d𝒚=𝔛2Nq(t;𝒙,𝒚)G(𝒙)G(𝒚)d𝒚\displaystyle=\int_{\mathfrak{X}_{2N}}q(t;\boldsymbol{x},\boldsymbol{y})\mathrm{d}\boldsymbol{y}=\int_{\mathfrak{X}_{2N}}q_{*}(t;\boldsymbol{x},\boldsymbol{y})\frac{G(\boldsymbol{x})}{G(\boldsymbol{y})}\mathrm{d}\boldsymbol{y}
=N1G(𝒙)t2N2𝔛2Nf2(𝒚)G(𝒚)exp(|𝒚|22t)d𝒚×(1+O(|𝒙|t))\displaystyle=\mathcal{I}_{N}^{-1}G(\boldsymbol{x})t^{-2N^{2}}\int_{\mathfrak{X}_{2N}}\frac{f_{2}(\boldsymbol{y})}{G(\boldsymbol{y})}\exp\left(-\frac{|\boldsymbol{y}|^{2}}{2t}\right)\mathrm{d}\boldsymbol{y}\times\left(1+O\left(\frac{|\boldsymbol{x}|}{\sqrt{t}}\right)\right)
=N1G(𝒙)t2N2×𝒥Nt3N2/2(1+O(|𝒙|t))\displaystyle=\mathcal{I}_{N}^{-1}G(\boldsymbol{x})t^{-2N^{2}}\times\mathcal{J}_{N}t^{3N^{2}/2}\left(1+O\left(\frac{|\boldsymbol{x}|}{\sqrt{t}}\right)\right)
=N1𝒥NG(𝒙)tN2/2(1+O(|𝒙|t)),as t,\displaystyle=\mathcal{I}_{N}^{-1}\mathcal{J}_{N}G(\boldsymbol{x})t^{-N^{2}/2}\left(1+O\left(\frac{|\boldsymbol{x}|}{\sqrt{t}}\right)\right),\quad\text{as }t\to\infty,

where

𝒥N=𝔛2Nf2(𝒙)G(𝒙)e12|𝒙|2d𝒙=𝔛2N1j<k2N(xkxj)12(3+(1)jk)e12|𝒙|2d𝒙.\displaystyle\mathcal{J}_{N}=\int_{\mathfrak{X}_{2N}}\frac{f_{2}(\boldsymbol{x})}{G(\boldsymbol{x})}e^{-\frac{1}{2}|\boldsymbol{x}|^{2}}\mathrm{d}\boldsymbol{x}=\int_{\mathfrak{X}_{2N}}\prod_{1\leq j<k\leq 2N}(x_{k}-x_{j})^{\frac{1}{2}(3+(-1)^{j-k})}e^{-\frac{1}{2}|\boldsymbol{x}|^{2}}\mathrm{d}\boldsymbol{x}.

This completes the proof of (4.5).

The same analysis in the proof of Corollary 3.5 implies the convergence as tt\to\infty of GFF(𝒙)\mathbb{P}_{\mathrm{GFF}}^{(\boldsymbol{x})} conditional on {T>t}\{T>t\} to (𝒙)\mathbb{P}_{*}^{(\boldsymbol{x})} for κ=4\kappa=4. As their induced measures on driving function, the law of the solution to (4.2) conditional on {T>t}\{T>t\} converges to the law of Dyson Brownian motion with parameter β=2\beta=2. ∎

5 Proof of Theorem 1.4

Fix κ(0,8)\kappa\in(0,8), 𝒙=(x1,,x2N)𝔛2N\boldsymbol{x}=(x_{1},\ldots,x_{2N})\in\mathfrak{X}_{2N}, αLPN\alpha\in\mathrm{LP}_{N} and suppose (γ1,,γN)α(𝒙)(\gamma_{1},\ldots,\gamma_{N})\sim\mathbb{P}_{\alpha}^{(\boldsymbol{x})} is a chordal NN-SLEκ\mathrm{SLE}_{\kappa} in (;x1,,x2N)(\mathbb{H};x_{1},\ldots,x_{2N}) associated to link pattern α\alpha.

Proof of Theorem 1.4.

First, we regard (γ1,,γN)(\gamma_{1},\ldots,\gamma_{N}) as a 2N2N-tuple of curves (η1,,η2N)(\eta_{1},\ldots,\eta_{2N}) as described in Section 1.1 and derive the relation between different time parameterizations of the curves. Suppose that for each 1j2N1\leq j\leq 2N, ηj\eta_{j} is parameterized with aa-usual parameterization. Then we get a 2N2N-time parameter 𝒕=(t1,,t2N)+2N\boldsymbol{t}=(t_{1},\ldots,t_{2N})\in\mathbb{R}_{+}^{2N}. Recall the normalized conformal maps gtjjg_{t_{j}}^{j}, g𝒕g_{\boldsymbol{t}}, g𝒕,jg_{\boldsymbol{t},j} in Section 3.2. A standard calculation gives that

d(𝒕)=aj=12Ng𝒕,j(Wtjj)2dtj.\mathrm{d}\aleph(\boldsymbol{t})=a\sum_{j=1}^{2N}g_{\boldsymbol{t},j}^{\prime}(W_{t_{j}}^{j})^{2}\mathrm{d}t_{j}.

For each 1j2N1\leq j\leq 2N, we have g𝒕,j(Wtjj)(0,1]g_{\boldsymbol{t},j}^{\prime}(W_{t_{j}}^{j})\in(0,1], and for 𝒕=(0,,0,tj,0,,0)+2N\boldsymbol{t}=(0,\ldots,0,t_{j},0,\ldots,0)\in\mathbb{R}_{+}^{2N}, we have (𝒕)=atj\aleph(\boldsymbol{t})=at_{j}. Therefore, for 𝒕=(t1,,t2N)+2N\boldsymbol{t}=(t_{1},\ldots,t_{2N})\in\mathbb{R}_{+}^{2N},

amax1j2Ntj(𝒕)a(t1++t2N).a\max_{1\leq j\leq 2N}t_{j}\leq\aleph(\boldsymbol{t})\leq a(t_{1}+\cdots+t_{2N}).

There exists a time change 𝒕=𝒕(t)=(t1(t),,t2N(t))\boldsymbol{t}=\boldsymbol{t}(t)=(t_{1}(t),\ldots,t_{2N}(t)) with tj(t)=g𝒕,j(Wtjj)2t_{j}^{\prime}(t)=g_{\boldsymbol{t},j}^{\prime}(W_{t_{j}}^{j})^{-2} for 1j2N1\leq j\leq 2N, such that (𝜼(t)=(η1(t1(t)),,η2N(t2N(t))),0t<T)(\boldsymbol{\eta}(t)=(\eta_{1}(t_{1}(t)),\ldots,\eta_{2N}(t_{2N}(t))),0\leq t<T) is parameterized with aa-common parameterization. In this case, we have (t)=2aNt\aleph(t)=2aNt, which gives the relation

max1j2Ntj2Ntt1++t2N,\max_{1\leq j\leq 2N}t_{j}\leq 2Nt\leq t_{1}+\cdots+t_{2N},

and from the time change, we have tj(t)tt_{j}(t)\geq t for 1j2N1\leq j\leq 2N. We may conclude the relation as

ttj(t)2Nt,for all 1j2N.t\leq t_{j}(t)\leq 2Nt,\quad\text{for all }1\leq j\leq 2N. (5.1)

Second, we compare the events {γjB(0,R),1jN}\{\gamma_{j}\cap\partial B(0,R)\neq\emptyset,1\leq j\leq N\} and {T>t}\{T>t\} for some proper t>0t>0. Precisely, we will prove the relation

{T>R2a}{γjB(0,R),1jN}{T>R28aN}.\left\{T>\frac{R^{2}}{a}\right\}\subset\{\gamma_{j}\cap\partial B(0,R)\neq\emptyset,1\leq j\leq N\}\subset\left\{T>\frac{R^{2}}{8aN}\right\}. (5.2)

Suppose that for each 1j2N1\leq j\leq 2N, ηj\eta_{j} is parameterized with aa-usual parameterization, and we define τj=min(inf{tj:ηj(0,tj]B(0,R)},tj(T))\tau_{j}=\min(\inf\{t_{j}:\eta_{j}(0,t_{j}]\cap\partial B(0,R)\neq\emptyset\},t_{j}(T)).

  • On the event {T>R2/a}\{T>R^{2}/a\}, we know from (5.1) that tj(T)T>R2/at_{j}(T)\geq T>R^{2}/a for any 1j2N1\leq j\leq 2N. Then we have hcap(η(0,tj])>R2\text{hcap}(\eta(0,t_{j}])>R^{2}, and by Lemma 2.1, ηj(0,tj]B(0,R)\eta_{j}(0,t_{j}]\cap\partial B(0,R)\neq\emptyset for any 1j2N1\leq j\leq 2N. This implies {γjB(0,R),1jN}\{\gamma_{j}\cap\partial B(0,R)\neq\emptyset,1\leq j\leq N\}.

  • On the event {γjB(0,R),1jN}\{\gamma_{j}\cap\partial B(0,R)\neq\emptyset,1\leq j\leq N\}, for each 1jN1\leq j\leq N, we have τajτbj<\tau_{a_{j}}\wedge\tau_{b_{j}}<\infty, and by Lemma 2.1, τajτbj>R2/4a\tau_{a_{j}}\wedge\tau_{b_{j}}>R^{2}/4a. Note that τjtj(T)\tau_{j}\leq t_{j}(T) implies the non-collision of driving functions for {ηj(0,τj),1j2N}\{\eta_{j}(0,\tau_{j}),1\leq j\leq 2N\}, then by (5.1), the lifetime TT of (η1,,η2N)(\eta_{1},\ldots,\eta_{2N}) has a lower bound

    T>12Nmax1j2Nτj>R28aN.T>\frac{1}{2N}\max_{1\leq j\leq 2N}\tau_{j}>\frac{R^{2}}{8aN}.

This completes the proof of (5.2).

Finally, we prove (1.8):

α(𝒙)[γjB(0,R),1jN]\displaystyle\mathbb{P}_{\alpha}^{(\boldsymbol{x})}[\gamma_{j}\cap\partial B(0,R)\neq\emptyset,1\leq j\leq N]\leq α(𝒙)[T>R2/(8aN)]\displaystyle\mathbb{P}_{\alpha}^{(\boldsymbol{x})}[T>R^{2}/(8aN)] (due to (5.2))
=\displaystyle= 1𝒥αGα(𝒙)(8aNR)A2N+(1+O(|𝒙|R))\displaystyle\mathcal{I}_{*}^{-1}\mathcal{J}_{\alpha}G_{\alpha}(\boldsymbol{x})\left(\frac{\sqrt{8aN}}{R}\right)^{A_{2N}^{+}}\left(1+O\left(\frac{|\boldsymbol{x}|}{R}\right)\right) (due to (1.5))
\displaystyle\asymp Gα(𝒙)RA2N+;\displaystyle G_{\alpha}(\boldsymbol{x})R^{-A_{2N}^{+}};
α(𝒙)[γjB(0,R),1jN]\displaystyle\mathbb{P}_{\alpha}^{(\boldsymbol{x})}[\gamma_{j}\cap\partial B(0,R)\neq\emptyset,1\leq j\leq N]\geq α(𝒙)[T>R2/a]\displaystyle\mathbb{P}_{\alpha}^{(\boldsymbol{x})}[T>R^{2}/a] (due to (5.2))
=\displaystyle= 1𝒥αGα(𝒙)(aR)A2N+(1+O(|𝒙|R))\displaystyle\mathcal{I}_{*}^{-1}\mathcal{J}_{\alpha}G_{\alpha}(\boldsymbol{x})\left(\frac{\sqrt{a}}{R}\right)^{A_{2N}^{+}}\left(1+O\left(\frac{|\boldsymbol{x}|}{R}\right)\right) (due to (1.5))
\displaystyle\asymp Gα(𝒙)RA2N+;\displaystyle G_{\alpha}(\boldsymbol{x})R^{-A_{2N}^{+}};

where the implicit constants depend only on κ,N,α\kappa,N,\alpha. This completes the proof. ∎

We expect that a stronger conclusion holds:444For two functions F1(𝒙;R)F_{1}(\boldsymbol{x};R) and F2(𝒙;R)F_{2}(\boldsymbol{x};R), we write F1(𝒙;R)F2(𝒙;R)F_{1}(\boldsymbol{x};R)\sim F_{2}(\boldsymbol{x};R) as RR\to\infty if the limit limRF1(𝒙;R)F2(𝒙;R)\lim_{R\to\infty}\frac{F_{1}(\boldsymbol{x};R)}{F_{2}(\boldsymbol{x};R)} exists and takes value in (0,)(0,\infty).

α(𝒙)[γjB(0,R),1jN]Gα(𝒙)RA2N+,as R,\mathbb{P}_{\alpha}^{(\boldsymbol{x})}[\gamma_{j}\cap\partial B(0,R)\neq\emptyset,1\leq j\leq N]\sim G_{\alpha}(\boldsymbol{x})R^{-A_{2N}^{+}},\quad\text{as }R\to\infty, (5.3)

This stronger estimate (5.3) is proved for κ(0,8),N=1\kappa\in(0,8),N=1 in [Law15, Theorem 1] and is proved for κ(0,8),N=2\kappa\in(0,8),N=2 in [Zha21, Theorem 1.1]. We are not able to prove (5.3) due to technical difficulty.

Finally, let us briefly explain how we extend [FWY24, Theorem 1.2] from κ(0,4]\kappa\in(0,4] to κ(0,8)\kappa\in(0,8) as in Remark 1.5. To prove (1.9), we need the extension to κ(4,8)\kappa\in(4,8) of the following three basic inputs:

  • Radon-Nikodym derivative of the law α(𝕌;𝒙)\mathbb{P}_{\alpha}^{(\mathbb{U};\boldsymbol{x})} of chordal NN-SLEκ\mathrm{SLE}_{\kappa} against the law of radial Loewner chains driven by 2N2N-radial Bessel process with parameter 2a2a.

  • Transition density for 2N2N-radial Bessel process and its convergence to invariant density as tt\to\infty ([HL21, Proposition 5.5]).

  • The quasi-invariant density for the argument process under α(𝕌;𝒙)\mathbb{P}_{\alpha}^{(\mathbb{U};\boldsymbol{x})}.

The first input is an analogue of Lemmas 3.1 and 3.2 in this article. Note that the extension to κ(4,8)\kappa\in(4,8) of SDE (2.11) for driving function due to recent development in [FPW24, FLPW24] can be modified to the radial case, by using Möbius covariance of pure partition functions and technicals of coordinate changes for SLE\mathrm{SLE} (see [SW05]) respectively. The second input of 2N2N-radial Bessel process is an analogue of Lemma 2.6, which is true when parameter 2a1/22a\geq 1/2 (i.e., κ(0,8]\kappa\in(0,8]). The third input is an analogue of Proposition 1.3, and can be now extended to κ(4,8)\kappa\in(4,8) due to Lemma 2.3. Based on the extension of these main ingredients, the proof of [FWY24, Theorem 1.2] can be extended to κ(4,8)\kappa\in(4,8) using the same analysis.

Acknowledgements.

We thank Zhonggen Su who suggested this topic to H. W. and provided helpful discussions. We thank Vadim Gorin for helpful discussion and in particular, for pointing out useful references in Lemma 2.6.

References

  • [AHP24] Osama Abuzaid, Vivian Olsiewski Healey, and Eveliina Peltola. Large deviations of Dyson Brownian motion on the circle and multiradial SLE0+. Preprint in arXiv:2407.13762, 2024.
  • [AHSY23] Morris Ang, Nina Holden, Xin Sun, and Pu Yu. Conformal welding of quantum disks and multiple SLE: the non-simple case. Preprint in arXiv:2310.20583, 2023.
  • [BBK05] Michel Bauer, Denis Bernard, and Kalle Kytölä. Multiple Schramm-Loewner evolutions and statistical mechanics martingales. J. Stat. Phys., 120(5-6):1125–1163, 2005.
  • [BPW21] Vincent Beffara, Eveliina Peltola, and Hao Wu. On the uniqueness of global multiple SLEs. Ann. Probab., 49(1):400–434, 2021.
  • [Car03] John Cardy. Stochastic Loewner evolution and Dyson’s circular ensembles. J. Phys. A, 36(24):L379–L386, 2003.
  • [CLM23] Andrew Campbell, Kyle Luh, and Vlad Margarint. Rate of convergence in multiple SLE using random matrix theory. Preprint in arXiv:2301.04722, 2023.
  • [CM22] Jiaming Chen and Vlad Margarint. Perturbations of multiple Schramm-Loewner evolution with two non-colliding Dyson Brownian motions. Stochastic Process. Appl., 151:553–569, 2022.
  • [dMS16] Andrea del Monaco and Sebastian Schleißinger. Multiple SLE and the complex Burgers equation. Math. Nachr., 289(16):2007–2018, 2016.
  • [Dub07] Julien Dubédat. Commutation relations for Schramm-Loewner evolutions. Comm. Pure Appl. Math., 60(12):1792–1847, 2007.
  • [FK15a] Steven M. Flores and Peter Kleban. A solution space for a system of null-state partial differential equations: Part 2. Comm. Math. Phys., 333(1):435–481, 2015.
  • [FK15b] Steven M. Flores and Peter Kleban. A solution space for a system of null-state partial differential equations: Part 3. Comm. Math. Phys., 333(2):597–667, 2015.
  • [FLPW24] Yu Feng, Mingchang Liu, Eveliina Peltola and Hao Wu. Multiple SLEs for κ(0,8)\kappa\in(0,8): Coulomb gas integrals and pure partition functions. Preprint in arXiv:2406.06522, 2024.
  • [FPW24] Yu Feng, Eveliina Peltola, and Hao Wu. Crossing probabilities of multiple FK-Ising interfaces. Probability Theory and Related Fields, 189(1):281–367, 2024.
  • [FWY24] Yu Feng, Hao Wu, and Lu Yang. Multiple Ising interfaces in annulus and 2NN-sided radial SLE. Int. Math. Res. Not. IMRN, (6):5326–5372, 2024.
  • [HK18] Ikkei Hotta and Makoto Katori. Hydrodynamic limit of multiple SLE. J. Stat. Phys., 171:166–188, 2018.
  • [HL21] Vivian Olsiewski Healey and Gregory F. Lawler. N-sided radial Schramm-Loewner evolution. Probab. Theory Related Fields, 181(1-3):451–488, 2021.
  • [HS21] Ikkei Hotta and Sebastian Schleißinger. Limits of radial multiple SLE and a Burgers-Loewner differential equation. J. Theoret. Probab., 34(2):755–783, 2021.
  • [KK21] Makoto Katori and Shinji Koshida. Three phases of multiple SLE driven by non-colliding Dyson’s Brownian motions. J. Phys. A, 54(32):Paper No. 325002, 19, 2021.
  • [KL07] Michael J. Kozdron and Gregory F. Lawler. The configurational measure on mutually avoiding SLE\mathrm{SLE} paths. Fields Inst. Commun., 50:199–224, 2007.
  • [KP16] Kalle Kytölä and Eveliina Peltola. Pure partition functions of multiple SLEs. Comm. Math. Phys., 346(1):237–292, 2016.
  • [Law15] Gregory F. Lawler. Minkowski content of the intersection of a Schramm-Loewner evolution (SLE) curve with the real line. J. Math. Soc. Japan, 67(4):1631–1669, 2015.
  • [Meh04] Madan Lal Mehta. Random matrices, volume 142 of Pure and Applied Mathematics (Amsterdam). Elsevier/Academic Press, Amsterdam, third edition, 2004.
  • [MS16] Jason Miller and Scott Sheffield. Imaginary geometry I: Interacting SLEs. Probab. Theory Related Fields, 164(3-4):553–705, 2016.
  • [PW19] Eveliina Peltola and Hao Wu. Global and local multiple SLEs for κ4\kappa\leq 4 and connection probabilities for level lines of GFF. Comm. Math. Phys., 366(2):469–536, 2019.
  • [PW24] Eveliina Peltola and Yilin Wang. Large deviations of multichordal SLE0+, real rational functions, and zeta-regularized determinants of Laplacians. J. Eur. Math. Soc., 26(2):469–535, 2024.
  • [Rös98] Margit Rösler. Generalized Hermite polynomials and the heat equation for Dunkl operators. Comm. Math. Phys., 192(3):519–542, 1998.
  • [She07] Scott Sheffield. Gaussian free fields for mathematicians. Probab. Theory Related Fields, 139(3-4):521–541, 2007.
  • [SS13] Oded Schramm and Scott Sheffield. A contour line of the continuum Gaussian free field. Probab. Theory Related Fields, 157(1-2):47–80, 2013.
  • [SW05] Oded Schramm and David B. Wilson. SLE coordinate changes. New York J. Math., 11:659–669, 2005.
  • [Voi19] Michael Voit. Central limit theorems for multivariate Bessel processes in the freezing regime. J. Approx. Theory, 239:210–231, 2019.
  • [Wu20] Hao Wu. Hypergeometric SLE: conformal Markov characterization and applications. Comm. Math. Phys., 374(2):433–484, 2020.
  • [WW17] Menglu Wang and Hao Wu. Level lines of Gaussian Free Field I: Zero-boundary GFF. Stochastic Process. Appl., 127(4):1045–1124, 2017.
  • [Zha21] Dapeng Zhan. Two-curve Green’s function for 2-SLE: the boundary case. Electron. J. Probab., 26:Paper No. 32, 58, 2021.
  • [Zha23] Dapeng Zhan. Existence and uniqueness of nonsimple multiple SLE. Preprint in arXiv:2308.13886, 2023.