by Michael Cranston
Through his works and words, Kai Lai Chung has been the spur for substantial developments in the understanding of conditional Brownian motion and its application to the theory of Schrödinger operators. Many in the field received mail or phone calls from Chung with interesting and provocative questions on the subject. At the Spring 1982 meeting of the Seminar on Stochastic Processes, he posed an interesting question on the lifetime of conditional Brownian motion. The resolution of this question (described below) has led to wide-ranging developments. His foundational work with Rao on the gauge theorem, to mention just one of his many works in this area, has served as the motivation for many developments in the understanding of Schrödinger operators and their semigroups. And, of course, From Brownian Motion to Schrödinger’s Equation [1] with Zhongxin Zhao has served as guide to developments in the field. In this short, semi-accurate, historical note, I’d like to outline a few developments that trace their origins to the encouragement of Chung. I’d like to apologize in advance for the many works which are not mentioned here, due in large part to an interest in brevity.
First, an introduction is in order. Let \( B \) denote Brownian motion on \( \mathbb{R}^d \) defined on the probability space \( (\Omega, \mathcal F_t, \{P_x\}_{x\in \mathbb{R}^d}) \). For \( D\subset \mathbb{R}^d \), let \( p(t,x,y) \) be its transition density when killed at time \[ \tau^{}_D=\inf\{t > 0:B_t\notin D\} \] (the heat kernel on \( D \).) Given a positive super-harmonic function \( h \) on \( D \), define \[ p^h(t,x,y)=\frac{p(t,x,y)\,h(y)}{h(x)} .\] This is the transition density for a new diffusion, called the \( h \)-process or conditional Brownian motion. We denote by \( P^h_x \) the measure on path-space corresponding to Brownian motion started at \( x \) and with transition density \( p^h(t,x,y) \). In the case when \( h \) is the Martin kernel with pole at the Martin boundary point \( \xi \), then the conditional Brownian motion exits the domain at the boundary point \( \xi \), in the sense that \( B_t \) converges \( P_x^h \)-a.s. to \( \xi \) in the Martin topology as \( t \) approaches the path lifetime, \( \tau^{}_D \). If \( h(\,\cdot\,)=G_D(\,\cdot\,,y) \), where \( G_D \) is the Green function for \( D \) and \( y\in D \), then the \( h \)-process will converge to \( y \) as \( t \) approaches the path lifetime. When \( h \) is the Martin kernel with pole at \( \xi \), we denote the resulting measure by \( P^\xi_x \) and, when \( h(\,\cdot\,)=G(\,\cdot\,,y) \), by \( P_x^y \). These were developments due to Doob [e1] in his study of probabilistic versions of the Fatou boundary-limit results for harmonic functions. Now it’s been known for some time that, if \( D \) is bounded, then, for unconditioned Brownian motion, \begin{equation}\label{ceqn1} E_x[\tau^{}_D] < c_d \operatorname{vol}(D)^{2/d}, \end{equation} and, if \( \lambda_1 \) is the first Dirichlet eigenvalue for \( \frac12\Delta \) on \( D \), then \begin{equation}\label{ceqn2} \lim_{t\rightarrow \infty}\frac{1}{t}\log P_x(\tau^{}_D > t)=-\lambda_1. \end{equation} Using the Martin boundary, denoted here by \( \partial_M D \), the expected lifetime can be expressed as \[ E_x[\tau^{}_D]=\int_{\partial_MD}E_x^\xi[\tau^{}_D]\,\omega_x(d\xi) ,\] where \( \omega_x \) is the exit distribution of Brownian motion on \( \partial_MD \), also known as the harmonic measure. So, by Fubini, \( E_x^\xi[\tau^{}_D] \) is finite \( \omega_x \)-almost surely. Chung’s question is this: \[ \text{When is } E_x^\xi[\tau^{}_D] \text{ bounded uniformly in } x \text{ and } \xi? \] Or, more generally, when is \( E_x^\xi[\tau^{}_D] \) finite? This innocuous-sounding question turned out to have quite broad implications. It led to the introduction of some very interesting ideas from analysis into probability theory, such as the boundary Harnack principle, Whitney chains, Littlewood–Paley \( g \)-function and intrinsic ultracontractivity.
The first result on this question, due to McConnell and the author [e5], was that there is a positive constant \( c \) so that, if \( D\subset \mathbb{R}^2 \) and \( h \) is a positive harmonic function on \( D \), then \begin{equation} \label{ceqn3} E_x^h[\tau^{}_D]\le c\operatorname{vol}(D). \end{equation} This is the analog then of \eqref{ceqn1} in \( d=2 \). An example was given of a bounded \( D\subset \mathbb{R}^3 \) with a \( \xi\in \partial_MD \) for which \( E_x^\xi[\tau^{}_D]=\infty \). Thus, the analog of \eqref{ceqn1} cannot hold for \( d > 2 \) without further assumptions. First, a word or two on the proof of \eqref{ceqn3}. This relies on decomposing the domain \( D \) into subregions by means of the \( 2^m \)-level sets of the function \( h \). That is, \[ D=\bigcup_{m=-\infty}^{\infty} D_m \quad\text{ where }\quad D_m=\{x\in D:2^{m-1} < h(x) < 2^{m+1}\} .\] The conditional Brownian motion viewed at the successive hitting times to \[ C_m=\{x\in D:h(x)=2^m\} \] forms a birth and death Markov chain on \( \{2^m:m\in \mathbb{Z}\} \), with probability \( 2/3 \) of going up and \( 1/3 \) of going down. This implies that the number of visits to the \( C_m \) are geometrically distributed random variables. These have finite expectation with a value independent of \( m \). The other key observation is that the expected amount of time the conditional Brownian motion spends in \( D_m \) starting from \( C_m \) is equivalent (since \( 1/2 \le h(y)/h(x) \le 2 \) for \( x\in C_m \), \( y \in D_m \)) to the amount of time standard Brownian motion spends in \( D_m \) starting from \( C_m \). Combining this observation with \eqref{ceqn1} gives that the expected time spent in \( D_m \) starting on \( C_m \) by the \( h \)-process is bounded by \( C_d \operatorname{vol}(D_m)^{2/d} \). Using the strong Markov property and summing leads to an upper bound of \[E_x^h[\tau^{}_D]\le C_d \sum_{m=-\infty}^{\infty}\operatorname{vol}(D_m)^{2/d}. \] In case \( d=2 \), the sum is bounded by \( 2\operatorname{vol}(D) \), leading to the result that there is a constant \( c \) such that \eqref{ceqn3} holds for \( D\subset \mathbb{R}^2 \), \( x\in D \), and \( h \) any positive superharmonic on \( D \). Since \( 2/d < 1 \) for \( d\ge 3 \), the finiteness of \( \sum_{m=-\infty}^{\infty}\operatorname{vol}(D_m)^{2/d} \) does not generally hold, and leads to interesting questions about the influence of the regularity of the boundary and its effect on the size of the sets \( D_m \). (The relation between boundary regularity and the growth of harmonic functions is a key issue in the subject.) This question was addressed by Bañuelos [e9], Falkner [e10], Bass and Burdzy [e23], DeBlassie [e13], Kenig and Pipher [e17], and myself [e7], among others in the higher-dimensional case. The results of Bañuelos [e9] incorporated many of the types of domains encountered in analysis, namely Lipschitz, NTA (non-tangentially accessible), John- and BMO-extension (uniform) domains. In order to describe the results in [e9], we consider a Whitney decomposition of \( D \). This is a collection of closed squares \( Q_j \) with sides parallel to the coordinate axes and \( D=\bigcup_j Q_j \) with the properties \begin{align*} & Q_j^o\cap Q_k^o=\emptyset \quad\text{if }j\neq k,\\ & \frac{1}{4}\le\frac{l(Q_j)}{l(Q_k)}\le 4 \quad\text{if } Q_j\cap Q_k\neq\emptyset,\\ & 1\le\frac{d(Q_j,\partial D)}{l(Q_j)}\le 4\sqrt d \quad\text{for all }j. \end{align*} A Whitney chain connecting \( Q_j \) and \( Q_k \) is a sequence of Whitney squares \( \{Q_{m_i}\}_{i=0}^n \) with \( Q_{m_0}=Q_j \), \( Q_{m_n}=Q_k \) and \( Q_{m_i}\cap Q_{m_{i+1}}\neq \emptyset \). An important fact about Whitney squares is that there is a positive constant \( c \) so that, for any positive harmonic function \( h \) in \( D \) and adjacent Whitney squares \( Q_j\cap Q_k\neq \emptyset \), we have \( h(x) < c\,h(y) \) for \( x\in Q_j \), \( y\in Q_k \). Whitney chains are very well suited to the study of conditional Brownian motion. The reason is that, due to Harnack’s inequality, any positive harmonic function will be “flat” on the Whitney square \( Q_j \). This means that the transition densities \( p^h(t,x,y) \) and \( p(t,x,y) \) will be equivalent on \( Q_j \), which means the behavior of ordinary and conditional Brownian motion will be comparable on \( Q_j \). Now, define the quasi-hyperbolic distance from \( x\in Q_j \) to \( x_0 \) by first setting \( d(x)=\operatorname{dist}(x,\partial D) \) and then putting, \[\rho^{}_D(x_1,x_2)=\inf_\gamma \int_\gamma \frac{ds}{d(\gamma(s))},\] with the \( \inf \) being taken over all rectifiable curves in \( D \) from \( x_1 \) to \( x_2 \). Taking points \( x_1\in Q_j \), \( x_2\in Q_k \) we have \[\rho^{}_D(x_1,x_2)\,\approx\, \text{length of shortest Whitney chain from }Q_j\text{ to }Q_k .\] Note that repeated applications of Harnack’s inequality in successive squares in a Whitney chain implies that \( h(x_1)\le c^{\rho^{}_D(x_1,x_2)}h(x_2) \). If we fix an \( x_0\in D \) and write \( \rho^{}_D(x)=\rho^{}_D(x,x_0) \), this implies that, for some constant \( C \), \[ D_m\subset \{x\in D:\rho^{}_D(x) > C|m|\}. \] Now, a result of Smith and Stegenga [e19] implies that, for a class \( H(0) \) of domains, called Hölder of order 0 (which includes Lipschitz, NTA, John- and BMO-extension domains), one has \( \rho^{}_D(\,\cdot\,,x_0)\in L^p(D) \) for any \( 0 < p < \infty \). Using this, Bañuelos obtains for \( D \) a bounded Hölder of order 0 domain so that \[ \sum_{m=-\infty}^{\infty}\operatorname{vol}(D_m)^{2/d} < \infty. \] This implies that \( H(0) \) domains are regular enough so that an analog of \eqref{ceqn3} holds for them in all dimensions. There are also beautiful connections in simply connected planar domains between the behavior of conditional Brownian motion and the hyperbolic geometry of the region. This was developed in Bañuelos and Carrol [e25], and Davis [e1]. We start our exposition of this connection with an observation of Bañuelos [e24]. If \( D \) is a simply connected planar domain, and \( \varphi: B(0,1)\rightarrow D \) maps the unit disc \( B(0,1) \) of the complex plane conformally onto \( D \) with \( \varphi(0)=x \), then \[ g^2_*(\varphi)(\theta)=\frac{1}{\pi}\int_{B(0,1)}\log\Bigl(\frac{1}{|z|}\Bigr)\,\frac{1-|z|^2}{|z-e^{i\theta}|^2}\,|\varphi^{\prime}(z)|^2 \,dz \] is the Littlewood–Paley square function. Recalling that the Green function of \( B(0,1) \) with pole at the origin is \( \log(1/|z|) \), and that the Green function is preserved by conformal mappings, it is easy to deduce that, for \( h \) a positive harmonic function on \( D \) with the representation \[ h(z)=\int_0^{2\pi}\frac{1-|z|^2}{|z-e^{i\theta}|}\,d\mu(\theta) \] with \( \mu \) a positive Borel measure on \( \partial B(0,1) \), that \begin{align}\label{ceqn4} E^h_x[\tau^{}_D] &=\frac{1}{h(x)}\int_D G_D(x,y)\,h(y)\,dy \\ &= \frac{1}{h(x)}\int_{B(0,1)}\log\Bigl(\frac{1}{|z|}\Bigr)\,h(\varphi(z))\,|\varphi^{\prime}(z)|^2\,dx\,dy \nonumber \\ &= \frac{1}{h(x)}\int_0^{2\pi}g_*^2(\varphi)(\theta)\,d\mu(\theta). \nonumber \end{align} Since \( \mu([0,2\pi])=h(x) \) and \[ g^2_*(\varphi)(\theta)\le C\int_{B(0,1)}|\varphi^{\prime}(z)|^2\,dz \le C\operatorname{vol}(D), \] it follows that \[ E^h_x[\tau^{}_D]\le C\operatorname{vol}(D), \] thus giving another derivation of the lifetime estimate in the special case of simply connected planar domains. But this gives additional information, as developed in Bañuelos and Carrol [e25]. There, the authors observed that, if \( K(z,\xi) \) is the Poisson kernel for \( B(0,1) \) with pole at \( \xi\in\partial B(0,1) \), then there are positive constants \( c \) and \( C \) such that \[ c\sup_\varphi g^2_*(\varphi)(0) \,\le\, \sup_\varphi \int_{B(0,1)}\!\!\! K(z,1)K(z,-1)\,|\varphi^{\prime}(z)|^2\,dx\,dy \,\le\, C\sup_\varphi g^2_*(\varphi)(0), \] where the \( \sup \) is taken over all conformal mappings \( \varphi: B(0,1)\rightarrow D \) with \( \varphi(0)=x \). But another equivalence holds for \( K(z,1)K(z,{-}1) \). Namely, if \( d(z,\Gamma) \) denotes the hyperbolic distance in \( B(0,1) \) from \( z \) to the geodesic \( \tilde{\Gamma}=[-1,1] \), then \[ \tfrac{1}{4}K(z,1)K(z,-1)\le e^{-2d(z,\tilde{\Gamma})}\le K(z,1)K(z,-1). \] Using the conformal invariance of the hyperbolic metric, writing \( d^{}_D \) for the hyperbolic metric in \( D \), and putting these two equivalences together yields the existence of two positive constants \( c \) and \( C \) such that \[ c\sup_{\Gamma}\int_De^{-2d^{}_D(z,\Gamma)} \le \sup_{x,h}E_x^h[\tau^{}_D] \le C\sup_{\Gamma}\int_De^{-2d^{}_D(z,\Gamma)}. \] This has a beautiful corollary involving the Whitney decomposition mentioned above. Let \( Q \) be a Whitney cube with center \( z^{}_Q \), and let \( T_Q \) be the total amount of time spent in \( Q \) before \( \tau^{}_D \). Then, for Martin boundary points \( \xi_1,\xi_2 \) and for \( \Gamma \) the hyperbolic geodesic connecting them, there are positive constants \( c \) and \( C \) such that \[ \tfrac{1}{4}e^{-Cd^{}_D(z^{}_Q,\Gamma)}\le E_{\xi_1}^{\xi_2}[T_Q]\le e^{-cd^{}_D(z^{}_Q,\Gamma)}. \] This is a quantitative statement about how closely the conditional Brownian motion from \( \xi_1\,\text{to}\,\xi_2 \) follows the hyperbolic geodesic from \( \xi_1\,\text{to}\,\xi_2 \). Davis [e14] pursued this connection further in estimating the variance of \( \tau^{}_D \) under the measure \( E_{\xi_1}^{\xi_2} \). If \( Q \) and \( R \) are Whitney squares, then setting \[ P_Q=P_{\xi_1}^{\xi_2}(\tau^{}_{D\cap Q^c} < \tau^{}_D) \quad\text{and}\quad P_R=P_{\xi_1}^{\xi_2}(\tau^{}_{D\cap R^c} < \tau^{}_D) \] and letting \( \delta(D) \) be the area of the largest disc which can be inscribed in \( D \) yields \begin{align*} & |\operatorname{Cov}_{\xi_1}^{\xi_2}(T_Q,T_R)| \le Ce^{-c\delta^{}_D(z^{}_Q, z^{}_R)} \operatorname{vol}(Q) \operatorname{vol}(R)\,(P_Q+P_R) \quad\text{and} \\ & \operatorname{Var}_{\xi_1}^{\xi_2}(\tau^{}_D) \le \delta(D)\,E_{\xi_1}^{\xi_2}[\tau^{}_D]. \end{align*} The first of these shows exactly how the decay of the dependence between the occupation times \( T_Q \) and \( T_R \) depends on the hyperbolic distance between \( Q \) and \( R \). The second confirms the intuition that the conditional Brownian motion speeds up when traversing narrow channels. (If \( D \) is a rectangle of length \( n \) and width \( 1/n \), then, for \( {\xi_1} \) and \( {\xi_2} \) on opposite ends of the long side of the rectangle, \[ E_{\xi_1}^{\xi_2}[\tau^{}_D]\le c \quad\text{and}\quad \operatorname{Var}_{\xi_1}^{\xi_2}[\tau^{}_D]\le c/n .\] Thus, the conditional motion must go a distance \( n \) in a time with bounded expectation, independent of \( n \), but with variance bounded by \( 1/n \). This means the mass of the measure \( P_{\xi_1}^{\xi_2} \) is concentrating on paths which make the length-\( n \) trip in a time which is some constant that doesn’t depend on \( n \).)
Refinements and further progress in these directions can be found in the works of Griffin, McConnell and Verchota [e27], Griffin, Verchota and Vogel [e26], Zhang [e30], Davis and Zhang [e28], and Xu [e20], to name but a few.
Now, let’s turn our attention to the problem of deciding to what extent the analog of \eqref{ceqn2} holds for conditional Brownian motion. From the case of a ball \( D=\{x:|x| < r\} \) in Euclidean space where \[ P_0(\tau^{}_D > t)=P_0^\xi(\tau^{}_D > t) \] for every boundary point \( \xi \), one might suspect that, with some smoothness in \( d > 2 \) and maybe even with \( \operatorname{vol}(D) < \infty \) in \( d=2 \), if \( H^+(D) \) is the class of positive harmonic functions on \( D \), then \begin{equation} \label{ceqn5} \lim_{t\rightarrow \infty}\frac{1}{t}\log P^h_x(\tau^{}_D > t)=-\lambda_1 \quad\text{for }x\in D\text{ and }h\in H^+(D). \end{equation} This was addressed in De Blassie [e13] where it was proved that \eqref{ceqn5} holds provided \( D \) is a Lipschitz domain with sufficiently small Lipschitz constant. Later, Kenig and Pipher [e17] extended this result to Lipschitz domains and NTA domains. Perhaps the nicest approach is due to Bañuelos [e21] and Bañuelos and Davis [e15], which illuminates the relation between the tail behavior of the lifetime of conditional Brownian motion and intrinsic ultracontractivity. The notion of intrinsic ultracontractivity is defined in Davies and Simon [e8] as the property that the semigroup of the ground-state transformation of an operator maps \( L^2 \) to \( L^\infty \). To make this definition precise in the current setting, if \( \varphi_1 \) is the first Dirichlet eigenfunction for \( \frac{1}{2}\Delta \) on \( D \), define a semigroup on \( L^2(\varphi_1^2\,dx) \) by \[ P^{\varphi_1}_t f(x) =\int_D\frac{e^{\lambda_1 t}p(t,x,y)}{\varphi_1(x)\varphi_1(y)} f(y)\,\varphi_1^2(y)\,dy \quad\text{for }f \in L^2(\varphi_1^2\,dx). \] Then, the domain \( D \) is defined to be intrinsically ultracontractive (IU) if there exist constants \( C_t \) such that \[ |P^{\varphi_1}_t f(x)|\le C_t\|f\|_{L^2(\varphi_1^2dx)} \quad\text{for }t > 0 .\] An important consequence of IU is that for any \( \varepsilon > 0 \) there is a \( t(\varepsilon) \) such that \begin{equation}\label{ceqn6} (1-\varepsilon)\,e^{-\lambda_1t}\varphi_1(x)\varphi_1(y) \le p(t,x,y) \le(1+\varepsilon)\,e^{-\lambda_1t}\varphi_1(x)\varphi_1(y). \end{equation} Since, for any \( h\in H^+(D) \), \[ P_x^h(\tau^{}_D\ge t)=\frac{1}{h(x)}\int_Dp(t,x,y)h(y)\,dy \le 1,\] it follows easily from \eqref{ceqn5} that \[ \lim_{t \rightarrow \infty}\frac{1}{t}\log P_x^h(\tau^{}_D\ge t)=-\lambda_1,\] giving the Bañuelos analog of \eqref{ceqn2} for conditional Brownian motion on IU domains. In the case of planar domains of finite area, Bañuelos and Davis [e15] proved the following analog of IU for each \( x\in D \): \[ \lim_{t\rightarrow \infty}\frac{e^{\lambda_1t}p(t,x,y)}{\varphi_1(x)\varphi_1(y)}=1 \text{ uniformly in }y\in D. \] This implies that the analog of \eqref{ceqn2} for conditional Brownian motion holds for planar domains of finite area.
Another application of conditional Brownian motion, which has been an area of research to which Professor Chung has made many contributions, is to the study of the Schrödinger equation by means of the Feynman–Kac formula. A seminal paper on the subject was that of Aizenman and Simon [e2], who used path-integral techniques (the Feynman–Kac formula) to prove Harnack’s inequality for Schrödinger operators. Consider, with \( d > 2 \) for ease of presentation, a potential \( V \) satisfying \[ \lim_{r \rightarrow 0}\sup_{x \in \mathbb{R}^d}\int_{|x-y| < r}\frac{|V(y)|\ }{|x-y|^{d-2}}\,dy =0. \] The class of such potentials is called the Kato class, and is denoted by \( K_d \). They are particularly well suited to the Newtonian potential, and thus as well to the occupation properties of Brownian motion. Now, for \( f\in C(\partial D) \), consider the Dirichlet problem \begin{align} \label{ceqn7} \tfrac{1}{2}\Delta u(x)+V(x)u(x) &=0 \quad\text{for }x\in D,\\ u(x) &=f(x) \quad\text{for }x\in \partial D. \nonumber \end{align} The Gauge Theorem of Chung and Rao (see the article of Ruth Williams in this volume) says that either \( E_x\bigl[e^{\int_0^{\tau^{}_D}V(B_s)\,ds}\bigr]\equiv \infty \) on \( D \), or this quantity (called the gauge) is bounded on \( D \). Let’s assume that the second alternative of this dichotomy holds. Then, by Feynman–Kac, the solution of \eqref{ceqn7} is given by \begin{eqnarray*} u(x)= E_x\bigl[e^{\int_0^{\tau^{}_D}V(B_s)\,ds}f(B_{\tau^{}_D})\bigr]. \end{eqnarray*}
Let’s suppose now that \( D \) is a Lipschitz domain so that the Euclidean and Martin boundary of \( D \) are the same. Decompose the Feynman–Kac formula using conditional Brownian motion, \begin{equation}\label{ceqn8} u(x)= \int_{\partial D} E_x^y\bigl[e^{\int_0^{\tau^{}_D}V(B_s)\,ds}\bigr] f(y)\,P_x(B_{\tau^{}_D}\in dy). \end{equation} The analog of Chung’s question regarding the finiteness of the expected lifetime of conditional Brownian motion, as well as his question regarding the finiteness of the gauge, is: \[ \text{When is }\,E_x^y\bigl[e^{\int_0^{\tau^{}_D}V(B_s)\,ds}\bigr] < \infty\,\text{?} \] The quantity \[ u(x,y)=E_x^y\bigl[e^{\int_0^{\tau^{}_D}V(B_s)\,ds}\bigr] \] is known as the conditional gauge. Under the conditions set down above, namely that \( D \) be a Lipschitz domain and \( V\in K_d \), a dichotomy (similar to the Gauge Theorem) holds: either \[ E_x^y\bigl[e^{\int_0^{\tau^{}_D}V(B_s)\,ds}\bigr]\equiv \infty\] or there are positive constants \( c \) and \( C \) such that \begin{equation} \label{ceqn9} c\le E_x^y\bigl[e^{\int_0^{\tau^{}_D}V(B_s)\,ds}\bigr]\le C \quad\text{for all }x,y\in D \cup \partial D. \end{equation} This is called the Conditional Gauge Theorem (CGT). It can be viewed as a statement on the mixing properties of conditional Brownian motion. The potential \( V \) may possess singularities. The CGT says that these singularities can’t be so bad that \( P_x^y \)-paths would miss them, in the sense that \[ E_x^y\bigl[e^{\int_0^{\tau^{}_D}V(B_s)\,ds}\bigr] < \infty \] for one pair of points \( x,y \), but \[ E_z^w\bigl[e^{\int_0^{\tau^{}_D}V(B_s)\,ds}\bigr]=\infty \] for another pair \( z,w \). That is, under both measures \( P_x^y \) and \( P_z^w \), the occupation distributions of paths are similar enough that they will simultaneously give a finite answer or an infinite answer when asked about the value of the conditional gauge. This requires some smoothness of \( \partial D \) with its resulting effect on the behavior of the Green function. Early results on the subject were those of Falkner [e4] and Zhao [e3], [e6]. In the fundamental works of Zhao, the CGT was proved for Kato-class potentials on the ball, and then domains with \( C^2 \) boundary. For Lipschitz domains and Kato potentials, the result was proven in Cranston, Fabes and Zhao [e12]. The extension to Lipschitz domains of the CGT used the so-called 3\( G \)-theorem. This result says that, if \( G \) is the Green function for \( \tfrac{1}{2}\Delta \) on \( D \), then there is a positive constant \( C \) such that \begin{equation} \label{ceqn10} \frac{G(x,z)G(z,y)}{G(x,y)}\le C\Bigl(\frac{1}{|x-z|^{d-2}}+\frac{1}{|y-z|^{d-2}}\Bigr). \end{equation} The left-hand side in \eqref{ceqn10} is the Green function for conditional Brownian motion started at \( x \) and conditioned to exit \( D \) at \( y \). This is the occupation density for conditional Brownian motion in \( D \), in the sense that the total expected amount of time spent by \( B \) in \( A\subset D \) with respect to the measure \( P_x^y \) is \[ \int_A \frac{G(x,z)G(z,y)}{G(x,y)}\,dz .\] The right-hand side of \eqref{ceqn10} is the sum of the Newtonian potentials with poles at \( x \) and \( y \), respectively. These are the occupation densities for unconditioned Brownian motion in \( \mathbb{R}^d \) started at \( x \) and \( y \). The 3\( G \)-inequality says that, if \( V \) is in \( K_d \) and thus well adapted to the occupation measure of (unconditioned) Brownian motion, then it is also well adapted to the occupation measure of conditional Brownian motion. In the case when the conditional gauge is finite, the CGT permits comparisons between potential theoretic quantities for the two operators \( -\frac{1}{2}\Delta \) and \( -\frac{1}{2}\Delta+ V \). This lies close to the original motivation of Aizenman and Simon [e2]. For example, suppose that for some \( f\in C(\partial D) \), \begin{align*} \tfrac{1}{2}\Delta v(x) &=0 \quad\text{for } x \in D,\\ v(x) &=f(x) \quad\text{for } x\in \partial D. \end{align*} and \begin{align*} -\tfrac{1}{2}\Delta u(x)+V(x)u(x) &=0 \quad\text{for } x \in D,\\ u(x) &=f(x) \quad\text{for } x\in \partial D. \end{align*} Then, \( v(x)= E_x[f(B_{\tau^{}_D})] \) and, since \[ c\le E_x^y[e^{\int_0^{\tau^{}_D}V(B_s)\,ds}]\le C ,\] it follows from \eqref{ceqn4} that \( cv(x)\le u(x)\le Cv(x),\,x \in D \). With this equivalence, Harnack’s inequality, and even the boundary Harnack inequality, can be deduced for positive solutions in \( D \) of \[ -\tfrac{1}{2}\Delta u(x)+V(x)u(x)=0 .\] Many other similar conclusions follow in an equally easy manner. Using the simple formula \begin{equation} \label{ceqn11} u(x,y)=\frac{G_V(x,y)}{G(x,y)}, \end{equation} it follows that \[cG(x,y)\le G_V(x,y) \le CG(x,y) \quad\text{for }x,y\in D,\] where \( G_V \) is the Green function for \( -\frac{1}{2}\Delta+ V \). Since the Martin kernels \( K(x,\xi) \) and \( K_V(x,\xi) \) are the limit of ratios of the Green functions \( G(x,\xi) \) and \( G_V(x,\xi) \), it follows as well that \[ cK(x,y)\le K_V(x,y) \le CK(x,y) \quad\text{for } x,y\in D,\] where \( K \) and \( K_V \) are the Martin kernels for \( \frac{1}{2}\Delta \) and \( -\frac{1}{2}\Delta+ V \), respectively. Two-dimensional versions of these results appeared in Bass and Burdzy [e29], Cranston [e16], McConnell [e18], and Zhao [e11]. Results similar in flavor and which also incorporate the notion of IU above are due to Bañuelos [e24], who proved that, when the conditional gauge is finite and \( D \) is a Lipschitz or NTA domain, there exist positive constants \( c_t \) and \( C_t \) such that \[c_tp(t,x,y)\le p_V(t,x,y)\le C_tp(t,x,y) \quad\text{for }t > 0\text{ and }x,y\in D, \] where \( p_V \) is the heat kernel for \( -\frac{1}{2}\Delta+V \). An additional result of Bañuelos in this connection is that, if the conditional gauge is finite and \( D \) is an \( H(0) \) domain (as described earlier), then the operator \( -\frac{1}{2}\Delta+ V \) is IU. It’s interesting to note that the proofs used log-Sobolev inequalities. Further developments appear in a series of papers by Chen and Song [◊], and Chen [e32], among others. In [e31], the authors follow the developments of Bañuelos [e24], and consider the conditional gauge problem for the fractional Laplacian, \( (-\Delta)^\alpha \) for \( 0 < \alpha < 2 \), and potentials in the suitably modified Kato class \( K_{\alpha,d} \), where \( V\in K_{\alpha,d} \) if \[\lim_{r \rightarrow 0}\sup_{\{x\in \mathbb{R}^d\}} \int_{|x-y| < r}\frac{|V(y)|\ }{\ |x-y|^{d-\alpha}} \,dy=0. \] In this paper, Chen and Song [e31] deduced the CGT on \( C^{1,1} \) domains for the operator \( (-\Delta)^\alpha \) and \( K_{\alpha,d} \) potentials. The proper process to use in the Feynman–Kac representation in this case is the symmetric stable process of order \( \alpha \), \( X \), rather than the Brownian motion used when considering \( \Delta \). Their approach was to split the potential, writing \( V=V_1+V_2 \) for \( V_2\in \mathbf{L}^\infty \) and \( V_1 \) with a small Kato norm, that is, with small \[ \sup_{x\in D}\int_D \frac{|V_1(y)|}{\ |x-y|^{d-\alpha}}dy .\] Then, by a simple lemma of Khasminski, they show that the Green functions \( G^\alpha_{V_1} \) for \( (-\Delta)^\alpha+V_1 \) and \( G^\alpha \) for \( (-\Delta)^\alpha \) on \( D \) satisfy \( G^\alpha_{V_1}\approx G^\alpha \), in the sense that there are positive constants \( c \) and \( C \) such that \( cG^\alpha\le G^\alpha_{V_1}\le C G^\alpha \). This equivalence can then be used to prove that \( (-\Delta)^\alpha+V \) is IU. However, \( (-\Delta)^\alpha+V \) being IU implies that \( G^\alpha \approx G^\alpha_{V} \). Now, using a formula analogue to \eqref{ceqn7}, the finiteness of the right-hand side follows from the inequality \( G^\alpha_{V_1}\le C G^\alpha \). From the 3\( G \)-Theorem for the Green function \( G^\alpha \) on \( D \), the CGT follows. This was extended in [e33] to \( H(0) \) domains, again by an approach inspired by [e24].
Relations between subcriticality and boundedness of the conditional gauge have been investigated by Zhao [e22]. Consider the class \[ B_c=\{q:\mathbb{R}^d\to\mathbb{R}:\operatorname{supp}q \text{ is compact}\} \cap \mathbf{L}^\infty .\] The operator \( -\frac{1}{2}\Delta +V \) is called subcritical if \[\text{for all }q\in B_c \text{ there exists } \varepsilon > 0 \text{ such that } -\tfrac{1}{2}\Delta +V+\varepsilon q\ge 0 . \] This amounts to a strict positivity of \( -\frac{1}{2}\Delta +V \).
For a subclass of \( K_d \) potentials which satisfy a condition at \( \infty \), Zhao [e22] proved that subcriticality is equivalent to \[ u(x,y)=E_x^y\bigl[e^{\int_0^\infty V(B_s)\,ds}\bigr] \text{ is bounded on } \mathbb{R}^d\times \mathbb{R}^d. \] There were many other equivalences in that work which go a long way toward establishing the power of the approach in investigating the Schrödinger operator.
Generalizations of the Conditional Gauge Theorem to broader classes of Markov processes and potentials, including measures, have been carried out in Chen and Song [e31] and Chen [e32]. In the last work, Chen has proved gauge and conditional gauge theorems for a new class of Kato potentials, which even includes singular measures and general transient Borel right processes. And, most strikingly, following a suggestion of Chung, he proved that the CGT is actually the Gauge Theorem for the conditional process!
In this review, we’ve examined some of the many results which have connections with the works of Chung to be found in this volume. While we haven’t explicitly drawn the connections, we hope that these ties will become obvious to any reader of this volume. Finally, the author would like to express his gratitude to Professor Chung for introducing him to the fascinating problems in this area.