October 02, 2025
Convex optimization is the powerhouse behind the theory and practice of optimization. We introduce a quantum analogue of unconstrained convex optimization: computing the minimum eigenvalue of a Schrödinger operator \(h = -\Delta + V\) with convex potential \(V:{\mathbb{R}^n}\rightarrow \mathbb{R}_{\ge 0}\) such that \(V(x)\rightarrow\infty\) as \(\|x\|\rightarrow\infty\). For this problem, we present an efficient quantum algorithm, called the Fundamental Gap Algorithm (FGA), that computes the minimum eigenvalue of \(h\) up to error \(\epsilon\) in polynomial time in \(n\), \(1/\epsilon\), and parameters that depend on \(V\). Adiabatic evolution of the ground state is used as a key subroutine, which we analyze with novel techniques that allow us to focus on the low-energy space. We apply the FGA to give the first known polynomial-time algorithm for finding the lowest frequency of an \(n\)-dimensional convex drum, or mathematically, the minimum eigenvalue of the Dirichlet Laplacian on an \(n\)-dimensional region that is defined by \(m\) linear constraints in polynomial time in \(n\), \(m\), \(1/\epsilon\) and the radius \(R\) of a ball encompassing the region.
Convex optimization is the infrastructure supporting the theory and practice of optimization. The study of minimization of convex functions has led to a wide range of tools that are essential even for nonconvex and combinatorial problems [1], [2]. For optimization problems over quantum states, however, such a versatile framework is yet to be developed.
In this work, we introduce the problem of computing the minimum eigenvalue of a Schrödinger operator \[\begin{align} h \;= \;-\Delta \; + \;V \;= \;- \sum_{i\in[n]} \frac{\partial^2}{\partial x_i^2} \;+ \;V \end{align}\] with convex potential \(V:{\mathbb{R}^n}\rightarrow \mathbb{R}_{\ge 0}\) such that \(V(x)\rightarrow\infty\) as \(\|x\|\rightarrow\infty\) as a quantum analogue of unconstrained convex optimization. For this problem, we present an efficient quantum algorithm, called the Fundamental Gap Algorithm (FGA), that computes the minimum eigenvalue of \(h\) up to error \(\epsilon\) in polynomial time in \(n\), \(1/\epsilon\), and parameters that depend on \(V\). Adiabatic evolution of the ground state [3], [4] is used as a key subroutine, which we analyze with novel techniques that allow us to focus on the low-energy space. We apply the FGA to give the first known polynomial-time algorithm for finding the lowest frequency of an \(n\)-dimensional convex drum up to error \(\epsilon\). More precisely, we compute the minimum eigenvalue of the Dirichlet Laplacian on an \(n\)-dimensional region that is defined by \(m\) linear constraints in polynomial time in \(n\), \(m\), \(1/\epsilon\), and the radius \(R\) of a ball encompassing the region.
The Schrödinger operator \(h\) acts on a function \(\psi :{\mathbb{R}^n}\rightarrow \mathbb{C}\) to give another function \(h\psi:{\mathbb{R}^n}\rightarrow\mathbb{C}\) such that \[\begin{align} (h\psi)(x) = -\sum_{i\in[n]}\frac{\partial ^2}{\partial x_i^2}\psi(x) + V(x)\psi(x). \end{align}\] The minimum eigenvalue problem is to compute the smallest number \(\lambda_0\in\mathbb{R}\) for which there exists \(\psi\ne0\) such that \(h\psi = \lambda_0\psi\). Since the eigenvectors of \(h\) form a complete set of basis, we have an equivalent characterization \[\begin{align} \lambda_0 = \min_{\langle \psi|\psi\rangle = 1} \langle \psi| h\psi\rangle , \qquad \text{where} \qquad \langle \phi|\psi\rangle = \int_{\mathbb{R}^n}\overline{\phi(x)}\psi(x) \text{d} x. \end{align}\] Physically, the wavefunction \(\psi\) such that \(\langle\psi|\psi\rangle =1\) represents a quantum particle in \({\mathbb{R}^n}\). The total mechanical energy of the particle is \(\langle \psi|h\psi\rangle\), which is the sum of the kinetic energy \(-\langle \psi|\Delta\psi\rangle\) and the the potential energy \(\langle \psi|V\psi\rangle\). Our goal is to find the minimum mechanical energy that a quantum particle can have under the convex potential \(V\).
Why is this problem a quantum analogue of unconstrained convex optimization? Minimizing each of \(-\langle \psi|\Delta\psi\rangle\) and \(\langle \psi|V\psi\rangle\) reduces to classical convex optimization in the frequency and the position domain, respectively. When they are added together, however, the spectrum becomes discretized (quantized) and the minimizer is no longer a classical state; a quantum problem emerges from two classical objectives. Alone classical, together quantum.
Another reason is a physical interpretation of accelerated gradient descent [5], a foundational algorithm for convex optimization: the algorithm solves convex optimization by minimizing the classical mechanical energy of a particle under a convex potential. Therefore, finding the minimum quantum mechanical energy under convex potential is natural quantum analogue of convex optimization.
Previous works designed quantum algorithms for classical optimization problems by quantizing the classical dynamics of various gradient descent algorithms [6]–[10], [10]–[16] and friction [17]. Our work departs from these approaches in that we optimize quantum objectives, rather than applying a quantum algorithm for a classical objective.
To clarify what we mean by a quantum objective function, we adapt the prover-verifier definition of optimization [18]. A pair of algorithms \((p,v)\) minimizes a function \(f\) to the value \(\alpha\), if \(v\) is a verifier for the statement “\(\min f\le \alpha\)", and \(p\) generates a witness \(|w\rangle\) that \(v\) accepts (formal: Definition 1). The operator \(h\) is a genuinely quantum objective in the sense that a low eigenvalue can be efficiently verified by a quantum algorithm (e.g. phase estimation, as we show in the paper), whereas an efficient classical verifier is less plausible to exist.
As far as we are aware, only few known rigorous optimization algorithms require quantum verification. For example, dissipation-based algorithms for optimizing quantum Hamiltonians in polynomial time [19], [20] and exponential time [21] are recently introduced. More closely related to our work is a quantum algorithm computing the ground energy of a Schrödinger operator on a bounded domain \([0,1]^n\) with the Dirichlet boundary condition [22]. This algorithm prepares the ground state of an initial Hamiltonian, and then repeatedly measures in the energy basis of a slowly varying Hamiltonian. For the correctness of the algorithm, the Fundamental Gap Theorem [23] is employed to assert the spectral gap of the time-dependent Hamiltonian.
We adapt the key ideas from [22]. We replace the repeated measurements with a time-dependent Hamiltonian simulation, while the Fundamental Gap Theorem [23] gives the spectral gap for the performance guarantee. At the same time, we face an additional challenge that \(h\) is over the infinite domain \(\mathbb{R}^n\). To address this issue, we develop a set of tools that allow us to ignore high-energy parts of a Hamiltonian. We define a low-energy truncation of \(h\), which represents the operator up to its low-energy sector with error \(\epsilon\) (Definition 6). We show that if two Hamiltonians admit similar truncations, then their spectral gaps and the ground energies are close to each other. For the analysis, we construct a series of intermediate Hamiltonians on different domains that are treated in a modularized manner and connected via the truncation techniques.
As an application of the FGA, we give the first known polynomial-time algorithm for finding the minimum frequency of an \(n\)-dimensional convex drum specified by linear constraints. Mathematically, the problem is to compute the minimum eigenvalue of the Laplacian under the Dirichlet boundary condition on an \(n\)-dimensional polytope. The runtime of classical solvers such as finite element methods [24] grows exponentially to \(n\), exhibiting the curse of dimensionality. Information theoretic arguments [25] show that no classical algorithm runs in time polynomial in both \(n\) and \(1/\epsilon\).
Our objective function \(h\) is defined on the function space \[\begin{align} L^2({\mathbb{R}^n}):= \left\{\psi:{\mathbb{R}^n}\rightarrow\mathbb{C} \;| \; \langle \psi | \psi\rangle <\infty \right\}. \end{align}\] An optimization problem is often stated as: find the best element from a given set. It is unclear what it means to find the best element over the inifinite-dimensional space \(L^2({\mathbb{R}^n})\), when we are bound to a physical machine hosting only a finite number of qubits.
The infinity of the domain is already a familiar issue in classical optimization. For instance, we know \(\min_{x\in\mathbb{R}} x^2 =0\) at \(x = 0\) even though \(\mathbb{R}\) is infinite. More generally, if we are given certificate for \(f(x^*)\), we know that \(\min_x f(x)\le f(x^*)\). This idea of a certificate of optimization can be formalized as follow.
Definition 1 (generalization of [18]). Given an objective function \(f:X \rightarrow\mathbb{R}\), a pair of algorithms \((p,v)\) minimizes \(f\) to a value \(\alpha\) up to error \(\epsilon\), if the following conditions hold.
The algorithm \(v\) correctly verifies \(\min_X f\le \alpha\):
If \(\text{min}_{x\in X}f(x) \le \alpha\): \(\exists\) \(|w\rangle\) on finitely many (qu)bits such that \(v(|w\rangle) =1\) w.p. \(\ge 2/3\).
If \(\text{min}_{x\in X}f(x) \ge \alpha +\epsilon\): \(\forall\) \(|w\rangle\) on finitely many (qu)bits, we have \(v(|w\rangle) =1\) w.p. \(\le 1/3\).
The algorithm \(p\), w.p. \(\ge 2/3\), outputs a finite length \(|w'\rangle\) such that \(v(|w'\rangle)=1\).
In this paper, we find the minimum of objective functions in polynomial time in the following sense.
Definition 2. Given an objective function \(f:X \rightarrow\mathbb{R}\), a pair of algorithms \((p,v)\) finds the minimum of \(f\) up to error \(\epsilon\) in time \(T\), if \((p,v)\) minimizes \(f\) to the value \(\min_{x\in X} f(x)\) up to error \(\epsilon\) and both \(p\) and \(v\) halt in time \(T\).
Whether \(p\), \(v\), and \(|w\rangle\) are classical or quantum is left unspecified by the definition; all classical/quantum combinations are allowed. In our case, \(p, v\) and \(|w\rangle\) are all quantum.
The main problem of this paper is to find the minimum nonzero eigenvalue \(\lambda_0\) of a Schrödinger operator \(h = -\Delta + V\) where \(V(x)\rightarrow\infty\) as \(\|x\|\rightarrow\infty\), and \(V\) is convex.
Problem 1 (Schrödinger). Given a Schrödinger operator \(h=-\Delta + V\) with a smooth convex function \(V:{\mathbb{R}^n}\rightarrow\mathbb{R}_{\ge 0}\) such that \(V(x)\rightarrow\infty\) as \(\|x\|\rightarrow\infty\), \[\begin{align} \text{minimize}& \qquad \lambda_0 \\ \text{subject to}& \qquad \psi:{\mathbb{R}^n}\rightarrow\mathbb{C},\\ &\qquad \lambda_0 \psi = h\psi, \\ & \qquad \psi\ne 0. \end{align}\]
We have an equivalent formulation in terms of energy, since \(h\) has a purely discrete spectrum. We define the energy functional as follows.
Definition 3 (Energy, or Rayleigh quotient). Let \(h\) be a self-adjoint operator on a Hilbert space \(\mathcal{H}\) with operator domain \(\mathcal{D}(h)\subseteq \mathcal{H}\). The energy of a nonzero vector \(\psi \in \mathcal{D}(h)\) with respect to \(h\) is denoted \[h[\psi] := \frac{\langle \psi | h \psi \rangle}{\langle \psi | \psi \rangle}.\]
The Hilbert space \(\mathcal{H}\) is \(L^2({\mathbb{R}^n})\) for the Schrödinger problem. The operator domain \(\mathcal{D}(h)\subseteq \mathcal{H}\) is the subset on which the operator \(h\) is defined and \(h\psi \in \mathcal{H}\). For more explanation, see Section 4.
Problem 2 (Schrödinger). Given a Schrödinger operator \(h=-\Delta + V\) with a smooth convex function \(V:{\mathbb{R}^n}\rightarrow\mathbb{R}_{\ge 0}\) such that \(V(x)\rightarrow\infty\) as \(\|x\|\rightarrow\infty\), \[\begin{align} \text{minimize}& \qquad h[\psi]. \end{align}\]
There is an ambiguity as to how \(V\) is given to us. We assume that we have access to a circuit that computes \(V(x)\) when \(\|x\|_\infty \le L/2\). The Fundamental Gap Algorithm efficiently solves Problem 1 for \(V\) satisfying certain conditions.
Theorem 1 (main). Algorithm 1 computes the lowest eigenvalue of \(h = -\Delta + V\) within error \(\epsilon_0\) with probability \(\ge 2/3\) in polynomial time in \(n, 1/\epsilon_0,L,c\), if \(V\) satisfies the following conditions, where \(B^q_d:=\{x\in{\mathbb{R}^n}\;| \; \|x\|_q\le d\}\) and \(V_E^{-1}:= V^{-1}([0,E])\):
Purely discrete spectrum/ and convexity. We have \(V(x)\rightarrow\infty\) as \(\|x\|\rightarrow\infty\) and \(V\) is convex on \(B^2_{r+1}\).
“Bowl-shaped” in \(B^\infty_{L/2}\). There exist \(a\le b\le c\), and \(1 \le r < L/2\) such that \[\begin{align} B^2_1\subset V^{-1}_{a } \subset V^{-1}_{b}\subset B^2_r \subset B^2_{r+1} \subset V^{-1}_{c}\subset V^{-1}_{c+1} \subset B^\infty_{L/2} \end{align}\] where
(i) \(b = \Theta(E_0^7/\sigma^6)\)
(ii) \(E_0=10(n(n+3)\pi^2 +a) = \Theta(n^2+a)\)
(iii) \(\sigma = \Theta(\epsilon_0/r^4 E_0^{1.5})\)
(iv) \(\epsilon_0<1 < r,E_0,c\).
Bounded high-order derivatives. The potential \(V\) is smooth and \[\begin{align} \max \left\{ { \frac{1}{l}\log|\partial^l_j V(x)| } \; \bigg| \;{l\in[m], j\in[n], x\in [-L/2,L/2]^n }\right\} \le \log p(m) \end{align}\] for some polynomial \(p\).
Access to \(V\). There exists a circuit that computes \(V(x)\) for all \(x \in B^\infty_{L/2}\) with exponential small error in the complexity parameters.
Proof. See Section 6. ◻
Note that \(V\) is not required to be convex only in \(B^2_{r+1}\), and also, not every convex \(V\) satisfy the conditions.
We provide an example to demonstrate that Theorem 1 is not void, and indeed useful. The example is to compute the minimum eigenvalue of the Laplacian on an \(n\)-dimensional domain \(\Omega\). We impose the Dirichlet boundary condition, which enforces that \[\begin{align} \psi(x) &= 0 \qquad\qquad \forall x\in\partial \Omega. \end{align}\] We assume that we are given \[\begin{align} \Omega = \{ x\in | \;a_j \cdot x \le b_j , \;\forall j \in [m]\} \subset {\mathbb{R}^n}, \end{align}\] where \(a_j \in \mathbb{R}^n\) and \(\|a_j\|_2 = 1\) for each \(j\). Additionally, we assume that \(b_j\ge 1\) and also that \(\Omega\subset B^2_R\), giving that \[\begin{align} B^2_1\subset \Omega\subset B^2_R. \end{align}\]
The task is to solve the following eigenvalue problem.
Problem 3 (Drum). Given a polytope \(\Omega = \{ x\in \mathbb{R}^n | \;a_i \cdot x \le b_i , \;\forall i \in [m]\},\) \[\begin{align} \text{minimize}& \qquad \lambda_0 \\ \text{subject to} & \qquad \psi:\Omega \rightarrow \mathbb{C}, \\ & \qquad \lambda_0 \psi(x) = -\Delta\psi(x) \qquad \forall x \in \Omega \setminus \partial \Omega,\\ & \qquad \psi(x) = 0 \qquad \qquad \forall x \in \partial \Omega, \\ & \qquad \psi \ne 0. \end{align}\]
Similarly to the Schrödinger problem, we have an equivalent problem in terms of the energy functional (see Definition 3). Here \(\Delta^D_\Omega\) is the Laplacian operator on \(\Omega\) with the Dirichlet boundary condition. For the Drum problem, \(\mathcal{H} = L^2(\Omega)\) is the appropriate Hilbert space.
Problem 4 (Drum). Given a polytope \(\Omega = \{ x\in \mathbb{R}^n | \;a_i \cdot x \le b_i , \;\forall i \in [m]\},\) \[\begin{align} \text{minimize} & \qquad - \Delta^D_\Omega[ \psi] . \end{align}\]
Theorem 2. Given \(\Omega = \{ x\in \mathbb{R}^n | \;a_i \cdot x \le b_i , \;\forall i \in [m]\}\) satisfying \(B^2_1\subset \Omega\subset B^2_R\), Algorithm 2 solves Problem 3 up to error \(\epsilon_0\) with probability \(\ge2/3\) in polynomial time in \(n,m,R,\epsilon_0\).
Proof. See Section 6. ◻
Our algorithm applies an adiabatic evolution under the time‑dependent Hamiltonian \[\begin{align} \label{eq:adiabatic95path} H_Q(t) \;:=\; K_Q \;+\; V_Q(t), \qquad t\in[0,T], \end{align}\tag{1}\] acting on \(n\) registers of \(\log N\) qubits each. The kinetic term \(K_Q\) is diagonal in the discrete Fourier basis and mimics the continuous operator \(-\Delta\). The potential term \(V_Q(t)\) is diagonal in the computational basis and mimics a continuous potential operator. This discretization of kinetic and potential operators follows prior work in quantum simulation [26]–[29].
The time-dependent \(V_Q(t)\) interpolates between an initial potential \(V_Q(0)\) and a final potential \(V_Q(T)\). The initial potential is chosen so that \(K_Q + V_Q(0)\) decomposes as a sum of single‑register Hamiltonians \(H_N\). Consequently, the ground state of \(H_Q(0)\) can be prepared in polynomial time as the product of the ground states of the \(H_N\). The final potential is defined so that \(\lambda_0\big(H_Q(T)\big)\) is close to \(\lambda_0(h)\), where \(\lambda_0(A)\) is the minimum eigenvalue of a self-adjoint \(A\).
We first describe how we discretize Schrödinger operators so that they are simulable on a quantum computer, and then state our two algorithms.
It is more natural to discretize a Schrödinger operator \(h_{\mathbb{T}^n}= -\Delta_{\mathbb{T}^n}+ V_{\mathbb{T}^n}\) on the torus \[\begin{align} {\mathbb{T}^n}: = {\mathbb{R}^n}/ L\mathbb{Z}^n \end{align}\] than \(h\) on \({\mathbb{R}^n}\). Here, \(\Delta_{\mathbb{T}^n}= - \sum_{i\in[n]} \partial^2/\partial x_i^2\) is the Laplacian on the flat torus, and \(V_{\mathbb{T}^n}:{\mathbb{T}^n}\rightarrow\mathbb{R}_{\ge 0}\) is smooth. We discretize \({\mathbb{T}^n}\) to the grid \(\{\, yL/N \;|\; y\in\mathcal{N}^n \,\}\), where \[\begin{align} \mathcal{N} \;:=\; \left\{ -\frac{N}{2},\ldots,-1,0,1,\ldots,\frac{N}{2}-1 \right\}. \label{eq:grid95label} \end{align}\tag{2}\] We assume \(N\) is a power of 2, and identify the labels for the computational basis states modulo \(N\), so that, for \(y,y'\in \mathbb{Z}^n\), \[\begin{align} |y\rangle = |y'\rangle \qquad \text{whenever} \qquad y = y' \mod N. \end{align}\]
We define the discretized Schrödinger, kinetic, and potential operators to be \[\begin{align} H_Q &:= K_{Q} + V_Q, \\ K_Q &:= \sum_{i=1}^n I_N^{\otimes(i-1)} \otimes K_N \otimes I_N^{\otimes (n-i)}, \\ V_Q &:= \sum_{y\in \mathcal{N}^n} |y\rangle \; V_{{\mathbb{T}^n}}\!\left(\frac{yL}{N}\right)\! \langle y |, \end{align}\] where \(I_N\) is the \(N\times N\) identity, \[\begin{align} K_N \;:=\; \sum_{k_0 \in \mathcal{N}} U_N |k_0 \rangle \; \frac{4\pi^2 k_0^2}{L^2}\; \langle k_0 |\, U_N^\dagger , \end{align}\] and \(U_N\) is the \(N\)‑dimensional quantum Fourier transform acting on \(|k_0\rangle\) for \(k_0\in \mathcal{N}\) by \[\begin{align} U_N|k_0\rangle \;=\; \frac{1}{\sqrt{N}} \sum_{y \in \mathcal{N}} e^{\,i 2\pi\, k_0 y/N}\, | y\rangle. \end{align}\] Equivalently, we can also write \[\begin{align} K_Q \;:=\; \sum_{k \in \mathcal{N}^n} U_{\mathcal{F}} |k \rangle \; \frac{4\pi^2 \|k\|_2^2}{L^2}\; \langle k |\, U_{\mathcal{F}}^\dagger \end{align}\] where \(U_{\mathcal{F}} = \bigotimes_{i=1}^n U_N\) is the discrete Fourier transform on \(\mathcal{N}^n\).
We define the time-dependent Hamiltonian \[\begin{align} H_Q(t) \;:= \;K_Q + \frac{T-t}{T}V_Q(0) + \frac{ t }{T }V_Q(T), \end{align}\] where \[\begin{align} V_{Q}(0) &:= b\sum_{ y\in{\mathcal{N}^n}} |y\rangle \; \sum_{i\in[n]} \left( 1 - \text{Cut}_{\frac{1}{4\sqrt n},\frac{1}{4\sqrt n}}\!\left(\left|\tfrac{y_i L}{N} \right| \right) \right)\langle y |,\tag{3}\\ V_{Q}(T) &:= \sum_{y \in{\mathcal{N}^n}} |y\rangle \; \text{Sat}_{c,1}\!\circ\! V\big(\tfrac{yL}{N}\big) \langle y| \tag{4}, \end{align}\] and \(b,c \in\mathbb{R}_{\ge 0}\). Here, \(\text{Cut}_{\frac{1}{4\sqrt n },\frac{1}{4\sqrt n }}:\mathbb{R}\rightarrow \mathbb{R}\) is a smooth cutoff function and \(\text{Sat}_{c,1}:\mathbb{R}\rightarrow\mathbb{R}\) is a smooth saturating function (Definition 5) such that \[\begin{align} \text{Cut}_{\frac{1}{4\sqrt n },\frac{1}{4\sqrt n }}(x) & \; \begin{cases} = 1 \qquad\qquad & x \le \frac{1}{4 \sqrt n}, \\ \in[0,1] \qquad\qquad &x\in [\frac{1}{4 \sqrt n}, \frac{1}{2 \sqrt n}],\\ = 0 \qquad\qquad & x \ge \frac{1}{2 \sqrt n}, \end{cases}\\ \text{Sat}_{c,1}(x) & \; \begin{cases} = x \qquad & x \le c, \\ :\text{monotone increasing} \qquad &x\in [c, c+1],\\ : \text{some constant in} \; [c,c+1] \qquad & x \ge c+1. \end{cases} \end{align}\] Note that the initial Hamiltonian \(H_Q(0)= K_Q + V_Q(0)\) is a sum of single-register Hamiltonians: \[\begin{align} H_Q(0) = &\sum_{i=1}^n I_N^{\otimes(i-1)} \otimes H_N \otimes I_N^{\otimes (n-i)}, \nonumber\\ \text{where}\qquad H_N := &K_N \;+\; b\sum_{y_0\in \mathcal{N}} |y_0\rangle \left( 1 - \text{Cut}_{\frac{1}{4\sqrt n}, \frac{1}{4\sqrt n}} \!\left( \left|\frac{y_0L}{N}\right| \right) \right)\! \langle y_0|, \label{eq:hn} \end{align}\tag{5}\]
We present our main algorithm that we call the Fundamental Gap Algorithm (FGA).
Figure 1: .
Since we are only aiming for a polynomial-time algorithm, any off-the-shelf Hamiltonian algorithm suffices. For instance, we can use the algorithm from [30].
Our approach for the drum problem is to run FGA for a Schrödinger operator on \({\mathbb{R}^n}\) with the potential defined by a barrier function that penalizes going outside of \(\Omega\). We use a smooth barrier function \(\text{Bar}_\epsilon\) (see Definition 5) such that \[\begin{align} \text{Bar}_\epsilon(x) \; \begin{cases} = 0 \qquad& \qquad x \le 0 \\ : \; \text{monotone increasing} \qquad & \qquad x \in[0,\epsilon] \\ : \;\text{increasing with slope } 1 \qquad & \qquad x \ge \epsilon \end{cases} . \end{align}\]
Figure 2: .
We adapt the language and tools of operator theory and PDE. This section does not aim to provide a complete survey. We refer interested readers to [31]–[33].
In this paper, we are concerned with functions that represent quantum states on \(\Omega\in\{{\mathbb{R}^n},B, {\mathbb{T}^n}\}\), where \(B\subset{\mathbb{R}^n}\) is a compact measurable subset.
A Hilbert space is a complete metric space under the metric induced by an inner product. For example, the space \[\begin{align} L^2(\Omega) = \{ \;\psi \;| \;\psi :\mathbb{R}^n \rightarrow \mathbb{C}, \;\int_{\Omega } |\psi(x)|^2 \text{d} x < \infty \} \end{align}\] is a Hilbert space, equipped with the inner product between \(\psi,\phi\in L^2(\Omega)\) given by \[\begin{align} \langle \psi| \phi\rangle :=\int_{ \Omega} \overline{\psi(x)} \phi(x) \;\text{d} x, \end{align}\] where \(\overline{a}\) is the complex conjugate of \(a\in \mathbb{C}.\)
An operator \(A\) on Hilbert space \(\mathcal{H}\) is a pair \((A , \mathcal{D}(A))\) where \(A\) is a linear map that is defined by its operator domain \(\mathcal{D}(A)\le \mathcal{H}\) and its action \(A:\mathcal{D}(A) \rightarrow \mathcal{H}.\) For example, the Laplacian operator \[\begin{align} \Delta:\psi\rightarrow \sum_{i \in [n]}\frac{\partial^2 \psi}{\partial x_i^2} \end{align}\] is an operator on \(\mathcal{H} = L^2(\mathbb{R}^n)\) that is defined on the domain \(\mathcal{D}({\Delta})\). Note that \(\Delta\psi\) is only defined on twice-differentiable functions as per the definition above, therefore, \(\mathcal{D}(\Delta) \subset L^2({\mathbb{R}^n})\cap C^2({\mathbb{R}^n})\). Furthermore, not every twice-differentiable \(\psi\in L^2({\mathbb{R}^n})\) yields \(\Delta \psi \in L^2({\mathbb{R}^n}).\) In many cases, we want an operator domain on which the operator is self-adjoint (see below). Finding such a domain is a nontrivial task that can be found in the mathematical physics literature [31].
An important family of operators is multiplication operators. Given a function \(f:\mathbb{R}^n \rightarrow \mathbb{C}\), multiplication operator \(M_f\) is defined on \(L_2(\mathbb{R}^n)\) so that for \(x\in\mathbb{R}^n\), \[\begin{align} M_f\psi (x) := f(x) \psi(x). \end{align}\] For example, the potential operator \(M_V\) is a multiplication operator. In this paper, we abuse the notation and denote \(V\) instead of \(M_V\), following the notation in physics.
A unitary operator is a surjective linear operator \(U:\mathcal{H} \rightarrow\mathcal{H}\) that preserves the inner product \[\begin{align} \langle\phi|\psi\rangle = \langle U\phi|U\psi\rangle \qquad \forall \phi,\psi\in\mathcal{H}. \end{align}\]
Define operator norm \(\|A\|\) of an operator \(M\) on \(\mathcal{H}\) to be \[\begin{align} \|A\|:=\sup_{\psi \in \mathcal{D}(A)} \frac{\|A\psi\|}{\|\psi\|} = \sup_{\psi \in\mathcal{D}, \|\psi\| =1} \|A\psi\|. \end{align}\] For a multiplication operator \(M_V\), we have \(\|M_f\| = \|f\|_{\infty}.\)
An operator \(A\) is called symmetric if its domain \(\mathcal{D} (A)\) is dense in \(\mathcal{H}\), and for all \(\psi, \phi \in \mathcal{D}(A)\), we have \(\langle A\psi |\phi\rangle = \langle \psi |A\phi\rangle.\)
Given that \(\mathcal{D}(A)\) is dense in \(\mathcal{H}\), its adjoint operator \(A^\dagger\) is defined so that \(A^\dagger \psi = z\) where \(z\) is the unique vector satisfying \(\langle A\phi |\psi \rangle=\langle \phi | z \rangle \;\forall \phi \in \mathcal{D}(A)\). The domain \(\mathcal{D}(A^\dagger)\) is the set of \(\psi\) for which such \(z\) exists.
The operator \(A\) is self-adjoint if \(\mathcal{D}(A) = \mathcal{D}(A^\dagger)\) and \(A\psi = A^\dagger\psi\) for all \(\psi \in \mathcal{D}(A)\).
A vector \(\psi \in\mathcal{D}(A)\setminus \{0\}\) is an eigenvector of \(A\) if there exists \(\lambda \in \mathbb{C}\) such that \(A\psi = \lambda \psi.\) The value \(\lambda\) is called the eigenvalue of \(\psi.\)
The spectrum of \(A\) is the set \[\begin{align} \text{Spec}(A) :=\{\lambda \;|A - \lambda I \text{ does not have an inverse with a bounded norm\}}. \end{align}\] If \(\lambda\) is an eigenvalue, then \(\lambda \in \text{Spec}(A)\). The inverse of this statement is not necessarily true.
A spectrum \(\text{Spec}(A)\) is purely discrete if each \(\lambda\in \text{Spec}(A)\) is an eigenvalue, the multiplicity of \(\lambda\), defined as \(\dim \ker ( A - \lambda I)\), is finite, and \(\text{Spec}(A)\) accumulates at no other point than \(\infty\).
A Schrödinger operator \(h=-\Delta + V\) is a linear operator that is the sum of the negative Laplacian \(-\Delta\), and a multiplication operator \(V\). Its action on \(\psi\in \mathcal{D}(h)\) is defined by \[\begin{align} h\psi(x)= -\sum_{i\in[n]} \frac{\partial^2}{\partial x_i^2} \psi(x) + V(x)\psi(x). \end{align}\] If \(\Omega = B\subset{\mathbb{R}^n}\) is a bounded domain, we always impose the Dirichlet boundary condition, which asserts that an operator acts only on \(\psi\in L^2(\Omega)\) such that \[\begin{align} \psi(x) = 0 \qquad \forall x\in\partial B. \end{align}\] We use the superscript \(D\) to denote that the operator is under the Dirichlet boundary condition. For example, \(\Delta_B^D\) is the Dirichlet Laplacian operator on the domain \(B\).
In this paper, we are interested in Schrödinger operators in the following forms.
Definition 4 (Schrödinger operators). We define sets of Schrödinger operators on the real space \(\mathbb{R}^n\), on a compact subset with smooth boundary \(\Omega \subset \mathbb{R}^n\) with the Dirichlet boundary condition, and on the length \(L\) torus \(\mathbb{T}^n :=\mathbb{R}^n/L\mathbb{Z}^n\) as follows: \[\begin{align} S_{\mathbb{R}^n} :=& \{ h = -\Delta_{\mathbb{R}^n} + V \;| \;V:\mathbb{R}^n \rightarrow \mathbb{R}_{\ge 0}, V\in \mathcal{C}^\infty(\mathbb{R}^n), \; V(x)\rightarrow\infty \text{ as } |x| \rightarrow\infty \} \\ S_{\Omega}^D :=& \{h= - \Delta_{\Omega}^D +V \; | \; V: \Omega\rightarrow \mathbb{R} _{\ge 0}, \;V\in C^\infty (\Omega\mathcal{)}, \;\text{with the Dirichlet boundary condition} \}, \\ S_{\mathbb{T}^n} := &\{h= - \Delta_{\mathbb{T}^n} +V \;|\; V: \mathbb{T}^n \rightarrow \mathbb{R} _{\ge 0}, \;V\in \mathcal{C}^\infty (\mathbb{T}^n) \}, \end{align}\] where \(\Delta_{\mathbb{R}^n}\), \(\Delta^D_{\Omega}\), \(\Delta_{\mathbb{T}^n}\) denote the Laplacian operators \(\sum_{i\in[n]}\partial^2/\partial x_i^2\) on the respective domain with the respective boundary condition. For each \(h \in S_{\mathbb{R}^n}\cup S^D_\Omega \cup S_{\mathbb{T}^n}\), we assume that \(h\) is self-adjoint on its domain \(\mathcal{D}(h)\), and \(\mathcal{D}(h)\) contains \(C^\infty_0(\Omega)\), the set of smooth functions with compact support in \(\Omega\).
Remark 1. It is a non-trivial but well-known fact that for an operator of the form \(h \in S_{\mathbb{R}^n}\cup S^D_\Omega \cup S_{\mathbb{T}^n}\), we can define \(\mathcal{D}(h)\) so that \(h\) is self-adjoint. See Theorem X.28 of [31] for \({\mathbb{R}^n}\), Theorem 3.8 of [34] for a bounded \(\Omega \subset {\mathbb{R}^n}\), and the Sobolev space \(H^2({\mathbb{T}^n})\) forms an operator domain on which \(\Delta_{\mathbb{T}^n}\) is self-adjoint.
It is important to us that each Schrödinger operator in which we are interested has a purely discrete spectrum. Eigenfunctions of an operator with a purely discrete spectrum form a complete basis set of the Hilbert space. Furthermore, an eigenfunction of a Schrödinger operator is smooth as a function.
Lemma 1 (Theorem 3.10 of [34]). Let \(A\) be a self-adjoint operator with a purely discrete spectrum on the Hilbert space \(\mathcal{H}\). Then, the eigenvectors of \(A\) form an orthonormal basis in \(\mathcal{H}\).
Lemma 2 (Spectrum of a Schrödinger operator). Suppose \(h \in S_{\mathbb{R}^n}\cup S^D_\Omega \cup S_{\mathbb{T}^n}\). Then, \(\text{Spec}(h)\) is purely discrete, and the eigenvectors form a complete orthogonal basis.
Proof. By Lemma 1, it is enough to show that \(h\) has a purely discrete spectrum. For \(h \in S_{\mathbb{R}^n}\), see [35]. For \(h \in S^D_\Omega \cup S_{\mathbb{T}^n}\), Weyl’s essential spectrum theorem applied to the corresponding Laplacian operators gives the purely discreteness of the spectrum. ◻
Lemma 3 (Smoothness of eigenfunctions). Let \(h\in S_{\mathbb{R}^n}\cup S^D_\Omega \cup S_{\mathbb{T}^n}\). Let \(\psi \in \mathcal{D}(h)\) be an eigenfunction of \(h\). Then \(\psi\) is a smooth function.
Proof. An eigenfunction of \(h\) is a solution of an elliptic PDE, whose smoothness is a well-studied problem. The hypoellipticity of Schrödinger operators on \({\mathbb{R}^n}\) [36], the elliptic regularity on a compact, smooth-boundary domain with the Dirichlet boundary [33] and the elliptic regularity on \({\mathbb{T}^n}\) [37] give the smoothness of the eigenfunction of \(h\) in each case. ◻
Definition 5 (useful smooth functions).
We define a smooth bump function, a smooth cutoff function, a smooth saturating function, and a smooth barrier function \[\begin{align} \text{Bump}(x) \;& = \; \begin{cases} \;\exp \left(\frac{1}{x(x-1)}\right) \qquad &x \in[0,1], \\ \;0 \qquad &x \notin [0,1], \end{cases} \\ \text{Cut}_{\alpha,\beta} (x) \;& = \;1 - {\int_\alpha^x \text{Bump}\big(\frac{y - \alpha}{\beta } \big) \;\text{d} y} \;\Big/ {\int_\alpha^{\alpha + \beta} \text{Bump}\big(\frac{y - \alpha}{\beta } \big) \;\text{d} y } ,\\ \text{Sat}_{\alpha, \beta}(x) \; &= \;\int_0^x \text{Cut}_{\alpha,\beta}(y) \;\text{d} y, \\ \text{Bar}_\epsilon(x) \;& = \;\int_0^x(1 - \text{Cut}_{0,\epsilon}(y)) \;\text{d} y. \end{align}\]
The analysis achieves two main goals. First, we want an inverse-polynomial spectral gap in \(H_Q(t)\) throughout \(t\in[0,T]\). Second, we want the lowest eigenvalue of \(h\) to be close to that of the final qubit Hamiltonian \(H_Q(T)\). Once we have established these points, the correctness of the algorithm follows from the Adiabatic Theorem.
In general, proving a lower bound on the spectral gap can be challenging. In our case, we leverage the Fundamental Gap Theorem, which ensures an inverse polynomial gap for a Dirichlet Schr̈odinger operator with a convex potential on a bounded convex domain.
We consider the four pairs of domains and Hamiltonians \[({\mathbb{R}^n},h),\quad (B, h_B^D),\quad ({\mathbb{T}^n},h_{{\mathbb{T}^n}}),\quad (\mathcal{N}^n,H_Q),\] where \(B\) is a Euclidean ball in \({\mathbb{R}^n}\). Each plays a different role in the analysis: \(h\) on \({\mathbb{R}^n}\) is the objective, \(h^D_B\) on \(B\) provides the spectral gap via the Fundamental Gap Theorem, \(h_{\mathbb{T}^n}\) on \({\mathbb{T}^n}\) plays the role of a “gearbox” that connects other Hamiltonians, and finally \(H_Q\) on the finite set \(\mathcal{N}^n\) is efficiently simulable on a quantum computer. We connect them through truncation lemmas.
The central fact in proving the correctness of our algorithm is the Adiabatic Theorem. For the rest of the paper, \(\lambda_0(H), \lambda_1(H)\) denote the lowest and the second lowest eigenvalues of \(H\), and \[\begin{align} \text{gap}(H):=\lambda_1(H) - \lambda_0(H). \end{align}\] We apply this theorem to the time-dependent qubit theorem \(H_Q(t)\) of Algorithm 1.
Theorem 3 (The Adiabatic Theorem (adapted from [38], [39])). Let \(H_{ init}\) and \(H_{final}\) be two Hamiltonians acting on a finite-dimensional quantum system. Consider the time-dependent Hamiltonian \[H(t) := (1 - t/T) H_{ init} + (t/T) H_{final},\] that has a unique ground state for all \(t\in [0,T]\).
Then the final state of an adiabatic evolution according to \(H(t)\) for \(t\in[0,T]\) is \(\epsilon\)-close to the ground state of \(H_{final}\), if the total evolution time \[T \ge \Omega\!\left( \frac{\|H_{\rm final} - H_{\rm init}\|^{2}}{\epsilon (\min_{t \in [0,T]} \text{gap}(H(s)))^3} \right).\] The matrix norm is the spectral norm \(\|H\| := \max_{\|w\|=1} \|H w\|\).
The Dirichlet Schrödinger operator \(h^D_B\) is a Schrödinger operator on \(B\) with the Dirichlet boundary. The Fundamental Gap Theorem [23] provides a lower bound on the spectral gap of \(h^D_B\) as an inverse polynomial in the diameter of \(B\).
Theorem 4 (Fundamental Gap Theorem [23]). Let \(\Omega \subset \mathbb{R}^n\) be a bounded convex domain of diameter \(R\), and \(V\) a convex potential. Then the eigenvalues of the Dirichlet Schrödinger operator \(h^D_\Omega = -\Delta^D_\Omega + V\) satisfy \[\begin{align} \text{gap}(h^D_\Omega) \ge \frac{3\pi^2}{R^2}, \end{align}\] where \(-\Delta^D_\Omega\) denotes the Dirichlet Laplacian operator on \(\Omega\).
A disconnection between Theorem 3 and Theorem 4 is that we need a bound on \(\text{gap}(H_Q(t))\), while we have a bound on \(\text{gap}(h^D_B)\).
We relate the properties of each Hamiltonian through Truncation Lemmas. The key observation is that only low‑energy properties need to be preserved; the rest of the Hilbert space can be ignored, even if infinite‑dimensional. For the adiabatic simulation, it suffices to approximate the first two eigenvalues. To formalize this, we introduce \((E,\epsilon)\)‑truncated domains, capturing the low‑energy subspace up to an error \(\epsilon\).
We then prove the Truncation Lemmas showing that if \((E,\epsilon)\)‑truncated domains of two Hamiltonians admit an isomorphism between them that approximately preserves energy, then their first two eigenvalues must be close to each other. Our truncation framework is tailored to adiabatic optimization and accommodates unbounded operators.
Definition 6 (Truncation, and Equivalence). For \(\epsilon>0\), a normalized vector \(\widetilde{\psi}\in\mathcal{D}(h)\) is an \(\epsilon\)-truncation of \(\psi\) with respect to \(h\), if \[\begin{align} i) & \qquad h[\widetilde{\psi}] \; \leq \; h[\psi] + \epsilon \qquad \qquad \text{(energy)} \\ ii) & \qquad \|\widetilde{\psi} - \psi\| \; \leq \;\epsilon \qquad \qquad \text{(norm)}. \end{align}\]
A subspace \(\widetilde{\mathcal{D}} \le \mathcal{H}\) is an \((E, \epsilon)\)-truncated domain of \(h\), if, for every normalized \(\psi \in \mathcal{D}(h)\) such that \(h[\psi]\le E\), there exists an \(\epsilon\)-truncation of \(\psi\) in \(\widetilde{ \mathcal{D}}\).
Self-adjoint operators \(h_a\) and \(h_b\) are \((E,\epsilon)\)-equivalent if there exist \((E,\epsilon)\)-truncated domains \(\widetilde{\mathcal{D}}_a\) of \(h_a\) and \(\widetilde{\mathcal{D}}_b\) of \(h_b\) that are isomorphic via a unitary \(U: \widetilde{\mathcal{D}}_a \rightarrow \widetilde{\mathcal{D}}_b\) such that \[\begin{align} \left| h_a[\widetilde{\psi}] - h_b[U\widetilde{\psi}] \right| \leq \epsilon \qquad \qquad \forall \;\widetilde{\psi} \in \widetilde{\mathcal{D}}_a . \end{align}\]
Lemma 4 (\(\lambda_0,\lambda_1\) approximation). Let \(h_a,h_b\ge 0\) be a self-adjoint linear operator with purely discrete spectrum. Suppose \(h_a\) and \(h_b\) are \((E,\sigma)\)-equivalent and the following conditions hold: \[\begin{align} &\text{gap}(h_a)\ge g \;> 0, \nonumber \\ &2({\lambda_1(h_a)} +1) \; \leq \; E,\quad \\ &\epsilon, E^{-1} \;< \;1 . \nonumber \end{align}\] If \(\epsilon \in [0,c]\) for a sufficiently small universal constant \(c\) and \[\begin{align} \sigma = O \left(\frac{\epsilon g}{E^{1.5}} \right), \end{align}\] then \[\begin{align} |\lambda_0(h_a) - \lambda_0(h_b)|, \; |\lambda_1(h_a) - \lambda_1(h_b)| \leq \epsilon . \end{align}\]
Proof. See Appendix 9. ◻
We explain how we apply the idea of truncation to Schrödinger operators of our interest. We assume that \(V\) is of a bowl shape: low around the center (the condition \(B^2_1\subset V^{-1}_a\)), and high outside the radius \(r\) (the condition \(V^{-1}_b \subset B^2_{r}\)). This assumption enables Markov’s inequality, which is employed to show that a wavefunction with a low energy has a low weight outside the ball \(B^2_{r}\), due to the high potential. We apply this position-based Markov’s inequality to truncate the Schrödinger operators on \({\mathbb{R}^n}\), \(B:=B^2_{r+1}\), and \({\mathbb{T}^n}\). These truncations in fact \((E,\epsilon)\)-equivalent, showing that the first two eigenvalues of the three Hamiltonians close to each other.
Lemma 5. Suppose \(h_1, h_2, h_3\) are Schrödinger operators on \({\mathbb{R}^n}, B :=B^2_{r+1}, {\mathbb{T}^n}\), respectively, and \(h_2\) is under the Dirichlet boundary condition \[\begin{align} h_1&:=-\Delta_{\mathbb{R}^n} + V_1 \in S_{\mathbb{R}^n}, \\ h_2&:=- \Delta_B^D + V_2 \in S_{B}^D, \\ h_3 &:= - \Delta_{\mathbb{T}^n} + V_3 \in S_{\mathbb{T}^n} , \end{align}\] where \(2(r+1)< L\). Furthermore, assume that \(\epsilon<1\) and \[\begin{align} \begin{cases} V_1(x)=V_2(x)=V_1(x) \;\;\;\;&\text{if} \;\;\|x\|_2\le r+1, \\ V_1(x),V_2(x),V_3(x) \ge b \;\;\;\;&\text{if} \;\;\|x\|_2\ge r. \end{cases} \end{align}\]
If \(b= \Omega( E^7/\epsilon^6)\), then \(h_i\) and \(h_j\) are \((E,\epsilon)\)-equivalent for any \(i,j\in \{1,2,3\}.\)
Proof. See Section 10. ◻
We can similarly apply Markov’s inequality in the frequency domain. The kinetic term on the torus penalizes high frequency components, therefore, a low-energy state should have low weights on the high frequency components. Hence, the low-frequency spaces are natural truncated domains for the torus and qubit Schrd̈oinger operators.
The discretization error arises in this step, where we upper bound it by a quantifiable notion of smoothness \(\log_\partial\), that we explain shortly in the next subsection. Intuitively, the following lemma states that we need more qubits per spatial dimension as \(L\) gets bigger, \(V_{\mathbb{T}^n}\) gets less smooth, and the target error gets smaller.
Lemma 6 (Equivalence of torus and qubit Hamiltonians). Let \(V_{\mathbb{T}^n}:\mathbb{T}^n \rightarrow \mathbb{R}\) be a smooth potential on the torus, and \[\begin{align} V_Q:=\sum_{y \in \mathcal{N}^n}|y\rangle V_{\mathbb{T}^n}\big(\frac{yL}{N}\big)\langle y |. \end{align}\] Suppose \[\begin{align} h_{\mathbb{T}^n} & : = -\Delta_{\mathbb{T}^n} +V_{\mathbb{T}^n}\qquad \qquad \qquad \qquad \text{on \;}\mathbb{T}^n, \\ h_{Q} & : = K_{Q} +V_{Q} \qquad \qquad \qquad \qquad \text{on \;}\log N\times n \;\text{qubits,} \end{align}\] where \(V(\mathbb{T}^n)\subset [0, V_{\max}]\) and \(\log_\partial (V_{\mathbb{T}^n}, \mathbb{T}^n, m) \le \log p(m)\) for some polynomial \(p\).
For \(\epsilon<1\), we have that \(h_{\mathbb{T}^n}\) and \(h_Q\) are \((E,\epsilon)\)-equivalent, if \[\begin{align} N\ge \Omega\left( \frac{L^2 (p(2n))^2 (E+V_{\max})^{3/2} }{\epsilon }\right). \end{align}\]
Proof. See Section 11. ◻
By considering the discrepancy between a Fourier base state on \({\mathbb{T}^n}\) and its discretization on \(\mathcal{N} ^n\), we naturally arrive at the following definition of a quantifiable smoothness.
Definition 7. Given a smooth function \(g:A \rightarrow \mathbb{R}\) on a subset \(A\) of \(\mathbb{R}^n\) or \(\mathbb{T}^n\), we define the smootheness factor of \(g\) to be \[\begin{align} \log_\partial (g,A,m) =\max_{l\in[m], j\in[n], x\in A} \frac{\left( \log |\partial^{l}_j g(x)| \right)^+}{l} , \end{align}\] where we use the notation \((a)^+ = \max(a,0)\) for \(a \in \mathbb{R}\), and assume that the support of \(\partial_j g\) is compact for all \(j\).
Intuitively, the smoothness factor indicates how “rough" a function is. In this paper, it is desired that \(V_{\mathbb{T}^n}\) has a logarithmic smoothness factor, so that there exists some polynomial \(p(m)\) such that \[\begin{align} \log_\partial (g,A,m) \le \log p(m). \end{align}\]
Having a logarithmic smoothness factor is closed under summation, scalar multiplication, and most importantly, composition.
Lemma 7. The following statements are true:
(Scalar multiplication and summation) Let \(f,f_k:\mathbb{T}^n \rightarrow\mathbb{R}\) be smooth for \(k\in[r]\), and \(c>0\). Then we have \[\begin{align} \log_\partial( cf, \mathbb{T}^n, m) &\; \le \; (\log c)^+ \;+ \; \log_\partial( f, \mathbb{T}^n, m) , \\ \log_\partial(\sum_{i\in[r]} f_i, \mathbb{T}^n, m) &\; \le \; \log r \;+ \; \sum_{i\in[r]} \log_\partial( f_i, \mathbb{T}^n, m) . \end{align}\]
(Composition) Let \(g:\mathbb{T}^n \rightarrow\mathbb{R}\) and \(f:g(\mathbb{T}^n) \rightarrow \mathbb{R}\) be smooth, where \(g(\mathbb{T}^n)\) is the range of \(g\). Then we have \[\begin{align} \log_\partial( f\circ g, \mathbb{T}^n, m) \; \le \; 2 \log m \;+ \; \log_\partial(f, g(\mathbb{T}^n), m) \;+ \; \log_\partial (g, \mathbb{T}^n, m). \end{align}\]
(The four smooth functions) We have \[\begin{align} \log_\partial (\text{Bump}, \mathbb{R},m) &\le \log O(m^4), \\ \log_\partial (\text{Sat}_{\alpha,\beta}, \mathbb{R},m), \log_\partial (\text{Cut}_{\alpha,\beta}, \mathbb{R},m),\log_\partial (\text{Bar}_{\beta}, \mathbb{R},m) &\le \log O(1/\beta + m^4), \end{align}\]
Our approach is to run FGA for a Schrödinger operator on \({\mathbb{R}^n}\) with the potential defined by a barrier function that penalizes going outside of \(\Omega\), namely \[\begin{align} V =\frac{3E}{\mu^6} \sum_{i\in[m]} \text{Bar}_\epsilon( a_i \cdot x -b_i ). \end{align}\]
We then construct a Schrödinger operator \(h\) on \({\mathbb{R}^n}\) with \(V\) as the potential. We show that \(\lambda_0(h)\) and \(\lambda_0(-\Delta^D_\Omega)\) are close to each other using the truncation lemmas. Therefore, it is enough to compute \(\lambda_0(h)\) through Algorithm 1. The parameters in the algorithm for \(V\) are polynomial in \(n,m,1/\epsilon,R\).
We first provide some facts on the eigenvalues of the Schrödinger operators and prove the two theorems.
The following lemma characterizes the \(i\)-th eigenvalue of \(h\).
Theorem 5 (Min-Max). [40] Let \(h\) be self-adjoint with a purely discrete spectrum. Let \(\lambda_0(h)\le \lambda_1(h)\le \cdots\) be the eigenvalues of \(h\).
Then, we have \[\begin{align} \lambda_n(h) \;=\; \min_{\psi_0,\dots,\psi_{n-1}} \max \{\;\langle \psi|h\psi\rangle \;|\;\|\psi\|=1, \psi\perp\psi_i \;\forall i\in\{0,\dots, n-1\}\}. \end{align}\]
As a corollary, we have a comparison of each eigenvalue of two Schrödinger operators, if their potentials are comparable [40].
Lemma 8 (Potential comparison). Let \(h_1,h_2 \in S^D_\Omega\) with potentials \(V_1,V_2:\Omega\rightarrow \mathbb{R}_{\ge0}\) such that \(V_1(x)\ge V_2(x)\) for all \(x\in\Omega\). Then, \(\lambda_k(h_1)\ge \lambda_k(h_2)\) for all \(k\in\mathbb{Z}_{\ge 0}\).
The following Lemma is a standard result in the spectral theory [41] that says, for any \(k\). \(k\)-th eigenvalue decreases as the domain increases. Intuitively, one could view the Dirichlet Schrödinger operators on \(\Omega\) as a Schrödinger operator on \({\mathbb{R}^n}\) with a potential \(V\) such that \(V(x)=\infty\) for \(x\notin \Omega\).
Lemma 9 (Domain monotonicity). Suppose for bounded \(\Omega_1\subseteq\Omega_2 \subset {\mathbb{R}^n}\), \[\begin{align} h &:=-\Delta_{{\mathbb{R}^n}} + V \in S_{{\mathbb{R}^n}}, \\ h_1 &:=-\Delta^D_{\Omega_1} +V\big|_{\Omega_1} \in S^D_{\Omega_1}, \\ h_2 &:=-\Delta^D_{\Omega_2} +V\big|_{\Omega_2} \in S^D_{\Omega_2}. \end{align}\] Then, for any \(k\in\mathbb{Z}_{\ge 0}\), we have \[\begin{align} \lambda_k(h)\le \lambda_k(h_1) \le \lambda_k(h_2). \end{align}\] If \(V =0\), then we have the special case of the domain monotonicity on Dirichlet Laplacian: \[\begin{align} \lambda_k(-\Delta^D_{\Omega_1}) \le \lambda_k(-\Delta^D_{\Omega_2}) . \end{align}\]
The following lemma is what we want to use in the analysis of the main theorem.
Lemma 10. Let \(h =t h_1 +(1-t)h_2\) for \(t\in[0,1]\) be a convex combination of two Dirichlet Schrödinger operators \(h_1,h_2\) on \(\Omega\subset\mathbb{R}^n\) with smooth potentials \(V_1,V_2:\Omega \rightarrow\mathbb{R}_{\ge0}\). Suppose \(V_2(x) =0\) for all \(x\in \Omega'\subset \Omega\), where \(\Omega'\) is compact.
Then, \[\begin{align} \lambda_1(h) \le \lambda_1(-\Delta^D_{\Omega'}) +\max_{x\in\Omega'} V_1(x). \end{align}\]
Proof. We have \(h = -\Delta^D_\Omega + tV_1+(1-t)V_2\). By the domain monotonicity (Lemma 9), we have \(\lambda_1(h)\le \lambda_1(h^D_{\Omega'})\), where \(h^D_{\Omega'}:=-\Delta^D_{\Omega'} +V_1\big|_{\Omega'}\) is a Dirichlet Schrödinger operator on \(\Omega'\) with the potential given as the restriction of \(V_1\).
By Lemma 8, we have \[\begin{align} \lambda_1(h^D_{\Omega'}) \le \lambda_1(-\Delta^D_{\Omega'} +\max_{x\in \Omega'} V_1(x)) =\lambda_1(-\Delta^D_{\Omega'}) +\max_{x\in \Omega'} V_1(x), \end{align}\] since \(V_1(y)\le \max_{x\in \Omega'}V(x)\) at all \(y\in\Omega'\). Therefore, we have \[\begin{align} \lambda_1(h)\le \lambda_1(h^D_{\Omega'}) \le \lambda_1(-\Delta^D_{\Omega'}) +\max_{x\in\Omega'} V_1(x). \end{align}\] ◻
(restate). Algorithm 1 computes the lowest eigenvalue of \(h = -\Delta + V\) within error \(\epsilon_0\) with probability \(\ge 2/3\) in polynomial time in \(n, 1/\epsilon_0,L,c\), if \(V\) satisfies the following conditions, where \(B^q_d:=\{x\in{\mathbb{R}^n}\;| \; \|x\|_q\le d\}\) and \(V_E^{-1}:= V^{-1}([0,E])\):
Self-adjointness and convexity. We have \(V(x)\rightarrow\infty\) as \(\|x\|\rightarrow\infty\) and \(V\) is convex on \(B^2_{r+1}\).
“Bowl-shaped” in \(B^\infty_{L/2}\). There exist \(a\le b\le c\), and \(1 \le r < L/2\) such that \[\begin{align} B^2_1\subset V^{-1}_{a } \subset V^{-1}_{b}\subset B^2_r \subset B^2_{r+1} \subset V^{-1}_{c}\subset V^{-1}_{c+1} \subset B^\infty_{L/2} \end{align}\] where
(i) \(b = \Theta(E_0^7/\sigma^6)\)
(ii) \(E_0=10(n(n+3)\pi^2 +a) = \Theta(n^2+a)\)
(iii) \(\sigma = \Theta(\epsilon_0/r^4 E_0^{1.5})\)
(iv) \(\epsilon_0<1 < r,E_0,c\).
Bounded high-order derivatives. The potential \(V\) is smooth and \[\begin{align} \max \left\{ { \frac{1}{l}\log|\partial^l_j V(x)| } \; \bigg| \;{l\in[m], j\in[n], x\in [-L/2,L/2]^n }\right\} \le \log p(m) \end{align}\] for some polynomial \(p\).
Access to \(V\). There exists a circuit that computes \(V(x)\) for all \(x \in B^\infty_{L/2}\) with a negligible error.
Proof of Theorem 1. We define functions \(V_{init},V_{final}, W_t:B^\infty_{L/2} \rightarrow\mathbb{R}_{\ge 0}\) for \(t \in [0,T]\) to be \[\begin{align} V_{init} &:= b \sum_{i\in[n]} \left( 1 - \text{Cut}_{\frac{1}{4\sqrt n},\frac{1}{4\sqrt n}}\!\left(\left| x_i \right| \right) \right) , \\ V_{final} &:= \text{Sat}_{c,1}\!\circ\! V(x), \\ W_t(x) &: = \left(1- \frac{t}{T}\right) V_{init}(x) + \frac{t}{T}V_{final} (x). \end{align}\]
Also define time-dependent Hamiltonians on \(B:=B^2_{r+1}, \mathbb{T}^n :=\mathbb{R}^n/L\mathbb{Z}^n\), and \(\mathcal{N}^n\) \[\begin{align} H_{B}(t) &: = -\Delta^D_B + W_{B,t} , \\ H_{\mathbb{T}^n}(t) &: = -\Delta_{\mathbb{T}^n} + W_{\mathbb{T}^n,t}, \\ H_{Q}(t) &:= \; \; K_Q + \sum_{y \in \mathcal{N}} |y\rangle W_t\big({yL}/{N}\big) \langle y |, \end{align}\] where \(W_{B,t}:= W_t|_B\) is the restriction of \(W_t\) on \(B\), and \(W_{\mathbb{T}^n,t}(x) := W_t(x')\) for \(x '\in[-L/2,L/2]^n\), such that \(x =x'\mod L\mathbb{Z}^n\). Note that \(H_Q(t)\) is defined consistently as in Algorithm 1.
We use the notation \(g:={3\pi^2}/{(2(r+1))^2}\) for the rest of the proof.
show \(\text{gap}(H_{B}(t))\ge g\) and \(2(\lambda_1(H_B(t))+1)\le E_0\) for all \(t\in[T]\). At an arbitrary time \(t \in [0,T]\), we know from Theorem 4 that \[\begin{align} \text{gap}(H_B(t)) \ge \; \frac{3\pi^2}{(2(r+1))^2}=g, \end{align}\] since the diameter of \(B\) is \(2(r+1).\)
To upper bound \(\lambda_1(H_B(t))\), we use the fact that \(B\) includes the box \(B':=B^\infty_{1/2\sqrt{n}}\), where the potential is at most \(a\). We define \[\begin{align} H_{B'}(t):= -\Delta^D_{B'} + W_{t}|_{B'} \end{align}\]
By the domain monotonicity (Lemma 9), we have \(\lambda_1(H_B(t))\le \lambda_1(H_{B'}(t))\). We have \(W_{t}|_{B'}(x) \le a\) for \(x \in B'\), by the condition \(B^2_1 \subset V^{-1}_{a}\). Therefore, \[\begin{align} \lambda_1(H_B (t)) &\le \lambda_1(H_{B'}(t)) \\ &\le \lambda_1(-\Delta^D_{B'}) + a \\ &= n(n+3) \pi^2 + a\\ &= E_0/10. \end{align}\]
show \(\text{gap}(H_{\mathbb{T}^n}(t))\ge 0.99g\), and \(2(\lambda_1(H_{\mathbb{T}^n}(t))+1)\le E_0\) for all \(t\in[T]\). We first show that \(H_B(t)\) and \(H_{{\mathbb{T}^n}}(t)\) are \((E_0,\sigma)\)-equivalent, and then, transfer the large gap from \(B\) to \(\mathbb{T}^n\) via Lemma 14.
For all \(x\) with \(\|x\|_2\ge r\), we have \(V_{init}(x) >b\) (due to the condition \(V^{-1}_b\subset B^2_r\)) and \[\begin{align} V_{final}(x) = \text{Sat}(V(x))\ge \min(V(x), c) > b, \end{align}\] where we denote \(\text{Sat}: = \text{Sat}_{c,1}\). Therefore, we have \(W_t(x)\ge \min(V_{init}(x),V_{final}(x))> b\). By Lemma 5, it follows that \(H_{B}(t)\) and \(H_{\mathbb{T}^n}(t)\) are \((E_0,\sigma)\)-equivalent.
In addition, by Lemma 13, we have that \(\lambda_1(H_{\mathbb{T}^n}(t)) \le E_0/10 + O(\epsilon_0)\). Therefore, we have \(2(\lambda_1(H_{\mathbb{T}^n}(t) +1)\le E_0\), and Lemma 14 is applicable.
Since \(\sigma = \Theta(\epsilon_0/r^4 E_0^{1.5}) = O(g^2/E_0^{1.5})\), by Lemma 14, we have \[\begin{align} \text{gap}(H_{\mathbb{T}^n}(t)) \ge \text{gap}(H_{B}(t)) - 0.01g \ge 0.99g. \label{eq:main95torus95gap} \end{align}\tag{6}\]
show that \(H_Q(t)\ge0.98g\) for all \(t\in[0,T]\). Again, we show that \(H_{{\mathbb{T}^n}}(t)\) and \(H_{Q}(t)\) are \((E_0,\sigma)\)-equivalent and then apply Lemma 14 to lower bound the gap.
To argue equivalence between \(H_{\mathbb{T}^n }(t)\) and \(H_{Q}(t)\) via Lemma 6, we need an upper bound on \(\log_\partial(W_{\mathbb{T}^n,t},\mathbb{T}^n, m)\). We first show that the function \(W_{\mathbb{T}^n,t}\) is smooth at all \(t\in[0,T]\), which amounts to showing smoothness of \(V_{final}\) and \(V_{init}\). For \(x\) with \(\|x\|_\infty < L/2\), \(\text{Sat}\circ V\) is smooth because it is a composition of two smooth functions. For \(x\) with \(\|x\|_\infty = L/2\), we have \(V(x) \ge c +1\) by the condition \(V^{-1}_{c+1}\subset B^\infty_{L/2}\); hence \(\text{Sat}(V(x))\) is constant, and all of its derivatives vanish. Also, \(V_{init}\) is smooth, because each \(\text{Cut}(|x_i|)\) is smooth on the torus, where we denote \(\text{Cut}:=\text{Cut}_{1/4\sqrt n,1/4\sqrt n}\). Therefore, \(W_{\mathbb{T}^n,t}\) is smooth on the torus at all \(t\).
By Condition 2, we have \(V_{final} (\mathbb{T}^n)\subset [ 0, c +1]\), and we have \(V_{init}(\mathbb{T}^n) \subset [0,n b]\) by definition. Therefore, \(W_{\mathbb{T}^n,t}(\mathbb{T}^n) \subset [0,O(c + nb )].\)
We bound the smoothness factor of \(W_{\mathbb{T}^n,t }\), using Lemma 20, \[\begin{align} \log_\partial( W_{\mathbb{T}^n,t},\mathbb{T}^n, m) &= \log_\partial( (1 -t/T) V_{init} + (t/T)V_{final}, \mathbb{T}^n,m) \\ & \le \log 2 + \log_{\partial}(V_{init}, \mathbb{T}^n, m) + \log_{\partial}(V_{final} , \mathbb{T}^n, m). \end{align}\] By Lemmas 20 and 22, we get \[\begin{align} \log_{\partial}(V_{init} , \mathbb{T}^n, m) & \;\le \; \log n + \log b +\log_\partial (\text{Cut}(|x_i|) , \mathbb{T}^n, m) \\ & \;= \; \log n + \log b +\log_\partial (\text{Cut}(|x_i|) , \mathbb{R} , m) \\ & \;= \; \log O (bn^{1.5} m^4). \end{align}\] Also, by Lemma 21, \[\begin{align} \log_{\partial}(V_{final} , \mathbb{T}^n, m) & \;\le \; 2 \log m + \log_\partial (\text{Sat}, [0, c+1], m ) + \log_\partial (V, [-L/2,L/2]^n, m ) \\ & \;\le \; \log (\Theta(m^2 \cdot m^4\cdot p(m))). \end{align}\] Combining the two smoothness factors gives \[\begin{align} \log_\partial( W_{\mathbb{T}^n,t},\mathbb{T}^n, m) \;\le \;\log(\Theta(bn^{1.5} m^{10} p(m)). \end{align}\] Therefore, by Lemma 6, encoding each spatial dimension with \[\begin{align} N \ge \Theta \left( \frac{L^2 \cdot b^2 n^{11.5} (p(2n))^2\cdot (V_{\max } + E_0^6/\sigma^7)^{1.5} }{\epsilon} \right) \end{align}\] dimensional quantum space gives that \(H_{\mathbb{T}^n}(t)\) and \(H_Q(t)\) are \((E_0,\sigma)\)-equivalent.
We have established that \(2(\lambda_1(H_{\mathbb{T}^n}(t))+1)\le E_0\) in Step 2. Therefore, by Lemma 14, it follows that \[\begin{align} \text{gap}(H_{Q}(t))\ge 0.98g. \end{align}\]
show \(|\lambda_0(H_{\mathbb{R}^n}(T))-\lambda_0(H_{Q}(T))|\le 0.02\epsilon_0\). By Lemma 5, the condition \(V^{-1}_b \subset B^2_r\) gives that \(H_{\mathbb{R}^n}(T)\) and \(H_{\mathbb{T}^n}(T)\) are \((E_0,\sigma)\)-equivalent. Lemma 12 implies that \[\begin{align} |\lambda_0(H_{\mathbb{R}^n}(T) ) -\lambda_0(H_{\mathbb{T}^n}(T) )| \le \sigma \le 0.01\epsilon_0. \end{align}\]
By Step 3, \(H_{\mathbb{T}^n }(T)\) and \(H_Q(T)\) are \(\sigma\)-equivalent under our choice of parameters. Lemma 12 implies that \[\begin{align} |\lambda_0(H_{\mathbb{T}^n}(T) ) -\lambda_0(H_{ Q}(T) )| \le 2\sigma \le 0.01\epsilon_0. \end{align}\] Therefore, by the triangle inequality, we have \[\begin{align} |\lambda_0(H_{\mathbb{R}^n} (T)) - \lambda_0(H_Q(T))| \le 0.02 \epsilon_0 \end{align}\]
finish the proof.
In the first step of the algorithm, each register measures the ground state with probability \(1- O(1/n)\). Therefore \(|\Psi_{init}\rangle\) is correctly prepared with probability \(\Omega(1)\).
We showed in Step 3 that the spectral gap of \(H_Q(t)\) is at least \(0.98\cdot3\pi^2/(2(r+1))^2\). By Adiabatic Theorem 3, after evolving \(|\Psi_{init}\rangle\) under the time-dependent Hamiltonian \(H_Q(t)\) for time \[T = O\!\left( \frac{\|H_Q(0) - H_Q(T)\|^{2}}{0.01 \cdot (\min_{t \in [0,T]} \text{gap}(H_Q(t)))^3} \right) \le O\! \left( {r^6 (N^2 + c + nb)^2 } \right),\] the final state \(|\Psi_{final}\rangle\) is \(0.01\)-close to the ground state of \(H_Q(T)\). The evolution time \(T\) is polynomial in the complexity parameters and so is the norm \(\|H_Q(t)\|\). Therefore, any off-the-shelf Hamiltonian simulation algorithm, such as [30], simulates \(H_Q\) in polynomial time.
The Hamiltonian simulation results in a state that is 0.02-close to the ground state of \(H_Q(T)\), which we measure in the final step of the Algorithm with probability \(\ge 96\%\). The energy we measure deviates from \(\lambda_0(h)\) by at most \(0.02\epsilon_0\) (Step 4 of this proof) plus the error from the phase estimation, which can be made negligible. ◻
Proof of Theorem 2. We specify the parameters \(a,b,c,E, \sigma,r, L\) in Theorem 1, and show that they are polynomials in \(n,m, R\). First, we set \[\begin{align} a = 0, \quad r = 2R,\quad L= 3R, \quad E = \Theta(n^2) , \qquad \sigma = \Theta(\epsilon_0/r^4E_0^{1.5}). \end{align}\]
Condition 2 of Theorem 1 enforces that \[\begin{align} b = \Theta(E^7/\sigma^6) = \Theta(E^{16} R^{24}/ \epsilon_0^6) = \Theta(n^{32}R^{24}/\epsilon_0^6) \end{align}\] and at the same time, \(V^{-1}_b \subset B^2_r\). We satisfy these conditions by choosing a sufficiently small \(\mu\) in [eq:DL-potential].
Note that, for any \(x\in{\mathbb{R}^n}\) such that \(\|x\|=R\), there exists \(j\in[m]\) such that \(a_j\cdot x - b_j \ge 0\). Therefore, for \(x\) such that \(\|x\| = 2R\), we have \[\begin{align} V(x) \ge \frac{E}{\mu^6}\text{Bar}_\epsilon(R) = \Theta\left(\frac{n^2 R}{\mu^6}\right), \end{align}\] and for \[\begin{align} \label{eq:DL95sigma95first} \mu = O \left(\frac{\epsilon_0}{n^{5}R^{23/6} }\right), \end{align}\tag{7}\] we have \(V^{-1}_b \subset B^2_r\).
On the other hand for \(\sigma,\) it should be chosen that \(|\lambda_0(-\Delta^E_\Omega) - \lambda_0(h)| \le O(\epsilon_0)\). By 48 , we define the expanded region \(\Omega\) with \(\epsilon = O(\epsilon_0/n^2)\) to obtain \[\begin{align} |\lambda_0(h^D) - \lambda_0(-\Delta^D_\Omega)| = \epsilon_0/3. \end{align}\] By Lemma 12, we need \((E,O(\epsilon_0))\)-equivalence between \(h\) and \(h^D\) to obtain \[\begin{align} |\lambda_0(h) - \lambda_0(h^D)| =\epsilon_0/3, \end{align}\] where \(E=\Theta(\lambda_0(h)) \le \Theta(n^2)\). By Lemma 24, setting \[\begin{align} \label{eq:DL95sigma95second} \mu = O\left(\frac{\epsilon_0}{E (mn^2)^{1/3}}\right) = O\left(\frac{\epsilon_0}{( m n^8)^{1/3}}\right) \end{align}\tag{8}\] gives \((E_0,O(\epsilon_0))\)-equivalence between \(h\) and \(h^D\), finally yielding \[\begin{align} |\lambda_0(-\Delta^D_\Omega) - \lambda_0(h) | = \epsilon_0/3. \end{align}\] Our algorithm estimates \(\lambda_0(-\Delta^D_\Omega)\) up to \(\epsilon_0/3\) when it estimates \(\lambda_0(h)\) up to \(\epsilon_0/3\).
To satisfy both 7 and 8 , we set \[\begin{align} \mu = O\left(\frac{\epsilon_0}{n^5 m^{1/3} R^{23/6}}\right) . \end{align}\]
We are only left with \(c\), whose upper bound we find by \[\begin{align} c \le \max_{x:\|x\|_\infty \le L} V(x) \le \max_{x:\|x\|_2 \le \sqrt n L} V(x) \le \sum_{j\in[m]} \frac{E}{\mu^6} \sqrt n L = O\left( \frac{ELm\sqrt n}{\mu^6 }\right) \end{align}\] Therefore, each parameter in Theorem 1 is polynomial in \(n,m,R, \epsilon_0\).
Also, since \(V\) is a sum of \(\text{Bar}_\epsilon\), which has a logarithmic smoothness factor by Lemma 22, the algorithm requires only polynomial time and qubits in the parameters. ◻
An interesting notion of noncommutative convexity arises in \(h\). The functional \(|\psi\rangle \rightarrow\langle \psi|h\psi\rangle\) is convex under the unitaries of translation \(T_a\) and modulation \(M_b\) \[T_a(\psi)(x) := \psi(x+a), \qquad M_b(\psi)(x) := e^{i b\cdot x}\,\psi(x),\] and more generally, under the Weyl operators \[\begin{align} W(a,b) : = e^{\frac{i}{2}a \cdot b} M_b T_a \end{align}\] that are noncommuting. This notion is analogous to the fact that a classical convex function is convex under translation of a fixed point.
Geodesic convex optimization [42]–[45] also studies convex objectives in noncommutative directions, but our case differs in that the orbit \(\{W(a,b)\psi\}_{(a,b)\in\mathbb{R}^2 }\) of any \(\psi\in L^2({\mathbb{R}^n})\) is just a small subset of the feasible set; the envelope \(x\rightarrow|\psi(x)|\) is preserved up to translation. Another notion of noncommutative convex optimization appears in Noncommutative Polynomial Optimization (NcPO) [46]–[49]. When the potential \(V\) is a polynomial, our formulation can be viewed as an infinite-dimensional instantiation of NcPO.
We formalize the idea mentioned above that \(h[\cdot]\) is convex under Weyl operators.
Definition 8 (Weyl convexity). A functional \(f:L^2({\mathbb{R}^n}) \rightarrow \mathbb{R}\) is Weyl-convex if for any fixed \(\psi \in \mathcal{D}(f)\), the function \[\begin{align} (a,b)\rightarrow f(W_{a,b}\psi) \end{align}\] is convex over \((a,b) \in \mathbb{R}^n \times \mathbb{R}^n\).
Lemma 11. Given a Schrödinger operator \(h = -\Delta + V\) with convex \(V:{\mathbb{R}^n}\rightarrow\mathbb{R}\), the energy functional \(h[\psi]\) is Weyl-convex.
Proof. Let \(p:=-i\nabla\), so \(-\Delta=p^2\). The conjugations are well-known to be \[T_a^\dagger p\,T_a=p,\quad T_a^\dagger V(x)\,T_a=V(x-a),\qquad M_b^\dagger p\,M_b=p+b,\quad M_b^\dagger V(x)\,M_b=V(x).\] Hence \[W_{a,b}^\dagger\,p^2\,W_{a,b}=(p+b)^2.\] Since \(W_{a,b}\) is unitary, \(\|W_{a,b}\psi\|=\|\psi\|\), and therefore \[h[W_{a,b} \psi]=\frac{\langle\psi|((p+b)^2 \psi\rangle + \langle T_a \psi | V |T_a \psi\rangle}{\langle\psi|\psi\rangle}.\] Set \[\bar p:=\left( \frac{\langle\psi | p_j \psi\rangle}{\langle\psi| \psi\rangle}\right)_{j\in[n]} \in\mathbb{R}^n,\qquad K:=-\frac{\langle\psi|\Delta\psi\rangle}{\langle\psi| \psi\rangle}<\infty.\] Expanding gives \[h[W(a,b)\psi]=\underbrace{\big(\|b\|^2+2\,b\cdot\bar p+K\big)}_{=:q(b)} \;+\;\underbrace{\int_{\mathbb{R}^n} V(x-a)\,|\psi(x)|^2\;\text{d} x}_{=:G(a)}.\] The function \(q(b)\) has Hessian \(2I_n\succeq 0\), hence is (strictly) convex in \(b\). For each fixed \(x\), the map \(a\mapsto V(x-a)\) is convex; integrating convex functions preserves convexity, so \(G(a)\) is convex in \(a\). Therefore \(h[W(a,b)\psi]=q(b)+G(a)\) is jointly convex in \((a,b)\). ◻
The results on the complexity of the Schrödinger operators are sparse. As far as we are aware, Zheng et al. [50] is the only work on the topic. They prove that the Schrödinger operator is StoqMA-hard for general smooth potential on a bounded Dirichlet domain, where simulation of Schrödinger operator is BQP-hard. Is Schrödinger operators with convex potential BQP-complete? What is the complexity of Schrödinger operators with the fermionic symmetry? What is the complexity of finding the ground energy of a molecule? What is the complexity of the convex drum problem?
Let us denote \(\widehat x:=(\widehat x_1,\dots,\widehat x_n)\), where \(\widehat x_i\) is the multiplicative operator associated with the coordinate function \(x_i\) in \({\mathbb{R}^n}\). Similarly, let us denote \(\widehat p:=(\widehat p_1,\dots,\widehat p_n)\), where \(\widehat p_i:=U_{\mathcal{F}}^\dagger \widehat x_i U_{\mathcal{F}}\) for the Fourier transformation unitary \(U_{\mathcal{F}}\) on \(L^2({\mathbb{R}^n})\).
In this paper, we gave a rudimentary algorithm for computing \[\begin{align} \min_{\|\psi\| =1} \langle\psi| \; \|\widehat p\|^2 + C(\widehat x) \;|\psi\rangle, \end{align}\] for a convex \(C\). A natural extension is to find an algorithm for \[\begin{align} \min_{\|\psi\| =1} \langle\psi| \; C_1(\widehat p) + C_2(\widehat x) \;|\psi\rangle, \end{align}\] where \(C_1,C_2\) are convex. A step further is to consider \[\begin{align} \min_{\|\psi\| =1} \langle\psi| \; C_3(\widehat p,\widehat x) \;|\psi\rangle, \end{align}\] for a Weyl-convex \(C_3\).
A more ambitious goal is to construct a quantum theory that parallels classical mathematical optimization. For instance, we can aim to solve constraint problems \[\begin{align} \min_{\|\psi\| =1} & \langle\psi| \; C(\widehat p,\widehat x) \;|\psi\rangle, \\ \text{subject to } &\langle\psi| \; D_i(\widehat p,\widehat x) \;|\psi\rangle \le a_i \quad \forall i\in[m]. \end{align}\] that hopefully are as useful and versatile as classical optimization programming, such as linear and semidefinite programming.
The optimization problems we solve in this paper are instantiations of the calculus of variations [51]. We propose the calculus of variations as a venue for exponential quantum speedups.
A typical problem is \[\begin{align} \text{minimize}\qquad J[u]&:=\int_{x\in\Omega} L(x,u(x),\nabla u(x))\;\text{d} x \\ \text{subject to}\qquad 0 & \; =\int_{x\in\Omega} C_i (x,u(x),\nabla u(x))\;\text{d} x \qquad \forall i \in[m], \\ u&: \;\Omega \subset {\mathbb{R}^n}\rightarrow \mathbb{R}, \end{align}\] where \(L,C_i: {\mathbb{R}^n}\times \mathbb{R}\times {\mathbb{R}^n}\).
The problem is a good candidate for an exponential quantum advantage, since classically solving it, for example via finite element method, requires exponential time in \(n\). A quantum computer has an advantage in that the exponentially large data \(u\) can be efficiently stored and manipulated in a number of qubits increasing polynomially as the dimension \(n\) grows.
The idea refers to the fact that two classically minimizable functionals, namely the kinetic and momentum terms, are added to give a quantum objective that is efficiently optimized by a quantum algorithm.
Does this phenomenon appear in a more general setting? Can we add more than two classical functionals to get an optimizable quantum Hamiltonian? Can we add two optimizable qubit Hamiltonians in the \(X\) and \(Z\) basis to get an optimizable Hamiltonian? Can we add classically approximateable Hamiltonians to get a quantum approximable Hamiltonian? Is there a unified framework to solve such problems?
For a concrete example, the Quantum Max Cut [52] Hamiltonian \(\sum_{ij\in E} I-X_iX_j-Y_iY_j-Z_iZ_j\) is the summation of the Max Cut Hamiltonians in the \(X,Y,Z\) basis, each of which can be optimally approximated by the Goemans-Williamson algorithm [53].
We observe that efficient quantum optimization algorithms requiring a quantum verifier roughly falls under two categories: dissipation [19], [20], and adiabatic evolution [22].
We conjecture there is a unified framework encompassing both. Our algorithm solves the Schrödinger problem via adiabatic evolution, but intuitively, a dissipation-based algorithm makes more sense for the problem; it is more natural to imagine a particle in nature minimizes its kinetic energy via dissipation, rather than by some adiabatic process. Furthermore, [17] devices a time-dependent Hamiltonian in order to simulate quantum friction, showing that friction and adiabatic evolution are interchangeable.
The author is supported by a KIAS Individual Grant CG093802 at Korea Institute for Advanced Study. The author appreciates helpful discussions with Hyukjoon Kwon and Isaac Kim.
In addition to proving Lemma 4, we prove Lemma 14, which relates the spectral gaps of two Hamiltonians.
Lemma 12 (\(\lambda_0\) approximation). Let \(h_a,h_b\) be self-adjoint linear operators with purely discrete spectrum and \(\lambda_0(h_a),\lambda_0(h_b)\ge 0\). Suppose \(h_a\) and \(h_b\) are \((E,\epsilon)\)-equivalent. There exists a constant \(c>0\) such that if \(\epsilon \in[0,c]\) and \[\begin{align} \lambda_0(h_a) + 1 \leq E, \end{align}\] then we have \[|\lambda_0(h_a) - \lambda_0(h_b)| \leq 2\epsilon.\]
Proof. Let \(\widetilde{\mathcal{D}}_a, \widetilde{\mathcal{D}}_b\) be \((E,\epsilon)\)-truncated domains of \(h_a,h_b\) that are isomorphic to each other via a unitary \(U:\widetilde{\mathcal{D}}_a\rightarrow\widetilde{\mathcal{D}}_b\) such that \(|h_a[\widetilde{\psi}]-h_b[U\widetilde{\psi}]|\le \epsilon\) for all \(\psi \in\widetilde{D}_a\). Suppose \(\lambda_0(h_a) = h_a[\psi_0]\) for a normalized \(\psi_0 \in \mathcal{D}(h_a)\). Because \(h_a[\psi_0] < E\), there is an \(\epsilon\)-truncation of \({\psi_0}\), denoted \(\widetilde{\psi}_0\in\widetilde{\mathcal{D}}_a\). By the definition of \((E,\epsilon)\)-equivalence, we have \[\lambda_0(h_b) \le h_b[U\widetilde{\psi}_0]\le h_a[\widetilde{\psi}_0] + \epsilon\le h_a[\psi_0]+2\epsilon = \lambda_0(h_a) +2\epsilon,\] and therefore \[\begin{align} \lambda_0(h_b)- \lambda_0(h_a) \le 2\epsilon. \label{eq:ground95energy95thm} \end{align}\tag{9}\]
Similarly, let \(\mu_0\in\mathcal{D}_b\) be a vector such that \(h_b[\mu_0] =\lambda_0(h_b).\) The fact that \(\lambda_0(h_b) \le \lambda_0(h_a)+2\epsilon \le E\), for \(\epsilon \in[0,1/2]\), allows an \(\epsilon\)-truncation of \(\mu_0\), which we denote \(\widetilde{\mu}_0\in\widetilde{\mathcal{D}}_b\). By the definition of \((E,\epsilon)\)-equivalence, we have \[\begin{align} \lambda_0(h_a)\le h_a[U^{-1}\widetilde{\mu}_0]\le h_b[\widetilde{\mu}_0] + \epsilon\le h_b[\mu_0]+2\epsilon = \lambda_0(h_b) +2\epsilon, \end{align}\] and therefore \[\begin{align} \lambda_0(h_a)- \lambda_0(h_b) \le 2\epsilon. \end{align}\] Together with 9 , we prove the lemma. ◻
Lemma 13 (\(\lambda_1\) approximation). Let \(h_a,h_b\ge 0\) be a self-adjoint linear operator with purely discrete spectrum, such that \(\lambda_1(h_a)>\lambda_0(h_a) \ge0\). Suppose \(h_a\) and \(h_b\) are \((E,\sigma)\)-equivalent, and also satisfy the conditions \[\begin{align} &g:= \text{gap}(h_a) \;> \;0, \nonumber \\ &2({\lambda_1(h_a)} +1) \; \leq \; E,\quad \label{eq:thm95gap95lambda195condi} \\ &\epsilon, E^{-1} \;< \;1 . \nonumber \end{align}\tag{10}\]
Suppose \(c>0\) is a sufficiently small universal constant. If \(\epsilon \in [0,c]\) and \[\begin{align} \sigma = O \left(\frac{\epsilon g}{E^{1.5}} \right), \end{align}\] then \[\begin{align} |\lambda_1(h_a) - \lambda_1(h_b)| \leq \epsilon . \label{eq:thm95gap95ineq} \end{align}\tag{11}\]
Proof. Let \(\widetilde{\mathcal{D}}_a, \widetilde{\mathcal{D}}_b\) be \((E,\epsilon)\)-truncated domains of \(h_a,h_b\) that are isomorphic to each other via a unitary \(U:\widetilde{\mathcal{D}}_a\rightarrow\widetilde{\mathcal{D}}_b\)
Let \(\mu_0\in \mathcal{D}(h_b)\) be a normalized vector such that \(h_b[\mu_0] = \lambda_0(h_b) .\) From Lemma 12, we know that \(h_b[\mu_0]\le \lambda_0(h_a) + 2 \sigma \le E.\) Let \(\widetilde{\mu}_0\in \widetilde{\mathcal{D}}_b\) be a \(\sigma\)-truncation of \(\mu_0\), and let \(\widetilde{\alpha}_0:=U^{-1}\widetilde{\mu}_0 \in \widetilde{\mathcal{D}}_a\). Then \(h_a[\widetilde{\alpha}_0]\) is a good approximation of \(\lambda_0(h_a)\): \[\begin{align} h_a[\widetilde{\alpha}_0] \le h_b[\widetilde{\mu}_0] +\sigma \le h_b[{\mu}_0] +2\sigma \le \lambda_0(h_a) + 4\sigma. \label{eq:alpha95energy} \end{align}\tag{12}\]
Let \(\psi_0\in \mathcal{D}(h_a)\) be a normalized vector such that \(h_a[\psi_0] = \lambda_0(h_a)\). We show that \(\|\widetilde{ \alpha}_0 - \psi_0\|\) is small. Since \(h_a\) has a purely discrete spectrum, its eigenvectors form a complete set of basis. Therefore, we can write \[\widetilde{\alpha}_0 = \sqrt{1- t^2} \psi_0 + t {\psi}_0^\perp\] for some \(t\ge 0\) and a normalized vector \({\psi}_0^\perp \perp \psi_0\) in \(\mathcal{D}(h_a)\). We are assuming that \(\lambda_0(h_a)\) has a nonzero gap \(g\), so we have \(h_a[\psi_0^\perp] \ge \lambda_1(h_a) = \lambda_0(h_a) + g.\) Since \(\langle \psi_0|h_a|\psi_0^\perp\rangle = 0\), we have \[\begin{align} h_a[\widetilde{\alpha}_0] = (1- t^2) \lambda_0(h_a) + t^2 h_a[{\psi}_0^\perp] \ge (1- t^2) \lambda_0(h_a) + t^2(\lambda_0(h_a) + g) = \lambda_0(h_a) + t^2 g. \label{eq:alpha95energy95lowerbound} \end{align}\tag{13}\] Combining 12 with 13 , we get \[t^2 \le 4\sigma/g.\] Because \(1 - \sqrt{1-t^2} = t^2/({1+\sqrt{1-t^2}})\le t^2\), we have \[\begin{align} \|\widetilde{\alpha}_0 - \psi_0\| = \sqrt{(1 - \sqrt{1-t^2})^2 +t^2} \le \sqrt{ t^4 + t^2} \le \sqrt{2}t \le 2\sqrt {2 \sigma/g}. \label{eq:gap95alpha95distance} \end{align}\tag{14}\]
Let \(\psi_1\in \mathcal{D}(h_a)\) be a normalized eigenvector such that \(h_a[\psi_1] = \lambda_1(h_a)\). Since \(\lambda_1(h_a) < E\), there exists an \(\sigma\)-truncation of \(\psi_1\), which we denote \(\widetilde{\psi}_1 \in \widetilde{\mathcal{D}}_a\). We know that \(\psi_1\) is orthogonal to \(\psi_0\). We now show that \(\widetilde{\beta}_1:= U\widetilde{\psi}_1\) is approximately orthogonal to \(\mu_0\), which in turn will approximately lower bound \(\lambda_1(h_b)\) by \(h_1[\widetilde{\beta}_1]\): \[\begin{align} |\langle\widetilde{\beta}_1|\mu_0\rangle| &\;=\; |\langle\widetilde{\beta}_1|((|\mu_0\rangle -|\widetilde{\mu}_0\rangle) +| \widetilde{\mu}_0\rangle)| \nonumber \\ &\;\le\; |\langle\widetilde{\beta}_1| \widetilde{\mu}_0\rangle | + \|\langle\widetilde{\beta}_1| \| \cdot \| |\mu_0\rangle -|\widetilde{\mu}_0\rangle \| \nonumber \\ &\;\le\; |\langle\widetilde{\beta}_1|\widetilde{ \mu}_0\rangle| + \sigma \tag{15}\\ &\;=\; |\langle U^{-1} \widetilde{\beta}_1|U^{-1}\widetilde{ \mu}_0 \rangle| + \sigma \nonumber \\ &\;=\; |\langle \widetilde{\psi}_1|\widetilde{ \alpha}_0 \rangle| + \sigma \nonumber \\ &\;\le\; |\langle {\psi}_1|\widetilde{ \alpha}_0 \rangle| + 2\sigma \tag{16}\\ &\;\le\; |\langle {\psi}_1|\psi_0 \rangle| + 2\sigma + 2\sqrt{2\sigma/g} \tag{17} \\ &\; = \; 2\sigma + 2\sqrt{2\sigma/g} \tag{18} \end{align}\] Here, 15 and 16 follow from Cauchy-Schwarz and the norm condition of \(\sigma\)-truncation, and 17 follows from 14 . Since \(\epsilon, E^{-1}, g/E<1\), we have \[\begin{align} \label{eq:gaps95trunc95sigma95choice} \sigma = O(\epsilon g/E^{1.5}) =O(e \cdot E^{-0.5} \cdot g/E) \le O(\min(\epsilon,\sqrt{\epsilon/E}, \epsilon g/E)). \end{align}\tag{19}\] It follows that 18 is less than 1.
Therefore, we can write \(\widetilde{\beta}_1 = s\mu_0 + \sqrt{1- |s|^2} {\mu}_0^\perp\) for some \(s\in \mathbb{C}\) such that \(|s|\le 2\sigma + 2\sqrt{2\sigma/g}\) and a normalized vector \({\mu}_0^\perp\in \mathcal{D}(h_b)\) such that \(\mu_0^\perp \perp \mu_0\). Then we have \[\begin{align} (1-|s|^2)h_b[\mu_0^\perp] \; \le \; |s|^2 \lambda_0(h_b) + (1 - |s|^2)h_b[\mu_0^\perp] &\;=\; h_b[\widetilde{\beta}_1]\\ &\;=\; h_b[U\widetilde{\psi}_1] \\ &\;\le\; h_a[\widetilde{\psi}_1] + \sigma \\ &\;\le\; h_a[\psi_1] + 2\sigma \\ &\;=\; \lambda_1(h_a) + 2\sigma, \end{align}\] and therefore \[\begin{align} \lambda_1(h_b) \le h_b[\mu_0^\perp] &\;\le\; \frac{\lambda_1(h_a) +2\sigma}{1-|s|^2} \nonumber \\ &\;\le\; \frac{\lambda_1(h_a) + 2\sigma}{1- (2\sigma + 2\sqrt{2\sigma/g})^2} \tag{20} \\ & \;\le \;2{\lambda_1(h_a)+1} \nonumber\\ &\;\le\; E , \tag{21} \end{align}\] by 19 the condition 10 . From Inequalities 19 and 20 , we have \[\begin{align} \lambda_1(h_b) - \lambda_1(h_a) &\;\le \; \frac{(2\sigma +2 \sqrt{2\sigma/g})^2\lambda_1(h_a) + 2\sigma}{1-(2\sigma +2 \sqrt{2\epsilon/g})^2} \nonumber\\ &\;\le \;(2\sigma +2 \sqrt{2\sigma/g})^2E + \frac{ 2\sigma}{1-(2\sigma +2 \sqrt{2\sigma/g})^2} \nonumber\\ &\;\le \;\epsilon . \label{eq:thm95gap95lambda1hb-lambda1ha} \end{align}\tag{22}\]
Now we upper bound \(\lambda_1(h_a) - \lambda_1(h_b)\) in a similar manner. Let \(\mu_1\in \mathcal{D}(h_b)\) be a normalized eigenvector of \(h_b\) with the second lowest eigenvalue \(\lambda_1(h_b).\) From 21 , we know that there exists an \(\sigma\)-truncation of \(\mu_1\), denoted \(\widetilde{\mu}_1 \in \widetilde{\mathcal{D}}_a\). Let \[\begin{align} \widetilde{\alpha}_1 \;: =\; U^{-1}\widetilde{\mu}_1 \;= \; t'\psi_0 + \sqrt{1-t'^2}\psi_0'^\perp, \end{align}\] for some \(t'\) such that \(|t'|\in [0,1]\) and \(\psi_0'^\perp \in \mathcal{D}(h_a)\) such that \({\psi_0'} ^\perp \perp \psi_0\). By 14 and the norm condition of \(\epsilon\)-truncation, \[\begin{align} |t'| &\;=\; |\langle\psi_0| \widetilde{\alpha}_1\rangle| \nonumber\\ &\;\le\; |\langle\widetilde{\alpha}_0| \widetilde{\alpha}_1\rangle| + 2\sqrt{2\sigma/g} \nonumber\\ &\;=\; |\langle\widetilde{\mu}_0| \widetilde{\mu}_1\rangle| + 2\sqrt{2\sigma/g} \nonumber\\ &\;\le\; |\langle{\mu}_0| {\mu}_1\rangle| +2\sigma + 2\sqrt{2\sigma/g} \nonumber\\ &\;=\; 2\sigma + 2\sqrt{2\sigma/g}. \nonumber \end{align}\] By similar arguments as before, we get \[\begin{align} (1 - |t'|^2)\lambda_1(h_a) \le |t'|^2\lambda_0(h_a) + (1 - |t'|^2)\lambda_1(h_a) = h_a[\widetilde{\alpha}_1] \le h_b[\widetilde{\mu}_1] + \sigma \le \lambda_1(h_b) + 2\sigma \end{align}\] and, \[\begin{align} \frac{\lambda_1(h_b) + 2\sigma}{ 1- (2\sigma + 2\sqrt{2\sigma/g})^2} \ge \lambda_1(h_a). \end{align}\] By similar calculation that led to 22 , we get \[\begin{align} \lambda_1(h_a) - \lambda_1(h_b) \le \epsilon, \nonumber \end{align}\] and combining with 22 gives \[\begin{align} |\lambda_1(h_a) - \lambda_1(h_b) | &\;\le \;\epsilon. \nonumber \end{align}\] ◻
Lemma 14 (gap approximation). Let \(h_a,h_b\ge 0\) be a self-adjoint linear operator with purely discrete spectrum. Suppose \(h_a\) and \(h_b\) are \((E,\sigma)\)-equivalent and the following conditions hold: \[\begin{align} &\text{gap}(h_a) \; \ge g \;> 0, \nonumber \\ &2({\lambda_1(h_a)} +1) \; \leq \; E,\quad \\ &\epsilon, E^{-1} \;< \;1 . \nonumber \end{align}\]
Then there exist a universal constant \(c>0\) such that, if \(\epsilon \in \cap[0,c]\) and \[\begin{align} \sigma \le \Theta \left(\frac{g^2}{E^{1.5}} \right) \end{align}\] with a sufficiently small constant factor, then \[\begin{align} |\text{gap}(h_a) - \text{gap}(h_b) | \le 0.01g . \end{align}\]
Proof. We have \(\sigma = O(g)\), since \(\sigma \le \Theta(g\cdot g/E\cdot 1/E^{0.5})\) and \(g\le \lambda_1(h_a) < E\). By Lemma 12, \(\sigma\le \Theta(g)\) implies that \[\begin{align} |\lambda_0(h_a) - \lambda_0(h_b)| \le 0.01g/2. \end{align}\] By Lemma 13, \(\sigma\le \Theta(g\cdot g/E^{1.5})\) implies that \[\begin{align} |\lambda_1(h_a) - \lambda_1(h_b)| \le 0.01g/2. \end{align}\] Therefore, the triangle inequality gives \[\begin{align} |\text{gap}(h_a) - \text{gap}(h_b) | \le |\lambda_0(h_a) - \lambda_0(h_b)| + |\lambda_1(h_a) - \lambda_1(h_b)| \le 0.01g. \end{align}\] ◻
Proof of Lemma 5. Let us denote \(L^2(B\subset \Omega_i)\) to be the space of \(L^2\) functions on \(\Omega_i\in\{\mathbb{R}^n, B, \mathbb{T}^n\}\) that is supported on \(B\subset \Omega_i\). By Lemma 16, \(L^2(B\subset \Omega_i)\) is an \((E,\epsilon)\)-truncated domain for \(h_i\) for all \(i\in\{1,2,3\}\). The two \((E,\epsilon)\)-truncated spaces \(L^2(B\subset \Omega_i)\) and \(L^2(B\subset \Omega_j)\) are identified by the unitary \(U: L^2(B\subset \Omega_i)\rightarrow L^2(B\subset \Omega_j)\), such that \[\begin{align} U\psi(x) =\begin{cases} \psi(x) \qquad &\|x\|_2\le r+1 \\ 0 \qquad & \|x\|_2> r+1. \end{cases} \end{align}\]
We have \(h_i[\psi] = h_j[U\psi]\) because \(V_i(x) = V_j(x)\) when \(x \in B.\) Therefore, \(h_i\) and \(h_j\) are \(\epsilon\)-equivalent. ◻
Lemma 15 (Markov’s in position). Let \(h :=-\Delta + V \in S_{\mathbb{R}^n}\cup S^D_{B} \cup S_{\mathbb{T}^n}\) be a Schrödinger operator on the domain \(\Omega\in\{\mathbb{R}^n, B, \mathbb{T}^n\}\), where \(B:=B^2_{r+1}\) and \(L>2(r+1)\). Let \(\psi \in \mathcal{D}(h)\) be a normalized state with \(h[\psi] \le E\). Let \(f:\Omega \rightarrow \mathbb{[}0,1]\) be any measurable function such that \[\begin{align} \begin{cases} f(x) \; = \; 1 & \text{if }\;\;\|x\|_2 \le r, \\ f(x) \;\in \;[0,1] &\text{if }\;\;\| x\|_2 \in [r, r+1],\\ f(x) \;=0 &\text{if } \;\; \|x\|_2\ge r+1. \end{cases} \end{align}\] For a parameter \(\sigma \in[0,1/2]\), suppose \(V(x)\ge E/\sigma^2\) for all \(x\) with \(\|x\|_2 \ge r\). Then, we have \[\begin{align} \left\| \psi - f\psi \right\| \; &\le \; \sigma. \label{eq:401-f41psi} \end{align}\tag{23}\]
Proof. The proof is by Markov’s inequality. In each case of \(\Omega \in \{\mathbb{R}^n, B, \mathbb{T}^n\}\), we have \[\begin{align} \langle \psi| (-\Delta)|\psi\rangle \;= \; -\int_\Omega \overline{\psi}\nabla \cdot \nabla \psi \;\text{d} x \;= \;\int_\Omega |\nabla \psi|^2 \;\text{d} x \;\ge \; 0, \end{align}\] by integration by parts. Since \(\psi\) is normalized, we have \[\begin{align} E \;= \;h[\psi] \; = \int_\Omega \overline{\psi}(x) (-\Delta + V(x) ) \psi(x) \;\text{d} x \;\ge \int |\psi(x) |^2 V(x) \;\text{d} x \;\ge \frac{E}{\sigma^2}\int_{\|x\| \ge r} |\psi(x )|^2 \;\text{d} x \end{align}\] Therefore, \[\begin{align} \sigma^2 \; \ge \; \int _{\|x\| \ge r} |\psi(x)|^2 \;\text{d} x. \end{align}\] Now we can bound \[\begin{align} \|\psi - f_r\psi\|^2 & \;= \;\int \;\;\;\;\;(1 - f_r(x))^2 |\psi(x)|^2 \;\text{d} x \\ & \;= \; \int_{\| x\| \ge r} (1 - f_r(x))^2 |\psi(x)|^2 \;\text{d} x \\ & \;\le \;\int_{\|x\| \ge r} |\psi(x)|^2 \;\text{d} x \\ & \;\le \;\sigma^2. \end{align}\] ◻
Lemma 16 (Truncation in position). Let \(h :=-\Delta + V \in S_{\mathbb{R}^n}\cup S^D_{B} \cup S_{\mathbb{T}^n}\) be a Schrödinger operator on the domain \(\Omega\in\{\mathbb{R}^n, B, \mathbb{T}^n\}\), where \(B:=\{x\in\Omega \;| \;\|x\|_2\le r+1 \}\) and \(L>2(r+1)\).
Suppose \(V^{-1}([0,E/\sigma^6])\subset B^2_r\). Then, the Hilbert space \(\widetilde{\mathcal{D}}:= L^2(B)\) is an \((E,\epsilon)\)-truncated domain of \(h\), where \(E>1\), and \(\sigma = O(\epsilon/E)\).
Proof. By Theorem 1 and Lemma 3, there exists an orthonormal basis \(\{ \; \psi_i \;| \;h\psi_i = \lambda_i\psi _i, \;i \in \mathbb{Z}_{\ge 0} \;\}\) for the Hilbert space such that \(0\le \lambda_0\le \lambda_1 \le \lambda_2 \le \cdots\), and each \(\psi_i\) is smooth. Therefore, for a normalized vector \(\psi \in \mathcal{D}(h)\) with \(h[\psi]\le E\), we can write \[\begin{align} \psi = \sum_{i\in\mathbb{Z}_{\ge 0}} c_i\psi_i \end{align}\] for \(c_i\in\mathbb{C}\) such that \(\sum_i |c_i|^2 = 1\). We apply the Markov’s inequality on \(\{|c_i|^2\}_i\) as follows. For \(\sigma>0\), we have \[\begin{align} E\ge h[\psi] = \sum_{i \in \mathbb{Z}_{\ge 0}} |c_i|^2\lambda_i \ge \frac{E}{\sigma^2} \sum_{i: \lambda_i >E/\sigma^2} |c_i|^2 , \end{align}\] which gives us \[\begin{align} \sigma^2 \ge \sum_{i: \lambda_i >E/\sigma^2} |c_i|^2 = \|\psi - \psi^\downarrow\|^2, \; \;\;\;\;\;\text{where} \;\;\; \;\; \; \psi^\downarrow:=\sum_{i: \lambda_i \le E/\sigma^2 } c_i\psi_i. \label{eq:position95trunc95markov95ortho} \end{align}\tag{24}\] Since the spectrum of \(h\) is purely discrete, the number of \(i\) such that \(\lambda_i\le E/\sigma^2\) is finite. Otherwise, there would be an accumulation point of \(\lambda_i\)’s in the compact set \([0,E/\sigma^2]\), which is a contradiction to the fact that purely discrete spectrum accumulates only at \(\infty\). Therefore, \(\psi^\downarrow\) is smooth.
To prove the theorem, it is sufficient to prove the following claim: the vector \(\widetilde{\psi}\in \widetilde{\mathcal{D}}\) is an \(\sigma\)-truncation of \(\psi\), where \[\begin{align} \widetilde{\psi} \;& := \;\frac{f\psi^\downarrow }{\|f \psi^\downarrow \|} , \\ f(x)\;&:= \text{Cut}_{r,1}(\|x\|_2). \end{align}\] The function \(\widetilde{\psi}\) is supported on \(B\), and \(\|f\psi^\downarrow\| \le \|\psi^\downarrow\| \le \|\psi\|\). Therefore, we have \(\psi^\downarrow \in \widetilde{\mathcal{D}}\). We show that \(\widetilde{\psi}\) satisfies the norm and energy conditions of \(\sigma\)-trunction in the rest of the proof.
We first verify the norm condition of \(\sigma\)-truncation. By the triangle inequality, \[\begin{align} \| \psi - \widetilde{\psi} \| &\;\le\;\|\psi -f\psi^\downarrow\| + \|f\psi^\downarrow -f\psi^\downarrow /\|f\psi^\downarrow \|\|\;\nonumber \\ &\;=\;\|\psi -f\psi^\downarrow \| + |\|f\psi^\downarrow\| - 1| \\ &\;=\;\|\psi -f\psi^\downarrow\| + |\|f\psi^\downarrow\| - \|\psi\|| \nonumber \\ &\;\le\;2\|\psi - f\psi^\downarrow\| \\ \label{eq:thm95position95trunc95psitilde95nrom} & \; \le \; 2(\|\psi - \psi^\downarrow\| + \| \psi^\downarrow - f\psi^\downarrow\| \; ) \end{align}\tag{25}\] We have \(\|\psi - \psi^\downarrow\| \le \sigma\) due to 24 . Also, we have \(h[\psi^\downarrow] \le h[\psi]<E\), because \(h[\psi]\) is a convex combination of \(h[\psi^\downarrow]\) and \(h[\psi - \psi^\downarrow]\), and we know \(h[\psi^\downarrow]\le E/\sigma^2 \le h[\psi - \psi^\downarrow]\) by definition. By applying Lemma 15 on \(\psi^\downarrow/\|\psi^\downarrow\|\), we get \(\|\psi^\downarrow - f\psi^\downarrow\|/\|\psi^\downarrow\| \le \sigma^3.\) Therefore, we verify that \(\sigma =O(\epsilon/E)\) for a small enough constant factor satisfies the norm condition of \((E,\epsilon)\)-truncation as \[\begin{align} \|\psi - \widetilde{\psi} \| \le 2(\sigma + \sigma^3\|\psi^\downarrow\|) \le 2(\sigma + \sigma^3) = O(\epsilon/E)\le \epsilon. \label{eq:position95based95norm95condi} \end{align}\tag{26}\]
Now we prove the energy condition. Since \(\nabla f^2 =2f \nabla f = 0\) on the boundary of \(B\), we can apply the divergence theorem to \(|\psi^\downarrow|^2\nabla f^2\) on \(B\) and get \[\begin{align} 0 =\oint_{\partial B} (|\psi^\downarrow|^2\nabla f^2 )\cdot \frac{ x }{\|x\|} \; \text{d}^{n-1}x &= \int_{B} \nabla \cdot( |\psi^\downarrow|^2\nabla f^2) \; \text{d}^{n}x \\ &= \int_{B} 2\text{Re}(\overline{\psi^\downarrow} \nabla \psi^\downarrow) \cdot \nabla f^2 + |\psi^\downarrow|^2\Delta f^2 \; \text{d}^{n}x. \end{align}\] Since \(f\) is 0 outside the ball, \[\begin{align} -\int_{\Omega} \;\;\text{Re}(\overline{\psi^\downarrow}\nabla \psi^\downarrow) \cdot \nabla f^2 \;\text{d} x & \;=\; \nonumber \\ -\int_{B} \text{Re}( \overline{\psi^\downarrow}\nabla \psi^\downarrow) \cdot \nabla f^2 \;\text{d} x & \;= \;\frac{1}{2} \int_{B} |\psi^\downarrow|^2\Delta f^2 \; \text{d} x \nonumber \\ & \;= \;\frac{1}{2} \int_{\Omega} \;\; |\psi^\downarrow|^2\Delta f^2 \; \text{d} x \;= \;\langle \psi^\downarrow |\Delta f^2| \psi^\downarrow\rangle/2 \label{eq:divergence}. \end{align}\tag{27}\] Since \(h[\cdot]\) always gives a real number, we have \[\begin{align} &\;\;\;\;\; \;\|f\psi^\downarrow\| \;h[f\psi^\downarrow] \\ \nonumber & \;= \; \text{Re}(\|f\psi^\downarrow\| \;h[f\psi^\downarrow] ) \\ & \;= \; \text{Re}\int -f \overline{\psi^\downarrow} \Delta (f\psi^\downarrow) + f\overline{\psi^\downarrow} V f\psi^\downarrow \;\text{d}x \nonumber &\\ &\;= \; \int -f |\psi^\downarrow|^2 \Delta f - 2\text{Re}(f\overline{\psi^\downarrow} \nabla f\cdot \nabla \psi^\downarrow) - \text{Re}( f^2 \overline{\psi^\downarrow} \Delta {\psi^\downarrow}) + f^2\overline{\psi^\downarrow} V\psi^\downarrow \;\text{d}x \nonumber &\\ &\;= \; \int -f |\psi^\downarrow|^2 \Delta f - \text{Re}(\overline{\psi^\downarrow}\nabla \psi^\downarrow)\cdot \nabla f^2 - \text{Re}( f^2 \overline{\psi^\downarrow} \Delta \psi^\downarrow) + f^2\overline{\psi^\downarrow} V\psi^\downarrow \;\text{d}x \nonumber &\\ &\;= \; \int |\psi^\downarrow|^2 ( - f\Delta f + \Delta f^2/2) + \text{Re}(f^2 \overline{\psi^\downarrow}(- \Delta +V) \psi^\downarrow) \;\text{d}x \nonumber &\\ &\;= \; \langle \psi^\downarrow | \|\nabla f\|^2 |\psi^\downarrow\rangle + \text{Re}\langle f^2\psi^\downarrow|h\psi^\downarrow\rangle ,\label{eq:thm95position95trunc95hcmu} \end{align}\tag{28}\] where the second to the last equation follows from 27 , and the last equation is by applying the identity \(\Delta f^2 = 2|\nabla f|^2 + 2f \Delta f\).
We will show that the first term of 28 is small, and the second term is approximately \(\langle\psi^\downarrow|h\psi^\downarrow\rangle.\) To upper bound the first term, note that \(\nabla f\) is nonzero only outside \(B^2_r\) where we show \(\psi^\downarrow\) has a low weight. We define a multiplication operator \(p_r:\mathcal{H} \rightarrow \mathcal{H}\) such that for \(\phi \in \mathcal{H},\) \[\begin{align} (p_r\phi)(x) = \begin{cases} \phi(x) \; & \; \text{if } \;\|x\|_2 \le r, \\ 0 \; & \; \text{if } \;\|x\|_2> r. \end{cases} \end{align}\] We apply Lemma 15 on \(\psi^\downarrow/\| \psi^\downarrow\|\) and \(p_r\), and obtain \[\begin{align} \|(1-p_r)\psi^\downarrow\| = \left\|(1-p_r) \frac{\psi^\downarrow}{ \|\psi^\downarrow\|} \right\| \|\psi^\downarrow\| \le \sigma^3 \|\psi^\downarrow\| \le \sigma^3, \end{align}\] which is equivalent to \[\begin{align} \int_{x: \|x\|_2\ge r} |\psi^\downarrow(x) |^2 \;\text{d} x \le \sigma^6. \end{align}\] Therefore, we have \[\begin{align} \langle \psi^\downarrow | \|\nabla f\|^2 |\psi^\downarrow\rangle = \int_{x : \|x\|_2\ge r} \| \nabla f(x)\|^2 |\psi^\downarrow(x)|^2 \;\text{d} r \;\le \sigma^6 \max_{x: \|x\|_2\ge r}\left|\nabla f(x) \right|^2 . \label{eq:thm95position95trunc95nablaf95upperbound} \end{align}\tag{29}\] We compute the gradient in the polar coordinate with the radius \(t = \|x\|_2\) and get \[\begin{align} \max_{\|x\|\ge r} \|\nabla f(x)\|^2 &= \max_{t\ge r} \left(\frac{\text{d} \text{Cut}_{r,1}(t)}{\text{d} t}\right)^2 = \left(\frac{ \text{Bump}(t - r) }{\int_r^{r+1} \text{Bump}( {y - r}) \;\text{d} y } \right)^2 \le \left( \frac{\max_{t \in [0,1]} \text{Bump}(t) }{\int_0^1 \text{Bump}(t) \text{d} t } \right)^2 \\ &= {C}, \end{align}\] where \(C\) is a universal constant. Applying this upper bound to 29 gives \[\begin{align} \langle\psi^\downarrow|\|\nabla f\|^2|\psi^\downarrow\rangle \le {\sigma^6 C}{}. \label{eq:thm95position95trunc95first95term} \end{align}\tag{30}\]
Now we show that the second term of 28 is close to \(\langle \psi^\downarrow | h\psi^\downarrow\rangle\). By the Cauchy-Schwarz inequality, \[\begin{align} |\langle f^2\psi^\downarrow|h\psi^\downarrow\rangle -\langle \psi^\downarrow | h\psi^\downarrow\rangle | \; = \; |(\langle f^2\psi^\downarrow| -\langle \psi^\downarrow |)|h\psi^\downarrow\rangle | \; \le \; \|f^2 \psi^\downarrow - \psi^\downarrow\|\|h\psi^\downarrow\| \label{eq:thm95position95trunc95secondterm} \end{align}\tag{31}\] By applying Lemma 15 on \(\psi^\downarrow/\|\psi^\downarrow\|\) with respect to \(1 -f^2\), we bound the first norm \[\begin{align} \|f^2\psi^\downarrow - \psi^\downarrow\| \le \sigma^3 \| \psi^\downarrow\| \le \sigma^3. \label{eq:thm95position95trunc95secondterm95firstterm} \end{align}\tag{32}\] The second norm is bounded as: \[\begin{align} &\|h\psi^\downarrow\| \;=\; \sqrt{\sum_{i: \lambda_ i \le E/\sigma^2}\lambda_i^2 |c_i|^2} \le \frac{E}{\sigma^2}. \label{eq:thm95position95trunc95secondterm95secondterm} \end{align}\tag{33}\]
Plugging 32 , and 33 into 31 gives \[\begin{align} |\langle f^2\psi^\downarrow|h\psi^\downarrow\rangle -\langle \psi^\downarrow | h\psi^\downarrow\rangle | \; \le \; E\sigma \end{align}\] Since \(\langle \psi ^\downarrow|h\psi^\downarrow\rangle\) is real, we have \[\begin{align} |\text{Re}\langle f^2\psi^\downarrow|h\psi^\downarrow\rangle -\langle \psi^\downarrow |h\psi^\downarrow\rangle| \le |\langle f^2\psi^\downarrow|h\psi^\downarrow\rangle -\langle \psi^\downarrow | h\psi^\downarrow\rangle | , \end{align}\] and therefore \[\begin{align} \text{Re}\langle f^2\psi^\downarrow|h\psi^\downarrow\rangle \;\le \;& \langle \psi^\downarrow|h\psi^\downarrow\rangle + \sigma E \nonumber \\ \;= \;& h[\psi^\downarrow]\|\psi^\downarrow\|^2 + \sigma E \nonumber \\ \;\le \;& h[\psi] + \sigma E . \label{eq:thm95position95trunc95second95term} \end{align}\tag{34}\] By 28 , 30 , and 34 , we get \[\begin{align} \|f\psi^\downarrow\| \;h[f\psi^\downarrow] &\;\le \; h[\psi] + \sigma E + {\sigma^6 C}{} . \end{align}\] The triangle inequality implies \[\begin{align} \|f\psi^\downarrow\| \;\ge \;\|\psi^\downarrow\| -\|(1 - f) \psi^\downarrow\| \;\ge \; 1 - \sigma - \sigma^3. \end{align}\] Therefore, we have \[\begin{align} h[\widetilde{\psi}] = h[f\psi^\downarrow] \le \frac{ h[\psi] + \sigma E+ {\sigma^6 C}{} }{ 1 - \sigma - \sigma^3} \le h[\psi] + \epsilon \end{align}\] for \(\sigma = O(\epsilon/E)\). ◻
In this section, we show that \(h_{\mathbb{T}^n}\) and \(h_Q\) are equivalent. We first show that \[\begin{align} {\widetilde{\mathcal{D}}}_{\mathbb{T}^n,f} &:= \text{span}\left\{ \; \omega_k \;| \;k\in \mathbb{Z}^n, \;\|k\|_2\le K \right\} \\ {\widetilde{\mathcal{D}}}_{Q} &:= \text{span}\{ \;|p_k\rangle \;| \; k\in \mathbb{Z}^n,\|k\|_2\le K \}, \end{align}\] are \((E,\epsilon)\)-truncated domains of \(h_{\mathbb{T}^n}\) (Lemma [lem:freq95trunc95torus]) and \(h_Q\) (Lemma 18) respectively, where \[\begin{align} \omega_k(x) \;:=& \; \frac{1}{\sqrt {L^n}}e^{i2\pi k\cdot x/L} \;\;\;\text{for } x \in \mathbb{T}^n, \\ |p_k\rangle &\; := \frac{1}{\sqrt{N^n}} \sum_{y \in\mathcal{N}^n} e^{i 2\pi k\cdot y/N } | y\rangle \end{align}\] are the Fourier basis of the respective space. They are identified by the unitary \(U:{\widetilde{\mathcal{D}}}_{\mathbb{T}^n,f}\rightarrow {\widetilde{\mathcal{D}}}_{Q}\) such that \(U\omega_k = |p_k\rangle.\)
Proof of Lemma 6. By Lemma [lem:freq95trunc95torus] and Lemma 18, \[\begin{align} {\widetilde{\mathcal{D}}}_{\mathbb{T}^n,f} &:= \text{span}\left\{ \; \omega_k \;| \;k\in \mathbb{Z}^n, \;\|k\|_2\le K \right\}, \\ {\widetilde{\mathcal{D}}}_{Q} &:= \text{span}\{ \;|p_k\rangle \;| \; k\in \mathbb{Z}^n,\|k\|_2\le K \} \end{align}\] are \((E,\epsilon)\)-truncated domains of \(h_{\mathbb{T}^n}\) and \(h_Q\), respectively, provided that \[\begin{align} K = \Omega\left(\frac{L (E+V_{\max})^{3/2}) }{ \epsilon }\right). \end{align}\] By Lemma 19, the unitary \(U:\widetilde{\mathcal{D}}_{\mathbb{T}^n,f}\rightarrow {\widetilde{\mathcal{D}}}_{Q}\) defined by \(U \omega_k= |p_k\rangle\) satisfies \(|h[\psi] - h[U\psi]|\le \epsilon\) for all \(\psi\in\widetilde{\mathcal{D}}_{\mathbb{T}^n,f}\), provided that \[\begin{align} N\ge 4 K \ge 16\left(\frac{L p(2n) }{2\pi} \right)^2 \frac{1}{e^{1/n}}. \end{align}\] Therefore, \[\begin{align} N = 4K = \Omega\left( L^2 (E + V_{\max})^{1.5}(p(2n))^2/\epsilon \right) \ge \Omega\left( \max \left\{ {L (E+V_{\max})^{1.5}) }/{ \epsilon } , ({L p(2n) } )^2 /{e^{1/n}} \right\} \right) \end{align}\] gives that \(h_{\mathbb{T}^n}\) and \(h_Q\) are \((E,\epsilon)\)-equivalent. ◻
Lemma 17 (Frequency truncation on torus). Let \(\mathbb{T}^n :=\mathbb{R}^n/L\mathbb{Z}^n\) be the \(n\)-dimensional torus of length \(L\), \(h = -\Delta + V\) be a Schrödinger operator on \(\mathbb{T}^n\) such that \(V(\mathbb{T}^n ) \subset [0,V_{\max} ]\) for some \(V_{\max}>1\). Then \[\begin{align} &{\widetilde{\mathcal{D}}}_{\mathbb{T}^n,f} := \text{span}\left\{ \; \omega_k \;| \;k\in \mathbb{Z}^n, \;\|k\|_2\le K \right\} \end{align}\] is an \((E, \epsilon)\)-truncated domain of \(h\), where \[\begin{align} K \;=& \; \Omega\left( \frac{L (E+V_{\max})^{3/2})}{\epsilon} \right). \end{align}\]
Proof. Let \(\psi \in \mathcal{D}(h)\) be a normalized vector with energy \(h[\psi] \le E\). We show that there exists an \((E,\epsilon)\)-truncation of \(\psi\) in \(\widetilde{ \mathcal{D}} _{\mathbb{T}^n, f}\). Since \(\psi\) is also in \(L^2(\mathbb{T}^n)\), \(\psi\) has the Fourier decomposition \[\begin{align} \psi = \sum_{k\in \mathbb{Z}^n} \widehat{\psi}_k\omega_k. \end{align}\] Since \(\langle \psi |V|\psi\rangle \ge 0\), we have \[\begin{align} E \ge \langle \psi | (-\Delta)\psi\rangle = \sum_{k \in \mathbb{Z}^n} \frac{4\pi^2\|k\|_2^2 \widehat{\psi}_k}{L^2} \langle\psi| \omega_k \rangle = \sum_{k \in \mathbb{Z}^n} \frac{4 \pi^2 \|k\|^2 }{L^2} | \widehat{\psi}_k |^2 . \label{eq:thm95frequency95trunc95kinetic} \end{align}\tag{35}\] Given the parameter \(K\), we define low- and high-frequency vectors \[\begin{align} \psi^\downarrow \; &: = \sum_{\|k\|\le K} \widehat{\psi}_k\omega_k \in \widetilde{\mathcal{D}}_{\mathbb{T}^n, f} , \\ \psi^\uparrow \; &: = \sum_{\|k\| > K} \widehat{\psi}_k\omega_k = \psi - \psi^\downarrow. \end{align}\] Since \(\{\omega_k\}_k\) forms a complete orthonormal basis set in \(L^2(\mathbb{T}^n)\), Parseval’s theorem gives \(\sum_{k \in \mathbb{Z}^n} |\widehat{\psi}_k |^2 = 1\). By the Markov’s inequality on \(|\widehat \psi_k|^2\), we obtain \[\begin{align} \|\psi^\uparrow\|^2 \;= \; \|\psi^\downarrow - \psi \|^2 \;= \;\sum_{\|k\| \ge K} |\widehat{\psi}_k |^2 \;\le \;\frac{EL^2}{4\pi^2 K^2} :=\sigma^2 \label{eq:thm95frequency95trunc95sigma} \end{align}\tag{36}\] where \(\sigma \ge 0.\) We claim that the vector \(\widetilde{\psi}:=\psi^\downarrow/\|\psi^\downarrow\|\) is an \(\epsilon\)-truncation of \(\psi.\) For the norm condition: \[\begin{align} \left\|\frac{\psi^\downarrow}{\|\psi^\downarrow\|} - \psi \right\| & \; \le \; \left\|\frac{\psi^\downarrow}{\|\psi^\downarrow\|} - \psi^\downarrow \right\| + \| \psi ^\downarrow - \psi \| \nonumber \\ & \;= \; \left|\| \psi\| -{\|\psi^\downarrow\|} \right| + \| \psi ^\downarrow - \psi \| \nonumber \\ & \;\le \; 2 \| \psi ^\downarrow - \psi \| \nonumber \\ &\; \le \; 2\sigma. \label{eq:thm95frequency95trunc95norm} \end{align}\tag{37}\] We choose \(\sigma =O( {\epsilon}/{(E + V_{\max} )} )\), so that \(2\sigma\le \epsilon\).
For the energy condition of \((E,\epsilon)\)-truncation, we first bound the Laplacian term. From 35 , we have \[\begin{align} \langle \psi^\downarrow |(-\Delta)|\psi^\downarrow\rangle - \langle \psi |(-\Delta)|\psi\rangle = - \sum_{k :\|k\|_2 > K} \frac{4\pi^2\|k\|^2 }{L^2}|\widehat{\psi}_k |^2 \le 0. \label{eq:thm95frequency95trunc95torus95diff95kinetic} \end{align}\tag{38}\] In order to upper bound \(\langle \widetilde{ \psi}|V|\widetilde{\psi}\rangle - \langle \psi|V|\psi\rangle\), we use the fact \(\| V|\psi\rangle \| \;\le \;V_{\max} \; \| \psi \|.\) By linearity and norm considerations, we have \[\begin{align} | \langle \psi^\downarrow|V|{\psi}^\downarrow\rangle - \langle \psi|V|\psi\rangle| & \; = \; |\langle \psi^\downarrow|V|{\psi}^\downarrow\rangle - (\langle \psi^\downarrow | + \langle \psi^\uparrow | )V (| \psi^\downarrow \rangle + | \psi^\uparrow \rangle ) | \nonumber \\ & \;= \; | - \langle \psi^\downarrow|V|{\psi}^\uparrow\rangle - \langle \psi^\uparrow|V|{\psi}^\downarrow\rangle - \langle \psi^\uparrow|V|{\psi}^\uparrow\rangle | \nonumber \\ & \;\le \; V_{\max} (\|\psi^\downarrow \| \|\psi^\uparrow \| + \|\psi^\uparrow \| \|\psi^\downarrow\| +\|\psi^\uparrow \| \|\psi^\uparrow \| ) \nonumber \\ & \;\le \; V_{\max}(2\sigma + \sigma^2). \label{eq:thm95frequency95trunc95torus95diff95potential} \end{align}\tag{39}\] Combining 38 and 39 gives \[\begin{align} \langle \psi^\downarrow |h|\psi^\downarrow\rangle - \langle \psi |h|\psi\rangle & \;= \; \langle \psi^\downarrow |(-\Delta)|\psi^\downarrow\rangle - \langle \psi |(-\Delta)|\psi\rangle + \langle \psi^\downarrow |V|\psi^\downarrow\rangle - \langle \psi |V|\psi\rangle \\ & \;\le \;V_{\max}(2\sigma + \sigma^2). \end{align}\] By the fact that \(\|\psi^\downarrow\| \ge \|\psi\| - \|\psi^\downarrow - \psi\| \ge 1 - \sigma\), we have \[\begin{align} h[\psi^\downarrow] \;= \; \frac{\langle \psi^\downarrow |h|\psi^\downarrow\rangle}{\| \psi^\downarrow\|^2} & \;\le \; \frac{h[\psi] + V_{\max}(2\sigma + \sigma^2 )}{\|\psi^\downarrow\|^2} \nonumber\\ & \;\le \; \frac{h[\psi] + V_{\max}(2\sigma + \sigma^2)}{( 1- \sigma)^2} \nonumber\\ & \;= \; h[\psi] + \frac{h[\psi](2\sigma - \sigma^2) + V_{\max}(2\sigma + \sigma^2)}{( 1- \sigma)^2} \nonumber \\ & \;\le \; h[\psi] + \frac{E(2\sigma - \sigma^2) + V_{\max}(2\sigma + \sigma^2)}{( 1- \sigma)^2} \nonumber \\ & \;\le \; h[\psi] + \frac{2\sigma + \sigma^2 }{( 1- \sigma)^2} (E + V_{\max}) \nonumber \\ & \;\le \;\epsilon \nonumber, \end{align}\] where the last line is by our choice \(\sigma =O( {\epsilon}/{(E + V_{\max} )} )\). By 36 , we have \(\sigma \le L\sqrt{E + V_{\max}} /( 2\pi K)\). Therefore, we have that \(\widetilde{ \mathcal{D}} _{\mathbb{T}^n, f}\) is an \((E,\epsilon)\)-truncated domain, when we have \[\begin{align} K =\Omega\left(\frac{L (E+V_{\max})^{3/2}) }{ \epsilon }\right). \end{align}\] ◻
Lemma 18 (Frequency truncation on qubit). Let \(h_Q\) be a qubit Schrödinger operator on \(N^n\)-dimensional space such that \(V(x) \in [0,V_{\max} ]\) for all \(x\in \mathbb{T}^n\) and \(V_{\max}>1\). Let \(\mathcal{D}_Q\) be the \(N^n\)-dimensional Hilbert space and let \(|p_k\rangle\) be defined as above.
Then \[\begin{align} &{\widetilde{\mathcal{D}}}_{Q} := \text{span}\{ \;|p_k\rangle \;| \; k\in \mathbb{Z}^n,\|k\|_2\le K \}, \end{align}\] is an \((E, \epsilon)\)-truncation of \(\mathcal{D}_Q\) with respect to \(h_Q\), where \[\begin{align} K \;=& \; \Theta\left( \frac{L (E+V_{\max})^{3/2} }{\epsilon} \right) \end{align}\] for some constant factor.
Proof. Let us assume that an element \(|\psi\rangle \in \mathcal{D}_{Q}\) has energy \(h[|\psi\rangle] \le E\). We write \(|\psi\rangle\) in the Fourier basis: \[\begin{align} |\psi\rangle = \sum_{k\in \mathbb{Z}^n} \widehat{\psi}_k |p_k\rangle . \end{align}\] Because \(\langle \psi |V|\psi\rangle \ge 0\), we have \[\begin{align} E \ge \langle \psi | (-\Delta_Q)|\psi\rangle = \sum_{k \in \mathbb{Z}^n} \frac{4 \pi^2 \|k\|^2 }{L^2} | \widehat{\psi}_k |^2 . \nonumber \end{align}\] After defining vectors \[\begin{align} |\psi^\downarrow\rangle \; &: = \sum_{\|k\|\le K} \widehat{\psi}_k |p_k\rangle \in \widetilde{\mathcal{D}}_{Q} , \\ |\psi^\uparrow\rangle \; &: = \sum_{\|k\| > K} \widehat{\psi}_k |p_k\rangle = \psi - \psi^\downarrow, \end{align}\] the rest of the proof is verbatim from the proof of Theorem [lem:freq95trunc95torus] and therefore omitted. ◻
Our main goal in this section is to show that \((h_\mathbb{T}, \widetilde{\mathcal{D}}_{\mathbb{T},2})\) and \((h_Q, \widetilde{\mathcal{ D}}_Q)\) are \(\epsilon\)-equivalent, given that \(V\) is “smooth enough” according to a quantifiable notion of smoothness that we call smoothness factor. The Fourier decomposition for \(V_\mathbb{T}\) plays a crucial role in doing so.
Lemma 19. Let \(V:\mathbb{T}^n \rightarrow \mathbb{R}\) be a smooth potential on the torus. Suppose \[\begin{align} h & : = -\Delta +V \qquad \qquad \qquad \qquad \quad \text{on \;}\mathbb{T}^n, \\ h_{Q} & : = \; K_{Q} +V_{Q} \qquad \qquad \qquad \qquad \text{on \;}\log N\times n \;\text{qubits,} \\ \widetilde{\mathcal{D}}_{\mathbb{T}^n,f} & := \text{span}\{\omega_k | \; \;k \in \mathbb{Z}^n, \;\| k\|_2 \le K \} \\ \widetilde{ D}_{Q} & := \text{span}\{|p_k\rangle | \; \;k \in \mathbb{Z}^n, \;\| k\|_2 \le K \}. \end{align}\] Furthermore, suppose \(\log_\partial (V_{\mathbb{T}}, \mathbb{T}^n, m) \le \log p(m)\) for some polynomial \(p\), Define a unitary \(U:\widetilde{\mathcal{D}}_{\mathbb{T}^n,f}\rightarrow \widetilde{ D}_{Q}\) by \(U\omega_k:= |p_k\rangle\).
Then, for all \(\psi\in\widetilde{\mathcal{D}}_{\mathbb{T}^n,f}\), we have \(|h_{\mathbb{T}^n}[\psi] - h_Q[U\psi] | \le \epsilon\), given tha \[\begin{align} N\ge 4 K \ge 16\left(\frac{L p(2n) }{2\pi} \right)^2 \frac{1}{e^{1/n}}. \end{align}\]
Proof. For a normalized state \(\psi = \sum_{k: \|k\|_2 \le K} c_k \omega_k \in \widetilde{\mathcal{D}}_{\mathbb{T}^n,f}\) and \(|\phi\rangle :=U\psi= \sum_{k: \|k\|_2 \le K} c_k |p_k\rangle \in \widetilde{\mathcal{D}}_Q\), we prove that \[\begin{align} |h[\psi] - h_Q[\phi]| \le \epsilon. \end{align}\]
We first compute \[\begin{align} h_{{\mathbb{T}}}[\psi] \;&= \; \langle \psi|(-\Delta) | \psi \rangle + \langle \psi | V|\psi\rangle \\ \;&= \; \sum_{k:\|k\|_2\le K} |c_k|^2 \frac{4\pi^2 \|k\|_2^2}{L^2} + \sum_{\|k\|,\|k'\|\le K} \overline{c_{k'}} c_k \langle \omega_{k'}|V| \omega_k\rangle \\ \;&= \; \sum_{k:\|k\|_2\le K} |c_k|^2 \frac{4\pi^2 \|k\|_2^2}{L^2} + \sum_{\|k\|,\|k'\|\le K} \overline{c_{k'}} c_k \int_{x\in\mathbb{T}^n} \frac{V(x) e^{i 2\pi (k - k')\cdot x/L}}{L^n} \; \text{ d} x \\ \;&= \; \sum_{k:\|k\|_2\le K} |c_k|^2 \frac{4\pi^2 \|k\|_2^2}{L^2} + \sum_{\|k\|,\|k'\|\le K} \overline{c_{k'}} c_k \widehat{V}_{k' - k} , \end{align}\] and similarly, \[\begin{align} h_{Q}[\phi] \;&= \; \langle \phi|K_Q | \phi \rangle + \langle \phi | V_Q|\phi\rangle \\ \;&= \; \sum_{k:\|k\|_2\le K} |c_k|^2 \frac{4\pi^2 \|k\|_2^2}{L^2} + \sum_{\|k\|,\|k'\|\le K} \overline{c_{k'}} c_k \langle p_{k'}|V_Q| p_k\rangle \\ \;&= \; \sum_{k:\|k\|_2\le K} |c_k|^2 \frac{4\pi^2 \|k\|_2^2}{L^2} + \sum_{\|k\|,\|k'\|\le K} \overline{c_{k'}} c_k \sum_{y\in\mathbb{[}N]^n} \frac{V(yL/N) e^{i 2\pi (k - k')\cdot y/N}}{N^n} . \end{align}\] Therefore, \[\begin{align} |h[\psi] - h_Q[\phi]| & \;= \; \left |\sum_{\|k\|,\|k'\|\le K} \overline{c_{k'}} c_k \bigg( \widehat{V}_{k' - k} - \sum_{y\in\mathbb{[}N]^n} \frac{V(yL/N) e^{i 2\pi (k - k')\cdot y/N}}{N^n} \bigg) \right| \nonumber \\ & \;\le \; \sum_{\|k\|,\|k'\|\le K} |c_{k'}| |c_k| \max_{\|k\|,\|k'\| \le K} \left| \widehat{V}_{k' - k} - \sum_{y\in\mathbb{[}N]^n} \frac{V(yL/N) e^{i 2\pi (k - k')\cdot y/N}}{N^n} \right| \nonumber \\ & \;\le \; \max_{\|k\|,\|k'\| \le K} \left| \widehat{V}_{k' - k} - \sum_{y\in\mathbb{[}N]^n} \frac{V(yL/N) e^{i 2\pi (k - k')\cdot y/N}}{N^n} \right| \tag{40} \\ & \;\le \; \max_{\|k\| \le 2K} \left| \widehat{V}_{k} - \sum_{y\in\mathbb{[}N]^n} \frac{V(yL/N) e^{-i 2\pi k\cdot y/N}}{N^n} \right| \tag{41} \\ & \;= \; \max_{\|k\|_2 \le 2K} \left| \int_{\mathbb{T}^n} \frac{V (x)e^{-i2\pi k\cdot x /L} }{L^n} \;\text{d} x - \sum_{y\in\mathbb{[}N]^n} \frac{V(yL/N) e^{-i 2\pi k\cdot y/N}}{N^n} \right| \nonumber \end{align}\] where 40 is by the Cauchy-Schwarz inequality on \(|c_k|\), and 41 is by the change of variables \(k'-k\rightarrow k.\) The expression in the last line is the maximum discrepancy between the integral average and the corresponding Riemann sum average of \(V(x)e^{-i2\pi k \cdot x /L}\). The key observation for bounding this quantity is that the discrepancy vanishes whenever the function has no high-frequency Fourier components. We consider the decomposition \(V = V^\downarrow + V^\uparrow\), where \[\begin{align} V ^\downarrow(x) = \sum_{l: \|l\|_\infty \le K} \widehat {V}_l e^{i2\pi l\cdot x/L}, \\ V ^\uparrow(x) = \sum_{l: \|l\|_\infty > K} \widehat {V}_l e^{i2\pi l\cdot x/L}. \end{align}\] Then for \(k\) such that \(\|k\|_2 \le 2K\), \[\begin{align} \sum_{y\in\mathbb{[}N]^n} \frac{V^\downarrow (yL/N) e^{-i2\pi k\cdot y /N}}{N^n} \;&= \; \sum_{l: \|l\|_\infty \le K } \sum_{y\in\mathbb{[}N]^n} \frac{\widehat{V}_l e^{i2\pi l\cdot y/N} e^{-i2\pi k\cdot y /N}}{N^n} \nonumber \\ \;&= \; \sum_{l: \|l\|_\infty \le K } \widehat{V}_l \sum_{y\in\mathbb{[}N]^n} \frac{ e^{i2\pi (l-k)\cdot y/N} }{N^n} \nonumber \\ \;&= \; \sum_{\substack{l:\|l\|_\infty \le K \\ l = k \mod N} } \widehat{V}_k. \nonumber \\ \;&= \; \widehat{V}_k. \nonumber \end{align}\] The last equation holds because we enforce \(N\ge3K+1\). Since \(\|k \| _\infty \le \|k\|_2 \le 2K\) and \(\|l\|_\infty \le K\), we have \(\|k - l\|_\infty\le \|k\| _\infty + \|l\|_\infty \le 3K.\) Therefore, \(k=l \mod N\) if and only if \(k=l \in \mathbb{Z}^n\).
We conclude that \[\begin{align} |h[\psi] - h_Q[\phi]| & \;\le \; \max_{\|k\| \le 2K} \left| \widehat{V}_{k} - \sum_{y\in\mathbb{[}N]^n} \frac{V(y L/N) e^{-i 2\pi k\cdot y/N}}{N^n} \right| \\ & \;= \; \max_{\|k\| \le 2K} \left| \widehat{V}_{k} - \sum_{y\in\mathbb{[}N]^n} \frac{V^\downarrow(yL/N) e^{-i 2\pi k\cdot y/N}}{N^n} - \sum_{y\in\mathbb{[}N]^n} \frac{ V^\uparrow(yL/N) e^{-i 2\pi k\cdot y/N}}{N^n} \right| \\ & \;= \; \max_{\|k\| \le 2K} \left| \sum_{y\in\mathbb{[}N]^n} \frac{ V^\uparrow(yL/N) e^{-i 2\pi k\cdot y/N}}{N^n} \right| \\ & \;\le \; \max_{\|k\| \le 2K} \sum_{y\in\mathbb{[}N]^n} \frac{ |V^\uparrow(yL/N) | |e^{-i 2\pi k\cdot y/N}|}{N^n} \\ & \;= \; \max_{y\in[N]^n} V^\uparrow(yL/N) \; \le\; \max_{x\in[0,L]^n} V^\uparrow(x) \\ & \;= \; \max_{x\in[0,L]^n} \sum_{l:\|l\|_\infty >K} \widehat{V}_k e^{i2\pi k\cdot x/L} \\ &\; \le \;\sum_{l:\|l\|_\infty>K} |\widehat{V}_k|. \end{align}\]
We write \(x = (x_1,\dots,x_n) = (x_j, x_{ j*}) \in [0,L]^n\), \(l = (l_1,\dots,l_n) = (l_j, l_{j*}) \in \mathbb{Z}^n\) where \(x_{ j*} := (x_1,\dots, x_{j-1},x_{j+1},\dots, x_n)\), and \(l_{j*}\) is defined similarly. Then \[\begin{align} \widehat V_l &= \frac{1}{L^n} \int_{x\in\mathbb{[}0,L]^n} V(x) e^{-i2\pi l\cdot x/L} \;\text{d} x \nonumber \\ & = \frac{1}{L^n}\int_{ x_{j*} \in [0,L]^{n-1} } \;e^{-i2\pi l_{j*} \cdot \; x_{j*} /L} \int_{x_j \in [0,L]} V(x_j , x_{j*}) e^{-i2\pi l_j x_j /L} \;\text{d} x_j \;\text{d} x_{j*} \label{eq:vl95integration} \end{align}\tag{42}\] Fubini’s theorem lets us integrate the whole expression first in the variable \(x_j\) and then in the other variables sequentially. By repeatedly integrating by parts in \(x_j\) and assuming \(l_j \ne 0,\) we have, with the notation \(\partial_j^m:=\frac{\partial^m}{\partial x_j^m}\), \[\begin{align} & \int_{0}^L V(x_j , x_{j*}) e^{-i2\pi l_j x_j /L} \;\text{d} x_j \\ = & \frac{L}{-i2\pi l_j} V(x_j, x_{j*})e^{-2\pi l_j x_j/L} \Bigg|_0^L - \frac{L}{-i2\pi l_j}\int_{0}^L e^{-i2\pi l_j x_j /L} {\partial_j V(x_j,x_{j*})} \;\text{d} x_j \\ = & \frac{L}{i2\pi l_j}\int_{0}^L e^{-i2\pi l_j x_j /L} \;{\partial_j V(x_j,x_{j*})} \; \text{d} x_j \\ = & \left(\frac{L}{i2\pi l_j} \right)^2 \int_{0}^L e^{-i2\pi l_j x_j /L} \; {\partial_j^2 V (x_j,x_{j*} )} \;\text{d} x_j \\ &\vdots \\ = & \left(\frac{L}{i2\pi l_j} \right)^m \int_{0}^L e^{-i2\pi l_j x_j /L} \;\partial_j^m V(x_j,x_{j*}) \; \text{d} x_j \end{align}\] Plugging the result into 42 , we get \[\begin{align} \widehat V_l & \;= \; \frac{1}{L^n} \left(\frac{L}{i2\pi l_j} \right)^m \int_{ x \in [0,L]^n } e^{-i2\pi l\cdot x /L} \;{\partial^m_j V(x)} \; \text{d} x \end{align}\] for all \(m \in \mathbb{N}\). Since \(j\in [n]\) is arbitrary, we take \(j\) such that \(l_j = \|l\|_\infty\) and obtain \[\begin{align} | \widehat V_l | & \;\le \; \frac{1}{L^n} \left(\frac{L}{2\pi \|l\|_\infty} \right)^m \int_{ x \in [0,L]^n } | {\partial^m_j V(x)} | \; \text{d} x \\ & \;\le \; \left(\frac{L}{2\pi \|l\|_\infty} \right)^m \max_{x\in [0,L]^n} | {\partial^m_j V(x)} | \\ & \;\le \; \left(\frac{L}{2\pi \|l\|_\infty} \right)^m V_{\max} ^{(m)} \end{align}\] for all \(m \in \mathbb{N}\) where \(V^{(m)}_{\max} := \max \{|\partial_j^m V(x)| \;| \; x\in \mathbb{T}^n, j\in[n]\}\). Assuming \(m\ge n+1\), we have \[\begin{align} \sum_{l:\|l\|_\infty > K} |\widehat V_l| &\le V^{(m)}_\text{max} \sum_{\|l\|_\infty > K} \left(\frac{L}{2\pi \|l\|_{\infty}}\right)^m \\ &= V^{(m)}_\text{max} \sum_{t= K+1}^\infty \sum_{l: \|l\|_\infty = t} \left(\frac{L}{2\pi \|l\|_{\infty}}\right)^m \\ & \le V^{(m)}_\text{max} \sum_{t = K+1}^\infty 2n(2t+1)^{n-1} \left(\frac{L}{2\pi t}\right)^m \\ & \le 2n\cdot 4^{n-1} V^{(m)}_\text{max} \left(\frac{L}{2\pi }\right)^m \sum_{t =K+1}^\infty \frac{1}{t^{m-n+1 }} \\ & \le 4^{n}n V^{(m)}_\text{max} \left(\frac{L}{2\pi }\right)^m \frac{1}{(m-n) {K}^{m-n }}, \end{align}\] where the last inequality is by integration over \(t \in [K,\infty)\). We set \(m =2n\). We want the last quantity to be less than \(\epsilon,\) or equivalently \[\begin{align} n\log 4 + \log V^{(2n)}_{\max} + m \log \left( \frac{L}{2\pi} \right) - n\log K \le \log \epsilon. \end{align}\] This inequality is satisfied by choosing \(K\) so that \[\begin{align} & 2 \frac{\log V^{(2n)} _{\max}}{2n} + \log 4 + 2\log\left(\frac{L}{2\pi}\right) + \log \frac{1}{\epsilon^{1/n}} \\ \; \le \; &2 \log_\partial(V, \mathbb{T}^n, 2n) + \log 4 + 2\log\left(\frac{L}{2\pi}\right) + \log \frac{1}{\epsilon^{1/n}} \\ \; \le \; &2 \log p(2n) + \log 4 + 2\log\left(\frac{L}{2\pi}\right) + \log \frac{1}{\epsilon^{1/n}} \\ \;\le \;&\log K. \end{align}\] Therefore, we set \(K =\Theta\left(\left( p(2n) \frac{L}{2\pi}\right)^2 \frac{1}{\epsilon^{1/n}} \right)\) to get the desired bound ◻
The following lemmas state that having a logarithmic smoothness factor is closed under sum and composition.
Lemma 20. Let \(f,f_k:\mathbb{T}^n \rightarrow\mathbb{R}\) be smooth for \(k\in[r]\), and \(c>0\). Then we have \[\begin{align} \log_\partial( cf, \mathbb{T}^n, m) &\; \le \; (\log c)^+ \;+ \; \log_\partial( f, \mathbb{T}^n, m) , \\ \log_\partial(\sum_{i\in[r]} f_i, \mathbb{T}^n, m) &\; \le \; \log r \;+ \; \sum_{i\in[r]} \log_\partial( f_i, \mathbb{T}^n, m) . \end{align}\]
Proof. The proof is by manipulating the maximums: \[\begin{align} \log_\partial( cf, \mathbb{T}^n, m) = &\max_{l\in[m], j\in[n],x \in\mathbb{T}^n} \frac{(\log c| \partial^l_j f(x)|)^+}{l } \\ = & \max_{l\in[m], j\in[n],x \in\mathbb{T}^n} \frac{(\log c + \log | \partial^l_j f(x)|)^+}{l } \\ \le & (\log c)^+ + \max_{l\in[m], j\in[n],x \in\mathbb{T}^n} \frac{( \log | \partial^l_j f(x)|)^+}{l } = (\log c)^+ + \log_\partial ( f, \mathbb{T}^n, m), \end{align}\] and \[\begin{align} \log_\partial( \sum_{k \in[r]} f_k, \mathbb{T}^n, m) \;=\; &\max_{l\in[m], j\in[n],x \in\mathbb{T}^n} \frac{(\log | \sum_{k\in [r]} \partial^l_j f_k(x)|)^+}{l } \\ \;\le \;& \max_{l\in[m], j\in[n],x \in\mathbb{T}^n} \frac{(\log (\sum_{k\in [r]} | \partial^l_j f_k (x)| )^+ }{l } \\ \; \le \;& \max_{l\in[m], j\in[n],x \in\mathbb{T}^n} \frac{(\log ( r\max_{k\in[r]} | \partial^l_j f_k (x)| ) )^+ }{l } \\ \; \le \;& \max_{l\in[m], j\in[n],x \in\mathbb{T}^n} \frac{(\log ( \max_{k\in[r]} | \partial^l_j f_k (x)| ) )^+ + \log r }{l } \\ \; \le \;& \log r + \max_{l\in[m], j\in[n],x \in\mathbb{T}^n} \frac{\max_{k\in[r]} (\log | \partial^l_j f_k (x)| )^+ }{l } \\ \; \le \;& \log r + \max_{k\in[r]} \max_{l\in[m], j\in[n],x \in\mathbb{T}^n} \frac{(\log | \partial^l_j f_k (x)| )^+ }{l } \\ = & \log r + \max_{k\in[r]} \log_\partial(f_k, \mathbb{T}, m) . \end{align}\] ◻
Lemma 21. Let \(g:\mathbb{T}^n \rightarrow\mathbb{R}\) and \(f:g(\mathbb{T}^n) \rightarrow \mathbb{R}\) be smooth, where \(g(\mathbb{T}^n)\) is the range of \(g\). Then we have \[\begin{align} \log_\partial( f\circ g, \mathbb{T}^n, m) \; \le \; 2 \log m \;+ \; \log_\partial(f, g(\mathbb{T}^n), m) \;+ \; \log_\partial (g, \mathbb{T}^n, m). \end{align}\]
Proof. Faà di Bruno’s formula gives \(\;\forall x \in \mathbb{T}^n\) and \(\forall l \in \mathbb{N},\) \[\begin{align} \partial_j^l f(g(x)) &= \sum_{} {\frac{l!}{a_{1}!\,a_{2}!\,\cdots \,a_{l}!}}\cdot f^{(a_{1}+\cdots +a_{l})}(g(x))\cdot \prod _{t=1}^{l}\left({\frac{\partial_j^t g^{}(x)}{t!}}\right)^{a_{t}}, \end{align}\] and therefore \[\begin{align} |\partial_j^l f(g(x))| &\le \sum_{} {\frac{l!}{a_{1}!\,a_{2}!\,\cdots \,a_{l}!}} \left| f^{(a_{1}+\cdots +a_{l})}(g(x)) \right| \prod _{t=1}^{l} \left|{\frac{\partial_j^t g^{}(x)}{t!}}\right|^{a_{t}}, \end{align}\] where the sum is over \((a_1,\cdots, a_l)\in\mathbb{Z}_{\ge 0}^l\) such that \(\sum_{t=1}^l t\cdot a_t = l\). There are at most \(l^l\) such \(l\)-tuples. The first factor is at most \(l^l\). The second factor is at most \(\max_{} \{\;| f^{(l')} (y) | \; | \;l'\in [l] , y\in g(\mathbb{T}^n)\}\). The last factor is at most \(\prod_{t=1}^l | \partial_j^t g(x) |^{a_t}\). Therefore, we have \[\begin{align} |\partial^l _j f(g(x)) | &\le l^{2l} \max_{\substack{ y \in g(\mathbb{T}^n),\;l' \in[l]}} |f^{(l')} (y)| \prod_{t=1}^l | \partial_j^t g(x) |^{a_t}. \end{align}\] By taking the log of both sides, \[\begin{align} \log |\partial^l _j f(g(x)) | &\le 2l \log l + \log \max_{\substack{ y \in g(\mathbb{T}^n),\;l' \in[l]}} |f^{(l')} (y)| + \sum_{t =1}^m \log | \partial_j^t g(x) |^{a_t} \nonumber \\ & = 2l \log l + \max_{\substack{ y \in g(\mathbb{T}^n),\;l' \in[l]}} \log |f^{(l')} (y)| + \sum_{t =1}^m \log | \partial_j^t g(x) |^{a_t}. \label{eq:logpartial95composition95main95eq} \end{align}\tag{43}\] We bound the second term of 43 as follows: \[\begin{align} \;\max_{\substack{ y \in g(\mathbb{T}^n),\;l' \in[l]}} \log |f^{(l')} (y)| \nonumber \; \le & \; \max _{\;l' \in[l]} \max_{\substack{ y \in g(\mathbb{T}^n)}} (\log |f^{(l')} (y)|)^+ \nonumber \\ = & \; \max_{l' \in [l] } \;l' \; \max_{\substack{ y \in g(\mathbb{T}^n)}} \frac{(\log |f^{(l')} (y)|)^+}{l'} \nonumber \\ \le & \; \left( \max_{l'' \in [l] } \;l''\right) \;\left( \max_{\substack{ y \in g(\mathbb{T}^n) , \;l'\in[l]}} \frac{(\log |f^{(l')} (y)|)^+}{l'} \right) \nonumber \\ = & \;l \max_{\substack{ y \in g(\mathbb{T}^n),\;l' \in[l]}} \frac{(\log |f^{(l')} (y)|)^+}{l'} \\ = & \;l \log_\partial ( f, g(\mathbb{T}^n), l). \label{eq:logpartial95composition95f} \end{align}\tag{44}\] Since \(\log\) is an increasing function and \(\sum_{t=1}^l t \;a_t /l =1\), we bound the third term of 43 as follows: \[\begin{align} &\;\sum_{t =1}^l \log | \partial_j^t g(x) |^{a_t} \nonumber \\ = & \; l \sum_{t =1}^l \frac{t a_t}{l} \frac{\log | \partial_j^t g(x)| }{t} \nonumber \\ \le& \; l \max_{t\in[l]} \frac{\log | \partial_j^t g(x)| }{t} \nonumber \\ \le & \; l \max_{t\in[l],y \in \mathbb{T}^n} \frac{\log | \partial_j^t g(y)| }{t} \; \le \;l \log_\partial (g, \mathbb{T}^n, l) \label{eq:logpartial95composition95g} . \end{align}\tag{45}\] By inserting 44 and 45 into 43 and dividing by \(l\), we get \[\begin{align} \frac{\log|\partial _j^l f(g(x))|}{l} \le 2 \log l + \log_\partial (f, g(\mathbb{T}^n), l) + \log_\partial (g, \mathbb{T}^n, l). \end{align}\] Since the RHS is always nonnegative, we have \[\begin{align} \frac{(\log|\partial _j^l f(g(x))|)^+}{l} \le 2 \log l + \log_\partial (f, g(\mathbb{T}^n), l) + \log_\partial (g, \mathbb{T}^n, l). \end{align}\] The inequality still holds if we take the maximum over \(x \in \mathbb{T}^n\), \(j\in[n]\), and \(l \in [m]\) on both sides. Therefore, \[\begin{align} \log_\partial(f\circ g, \mathbb{T}^n , m) & = \\ \max_{j\in[n], l \in[m], x\in \mathbb{T}^n } \frac{(\log|\partial _j^l f(g(x))|)^+}{l} &\le 2 \log m + \max_{l \in[m]} \log_\partial (f, g(\mathbb{T}^n), l) + \max_{l \in[m]} \log_\partial (g, \mathbb{T}^n, l) \\ &= 2\log m + \log_\partial (f, g(\mathbb{T}^n),m ) + \log_\partial (g, \mathbb{T}^n, m). \end{align}\] ◻
Lemma 22. We have \[\begin{align} \log_\partial (\text{Bump}, \mathbb{R},m) &\le \log O(m^4), \\ \log_\partial (\text{Sat}_{\alpha,\beta} \mathbb{R},m), \log_\partial (\text{Cut}_{\alpha,\beta}, \mathbb{R},m),\log_\partial (\text{Bar}_{\beta}, \mathbb{R},m) &\le \log O(1/\beta + m^4), \end{align}\]
Proof. The following is a well-known smooth function: \[f_0(x):= \begin{cases} \;\exp \left(-1/x\right) & x> 0 \\ \;0 &x\le 0. \end{cases}\] We use the fact that \(\text{Bump}(x) = f_0(x) f_0(1-x)\) and the Leibniz rule \[\begin{align} (fg)^{(m)} = \sum_{j = 0}^m \begin{pmatrix} m \\ j \end{pmatrix} f^{(j)} g^{(m-j)} \end{align}\] to get an upper bound \[\begin{align} \label{eq:bump95function95nth95diff95summation} \max_{x\in \mathbb{R}} |\text{Bump}^{(m)} (x) | \le \sum_{j = 0}^m \begin{pmatrix} m \\ j \end{pmatrix} \max_{x\in \mathbb{R}} | f_0^{(i)}(x)| \; \max_{x\in \mathbb{R}} | f_0^{(m-i)}(x) | . \end{align}\tag{46}\] We are left with upper bounding \(|f_0^{(j)}|\). Note that for \(x>0\), \(f_0'(x) =(1/x)^2 e^{-1/x} = t^2 e^{-t}\), where \(t:=1/x.\)
We prove by induction that there exists a polynomial such that \(f_0^{(j)}(x) = p_j(t) e^{-t}\), for all \(j \in \mathbb{N}\). We showed above the statement for \(j=1.\) Suppose the hypothesis is true for \(j\in \mathbb{N}\). Then for \(j+1\): \[\begin{align} f^{(j+1)}_0 (x) = \frac{dt}{dx} \frac{d}{dt} (p_{j}(t) e^{-t}) =-t^2\left(p_j'(t) - p_j(t)\right) e^{-t}:=p_{j+1}(t)e^{-t}. \end{align}\] Therefore, the induction hypothesis is true for all \(j \in \mathbb{N}\).
The maximum of \(t^j e^{-t}\), for \(t>0\), is \((j/e)^j\) at the maximizer \(t = j\). Observe that this quantity increases as a function of \(j\) for \(j\ge 1.\) For an arbitrary polynomial \(p(t):=\sum_{j=0}^{\deg p} a_j t^j,\) define sum of the absolute values of its coefficients \(\| p\|_c:= \sum_{j=0}^{\deg p} |a_j|\). Then, if \(p\) does not have a constant term, \[\begin{align} \max_{t>0}| p_j(t)e^{-t}| \le \sum_{j = 1}^{\deg p_j} |a_j| (j/e)^j\le \|p\|_c (\deg p /e)^{\deg p}. \end{align}\] Therefore, \[\begin{align} \max_{t >0} \left| f_0^{(j)}(t) \right| \le \|p_j\| \left(\frac{\deg p_j}{e}\right)^{\deg p_j}. \end{align}\] From the recursive formula \(p_{j+1} (t) = t^2(p_j(t) - p'_j(t))\) with the initial condition \(p_1 = t^2\), we see that \(\deg p_j = 2j\) and \(p_j\) is without a constant term, for all \(j\in \mathbb{N}.\) From the recursive formula, we also have \[\begin{align} \|p_j\|_c &\le \|t^2 p_j\|_c + \|t^2 p'_j\|_c \\ &\le \|p_{j-1}\| + 2(j-1)\|p_{j-1}\| \\& = (2j-1)\|p_{j-1}\| \\ & \le (2j-1)(2j-3)\cdots 3\cdot 1\cdot \| p_{1}\| \end{align}\] Therefore, \(\|p_j\| <(2j)^j\), and we have \[\begin{align} \max_{x \in\mathbb{R}} |f_0^{(j)}(x)| = \max_{t>0} | p(t) e^{-t}| \le (2j^2/e)^j \le (2m^2/e)^m, \end{align}\] for \(0\le j\le m\). Finally, we can plug this inequality into 46 and use the binomial identity \(\sum_{j=0}^m \begin{pmatrix} m \\ j \end{pmatrix} = 2^m\) to get \(\max_{x\in \mathbb{R}}|\text{Bump}^{(m)}(x)| \le 2^m(4m^4/e^2)^m,\) and therefore, \[\begin{align} \log_\partial (\text{Bump}, \mathbb{R}, m ) &= \sup_{l\in[m], x \in\mathbb{R} } \frac{(\log|\text{Bump}^{(l)}(x)|)^+ }{l} \\ &\le \max_{l\in[m]}\frac{l\log 2 + l\log(4l^4/e^2) }{l} \\ &= \log O(m^4). \end{align}\]
Since \(\text{Cut}_{\alpha,\beta} '(x) = - \text{Bump}(\frac{x - \alpha}{\beta})/\int_{\alpha}^{\alpha + \beta}\text{Bump}(\frac{y-\alpha}{\beta})\text{d} y\), we have \[\begin{align} \text{Cut}_{\alpha, \beta}^{(m)} (x) = - \frac{1}{\beta^{m-1} } \text{Bump}^{(m-1)} \big(\frac{x-\alpha}{\beta} \big) \bigg/ {\int_\alpha^{\alpha + \beta} \text{Bump}\big((y-\alpha)/\beta \big) \; \text{d} y } . \end{align}\] Therefore, \[\begin{align} \log_\partial(\text{Cut}_{\alpha,\beta},\mathbb{R}, m) = \sup_{l\in[m],x\in\mathbb{R}}\frac{( \log|\text{Cut}^{(l)}_{\alpha, \beta}(x)| )^+ }{l} = \log O(1/\beta + m^4). \end{align}\] Since \(\text{Sat}_{\alpha, \beta}' = \text{Cut}_{\alpha,\beta}\) and \(\text{Bar}_\epsilon' = 1- \text{Cut}_{0,\epsilon}(x)\), we apply a similar analysis to obtain \[\begin{align} \log_\partial(\text{Sat}_{\alpha,\beta},\mathbb{R},m) = \log O(1/\beta + m^4), \\ \log_\partial(\text{Bar}_{\epsilon},\mathbb{R},m) = \log O(1/\epsilon + m^4). \end{align}\] ◻
We define a slightly expanded region of \(\Omega\), \[\begin{align} \Omega' := (1+\epsilon)\Omega = \{ x\in \mathbb{R}^n | \;a_i \cdot x \le b_i(1+\epsilon) , \;\forall i \in [m]\}, \end{align}\] and two associated Schrödinger operators \[\begin{align} \label{eq:DL95hamiltonians} h &:=-\Delta + V \\ h^D &:=-\Delta^D_{\Omega' } + V|_{\Omega'}, \end{align}\tag{47}\] where \(h\) is a Schrödinger operator on \({\mathbb{R}^n}\), and \(h^D\) is its Dirichlet restriction on \(\Omega'\) .
Lemma 23. Suppose \(h^D\) is defined as above. If \(\epsilon = O({\epsilon_0}/{n^2})\), then we have \[\begin{align} |\lambda_0(h^D) - \lambda_0(-\Delta^D_\Omega)| \le O(\epsilon_0). \end{align}\]
Proof. By domain monotonicity (Lemma 9), we have \(\lambda_0(h^D) \le \lambda_0(-\Delta^D_{\Omega})\), and by potential comparison (Lemma 8), we have \(\lambda_0(-\Delta^D_{\Omega'}) \le \lambda_0(h^D)\), giving \[\begin{align} \lambda_0(-\Delta^D_{\Omega'}) \le \lambda_0(h^D) \le \lambda_0(-\Delta^D_{\Omega}). \end{align}\] Additionally, for every eigenfunction \(\psi\) of \(\Delta_{\Omega}\), we have that \(\psi'\) as an eigenfunction of \(\Delta_{\Omega'}\), where \(\psi'(x) = \psi(x/(1+\epsilon))\). Therefore, \[\begin{align} \lambda_0(-\Delta^D_{\Omega'}) = \lambda_0(-\Delta^D_\Omega)/(1+\epsilon)^2\ge \lambda_0(-\Delta^D_\Omega) -O(\epsilon\lambda_0(-\Delta^D_\Omega)), \end{align}\] which gives us a sandwich \[\begin{align} \lambda_0(-\Delta^D_{\Omega})- O(\epsilon\lambda_0(-\Delta^D_\Omega)) \le \lambda_0(h^D) \le \lambda_0(-\Delta^D_{\Omega}). \end{align}\] Hence, we have \[\begin{align} \label{eq:laplacian95nsquare} |\lambda_0(h^D) - \lambda_0(-\Delta^D_\Omega)| \le O(\epsilon\lambda_0(-\Delta^D_\Omega) ) \le O(\epsilon n^2), \end{align}\tag{48}\] where the last inequality is from the domain monotonicity (Lemma 9) of \[\begin{align} B^\infty_{1/2\sqrt n} \subset B^2_1 \subset \Omega. \end{align}\] We set \(\epsilon = O({\epsilon_0}/{n^2})\) and obtain \(|\lambda_0(h^D) - \lambda_0(-\Delta^D_\Omega)| \le O(\epsilon_0).\) ◻
We aim to use Markov’s to establish equivalence between \(h\) and \(h^D\) and eventually claim that \(\lambda_0(-\Delta^D_\Omega)\) and \(\lambda_0(h)\) are close. Suppose \(x \in \partial \Omega'\). Then, there exists \(j\in[m]\) such that \(a_j \cdot x = b_j(1+\epsilon)\). Therefore, \[\begin{align} \label{eq:app95boundary95potential} V(x) \ge \frac{3E}{\mu^6}\text{Bar}_\epsilon(a_j\cdot x-b_j) \ge \frac{3E \;\text{Bar}_\epsilon(\epsilon b_j)}{\mu^6} = \frac{3E}{\mu^6} \qquad \forall\; x \in \partial \Omega. \end{align}\tag{49}\]
Lemma 24. If \(\mu = O(\frac{\epsilon_0}{E (mn^2)^{1/3}})\), then \(h\) and \(h^D\) are \((E,O(\epsilon_0))\)-equivalent.
Proof. We claim that \(L^2(\Omega')\) is an \((E,\epsilon)\)-truncated domain for both \(h^D\) and \(h\). The claim would prove that \(h\) and \(h^D\) are \((E,\epsilon)\)-equivalent, since the identity unitary takes \(L^2(\Omega')\) to itself, and the restrictions of \(h\) and \(h^D\) on \(L^2(\Omega')\) are identical.
For \(h^D\), we have \(\mathcal{D}(h^D)\subset L^2(\Omega')\), and therefore \(L^2(\Omega')\) is an \((E,\epsilon)\)-truncated domain for \(h^D\). Therefore, we only need to show that \(L^2(\Omega')\) is an \((E,\epsilon)\)-truncated domain for \(h\).
The derivation is almost the same as in the proof of Lemma 16, except that we use a different cutoff function \(f\). We will refer to the proof of Lemma 16, when the calculation is the same.
Suppose \(\psi\in L^2 ({\mathbb{R}^n})\) has energy \(h[\psi]\le E\). We define \[\begin{align} \psi^\downarrow:=\sum_{i: \lambda_i \le E/\mu^2 } c_i\psi_i, \end{align}\] where \(\psi_i\) is the \(i\)-th eigenvectors of \(h\) with eigenvalue \(\lambda_i\). We define sets \[\begin{align} A &:=V^{-1}([0,E/\mu^6]) , \\ B &:=V^{-1}([0,2E/\mu^6]). \end{align}\] Then we have \(A\subset B\subset \Omega'\), because of 49 .
The cutoff function \(f:= \text{Cut}_{E/\mu^6,E/\mu^6}\circ V\) satisfies that \[\begin{align} f(x) \begin{cases} = 1 & x \in A \\ \in[0,1] & x \in B\setminus A\\ = 0 & x \notin B. \end{cases} \end{align}\]
We claim that \[\begin{align} \widetilde{\psi}:=\frac{f\psi^\downarrow}{\|f\psi^\downarrow\|} \end{align}\] is an \(\mu\)-truncation of \(\psi\). The claim implies the lemma, since \(\widetilde{\psi}\in L^2 (\Omega')\).
We first show the norm condition of \(\mu\)-truncation. Since \(V(x)\ge E/\mu^6\) for all \(x\) such that \(f(x) = 0\), by Markov’s inequality (see the proof of Lemma 15), we have \[\begin{align} \|\psi^\downarrow - f\psi^\downarrow\|\le \mu^3 \|\psi^\downarrow\|. \end{align}\] By the same derivation that led to 26 , we have \[\begin{align} \|\psi - \widetilde{\psi}\| \le O(\mu) \le O(\epsilon_0). \end{align}\]
For the rest of the proof, we show the energy condition. By the same calculation that led to 28 , we have \[\begin{align} \|f\psi^\downarrow \| h[f\psi^\downarrow\| = \langle \psi^\downarrow | \|\;\nabla f\|^2 |\psi^\downarrow\rangle + \text{Re}\langle f^2 \psi^\downarrow | h\psi ^\downarrow \rangle . \end{align}\] We upper bound the two terms similarly to the proof of Lemma 16. We have \[\begin{align} \langle \psi^\downarrow | \|\;\nabla f\|^2 |\psi^\downarrow\rangle =&\int_{x \in {\mathbb{R}^n}} |\psi^\downarrow(x)|^2 \|\nabla f(x)\|^2 \;\text{d} x \\ \le & \; \max_{x \notin A} \|\nabla f(x)\|^2 \int_{x\notin A} |\psi^\downarrow(x)|^2 \;\text{d} x \\ \le & \; \max_{x \notin A} | \text{Cut}'_{E/\mu^6,E/\mu^6} ( V(x))|^2 \;\| \nabla V(x)\|^2 \mu^6 \\ \le & \; O\left(\frac{\mu^{12}}{E^2} \cdot \frac{m ^2 E^2}{\epsilon^2\mu ^{12}}\cdot \mu^6 \right) \\ = & \;O \left( \frac{m^2 n^4 \mu^{6} }{\epsilon_0^2}\right) \\ \le & \; O(\epsilon^4/ E^4). \end{align}\] Also, by the calculation that led to 34 , we have \[\begin{align} \text{Re}\langle f^2\psi^\downarrow|h\psi^\downarrow\rangle \le h[\psi] + \mu E \le h[\psi] + \epsilon_0. \end{align}\] Therefore, we have \[\begin{align} \|f\psi^\downarrow \| h[f\psi^\downarrow\| \le h[\psi] + O(\epsilon_0) + O(\epsilon_0^4 E^{4}). \end{align}\] Finally, since \[\begin{align} \|f\psi^\downarrow\| \;\ge \;\|\psi^\downarrow\| -\|(1 - f) \psi^\downarrow\| \;\ge \; 1 - \mu - \mu^3, \end{align}\] we have \[\begin{align} h[\widetilde{\psi}] = h[f\psi^\downarrow] \le \frac{ h[\psi] + O(\epsilon_0) + O(\epsilon_0^4 E^{4}) }{ 1 - \mu - \mu^3} \le h[\psi] + O(\epsilon_0) . \end{align}\] ◻