|
| 1 | +\documentclass[11pt,a4paper]{article} |
| 2 | +\usepackage[utf8]{inputenc} |
| 3 | +\usepackage[T1]{fontenc} |
| 4 | +\usepackage{amsmath, amssymb, amsfonts, bm} |
| 5 | +\usepackage{graphicx} |
| 6 | +\usepackage{hyperref} |
| 7 | +\usepackage{physics} |
| 8 | +\usepackage{braket} |
| 9 | +\usepackage{geometry} |
| 10 | +\geometry{margin=2.5cm} |
| 11 | +\usepackage{setspace} |
| 12 | +\onehalfspacing |
| 13 | + |
| 14 | +\title{Lindlad, Noise and possible links with PMMs} |
| 15 | +\author{Collaboration on Noise, Open quantum systems and more} |
| 16 | +\date{November 2} |
| 17 | + |
| 18 | +\begin{document} |
| 19 | +\maketitle |
| 20 | + |
| 21 | +\section{Introduction} |
| 22 | + |
| 23 | +This note provides an extremely abridged, and perhaps overly |
| 24 | +practical, introduction to open quantum systems. The goal is to |
| 25 | +provide the basis for understanding and implementing an open quantum |
| 26 | +system using the Lindblad master equation. And then hopefully link this with PMMs. |
| 27 | + |
| 28 | +\section{Open?} |
| 29 | + |
| 30 | +In quantum mechanics, a system is considered \emph{closed} if it is |
| 31 | +isolated from its environment. In contrast, an \emph{open quantum |
| 32 | +system} interacts with its environment, which can cause the system to |
| 33 | +lose coherence and entanglement. |
| 34 | + |
| 35 | +All this boils down to whether or not energy is conserved. In a closed |
| 36 | +system, energy is conserved, while in an open system, energy is not |
| 37 | +conserved. |
| 38 | + |
| 39 | + |
| 40 | +\section{Density Matrices} |
| 41 | + |
| 42 | +As opposed to the wavefunction (a so-called pure state) in closed |
| 43 | +(Hermitian) quantum mechanics, the density matrix uniquely describes |
| 44 | +the state of a quantum system in an open system, as well as in closed |
| 45 | +systems. |
| 46 | + |
| 47 | +With pure states, the state of a quantum system is deterministic and |
| 48 | +leads to probabilities of outcomes of measurements. In contrast, what |
| 49 | +if the state itself is uncertain and therefore probabilistic? This is |
| 50 | +what density matrices describe.\footnote{A natural question might be: |
| 51 | +“What about probabilistic mixtures of density matrices?” Luckily, such |
| 52 | +systems are also just described by density matrices.} |
| 53 | + |
| 54 | +The important properties are summarized here: |
| 55 | +\begin{itemize} |
| 56 | + \item The density matrix of a pure state $\ket{\psi}$ is $\rho = \ket{\psi}\bra{\psi}$. |
| 57 | + \item $\rho$ is Hermitian, positive semi-definite (all eigenvalues non-negative), and has unit trace: |
| 58 | + \[ |
| 59 | + \rho = \rho^\dagger, \quad \lambda(\rho) \ge 0, \quad \Tr(\rho) = 1. |
| 60 | + \] |
| 61 | + \item Given a probabilistic mixture of states with probabilities $\{p_i\}$ and states $\{\ket{\psi_i}\}$, |
| 62 | + \[ |
| 63 | + \rho = \sum_i p_i \ket{\psi_i}\bra{\psi_i}. |
| 64 | + \] |
| 65 | + \item For a $d \times d$ density matrix $\rho$, there exists an orthonormal basis $\{\ket{b_i}\}$ such that |
| 66 | + \[ |
| 67 | + \rho = \sum_i \lambda_i \ket{b_i}\bra{b_i}, |
| 68 | + \] |
| 69 | + where $\lambda_i$ are probabilities. |
| 70 | + \item The expectation value of an observable $X$ is |
| 71 | + \[ |
| 72 | + \langle X \rangle = \Tr(\rho X) = \sum_i \lambda_i \bra{b_i} X \ket{b_i}. |
| 73 | + \] |
| 74 | + \item The time evolution in a closed system with Hamiltonian $H$ is |
| 75 | + \[ |
| 76 | + \rho(t) = e^{-iHt} \rho(0) e^{iHt}. |
| 77 | + \] |
| 78 | +\end{itemize} |
| 79 | + |
| 80 | +\subsection{One Qubit Example} |
| 81 | + |
| 82 | +The codes for these various examples are included as a separate Python code.. |
| 83 | +Consider a single qubit system that has a 50\% chance of being in |
| 84 | +$\ket{0}$, a 25\% chance of being in $\ket{+} = |
| 85 | +\frac{1}{\sqrt{2}}(\ket{0}+\ket{1})$, and a 25\% chance of being in |
| 86 | +$\ket{-} = \frac{1}{\sqrt{2}}(\ket{0}-\ket{1})$. Then |
| 87 | + |
| 88 | +\[ |
| 89 | +\rho = \frac{1}{2}\ket{0}\bra{0} + \frac{1}{4}\ket{+}\bra{+} + \frac{1}{4}\ket{-}\bra{-} = |
| 90 | +\begin{bmatrix} |
| 91 | +3/4 & 0 \\[4pt] |
| 92 | +0 & 1/4 |
| 93 | +\end{bmatrix}. |
| 94 | +\] |
| 95 | + |
| 96 | +\subsection{Two Qubit Example} |
| 97 | + |
| 98 | +Consider a system of two qubits that has a 90\% chance of being in the Bell state |
| 99 | +\[ |
| 100 | +\ket{B_+} = \frac{1}{\sqrt{2}}(\ket{00} + \ket{11}). |
| 101 | +\] |
| 102 | +The remaining 10\% is distributed among $\ket{01}$, $\ket{10}$, and $\ket{B_-} = \frac{1}{\sqrt{2}}(\ket{00}-\ket{11})$ in the ratio 3:2:1. Then |
| 103 | +\[ |
| 104 | +\rho = \frac{9}{10}\ket{B_+}\bra{B_+} + \frac{1}{20}\ket{01}\bra{01} + \frac{1}{30}\ket{10}\bra{10} + \frac{1}{60}\ket{B_-}\bra{B_-}. |
| 105 | +\] |
| 106 | + |
| 107 | +\section{Subsystems and Partial Traces} |
| 108 | + |
| 109 | +An open quantum system is a subsystem of a closed system composed of system + environment. To describe only the system, we trace out the environment: |
| 110 | +\[ |
| 111 | +\rho_A = \Tr_B(\rho). |
| 112 | +\] |
| 113 | +For a pure state $\ket{\psi} = \ket{a}\otimes\ket{b}$, the reduced density matrix is |
| 114 | +\[ |
| 115 | +\rho_A = \sum_j (I_A \otimes \bra{j}) \rho (I_A \otimes \ket{j}). |
| 116 | +\] |
| 117 | + |
| 118 | +\subsection{Two Qubit Example} |
| 119 | + |
| 120 | +From the two-qubit example, |
| 121 | +\[ |
| 122 | +\rho_1 = \Tr_2(\rho) = |
| 123 | +\begin{bmatrix} |
| 124 | +61/120 & 0 \\[4pt] |
| 125 | +0 & 59/120 |
| 126 | +\end{bmatrix}, |
| 127 | +\qquad |
| 128 | +\rho_2 = \Tr_1(\rho) = |
| 129 | +\begin{bmatrix} |
| 130 | +59/120 & 0 \\[4pt] |
| 131 | +0 & 61/120 |
| 132 | +\end{bmatrix}. |
| 133 | +\] |
| 134 | + |
| 135 | +\section{von Neumann Equation} |
| 136 | + |
| 137 | +For the total density matrix $\rho_T$ of system plus environment with total Hamiltonian $H_T$, |
| 138 | +\[ |
| 139 | +\frac{d\rho_T}{dt} = -i[H_T, \rho_T]. |
| 140 | +\] |
| 141 | + |
| 142 | +\section{Lindblad Master Equation} |
| 143 | + |
| 144 | +Writing $H_T = H + H_E + H_I$, where $H$ is the system Hamiltonian, $H_E$ the environment Hamiltonian, and $H_I$ their interaction, tracing out the environment and assuming weak coupling yields: |
| 145 | +\[ |
| 146 | +\frac{d\rho}{dt} = -i[H,\rho] + \sum_k \gamma_k \left( L_k \rho L_k^\dagger - \frac{1}{2}\{L_k^\dagger L_k, \rho\} \right), |
| 147 | +\] |
| 148 | +where $\rho$ is the reduced density matrix, $\gamma_k \ge 0$ are decay rates, and $L_k$ are jump operators. |
| 149 | + |
| 150 | +The Heisenberg picture form is |
| 151 | +\[ |
| 152 | +\frac{dX}{dt} = i[H,X] + \sum_k \gamma_k \left( L_k^\dagger X L_k - \frac{1}{2}\{L_k^\dagger L_k, X\} \right), |
| 153 | +\] |
| 154 | +and the identity operator satisfies $\frac{dI}{dt}=0$. |
| 155 | + |
| 156 | +\subsection{Liouvillian Superoperator} |
| 157 | + |
| 158 | +Define the superoperator $\mathcal{L}$ by |
| 159 | +\[ |
| 160 | +\mathcal{L}[\rho] = -i[H,\rho] + \sum_k \gamma_k \left( L_k\rho L_k^\dagger - \frac{1}{2}\{L_k^\dagger L_k, \rho\} \right), |
| 161 | +\] |
| 162 | +so that |
| 163 | +\[ |
| 164 | +\frac{d\rho}{dt} = \mathcal{L}[\rho]. |
| 165 | +\] |
| 166 | + |
| 167 | +\subsection{Vectorization: The Fock–Liouville Space} |
| 168 | + |
| 169 | +Define $\ket{A\!\rangle\!\rangle}$ as the column-stacked vectorization of an operator $A$. The inner product is |
| 170 | +\[ |
| 171 | +\braket{\!\braket{A|B}} = \Tr(A^\dagger B), |
| 172 | +\] |
| 173 | +and |
| 174 | +\[ |
| 175 | +\frac{d}{dt}\ket{\!\braket{\rho}} = \hat{\mathcal{L}}\ket{\!\braket{\rho}}, |
| 176 | +\] |
| 177 | +with solution |
| 178 | +\[ |
| 179 | +\ket{\!\braket{\rho(t)}} = e^{\hat{\mathcal{L}}t} \ket{\!\braket{\rho(0)}}. |
| 180 | +\] |
| 181 | +Diagonalizing $\hat{\mathcal{L}}$ gives eigenvalues $\lambda_i$ and eigenvectors $\ket{\!\braket{r_i}}$ and $\bra{\!\braket{l_i}}$, allowing |
| 182 | +\[ |
| 183 | +\ket{\!\braket{\rho(t)}} = \sum_i e^{\lambda_i t} \braket{\!\braket{l_i|\rho(0)}} \ket{\!\braket{r_i}}. |
| 184 | +\] |
| 185 | + |
| 186 | +\subsection{Steady State} |
| 187 | + |
| 188 | +The steady state $\ket{\!\braket{\rho_{\text{ss}}}}$ satisfies |
| 189 | +\[ |
| 190 | +\hat{\mathcal{L}}\ket{\!\braket{\rho_{\text{ss}}}} = 0. |
| 191 | +\] |
| 192 | + |
| 193 | +\subsection{Harmonic Oscillator Coupled to a Bath} |
| 194 | + |
| 195 | +For $H = \omega a^\dagger a$ and jump operators $L_1 = \sqrt{\gamma(\tau+1)}\,a$, $L_2 = \sqrt{\gamma\tau}\,a^\dagger$, the Lindblad equation reads: |
| 196 | +\[ |
| 197 | +\frac{d\rho}{dt} = -i[\omega a^\dagger a, \rho] + \gamma(\tau+1)\left[a\rho a^\dagger - \frac{1}{2}\{a^\dagger a,\rho\}\right] |
| 198 | ++ \gamma\tau\left[a^\dagger\rho a - \frac{1}{2}\{aa^\dagger,\rho\}\right]. |
| 199 | +\] |
| 200 | + |
| 201 | +See separate Python code |
| 202 | + |
| 203 | +\section{Project 1: studies of Markovian and non-Markovian noise} |
| 204 | + |
| 205 | +More text to come here. |
| 206 | + |
| 207 | +\section{Project 2: roadmap for beyond zero noise extrapolation for quantum computing using PMMs} |
| 208 | + |
| 209 | +{\bf This text will be improved upon.} |
| 210 | + |
| 211 | + |
| 212 | +In the Noisy Intermediate Scale Quantum (NISQ) computing era, error |
| 213 | +mitigation is vital to obtain meaningful results from quantum |
| 214 | +algorithms running on existing hardware. This project aims to develop |
| 215 | +a Parametric Matrix Model (PMM) which learns the underlying open |
| 216 | +quantum system nature of today’s noisy quantum computers in order to |
| 217 | +better understand the sources of error and directly extrapolate to |
| 218 | +results with zero noise. |
| 219 | + |
| 220 | + |
| 221 | + |
| 222 | +Quantum gates are often modeled idealistically as unitary operators |
| 223 | +acting on a quantum state. However, in practice, quantum gates are |
| 224 | +subject to noise which can be modeled as a quantum channel. The noise |
| 225 | +can be characterized by a superoperator acting on the density matrix |
| 226 | +of the quantum state. There are many equivalent formalisms including |
| 227 | +the Liouville superoperator and the Kraus operator |
| 228 | +sum~\cite{wood2015tensor}. |
| 229 | + |
| 230 | +A simplified mathematical overview is that if a noiseless operation can be represented as a unitary operator \( U \) acting on a pure state \( |\psi\rangle \), |
| 231 | +\[ |
| 232 | +|\psi\rangle \rightarrow U |\psi\rangle, |
| 233 | +\] |
| 234 | +then a noisy operation can be represented as a superoperator \( S \) acting on a vectorized density matrix \( |\rho\rangle\rangle \), |
| 235 | +\[ |
| 236 | +|\rho\rangle\rangle \rightarrow S |\rho\rangle\rangle. |
| 237 | +\] |
| 238 | +Vectorization is a central operation in the Liouville formalism. The vectorization of a matrix \( A \) is formed by stacking the columns of \( A \) into a single column vector. |
| 239 | + |
| 240 | +Any noiseless quantum circuit can be represented by a product of unitary operators \( U_1, U_2, \ldots, U_n \) acting on the initial state \( |\psi_0\rangle \), |
| 241 | +\[ |
| 242 | +|\psi_0\rangle \rightarrow U_n U_{n-1} \cdots U_1 |\psi_0\rangle. |
| 243 | +\] |
| 244 | +The corresponding noisy circuit can be represented by a product of superoperators \( S_1, S_2, \ldots, S_n \) acting on the vectorized initial density matrix \( |\rho_0\rangle\rangle \), |
| 245 | +\[ |
| 246 | +|\rho_0\rangle\rangle \rightarrow S_n S_{n-1} \cdots S_1 |\rho_0\rangle\rangle. |
| 247 | +\] |
| 248 | + |
| 249 | +Unitary operators have many nice properties such as: |
| 250 | +\begin{itemize} |
| 251 | + \item \( U^\dagger U = U U^\dagger = I \) (unitarity), |
| 252 | + \item \( \lambda(U) = \{ e^{i\theta} : \theta \in \mathbb{R} \} \) (eigenvalues on the unit circle), |
| 253 | + \item \( \|Ux\|_2 = \|x\|_2 \) (norm preservation). |
| 254 | +\end{itemize} |
| 255 | + |
| 256 | +In contrast, superoperators for quantum channels must be \emph{completely positive and trace preserving} (CPTP). For a map \( \Phi : L(H) \rightarrow L(H) \), this means: |
| 257 | +\begin{itemize} |
| 258 | + \item For any positive semidefinite operator \( X \), \( \Phi(X) \) is also positive semidefinite. |
| 259 | + \item For any Hilbert space \( H' \), \( \Phi \otimes I_{H'} \) is also positive. |
| 260 | + \item \( \mathrm{tr}(\Phi(X)) = \mathrm{tr}(X) \) (trace preserving). |
| 261 | +\end{itemize} |
| 262 | + |
| 263 | +Completely positive maps are always positive, and every CPTP map |
| 264 | +represents a valid quantum channel. The matrix representation of a |
| 265 | +superoperator can be reshaped into a positive semidefinite matrix, |
| 266 | +guaranteeing complete positivity. Other properties, such as trace |
| 267 | +preservation and Hermiticity preservation, arise from related |
| 268 | +transformations. |
| 269 | + |
| 270 | +Choosing a formalism and parameterization that respects these |
| 271 | +properties is crucial for any PMM that aims to learn the underlying |
| 272 | +noise model of a quantum computer. |
| 273 | + |
| 274 | +A possible goal of this project is to develop a PMM that can learn (perhaps a |
| 275 | +low-dimensional representation of) the underlying open quantum system |
| 276 | +nature of a real quantum circuit on a real quantum computer. Using |
| 277 | +this, we aim to demonstrate zero-noise extrapolation using the PMM |
| 278 | +that is more physically constrained and potentially more accurate than |
| 279 | +current state-of-the-art ZNE (Zero-noise extrapolation) techniques. |
| 280 | + |
| 281 | + |
| 282 | +By learning the underlying noise model of each gate on a physical |
| 283 | +quantum computer, we can potentially use this information to construct |
| 284 | +circuits that are more robust to noise. |
| 285 | + |
| 286 | + |
| 287 | +\subsection{Implementing Hamiltonians onto a Circuit} |
| 288 | + |
| 289 | +Here one could use Qiskit to implement a Hamiltonian onto a circuit. As an examnple, we could start with a classic, the Ising Model: |
| 290 | +\[ |
| 291 | +H = B \sum_i X_i + J \sum_i Z_i Z_{i+1}, |
| 292 | +\] |
| 293 | +where \( X \) and \( Z \) are Pauli operators. See Ref.~\cite{smith2019simulating} for guidance on encoding many-body Hamiltonians. |
| 294 | + |
| 295 | +\subsection{Zero-Noise Extrapolation} |
| 296 | + |
| 297 | +One could then |
| 298 | +study sources of noise during computation and the presence of |
| 299 | +different noise channels. A useful reference is Nielsen and Chuang’s |
| 300 | +\emph{Quantum Computation and Quantum Information}. The depolarizing |
| 301 | +noise channel is typically used in ZNE. Implement ZNE following |
| 302 | +Ref.~\cite{giurgica2020digital}, or use the \texttt{mitiq} Python |
| 303 | +library (\url{https://mitiq.readthedocs.io/}). |
| 304 | + |
| 305 | +Note to self: need to perform a literature search on ZNE techniques enhanced |
| 306 | +with neural networks (recent results within the last 1–2 years) to |
| 307 | +possibly include as comparison. |
| 308 | + |
| 309 | +\subsection{Exact Simulation of Simple Circuit} |
| 310 | + |
| 311 | +One could then implement an exact unitary simulation of a simple |
| 312 | +circuit, such as Bell-state generation and measurement. Then extend it |
| 313 | +to an exact noisy simulation. |
| 314 | + |
| 315 | +\subsection{PMM for Noiseless Simple Circuit} |
| 316 | + |
| 317 | +The next step is to implement a PMM for the noiseless circuit and |
| 318 | +train it to reproduce observables or states. Explore larger circuits |
| 319 | +to study potential dimensionality reduction. |
| 320 | + |
| 321 | +\subsection{PMM for Noisy Simple Circuit} |
| 322 | + |
| 323 | + |
| 324 | +Develop a PMM for the noisy circuit. Demonstrate that it can learn the |
| 325 | +underlying noise model and extrapolate to the noiseless case. Explore |
| 326 | +larger circuits as before. |
| 327 | + |
| 328 | +\subsection{Comparison with Traditional ZNE} |
| 329 | + |
| 330 | +Compare PMM performance with traditional ZNE methods for the simple |
| 331 | +circuit. This may involve real quantum hardware or realistic noise |
| 332 | +simulators. Here one could |
| 333 | +demonstrate the PMM method on real quantum hardware, comparing it with existing ZNE techniques such as those in \texttt{mitiq}. One could extend this to include tensor-network approaches (MPS, DMRG, TEBD) to reduce computational complexity. |
| 334 | + |
| 335 | +\section*{References} |
| 336 | + |
| 337 | +\begin{thebibliography}{9} |
| 338 | + |
| 339 | +\bibitem{manzano2019} D. Manzano, \emph{A Short Introduction to the Lindblad Master Equation}, \href{https://arxiv.org/abs/1906.04478}{arXiv:1906.04478}, 2020. |
| 340 | +\bibitem{yuen2022} H. Yuen, \emph{Lecture 2: Mixed States and Density Matrices}, \href{https://www.henryyuen.net/spring2022/lec2-mixed-states.pdf}{2022}. |
| 341 | + |
| 342 | +\bibitem{giurgica2020digital} |
| 343 | +T.~Giurgica-Tiron, Y.~Hindy, R.~LaRose, A.~Mari, and W.~J.~Zeng, |
| 344 | +\newblock ``Digital zero noise extrapolation for quantum error mitigation,'' |
| 345 | +\newblock in \emph{IEEE International Conference on Quantum Computing and Engineering (QCE)}, pp. 306–316, 2020. |
| 346 | + |
| 347 | +\bibitem{smith2019simulating} |
| 348 | +A.~Smith, M.~S.~Kim, F.~Pollmann, and J.~Knolle, |
| 349 | +\newblock ``Simulating quantum many-body dynamics on a current digital quantum computer,'' |
| 350 | +\newblock \emph{npj Quantum Information}, 5(1):106, 2019. |
| 351 | + |
| 352 | +\bibitem{wood2015tensor} |
| 353 | +C.~J.~Wood, J.~D.~Biamonte, and D.~G.~Cory, |
| 354 | +\newblock ``Tensor networks and graphical calculus for open quantum systems,'' 2015. |
| 355 | + |
| 356 | +\end{thebibliography} |
| 357 | + |
| 358 | +\end{document} |
| 359 | + |
| 360 | + |
0 commit comments