Fact-checked by Grok 2 weeks ago

Dephasing

Dephasing is a type of in which a loses between its superposition states due to environmental interactions, leading to the of off-diagonal elements in the 's while preserving the diagonal populations corresponding to energy eigenstates. This process can be modeled as a , often represented by Kraus operators that act diagonally in the 's preferred basis, such as the computational basis for qubits. In physical realizations, dephasing arises from mechanisms like fluctuating magnetic fields, charge noise, or spin interactions, without involving energy exchange between the and bath. In and , dephasing is a primary source of error that limits coherence times and , often dominating over other processes like amplitude in systems such as superconducting qubits. It disrupts quantum interference essential for algorithms like Grover's search or Shor's factoring, necessitating error correction techniques or dynamical decoupling pulses to mitigate its effects. Research focuses on characterizing dephasing rates through metrics like the phase memory time T_M, which can be influenced by factors such as , lattice relaxation, and hyperfine coupling, with typical values ranging from microseconds in solid-state systems to milliseconds in trapped ions. Beyond computing, dephasing plays a key role in spectroscopic techniques, where it contributes to linewidth broadening in nuclear magnetic resonance (NMR) via T2 relaxation, causing the free induction decay (FID) signal to decay and resulting in broader spectral peaks. In optical and vibrational spectroscopy, it manifests as pure dephasing of electronic or molecular states, with rates scaling from $10^{12} s^{-1} for vibrations to $10^{14} s^{-1} for electronic transitions in condensed phases, affecting signal resolution and quantum control experiments. Understanding and controlling dephasing is thus crucial for advancing quantum technologies and precision measurements across physics and chemistry.

Definition and Basics

Definition

Dephasing is a fundamental process in whereby the relative phase information in a superposition of quantum states is progressively lost due to environmental perturbations, resulting in the degradation of quantum coherence and the emergence of classical-like behavior without net energy exchange between the system and its surroundings. This non-dissipative interaction leads to the randomization of phase relationships among quantum amplitudes, effectively suppressing interference effects that are hallmarks of quantum superpositions. The concept of dephasing was first rigorously described in the mid-20th century within the framework of (NMR), where it manifests as the transverse relaxation of spin ensembles. In their seminal 1948 paper, Bloembergen, Purcell, and Pound introduced the phenomenological relaxation times T1 and T2 to model these processes, with T2 specifically capturing the dephasing-induced decay of transverse magnetization. Key developments extended this understanding to in the 1970s, incorporating dephasing into the theory of open quantum systems and Markovian dynamics, as formalized in works like the Lindblad master equation. Pure dephasing refers to a specific form of this process characterized by decoherence that occurs without any between distinct levels of the , arising instead from fluctuating shifts due to . In this regime, the diagonal elements of the 's remain unchanged, while off-diagonal elements decay, marking the loss of . A central quantity associated with dephasing is the time T2, which quantifies the timescale over which information persists before environmental interactions cause significant loss. This is distinguished from the relaxation time T1, which governs the return of the to via dissipation; in many s, T2 is limited by both T1 processes and additional pure dephasing contributions, often satisfying T2 ≤ 2T1.

Quantum Mechanical Description

In , dephasing describes the randomization of relative in a superposition state of a quantum system due to fluctuating environmental fields, resulting in the loss of quantum without altering the populations of the basis states. For a two-level system, such as a , the initial pure state can be expressed as |\psi\rangle = \alpha |0\rangle + \beta |1\rangle, where \alpha and \beta are complex coefficients satisfying |\alpha|^2 + |\beta|^2 = 1. Under dephasing, this evolves into |\psi(t)\rangle = \alpha |0\rangle + \beta e^{i\phi(t)} |1\rangle, where \phi(t) represents a phase shift accumulated over time from noise sources. This phase randomization disrupts the between the superposition components, causing the system to behave more classically as the quantum encoded in the phases is lost. A concrete example occurs in a two-level subjected to , where the \phi(t) accumulates randomly according to the noise power spectrum. The , quantified by the off-diagonal element of the , then averages over noise realizations to \langle e^{i\phi(t)} \rangle = e^{-\langle \phi^2(t) \rangle / 2}, which decays to zero for sufficient time, eliminating interference patterns observable in measurements like . This process exemplifies how dephasing erodes —the ability of the system to retain and utilize the delicate phase relationships essential for quantum computations and simulations. The provides a geometric of this evolution for a two-level , mapping pure states to points on a where the z-axis represents the population difference and the equatorial () encodes information in superpositions. Dephasing manifests as a time-dependent contraction of the equatorial , shrinking the Bloch vector's transverse components toward the z-axis while leaving the polar angle unchanged, thereby preserving energy populations but diminishing . This radial shrinkage highlights the selective degradation of phase-dependent quantum features, distinguishing dephasing from other noise processes. Dephasing typically arises from environmental interactions introducing these fluctuating fields.

Physical Mechanisms

Environmental Interactions

Dephasing arises from the between a quantum and its surrounding environment, often modeled through an Hamiltonian of the form H_{\text{int}} = S \otimes B, where S is a and B is a bath . In pure dephasing scenarios, S is typically diagonal in the 's energy basis, such as the Pauli \sigma_z for a two-level , ensuring that the perturbs the phase without inducing transitions between energy eigenstates. This coupling facilitates an exchange of information between the and the , where the effectively "measures" the 's through weak interactions, leading to of the relative phases among superposed states without net to or from the . Such weak system- interactions result in the acquiring correlations that encode the 's off-diagonal elements, thereby suppressing quantum over time. Representative examples of these interactions include in solid-state s, such as quantum dots, where lattice vibrations couple longitudinally to the electronic spin or states, causing fluctuations due to fluctuating from the phonons. In , photon interactions within can similarly induce dephasing, as quantum fluctuations in the photon number of the couple to the qubit's via dispersive interactions, randomizing the qubit's . To model these effects theoretically, the Born-Markov approximation is commonly employed, assuming weak system-bath coupling and a memoryless bath response, which simplifies the dynamics to a Markovian where correlations in the bath decay rapidly compared to the system's timescales. This approximation captures the essential dephasing behavior in many regimes by treating the interaction as a perturbative, uncorrelated source.

Types of Noise

Noise sources inducing dephasing in quantum systems are broadly classified into classical and quantum categories based on their origins and statistical properties. Classical , often modeled as Gaussian processes arising from thermal baths, typically leads to Markovian dynamics where the environment's memory effects are negligible. In contrast, quantum involves non-Markovian fluctuations from quantum baths, such as bosonic oscillators, resulting in asymmetric spectral densities and temperature-dependent decoherence rates that can exhibit power-law or behaviors. Specific types of noise are distinguished by their power densities S(\omega), which dictate their impact on . White , characterized by a flat S(\omega) \approx \ constant for relevant frequencies, originates from high-frequency environmental baths and induces exponential dephasing with a \Gamma_\phi proportional to the , as the fluctuations average rapidly without long-term correlations. 1/f (pink) , with S(\omega) \propto 1/|\omega|, stems from ensembles of charge traps or defects that produce flicker-like fluctuations in the qubit's local environment; in flux qubits, this manifests as 1/f flux with around (10^{-6} \Phi_0)^2 / \mathrm{Hz} at 1 Hz, leading to Gaussian decay. Telegraph arises from sparse two-level fluctuators that randomly switch states, generating bistable jumps in the system's frequency \omega(t) = \omega_0 + \xi(t) where \xi(t) = \pm \nu, and results in non-Gaussian, non-Markovian dephasing with potential revivals depending on the switching \lambda and \nu. Low-frequency components of the noise spectrum dominate long-time dephasing, as the accumulated phase variance \langle \phi^2(t) \rangle scales with \int_0^\infty d\omega \, S(\omega) \frac{4 \sin^2(\omega t / 2)}{\omega^2}, approximating t^2 \int d\omega \, S(\omega) / \omega^2 for \omega t \ll 1, which diverges for 1/f and severely limits times. This sensitivity arises because low-frequency fluctuations cause quasi-static shifts in the energy splitting, accumulating phase errors linearly over time. In spin systems, such as electron spins in quantum dots, fluctuations from nuclear spin baths produce Overhauser field that drives dephasing through hyperfine interactions. Similarly, in superconducting , voltage —manifesting as charge fluctuations on gate electrodes—couples longitudinally to the qubit frequency, contributing to 1/f-like dephasing with amplitudes around 1-10 \mu eV at 1 Hz.

Distinction from Other Processes

Dephasing vs Relaxation

In , energy relaxation, often denoted as the T1 process, refers to the dissipation of energy from an to the through interactions with the environment, leading to a of in the . This process typically occurs via mechanisms such as of photons in atomic or superconducting systems, or absorption and emission of phonons in solid-state environments like quantum dots or spin ensembles. As a result, the longitudinal component of the or the difference along the quantization axis (z-direction) relaxes exponentially back to with a characteristic T1. In contrast, dephasing, characterized by the T2 process, involves the loss of phase coherence between quantum states without net energy exchange, preserving the populations but randomizing the relative phases in the transverse (x-y) plane. This pure dephasing arises from fluctuating environmental fields that cause phase diffusion, and the total coherence time satisfies T2 ≤ 2 T1, with equality holding in the absence of additional dephasing mechanisms. Inhomogeneous broadening, often contributing to an effective T2* ≤ T2, stems from static variations in local fields or transition frequencies across an ensemble, such as magnetic field inhomogeneities or site-specific disorder, which accelerate the apparent dephasing beyond intrinsic homogeneous limits. The distinctions between these processes are formalized in the Bloch equations, which describe the time evolution of the magnetization vector \vec{M} = (M_x, M_y, M_z) in the presence of a and relaxation. The longitudinal relaxation term affects only the z-component, driving M_z toward as \frac{dM_z}{dt} = -\frac{M_z - M_0}{T_1}, while the transverse relaxation term damps the x-y components as \frac{dM_x}{dt} = -\frac{M_x}{T_2} and \frac{dM_y}{dt} = -\frac{M_y}{T_2}, without altering populations. These equations highlight how T1 governs energy equilibration, whereas T2 captures phase randomization, with the full dynamics given by: \frac{d\vec{M}}{dt} = \gamma \vec{M} \times \vec{B} - \frac{M_x \hat{i} + M_y \hat{j}}{T_2} - \frac{(M_z - M_0) \hat{k}}{T_1}, where \gamma is the gyromagnetic ratio and \vec{B} is the effective field. The interplay between relaxation and dephasing is captured in the total decoherence rate, where the observed T2 incorporates contributions from both: \frac{1}{T_2} = \frac{1}{2 T_1} + \frac{1}{T_\phi}, with T_\phi representing the pure dephasing time due to elastic scattering or noise without energy transfer. This relation arises because population relaxation indirectly contributes to dephasing by randomly advancing or retarding phases during state changes, limiting the maximum coherence to twice the relaxation time in ideal cases. In the density matrix formalism, T1 decay affects diagonal elements, while T2 governs the exponential decay of off-diagonals, as detailed elsewhere.

Relation to Decoherence

refers to the loss of quantum superpositions arising from the entanglement of a with its , which effectively monitors certain observables and suppresses between incompatible states. This underlies the quantum-to-classical by rendering quantum correlations unobservable on macroscopic scales. Dephasing plays a central role in decoherence as the mechanism that selects pointer states—robust, classical-like states resilient to environmental perturbations—through the destruction of in superpositions. By inducing random fluctuations without exchange, dephasing preferentially preserves pointer states that align with the environment's monitoring basis, thereby eliminating fragile quantum interferences. This selection process is formalized as einselection (environment-induced superselection), where dephasing stabilizes classical states by redundantly encoding system information in the , enforcing an effective ban on non-classical superpositions across the . In einselection, the acts as a witness, dynamically favoring pointer states that minimize entanglement and maximize predictability, thus bridging to classical objectivity. In many quantum systems, dephasing precedes full decoherence, initiating the loss of coherence before amplitude damping or energy relaxation takes hold; for instance, in cavity quantum electrodynamics (QED) experiments with Rydberg atoms and microwave fields, initial dephasing due to environmental scattering disrupts photon superpositions, paving the way for complete decoherence. These observations highlight dephasing's hierarchical position in the decoherence cascade. Regarding the quantum Zeno effect in dephasing contexts, frequent measurements can accelerate decoherence through the anti-Zeno regime, where intermediate measurement intervals enhance dephasing rates by amplifying environmental correlations, contrasting the suppression seen in the standard Zeno limit. This dual behavior underscores dephasing's sensitivity to observation frequency in driving systems toward classicality.

Mathematical Formalism

Density Matrix Approach

The density matrix formalism provides a statistical description of , particularly useful for modeling dephasing in open systems where pure-state wavefunctions are insufficient. The \rho is defined as \rho = \sum_i p_i |\psi_i\rangle\langle\psi_i|, where p_i are the probabilities of the pure states |\psi_i\rangle in an ensemble, ensuring \sum_i p_i = 1 and \operatorname{Tr}(\rho) = 1. This representation captures both pure states (where \rho = |\psi\rangle\langle\psi| and \operatorname{Tr}(\rho^2) = 1) and mixed states (where \operatorname{Tr}(\rho^2) < 1), allowing for the treatment of incoherent superpositions arising from environmental interactions. In the context of dephasing, the off-diagonal elements of \rho, which encode quantum coherences, decay exponentially over time while the diagonal elements (populations) remain unchanged. Specifically, for a two-level system, the off-diagonal coherence \rho_{01}(t) evolves as \rho_{01}(t) = \rho_{01}(0) e^{-\Gamma t}, where \Gamma is the dephasing rate, often expressed in terms of the dephasing time T_2 such that \Gamma = 1/T_2. This process transforms a coherent superposition—initially a pure state with significant off-diagonal terms—into a classical mixture, where \rho becomes diagonal in the energy basis, reflecting the loss of phase information without energy exchange. For a , the under pure dephasing takes the form \rho(t) = \begin{pmatrix} \rho_{00}(0) & \rho_{01}(0) e^{-t/T_2} \\ \rho_{10}(0) e^{-t/T_2} & \rho_{11}(0) \end{pmatrix}, where \rho_{00} + \rho_{11} = 1 and the Hermitian conjugate ensures \rho_{10} = \rho_{01}^*. This evolution preserves the trace \operatorname{Tr}(\rho(t)) = 1 and the populations \rho_{00}, \rho_{11}, but systematically erodes the coherences, leading to a fully mixed \rho = \operatorname{diag}(1/2, 1/2) in the long-time limit. Such behavior highlights dephasing's role in degrading while maintaining marginal probabilities.

Master Equation

The dynamics of dephasing in open quantum systems is governed by that describe the evolution of the reduced density operator \rho of the system, accounting for interactions with an . In the Markovian regime, where environmental correlations decay much faster than the system's timescales, the Lindblad provides the most general form ensuring complete positivity and preservation. This is given by \frac{d\rho}{dt} = -i [H, \rho] + \sum_k \left( L_k \rho L_k^\dagger - \frac{1}{2} \{ L_k^\dagger L_k, \rho \} \right), where H is the Hamiltonian and the L_k are Lindblad operators representing dissipative channels. For pure dephasing in a two-level with H = \frac{\omega}{2} \sigma_z, the process is modeled by a single Lindblad operator L_z = \sqrt{\Gamma/2} \, \sigma_z, where \Gamma is the dephasing rate and \sigma_z is the Pauli-z matrix. This form preserves the populations \rho_{00} and \rho_{11} while causing the off-diagonal \rho_{01} to decay exponentially as \rho_{01}(t) = \rho_{01}(0) e^{-i \omega t - \Gamma t}, reflecting the loss of phase information without exchange. For systems with weak system-bath coupling, the offers a perturbative approach to derive the from the full system-bath dynamics, typically via second-order expansion in the . The general acts on \rho as \dot{\rho} = -i [H, \rho] + \mathcal{R}[\rho], where \mathcal{R} incorporates relaxation and dephasing terms proportional to bath correlation functions evaluated at the system's Bohr frequencies. In the context of dephasing for a two-level system, the secular approximation simplifies this to a dephasing superoperator that affects only the off-diagonals, yielding a rate \Gamma = 2 \pi J(0) k_B T / \hbar^2 in the high-temperature limit for baths with finite low-frequency spectral density J(0). This approximation bridges microscopic interactions, such as those from phonons or fluctuators, to observable coherence decay, though it assumes weak coupling and neglects non-secular terms that can lead to inaccuracies at short times. When environmental correlations persist over timescales comparable to or longer than the system's , non-Markovian effects arise, invalidating the memoryless assumption and leading to more complex master equations, often integrodifferential in form. These effects can manifest as revivals of or incomplete dephasing, where temporarily flows back from the bath to the . For a two-level under pure dephasing, coupled diagonally to a bosonic bath via the interaction V = \sigma_z \sum_k \lambda_k (b_k + b_k^\dagger), the exact solution for the is \rho_{01}(t) = \rho_{01}(0) \, e^{-i \omega t - \int_0^t \gamma(\tau) \, d\tau}, where \gamma(\tau) is a time-dependent dephasing derived from the bath C(\tau) = \langle V(\tau) V(0) \rangle, capturing memory effects such as those in low-temperature or structured baths. This form highlights how persistent correlations can slow or oscillate the dephasing process compared to Markovian predictions.

Applications and Examples

In Quantum Computing

In quantum computing, dephasing primarily affects qubits by disrupting the relative phase between computational states, leading to phase-flip errors equivalent to Pauli operators in the . These errors manifest as bit-flip errors when qubits are measured in the phase basis (after applying Hadamard gates), thereby corrupting superpositions essential for quantum algorithms. In flux-tunable superconducting qubits, such as transmons or fluxoniums, dephasing is often the dominant error source due to low-frequency flux noise coupling to the qubit's tunable frequency, which induces random phase accumulation. Modern superconducting quantum processors achieve dephasing-limited times T_2 ranging from 100 μs to over 1 ms as of 2025, which sets an upper bound on gate fidelities approaching or exceeding 99.9% for two-qubit operations without correction. Recent advancements, such as using tantalum-based materials, have extended these times beyond previous limits, improving . These T_2 times (distinct from energy relaxation T_1) constrain the depth of quantum circuits, as prolonged idling or multi-gate sequences amplify phase errors. Dephasing is modeled in frameworks as a pure dephasing , where the process applies Pauli Z errors with a probability proportional to the dephasing rate, enabling tailored codes like codes in the Z basis. The first experimental observation of qubit dephasing occurred in the late 1990s during pioneering (NMR) quantum computing demonstrations, where ensemble spin coherence decay was measured as T_2 relaxation in liquid-state systems implementing basic quantum gates. These early NMR experiments highlighted dephasing as a key limitation, paving the way for subsequent solid-state designs that continue to grapple with similar challenges.

In Spectroscopy

In nuclear magnetic resonance (NMR) and electron spin resonance (ESR) spectroscopy, dephasing contributes to the transverse relaxation time T₂, which characterizes the loss of phase coherence among spins due to local magnetic field fluctuations and spin-spin interactions. This process broadens the lineshape of spectral peaks, as the dephasing of transverse leads to a of the signal. In NMR, T₂ relaxation encompasses both irreversible spin-spin interactions and reversible dephasing effects, distinguishing it from the longitudinal relaxation time T₁. Similarly, in ESR, dephasing via T₂ mechanisms arises from hyperfine interactions and environmental noise, resulting in linewidth broadening that limits spectral resolution. In optical , pure dephasing plays a key role in the dynamics of , where it disrupts the phase relationship between the and excited states without energy exchange, leading to of absorption lines. This broadening is particularly evident in nanostructures and molecular aggregates, where exciton-phonon and exciton-exciton interactions accelerate dephasing, reducing the lifetime. For instance, in dichalcogenides, pure dephasing contributes significantly to the exciton linewidth, with times on the order of femtoseconds to picoseconds, as measured by two-dimensional . This process must be distinguished from population relaxation, though both influence the overall spectral response. Femtosecond spectroscopy techniques, such as photon echo methods, probe dephasing dynamics in molecular systems on ultrafast timescales, revealing coherence times typically around picoseconds due to solvent interactions and intramolecular vibrations. In these experiments, three-pulse photon echoes rephase the initial dephasing caused by environmental fluctuations, allowing direct measurement of the pure dephasing rate γ*, which quantifies phase-destroying collisions without population transfer. For dye molecules in solution, such as resorufin in ethanol, dephasing times of ~100-500 fs have been observed, highlighting the role of non-Markovian bath correlations in condensed phases. These measurements provide insights into the transition from quantum coherence to classical behavior in molecular excitations. In two-dimensional (2D) electronic , dephasing modulates the decay of off-diagonal peaks, enabling the mapping of pathways in light-harvesting complexes by distinguishing coherent from incoherent transport. For example, in the LHCII complex of , rapid electronic dephasing (~60 fs) broadens cross-peaks, revealing downhill energy funneling from higher- to lower-energy sites with transfer times of 200-700 fs. This dephasing-driven analysis highlights how environmental interactions facilitate efficient migration while suppressing long-lived coherences, as confirmed by global fitting of 2D spectra. Such observations underscore dephasing's role in optimizing without delving into off-diagonal details.

Measurement and Control

Experimental Measurement

Experimental measurement of dephasing rates typically involves pulse sequences that probe the decay of quantum coherence in controlled such as qubits or nuclear spins. One fundamental technique is , which quantifies the inhomogeneous dephasing time T_2^* by initializing the system in a superposition state using a \pi/2 pulse, allowing free evolution for a variable time \tau, and then applying a second \pi/2 pulse to read out the phase accumulation. The resulting interference fringes decay exponentially due to dephasing, with the coherence signal fitting to e^{-\tau / T_2^*}, providing a direct measure of environmental noise sensitivity. This method is widely used in superconducting qubits and trapped ions, where it reveals the combined effects of static and low-frequency noise. To isolate the homogeneous dephasing time T_2, which excludes inhomogeneous broadening, the technique, particularly the Hahn echo sequence, is employed. This involves a \pi/2 preparation pulse, a free evolution period \tau, a \pi refocusing pulse to reverse phase accumulations from static offsets, another evolution period \tau, and a final \pi/2 readout pulse. The echo signal at time $2\tau decays as e^{-2\tau / T_2}, allowing extraction of T_2 through exponential fitting, as the refocusing compensates for quasi-static noise while retaining sensitivity to fluctuating dephasing mechanisms. This approach has been foundational in (NMR) and extended to solid-state qubits for precise characterization of pure dephasing. Advanced techniques using dynamical decoupling pulses further refine dephasing measurements by probing the spectrum. Sequences like the Carr-Purcell-Meiboom-Gill (CPMG) apply multiple \pi pulses at varying inter-pulse spacings during free evolution, effectively filtering at specific frequencies and extending to reveal the bath's spectral density. By analyzing the rate as a of pulse number or timing, researchers reconstruct the environment, distinguishing between 1/ and contributions in systems like superconducting circuits. These methods enhance measurement accuracy beyond basic Ramsey or protocols, particularly for low-frequency dephasing sources. Dephasing times extracted from these experiments vary by platform: in superconducting qubits, T_2^* from often ranges from 10 to 100 \mus, while Hahn yields T_2 up to several ms; in NMR systems, T_2 typically spans milliseconds to seconds, reflecting weaker environmental coupling. These values are obtained via least-squares fitting to models, establishing key benchmarks for quantum device performance.

Mitigation Strategies

Dynamical decoupling (DD) techniques employ sequences of precisely timed pulses to refocus the phase evolution of , effectively averaging out low-frequency noise sources that cause dephasing. These methods are particularly effective against quasi-static or slowly varying environmental fluctuations, such as those from nuclear spins or inhomogeneities. A prominent example is the Uhrig dynamical decoupling (UDD) protocol, which optimizes pulse timings using a to minimize decoherence up to higher orders compared to earlier schemes like Carr-Purcell-Meiboom-Gill (CPMG). In experiments with superconducting qubits and trapped ions, UDD has extended times by factors of 2–10 under dephasing-dominant noise, demonstrating robustness across various bath models including spin-boson environments. Material engineering approaches target the reduction of intrinsic noise baths at the hardware level to inherently prolong dephasing times. In semiconductor quantum dots and donor-based qubits, isotopic purification removes nuclei with nonzero , such as depleting 29Si () in favor of 28Si (-0), which suppresses hyperfine interactions that drive dephasing. This technique has achieved electron coherence times exceeding 30 seconds at millikelvin temperatures in isotopically engineered , a dramatic improvement over natural abundance samples where T2 remains below 1 second due to nuclear fluctuations. Similar purification in and (e.g., 12C enrichment) has yielded comparable gains, enabling longer storage of in solid-state platforms. Quantum error correction (QEC) codes provide a scalable strategy to combat dephasing errors by encoding logical s across multiple physical ones, allowing detection and correction without direct measurement of the encoded state. Surface codes, in particular, exhibit high thresholds for dephasing-biased noise, where phase-flip (Z) errors dominate; modified variants can tolerate error rates up to 1% per cycle while suppressing logical errors exponentially with code distance. Effective implementation requires the dephasing time to exceed the extraction cycle time, typically on the order of microseconds for current gate fidelities, ensuring that errors accumulate slowly relative to correction operations. In demonstrations with superconducting processors, surface code patches have maintained logical fidelity above 99% for dephasing rates corresponding to ≈ 100 μs. Advanced feedback control methods leverage continuous weak measurements to monitor and dynamically adjust qubit phases in real time, stabilizing coherence against stochastic dephasing. By detecting phase drifts through dispersive readout and applying corrective pulses or bias shifts, these schemes can suppress noise equivalent to extending T2 by factors of 3–5 in transmon qubits. Theoretical frameworks based on stochastic master equations guide the feedback gain to minimize backaction while maximizing information extraction, with experimental validations showing reduced dephasing rates under fluctuating flux noise. Such approaches are especially promising for hybrid systems integrating measurement and control at cryogenic temperatures.

References

  1. [1]
    [PDF] Lecture Notes for Ph219/CS219: Quantum Information Chapter 3
    This is a model of dephasing because the coupling of the spin to the bath is diagonal in the spin's energy eigenstate basis. (Otherwise the physics of the model ...
  2. [2]
  3. [3]
    Spin Dephasing - Purdue Physics
    The dephasing is due to weak scattering processes which do not conserve the total electron spin. Two scattering processes which are of particular importance in ...
  4. [4]
    [PDF] Superconducting Qubits: Dephasing and Quantum Chemistry
    This work discusses dephasing in superconducting qubits, quantifying phase noise, and quantum algorithms for simulating chemical Hamiltonians.
  5. [5]
    Effect of Pure Dephasing Quantum Noise in the Quantum Search ...
    By evaluating pure dephasing quantum noise, we highlight the importance of considering relaxation times when designing optimal architectures for ...
  6. [6]
    [PDF] Basics of NMR Spectroscopy - UConn Health
    Nov 29, 2016 · NMR Relaxation – Dephasing in the xy plane. • T. 2 relaxation causes the FID to decay and the faster the decay the broader the NMR signal ...
  7. [7]
    Dephasing and Relaxation | Wright Group
    Condensed phase dephasing rates for NMR are ~103 sec-1, vibrational states are ~1012 sec-1, and electronic states are ~1014 sec-1.
  8. [8]
    Quantifying the nonclassicality of pure dephasing - Nature
    Aug 22, 2019 · Dephasing processes, caused by non-dissipative information exchange between quantum systems and environments, provides a natural platform for ...
  9. [9]
    [PDF] Introduction to dissipation and decoherence in quantum systems
    Sep 25, 2008 · These will be destroyed gradually by the cou- pling to the environment, an effect known as “decoherence” or “dephasing”. ... Mathematical ...
  10. [10]
    Relaxation Effects in Nuclear Magnetic Resonance Absorption
    Relaxation Effects in Nuclear Magnetic Resonance Absorption. N. Bloembergen*, E. M. Purcell, and R. V. Pound† ... 73, 679 – Published 1 April, 1948. DOI ...
  11. [11]
    Canonical Hamiltonian ensemble representation of dephasing ...
    May 11, 2021 · The ensemble average results in the blue arrow, whose dynamical behaviour is pure dephasing due to the random phase of the ensemble of red ...
  12. [12]
    [PDF] Case Study of Decoherence Times of Transmon Qubit - arXiv
    Sep 10, 2023 · Quantum decoherence is usually characterized by measuring two times: T1 (relaxation time) and T2 (dephasing time). Although in the new qubit ...
  13. [13]
    [PDF] Notes on noise - John Preskill
    In this model, the Hamiltonian for the bath and for the coupling of the bath to the system is. HB + HSB = X k ωka. † kak −. 1. 2 σz. X k gkak + g∗ ka. † k ! (22).
  14. [14]
    [PDF] Dephasing of a qubit due to quantum and classical noise
    We consider quantum noise generated by a dissipative quan- tum bath. A detailed comparative study with the results for a classical noise source such as gen-.
  15. [15]
    Decoherence of flux qubits due to 1/f flux noise - cond-mat - arXiv
    Jun 19, 2006 · We present a Gaussian decay function of the echo signal as evidence of dephasing due to 1/f flux noise whose spectral density is evaluated to be about (10^{-6} ...
  16. [16]
    Quantum dephasing induced by non-Markovian random telegraph ...
    Jan 9, 2020 · We theoretically study the dynamical dephasing of a quantum two level system interacting with an environment which exhibits non-Markovian random telegraph ...
  17. [17]
    [PDF] Decoherence of a superconducting qubit due to bias noise
    Mar 25, 2003 · The main purpose of this paper is to give a physical picture of decoherence, showing that it can be understood as a random fluctuation in the ...
  18. [18]
    (PDF) Controlling the Spontaneous Emission of a Superconducting ...
    Aug 6, 2025 · With advancements in quantum information engineering, there has been increasing focus on the relaxation and dephasing mechanisms in quantum ...<|control11|><|separator|>
  19. [19]
    Bloch Equations and Relaxation II - UCLA
    Jan 25, 2023 · Transverse or spin-spin relaxation. – Molecular interaction causes spin dephasing. – Typically, T2 < (10s ms).
  20. [20]
    [PDF] CHM 502 - Module 11 - The Optical Bloch Equations
    This will give rise to the “Optical Bloch equations.”    (15) accounting for both population relaxation (T1 = 1/γ1) and dephasing (T2 = 1/γ2). where ω is ...
  21. [21]
    12.4: Ensemble Averaging and Line-Broadening
    Mar 6, 2022 · These intrinsically molecular processes, often referred to as homogeneous broadening, are commonly assigned a time scale T 2 = Γ − 1 .
  22. [22]
    [PDF] Semiconductor Bloch Equations - Karlsruher Institut für Technologie
    1. (a) First we study the dynamics of atoms near resonance in the two–level ap- proximation and derive the optical Bloch equations ...
  23. [23]
    Decoherence, einselection, and the quantum origins of the classical
    May 24, 2001 · Decoherence is caused by the interaction with the environment. Environment monitors certain observables of the system, destroying interference ...
  24. [24]
    Observing the Progressive Decoherence of the ``Meter'' in a ...
    A mesoscopic superposition of quantum states involving radiation fields with classically distinct phases was created and its progressive decoherence observed.
  25. [25]
    [1402.5228] Zeno and anti-Zeno effects on Dephasing - arXiv
    Feb 21, 2014 · Title:Zeno and anti-Zeno effects on Dephasing. Authors:Adam Zaman Chaudhry, Jiangbin Gong. View a PDF of the paper titled Zeno and anti-Zeno ...
  26. [26]
    [PDF] CHM 502 - Module 10 - Density Matrices & Dephasing
    The density matrix (or density operator) ρ is critical infrastructure to treat statistical mixtures of quantum systems. We'll start by reviewing this formalism.
  27. [27]
  28. [28]
    [1101.0141] Relaxation and dephasing in open quantum systems ...
    Dec 30, 2010 · We examine two different representitive limits of the Lindblad equation (relaxation and pure dephasing) and are able to deduce a number of ...
  29. [29]
    Efficient Quantum Error Correction of Dephasing Induced by a ...
    Jan 17, 2020 · The standard QEC approach to correct dephasing uses E i s comprising Pauli Z operators on at most w qubits (and I on the rest).
  30. [30]
    Overlap junctions for high coherence superconducting qubits
    Coherence times of superconducting qubits have been increased significantly in both 2D and 3D geometries (∼10–100 μs).1–5 These relatively long coherence times ...Missing: T2 | Show results with:T2
  31. [31]
    [PDF] Ultrahigh Error Threshold for Surface Codes with Biased Noise - arXiv
    We show that a simple modification of the surface code can exhibit an enormous gain in the error correction threshold for a noise model in which Pauli Z errors ...
  32. [32]
    NMR techniques for quantum control and computation
    Jan 12, 2005 · This article surveys and summarizes a broad variety of pulse control and tomographic techniques which have been developed for, and used in, NMR quantum ...<|separator|>
  33. [33]
    [PDF] Lecture #3 Basics of Relaxation
    T1 relaxation depends on transverse fields having energy at the. Larmor frequency. • T2 relaxation depends on both J(0) and J(ω0. ). • Helps explain relaxation ...<|control11|><|separator|>
  34. [34]
    [PDF] Introduction to solution NMR - EMBL Hamburg
    Dec 6, 2012 · – Spin-spin relaxation. – Transversal relaxation → T2 relaxation. • Dephasing of magnetization in the x/y plane. NMR relaxation. 47. B0. B0. B1.
  35. [35]
    Intrinsic homogeneous linewidth and broadening mechanisms ... - NIH
    Sep 18, 2015 · The homogeneous linewidth is linked to population relaxation through γ=Γ/2+γ*, where γ* characterizes pure dephasing processes10 such as elastic ...
  36. [36]
    Pure optical dephasing dynamics in semiconducting single-walled ...
    Jan 19, 2011 · We report a detailed study of ultrafast exciton dephasing processes in semiconducting single-walled carbon nanotubes employing a sample ...
  37. [37]
    [PDF] Femtosecond photon echo measurements of electronic coherence ...
    Oct 1, 2003 · Photon echo and reverse transient grating measurements of the loss of electronic coherence for molecular iodine are presented.
  38. [38]
    Non-Markovian dephasing of molecules in solution measured with ...
    The electronic dephasing of large molecules in solution is investigated with three-pulse photon echoes generated by 6-fs optical pulses.<|separator|>
  39. [39]
    [PDF] Two-dimensional Electronic Spectroscopy of Light Harvesting ...
    We found that the electronic dephasing occurs within ∼ 60 fs and inhomogeneous broadening is approximately 120 cm−1. A three-dimensional global fit analysis ...
  40. [40]
    [PDF] Two-Dimensional Spectroscopy Can Distinguish between ...
    Nov 2, 2011 · In an ensemble measurement, the observed de- phasing of a coherent state can occur because of either disorder across the ensemble or decoherence ...
  41. [41]
    A quantum engineer's guide to superconducting qubits
    Jun 17, 2019 · The aim of this review is to provide quantum engineers with an introductory guide to the central concepts and challenges in the rapidly accelerating field of ...
  42. [42]
    Realization of a Rydberg-Dressed Ramsey Interferometer and ...
    Feb 6, 2019 · We present the experimental realization and characterization of a Ramsey interferometer based on optically trapped ultracold potassium atoms.
  43. [43]
    Decoherence benchmarking of superconducting qubits - Nature
    Jun 26, 2019 · In this paper, we benchmark the stability of decoherence properties of superconducting qubits: T1, (free-induction decay), Tϕ (pure dephasing), ...
  44. [44]
    Ultrafast Optical Spin Echo for Electron Spins in Semiconductors
    The simplest method to measure T 2 is the Hahn spin-echo sequence [20] , which consists of a π / 2 pulse, a period of free evolution, a π pulse, and then ...Missing: T2 | Show results with:T2
  45. [45]
    [PDF] Spin echoes - MRI Questions
    that T, can be measured directly as well as T2, according to the discussion in III-D. A measured value of T₂=0.023 sec. is obtained, which is in substantial ...
  46. [46]
    A noise-resisted scheme of dynamical decoupling pulses ... - Nature
    Sep 15, 2020 · Here we propose a scheme to design a noise-resisted pulse, which features high fidelity exceeding under realistic situations.
  47. [47]
    [PDF] Material matters in superconducting qubits - arXiv
    Coupling to flux bias line generally leads to shorter T1 and T2 times in these qubits over fixed frequency transmons, which are less ... reductions in T2 and ...
  48. [48]
    Keeping a Quantum Bit Alive by Optimized 𝜋 -Pulse Sequences
    An optimized π-pulse sequence, extending the CPMG cycle, is used for dynamic decoupling to maintain quantum bit coherence and suppress decoherence.Abstract · Article Text
  49. [49]
    Tailoring Surface Codes for Highly Biased Noise | Phys. Rev. X
    Nov 12, 2019 · The surface code, with a simple modification, exhibits ultrahigh error-correction thresholds when the noise is biased toward dephasing.