David J. Wineland’s research while affiliated with University of Oregon and other places

What is this page?


This page lists works of an author who doesn't have a ResearchGate profile or hasn't added the works to their profile yet. It is automatically generated from public (personal) data to further our legitimate goal of comprehensive and accurate scientific recordkeeping. If you are this author and want this page removed, please let us know.

Publications (63)


Raman Scattering Errors in Stimulated-Raman-Induced Logic Gates in Ba + 133
  • Article

August 2023

·

10 Reads

·

7 Citations

Physical Review Letters

Matthew J. Boguslawski

·

Zachary J. Wall

·

Samuel R. Vizvary

·

[...]

·

Wesley C. Campbell

Ba+133 is illuminated by a laser that is far detuned from optical transitions, and the resulting spontaneous Raman scattering rate is measured. The observed scattering rate is lower than previous theoretical estimates. The majority of the discrepancy is explained by a more accurate treatment of the scattered photon density of states. This work establishes that, contrary to previous models, there is no fundamental atomic physics limit to laser-driven quantum gates from laser-induced spontaneous Raman scattering.


FIG. 1. (a) 133 Ba + level structure. The SRS photon frequency, ωsc, varies depending on the laser frequency, ω , and the decay channel (wavy lines). (b) As the gate laser is detuned further to the red of the intermediate resonance (ω i , i = 1 → 3), the scattering rate to a metastable state (Γi, proportional to ω 3 sc ) decreases until the error channel is closed (see ω 3 sc Θ(ωsc) in Eqn. 1).
FIG. 3. The SRS branching fraction to the D 5/2 manifold, ηD 5/2 as a function of laser frequency. (Black, solid) Moore et al. treatment [13]. (Red, dashed) CDA model. (Blue, dashed) ω 3 -theory prediction. The black point shows the result of 21 measurements with the standard error displayed as error bars. The inset shows a zoom-in near the gate-laser frequency.
FIG. 4. Calculated two-qubit gate error versus gate-laser optical frequency. (Black, solid) Moore et al. theory using a three-beam configuration, K = 1 and motional sideband frequency of 2π × 5 MHz. (Red, dashed) CDA theory for three beams, which reproduces limit εD∞ = (1/2)1.46 · 10 −4 for this configuration. (Blue, dashed) ω 3 -theory prediction. (Black point) Inferred achievable two-qubit gate error, calculated from the measured scattering rate and Stark shifts at 532 nm.
Errors in stimulated-Raman-induced logic gates in 133^{133}Ba$^+
  • Preprint
  • File available

December 2022

·

95 Reads

{}^{133}\mathrm{Ba}^+$ is illuminated by a laser that is far-detuned from optical transitions, and the resulting spontaneous Raman scattering rate is measured. The observed scattering rate is lower than previous theoretical estimates. The majority of the discrepancy is explained by a more accurate treatment of the scattered photon density of states. This work establishes that, contrary to previous models, there is no fundamental limit to laser-driven quantum gates from laser-induced spontaneous Raman scattering.

Download

Schematic of the X‐junction trap. a) Schematic view of the top wafer of the trap, with DC (control) electrodes in orange and RF electrodes in blue. A second wafer below the top wafer has DC and RF electrodes swapped, as indicated in the cross‐section. The ion shown is located on the axis x=y=0 of a linear portion of the trap. The ions are held in three major experiment zones, labeled by S, H, and V, connected by the junction located at C. Trapping zones L, A, B, and R lying in the same linear region as S are used together with the zone S to perform operations such as separation, recombination, and individual addressing and detection. (See text for more details.) b) Pseudo‐potential along the linear channels connected by the junction in the plane equidistant to the two wafers defined to be y=0. The junction gives rise to four pseudo‐potential bumps (positions indicated by blue arrows) around C.
Individual addressing and detection sequence. Using the sequence depicted within the dashed lines, ions in each well can be manipulated and detected individually, with the shuttling operations in the dashed box taking ≈1ms. Ions a and b are first separated into zones A and B, and then a is shuttled into S while b moves to zone R. This formation allows for internal state manipulation and detection of a. When a is shifted to L, ion b enters S and can be manipulated and/or its state detected.
Reordering two ⁹Be⁺ ions using the X‐junction. a) Schematic representation of reordering sequence. Two ions a and b in the double well potential are shuttled sequentially through the junction to separated regions of the trap array, and then moved back to the initial well with their order swapped. The arrows (orange for ion a and green for ion b) indicate the trajectories of each ion (light blue circles) and the blue circles represent the end points of the primitives on the trajectories. b) Ion a is excited using a single‐qubit rotation in the configuration SaRb, and a Rabi oscillation is observed in SaLb (orange) after the reordering sequence. No population oscillation is observed for detection performed in RaSb (green), while ion b is ideally in S. c) Coherence of the internal states of ion a (orange) and b (green) is maintained in two corresponding Ramsey sequences enclosing an exchange of ion positions and addressing one of the ions respectively. The phase shifts (0.46(2) rad for ion a (orange) as the offset of the fringe minimum from 0 and 2.29(2) rad for ion b (green)) arise mainly from the durations that the two ions accumulate phase due to a frequency shift relative to the local oscillator. See text for more details. d,e) The temperatures of the two ⁹Be⁺ ions are probed on the red (red dots) and blue (blue dots) sidebands after the reordering sequence. Fits to the Rabi‐oscillation model outlined in the Supporting Information (solid lines) result in average occupation numbers of 1.1(1) for ion b and 1.7(1) for ion a.
Chip‐based multi‐qubit quantum devices. a) 1D chain of individually addressed ions. Tightly focused beams allow individual addressing of ions in the chain. b) 1D trap array. Ions are confined in separated wells created by an electrode array. Separation between the wells that is large compared to the electrode dimensions allows for isolated control of each well. Coupling of a certain pair of ions is realized by repeated swap operations or crystal rotation (black arrows) and separation/recombination. c) Multi‐dimensional trap lattices. Ions are confined in fixed potential wells, while the coupling between the ions is enabled by tuning the potentials (black arrow). d) Multi‐dimensional trap array. Couplings between the ions are realized by shuttling information carriers (blue) or messengers (orange) through dedicated sections of the array.
Ion Transport and Reordering in a 2D Trap Array

May 2020

·

57 Reads

·

23 Citations

Scaling quantum information processors is a challenging task, requiring manipulation of a large number of qubits with high fidelity and a high degree of connectivity. For trapped ions, this can be realized in a 2D array of interconnected traps in which ions are separated, transported, and recombined to carry out quantum operations on small subsets of ions. Here, functionality of a junction connecting orthogonal linear segments in a 2D trap array to reorder a two‐ion crystal is demonstrated. The secular motion of the ions experiences low energy gain and the internal qubit levels maintain coherence during the reordering process, therefore demonstrating a promising method for providing all‐to‐all connectivity in a large‐scale, 2D or 3D trapped‐ion quantum information processor.


Generating number states and number-state superpositions
a, Relevant energy levels and transitions for the creation of motional states. BSB pulses transfer population between ↓k\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\left|\downarrow \right\rangle \left|k\right\rangle $$\end{document} and ↑k+1\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\left|\uparrow \right\rangle \left|k+1\right\rangle $$\end{document}, whereas RSB pulses transfer population between ↑k\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\left|\uparrow \right\rangle \left|k\right\rangle $$\end{document} and ↓k+1\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\left|\downarrow \right\rangle \left|k+1\right\rangle $$\end{document}. The BSB does not couple to ↑0\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\left|\uparrow \right\rangle \left|0\right\rangle $$\end{document} (crossed-out, faded blue arrow) because there is no energy level below the ground state. Transitions between the states ↑\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\left|\uparrow \right\rangle $$\end{document} and aux\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\left|{\rm{aux}}\right\rangle $$\end{document} are driven by a microwave field (indicated in green), which does not change k. b, Pulse sequence for generating a harmonic oscillator number state. Alternating RSB and BSB π pulses are applied, with each pulse adding one quantum of motion (or more quanta on higher-order sidebands; see the main text) and flipping the spin of the state. To analyse the resulting state, an RSB pulse is applied for a variable duration (labelled ‘RSB flop’) and the final spin state is detected via state-selective fluorescence. c, Pulse sequences and trap frequencies for number-state interferometers. The first effective π/2 pulse creates 0+n\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\left|0\right\rangle +\left|n\right\rangle $$\end{document}, followed by a free-precession period during which the mode frequency is increased by Δω. An effective π pulse swaps the phase of the superposition, 0+eiφn→eiφ0+n\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\left|0\right\rangle +{{\rm{e}}}^{i\varphi }\left|n\right\rangle \to {{\rm{e}}}^{i\varphi }\left|0\right\rangle +\left|n\right\rangle $$\end{document}. After another free-precession period with the mode frequency reduced by Δω, a final effective π/2 pulse closes the interferometer. For the composition of effective pulses see the main text.
Sideband flopping on number states
a, RSB flopping on the first-order sideband of an ↑n=40\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\left|\uparrow \right\rangle \left|n=40\right\rangle $$\end{document} state prepared using first-order RSB and BSB pulses. The curve shows the probability of measuring ↓\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\left|\downarrow \right\rangle $$\end{document} as a function of pulse duration during first-order RSB flopping to ↓n=41\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\left|\downarrow \right\rangle \left|n=41\right\rangle $$\end{document}. Each point represents an average over 200 experiments. The error bars represent one standard deviation of the mean. Solid black lines show theory fits to the data, with the Rabi frequency, the initial contrast and an exponential decay constant as fit parameters. b, RSB flopping on the fourth-order sideband of ↑n=100\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\left|\uparrow \right\rangle \left|n=100\right\rangle $$\end{document}. The curve shows the probability of measuring ↓\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\left|\downarrow \right\rangle $$\end{document} as a function of pulse duration during fourth-order RSB flopping to ↓n=104\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\left|\downarrow \right\rangle \left|n=104\right\rangle $$\end{document}. Each point represents an average over 500 experiments. c, Same as b, but on the third-order sideband of ↑n=100↔↓n=103\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\left|\uparrow \right\rangle \left|n=100\right\rangle \leftrightarrow \left|\downarrow \right\rangle \left|n=103\right\rangle $$\end{document}. d, First- to fourth-order sideband Rabi frequencies. All first-order data points (blue triangles) are fitted to determine the Lamb–Dicke parameter η = 0.2632(2). Curves for higher-order sidebands are plotted for the same η. Measured Rabi frequencies for higher-order sidebands (symbols) are consistent with theory (solid lines) for all orders. The duration of the π pulse from n=0\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\left|n=0\right\rangle $$\end{document} to n=1\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\left|n=1\right\rangle $$\end{document} is approximately 13 μs; the pulse durations required to produce higher-number states can be calculated on the basis of this result and the plotted Rabi frequencies for higher-number states (see Methods).
Interference and sensitivity of different number-state superpositions
a, Interference fringes for number-state interferometers with n = 2, 4, 8, 12. Each data point is averaged over 250 experiments and uses a waiting time of 100 μs before and after the effective π pulse. The error bars represent one standard deviation of the mean. The fringe spacing is reduced as 1/n, as expected for Heisenberg scaling. At the same time, the fringe contrast is reduced with increasing n owing to the larger number of imperfect pulses and the higher susceptibility to mode-frequency changes, which are not stable over all of the 250 experiments for each data point. This reduces the fringe slopes for n > 12 below the maximal slope, which is reached for n = 12. Solid lines show theory fits using equation (2), with the fringe spacing, contrast and vertical offset as fit parameters. We attribute deviations from the expected sinusoidal behaviour (for example, near the centre of the 0+12\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\left|0\right\rangle +\left|12\right\rangle $$\end{document} fringe) to changes in the Raman sideband Rabi frequencies by a few per cent, temporarily reducing the contrast for some points. b, Experimentally determined noise-to-signal ratio, δϕ, as defined by equation (3), as a function of order n (dots). Also shown are theoretical lines for a perfect classical interferometer at 1/n\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$1/\sqrt{n}$$\end{document} and the 1/n Heisenberg limit; this limit is valid for ideal number-state interferometers.
Tracking of oscillator frequency using number-state interferometers
a, Allan standard deviation of tracked fractional trap frequencies versus averaging time for interleaved experiments with n = 2, n = 4, n = 6 and n = 8 interferometers. The repetition rate of a single run, comprising long (100 μs Ramsey time) and short (20 μs Ramsey time) auto-balance sequences on both sides of the fringe was approximately 7 s⁻¹. The n = 8 interferometer produces the lowest Allan deviation of the fractional frequency. Trap-frequency drifts begin to dominate the Allan deviation at tens of seconds. b, Uncertainty of the fractional mode frequency versus averaging time for two series of only n = 8 interferometer runs performed to maximize the measurement duty cycle. We achieve a minimal fractional-frequency Allan deviation of 2.6(2) × 10⁻⁶ at approximately 4 s of averaging time with tracking activated (red triangles) and an experiment rate of approximately 43 s⁻¹ as defined above. The Allan standard deviation for averaging times of up to 1 s, without tracking activated, is shown by the blue circles. The minimum is reached after 0.5 s, with the experiment rate increased to approximately 250 s⁻¹. The error bars represent one standard deviation of the mean.
Quantum-enhanced sensing of a single-ion mechanical oscillator

August 2019

·

146 Reads

·

142 Citations

Nature

Special quantum states are used in metrology to achieve sensitivities below the limits established by classically behaving states1,2. In bosonic interferometers, squeezed states3, number states4,5 and 'Schrödinger cat' states5 have been implemented on various platforms and have demonstrated improved measurement precision over interferometers using coherent states6,7. Another metrologically useful state is an equal superposition of two eigenstates with maximally different energies; this state ideally reaches the full interferometric sensitivity allowed by quantum mechanics8,9. Here we demonstrate the enhanced sensitivity of these quantum states in the case of a harmonic oscillator. We extend an existing experimental technique10 to create number states of order up to n = 100 and to generate superpositions of a harmonic oscillator ground state and a number state of the form [Formula: see text] with n up to 18 in the motion of a single trapped ion. Although experimental imperfections prevent us from reaching the ideal Heisenberg limit, we observe enhanced sensitivity to changes in the frequency of the mechanical oscillator. This sensitivity initially increases linearly with n and reaches a maximum at n = 12, where we observe a metrological enhancement of 6.4(4) decibels (the uncertainty is one standard deviation of the mean) compared to an ideal measurement on a coherent state with the same average occupation number. Such measurements should provide improved characterization of motional decoherence, which is an important source of error in quantum information processing with trapped ions11,12. It should also be possible to use the quantum advantage from number-state superpositions to achieve precision measurements in other harmonic oscillator systems.



Quantum gate teleportation between separated qubits in a trapped-ion processor

May 2019

·

126 Reads

·

149 Citations

Science

Teleporting a trapped-ion quantum gate Gating—controlling the state of one qubit conditioned on the state of another—is a key procedure in all quantum information processors. As the scale of quantum processors increases, the qubits will need to interact over larger and larger distances, which presents an experimental challenge in solid-state architectures. Wan et al. implemented the 20-year-old theoretical proposal of quantum gate teleportation that allows separated qubits to interact effectively. They deterministically teleported a controlled-NOT gate between two computational qubits in spatially separated zones in a segmented ion trap, demonstrating a feasible route toward scalable quantum information processors. Science , this issue p. 875


Coherently displaced oscillator quantum states of a single trapped atom

March 2019

·

66 Reads

·

12 Citations

Coherently displaced number states of a harmonically bound ion can be coupled to two internal states of the ion by a laser-induced motional sideband interaction. The internal states can subsequently be read out in a projective measurement via state-dependent fluorescence, with near-unit fidelity. This leads to a rich set of line shapes when recording the internal-state excitation probability after a sideband excitation, as a function of the frequency detuning of the displacement drive with respect to the ion’s motional frequency. We precisely characterize the coherent displacement based on the resulting line shapes, which exhibit sharp features that are useful for oscillator frequency determination from the single quantum regime up to very large coherent states with average occupation numbers of several hundred. We also introduce a technique based on multiple coherent displacements and free precession for characterizing noise on the trapping potential in the frequency range of 500 Hz–400 kHz. Signals from the ion are directly used to find and eliminate sources of technical noise in this typically unaccessed part of the spectrum.


Quantum gate teleportation between separated zones of a trapped-ion processor

February 2019

·

55 Reads

Large-scale quantum computers will inevitably require quantum gate operations between widely separated qubits, even within a single quantum information processing device. Nearly two decades ago, Gottesman and Chuang proposed a method for implementing such operations, known as quantum gate teleportation (1). It requires only local operations on the remote qubits, classical communication, and shared entanglement that is prepared before the logical operation. Here we demonstrate this approach in a scalable architecture by deterministically teleporting a controlled-NOT (CNOT) gate between two computational qubits in spatially separated zones in a segmented ion trap. Our teleported CNOT's entanglement fidelity is in the interval [0.845, 0.872] at the 95% confidence level. The implementation combines ion shuttling with individually-addressed single-qubit rotations and detections, same- and mixed-species two-qubit gates, and real-time conditional operations, thereby demonstrating essential tools for scaling trapped-ion quantum computers combined in a single device.


Figure 2. Spin-flip probability P π ↓ (See Sec. 4.2) of ion after 13 µs tickle excitation on |↓↓ versus detuning from ion oscillation frequency. The average occupation ¯ n of the ion motion in response to tickle excitation is mapped onto the spin state by performing a RSB pulse, which connects levels |↓↓ |n to |↑↑ |n − 1 for n > 0, while leaving population in |↓↓ |n = 0 unchanged. A subsequent microwave carrier π-pulse exchanges populations in |↑↑ and |↓↓ to reduce measurement projection noise. The solid line is a fit using Eq. (7) and free parameter ¯ n and an experimentally determined vertical offset of 0.05(1) to account for background counts and imperfect ground state cooling. The fit yields an on resonance average occupation of ¯ n = 0.61(1).
Coherently displaced oscillator quantum states of a single trapped atom

November 2018

·

101 Reads

Coherently displaced harmonic oscillator number states of a harmonically bound ion can be coupled to two internal states of the ion by a laser-induced motional sideband interaction. The internal states can subsequently be read out in a projective measurement via state-dependent fluorescence, with near-unit fidelity. This leads to a rich set of line shapes when recording the internal-state excitation probability after a sideband excitation, as a function of the frequency detuning of the displacement drive with respect to the ion's motional frequency. We precisely characterize the coherent displacement based on the resulting line shapes, which exhibit sharp features that are useful for oscillator frequency determination from the single quantum regime up to very large coherent states with average occupation numbers of several hundred. We also introduce a technique based on multiple coherent displacements and free precession for characterizing noise on the trapping potential in the frequency range of 500 Hz to 400 kHz. Signals from the ion are directly used to find and eliminate sources of technical noise in this typically unaccessed part of the spectrum.


Figure 3: Interference and sensitivity of different number-state superpositions. a, Interference fringes for number state interferometers with n= 2, 4, 8 and 12. Each data point is averaged over 250 experiments. The fringe spacing is reduced as 1/n as expected for Heisenberg scaling. At the same time, the fringe contrast is reduced with higher n due to the larger number of imperfect pulses and the higher susceptibility to mode-frequency changes that are not stable over all 250 experiments for each data point. This reduces the fringe slopes for n > 12 below the maximal slope reached for n = 12. b, Experimentally determined noise-to-signal ratio δφ as defined in Eq.(3) as a function of order n (colored dots) together with the theoretical lines for a perfect classical interferometer at 1/ √ n and the 1/n Heisenberg limit valid for ideal number-state interferometers.
Figure 4: Oscillator frequency tracking using number-state interferometers. a, Interleaved comparison of the Allan standard deviation of tracked fractional trap frequencies vs. averaging time found with n = 2, n = 4, n = 6 and n = 8 interferometers. The repetition rate of a single run, comprised of long and short autobalance sequences on both sides of the fringe respectively, was approximately 7/s. The n = 8 interferometer produces the lowest fractional frequency Allan deviation. Trap frequency drifts begin to dominate the Allan deviation at 10's of seconds. b, Fractional mode-frequency uncertainty vs. averaging time for two series of only n = 8 interferometer runs to maximize measurement duty cycle. We are able to achieve a minimal fractional frequency Allan deviation of 2.6(2) ×10 −6 at approximately 4 seconds of averaging time with tracking activated (red triangles) and an experiment rate of approximately 43/s as defined above. The Allan standard deviation for averaging times up to 1 s without tracking activated is shown by the blue circles. The minimum is reached after 0.5 s with the experiment rate increased to approximately 250/s.
Quantum-enhanced sensing of a mechanical oscillator

July 2018

·

143 Reads

The use of special quantum states in boson interferometry to achieve sensitivities below the limits established by classically-behaving coherent states has enjoyed immense success since its inception. Squeezed states, number states and cat states have been implemented on various platforms and have demonstrated improved measurement precision over coherent-state-based interferometers. Another state is an equal superposition of two eigenstates with maximally different energies; this state ideally reaches the full interferometric sensitivity allowed by quantum mechanics. We extend a technique to create number states up to n = 100 and to generate superpositions of a harmonic oscillator ground state and a number state of the form 1/2(0+n)1/\sqrt{2}(|0\rangle+|n\rangle) up to n = 18 in the motion of a single trapped ion. While experimental imperfections prevent us from reaching the ideal Heisenberg limit, we observe enhanced sensitivity to changes in the oscillator frequency that initially increases linearly with n, with maximal value at n = 12 where we observe 2.1(1) times higher sensitivity compared to an ideal measurement on a coherent state with the same average occupation number. The quantum advantage from using number-state superpositions can be leveraged towards precision measurements on any harmonic oscillator system; here it enables us to track the average fractional frequency of oscillation of a single trapped ion to approximately 2.6×1062.6 \times 10^{-6} in 5 s. Such measurements should enable improved characterization of imperfections and noise on trapping potentials, which can lead to motional decoherence, an important source of error in quantum information processing with trapped ions.


Citations (36)


... The detection of biomolecules holds paramount significance in fundamental molecular research, diagnostics [44,45], drug screening, and various biomedical applications. Raman spectroscopy, an analytical technique, relies on the scattering of photons by molecules within a sample to measure their vibrational and rotational modes [46,47]. This method is not only facile to execute but also rapid and non-destructive, circumventing interference from aqueous solutions. ...

Reference:

Noble Metal Nanoparticle-Based Photothermal Therapy: Development and Application in Effective Cancer Therapy
Raman Scattering Errors in Stimulated-Raman-Induced Logic Gates in Ba + 133
  • Citing Article
  • August 2023

Physical Review Letters

... Similarly, arrays of trapped Rydberg atoms have been realized, where atoms can be moved using optical tweezers and interact via induced dipole-dipole interactions [10][11][12][13][14][15]. But also other types of segmented traps are conceivable, where, e.g., 1D or 2D arrays of ions are manipulated by laser pulses [16,17], and where interaction between ions in different modules takes place via some distance-dependent coupling [18][19][20][21][22]. What most approaches have in common is the necessity to cool particles to their mechanical ground state, to enable their manipulation and gates between them with high fidelity [16,[23][24][25][26][27]. ...

Ion Transport and Reordering in a 2D Trap Array

... Quantum-enhanced metrology can be achieved by improving the sensing procedures, such as enhancing the sensitivity using the nonclassicality of the probe [50][51][52], optimizing the interaction time [53], or using effective quantum measurement [54][55][56]. For example, a superposition between a ground state |0 and a number state |n of atomic motion can be used to optimally estimate the changes in frequencies of mechanical oscillators [57]. A motional Fock state, on the other hand, has been reported to be an optimal state for sensing a phase-randomized displacement in trapped ions [52]. ...

Quantum-enhanced sensing of a single-ion mechanical oscillator

Nature

... A quantum state is jointly measured with one-half of the entangled resource in a given place and, as a result of such an operation, it is transferred to the other half in a remote location. Such a protocol, together with entanglement swapping 4,5 , is at the core of several quantum computation 6,7 and quantum communication schemes ranging from quantum repeaters 8,9 , quantum gate teleportation [10][11][12] , measurement-based quantum computing [13][14][15] as well as port-based quantum teleportation 16,17 . Over the years, several experiments have successfully shown the teleportation of unknown quantum states using different experimental setups and degrees of freedom 2,3,[18][19][20][21][22][23][24][25] . ...

Quantum gate teleportation between separated qubits in a trapped-ion processor
  • Citing Article
  • May 2019

Science

... These Rabi interactions also appear to be favorable to quantum interferometry of photon scattering events, termed cat-state spectroscopy as demonstrated on single qubits on a trapped ion crystal [68,69]. This idea has been extended to a driven interferometry of a spin-dependance force [70], and to a motion echo [71]. Recently, a quantum-sensing protocol was proposed that leverages the phase transition of the Rabi model focused on frequency estimation [72]. ...

Coherently displaced oscillator quantum states of a single trapped atom

... 12,25,26 For one, there still remain questions regard-ing the quantum-to-classical crossover point that will be revealed by such systems, but in addition, the precision readout of such devices may offer insights into quantum gravity and high frequency gravitational wave detection. 13,[27][28][29][30][31] Studies on these photon-phonon interactions employ various types of resonant photonic device architectures, such as superconducting circuits, 32 co-planar resonators, 33 3-D microwave cavity resonators, 34 and whispering gallery resonators, 21,35,36 combined with various types of mechanical architectures displaying high coupling rates, such as membrane-in-the-middle systems, 6 trapped ion particles, 37 and acoustic-mechanical modes such as bulk acoustic wave (BAW) resonators. 6,25,26,35 BAW resonators are low-loss mechanical resonators that exhibit long mechanical coherence times due to their specifically engineered convex surface, which helps trap the majority of phonons in the center of the resonator. ...

Hybrid quantum systems with trapped charged particles
  • Citing Article
  • August 2016

Physical Review A

... Conversely, the score N achievable in the AT Game, which we considered here, is defined without any reference to parametric time. Note that this operational viewpoint is also imposed on us when we talk about the accuracy of the best available clocks, e.g., atomic clocks [31][32][33]. Since there do not exist more accurate clocks, we have no reference to test against -but we can of course still count how many ticks two such clocks emit until they run out of synchronisation. ...

The evolution of the Frequency Standards and Metrology symposium and its physics

Journal of Physics Conference Series

... Various master equations, including the axiomatic Lindblad-Gorini-Kossakowski-Sudarshan formulation [13], the phenomenological post-Markovian counterpart [14], and the Stochastic Schrödinger equation [15,16], have undergone extensive investigation. These methodologies, have proven useful in quantum optics [17], quantum computation [18,19], condensed matter physics [20], and the theory of decoherence [21], and have broad applicability. ...

Dissipative Quantum Control of a Spin Chain

Physical Review Letters

... A resourceful way to protect quantum information against decoherence processes that act locally is to encode it nonlocally in the form of superpositions of coherent states [1]. These Schrödinger cat states [2,3,4] are the logical states of the so-called Kerr-cat qubit, which can be generated with driven Kerr parametric oscillators [5,6,7,8,9,10,11,12], as those experimentally realized in superconducting circuits [13]. The present work warns against 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 A c c e p t e d M a n u s c r i p t Driving superconducting qubits into chaos 2 an additional problem: the potential development of chaos if the parameters of the oscillators are pushed beyond a threshold. ...

Nobel Lecture: Superposition, entanglement, and raising Schrödinger’s cat
  • Citing Article
  • July 2013

Review of Modern Physics

... Coupled quantized mechanical oscillators and the exchange (hopping) of local phonons between them have been examined using two ions in double-well potentials [19,20]. Multiple potential wells, each containing a single ion, have been controlled to demonstrate the tunable coupling of phonons [21], their coherent coupling [22], and two-dimensional networks of vibrational modes [23,24]. Regarding the latter type of phonon, Porras and Cirac proposed using local phonons in the radial direction of a linear ion string for simulating the Bose-Hubbard model [25]. ...

Tunable spin-spin interactions and entanglement of ions in separate wells

Nature