Demonstration of Erasure Conversion in a Molecular Tweezer Array

Connor M. Holland Department of Physics, Princeton University, Princeton, New Jersey 08544 USA    Yukai Lu Department of Physics, Princeton University, Princeton, New Jersey 08544 USA Department of Electrical and Computer Engineering, Princeton University, Princeton, New Jersey 08544 USA    Samuel J. Li Department of Physics, Princeton University, Princeton, New Jersey 08544 USA    Callum L. Welsh Department of Physics, Princeton University, Princeton, New Jersey 08544 USA    Lawrence W. Cheuk [email protected] Department of Physics, Princeton University, Princeton, New Jersey 08544 USA
(June 4, 2024)
Abstract

Programmable optical tweezer arrays of molecules are an emerging platform for quantum simulation and quantum information science. For these applications, reducing and mitigating errors that arise during initial state preparation and subsequent evolution remain major challenges. In this paper, we present work on site-resolved detection of internal state errors and quantum erasures, which are qubit errors with known locations. First, using a new site-resolved detection scheme, we demonstrate robust and enhanced tweezer array preparation fidelities. This enables creating molecular arrays with low defect rates, opening the door to high-fidelity simulation of quantum many-body systems. Second, for the first time in molecules, we demonstrate mid-circuit detection of erasures using a composite detection scheme that minimally affects error-free qubits. We also demonstrate mid-circuit conversion of blackbody-induced errors into detectable erasures. Our demonstration of erasure conversion, which has been shown to significantly reduce overheads for fault-tolerant quantum error correction, could be useful for quantum information processing in molecular tweezer arrays.

I Introduction

Programmable optical tweezer arrays of ultracold molecules are an emerging platform for quantum science Kaufman and Ni (2021). They combine the microscopic detection and control capabilities offered by programmable optical tweezer arrays Endres et al. (2016); Labuhn et al. (2016) with useful properties of molecules such as rich internal structures and intermolecular electric dipolar interactions between long-lived states Yan et al. (2013); Christakis et al. (2023); Gregory et al. (2024). In particular, the internal structure of molecules provides additional ways to encode quantum information and can be exploited for precision measurement experiments Carr et al. (2009); Bohn et al. (2017); Safronova et al. (2018); Kozyryev and Hutzler (2017); Anderegg et al. (2023), while the intermolecular electric dipolar interaction is the key ingredient for high-fidelity quantum gates DeMille (2002); Ni et al. (2018); Yu et al. (2019) and simulation of novel quantum many-body Hamiltonians Micheli et al. (2006); Carr et al. (2009); Gadway and Yan (2016); Bohn et al. (2017); Kaufman and Ni (2021).

Starting with the first demonstrations of trap** and high-fidelity detection of individual molecules in optical tweezer traps Anderegg et al. (2019); Zhang et al. (2022); Holland et al. (2023a); Ruttley et al. (2023); Vilas et al. (2024), recent work has demonstrated highly coherent rotational qubits in tweezers Burchesky et al. (2021); Park et al. (2023); Gregory et al. (2024), observed coherent dipolar interactions between individually held molecules, and demonstrated on-demand entanglement and 2-qubit gates between rotational molecular qubits Holland et al. (2023b); Bao et al. (2023a). These works establish the building blocks for quantum information processing and quantum simulation with molecular tweezer arrays.

A major experimental challenge in any quantum science platform is reducing and mitigating errors that occur during the initial preparation of a quantum system and its subsequent evolution. In molecular tweezer arrays, one desires to deterministically prepare tweezer sites such that they are occupied by molecules from a single internal state. This preparation of molecular tweezer sites is challenging, and has been a bottleneck for applications in quantum simulation and information processing. In particular, the complex internal structure of molecules combined with substantial light shifts inside tweezer traps render high-fidelity optical pum** difficult Lu et al. (2024); Bao et al. (2023b), thereby limiting achievable internal state purities. Concerning errors that can arise during quantum evolution, Raman scattering of tweezer trap** light or blackbody radiation can excite molecules out of the desired quantum states Holland et al. (2023a). While directly reducing these errors through better control is an important research area that has attracted much interest, a complementary approach is to mitigate their effects via measurement and feedback. In particular, for programmable tweezer arrays, because local control is available, site-resolved error detection could in some cases enable their subsequent removal.

Site-resolved detection of preparation errors has been used in atomic tweezer arrays to increase internal state preparation fidelities Scholl et al. (2023a), increase robustness to radiative decays of high-lying Rydberg states Scholl et al. (2023a); Ma et al. (2023), and to lower motional temperatures Scholl et al. (2023b). In recent work with molecules that are assembled from two atoms and trapped in tweezer arrays, this has been used to enhance molecular formation efficiency and to reduce defects in molecular arrays Picard et al. (2024); Ruttley et al. (2024).

Restricting to the specific case where information is stored in an array, an information error whose location is known is called an erasure. For qubits, such errors are called quantum erasures Grassl et al. (1997); Bennett et al. (1997). Recently, it has been pointed out that leakage errors, where a particle hosting a qubit exits its computational space, can be converted and detected as an erasure Wu et al. (2022). Notably, erasure conversion substantially raises fault-tolerant quantum error correction thresholds to practically achievable levels Wu et al. (2022). This important insight has led to several proposed erasure conversion schemes across a variety of qubit platforms Wu et al. (2022); Kubica et al. (2023); Kang et al. (2023); Teoh et al. (2023). Recently, the first demonstrations of quantum erasure conversion were achieved in atomic tweezer arrays of alkaline-earth atoms  Ma et al. (2023); Scholl et al. (2023a, b) and superconducting circuits Chou et al. (2023); Levine et al. (2024).

In this paper, we present new site-resolved detection schemes to mitigate both state preparation errors and qubit leakage errors for laser-cooled molecular arrays. First, we demonstrate a new site-resolved detection scheme that enables robust and enhanced fidelities of preparing a tweezer site with a molecule in a single internal state. We achieve a record-level tweezer site preparation fidelity of 95.2(3)%95.2percent395.2(3)\%95.2 ( 3 ) %, with internal state purity of 99.5(1)%99.5percent199.5(1)\%99.5 ( 1 ) %, substantially surpassing the previous best reported fidelities of 80%absentpercent80\approx 80\%≈ 80 % Holland et al. (2023b); Lu et al. (2024). Second, using a new composite erasure detection scheme, we demonstrate quantum erasure conversion in molecules for the first time. By utilizing a new hyperfine qubit encoding that is both highly coherent and compatible with a previously demonstrated 2-qubit gate Holland et al. (2023b); Bao et al. (2023a), we demonstrate a mid-circuit detection scheme that minimally affects qubit population and coherence. We further show mid-circuit erasure conversion of leakage errors caused by blackbody radiation, achieving improved qubit lifetimes and coherence times.

Our new preparation scheme overcomes a major challenge in molecular tweezer arrays and opens the door to simulation of quantum many-body spin models with low defect rates. Our demonstration of quantum erasure conversion adds an important capability to the molecular quantum information processing toolkit, and opens the door to mid-circuit quantum error correction in molecular tweezer arrays.

II Framework for Site-Resolved Error Detection

In atomic and molecular tweezer array experiments, each tweezer site is either empty or occupied by one particle. The local Hilbert space \mathcal{H}caligraphic_H at each site thus consists of |ket\left|\emptyset\right\rangle| ∅ ⟩, the state corresponding to an empty site, and the set of all internal states (Fig. 1(a)). (We will ignore motional states of the particle within a tweezer here, although the same framework used here applies Scholl et al. (2023b).) For a particular application, only a target subset of these states 𝒯𝒯\mathcal{T}caligraphic_T is used. For example, in order to encode a qubit, at least two internal states are needed.

For site-resolved error detection, one uses a set of detectable internal states \mathcal{F}caligraphic_F to flag the presence of an error. The detection outcome is binarized and converted into an error flag {ei}subscript𝑒𝑖\{e_{i}\}{ italic_e start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT } where ei=1subscript𝑒𝑖1e_{i}=1italic_e start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 1 (ei=0subscript𝑒𝑖0e_{i}=0italic_e start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 0) indicates an error (error-free site) (Fig. 1(b)). {ei}subscript𝑒𝑖\{e_{i}\}{ italic_e start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT } could be used to remove errors mid-sequence, or to post-select for error-free instances by excising error-flagged data.

When at least one bit of information is encoded in 𝒯𝒯\mathcal{T}caligraphic_T, errors with a known location are called erasures Grassl et al. (1997); Bennett et al. (1997). In particular, population leakage out of 𝒯𝒯\mathcal{T}caligraphic_T into a set of disjoint internal states \mathcal{E}caligraphic_E can be transferred to \mathcal{F}caligraphic_F and subsequently detected, a process known as erasure conversion Wu et al. (2022).

For site-resolved error detection to be practically useful, it must not affect error-free sites. Specifically, for quantum erasure detection, all populations and coherences in 𝒯𝒯\mathcal{T}caligraphic_T should be minimally affected by detection. We will reserve the term mid-circuit to situations when 𝒯𝒯\mathcal{T}caligraphic_T encodes a qubit and is minimally affected by conversion or detection.

Refer to caption
Figure 1: Framework for Site-Resolved Error Detection. (a) \mathcal{H}caligraphic_H is the Hilbert space on a tweezer site; 𝒯𝒯\mathcal{T}caligraphic_T is the target space; \mathcal{F}caligraphic_F, the detection manifold, consists of detectable states used to flag errors. Leakage errors transfer population from 𝒯𝒯\mathcal{T}caligraphic_T to \mathcal{E}caligraphic_E, which can subsequently be converted to \mathcal{F}caligraphic_F for detection. Loss of a particle results in an empty tweezer (|ket|\emptyset\rangle| ∅ ⟩). (b) Detection of total population in \mathcal{F}caligraphic_F is binarized into a bitstring {bi}subscript𝑏𝑖\{b_{i}\}{ italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT }, from which the error flag {ei}subscript𝑒𝑖\{e_{i}\}{ italic_e start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT } is obtained.

III Our Platform: Programmable Optical Tweezer Arrays of Laser-Cooled Molecules

Our experiment utilizes laser-cooled CaF molecules trapped in a programmable optical tweezer array. A molecular beam of CaF molecules is created using a cryogenic buffer gas cell Hutzler et al. (2012), laser-slowed Truppe et al. (2017a), and trapped in a magneto-optical trap (MOT) Anderegg et al. (2017); Truppe et al. (2017b). Subsequently, the molecules are further laser-cooled Cheuk et al. (2018), compressed Li et al. (2024), optically trapped Anderegg et al. (2018), and eventually transferred into a programmable 1D array of optical tweezer traps Anderegg et al. (2019); Holland et al. (2023a).

We detect CaF molecules via fluorescence imaging; in particular, only molecules in the X2Σ(v=0,N=1)superscript𝑋2Σformulae-sequence𝑣0𝑁1X^{2}\Sigma(v=0,N=1)italic_X start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Σ ( italic_v = 0 , italic_N = 1 ) rovibrational manifold can be cooled and imaged, where v𝑣vitalic_v (N𝑁Nitalic_N) denote vibrational (rotational) quantum numbers. For fluorescence imaging, light addressing the X2Σ(v=0,N=1)A2Π1/2(v=0,J=1/2,+)superscript𝑋2Σformulae-sequence𝑣0𝑁1superscript𝐴2subscriptΠ12formulae-sequence𝑣0𝐽12X^{2}\Sigma(v=0,N=1)\rightarrow A^{2}\Pi_{1/2}(v=0,J=1/2,+)italic_X start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Σ ( italic_v = 0 , italic_N = 1 ) → italic_A start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Π start_POSTSUBSCRIPT 1 / 2 end_POSTSUBSCRIPT ( italic_v = 0 , italic_J = 1 / 2 , + ) transition is applied, and fluorescence emitted on the same transition is collected on a camera. During each image, the molecules explore all 12 hyperfine states in X2Σ(v=0,N=1)superscript𝑋2Σformulae-sequence𝑣0𝑁1X^{2}\Sigma(v=0,N=1)italic_X start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Σ ( italic_v = 0 , italic_N = 1 ).

In this paper, no tweezer rearrangement is utilized, but an initial non-destructive image is taken to identify occupied tweezers. This process leaves molecules in the set of detectable states =X2Σ(v=0,N=1)superscript𝑋2Σformulae-sequence𝑣0𝑁1\mathcal{F}=X^{2}\Sigma(v=0,N=1)caligraphic_F = italic_X start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Σ ( italic_v = 0 , italic_N = 1 ) with a small probability of molecular loss (leakage into |ket\left|\emptyset\right\rangle| ∅ ⟩). All measurements are conditioned on tweezer sites that are initially identified to be occupied.

IV New Site-Resolved Detection Scheme with Rapid Resonant Imaging

Compared to detection requirements needed for preparing defect-free arrays, our work here, which concerns enhancing internal state purities and quantum erasure detection, has much more stringent detection requirements. Rather than only needing to preserve the total population in \mathcal{F}caligraphic_F, we instead need to either preserve the population of a single quantum state or preserve both population and coherence within a qubit subspace. These requirements necessitate a new detection scheme, which we describe in this section.

We first discuss the choice of \mathcal{F}caligraphic_F and 𝒯𝒯\mathcal{T}caligraphic_T. Because a measurement projects a qubit, the detection manifold \mathcal{F}caligraphic_F must be disjoint from the target state(s) 𝒯𝒯\mathcal{T}caligraphic_T. In particular, detected population in \mathcal{F}caligraphic_F will correspond to an error ({ei}={bi}subscript𝑒𝑖subscript𝑏𝑖\{e_{i}\}=\{b_{i}\}{ italic_e start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT } = { italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT }). In order for a measurement of \mathcal{F}caligraphic_F to minimally affect 𝒯𝒯\mathcal{T}caligraphic_T, a viable approach is to choose internal states for \mathcal{F}caligraphic_F and 𝒯𝒯\mathcal{T}caligraphic_T such that they have widely separated optical transitions compared to their optical linewidths. This general approach has been proposed and demonstrated in alkaline-earth atomic tweezer arrays Wu et al. (2022); Ma et al. (2023); Scholl et al. (2023a, b), where \mathcal{F}caligraphic_F and 𝒯𝒯\mathcal{T}caligraphic_T use separate ground and metastable electronic manifolds.

Here, instead of using different electronic manifolds, we make use of two different long-lived rotational manifolds available in a molecule to encode \mathcal{F}caligraphic_F and 𝒯𝒯\mathcal{T}caligraphic_T. Specifically, in our scheme for CaF molecules, we use the set of optical cycling states X2Σ(v=0,N=1)superscript𝑋2Σformulae-sequence𝑣0𝑁1X^{2}\Sigma(v=0,N=1)italic_X start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Σ ( italic_v = 0 , italic_N = 1 ) as \mathcal{F}caligraphic_F, and states from the ground rovibrational manifold X2Σ(v=0,N=0)superscript𝑋2Σformulae-sequence𝑣0𝑁0X^{2}\Sigma(v=0,N=0)italic_X start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Σ ( italic_v = 0 , italic_N = 0 ) to form the target subspace 𝒯𝒯\mathcal{T}caligraphic_T (Fig. 2(a)). The states in 𝒯𝒯\mathcal{T}caligraphic_T are separated from the detection manifold states \mathcal{F}caligraphic_F by a frequency of Δ2π×20GHzΔ2𝜋20GHz\Delta\approx 2\pi\times 20\,\text{GHz}roman_Δ ≈ 2 italic_π × 20 GHz, which is much larger than the optical linewidth Γ2π×10MHzΓ2𝜋10MHz\Gamma\approx 2\pi\times 10\,\text{MHz}roman_Γ ≈ 2 italic_π × 10 MHz of the relevant imaging transition used to detect \mathcal{F}caligraphic_F.

We next describe considerations for fluorescence detection of \mathcal{F}caligraphic_F. We desire to minimize errors for 𝒯𝒯\mathcal{T}caligraphic_T when measuring population in \mathcal{F}caligraphic_F. These can occur due to off-resonant photon scattering of imaging light that affects both populations and coherences in 𝒯𝒯\mathcal{T}caligraphic_T. A useful figure of merit is the ratio η𝜂\etaitalic_η of the off-resonant scattering rate on 𝒯𝒯\mathcal{T}caligraphic_T to the fluorescence rate of \mathcal{F}caligraphic_F. For our choice of \mathcal{F}caligraphic_F and 𝒯𝒯\mathcal{T}caligraphic_T, η𝜂\etaitalic_η can be as low as 9Γ2/Δ2106similar-to9superscriptΓ2superscriptΔ2superscript1069\Gamma^{2}/\Delta^{2}\sim 10^{-6}9 roman_Γ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / roman_Δ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ∼ 10 start_POSTSUPERSCRIPT - 6 end_POSTSUPERSCRIPT for on-resonant imaging Sup . Because we scatter 1001000similar-toabsent1001000\sim 100-1000∼ 100 - 1000 imaging photons to achieve sufficient signal-to-noise, our scheme has a minimum error of 103similar-toabsentsuperscript103\sim 10^{-3}∼ 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT, which predominantly takes the form of leakage into other molecular states. These leakage errors could in principle be corrected with more complex schemes that involve error conversion to \mathcal{F}caligraphic_F. We note that our on-resonant imaging scheme is different than that previously used for detection in CaF tweezer arrays, where the primary requirement was to preserve molecular population. This was accomplished using light detuned from resonance by 4Γsimilar-toabsent4Γ\sim 4\,\Gamma∼ 4 roman_Γ Cheuk et al. (2018); Anderegg et al. (2019); Holland et al. (2023a).

An additional consideration for error detection is crosstalk, where population leaks between \mathcal{F}caligraphic_F and 𝒯𝒯\mathcal{T}caligraphic_T. In our scheme, crosstalk is protected by selection rules. Because the optical transitions used are all E1 transitions, and =X2Σ(v=0,N=1)superscript𝑋2Σformulae-sequence𝑣0𝑁1\mathcal{F}=X^{2}\Sigma(v=0,N=1)caligraphic_F = italic_X start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Σ ( italic_v = 0 , italic_N = 1 ) and 𝒯X2Σ(v=0,N=0)𝒯superscript𝑋2Σformulae-sequence𝑣0𝑁0\mathcal{T}\subset X^{2}\Sigma(v=0,N=0)caligraphic_T ⊂ italic_X start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Σ ( italic_v = 0 , italic_N = 0 ) are of opposite parity, no population leakage occurs even in the presence of off-resonant scattering from imaging light.

We next describe how we implement on-resonant imaging and optimize its performance. Specifically, to image molecules in \mathcal{F}caligraphic_F, we apply light resonant with all resolved hyperfine transitions of the X2Σ(v=0,N=1)A2Π1/2(v=0,J=1/2,+)superscript𝑋2Σformulae-sequence𝑣0𝑁1superscript𝐴2subscriptΠ12formulae-sequence𝑣0𝐽12X^{2}\Sigma(v=0,N=1)-A^{2}\Pi_{1/2}(v=0,J=1/2,+)italic_X start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Σ ( italic_v = 0 , italic_N = 1 ) - italic_A start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Π start_POSTSUBSCRIPT 1 / 2 end_POSTSUBSCRIPT ( italic_v = 0 , italic_J = 1 / 2 , + ) line. Laser light addressing the X2Σ(v=1,2,3,N=1)A2Π1/2(v=v1,J=1/2,+)superscript𝑋2Σformulae-sequence𝑣123𝑁1superscript𝐴2subscriptΠ12formulae-sequencesuperscript𝑣𝑣1𝐽12X^{2}\Sigma(v=1,2,3,N=1)\rightarrow A^{2}\Pi_{1/2}(v^{\prime}=v-1,J=1/2,+)italic_X start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Σ ( italic_v = 1 , 2 , 3 , italic_N = 1 ) → italic_A start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Π start_POSTSUBSCRIPT 1 / 2 end_POSTSUBSCRIPT ( italic_v start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT = italic_v - 1 , italic_J = 1 / 2 , + ) transitions acting as vibrational repumpers is also applied. We additionally apply light addressing the X2Σ(v=0,N=3)B2Σ(v=0,N=2)superscript𝑋2Σformulae-sequence𝑣0𝑁3superscript𝐵2Σformulae-sequencesuperscript𝑣0𝑁2X^{2}\Sigma(v=0,N=3)\rightarrow B^{2}\Sigma(v^{\prime}=0,N=2)italic_X start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Σ ( italic_v = 0 , italic_N = 3 ) → italic_B start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Σ ( italic_v start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT = 0 , italic_N = 2 ) transition, which acts as a rotational repumper Holland et al. (2023a). These beams propagate perpendicular to the imaging axis, and cause no observable background for typical imaging durations.

Refer to caption
Figure 2: Site-Resolved Detection using Rapid Resonant Imaging. (a) Imaging Scheme. Molecules in the detection manifold =X2Σ(v=0,N=1)superscript𝑋2Σformulae-sequence𝑣0𝑁1\mathcal{F}=X^{2}\Sigma(v=0,N=1)caligraphic_F = italic_X start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Σ ( italic_v = 0 , italic_N = 1 ) scatter light resonant with the X2Σ(v=0,N=1)A2Π1/2(v=0,J=1/2,+)superscript𝑋2Σformulae-sequence𝑣0𝑁1superscript𝐴2subscriptΠ12formulae-sequence𝑣0𝐽12X^{2}\Sigma(v=0,N=1)\rightarrow A^{2}\Pi_{1/2}(v=0,J=1/2,+)italic_X start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Σ ( italic_v = 0 , italic_N = 1 ) → italic_A start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Π start_POSTSUBSCRIPT 1 / 2 end_POSTSUBSCRIPT ( italic_v = 0 , italic_J = 1 / 2 , + ) transition (orange arrows). Vibrational (red solid arrows) and rotational repum** light (blue solid arrow) is applied. The imaging light is off-resonant for states in the target manifold 𝒯X2Σ(v=0,N=0)𝒯superscript𝑋2Σformulae-sequence𝑣0𝑁0\mathcal{T}\subset X^{2}\Sigma(v=0,N=0)caligraphic_T ⊂ italic_X start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Σ ( italic_v = 0 , italic_N = 0 ). (b) Error Identification Performance. The false positive probability ϵ01subscriptitalic-ϵ01\epsilon_{01}italic_ϵ start_POSTSUBSCRIPT 01 end_POSTSUBSCRIPT (red squares) and false negative probability ϵ10subscriptitalic-ϵ10\epsilon_{10}italic_ϵ start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT (green diamonds) are shown versus the classification threshold θEsubscript𝜃E\theta_{\text{E}}italic_θ start_POSTSUBSCRIPT E end_POSTSUBSCRIPT. Vertical dashed line marks θE=θE,0=4.8subscript𝜃Esubscript𝜃E,04.8\theta_{\text{E}}=\theta_{\text{E,0}}=4.8italic_θ start_POSTSUBSCRIPT E end_POSTSUBSCRIPT = italic_θ start_POSTSUBSCRIPT E,0 end_POSTSUBSCRIPT = 4.8, the value used in this work. (c) Population P0subscript𝑃0P_{0}italic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT in |X2Σ(v=0,N=0),J=1/2,F=1,mF=1𝒯ketformulae-sequencesuperscript𝑋2Σformulae-sequence𝑣0𝑁0𝐽12formulae-sequence𝐹1subscript𝑚𝐹1𝒯\left|{X^{2}\Sigma(v=0,N=0),J=1/2,F=1,m_{F}=-1}\right\rangle\in\mathcal{T}| italic_X start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Σ ( italic_v = 0 , italic_N = 0 ) , italic_J = 1 / 2 , italic_F = 1 , italic_m start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT = - 1 ⟩ ∈ caligraphic_T versus imaging duration t𝑡titalic_t. An exponential fit gives a 1/e1𝑒1/e1 / italic_e lifetime of τ=790(60)ms𝜏79060ms\tau=790(60)\,\text{ms}italic_τ = 790 ( 60 ) ms, much longer than the imaging duration of 3ms3ms3\,\text{ms}3 ms.

To optimize imaging for the purpose of identifying errors, we prioritize minimizing the false negative probability (ϵ10subscriptitalic-ϵ10\epsilon_{10}italic_ϵ start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT), since it is the probability that errors remain undetected. We note that for tweezer preparation, the false positive probability (ϵ01subscriptitalic-ϵ01\epsilon_{01}italic_ϵ start_POSTSUBSCRIPT 01 end_POSTSUBSCRIPT) only affects the data rate, since any site flagged as an error could be discarded during subsequent rearrangement or post-selection. We therefore seek to optimize ϵ10subscriptitalic-ϵ10\epsilon_{10}italic_ϵ start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT with respect to the imaging light intensity (X2Σ(v=0,N=1)A2Π1/2(v=0,J=1/2,+)superscript𝑋2Σformulae-sequence𝑣0𝑁1superscript𝐴2subscriptΠ12formulae-sequence𝑣0𝐽12X^{2}\Sigma(v=0,N=1)-A^{2}\Pi_{1/2}(v=0,J=1/2,+)italic_X start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Σ ( italic_v = 0 , italic_N = 1 ) - italic_A start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Π start_POSTSUBSCRIPT 1 / 2 end_POSTSUBSCRIPT ( italic_v = 0 , italic_J = 1 / 2 , + )), the imaging duration, and the tweezer depth. Specifically, we minimize ϵ10subscriptitalic-ϵ10\epsilon_{10}italic_ϵ start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT at a fixed threshold of θE=θE,0=4.8subscript𝜃Esubscript𝜃E,04.8\theta_{\text{E}}=\theta_{\text{E,0}}=4.8italic_θ start_POSTSUBSCRIPT E end_POSTSUBSCRIPT = italic_θ start_POSTSUBSCRIPT E,0 end_POSTSUBSCRIPT = 4.8, and achieve a false positive probability (ϵ10(θE,0)=0.033(4)subscriptitalic-ϵ10subscript𝜃E,00.0334\epsilon_{10}(\theta_{\text{E,0}})=0.033(4)italic_ϵ start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT ( italic_θ start_POSTSUBSCRIPT E,0 end_POSTSUBSCRIPT ) = 0.033 ( 4 )) at an imaging intensity of I=4.5mW/cm2𝐼4.5superscriptmW/cm2I=4.5\,\text{mW/cm}^{2}italic_I = 4.5 mW/cm start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, imaging duration of 3ms3ms3\,\text{ms}3 ms, and tweezer depth of kB×930(20)μKsubscript𝑘𝐵93020𝜇Kk_{B}\times 930(20)\,\mu\text{K}italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT × 930 ( 20 ) italic_μ K (Fig. 2(b)). In principle, for detectable errors, these values allow reduction of intrinsic error rates by up to a factor of 30similar-toabsent30\sim 30∼ 30. Henceforth, θEsubscript𝜃E\theta_{\text{E}}italic_θ start_POSTSUBSCRIPT E end_POSTSUBSCRIPT is fixed to θE,0subscript𝜃E,0\theta_{\text{E,0}}italic_θ start_POSTSUBSCRIPT E,0 end_POSTSUBSCRIPT unless noted. Because the background is low enough to be largely independent of the imaging duration, ϵ01subscriptitalic-ϵ01\epsilon_{01}italic_ϵ start_POSTSUBSCRIPT 01 end_POSTSUBSCRIPT, which determines the data rate, is independent of imaging parameters and only depends on camera noise.

To verify that our site-resolved detection scheme minimally affects population in 𝒯𝒯\mathcal{T}caligraphic_T, which is composed of states from X2Σ(v=0,N=0)superscript𝑋2Σformulae-sequence𝑣0𝑁0X^{2}\Sigma(v=0,N=0)italic_X start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Σ ( italic_v = 0 , italic_N = 0 ), we prepare molecules in |X2Σ(v=0,N=0),J=1/2,F=1,mF=1ketformulae-sequencesuperscript𝑋2Σformulae-sequence𝑣0𝑁0𝐽12formulae-sequence𝐹1subscript𝑚𝐹1|X^{2}\Sigma(v=0,N=0),J=1/2,F=1,m_{F}=-1\rangle| italic_X start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Σ ( italic_v = 0 , italic_N = 0 ) , italic_J = 1 / 2 , italic_F = 1 , italic_m start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT = - 1 ⟩ and measure the population loss per image (Fig. 2(c)). We find a loss rate of 3.8(3)×1033.83superscript1033.8(3)\times 10^{-3}3.8 ( 3 ) × 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT per image, confirming that our scheme minimally affects the population in X2Σ(v=0,N=0)superscript𝑋2Σformulae-sequence𝑣0𝑁0X^{2}\Sigma(v=0,N=0)italic_X start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Σ ( italic_v = 0 , italic_N = 0 ).

Refer to caption
Figure 3: Robust and Enhanced Tweezer Preparation Using Site-Resolved Error Detection. (a) Our enhanced internal state preparation procedure involves: (i) optical pum** to |Dketsuperscript𝐷|D^{\prime}\rangle| italic_D start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ⟩ (orange arrow), (ii) population transfer to 𝒯=|𝒯ket\mathcal{T}=|-\ranglecaligraphic_T = | - ⟩ via a microwave pulse (grey arrow), and (iii) error detection of left-over molecules in \mathcal{F}caligraphic_F. (b) Verifying Tweezer Preparation. A non-destructive image (top) identifies initially loaded sites. The detection-enhanced internal state preparation procedure is applied, resulting in an error image (middle) that flags sites with internal state errors. For verification, we state-selectively transfer molecules from 𝒯=|𝒯ket\mathcal{T}=\left|-\right\ranglecaligraphic_T = | - ⟩ to |ket\left|\uparrow\right\rangle| ↑ ⟩ for a final destructive measurement (bottom). (c) The detection-conditioned probability p~Esubscript~𝑝𝐸\tilde{p}_{E}over~ start_ARG italic_p end_ARG start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT of a molecule appearing in the final image versus optical pum** duration t𝑡titalic_t, which is used to control the bare optical pum** fidelity fOPsubscript𝑓OPf_{\text{OP}}italic_f start_POSTSUBSCRIPT OP end_POSTSUBSCRIPT. p~Esubscript~𝑝𝐸\tilde{p}_{E}over~ start_ARG italic_p end_ARG start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT for various error-detection image classification thresholds θE=(4.6,5.6,6.6,)subscript𝜃𝐸4.65.66.6\theta_{E}=(4.6,5.6,6.6,\infty)italic_θ start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT = ( 4.6 , 5.6 , 6.6 , ∞ ) (blue circles, red triangles, green squares, gray diamonds) are shown. Colored bands show corresponding theoretical predictions. (d) Error-excised tweezer fidelity fEsubscript𝑓𝐸f_{E}italic_f start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT (probability of being occupied and in the target internal state) versus bare tweezer fidelity f𝑓fitalic_f. Inset shows the error-excised internal state purity pEsubscript𝑝𝐸p_{E}italic_p start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT versus the optical pum** fidelity fOPsubscript𝑓OPf_{\text{OP}}italic_f start_POSTSUBSCRIPT OP end_POSTSUBSCRIPT.

V Robust and Enhanced Internal State Preparation Using Site-Resolved Detection of Errors

We next make use of our site-resolved error detection scheme to achieve robust and enhanced tweezer preparation fidelities. In previous work, the fidelity of having a tweezer loaded with a molecule in a target internal state was reported to be 80%absentpercent80\approx 80\%≈ 80 %, limited both by optical pum** fidelities and imperfect microwave transfers Holland et al. (2023b); Bao et al. (2023a).

Here, we present a scheme that allows internal state errors to appear as detectable errors, which can be subsequently corrected mid-sequence. Our scheme works as follows (Fig. 3(a)). Initially, molecules are distributed over all 12 hyperfine states in X2Σ(v=0,N=1)superscript𝑋2Σformulae-sequence𝑣0𝑁1X^{2}\Sigma(v=0,N=1)italic_X start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Σ ( italic_v = 0 , italic_N = 1 ), the laser-coolable manifold, which is also the detection manifold \mathcal{F}caligraphic_F. We optically pump the molecules into a single hyperfine state |D=|X2Σ(v=0,N=1),J=3/2,F=2,mF=2ketsuperscript𝐷ketformulae-sequencesuperscript𝑋2Σformulae-sequence𝑣0𝑁1𝐽32formulae-sequence𝐹2subscript𝑚𝐹2|D^{\prime}\rangle=|X^{2}\Sigma(v=0,N=1),J=3/2,F=2,m_{F}=-2\rangle\in\mathcal{F}| italic_D start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ⟩ = | italic_X start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Σ ( italic_v = 0 , italic_N = 1 ) , italic_J = 3 / 2 , italic_F = 2 , italic_m start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT = - 2 ⟩ ∈ caligraphic_F. Subsequently, we “shelve” them into the ground rotational manifold X2Σ(v=0,N=0)superscript𝑋2Σformulae-sequence𝑣0𝑁0X^{2}\Sigma(v=0,N=0)italic_X start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Σ ( italic_v = 0 , italic_N = 0 ). Specifically, we use a microwave pulse with a frequency of 20.5GHzabsent20.5GHz\approx 20.5\,\text{GHz}≈ 20.5 GHz to state-selectively transfer |Dketsuperscript𝐷|D^{\prime}\rangle| italic_D start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ⟩ to a single internal state |=|X2Σ(v=0,N=0),J=1/2,F=1,mF=1ketketformulae-sequencesuperscript𝑋2Σformulae-sequence𝑣0𝑁0𝐽12formulae-sequence𝐹1subscript𝑚𝐹1|-\rangle=|X^{2}\Sigma(v=0,N=0),J=1/2,F=1,m_{F}=-1\rangle| - ⟩ = | italic_X start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Σ ( italic_v = 0 , italic_N = 0 ) , italic_J = 1 / 2 , italic_F = 1 , italic_m start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT = - 1 ⟩, which is our target state (𝒯={|}𝒯ket\mathcal{T}=\{|-\rangle\}caligraphic_T = { | - ⟩ }). All internal state preparation errors leave the molecule in \mathcal{F}caligraphic_F, that is, =\mathcal{E}=\mathcal{F}caligraphic_E = caligraphic_F. Specifically, optical pum** errors leave a molecule in (|D)ketsuperscript𝐷(\mathcal{F}-\left|D^{\prime}\right\rangle)\subset\mathcal{F}( caligraphic_F - | italic_D start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ⟩ ) ⊂ caligraphic_F, and imperfect microwave transfer leaves a molecule in |Dketsuperscript𝐷\left|D^{\prime}\right\rangle\in\mathcal{F}| italic_D start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ⟩ ∈ caligraphic_F. Internal state preparation errors thus appear directly as detectable errors. They can then be identified by the site-resolved detection scheme described in Section IV, which minimally affects population in 𝒯={|}𝒯ket\mathcal{T}=\{|-\rangle\}caligraphic_T = { | - ⟩ } on error-free sites.

We next implement our scheme and examine its performance. Following internal state preparation into 𝒯𝒯\mathcal{T}caligraphic_T, we perform one error detection image. Subsequently, we transfer molecules from the target state |ket\left|-\right\rangle| - ⟩ to |=|X2Σ(v=0,N=1),J=1/2,F=0,mF=0)|{\uparrow}\rangle=\left|X^{2}\Sigma(v=0,N=1),J=1/2,F=0,m_{F}=0)\right\rangle| ↑ ⟩ = | italic_X start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Σ ( italic_v = 0 , italic_N = 1 ) , italic_J = 1 / 2 , italic_F = 0 , italic_m start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT = 0 ) ⟩, which is part of the detection manifold \mathcal{F}caligraphic_F (Fig. 3(b)). Finally, we destructively measure the total population in \mathcal{F}caligraphic_F.

To evaluate the robustness of our scheme, we intentionally vary the amount of optical pum** errors by varying the optical pum** duration t𝑡titalic_t. For each optical pum** duration, which serves as a proxy for the optical pum** fidelity, we extract p~Esubscript~𝑝𝐸\tilde{p}_{E}over~ start_ARG italic_p end_ARG start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT, the probability that we detect a molecule in the final image conditioned upon no identified error. In (Fig. 3(c)), we show p~Esubscript~𝑝𝐸\tilde{p}_{E}over~ start_ARG italic_p end_ARG start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT for various error detection thresholds θEsubscript𝜃𝐸\theta_{E}italic_θ start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT. As expected, identifying errors more aggressively by lowering θEsubscript𝜃𝐸\theta_{E}italic_θ start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT (smaller false-negative probability ϵ10subscriptitalic-ϵ10\epsilon_{10}italic_ϵ start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT and larger false-positive probability ϵ01subscriptitalic-ϵ01\epsilon_{01}italic_ϵ start_POSTSUBSCRIPT 01 end_POSTSUBSCRIPT) improves p~Esubscript~𝑝𝐸\tilde{p}_{E}over~ start_ARG italic_p end_ARG start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT.

To show the gains of our scheme directly, we compare the internal state purity conditioned upon no errors, pEsubscript𝑝Ep_{\text{E}}italic_p start_POSTSUBSCRIPT E end_POSTSUBSCRIPT, with the optical pum** fidelity fOPsubscript𝑓OPf_{\text{OP}}italic_f start_POSTSUBSCRIPT OP end_POSTSUBSCRIPT. We determine the experimentally achieved fOPsubscript𝑓OPf_{\text{OP}}italic_f start_POSTSUBSCRIPT OP end_POSTSUBSCRIPT for each pum** duration t𝑡titalic_t using p~Esubscript~𝑝𝐸\tilde{p}_{E}over~ start_ARG italic_p end_ARG start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT and the error imaging infidelities ϵ01subscriptitalic-ϵ01\epsilon_{01}italic_ϵ start_POSTSUBSCRIPT 01 end_POSTSUBSCRIPT and ϵ10subscriptitalic-ϵ10\epsilon_{10}italic_ϵ start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT Sup . As shown in Fig. 3(d, inset), error excision allows significantly improved internal state purity, especially at low optical pum** fidelities. Notably, we achieve an internal state purity of pE=0.995(1)subscript𝑝𝐸0.9951p_{E}=0.995(1)italic_p start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT = 0.995 ( 1 ) when the achieved bare optical pum** fidelity is fOP0.96subscript𝑓OP0.96f_{\text{OP}}\approx 0.96italic_f start_POSTSUBSCRIPT OP end_POSTSUBSCRIPT ≈ 0.96. Notably, pEsubscript𝑝𝐸p_{E}italic_p start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT remains high even when fOPsubscript𝑓𝑂𝑃f_{OP}italic_f start_POSTSUBSCRIPT italic_O italic_P end_POSTSUBSCRIPT is degraded. For example, at an optical pum** fidelity of fOP=0.46subscript𝑓OP0.46f_{\text{OP}}=0.46italic_f start_POSTSUBSCRIPT OP end_POSTSUBSCRIPT = 0.46, the error-excised internal state purity remains at pE0.96subscript𝑝𝐸0.96p_{E}\approx 0.96italic_p start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT ≈ 0.96. Finally, we extract the error-excised tweezer preparation fidelity fEsubscript𝑓𝐸f_{E}italic_f start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT, defined as the probability that a tweezer site is occupied with a molecule in the target internal state when conditioned upon an absence of a detected error. This is the practically useful metric for preparing low-defect arrays using tweezer rearrangement. In Fig. 3(d), we show fEsubscript𝑓𝐸f_{E}italic_f start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT versus the bare tweezer preparation fidelity f𝑓fitalic_f, which takes into account occupation errors. As expected, we find that error excision provides significant improvement and robustness. Notably, we achieve a tweezer preparation fidelity of fE=0.952(3)subscript𝑓𝐸0.9523f_{E}=0.952(3)italic_f start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT = 0.952 ( 3 ), with an estimated reduction in data rate of only 7.2(2)%7.2percent27.2(2)\%7.2 ( 2 ) %. This significantly improves upon previous reported values (f80%𝑓percent80f\approx 80\%italic_f ≈ 80 %Holland et al. (2023b); Bao et al. (2023a); Lu et al. (2024); Bao et al. (2023b) and importantly, our method is robust to day-to-day fluctuations in fOPsubscript𝑓OPf_{\text{OP}}italic_f start_POSTSUBSCRIPT OP end_POSTSUBSCRIPT.

We note although all error-conditioned measurements in this work are done via post-selection, where we excise sites that are error-flagged, our measurements indicate that our detection scheme minimally affects correctly prepared molecules. Thus, in combination with the ability to perform low-loss rearrangement (previously demonstrated to have losses 0.1%similar-toabsentpercent0.1\sim 0.1\%∼ 0.1 % Holland et al. (2023b)), our scheme can be used with another round of tweezer rearrangement to achieve similar final fidelities. This will allow the creation of arrays with low defect rates both in spatial configuration and internal state purity, which is crucial to quantum computation and simulation applications with large-scale molecular tweezer arrays. In particular, the ability to create arrays with few-percent level errors opens the door to simulating interacting quantum spin systems with low defect rates. Specifically, these low defect rates would allow practical post-selection for perfect arrays (possible with full state-resolved detection) for up to 10 sites, with data rates reduced by only 40%similar-toabsentpercent40\sim 40\%∼ 40 %.

VI A Quantum Erasure Detection Scheme For CaF Molecules

Refer to caption
Figure 4: Coherent Control and Coherence Times of New Hyperfine Qubit. (a) Previous rotational qubit encoding |ket\left|{\downarrow}\right\rangle| ↓ ⟩-|ket\left|{\uparrow}\right\rangle| ↑ ⟩ (boxed in blue) is not compatible with our error detection scheme, since |ket\left|\uparrow\right\rangle| ↑ ⟩ is in the detection manifold \mathcal{F}caligraphic_F. Our new qubit encoding |ket\left|{\downarrow}\right\rangle| ↓ ⟩-|0ket0\left|0\right\rangle| 0 ⟩ uses two hyperfine states in X2Σ(v=0,N=0)superscript𝑋2Σformulae-sequence𝑣0𝑁0X^{2}\Sigma(v=0,N=0)italic_X start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Σ ( italic_v = 0 , italic_N = 0 ) that are outside of \mathcal{F}caligraphic_F (boxed in green). (b) Coherent manipulations of |ket|{\downarrow}\rangle| ↓ ⟩-|0ket0|0\rangle| 0 ⟩ is performed using a two-photon Raman processes. Our two schemes rely on light near-detuned to the X𝑋Xitalic_X-A𝐴Aitalic_A and X𝑋Xitalic_X-B𝐵Bitalic_B optical transitions. (c) Molecules are prepared in |ket\left|\downarrow\right\rangle| ↓ ⟩ and driven using Raman light for a variable duration t𝑡titalic_t. The top (bottom) panels show Rabi oscillations of the |ket\left|\downarrow\right\rangle| ↓ ⟩ population when driven by X𝑋Xitalic_X-A𝐴Aitalic_A (X𝑋Xitalic_X-B𝐵Bitalic_B) Raman light. Solid lines show fits to damped sinusoids. (d) Hyperfine qubit coherence versus hold time is measured via a Ramsey sequence (top panel). Data points show measured Ramsey fringe amplitudes versus hold time t𝑡titalic_t, with solid lines showing fits to exponential decay curves. Data without (with) a spin-echo pulse are shown by blue circles (red triangles); data with XY8 dynamical decoupling pulses using X𝑋Xitalic_X-B𝐵Bitalic_B(X𝑋Xitalic_X-A𝐴Aitalic_A) light are shown by green squares (orange diamonds). The fitted 1/e1𝑒1/e1 / italic_e lifetimes are 19.5(7)ms19.57ms19.5(7)\,\text{ms}19.5 ( 7 ) ms, 288(15)ms28815ms288(15)\,\text{ms}288 ( 15 ) ms, 630(50)ms63050ms630(50)\,\text{ms}630 ( 50 ) ms, and 1100(70)ms110070ms1100(70)\,\text{ms}1100 ( 70 ) ms, respectively. The dashed green (orange) line shows the estimated population decay due to off-resonant scattering for the X𝑋Xitalic_X-B𝐵Bitalic_B (X𝑋Xitalic_X-A𝐴Aitalic_A) Raman configuration Sup .

In the second part of our work, we demonstrate erasure detection for the first time in molecules. Compared to enhancing tweezer preparation using site-resolved detection, where only the total population in \mathcal{F}caligraphic_F needs to be preserved, erasure detection is more challenging since both population and coherence in error-free qubits should be minimally affected.

VI.1 A New Hyperfine Qubit Encoding: Coherent Control and Qubit Coherence

To implement quantum erasure detection, we first seek a qubit encoding that is compatible with our site-resolved error detection scheme described in Section IV. In previous work with CaF molecules, qubits were encoded using two states from neighboring rotational manifolds, namely, |=|X2Σ(v=0,N=1),J=1/2,F=0,mF=0ketketformulae-sequencesuperscript𝑋2Σformulae-sequence𝑣0𝑁1𝐽12formulae-sequence𝐹0subscript𝑚𝐹0|{\uparrow}\rangle=|X^{2}\Sigma(v=0,N=1),J=1/2,F=0,m_{F}=0\rangle| ↑ ⟩ = | italic_X start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Σ ( italic_v = 0 , italic_N = 1 ) , italic_J = 1 / 2 , italic_F = 0 , italic_m start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT = 0 ⟩ and |=|X2Σ(v=0,N=0),J=1/2,F=1,mF=0ketketformulae-sequencesuperscript𝑋2Σformulae-sequence𝑣0𝑁0𝐽12formulae-sequence𝐹1subscript𝑚𝐹0|{\downarrow}\rangle=|X^{2}\Sigma(v=0,N=0),J=1/2,F=1,m_{F}=0\rangle| ↓ ⟩ = | italic_X start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Σ ( italic_v = 0 , italic_N = 0 ) , italic_J = 1 / 2 , italic_F = 1 , italic_m start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT = 0 ⟩ Burchesky et al. (2021); Holland et al. (2023b); Bao et al. (2023a). This rotational encoding was used to demonstrate a 2-qubit iSWAP gate, but is incompatible with our detection scheme because |ket|{\uparrow}\rangle| ↑ ⟩ is part of the detection manifold \mathcal{F}caligraphic_F. The detection of \mathcal{F}caligraphic_F, which is a projective measurement of its population, will destroy qubit coherence between |ket|{\uparrow}\rangle| ↑ ⟩ and |ket|{\downarrow}\rangle| ↓ ⟩.

Here, we use a new qubit encoding that makes use of two hyperfine states within the ground rovibrational manifold X2Σ(v=0,N=0)superscript𝑋2Σformulae-sequence𝑣0𝑁0X^{2}\Sigma(v=0,N=0)italic_X start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Σ ( italic_v = 0 , italic_N = 0 ), which are dark to the imaging light used for site-resolved error detection. Specifically, the hyperfine encoding uses the states |=|X2Σ(v=0,N=0),J=1/2,F=1,mF=0ketketformulae-sequencesuperscript𝑋2Σformulae-sequence𝑣0𝑁0𝐽12formulae-sequence𝐹1subscript𝑚𝐹0|{\downarrow}\rangle=|X^{2}\Sigma(v=0,N=0),J=1/2,F=1,m_{F}=0\rangle| ↓ ⟩ = | italic_X start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Σ ( italic_v = 0 , italic_N = 0 ) , italic_J = 1 / 2 , italic_F = 1 , italic_m start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT = 0 ⟩ and |0=|X2Σ(v=0,N=0),J=1/2,F=0,mF=0ket0ketformulae-sequencesuperscript𝑋2Σformulae-sequence𝑣0𝑁0𝐽12formulae-sequence𝐹0subscript𝑚𝐹0|0\rangle=|X^{2}\Sigma(v=0,N=0),J=1/2,F=0,m_{F}=0\rangle| 0 ⟩ = | italic_X start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Σ ( italic_v = 0 , italic_N = 0 ) , italic_J = 1 / 2 , italic_F = 0 , italic_m start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT = 0 ⟩ (Fig. 4(a)). These states are predicted to have long coherence times because of the absence of tensor ac Stark shifts and their low magnetic moments. In the framework of Section II, the detection manifold \mathcal{F}caligraphic_F remains the same (=X2Σ(v=0,N=1)superscript𝑋2Σformulae-sequence𝑣0𝑁1\mathcal{F}=X^{2}\Sigma(v=0,N=1)caligraphic_F = italic_X start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Σ ( italic_v = 0 , italic_N = 1 )), while the target subspace 𝒯={|,|0}𝒯ketket0\mathcal{T}=\{|{\downarrow}\rangle,|{0}\rangle\}caligraphic_T = { | ↓ ⟩ , | 0 ⟩ } now consists of the two hyperfine qubit states.

We first demonstrate coherent control over this hyperfine qubit, and that it has long coherence times. The two qubit states are connected by an M1 transition at 122MHzsimilar-toabsent122MHz\sim 122\,\text{MHz}∼ 122 MHz. We coherently manipulate the qubit with two-photon Raman transitions rather than directly driving the M1 transition. Specifically, we use Raman light near-detuned to either the XA𝑋𝐴X-Aitalic_X - italic_A or XB𝑋𝐵X-Bitalic_X - italic_B optical transitions (Fig. 4(b)). The two frequencies in the Raman light co-propagate and drive σ+superscript𝜎\sigma^{+}italic_σ start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT transitions. To demonstrate coherent control, we prepare molecules in |ket|{\downarrow}\rangle| ↓ ⟩, and apply Raman light for a variable duration. As shown in Fig. 4(c), we observe Rabi oscillations with Rabi frequencies as high as ΩR2π×10kHzsubscriptΩ𝑅2𝜋10kHz\Omega_{R}\approx 2\pi\times 10\,\text{kHz}roman_Ω start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT ≈ 2 italic_π × 10 kHz for both Raman schemes.

We next measure the qubit coherence time using a Ramsey spectroscopy sequence consisting of two Raman π/2𝜋2\pi/2italic_π / 2 pulses separated by a variable duration. In order to minimize the effect of inhomogeneous light shifts among the tweezer sites, we operate at a tweezer depth of Uhf=kB×39(1)μKsubscript𝑈hfsubscript𝑘𝐵391𝜇KU_{\text{hf}}=k_{B}\times 39(1)\,\mu\text{K}italic_U start_POSTSUBSCRIPT hf end_POSTSUBSCRIPT = italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT × 39 ( 1 ) italic_μ K. By measuring the decay rate of the amplitude of the Ramsey fringe, we find a bare coherence time of T2=19.5(7)mssuperscriptsubscript𝑇219.57msT_{2}^{*}=19.5(7)\,\text{ms}italic_T start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT = 19.5 ( 7 ) ms, which can be increased to T2=288(15)mssubscript𝑇228815msT_{2}=288(15)\,\text{ms}italic_T start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = 288 ( 15 ) ms with a single spin-echo. The coherence time can be further extended by applying dynamical decoupling. Specifically, using a XY8 pulse sequence (pulse separation of 33.3 ms), the coherence times increase to 630(50)ms63050ms630(50)\,\text{ms}630 ( 50 ) ms and 1100(70)ms110070ms1100(70)\,\text{ms}1100 ( 70 ) ms for the X𝑋Xitalic_X-B𝐵Bitalic_B Raman scheme and X𝑋Xitalic_X-A𝐴Aitalic_A Raman scheme, respectively (Fig. 4(d)).

In addition to long coherence times, we note that our hyperfine qubit encoding is also compatible with the previously used rotational qubit encoding Burchesky et al. (2021); Holland et al. (2023b); Bao et al. (2023a). We have verified that the three states involved can remain simultaneously coherent, and that we can coherently transfer information encoded in a rotational qubit (|,|ketket|{\uparrow}\rangle,|{\downarrow}\rangle| ↑ ⟩ , | ↓ ⟩) to a hyperfine encoding (|,|0ketket0|{\downarrow}\rangle,|{0}\rangle| ↓ ⟩ , | 0 ⟩Sup .

Refer to caption
Figure 5: Mid-Circuit Quantum Erasure Detection. (a) Our composite erasure detection scheme comprises a tweezer depth increase, an error detection image, and a mid-detection X𝑋Xitalic_X-B𝐵Bitalic_B Raman π𝜋\piitalic_π-pulse. (b) Ramsey fringe offset versus the number of composite erasure detections Nrepsubscript𝑁repN_{\text{rep}}italic_N start_POSTSUBSCRIPT rep end_POSTSUBSCRIPT. (c) Ramsey fringe amplitude versus Nrepsubscript𝑁repN_{\text{rep}}italic_N start_POSTSUBSCRIPT rep end_POSTSUBSCRIPT. For (b,c), green circles (orange triangles) are data obtained using the X𝑋Xitalic_X-B𝐵Bitalic_B (X𝑋Xitalic_X-A𝐴Aitalic_A) Raman scheme.(d) Population loss measurements from combinations of tweezer ramp (T), error detection light (E), and v=1𝑣1v=1italic_v = 1 repump (V). The plot shows population Psubscript𝑃P_{\uparrow}italic_P start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT versus Nrepsubscript𝑁repN_{\text{rep}}italic_N start_POSTSUBSCRIPT rep end_POSTSUBSCRIPT cycles of TEV (blue circles), EV (red triangles), T (purple squares). Brown diamonds are measured at a shallow tweezer depth without TEV. The brown, purple, and red arrows indicate contributions from background vacuum and blackbody loss, tweezer ramp, and erasure detection. Inset: The sum of these measured loss rate contributions (colored bars summing to gray pentagon) is consistent with the total loss (blue bar and circle).

VI.2 Composite Erasure Detection Sequence to Preserve Qubit Coherence

Having established long coherence times and coherent control of the hyperfine qubit, we next investigate whether qubit coherence is preserved after an error detection image. Compared to the tweezer depth Uhfsubscript𝑈hfU_{\text{hf}}italic_U start_POSTSUBSCRIPT hf end_POSTSUBSCRIPT used to obtain long hyperfine qubit coherence times, we find that a much higher tweezer depth of UED=kB×930(20)μKsubscript𝑈EDsubscript𝑘𝐵93020𝜇KU_{\text{ED}}=k_{B}\times 930(20)\,\mu\text{K}italic_U start_POSTSUBSCRIPT ED end_POSTSUBSCRIPT = italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT × 930 ( 20 ) italic_μ K is needed to obtain high error detection fidelities.

Therefore, for erasure detection, the tweezer trap needs to be ramped up by a factor of 30similar-toabsent30\sim 30∼ 30 from Uhfsubscript𝑈hfU_{\text{hf}}italic_U start_POSTSUBSCRIPT hf end_POSTSUBSCRIPT to UEDsubscript𝑈EDU_{\text{ED}}italic_U start_POSTSUBSCRIPT ED end_POSTSUBSCRIPT. At the deeper depth, the qubit frequency shifts by 2π×0.40(6)kHz2𝜋0.406kHz2\pi\times 0.40(6)\,\text{kHz}2 italic_π × 0.40 ( 6 ) kHz, which on average leads to additional phase accumulation in the hyperfine qubit. This phase accumulation is inconsequential if the tweezers had uniform depth. However, our tweezer array has depth inhomogeneity of 10%similar-toabsentpercent10\sim 10\%∼ 10 %, giving rise to an effective phase variation of 1radabsent1rad\approx 1\,\text{rad}≈ 1 rad at UEDsubscript𝑈EDU_{\text{ED}}italic_U start_POSTSUBSCRIPT ED end_POSTSUBSCRIPT over the imaging duration. To counteract this inhomogeneous dephasing, we implement a “composite erasure detection” scheme consisting of four steps (Fig. 5(a)): 1) the tweezer depth is ramped from Uhfsubscript𝑈hfU_{\text{hf}}italic_U start_POSTSUBSCRIPT hf end_POSTSUBSCRIPT to UEDsubscript𝑈EDU_{\text{ED}}italic_U start_POSTSUBSCRIPT ED end_POSTSUBSCRIPT, 2) local errors are detected using rapid resonant imaging, 3) a π𝜋\piitalic_π-pulse is applied to the qubit halfway through the image, and 4) the tweezer depth is returned to Uhfsubscript𝑈hfU_{\text{hf}}italic_U start_POSTSUBSCRIPT hf end_POSTSUBSCRIPT. The π𝜋\piitalic_π-pulse echos out the effects of inhomogeneous light shifts due to tweezer depth variations, along with Stark shifts arising from the imaging light.

To determine how well the qubit population and coherence are preserved during composite erasure detection, we repeatedly apply composite erasure detection between two π𝜋\piitalic_π/2 pulses in a Ramsey sequence. We find that a Ramsey fringe is observable even after multiple images. The offset of the Ramsey fringe is proportional to the total population, and the amplitude is proportional to both the total population and coherence of the qubit states. Per image, we determine a fractional population reduction of εp=3.3(2)×102subscript𝜀𝑝3.32superscript102\varepsilon_{p}=3.3(2)\times 10^{-2}italic_ε start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT = 3.3 ( 2 ) × 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT (Fig. 5(b)) and a fractional amplitude reduction of εc=4.6(3)×102subscript𝜀𝑐4.63superscript102\varepsilon_{c}=4.6(3)\times 10^{-2}italic_ε start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = 4.6 ( 3 ) × 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT (Fig. 5(c)). These measurements indicate that our composite erasure detection sequence affects the hyperfine qubit population and coherence only at the few percent level. To our knowledge, this is the first demonstration of erasure detection in molecules.

We note that for the purposes of composite erasure detection, we only use XB𝑋𝐵X-Bitalic_X - italic_B Raman light, which has much shorter population lifetimes and coherence times compared to our XA𝑋𝐴X-Aitalic_X - italic_A Raman scheme due to higher off-resonant scattering. This is because our XA𝑋𝐴X-Aitalic_X - italic_A Raman light (due to constraints on laser availability) is resonant with the detection manifold =X2Σ(v=0,N=1)superscript𝑋2Σformulae-sequence𝑣0𝑁1\mathcal{F}=X^{2}\Sigma(v=0,N=1)caligraphic_F = italic_X start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Σ ( italic_v = 0 , italic_N = 1 ). It resonantly heats a molecule in \mathcal{F}caligraphic_F out of a tweezer, and is therefore not compatible with erasure detection. This limitation can easily be overcome with a XA𝑋𝐴X-Aitalic_X - italic_A Raman light source at a more optimal detuning. As we describe in detail in the next section, we anticipate that errors at the 103superscript10310^{-3}10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT level are achievable.

VI.3 Technical and Fundamental Limits of Our Erasure Detection Scheme

To understand the origin of the 102similar-toabsentsuperscript102\sim 10^{-2}∼ 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT level errors per erasure image, we independently measure the contributions of each step in our composite erasure detection sequence. First, we quantify the effect of off-resonant photon scattering on the qubit states caused by the spin-echo π𝜋\piitalic_π-pulse. We perform a similar Ramsey measurement with only π𝜋\piitalic_π pulses applied (without tweezer ramps and imaging light). This gives a population error per image of εp,R=2.7(2)×102subscript𝜀𝑝𝑅2.72superscript102\varepsilon_{p,R}=2.7(2)\times 10^{-2}italic_ε start_POSTSUBSCRIPT italic_p , italic_R end_POSTSUBSCRIPT = 2.7 ( 2 ) × 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT and a coherence error per image of εc,R=4.4(4)×102subscript𝜀𝑐𝑅4.44superscript102\varepsilon_{c,R}=4.4(4)\times 10^{-2}italic_ε start_POSTSUBSCRIPT italic_c , italic_R end_POSTSUBSCRIPT = 4.4 ( 4 ) × 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT. Comparing these with the total composite erasure errors εpsubscript𝜀𝑝\varepsilon_{p}italic_ε start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT and εcsubscript𝜀𝑐\varepsilon_{c}italic_ε start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, we conclude that our present scheme is limited by off-resonant scattering of Raman light by the qubit states in 𝒯𝒯\mathcal{T}caligraphic_T.

Off-resonant scattering can be significantly improved by a more optimal Raman scheme. In particular, we expect that coherence and population errors could be reduced to the 103superscript10310^{-3}10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT level. To quantify the off-resonant scattering rate, we define a Raman π𝜋\piitalic_π-pulse quality factor Qπ=ΩR/πΓscsubscript𝑄𝜋subscriptΩ𝑅𝜋subscriptΓ𝑠𝑐Q_{\pi}=\Omega_{R}/\pi\Gamma_{sc}italic_Q start_POSTSUBSCRIPT italic_π end_POSTSUBSCRIPT = roman_Ω start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT / italic_π roman_Γ start_POSTSUBSCRIPT italic_s italic_c end_POSTSUBSCRIPT, where ΩRsubscriptΩ𝑅\Omega_{R}roman_Ω start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT is the two-photon Rabi frequency, and ΓscsubscriptΓ𝑠𝑐\Gamma_{sc}roman_Γ start_POSTSUBSCRIPT italic_s italic_c end_POSTSUBSCRIPT is the photon scattering rate. Because off-resonant scattering predominantly leads to leakage from the qubit subspace, ΓscsubscriptΓ𝑠𝑐\Gamma_{sc}roman_Γ start_POSTSUBSCRIPT italic_s italic_c end_POSTSUBSCRIPT can be directly inferred from population loss rates. Qπsubscript𝑄𝜋Q_{\pi}italic_Q start_POSTSUBSCRIPT italic_π end_POSTSUBSCRIPT is therefore directly measurable. In our present Raman scheme (X𝑋Xitalic_X-B𝐵Bitalic_B), the light is detuned halfway between the X2Σ(N=0)B2Σ(N=1,J=1/2)superscript𝑋2Σ𝑁0superscript𝐵2Σformulae-sequence𝑁1𝐽12X^{2}\Sigma(N=0)\rightarrow B^{2}\Sigma(N=1,J=1/2)italic_X start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Σ ( italic_N = 0 ) → italic_B start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Σ ( italic_N = 1 , italic_J = 1 / 2 ) and X2Σ(N=0)B2Σ(N=1,J=3/2)superscript𝑋2Σ𝑁0superscript𝐵2Σformulae-sequence𝑁1𝐽32X^{2}\Sigma(N=0)\rightarrow B^{2}\Sigma(N=1,J=3/2)italic_X start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Σ ( italic_N = 0 ) → italic_B start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Σ ( italic_N = 1 , italic_J = 3 / 2 ) transitions. This results in a single-photon detuning of ΔXB2π×500MHzsimilar-tosubscriptΔ𝑋𝐵2𝜋500MHz\Delta_{XB}\sim 2\pi\times 500\,\text{MHz}roman_Δ start_POSTSUBSCRIPT italic_X italic_B end_POSTSUBSCRIPT ∼ 2 italic_π × 500 MHz for optical transitions of the qubit states in 𝒯𝒯\mathcal{T}caligraphic_T. Qπsubscript𝑄𝜋Q_{\pi}italic_Q start_POSTSUBSCRIPT italic_π end_POSTSUBSCRIPT for this scheme is measured to be 37(3).

To partially investigate a better Raman scheme, we performed similar Ramsey measurements using an X𝑋Xitalic_X-A𝐴Aitalic_A Raman scheme, where the Raman light is detuned by ΔXA2π×20GHzsimilar-tosubscriptΔ𝑋𝐴2𝜋20GHz\Delta_{XA}\sim 2\pi\times 20\,\text{GHz}roman_Δ start_POSTSUBSCRIPT italic_X italic_A end_POSTSUBSCRIPT ∼ 2 italic_π × 20 GHz from the optical transitions of the hyperfine qubit. This light cannot be used for composite erasure detection because it is resonant with the detection manifold \mathcal{F}caligraphic_F. Nevertheless, we can quantify its effect on the hyperfine qubit. A Ramsey measurement reveals significantly lower population errors and coherence errors per π𝜋\piitalic_π-pulse. Specifically, εp,R=3.4(3)×103subscript𝜀𝑝𝑅3.43superscript103\varepsilon_{p,R}=3.4(3)\times 10^{-3}italic_ε start_POSTSUBSCRIPT italic_p , italic_R end_POSTSUBSCRIPT = 3.4 ( 3 ) × 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT and εc,R=10(1)×103subscript𝜀𝑐𝑅101superscript103\varepsilon_{c,R}=10(1)\times 10^{-3}italic_ε start_POSTSUBSCRIPT italic_c , italic_R end_POSTSUBSCRIPT = 10 ( 1 ) × 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT. When this scheme is employed in a composite erasure detection sequence (with tweezer ramps but no erasure image due to technical limitations), we obtain overall population and coherence errors of εp=5.2(2)×103subscript𝜀𝑝5.22superscript103\varepsilon_{p}=5.2(2)\times 10^{-3}italic_ε start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT = 5.2 ( 2 ) × 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT and εc=11(1)×103subscript𝜀𝑐111superscript103\varepsilon_{c}=11(1)\times 10^{-3}italic_ε start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = 11 ( 1 ) × 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT per image, respectively (Fig. 5(b,c)). These improvements agree with predictions due to off-resonant scattering Sup .

Our current X𝑋Xitalic_X-A𝐴Aitalic_A Raman scheme could thus be improved significantly and made compatible with composite erasure detection simply by using a different detuning. In particular, by detuning halfway between the X2Σ(v=0)A2Π1/2(v=0)superscript𝑋2Σ𝑣0superscript𝐴2subscriptΠ12𝑣0X^{2}\Sigma(v=0)\rightarrow A^{2}\Pi_{1/2}(v=0)italic_X start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Σ ( italic_v = 0 ) → italic_A start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Π start_POSTSUBSCRIPT 1 / 2 end_POSTSUBSCRIPT ( italic_v = 0 ) and the X2Σ(v=0)A2Π3/2(v=0)superscript𝑋2Σ𝑣0superscript𝐴2subscriptΠ32𝑣0X^{2}\Sigma(v=0)\rightarrow A^{2}\Pi_{3/2}(v=0)italic_X start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Σ ( italic_v = 0 ) → italic_A start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Π start_POSTSUBSCRIPT 3 / 2 end_POSTSUBSCRIPT ( italic_v = 0 ) transitions (1.1THzsimilar-toabsent1.1THz\sim 1.1\,\text{THz}∼ 1.1 THz from each), the Raman light will no longer be resonant with the detection manifold \mathcal{F}caligraphic_F. The number of off-resonantly scattered photons for a Raman π𝜋\piitalic_π-pulse would then decrease by a factor of 30 compared to our current X𝑋Xitalic_X-A𝐴Aitalic_A scheme. Conservatively, we estimate that population and coherence errors due to Raman light at an optimal detuning will be suppressed to εp,R1×104subscript𝜀𝑝𝑅1superscript104\varepsilon_{p,R}\approx 1\times 10^{-4}italic_ε start_POSTSUBSCRIPT italic_p , italic_R end_POSTSUBSCRIPT ≈ 1 × 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT and εc,R6×103subscript𝜀𝑐𝑅6superscript103\varepsilon_{c,R}\approx 6\times 10^{-3}italic_ε start_POSTSUBSCRIPT italic_c , italic_R end_POSTSUBSCRIPT ≈ 6 × 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT.

When the off-resonant Raman scattering errors are reduced to these levels, we must examine the remaining error sources. These include the detection light, the tweezer depth ramps, and loss to higher vibrational states due to blackbody radiation. We measure their individual contributions by comparing population loss from |ket|{\downarrow}\rangle| ↓ ⟩ with and without various combinations of detection light, tweezer ramps, and v=1𝑣1v=1italic_v = 1 vibrational repum** light (Fig. 5(d)). From these measurements, we isolate a population loss per image of 3.0(3)×1033.03superscript1033.0(3)\times 10^{-3}3.0 ( 3 ) × 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT due to detection light, and 2.6(3)×1032.63superscript1032.6(3)\times 10^{-3}2.6 ( 3 ) × 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT due to the tweezer ramps. The background loss in the absence of detection light and tweezer ramps is 1.2(1)×1031.21superscript1031.2(1)\times 10^{-3}1.2 ( 1 ) × 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT, consistent with the predicted excitation rate to higher vibrational states due to blackbody radiation at room temperature. Therefore, by using optimally detuned Raman light, we estimate that total population (coherence) errors of 7×1037superscript1037\times 10^{-3}7 × 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT (1.4×1021.4superscript1021.4\times 10^{-2}1.4 × 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT) per image can be achieved.

The population loss rates can be further improved. First, improving the photon collection efficiency of the imaging system allows shorter image durations and therefore lower population loss. Second, the population loss appears primarily as internal state leakage errors into the X2Σ(v=0,N=2)superscript𝑋2Σformulae-sequence𝑣0𝑁2X^{2}\Sigma(v=0,N=2)italic_X start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Σ ( italic_v = 0 , italic_N = 2 ) manifold and the |X2Σ(v=0,N=0),J=1/2,F=1,mF=±1ketformulae-sequencesuperscript𝑋2Σformulae-sequence𝑣0𝑁0𝐽12formulae-sequence𝐹1subscript𝑚𝐹plus-or-minus1\left|X^{2}\Sigma(v=0,N=0),J=1/2,F=1,m_{F}=\pm 1\right\rangle| italic_X start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Σ ( italic_v = 0 , italic_N = 0 ) , italic_J = 1 / 2 , italic_F = 1 , italic_m start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT = ± 1 ⟩ states. These errors can be erasure-converted to \mathcal{F}caligraphic_F via additional microwave and optical pulses. Thus, we estimate that the population error due to detection light can be suppressed to well below 103superscript10310^{-3}10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT. Furthermore, blackbody induced errors can be erasure-converted, as we will show, or eliminated in a cryogenic setup. Depending on the population loss mechanism of the tweezer ramps, which we did not explore, the corresponding population errors could potentially be converted and corrected. Therefore, population errors at the low 103superscript10310^{-3}10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT level could potentially be achievable.

Refer to caption
Figure 6: Erasure Conversion of Blackbody Errors. (a) Erasure Conversion Scheme. Blackbody radiation (black wavy arrow) predominantly excites molecules from 𝒯𝒯\mathcal{T}caligraphic_T to \mathcal{E}caligraphic_E. We erasure convert to \mathcal{F}caligraphic_F using v=1𝑣1v=1italic_v = 1 repum** light (red arrow). (b) Robustness of \mathcal{F}caligraphic_F to blackbody leakage errors. Blackbody leakage errors (black wavy) primarily populate X2Σ(v=0,N=3)superscript𝑋2Σformulae-sequence𝑣0𝑁3X^{2}\Sigma(v=0,N=3)italic_X start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Σ ( italic_v = 0 , italic_N = 3 ) states, which are converted back to \mathcal{F}caligraphic_F by the rotational repum** light present during erasure detection (solid green). (c) Hyperfine qubit population P0subscript𝑃0P_{0}italic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT versus hold time t𝑡titalic_t. Data without erasure excision (gray squares), with a single final error detection image (red triangles), and with excision using five mid-sequence images (blue circles) are shown; solid lines show fits to exponential decay curves. Inset: fitted 1/e1𝑒1/e1 / italic_e loss rates γpsubscript𝛾𝑝\gamma_{p}italic_γ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT are shown, with red band (±1σplus-or-minus1𝜎\pm 1\sigma± 1 italic_σ) indicating prediction from separately measured rates Sup . No dynamical decoupling Raman pulses are applied. (d) Ramsey fringe amplitude of the hyperfine qubit versus hold time with (red triangles) and without (gray squares) error excision. Dynamical decoupling with Raman pulses is applied during the hold time. Upper Inset: Amplitude loss rate γcsubscript𝛾𝑐\gamma_{c}italic_γ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT without error excision (gray square) is consistent with the expected combined loss rate from the X𝑋Xitalic_X-B𝐵Bitalic_B Raman beams and the N=0𝑁0N=0italic_N = 0 population loss (green dashed line). Lower inset shows the loss rate improvement γcsubscript𝛾𝑐\gamma_{c}italic_γ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT when excising errors. The improvement for the amplitude loss rate (green hexagon) is consistent with the improvement in population loss (purple diamond).

VII Quantum Erasure Conversion and Detection of Blackbody Errors

We next demonstrate that leakage errors induced by blackbody radiation can be converted and detected as erasures. In CaF, blackbody radiation at room-temperature predominantly drives Δv=1Δ𝑣1\Delta v=1roman_Δ italic_v = 1 vibrational-changing E1 transitions, and leads to population leakage from the vibrational ground manifold X2Σ(v=0)superscript𝑋2Σ𝑣0X^{2}\Sigma(v=0)italic_X start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Σ ( italic_v = 0 ) to the first excited vibrational manifold X2Σ(v=1)superscript𝑋2Σ𝑣1X^{2}\Sigma(v=1)italic_X start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Σ ( italic_v = 1 ) at a rate of Γ010.3s1subscriptΓ010.3superscripts1\Gamma_{01}\approx 0.3\,\text{s}^{-1}roman_Γ start_POSTSUBSCRIPT 01 end_POSTSUBSCRIPT ≈ 0.3 s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT Hou and Bernath (2018). Consequently, the hyperfine qubit states 𝒯={|,|0}X2Σ(v=0,N=0)𝒯ketket0superscript𝑋2Σformulae-sequence𝑣0𝑁0\mathcal{T}=\{|{\downarrow}\rangle,|0\rangle\}\subset X^{2}\Sigma(v=0,N=0)caligraphic_T = { | ↓ ⟩ , | 0 ⟩ } ⊂ italic_X start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Σ ( italic_v = 0 , italic_N = 0 ) predominantly leak to =X2Σ(v=1,N=1)superscript𝑋2Σformulae-sequence𝑣1𝑁1\mathcal{E}=X^{2}\Sigma(v=1,N=1)caligraphic_E = italic_X start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Σ ( italic_v = 1 , italic_N = 1 ). As we will show, these errors can be erasure converted by using v=1𝑣1v=1italic_v = 1 vibrational repum** light, which excites molecules from \mathcal{E}caligraphic_E to A2Π1/2(v=0,J=1/2,+)superscript𝐴2subscriptΠ12formulae-sequence𝑣0𝐽12A^{2}\Pi_{1/2}(v=0,J=1/2,+)italic_A start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Π start_POSTSUBSCRIPT 1 / 2 end_POSTSUBSCRIPT ( italic_v = 0 , italic_J = 1 / 2 , + ). The excited molecules decay back to \mathcal{F}caligraphic_F with 98%percent9898\%98 % probability (Fig. 6(a)).

We note that \mathcal{E}caligraphic_E spontaneously decays back to X2Σ(v=0)superscript𝑋2Σ𝑣0X^{2}\Sigma(v=0)italic_X start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Σ ( italic_v = 0 ) at a rate Γ105.3s1subscriptΓ105.3superscripts1\Gamma_{10}\approx 5.3\,\text{s}^{-1}roman_Γ start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT ≈ 5.3 s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT, which is much quicker than the blackbody excitation rate Γ01subscriptΓ01\Gamma_{01}roman_Γ start_POSTSUBSCRIPT 01 end_POSTSUBSCRIPT. If such a decay occurs, conversion via v=1𝑣1v=1italic_v = 1 vibrational repum** light is no longer possible. Therefore, in our scheme, one must erasure convert \mathcal{E}caligraphic_E to \mathcal{F}caligraphic_F at a rate much faster than Γ10subscriptΓ10\Gamma_{10}roman_Γ start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT. It is not necessary, however, to detect \mathcal{F}caligraphic_F faster than Γ10subscriptΓ10\Gamma_{10}roman_Γ start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT.

We next discuss the effect of blackbody radiation on the detection manifold \mathcal{F}caligraphic_F. Absorption of a blackbody photon brings a molecule from \mathcal{F}caligraphic_F to X2Σ(v=1,N=0,2)superscript𝑋2Σformulae-sequence𝑣1𝑁02X^{2}\Sigma(v=1,N=0,2)italic_X start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Σ ( italic_v = 1 , italic_N = 0 , 2 ). With a 1/e1𝑒1/e1 / italic_e decay time of 1/Γ10190ms1subscriptΓ10190ms1/\Gamma_{10}\approx 190\,\text{ms}1 / roman_Γ start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT ≈ 190 ms, these states spontaneously decay back to \mathcal{F}caligraphic_F and X2Σ(v=0,N=3)superscript𝑋2Σformulae-sequence𝑣0𝑁3X^{2}\Sigma(v=0,N=3)italic_X start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Σ ( italic_v = 0 , italic_N = 3 ). Molecules in the latter manifold can be pumped back to \mathcal{F}caligraphic_F via rotational repum** light, which is already present during fluorescence imaging. Therefore, even in the presence of blackbody radiation, there is minimal undetectable leakage from \mathcal{F}caligraphic_F (Fig. 6(b)).

Experimentally, we first confirm that population in \mathcal{F}caligraphic_F is robust to blackbody induced leakage errors. As expected, when a rotational repump is applied during detection, we find that the population lifetime (τ=13.5(5)s𝜏13.55s\tau=13.5(5)\,\text{s}italic_τ = 13.5 ( 5 ) s) exceeds the blackbody-limited X2Σ(v=0)superscript𝑋2Σ𝑣0X^{2}\Sigma(v=0)italic_X start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Σ ( italic_v = 0 ) lifetime by a factor of 4absent4\approx 4≈ 4. We attribute residual loss to background gas collisions and double blackbody excitation events into N5𝑁5N\geq 5italic_N ≥ 5, which are not rotationally repumped during detection Sup .

Next, we investigate detection of population leakage errors by preparing molecules in each of the two hyperfine levels in 𝒯𝒯\mathcal{T}caligraphic_T and comparing population loss rates with and without excising errors. Specifically, we hold the molecules for a variable duration, and perform mid-sequence conversion by applying v=1𝑣1v=1italic_v = 1 repum** light every 50ms50ms50\,\text{ms}50 ms. Subsequently, we perform site-resolved error detection, and destructively detect population remaining in the initially prepared hyperfine state. Without error excision, we measure a population loss rate of γp=0.30(1)s1subscript𝛾𝑝0.301superscripts1\gamma_{p}=0.30(1)\,\text{s}^{-1}italic_γ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT = 0.30 ( 1 ) s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT, consistent with the predicted blackbody excitation rate. With error excision, the loss rate decreases to γp,E=0.13(1)s1subscript𝛾𝑝𝐸0.131superscripts1\gamma_{p,E}=0.13(1)\,\text{s}^{-1}italic_γ start_POSTSUBSCRIPT italic_p , italic_E end_POSTSUBSCRIPT = 0.13 ( 1 ) s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT (Fig. 6(c)).

Using a loss rate model incorporating vacuum loss, blackbody excitation, erasure conversion efficiency (estimated to be 90%absentpercent90\approx 90\%≈ 90 %), identification errors, and the internal state preparation fidelity, we obtain loss rates consistent with the measurements in Fig. 6(c). We use this model and the measured lifetime to infer that 80%absentpercent80\approx 80\%≈ 80 % of blackbody errors are converted and detected correctly Sup . With near-term technical improvements, efficiencies approaching 95%percent9595\%95 % should be possible Sup .

Having explored conversion of blackbody population errors followed by a single error detection image, we next explore mid-sequence detection. We measure the erasure-excised loss rate with five images equally spaced during a variable hold time and measure a loss rate of γp,E5=0.12(1)s1subscript𝛾𝑝𝐸50.121superscripts1\gamma_{p,E5}=0.12(1)\,\text{s}^{-1}italic_γ start_POSTSUBSCRIPT italic_p , italic_E 5 end_POSTSUBSCRIPT = 0.12 ( 1 ) s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT, which is lower than that obtained with one final erasure image by 0.014(15)s10.01415superscripts10.014(15)\,\text{s}^{-1}0.014 ( 15 ) s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT. This shows that the interspersed detection images do not lead to additional observable loss. In fact, the data hints that because mid-sequence detection reduces population leakage from \mathcal{F}caligraphic_F to higher vibrational and rotational states (N5𝑁5N\geq 5italic_N ≥ 5), the lifetime of the detection states \mathcal{F}caligraphic_F is increased. We note that for these population loss investigations, we neither apply dynamical decoupling nor the spin-echo pulse in composite erasure detection, and are therefore free from the technical limitations of our Raman scheme (see Section VI.3).

Lastly, we demonstrate that coherence loss due to blackbody leakage errors can be reduced by mid-circuit erasure conversion. Using a Ramsey sequence, we measure the hyperfine qubit coherence with and without erasure excision. For these measurements, we erasure-convert blackbody excitations mid-circuit, and perform error detection at the end of the sequence. During the Ramsey hold time, XY8 dynamical decoupling using the X𝑋Xitalic_X-B𝐵Bitalic_B Raman configuration is applied. Without excision, the coherence loss rate is 1.64(4)s11.644superscripts11.64(4)\,\text{s}^{-1}1.64 ( 4 ) s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT, consistent with the hyperfine qubit coherence in the X𝑋Xitalic_X-B𝐵Bitalic_B Raman scheme, blackbody loss, and vacuum loss. In particular, off-resonant scattering from the Raman light contributes at least 50%absentpercent50\approx 50\%≈ 50 % of the decoherence, and can be substantially reduced (see discussion in Section VI.3). By excising data with erasures, we see a small but statistically significant (3σabsent3𝜎\approx 3\sigma≈ 3 italic_σ) decrease in the decoherence rate of 0.17(6)s10.176superscripts10.17(6)\,\text{s}^{-1}0.17 ( 6 ) s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT (Fig. 6(d)), consistent with the improvement the measured population loss (0.162(15)s1)0.16215superscripts1(0.162(15)\,\text{s}^{-1})( 0.162 ( 15 ) s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ). Taken together, our measurements show that blackbody radiation primarily causes population loss but not qubit dephasing. Although the improvement is small compared to the total coherence loss rate, which is limited by population loss due to the XY8 Raman pulses, we expect that these losses can be substantially decreased by up to a factor of 60 to 0.1s10.1superscripts10.1\,\text{s}^{-1}0.1 s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT with a more optimal Raman light source (see Section VI.3). At that point, blackbody radiation will become a non-negligible source of error that can be erasure-converted and detected.

VIII Conclusion and Outlook

In summary, we have demonstrated 1) a site-resolved detection scheme that enables robust and record-level fidelities for preparing tweezers loaded with molecules occupying a single internal state, 2) an erasure detection scheme for molecules for the first time, with error rates at the few percent level, and 3) mid-circuit erasure conversion of blackbody induced leakage errors. These results open several possibilities. First, our achieved tweezer preparation fidelity of 95%absentpercent95\approx 95\%≈ 95 % opens access to quantum many-body simulation of spin models with few percent-level defects. Second, our erasure conversion scheme adds a powerful capability that could be useful for quantum error correction. Third, our work on erasure conversion of blackbody errors shows that they do not fundamentally impose an upper limit in circuit depth in molecular tweezer arrays, since they can be corrected mid-circuit when combined with tweezer reloading and state initialization Singh et al. (2022); Shaw et al. (2023).

In addition to the possibilities above, our development of a new hyperfine qubit that is simultaneously coherent with a previously demonstrated rotational encoding opens the door to applications requiring 3-level systems, such as quantum information processing with qutrits, simulation of S=1𝑆1S=1italic_S = 1 quantum spin models, bosonic t𝑡titalic_t-J𝐽Jitalic_J models Homeier et al. (2023), and lattice gauge theory models Halimeh et al. (2023). Separately, our work on detecting blackbody excitations could be relevant for investigations of molecule-based temperature standards Norrgard et al. (2021). Lastly, our work on enhanced tweezer preparation and catching blackbody errors could also aid molecular tweezer-based precision measurement experiments Kozyryev and Hutzler (2017); Anderegg et al. (2023).

Acknowledgements.
We thank Jeff Thompson, Waseem Bakr, and their groups for fruitful discussions. This work is supported by the National Science Foundation under Grant No. 2207518. C. M. H. acknowledges support from a Joseph Taylor Graduate Student Fellowship. S. J. L. and C. L. W. acknowledge support from Princeton Quantum Initiative Graduate Student Fellowships. L. W. C. acknowledges support from the Alfred P. Sloan Foundation under Grant No. FG-2022-19104.

References

  • Kaufman and Ni (2021) A. M. Kaufman and K.-K. Ni, Nature Physics 17, 1324 (2021).
  • Endres et al. (2016) M. Endres, H. Bernien, A. Keesling, H. Levine, E. R. Anschuetz, A. Krajenbrink, C. Senko, V. Vuletić, M. Greiner,  and M. D. Lukin, Science 354, 1024 (2016).
  • Labuhn et al. (2016) H. Labuhn, D. Barredo, S. Ravets, S. de Léséleuc, T. Macrì, T. Lahaye,  and A. Browaeys, Nature 534, 667 (2016).
  • Yan et al. (2013) B. Yan, S. A. Moses, B. Gadway, J. P. Covey, K. R. Hazzard, A. M. Rey, D. S. **,  and J. Ye, Nature 501, 521 (2013).
  • Christakis et al. (2023) L. Christakis, J. S. Rosenberg, R. Raj, S. Chi, A. Morningstar, D. A. Huse, Z. Z. Yan,  and W. S. Bakr, Nature 614, 64 (2023).
  • Gregory et al. (2024) P. D. Gregory, L. M. Fernley, A. L. Tao, S. L. Bromley, J. Stepp, Z. Zhang, S. Kotochigova, K. R. A. Hazzard,  and S. L. Cornish, Nature Physics 20, 415 (2024).
  • Carr et al. (2009) L. D. Carr, D. DeMille, R. V. Krems,  and J. Ye, New Journal of Physics 11, 055049 (2009).
  • Bohn et al. (2017) J. L. Bohn, A. M. Rey,  and J. Ye, Science 357, 1002 (2017).
  • Safronova et al. (2018) M. Safronova, D. Budker, D. DeMille, D. F. J. Kimball, A. Derevianko,  and C. W. Clark, Reviews of Modern Physics 90, 025008 (2018).
  • Kozyryev and Hutzler (2017) I. Kozyryev and N. R. Hutzler, Physical review letters 119, 133002 (2017).
  • Anderegg et al. (2023) L. Anderegg, N. B. Vilas, C. Hallas, P. Robichaud, A. Jadbabaie, J. M. Doyle,  and N. R. Hutzler, Science 382, 665 (2023).
  • DeMille (2002) D. DeMille, Physical Review Letters 88, 067901 (2002).
  • Ni et al. (2018) K.-K. Ni, T. Rosenband,  and D. D. Grimes, Chem. Sci. 9, 6830 (2018).
  • Yu et al. (2019) P. Yu, L. W. Cheuk, I. Kozyryev,  and J. M. Doyle, New Journal of Physics 21, 093049 (2019).
  • Micheli et al. (2006) A. Micheli, G. K. Brennen,  and P. Zoller, Nature Physics 2, 341 (2006).
  • Gadway and Yan (2016) B. Gadway and B. Yan, Journal of Physics B: Atomic, Molecular and Optical Physics 49, 152002 (2016).
  • Anderegg et al. (2019) L. Anderegg, L. W. Cheuk, Y. Bao, S. Burchesky, W. Ketterle, K.-K. Ni,  and J. M. Doyle, Science 365, 1156 (2019).
  • Zhang et al. (2022) J. T. Zhang, L. R. B. Picard, W. B. Cairncross, K. Wang, Y. Yu, F. Fang,  and K.-K. Ni, Quantum Science and Technology 7, 035006 (2022).
  • Holland et al. (2023a) C. M. Holland, Y. Lu,  and L. W. Cheuk, Phys. Rev. Lett. 131, 053202 (2023a).
  • Ruttley et al. (2023) D. K. Ruttley, A. Guttridge, S. Spence, R. C. Bird, C. R. Le Sueur, J. M. Hutson,  and S. L. Cornish, Phys. Rev. Lett. 130, 223401 (2023).
  • Vilas et al. (2024) N. B. Vilas, P. Robichaud, C. Hallas, G. K. Li, L. Anderegg,  and J. M. Doyle, Nature 628, 282 (2024).
  • Burchesky et al. (2021) S. Burchesky, L. Anderegg, Y. Bao, S. S. Yu, E. Chae, W. Ketterle, K.-K. Ni,  and J. M. Doyle, Phys. Rev. Lett. 127, 123202 (2021).
  • Park et al. (2023) A. J. Park, L. R. B. Picard, G. E. Patenotte, J. T. Zhang, T. Rosenband,  and K.-K. Ni, Phys. Rev. Lett. 131, 183401 (2023).
  • Holland et al. (2023b) C. M. Holland, Y. Lu,  and L. W. Cheuk, Science 382, 1143 (2023b).
  • Bao et al. (2023a) Y. Bao, S. S. Yu, L. Anderegg, E. Chae, W. Ketterle, K.-K. Ni,  and J. M. Doyle, Science 382, 1138 (2023a).
  • Lu et al. (2024) Y. Lu, S. J. Li, C. M. Holland,  and L. W. Cheuk, Nature Physics , 1 (2024).
  • Bao et al. (2023b) Y. Bao, S. S. Yu, J. You, L. Anderegg, E. Chae, W. Ketterle, K.-K. Ni,  and J. M. Doyle, arXiv:2309.08706v1  (2023b).
  • Scholl et al. (2023a) P. Scholl, A. L. Shaw, R. B.-S. Tsai, R. Finkelstein, J. Choi,  and M. Endres, Nature 622, 273 (2023a).
  • Ma et al. (2023) S. Ma, G. Liu, P. Peng, B. Zhang, S. Jandura, J. Claes, A. P. Burgers, G. Pupillo, S. Puri,  and J. D. Thompson, Nature 622, 279 (2023).
  • Scholl et al. (2023b) P. Scholl, A. L. Shaw, R. Finkelstein, R. B.-S. Tsai, J. Choi,  and M. Endres, arXiv:2311.15580  (2023b).
  • Picard et al. (2024) L. R. B. Picard, G. E. Patenotte, A. J. Park, S. F. Gebretsadkan,  and K.-K. Ni, PRX Quantum 5, 020344 (2024).
  • Ruttley et al. (2024) D. K. Ruttley, A. Guttridge, T. R. Hepworth,  and S. L. Cornish, PRX Quantum 5, 020333 (2024).
  • Grassl et al. (1997) M. Grassl, T. Beth,  and T. Pellizzari, Phys. Rev. A 56, 33 (1997).
  • Bennett et al. (1997) C. H. Bennett, D. P. DiVincenzo,  and J. A. Smolin, Phys. Rev. Lett. 78, 3217 (1997).
  • Wu et al. (2022) Y. Wu, S. Kolkowitz, S. Puri,  and J. D. Thompson, Nature communications 13, 4657 (2022).
  • Kubica et al. (2023) A. Kubica, A. Haim, Y. Vaknin, H. Levine, F. Brandão,  and A. Retzker, Phys. Rev. X 13, 041022 (2023).
  • Kang et al. (2023) M. Kang, W. C. Campbell,  and K. R. Brown, PRX Quantum 4, 020358 (2023).
  • Teoh et al. (2023) J. D. Teoh, P. Winkel, H. K. Babla, B. J. Chapman, J. Claes, S. J. de Graaf, J. W. O. Garmon, W. D. Kalfus, Y. Lu, A. Maiti, K. Sahay, N. Thakur, T. Tsunoda, S. H. Xue, L. Frunzio, S. M. Girvin, S. Puri,  and R. J. Schoelkopf, Proceedings of the National Academy of Sciences 120, e2221736120 (2023).
  • Chou et al. (2023) K. S. Chou, T. Shemma, H. McCarrick, T.-C. Chien, J. D. Teoh, P. Winkel, A. Anderson, J. Chen, J. Curtis, S. J. de Graaf, et al., arXiv:2307.03169  (2023).
  • Levine et al. (2024) H. Levine, A. Haim, J. S. C. Hung, N. Alidoust, M. Kalaee, L. DeLorenzo, E. A. Wollack, P. Arrangoiz-Arriola, A. Khalajhedayati, R. Sanil, H. Moradinejad, Y. Vaknin, A. Kubica, D. Hover, S. Aghaeimeibodi, J. A. Alcid, C. Baek, J. Barnett, K. Bawdekar, P. Bienias, H. A. Carson, C. Chen, L. Chen, H. Chinkezian, E. M. Chisholm, A. Clifford, R. Cosmic, N. Crisosto, A. M. Dalzell, E. Davis, J. M. D’Ewart, S. Diez, N. D’Souza, P. T. Dumitrescu, E. Elkhouly, M. T. Fang, Y. Fang, S. Flammia, M. J. Fling, G. Garcia, M. K. Gharzai, A. V. Gorshkov, M. J. Gray, S. Grimberg, A. L. Grimsmo, C. T. Hann, Y. He, S. Heidel, S. Howell, M. Hunt, J. Iverson, I. Jarrige, L. Jiang, W. M. Jones, R. Karabalin, P. J. Karalekas, A. J. Keller, D. Lasi, M. Lee, V. Ly, G. MacCabe, N. Mahuli, G. Marcaud, M. H. Matheny, S. McArdle, G. McCabe, G. Merton, C. Miles, A. Milsted, A. Mishra, L. Moncelsi, M. Naghiloo, K. Noh, E. Oblepias, G. Ortuno, J. C. Owens, J. Pagdilao, A. Panduro, J.-P. Paquette, R. N. Patel, G. Peairs, D. J. Perello, E. C. Peterson, S. Ponte, H. Putterman, G. Refael, P. Reinhold, R. Resnick, O. A. Reyna, R. Rodriguez, J. Rose, A. H. Rubin, M. Runyan, C. A. Ryan, A. Sahmoud, T. Scaffidi, B. Shah, S. Siavoshi, P. Sivarajah, T. Skogland, C.-J. Su, L. J. Swenson, J. Sylvia, S. M. Teo, A. Tomada, G. Torlai, M. Wistrom, K. Zhang, I. Zuk, A. A. Clerk, F. G. S. L. Brandão, A. Retzker,  and O. Painter, Phys. Rev. X 14, 011051 (2024).
  • Hutzler et al. (2012) N. R. Hutzler, H.-I. Lu,  and J. M. Doyle, Chem. Rev. 112, 4803 (2012).
  • Truppe et al. (2017a) S. Truppe, H. J. Williams, N. J. Fitch, M. Hambach, T. E. Wall, E. A. Hinds, B. E. Sauer,  and M. R. Tarbutt, New Journal of Physics 19, 022001 (2017a).
  • Anderegg et al. (2017) L. Anderegg, B. L. Augenbraun, E. Chae, B. Hemmerling, N. R. Hutzler, A. Ravi, A. Collopy, J. Ye, W. Ketterle,  and J. M. Doyle, Phys. Rev. Lett. 119, 103201 (2017).
  • Truppe et al. (2017b) S. Truppe, H. J. Williams, M. Hambach, L. Caldwell, N. J. Fitch, E. A. Hinds, B. E. Sauer,  and M. R. Tarbutt, Nature Physics 13, 1173 (2017b).
  • Cheuk et al. (2018) L. W. Cheuk, L. Anderegg, B. L. Augenbraun, Y. Bao, S. Burchesky, W. Ketterle,  and J. M. Doyle, Phys. Rev. Lett. 121, 083201 (2018).
  • Li et al. (2024) S. J. Li, C. M. Holland, Y. Lu,  and L. W. Cheuk, Physical Review Letters  (2024), 2311.05447v1 .
  • Anderegg et al. (2018) L. Anderegg, B. L. Augenbraun, Y. Bao, S. Burchesky, L. W. Cheuk, W. Ketterle,  and J. M. Doyle, Nature Physics 14, 890 (2018).
  • (48) See Supplemental Material.
  • Hou and Bernath (2018) S. Hou and P. F. Bernath, Journal of Quantitative Spectroscopy and Radiative Transfer 210, 44 (2018).
  • Singh et al. (2022) K. Singh, S. Anand, A. Pocklington, J. T. Kemp,  and H. Bernien, Physical Review X 12, 011040 (2022).
  • Shaw et al. (2023) A. L. Shaw, P. Scholl, R. Finklestein, I. S. Madjarov, B. Grinkemeyer,  and M. Endres, Physical Review Letters 130, 193402 (2023).
  • Homeier et al. (2023) L. Homeier, T. J. Harris, T. Blatz, U. Schollwöck, F. Grusdt,  and A. Bohrdt, arXiv preprint arXiv:2305.02322  (2023).
  • Halimeh et al. (2023) J. C. Halimeh, L. Homeier, A. Bohrdt,  and F. Grusdt, arXiv preprint arXiv:2305.06373  (2023).
  • Norrgard et al. (2021) E. B. Norrgard, S. P. Eckel, C. L. Holloway,  and E. L. Shirley, New Journal of Physics 23, 033037 (2021).