Fast photon-mediated entanglement of continuously-cooled trapped ions
for quantum networking

Jameson O’Reilly Corresponding author: [email protected] Present Address: Department of Physics, University of Oregon, Eugene, OR 97331 Duke Quantum Center, Departments of Electrical and Computer Engineering and Physics, Duke University, Durham, NC 27708    George Toh Duke Quantum Center, Departments of Electrical and Computer Engineering and Physics, Duke University, Durham, NC 27708    Isabella Goetting Duke Quantum Center, Departments of Electrical and Computer Engineering and Physics, Duke University, Durham, NC 27708    Sagnik Saha Duke Quantum Center, Departments of Electrical and Computer Engineering and Physics, Duke University, Durham, NC 27708    Mikhail Shalaev Duke Quantum Center, Departments of Electrical and Computer Engineering and Physics, Duke University, Durham, NC 27708    Allison Carter Present Address: National Institute of Standards and Technology, Boulder CO 80305 Joint Quantum Institute, Departments of Physics and Electrical and Computer Engineering, University of Maryland, College Park, MD 20742    Andrew Risinger Present Address: Intel Corp., Hillsboro, OR 97124 Joint Quantum Institute, Departments of Physics and Electrical and Computer Engineering, University of Maryland, College Park, MD 20742    Ashish Kalakuntla Duke Quantum Center, Departments of Electrical and Computer Engineering and Physics, Duke University, Durham, NC 27708    Tingguang Li Duke Quantum Center, Departments of Electrical and Computer Engineering and Physics, Duke University, Durham, NC 27708    Ashrit Verma Duke Quantum Center, Departments of Electrical and Computer Engineering and Physics, Duke University, Durham, NC 27708    Christopher Monroe Duke Quantum Center, Departments of Electrical and Computer Engineering and Physics, Duke University, Durham, NC 27708 Joint Quantum Institute, Departments of Physics and Electrical and Computer Engineering, University of Maryland, College Park, MD 20742
(July 3, 2024)
Abstract

We entangle two co-trapped atomic barium ion qubits by collecting single visible photons from each ion through in-vacuo 0.8 NA objectives, interfering them through an integrated fiber-beamsplitter and detecting them in coincidence. This projects the qubits into an entangled Bell state with an observed fidelity lower bound of F>94%𝐹percent94F>94\%italic_F > 94 %. We also introduce an ytterbium ion for sympathetic cooling to remove the need for recooling interruptions and achieve a continuous entanglement rate of 250250250250 s-1.

Photonic interconnects between quantum processing nodes may be the only way to achieve large-scale quantum computers, and such an architecture has been proposed for the leading qubit platforms Duan and Monroe (2010); Bernien et al. (2013); Axline et al. (2018); Young et al. (2022). Using these connections to distribute remote entanglement between computing modules with high rates and near-unit fidelity should enable universal and fully-connected control over a substantially larger Hilbert space, greatly increasing the collective power of the quantum processors Gottesman and Chuang (1999); Jiang et al. (2007); Monroe et al. (2014). Interconnects between quantum memories, even without multi-qubit universal control, also offer diverse opportunities in quantum sensing Komar et al. (2014); Nichol et al. (2022), communication Nadlinger et al. (2022), and quantum simulation.

Trapped ions are attractive candidates for both quantum computing and networking due to their natural homogeneity, isolation from their environment, and indefinite idle coherence times Bruzewicz et al. (2019). These advantages, along with decades of technological development, have led to demonstrations of the highest-fidelity state preparation and measurement (SPAM) Ransford et al. (2021) and coherent operations Ballance et al. (2016); Srinivas et al. (2021); Clark et al. (2021), all performed in small systems of just one or two ions. Low errors have also been achieved in medium-sized chains Cetina et al. (2022); Chen et al. (2023), with limits due to weaker trap confinement and resulting motional mode-crowding and crosstalk concerns. Alternatively, smaller ion chains can be shuttled between interaction zones Kielpinski et al. (2002); Pino et al. (2021), but transport already dominates the time budget of current systems with up to 32 qubits Moses et al. (2023).

Photonic interconnects can avoid the overhead associated with controlling larger chains and finite shuttling speeds, but they rely on probabilistic excitation and photon emission protocols and photon collection efficiencies that have thus far been limited to a few percent. The current state-of-the-art photon-mediated entangling rate between trapped ion qubits is 182 s-1 Stephenson et al. (2020), on par with the mean entanglement rate in shuttling architectures Moses et al. (2023) but much slower than typical local entangling rates of 10-100 kHz Egan et al. (2021); Postler et al. (2023). This demonstration was mainly limited by a success probability of 2.18×1042.18superscript1042.18\times 10^{-4}2.18 × 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT in each attempt Stephenson et al. (2020) where the leading inefficiency is the use of 0.6 numerical aperture (NA) objectives that only collect 10%percent1010\%10 % of the photons from each ion. A higher success probability of 2.9×1042.9superscript1042.9\times 10^{-4}2.9 × 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT has been achieved by surrounding ions with optical cavities, but the requirement of a much lower attempt rate led to a success rate of just 0.43 s-1 Krutyanskiy et al. (2023). In these experiments, the attempt rate is limited by initialization steps, including periodic interruptions to recool the ions, as heating from photon recoil can reduce the collection efficiency and cause state measurement errors Drmota et al. (2023).

Refer to caption
Figure 1: Overview of the experiment. (a) Three co-trapped ions with two barium ions imaged by in-vacuo 0.8 NA objectives Carter et al. (2024) and an optional ytterbium ion for sympathetic cooling. Scattered light at 493 nm is coupled into single mode optical fibers and routed to a Bell state analyzer consisting of an in-fiber beamsplitter to erase which-path information and polarizing beam splitters (PBS) to measure the photon state Simon and Irvine (2003). (b) 138Ba+ level diagrams for each operation associated with ion-photon entanglement generation. Our qubit states are defined as |S1/2,mJ=1/2ketketsubscript𝑆12subscript𝑚𝐽12\mid\downarrow\rangle\equiv|S_{1/2},m_{J}=-1/2\rangle∣ ↓ ⟩ ≡ | italic_S start_POSTSUBSCRIPT 1 / 2 end_POSTSUBSCRIPT , italic_m start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT = - 1 / 2 ⟩ and |S1/2,mJ=+1/2ketketsubscript𝑆12subscript𝑚𝐽12\mid\uparrow\rangle\equiv|S_{1/2},m_{J}=+1/2\rangle∣ ↑ ⟩ ≡ | italic_S start_POSTSUBSCRIPT 1 / 2 end_POSTSUBSCRIPT , italic_m start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT = + 1 / 2 ⟩. (c) Timeline for entanglement generation attempts without sympathetic cooling. Each 1 μ𝜇\muitalic_μs-long attempt consists of optical pum**, pulsed excitation, single photon collection, and fast logic to check for a heralding detection pattern. If no such pattern occurs, we repeat attempts up to 50 times before breaking for Doppler cooling. After cooling, we repeat this cycle until success.

In this work, we utilize two 0.8 NA objectives to demonstrate photon-mediated entanglement between 138Ba+ ions with a success probability of 2.4(1)×1042.41superscript1042.4(1)\times 10^{-4}2.4 ( 1 ) × 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT and a fidelity F93.7(1.3)%𝐹93.7percent1.3F\geq 93.7(1.3)\%italic_F ≥ 93.7 ( 1.3 ) %. Then, we introduce 171Yb+ as a sympathetic coolant to achieve an uninterrupted attempt rate of 1 MHz and an ion-ion qubit entanglement rate of 250(8)2508250(8)250 ( 8 ) s-1. We choose to work with 138Ba+because it offers the longest-wavelength SP𝑆𝑃S-Pitalic_S - italic_P dipole transition of the commonly-trapped ion species at 493 nm and is similar in mass to 171Yb+, a well-established species for quantum computing Cetina et al. (2022); Moses et al. (2023); Chen et al. (2023). Photons at 493 nm can also be converted to telecom wavelengths for long-distance networking Hannegan et al. (2021).

We begin by trap** two 138Ba+ ions in a four-rod rf Paul trap and Doppler-cooling them with 493 and 650 nm light. Two 0.8 NA in-vacuo objectives collect the ion fluorescence with each lens aligned to a different ion and <105absentsuperscript105<10^{-5}< 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT crosstalk after coupling into single-mode optical fibers (see Figure 1(a)). The trap** and imaging systems are described in more detail in Ref. Carter et al. (2024). The fiber positions are optimized for maximum coupling before each measurement and are typically stable across multiple hours.

To generate entanglement between each ion and its emitted photon, we begin by optically pum** each 138Ba+ ion to |S1/2,mJ=1/2ketketsubscript𝑆12subscript𝑚𝐽12\mid\downarrow\rangle\equiv|S_{1/2},m_{J}=-1/2\rangle∣ ↓ ⟩ ≡ | italic_S start_POSTSUBSCRIPT 1 / 2 end_POSTSUBSCRIPT , italic_m start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT = - 1 / 2 ⟩ and then exciting to |P1/2,mJ=+1/2ketsubscript𝑃12subscript𝑚𝐽12|P_{1/2},m_{J}=+1/2\rangle| italic_P start_POSTSUBSCRIPT 1 / 2 end_POSTSUBSCRIPT , italic_m start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT = + 1 / 2 ⟩ with near-unit probability using a 3 ps pulse of σ+superscript𝜎\sigma^{+}italic_σ start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT 493 nm light produced by a frequency-doubled mode-locked Ti:Sapphire laser Moehring et al. (2007) (see Supplemental Material .1). When the ion returns to the S1/2subscript𝑆12S_{1/2}italic_S start_POSTSUBSCRIPT 1 / 2 end_POSTSUBSCRIPT state after spontaneous emission (lifetime 8similar-toabsent8\sim 8∼ 8 ns), it can decay to either ket\mid\downarrow\rangle∣ ↓ ⟩ or |S1/2,mJ=+1/2ketketsubscript𝑆12subscript𝑚𝐽12\mid\uparrow\rangle\equiv|S_{1/2},m_{J}=+1/2\rangle∣ ↑ ⟩ ≡ | italic_S start_POSTSUBSCRIPT 1 / 2 end_POSTSUBSCRIPT , italic_m start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT = + 1 / 2 ⟩, correlated with the photon polarization. When a 493 nm photon is collected perpendicular to the magnetic field axis and coupled into a single-mode fiber, the photon and its parent ion are projected to the state

|H+|V2,ket𝐻ketket𝑉ket2\frac{|{H}\rangle\mid\downarrow\rangle+|{V}\rangle\mid\uparrow\rangle}{\sqrt{2% }},divide start_ARG | italic_H ⟩ ∣ ↓ ⟩ + | italic_V ⟩ ∣ ↑ ⟩ end_ARG start_ARG square-root start_ARG 2 end_ARG end_ARG , (1)

where |Hket𝐻|{H}\rangle| italic_H ⟩ and |Vket𝑉|{V}\rangle| italic_V ⟩ represent orthogonal polarizations. The static phase of the above superposition is set to zero for convenience and without loss of generality.

If a single photon is detected during the 50 ns photon detection window after excitation, we proceed to state analysis and detection. Otherwise, we either repeat the attempt or break for 100 μ𝜇\muitalic_μs of Doppler cooling after 50 successive attempts, for a duty cycle of 33%percent3333\%33 %. Each attempt takes 1 μ𝜇\muitalic_μs, dominated by AOM latency and state preparation, and has independent single-photon success probabilities of ηA=2.3(1)%subscript𝜂𝐴2.3percent1\eta_{A}=2.3(1)\%italic_η start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT = 2.3 ( 1 ) % and ηB=2.2(1)%subscript𝜂𝐵2.2percent1\eta_{B}=2.2(1)\%italic_η start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT = 2.2 ( 1 ) % (see Supplemental Material .2) through each of the two ion imaging systems (hereafter labelled A𝐴Aitalic_A and B𝐵Bitalic_B). The full experimental sequence is displayed in Figure 1(c), which is not to scale.

After the photon exits the fiber, it passes through a quarter-wave plate (QWP) to compensate for any ellipticity. Then, we examine the ion-photon correlations (Figs. 2a,c) by scanning the angle of a half-wave plate (HWP) in the beam path and measuring both the photon polarization and the parent ion qubit state (see Supplemental Material .3). The resulting contrast in the correlation sets an upper bound on the fidelity overlap with Eq. 1 of FA<99.1(1)%subscript𝐹𝐴99.1percent1F_{A}<99.1(1)\%italic_F start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT < 99.1 ( 1 ) % and FB<99.1(7)%subscript𝐹𝐵99.1percent7F_{B}<99.1(7)\%italic_F start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT < 99.1 ( 7 ) %, which we attribute to residual polarization mixing in the imaging systems. We perform the measurements for each imaging system by physically blocking the other.

Refer to caption
Figure 2: Characterization of ion-photon entanglement using imaging systems A𝐴Aitalic_A and B𝐵Bitalic_B, each coupled to a unique ion. Each data point is based on 200 detection events, making the statistical error bars too small to be visible. All data was taken over the course of about 10 minutes. Red and blue data points correspond to the probability that the ion is in the bright state ket\mid\uparrow\rangle∣ ↑ ⟩ after a photon is detected in the V𝑉Vitalic_V or H𝐻Hitalic_H output mode of a polarizer, respectively. The solid lines are fits to sinusoidal functions.

To establish a lower bound for the fidelity of each ion-photon pair, we shrink the photon detection window to 3 ns and rotate each HWP by 22.5superscript22.522.5^{\circ}22.5 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT so that single photon detections herald each parent ion into the equal superposition states

j±eiϕjj2,plus-or-minussubscriptket𝑗superscript𝑒𝑖subscriptitalic-ϕ𝑗subscriptket𝑗2\frac{\mid\downarrow\rangle_{j}\pm e^{i\phi_{j}}\mid\uparrow\rangle_{j}}{\sqrt% {2}},divide start_ARG ∣ ↓ ⟩ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ± italic_e start_POSTSUPERSCRIPT italic_i italic_ϕ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_POSTSUPERSCRIPT ∣ ↑ ⟩ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG start_ARG square-root start_ARG 2 end_ARG end_ARG , (2)

where the sign depends on which detector the photon hits and the phase ϕjsubscriptitalic-ϕ𝑗\phi_{j}italic_ϕ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT (j=A,B)𝑗𝐴𝐵(j=A,B)( italic_j = italic_A , italic_B ) is given by static polarization rotations in the fiber. Then, we use a pair of 532 nm Raman beams to drive a π/2𝜋2\pi/2italic_π / 2 rotation of the atomic qubit with variable phase (Figs. 2b,d) Crocker et al. (2019). The contrast of the qubit state population with this phase sets a lower bound on the ion-photon fidelities of FA>98.1(1.4)%subscript𝐹𝐴98.1percent1.4F_{A}>98.1(1.4)\%italic_F start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT > 98.1 ( 1.4 ) % and FB>96.8(6)%subscript𝐹𝐵96.8percent6F_{B}>96.8(6)\%italic_F start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT > 96.8 ( 6 ) % Auchter et al. (2014). We also measure unmatched superposition phases of ϕA=5.00(2)subscriptitalic-ϕ𝐴5.002\phi_{A}=5.00(2)italic_ϕ start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT = 5.00 ( 2 ) rad and ϕB=0.48(2)subscriptitalic-ϕ𝐵0.482\phi_{B}=0.48(2)italic_ϕ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT = 0.48 ( 2 ) rad caused by different uncompensated birefringence along the two photon paths.

Based on the measured qubit coherence time of T2=550(27)superscriptsubscript𝑇255027T_{2}^{*}=550(27)italic_T start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT = 550 ( 27 ) μ𝜇\muitalic_μs, limited by magnetic field fluctuations, we attribute 0.26(3)%0.26percent30.26(3)\%0.26 ( 3 ) % of each infidelity to decoherence during the 40 μ𝜇\muitalic_μs before the analysis π/2𝜋2\pi/2italic_π / 2 pulse. Another 0.10(2)%0.10percent20.10(2)\%0.10 ( 2 ) % comes from averaging over different ion qubit phases at the start of the analysis pulse due to the spread of photon detection times within the detection window. We bound errors from double excitations, crosstalk between the imaging systems, and excitation laser background to the 105superscript10510^{-5}10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT level by measuring the ratio between one and two photon events.

Having established ion-photon entanglement through each imaging system, we can now entangle the two ions by sending both photons into a Bell state analyzer as shown in Figure 1, thereby performing entanglement swap**. An in-fiber 50:50 beamsplitter erases the “which-path” information, so if we detect one H𝐻Hitalic_H and one V𝑉Vitalic_V photon in the same trial the ions are ideally heralded into the entangled state

AB±ei(δt+ϕ)AB2.plus-or-minussubscriptket𝐴subscriptket𝐵superscript𝑒𝑖𝛿𝑡italic-ϕsubscriptket𝐴subscriptket𝐵2\frac{\mid\downarrow\rangle_{A}\mid\uparrow\rangle_{B}\pm e^{i(\delta t+\phi)}% \mid\uparrow\rangle_{A}\mid\downarrow\rangle_{B}}{\sqrt{2}}.divide start_ARG ∣ ↓ ⟩ start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT ∣ ↑ ⟩ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT ± italic_e start_POSTSUPERSCRIPT italic_i ( italic_δ italic_t + italic_ϕ ) end_POSTSUPERSCRIPT ∣ ↑ ⟩ start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT ∣ ↓ ⟩ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT end_ARG start_ARG square-root start_ARG 2 end_ARG end_ARG . (3)

Here, the sign is determined by whether a coincident detection occurred on the same or opposite sides of the beamsplitter Simon and Irvine (2003), δωBωA=2π×984(2)𝛿subscript𝜔𝐵subscript𝜔𝐴2𝜋9842\delta\equiv\omega_{B}-\omega_{A}=2\pi\times 984(2)italic_δ ≡ italic_ω start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT - italic_ω start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT = 2 italic_π × 984 ( 2 ) Hz is the qubit frequency difference between the ions, t𝑡titalic_t is the time elapsed after coincidence detection, and ϕϕBϕAitalic-ϕsubscriptitalic-ϕ𝐵subscriptitalic-ϕ𝐴\phi\equiv\phi_{B}-\phi_{A}italic_ϕ ≡ italic_ϕ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT - italic_ϕ start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT. This state suppresses the effect of common-mode noise Lidar et al. (1998); Monz et al. (2009) and we indeed measure an extended Bell state coherence time of T2=38(13)superscriptsubscript𝑇23813T_{2}^{*}=38(13)italic_T start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT = 38 ( 13 ) ms.

For this experiment, we measure the probability to generate one of the above maximally-entangled states to be 2.4(1)×1042.41superscript1042.4(1)\times 10^{-4}2.4 ( 1 ) × 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT, which is consistent with the product of the measured ion-photon efficiencies above: 12ηAηB=2.50(16)×10412subscript𝜂𝐴subscript𝜂𝐵2.5016superscript104\frac{1}{2}\eta_{A}\eta_{B}=2.50(16)\times 10^{-4}divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_η start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT italic_η start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT = 2.50 ( 16 ) × 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT, with the factor of 1/2 stemming from heralding only two of the four Bell states. The effective attempt rate of 333 kHz is the same as in the individual ion-photon measurements above, so the ion-ion entanglement rate is 79(3)79379(3)79 ( 3 ) s-1.

We measure both the populations and coherences of the heralded state of the ions by applying appropriate qubit rotations to both ions, as described in Supplemental Material .4. We measure the populations of the odd parity states to be P+P=97.6(5)%subscript𝑃absentsubscript𝑃absent97.6percent5P_{\downarrow\uparrow}+P_{\uparrow\downarrow}=97.6(5)\%italic_P start_POSTSUBSCRIPT ↓ ↑ end_POSTSUBSCRIPT + italic_P start_POSTSUBSCRIPT ↑ ↓ end_POSTSUBSCRIPT = 97.6 ( 5 ) % with coherences 2Re(ρ,+ρ,)=92.5(1.7)%2Resubscript𝜌absentabsentsubscript𝜌absentabsent92.5percent1.72\text{Re}\left(\rho_{\downarrow\uparrow,\uparrow\downarrow}+\rho_{\downarrow% \downarrow,\uparrow\uparrow}\right)=92.5(1.7)\%2 Re ( italic_ρ start_POSTSUBSCRIPT ↓ ↑ , ↑ ↓ end_POSTSUBSCRIPT + italic_ρ start_POSTSUBSCRIPT ↓ ↓ , ↑ ↑ end_POSTSUBSCRIPT ) = 92.5 ( 1.7 ) % Sackett et al. (2000); Slodička et al. (2013). Bounding the other possible coherence terms results in a SPAM-corrected fidelity with respect to Eq. 3 of F93.7(1.3)%𝐹93.7percent1.3F\geq 93.7(1.3)\%italic_F ≥ 93.7 ( 1.3 ) % Casabone et al. (2013). These results are based on the data displayed in Fig. 3.

Refer to caption
Figure 3: Ion-ion entanglement fidelity estimation based on a total of 9,649 entanglement attempts accrued over the course of 40,834,498 attempts and about 8 minutes of clock time. (a) Two-ion state populations. Our detection method cannot distinguish between \mid\downarrow\uparrow\rangle∣ ↓ ↑ ⟩ and \mid\uparrow\downarrow\rangle∣ ↑ ↓ ⟩ but we assume here that they contribute equally to the measured one-bright population. (b) Parity scan for bounding the off-diagonal elements of the ion-ion state. The red data correspond to scanning the phase of a single π/2𝜋2\pi/2italic_π / 2 pulse. The corresponding solid line is the fit and the dashed line represents the expected behavior of the ideal state Ψ+ketsuperscriptΨ\mid\Psi^{+}\rangle∣ roman_Ψ start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT ⟩. The blue data correspond to scanning the phase of a second π/2𝜋2\pi/2italic_π / 2 pulse after a π/2𝜋2\pi/2italic_π / 2 pulse with fixed phase ϕ=0italic-ϕ0\phi=0italic_ϕ = 0.

Based on the measured finite contrast of the spin-polarization correlations, we expect an infidelity of 2.9(1.6)%2.9percent1.62.9(1.6)\%2.9 ( 1.6 ) %, which is consistent with the measured populations. The extended two-qubit coherence is expected to contribute 0.3(1)%0.3percent10.3(1)\%0.3 ( 1 ) % and other sources including temporal mismatch and dark counts account for <0.3(1)%absent0.3percent1<0.3(1)\%< 0.3 ( 1 ) %. The total predicted infidelity of <3.5(1.6)%absent3.5percent1.6<3.5(1.6)\%< 3.5 ( 1.6 ) % (see Table 1) is thus within error of our measured infidelity. Notably, using an in-fiber beamsplitter avoids the percent-level error induced by imperfect free-space photon spatial mode overlap in prior experiments Hucul et al. (2015); Stephenson et al. (2020).

Error source Infidelity (×102absentsuperscript102\times 10^{-2}× 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT)
Polarization errors 2.9(1.6)2.91.62.9(1.6)2.9 ( 1.6 )
Ion decoherence (T2=38(13)superscriptsubscript𝑇23813T_{2}^{*}=38(13)italic_T start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT = 38 ( 13 ) ms) 0.3(1)0.310.3(1)0.3 ( 1 )
Timing mismatch (Δt=21(2)Δ𝑡212\Delta t=21(2)roman_Δ italic_t = 21 ( 2 ) ps) 0.13(1)0.1310.13(1)0.13 ( 1 )
Imperfect splitting ratio (r=0.515(1)𝑟0.5151r=0.515(1)italic_r = 0.515 ( 1 )) 0.09(1)0.0910.09(1)0.09 ( 1 )
Dark counts and double excitations <0.1absent0.1<0.1< 0.1
Total <3.5(1.6)absent3.51.6<3.5(1.6)< 3.5 ( 1.6 )
Table 1: Infidelity budget for the entangled ion-ion state. The total expected error is consistent with our measured fidelity lower bound of F>93.7(1.3)%𝐹93.7percent1.3F>93.7(1.3)\%italic_F > 93.7 ( 1.3 ) %.
Refer to caption
Figure 4: (a) New experimental sequence with an ytterbium sympathetic coolant present in the Coulomb crystal. For each requested event we initialize by Doppler cooling the barium ion for 100 μ𝜇\muitalic_μs and then execute attempts until we either succeed or reach N𝑁Nitalic_N failures. Meanwhile, we continuously Doppler cool the ytterbium ion. (b) The points (data) and solid lines (fits) represent average entanglement rates for different maximum consecutive attempts with (blue) and without (red) the ytterbium sympathetic coolant. Red data is calculated from a histogram of attempts before success in 13,739 loops and each blue point is a direct measurement based on an average of 3,741,332 attempts. For a more direct comparison, we ignore barium recooling time in the no cooling case. The final, black point is based on a fit to the blue data. Dashed lines are the cumulative success probability of a heralding event through attempt N𝑁Nitalic_N for the fitted probability distributions of the Yb-cooling and no-cooling runs.

Over the course of many consecutive entanglement generation attempts, recoil from optical pum** and pulsed excitation heats the ions, reducing the heralded success probability, as shown in Figure 4(b). Above, we capped the number of attempts without cooling at 50 to avoid this decay and to maintain high-fidelity state detection. We avoid these issues while maximizing the entanglement attempt rate by co-trap** a 171Yb+ ion for continuous sympathetic cooling.

We Doppler cool the 171Yb+ ion using 370 and 935 nm light, with sufficient spectral isolation as not to degrade the 138Ba+ state detection, cooling, or coherent operations. The relatively similar masses of barium and ytterbium and the small ratio of radial to axial confinement in the trap enable significant coupling between the radial modes of the different species Sosnova et al. (2021), which in turn allows for efficient sympathetic cooling Sakrejda et al. (2021).

With continuous sympathetic cooling, we are able to perform entanglement attempts without stop** for recooling, recovering our full attempt rate of 1 MHz. Although our hardware counter resets at 214=16384superscript214163842^{14}=163842 start_POSTSUPERSCRIPT 14 end_POSTSUPERSCRIPT = 16384 attempts, we estimate an average success probability of 2.50(8)×1042.508superscript1042.50(8)\times 10^{-4}2.50 ( 8 ) × 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT when allowing a maximum of N=20,000𝑁20000N=20,000italic_N = 20 , 000 attempts, corresponding to an entanglement rate of 250(8)2508250(8)250 ( 8 ) s-1. After 20,000 attempts, the cumulative success probability of heralding an entanglement event is >99%absentpercent99>99\%> 99 % (see Supplemental Material .6).

This rate, made possible by increased numerical aperture and the introduction of sympathetic cooling, surpasses any previous mark in a system with Bell state fidelities above 70%percent7070\%70 % Stockill et al. (2017). It could be improved by almost a factor of three by replacing AOMs with electro-optic control to reduce latency and by another factor of two by switching to an atomic species with a larger branching ratio to the ground state Hucul et al. (2015); Stephenson et al. (2020). Building a duplicate of this system and using both imaging systems of each chamber to collect light from a single ion would again double the success probability reported here, providing a road map to kHz-level remote entanglement rates between atomic memories. Further increases could be achieved using Purcell enhancement in short optical cavities or large-scale spatial multiplexing with integrated optics Ghadimi et al. (2017).

Achieving a better understanding of polarization map** errors will require a more careful study of imaging aberrations and inhomogeneous birefringence in the lenses and vacuum window. This will be enabled by performing full tomography of the entangled ion-photon and ion-ion states. The dominance of imperfect polarization encoding in our error budget suggests that alternative photonic-qubit encodings, such as frequency Maunz et al. (2009) and time-bin Ward and Keller (2022); Tchebotareva et al. (2019), may be beneficial for short and medium-distance networking in addition to their usual application across longer distances Bersin et al. (2024). The former could be available using the 137Ba+ or 133Ba+ isotopes while the latter benefits from the long D𝐷Ditalic_D state lifetimes in any barium isotope.

The continued maturation of photonic interconnects will enable a wide variety of quantum technologies including scalable ion-trap quantum computers Monroe et al. (2014), quantum-limited sensing networks Komar et al. (2014); Nichol et al. (2022), secure communication Nadlinger et al. (2022), and blind quantum computation Fitzsimons (2017). Many of these applications will require the integration of remote and local entangling operations, both of which benefit from the introduction of sympathetic cooling. Dual-species or omg Allcock et al. (2021) operation is already necessary in most trapped-ion computing and networking architectures and has been demonstrated in numerous experiments Cetina et al. (2022); Drmota et al. (2023); Moses et al. (2023), but the integration of these techniques into a multi-node network, with demonstrations of quantum repeaters and entanglement distillation, remains outstanding.

This work is supported by the DOE Quantum Systems Accelerator (DE-FOA-0002253) and the NSF STAQ Program (PHY-1818914). J.O. is supported by the National Science Foundation Graduate Research Fellowship (DGE 2139754).

References

Supplemental Material

.1 Pulsed excitation

To produce pulsed 493 nm light, we use a mode-locked Coherent Mira 900P Ti:Sapphire laser at 986 nm and subsequently frequency-double the light using second harmonic generation (SHG) to make 493 nm. The laser generates 3 ps pulses with a repetition rate of 76 MHz. The train of 986 nm pulses enters an electro-optic pulse picker, which transmits single, on-demand pulses with an extinction ratio of about 500:1. After frequency doubling with a MgO-doped, periodically-poled lithium niobate crystal, this extinction ratio increases to ¿100,000:1. Finally, we send the pulses through an AOM for further extinction and power control before routing them to the vacuum chamber via polarization-maintaining optical fiber.

.2 Photon collection efficiencies

In each attempt, we pump to ket\mid\downarrow\rangle∣ ↓ ⟩ with 96(2)%96percent296(2)\%96 ( 2 ) % fidelity and excite an average of 96(2)%96percent296(2)\%96 ( 2 ) % of the population to |P1/2,mJ=+1/2ketsubscript𝑃12subscript𝑚𝐽12|P_{1/2},m_{J}=+1/2\rangle| italic_P start_POSTSUBSCRIPT 1 / 2 end_POSTSUBSCRIPT , italic_m start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT = + 1 / 2 ⟩. Based on the branching ratio back to S1/2subscript𝑆12S_{1/2}italic_S start_POSTSUBSCRIPT 1 / 2 end_POSTSUBSCRIPT, a 493 nm photon is emitted in 73.2%percent73.273.2\%73.2 % De Munshi et al. (2015) of decay events. The photons are collected by a 0.8 NA objective that covers 20%percent2020\%20 % of the emission solid angle, but within that area only 97(1)%97percent197(1)\%97 ( 1 ) % of the photons make it past the trap rods and the lens has a transmission of 91(3)%91percent391(3)\%91 ( 3 ) % Carter et al. (2024). We measure a fiber coupling efficiency of 30(3)%30percent330(3)\%30 ( 3 ) % Carter et al. (2024) and detect photons with avalanche photo-detectors that have specified quantum efficiencies of 71%percent7171\%71 %. In total, from either imaging system, we expect a single photon detection in 2.5(3)%2.5percent32.5(3)\%2.5 ( 3 ) % of trials, which is consistent with our measured values of 2.3(1)2.312.3(1)2.3 ( 1 ) and 2.2(1)%2.2percent12.2(1)\%2.2 ( 1 ) %. We believe that the measured values are a bit lower due to ion recoil heating and additional photonic losses from polarizers, waveplates, and optical filters.

.3 Two-ion state detection

We detect the state of the qubit(s) at the end of an experiment by shelving the ket\mid\downarrow\rangle∣ ↓ ⟩ population in the |D5/2,mJ=1/2ketsubscript𝐷52subscript𝑚𝐽12|D_{5/2},m_{J}=-1/2\rangle| italic_D start_POSTSUBSCRIPT 5 / 2 end_POSTSUBSCRIPT , italic_m start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT = - 1 / 2 ⟩ state. We shelve using 1762 nm light produced by a thulium-doped fiber laser and fiber amplifier. This system produces 450 mW of 1762 nm light and is stabilized to <200absent200<200< 200 Hz by locking to a high-finesse optical cavity with an ultra-low expansion (ULE) glass spacer. After shelving, we apply all polarizations of 493 and 650 nm light, which causes unshelved ions in ket\mid\uparrow\rangle∣ ↑ ⟩ to fluoresce while shelved ions remain dark in the metastable D5/2subscript𝐷52D_{5/2}italic_D start_POSTSUBSCRIPT 5 / 2 end_POSTSUBSCRIPT manifold, see Figure 5(a-b). We collect and detect these fluorescence photons using the imaging system shown in Figure 1 in the main text Dietrich et al. (2009); Keselman et al. (2011). For the one-ion case, as in the ion-photon measurements, we achieve an average ion qubit detection fidelity of 99.5%similar-toabsentpercent99.5\sim 99.5\%∼ 99.5 %.

Refer to caption
Figure 5: Two-ion state detection. (a) We shelve the ket\mid\downarrow\rangle∣ ↓ ⟩ state of each ion to the metastable D5/2subscript𝐷52D_{5/2}italic_D start_POSTSUBSCRIPT 5 / 2 end_POSTSUBSCRIPT manifold. (b) When we subsequently apply 493 and 650 nm light, the shelved population remains dark while any ket\mid\uparrow\rangle∣ ↑ ⟩ population fluoresces. (c) We collect 493 nm scattered photons from the ions for 1 ms using the imaging system shown in Figure 1 in the main text and set threshold values that separate the photon count distributions for zero, one, or two bright ions with high fidelity. In this example, both bright corresponds to pum** both ions to ket\mid\uparrow\rangle∣ ↑ ⟩, both dark corresponds to pum** both to ket\mid\downarrow\rangle∣ ↓ ⟩ and shelving, and mixture is the result of applying a partial shelving pulse so that we get significant one-bright population. We repeated each experiment 20,000 times.

Collecting fluorescence for 1 ms provides well-resolved photon number histograms for the cases of no bright ions, one bright ion, and two bright ions, as shown in Figure 5(c). Imperfect shelving reduces the no-bright detection fidelity to 98.7(4)%98.7percent498.7(4)\%98.7 ( 4 ) % with erroneous events predominantly registering as one-bright events. The decay of fiber coupling due to heating during experiments reduces the two-bright average fidelity to 98.1(4)%98.1percent498.1(4)\%98.1 ( 4 ) %. We correct for these and their corresponding single-ion errors by applying the inverse transformations to the data Shen and Duan (2012); Seif et al. (2018).

.4 Ion-ion entanglement fidelity bound

The global nature of our Raman addressing system limits us to analyzing the state fidelity relative to the state

|Ψ++2ketsuperscriptΨketketketket2|{\Psi^{+}}\rangle\equiv\frac{\mid\downarrow\rangle\mid\uparrow\rangle+\mid% \uparrow\rangle\mid\downarrow\rangle}{\sqrt{2}}| roman_Ψ start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT ⟩ ≡ divide start_ARG ∣ ↓ ⟩ ∣ ↑ ⟩ + ∣ ↑ ⟩ ∣ ↓ ⟩ end_ARG start_ARG square-root start_ARG 2 end_ARG end_ARG (4)

because the singlet state |ΨketsuperscriptΨ|\Psi^{-}\rangle| roman_Ψ start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ⟩ is invariant under global rotations. We begin this process by measuring the state populations ρ+ρ=97.6(5)%subscript𝜌absentsubscript𝜌absent97.6percent5\rho_{\downarrow\uparrow}+\rho_{\uparrow\downarrow}=97.6(5)\%italic_ρ start_POSTSUBSCRIPT ↓ ↑ end_POSTSUBSCRIPT + italic_ρ start_POSTSUBSCRIPT ↑ ↓ end_POSTSUBSCRIPT = 97.6 ( 5 ) %. After waiting 210 μ𝜇\muitalic_μs for δt=ϕ𝛿𝑡italic-ϕ\delta t=-\phiitalic_δ italic_t = - italic_ϕ, we apply a global π2𝜋2\frac{\pi}{2}divide start_ARG italic_π end_ARG start_ARG 2 end_ARG rotation that converts |Ψ+ketsuperscriptΨ|\Psi^{+}\rangle| roman_Ψ start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT ⟩ into |Φ+=12(+)|\Phi^{+}\rangle=\frac{1}{\sqrt{2}}\left(\mid\downarrow\downarrow\rangle+\mid% \uparrow\uparrow\rangle\right)| roman_Φ start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT ⟩ = divide start_ARG 1 end_ARG start_ARG square-root start_ARG 2 end_ARG end_ARG ( ∣ ↓ ↓ ⟩ + ∣ ↑ ↑ ⟩ ) followed by a second π2𝜋2\frac{\pi}{2}divide start_ARG italic_π end_ARG start_ARG 2 end_ARG pulse with varying phase ϕsuperscriptitalic-ϕ\phi^{\prime}italic_ϕ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT. The maximum of the parity Pρ~+ρ~ρ~ρ~𝑃subscript~𝜌absentsubscript~𝜌absentsubscript~𝜌absentsubscript~𝜌absentP\equiv\tilde{\rho}_{\downarrow\downarrow}+\tilde{\rho}_{\uparrow\uparrow}-% \tilde{\rho}_{\downarrow\uparrow}-\tilde{\rho}_{\uparrow\downarrow}italic_P ≡ over~ start_ARG italic_ρ end_ARG start_POSTSUBSCRIPT ↓ ↓ end_POSTSUBSCRIPT + over~ start_ARG italic_ρ end_ARG start_POSTSUBSCRIPT ↑ ↑ end_POSTSUBSCRIPT - over~ start_ARG italic_ρ end_ARG start_POSTSUBSCRIPT ↓ ↑ end_POSTSUBSCRIPT - over~ start_ARG italic_ρ end_ARG start_POSTSUBSCRIPT ↑ ↓ end_POSTSUBSCRIPT of the rotated state in this scan, shown in Figure 3 in the main text, corresponds to 2Re(ρ,+ρ,)=92.5(1.7)%2Resubscript𝜌absentabsentsubscript𝜌absentabsent92.5percent1.72\text{Re}\left(\rho_{\downarrow\uparrow,\uparrow\downarrow}+\rho_{\downarrow% \downarrow,\uparrow\uparrow}\right)=92.5(1.7)\%2 Re ( italic_ρ start_POSTSUBSCRIPT ↓ ↑ , ↑ ↓ end_POSTSUBSCRIPT + italic_ρ start_POSTSUBSCRIPT ↓ ↓ , ↑ ↑ end_POSTSUBSCRIPT ) = 92.5 ( 1.7 ) % Sackett et al. (2000). We measure 2Re(ρ,)=2.7(1.8)%2Resubscript𝜌absentabsent2.7percent1.82\text{Re}(\rho_{\downarrow\downarrow,\uparrow\uparrow})=2.7(1.8)\%2 Re ( italic_ρ start_POSTSUBSCRIPT ↓ ↓ , ↑ ↑ end_POSTSUBSCRIPT ) = 2.7 ( 1.8 ) % by scanning the phase of a single π2𝜋2\frac{\pi}{2}divide start_ARG italic_π end_ARG start_ARG 2 end_ARG pulse Slodička et al. (2013), which allows us to calculate an ion-ion fidelity lower bound of F93.7(1.3)%𝐹93.7percent1.3F\geq 93.7(1.3)\%italic_F ≥ 93.7 ( 1.3 ) % Casabone et al. (2013).

.5 Yb-Ba-Ba collective motional modes

Coulomb forces between ions co-trapped in a harmonic potential U𝑈Uitalic_U lead to collective motional modes that are often used as an information bus for local entangling gates in trapped ion systems James (1998); Sørensen and Mølmer (2000); Schmidt-Kaler et al. (2003). In our application, the collective nature allows us to cool the barium ions via their coupling to the ytterbium ion’s motion. Using our measured secular frequencies for a single barium ion Carter et al. (2024), we can find the structure of the modes in our Yb-Ba-Ba chain by solving

i,j=1NUqiqj|0bim=ωm2mibimevaluated-atsuperscriptsubscript𝑖𝑗1𝑁𝑈subscript𝑞𝑖subscript𝑞𝑗0subscript𝑏𝑖𝑚superscriptsubscript𝜔𝑚2subscript𝑚𝑖subscript𝑏𝑖𝑚\sum_{i,j=1}^{N}\frac{\partial U}{\partial q_{i}\partial q_{j}}\bigg{|}_{0}b_{% im}=\omega_{m}^{2}m_{i}b_{im}∑ start_POSTSUBSCRIPT italic_i , italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT divide start_ARG ∂ italic_U end_ARG start_ARG ∂ italic_q start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ∂ italic_q start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG | start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_b start_POSTSUBSCRIPT italic_i italic_m end_POSTSUBSCRIPT = italic_ω start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_m start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_b start_POSTSUBSCRIPT italic_i italic_m end_POSTSUBSCRIPT (5)

where qisubscript𝑞𝑖q_{i}italic_q start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT is the position of ion i𝑖iitalic_i, ωmsubscript𝜔𝑚\omega_{m}italic_ω start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT is the secular frequency of mode m𝑚mitalic_m, and bimsubscript𝑏𝑖𝑚b_{im}italic_b start_POSTSUBSCRIPT italic_i italic_m end_POSTSUBSCRIPT is the participation eigenvector of ion i𝑖iitalic_i in mode m𝑚mitalic_m with ibimbin=δnmsubscript𝑖subscript𝑏𝑖𝑚subscript𝑏𝑖𝑛subscript𝛿𝑛𝑚\sum_{i}b_{im}b_{in}=\delta_{nm}∑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_b start_POSTSUBSCRIPT italic_i italic_m end_POSTSUBSCRIPT italic_b start_POSTSUBSCRIPT italic_i italic_n end_POSTSUBSCRIPT = italic_δ start_POSTSUBSCRIPT italic_n italic_m end_POSTSUBSCRIPT and mbimbjm=δijsubscript𝑚subscript𝑏𝑖𝑚subscript𝑏𝑗𝑚subscript𝛿𝑖𝑗\sum_{m}b_{im}b_{jm}=\delta_{ij}∑ start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT italic_b start_POSTSUBSCRIPT italic_i italic_m end_POSTSUBSCRIPT italic_b start_POSTSUBSCRIPT italic_j italic_m end_POSTSUBSCRIPT = italic_δ start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT .

171Yb+ 138Ba+ 138Ba+ ωm/2πsubscript𝜔𝑚2𝜋\omega_{m}/2\piitalic_ω start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT / 2 italic_π (kHz)
Axial Mode 1 0.614 0.640 0.300 353
Axial Mode 2 0.567 -0.126 -0.840 604
Axial Mode 3 0.549 -0.758 0.453 872
Radial Mode 1 0.178 0.412 0.847 868
Radial Mode 2 0.587 0.672 -0.512 737
Radial Mode 3 0.790 -0.615 0.144 606
Table 2: Collective secular motional mode participation eigenvector matrix bimsubscript𝑏𝑖𝑚b_{im}italic_b start_POSTSUBSCRIPT italic_i italic_m end_POSTSUBSCRIPT of the Yb-Ba-Ba Coulomb crystal that we trap for sympathetic cooling experiments. Ions with different charge-to-mass ratios typically have good mutual participation in axial modes, and we also maintain strong coupling in the radial modes thanks to our relatively weak radial confinement Sosnova et al. (2021).

The excitation and pum** beams are delivered at 45superscript4545^{\circ}45 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT relative to the trap axis and emission is isotropic, so we need to sympathetically cool both the radial and axial directions. Multi-species ion traps often suffer from weak radial coupling between the species Home (2013); Sosnova et al. (2021), but this is circumvented by using a high ratio of axial to radial confinement (see Table 2).

.6 Deriving p¯(N)¯𝑝𝑁\bar{p}(N)over¯ start_ARG italic_p end_ARG ( italic_N )

In a system where the probability of success on the nthsuperscript𝑛thn^{\text{th}}italic_n start_POSTSUPERSCRIPT th end_POSTSUPERSCRIPT trial p(n)p𝑝𝑛𝑝p(n)\equiv pitalic_p ( italic_n ) ≡ italic_p is constant Drmota et al. (2023), we expect an exponential distribution of required trials n𝑛nitalic_n before success each time we attempt to generate entanglement: PDF(n)=penp𝑛𝑝superscript𝑒𝑛𝑝(n)=pe^{-np}( italic_n ) = italic_p italic_e start_POSTSUPERSCRIPT - italic_n italic_p end_POSTSUPERSCRIPT. Instead, we observe a success probability that decays to a steady-state value p(n)=AeBn+C𝑝𝑛𝐴superscript𝑒𝐵𝑛𝐶p(n)=Ae^{-Bn}+Citalic_p ( italic_n ) = italic_A italic_e start_POSTSUPERSCRIPT - italic_B italic_n end_POSTSUPERSCRIPT + italic_C (Fig. 4(b) in the main text), stemming from the increased Doppler temperature of the high-intensity optical pum** beam. In this case, we extend the trial number n𝑛nitalic_n as a continuous variable and use

PDF(n)=p(n)(10nPDF(n)dn)PDF𝑛𝑝𝑛1superscriptsubscript0𝑛PDFsuperscript𝑛dsuperscript𝑛\text{PDF}(n)=p(n)\left(1-\int_{0}^{n}\text{PDF}(n^{\prime})\text{d}n^{\prime}\right)PDF ( italic_n ) = italic_p ( italic_n ) ( 1 - ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT PDF ( italic_n start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) d italic_n start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) (6)

to find

PDF(n)=exp[AB(eBn1)Cn](AeBn+C).PDF𝑛𝐴𝐵superscript𝑒𝐵𝑛1𝐶𝑛𝐴superscript𝑒𝐵𝑛𝐶\text{PDF}(n)=\exp\left[\frac{A}{B}(e^{-Bn}-1)-Cn\right](Ae^{-Bn}+C).PDF ( italic_n ) = roman_exp [ divide start_ARG italic_A end_ARG start_ARG italic_B end_ARG ( italic_e start_POSTSUPERSCRIPT - italic_B italic_n end_POSTSUPERSCRIPT - 1 ) - italic_C italic_n ] ( italic_A italic_e start_POSTSUPERSCRIPT - italic_B italic_n end_POSTSUPERSCRIPT + italic_C ) . (7)

Integrating this from 0 to N𝑁Nitalic_N, we find the cumulative density function, or the probability of success up through N𝑁Nitalic_N trials,

CDF(N)=1exp[AB(eBN1)CN].CDF𝑁1𝐴𝐵superscript𝑒𝐵𝑁1𝐶𝑁\text{CDF}(N)=1-\exp\left[\frac{A}{B}(e^{-BN}-1)-CN\right].CDF ( italic_N ) = 1 - roman_exp [ divide start_ARG italic_A end_ARG start_ARG italic_B end_ARG ( italic_e start_POSTSUPERSCRIPT - italic_B italic_N end_POSTSUPERSCRIPT - 1 ) - italic_C italic_N ] . (8)

Finally, we arrive at the average success probability by dividing the probability of success up through N𝑁Nitalic_N trials by the total attempts that have been executed up to the Nthsuperscript𝑁thN^{\text{th}}italic_N start_POSTSUPERSCRIPT th end_POSTSUPERSCRIPT in each loop, resulting in

p¯(N)=CDF(N)N+10NCDF(n)dn.¯𝑝𝑁CDF𝑁𝑁1superscriptsubscript0𝑁CDF𝑛d𝑛\bar{p}(N)=\frac{\text{CDF}(N)}{N+1-\int_{0}^{N}\text{CDF}(n)\text{d}n}.over¯ start_ARG italic_p end_ARG ( italic_N ) = divide start_ARG CDF ( italic_N ) end_ARG start_ARG italic_N + 1 - ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT CDF ( italic_n ) d italic_n end_ARG . (9)

While we could not find an analytic solution for the integral, we were able to fit the blue data points in Figure 4(c) in the main text to this equation by integrating numerically.