8
Benchmarking an 11-qubit quantum computer K. Wright, 1, * K. M. Beck, 1 S. Debnath, 1 J. M. Amini, 1 Y. Nam, 1 N. Grzesiak, 1 J.-S. Chen, 1 N. C. Pisenti, 1 M. Chmielewski, 1, 2 C. Collins, 1 K. M. Hudek, 1 J. Mizrahi, 1 J. D. Wong-Campos, 1 S. Allen, 1 J. Apisdorf, 1 P. Solomon, 1 M. Williams, 1 A. M. Ducore, 1 A. Blinov, 1 S. M. Kreikemeier, 1 V. Chaplin, 1 M. Keesan, 1 C. Monroe, 1, 2 and J. Kim 1, 3 1 IonQ, Inc., College Park, MD 20740, USA 2 Joint Quantum Institute and Department of Physics, University of Maryland, College Park, MD 20742, USA 3 Department of Electrical and Computer Engineering, Duke University, Durham, NC 27708, USA The field of quantum computing has grown from concept to demonstration devices over the past 20 years. Universal quantum computing offers efficiency in approaching problems of scientific and com- mercial interest, such as factoring large numbers,[1] searching databases,[2] simulating intractable models from quantum physics,[3] and optimizing complex cost functions.[4] Here, we present an 11- qubit fully-connected, programmable quantum computer in a trapped ion system composed of 13 171 Yb + ions. We demonstrate average single-qubit gate fidelities of 99.5%, average two-qubit-gate fidelities of 97.5%, and state preparation and measurement errors of 0.7%. To illustrate the capa- bilities of this universal platform and provide a basis for comparison with similarly-sized devices, we compile the Bernstein-Vazirani (BV)[5] and Hidden Shift (HS)[6, 7] algorithms into our native gates and execute them on the hardware with average success rates of 78% and 35%, respectively. These algorithms serve as excellent benchmarks for any type of quantum hardware, and show that our system outperforms all other currently available hardware. Small universal quantum computers that can execute textbook quantum circuits exist in both academic[8– 12] and industrial[13–17] settings. With a range of two to seventy-two qubits and sufficient fidelity for only tens of entangling gates, these devices and the under- lying qubit implementations can be difficult to com- pare. Even within the trapped ion platform, there is large diversity in atomic species, system architec- tures, and gate implementations. Trapped ion systems with one to two qubits have shown single-qubit gate fidelities of 99.9999%[18] with microwave-based opera- tions and better than 99.99% fidelity with laser based operations,[19, 20] state preparation and measurement (SPAM) error below 10 -4 ,[18, 21] and two-qubit gates with fidelities exceeding 99.9%.[19, 20] Algorithms have been executed on up to seven trapped-ion qubits[22] and, while not optimized for universal quantum comput- ing, quantum simulators with more than 50 ions have modeled fundamental quantum systems including Ising chains[23] and quantum magnetism.[24] Benchmarking across implementations needs to be both universal across platforms and agnostic to the dif- ferences in the underlying hardware. In traditional com- puting, the performance of computers is measured by ex- ecuting a set of benchmark problems representing var- ious use-case scenarios, to provide users with an esti- mate of how the computers would perform in their spe- cific applications. Canonical quantum algorithms demon- strate unambiguous advantage of quantum computers over classical computation, and provide verifiable out- comes to assess successful execution of the algorithm. Therefore, they can serve as ideal candidate problems * [email protected] ION CHAIN DETECTION OPTICS GLOBAL ADDRESSING BEAM DIFFRACTIVE OPTICAL ELEMENT MULTI-CHANNEL AOM INDIVIDUAL ADDRESSING BEAMS FIG. 1. Schematic of the hardware. A linear chain of ions is trapped near a surface electrode trap (trap is not shown). Lasers at 369 nm and 935 nm (not shown) illuminate all of the ions during cooling, initialization, and detection. Each ion’s fluorescence is imaged through a 0.6 numeric aperture lens (“detection optics”) and directed onto individual photomul- tiplier tube channels. Two linearly-polarized counterpropa- gating 355 nm Raman beams are aligned to each qubit-ion, a globally addressing beam that couples to all of the qubits (red) and an individual addressing beam that is focused onto each ion (blue). Acousto-optic modulators (AOMs) modulate the frequency and amplitude of each of these beams to gen- erate single-qubit rotations and XX-gates between arbitrary pairs of qubit ions. for benchmarking the performance of any quantum com- puters. These benchmark algorithms exercise the full hardware/software stack. A hardware-specific compiler breaks down algorithms into the target hardware’s native gate set, optimizing for qubit connectivity, gate times, and coherence[25] to enhance the system’s performance. After execution on the hardware, the measurements can be directly compared with the expected output state to determine the accuracy of the device. This accuracy can then be compared with other devices that have compiled arXiv:1903.08181v1 [quant-ph] 19 Mar 2019

arXiv:1903.08181v1 [quant-ph] 19 Mar 2019 · 3 FIG. 3. Bernstein-Vazirani (BV) algorithm. (a) shows a textbook implementation of the BV algorithm with hidden bit string 1010101010

Embed Size (px)

Citation preview

Benchmarking an 11-qubit quantum computer

K. Wright,1, ∗ K. M. Beck,1 S. Debnath,1 J. M. Amini,1 Y. Nam,1 N. Grzesiak,1 J.-S. Chen,1 N. C. Pisenti,1 M.

Chmielewski,1, 2 C. Collins,1 K. M. Hudek,1 J. Mizrahi,1 J. D. Wong-Campos,1 S. Allen,1 J. Apisdorf,1 P. Solomon,1

M. Williams,1 A. M. Ducore,1 A. Blinov,1 S. M. Kreikemeier,1 V. Chaplin,1 M. Keesan,1 C. Monroe,1, 2 and J. Kim1, 3

1IonQ, Inc., College Park, MD 20740, USA2Joint Quantum Institute and Department of Physics,University of Maryland, College Park, MD 20742, USA

3Department of Electrical and Computer Engineering, Duke University, Durham, NC 27708, USA

The field of quantum computing has grown from concept to demonstration devices over the past 20years. Universal quantum computing offers efficiency in approaching problems of scientific and com-mercial interest, such as factoring large numbers,[1] searching databases,[2] simulating intractablemodels from quantum physics,[3] and optimizing complex cost functions.[4] Here, we present an 11-qubit fully-connected, programmable quantum computer in a trapped ion system composed of 13171Yb+ ions. We demonstrate average single-qubit gate fidelities of 99.5%, average two-qubit-gatefidelities of 97.5%, and state preparation and measurement errors of 0.7%. To illustrate the capa-bilities of this universal platform and provide a basis for comparison with similarly-sized devices,we compile the Bernstein-Vazirani (BV)[5] and Hidden Shift (HS)[6, 7] algorithms into our nativegates and execute them on the hardware with average success rates of 78% and 35%, respectively.These algorithms serve as excellent benchmarks for any type of quantum hardware, and show thatour system outperforms all other currently available hardware.

Small universal quantum computers that can executetextbook quantum circuits exist in both academic[8–12] and industrial[13–17] settings. With a range oftwo to seventy-two qubits and sufficient fidelity for onlytens of entangling gates, these devices and the under-lying qubit implementations can be difficult to com-pare. Even within the trapped ion platform, thereis large diversity in atomic species, system architec-tures, and gate implementations. Trapped ion systemswith one to two qubits have shown single-qubit gatefidelities of 99.9999%[18] with microwave-based opera-tions and better than 99.99% fidelity with laser basedoperations,[19, 20] state preparation and measurement(SPAM) error below 10−4,[18, 21] and two-qubit gateswith fidelities exceeding 99.9%.[19, 20] Algorithms havebeen executed on up to seven trapped-ion qubits[22]and, while not optimized for universal quantum comput-ing, quantum simulators with more than 50 ions havemodeled fundamental quantum systems including Isingchains[23] and quantum magnetism.[24]

Benchmarking across implementations needs to beboth universal across platforms and agnostic to the dif-ferences in the underlying hardware. In traditional com-puting, the performance of computers is measured by ex-ecuting a set of benchmark problems representing var-ious use-case scenarios, to provide users with an esti-mate of how the computers would perform in their spe-cific applications. Canonical quantum algorithms demon-strate unambiguous advantage of quantum computersover classical computation, and provide verifiable out-comes to assess successful execution of the algorithm.Therefore, they can serve as ideal candidate problems

[email protected]

ION CHAIN

DETECTION

OPTICS

GLOBAL

ADDRESSING

BEAM

DIFFRACTIVE

OPTICAL

ELEMENT

MULTI-CHANNEL

AOM

INDIVIDUAL

ADDRESSING

BEAMS

FIG. 1. Schematic of the hardware. A linear chain of ionsis trapped near a surface electrode trap (trap is not shown).Lasers at 369 nm and 935 nm (not shown) illuminate all of theions during cooling, initialization, and detection. Each ion’sfluorescence is imaged through a 0.6 numeric aperture lens(“detection optics”) and directed onto individual photomul-tiplier tube channels. Two linearly-polarized counterpropa-gating 355 nm Raman beams are aligned to each qubit-ion,a globally addressing beam that couples to all of the qubits(red) and an individual addressing beam that is focused ontoeach ion (blue). Acousto-optic modulators (AOMs) modulatethe frequency and amplitude of each of these beams to gen-erate single-qubit rotations and XX-gates between arbitrarypairs of qubit ions.

for benchmarking the performance of any quantum com-puters. These benchmark algorithms exercise the fullhardware/software stack. A hardware-specific compilerbreaks down algorithms into the target hardware’s nativegate set, optimizing for qubit connectivity, gate times,and coherence[25] to enhance the system’s performance.After execution on the hardware, the measurements canbe directly compared with the expected output state todetermine the accuracy of the device. This accuracy canthen be compared with other devices that have compiled

arX

iv:1

903.

0818

1v1

[qu

ant-

ph]

19

Mar

201

9

2

and run the same algorithm.[26]

two-qubit fidelity1.000.990.980.970.960.950.940.93

single-qubit fidelity1.0000.9950.9900.985

FIG. 2. Fidelity of native gates. For each qubit pair, we per-form an XX-gate and measure the joint populations of thequbit pair as a function of an analysis pulse phase angle. Thefidelity of two-qubit gates are plotted as a color scale in the il-lustration of our all-to-all connectivity in (a). For each qubit,we perform randomized benchmarking to determine the fi-delity of the single-qubit gates shown in (b), which are plot-ted as the color scale of the nodes in (a). We use maximumlikelihood estimation to extract fidelities from the parity andjoint-population measurement shown in (c). The average two-qubit raw fidelity is 97.5% and all two-qubit gates perform inthe range [95.1%, 98.9%]. The distribution of these fidelitiesare depicted in the histograms above the color bars shown in(b,c). The fidelity of all single-qubit gates are enumerated inTable I and all two-qubit pairs are enumerated in Table II ofthe extended data.

We benchmark two algorithms on an IonQ trappedion quantum computer, shown schematically in Figure 1.Our qubit register is comprised of a chain of trapped171Yb+ ions, spatially confined near a microfabricatedsurface electrode trap.[27] For this work, we load 13 ions,the middle 11 of which are used as qubits. The two endions allow for a more uniform spacing of the central 11ions. The ions are laser-cooled close to their motionalground state using a combination of Doppler and re-solved sideband cooling. We encode quantum informa-

tion into the hyperfine sublevels, |0〉 ≡ |F = 0,mF = 0〉and |1〉 ≡ |F = 1,mF = 0〉 of the 2S1/2 ground state. Atthe beginning of each computation, each qubit is initial-ized to |0〉 via optical pumping with high accuracy. Af-ter qubit operations (described below), we read out thestate of all of the qubits simultaneously by directing laserlight resonant with the 2S1/2 |F = 1〉 to 2P1/2 transition,imaging each ion onto a detector and thresholding thephoton counts to determine if each qubit was in the |1〉(spin up) or |0〉 (spin down) state. Thresholding is doneby taking a histogram of the collected photons and dis-criminating between collecting on average zero photonsfor the |0〉 state and ten photons on average for the |1〉state.

A two-photon Raman transition drives single- and two-qubit coherent operations by applying a pair of counter-propagating beams from a mode-locked pulsed 355 nmlaser.[28] One of these beams globally addresses all ofthe ions simultaneously, while the other beam addressesany of the ions individually (Figure 1). The individually-addressing beams pass through a multi-channel acousto-optic modulator (AOM), which allows for the simultane-ous modulation of the phase, frequency and amplitude ofeach beam. To perform a single-qubit gate, we tune thefrequency difference between Raman beams to resonantlydrive a spin flip transition (|1〉 ↔ |0〉). In order to per-form a two-qubit gate, we off-resonantly drive motionalsideband transitions to generate an XX-interaction.[29]Both the global and individual beams are directed overthe trap surface perpendicular to the axis of the ion chainto excite one principal axis of motion transverse to thechain axis. Individual addressing allows us to performsingle- and two-qubit gates on any targeted qubits.

Native two-qubit entangling XX-gates are achieved bydriving a spin dependent force.[29] Using an amplitudemodulated (AM) pulse on any selected pair of qubits, weaddress multiple transverse motional modes of the ionchain to mediate a spin-spin Ising interaction betweenqubits.[30] To achieve high fidelity, the amplitude mod-ulation is calculated to simultaneously decouple all mo-tional modes from the spin at the end of the gate op-eration. Additionally, these pulse shapes are designedto provide robustness against frequency drift of motionalmodes and suppress residual off-resonant carrier excita-tion during the XX-gate.[30, 31] This gate, in conjunc-tion with single-qubit rotations, forms a universal gateset for performing circuit model quantum computation.Since the XX-gates are mediated by the collective motionof the ion chain, we have all-to-all connectivity betweenqubits, allowing two-qubit gates to be executed betweenany qubit pair (Figure 2a).

We perform randomized benchmarking[32] to charac-terize the single-qubit operations on each ion of the 11-qubit chain. We apply a randomly chosen sequence ofπ/2 gates with length L about the X and Y axes. Inbetween each of these π/2 gates, we either add a π rota-tion about the X, Y , or Z axis, or an identity operation(leaving the qubit idle for the duration of a gate). A

3

FIG. 3. Bernstein-Vazirani (BV) algorithm. (a) shows a textbook implementation of the BV algorithm with hidden bit string1010101010. (b) shows the full output distribution for all 1024 oracle implementations calculated from 500 iterations of eachoracle after conditioning on the ancilla. (c) shows the probability (inset plot) of detecting the encoded hidden bit string for all1024 oracle implementations, as a function of the number of ones in the binary representation of the unknown bit string, whichis equivalent to the number of two-qubit gates (n). The boxplots highlight the minimum, first quartile, median, third quartile,and maximum of the data. Note that there is only one oracle implementation for n = 0, 10. The shaded area spans the expected

fidelity (excluding crosstalk errors) Fn2QF

2(n+1)1Q F10

SPAM (where F2Q is the fidelity of two-qubit gates, F1Q is the fidelity of single-qubit gates, and FSPAM is the average SPAM fidelity) if all of our gates share the best measured fidelity or, alternatively, allshare the worst fidelity. The result of a shared average fidelity is plotted as a dashed line. The average probability of successis 78% with 899 out of the 1024 oracle implementations exceeding the 2/3 BQP single-shot success threshold.

final π/2 gate is chosen such that the final state is inthe Z computational basis (i.e. |0〉 or |1〉). We measurethe overlap between the measured and expected outputstates across 500 iterations for at least 24 sequences foreach L ∈ {2, 4, 6, 8, 10, 12}. The fidelity of our single-qubit π/2 gate is then determined by fitting the resultingoverlap as a function of sequence length to a power law,BpL + 1

2 . Here, the base p is the gate fidelity and the

intercept B + 12 is the SPAM fidelity, equivalent to mea-

suring the ion after a single π rotation when it is in eitherstate |0〉 or |1〉. For a chain of 11 qubits, we measure anaverage single-qubit fidelity of 99.5% (Figure 2b) and anaverage SPAM fidelity of 99.3%.

To quantify the performance of our two-qubit gatesand estimate their fidelity, we measure the state fidelityof the Bell state 1√

2(|00〉+ eiφ|11〉) prepared using a sin-

gle XX-gate by performing partial tomography of thestate.[19, 20] The diagonal terms of the two-qubit densitymatrix are extracted by measuring the populations in theeven parity states. The populations are measured whenthe overall AM pulse height for the XX-gate is tuned toachieve maximal entanglement such that the even-parity

two-qubit states, P00 and P11, are equal (P00 = P11).The off-diagonal elements are obtained from the ampli-tude Φ of a parity oscillation, where the parity is given byP00 +P11−P01−P10 (P01 and P10 are the populations ofthe odd parity two-qubit states). The fidelity can then becalculated as F = 1

2 (P00+P11+Φ).[20] We use maximumlikelihood estimation on experimentally observed data toextract the parameters of the fidelity expression.[20] Wehave performed this analysis for all 55 pairs of qubitsin the 11-qubit chain (Figure 2c) and measure an av-erage fidelity of 97.5% with a minimum and maximumfidelity of 95.1+0.5

−0.7% and 98.9+0.1−0.3%, respectively. The

uncertainty here is determined by a statistical confidenceinterval on the maximum likelihood estimation. The re-ported fidelity represents a lower bound of the Bell statecreation as we do not correct for SPAM errors on thetwo-qubit states or errors in single-qubit rotations usedto observe the parity oscillations of the Bell state, whichon average are 0.7% and 0.5% respectively.

To benchmark our system, we implement two well-known algorithms: the Bernstein-Vazirani (BV) andHidden Shift (HS). Both of these algorithms havepreviously been run on trapped-ion[10, 25, 33] and

4

FIG. 4. Hidden Shift (HS) algorithm implementation on 10 qubits. (a) shows a textbook implementation of the HS algorithmwith hidden shift 1111101010. The circuit for each oracle was measured at least 50 times. We trace out the spectator ion andinterpret the binary output state of the 10-qubit register as an integer. The full output distribution is shown in (b).(c) showsthe probability of detecting the encoded shift s for each of the 1024 oracle implementations versus the number of single-qubitgates (m). The shaded area represents the expected fidelity F10

2QFm1QF10

SPAM (where F2Q is the fidelity of two-qubit gates, F1Q

is the fidelity of single-qubit gates, and FSPAM is the average SPAM fidelity) if all of our gates share the best measured fidelityor, alternatively, all share the worst fidelity. The result of a shared average fidelity is plotted as a dashed line. The averageprobability of success is 35%, and 1017 of the 1024 oracle implementations correctly return the hidden shift as the maximalprobability state.

superconducting[11, 25, 26] systems of up to 5 qubits.By comparing the results of this algorithm to the idealresult, we obtain a direct measure of the system per-formance, which accounts for our native gates, connec-tivity, coherence times, gate duration, and all otherisolated metrics of system performance. These resultscan be used as part of a suite of algorithms to com-pare our hardware with other systems. The qubit num-ber in these results is higher than any comparable pub-lished BV or HS results using a programmable quantumcomputer.[10, 11, 25, 26, 33]

The BV algorithm is an oracle problem in which theuser tries to determine an unknown bit string c of size N ,implemented by a specific oracle. The algorithm takes abinary input string x and performs a controlled inversionof an ancillary bit or qubit based on the bit-wise productof the input and the unknown bit string c modulo two,f(x) = c · x (mod 2).[5] For a quantum BV implemen-tation (example shown in Figure 3a), a single quantumquery is sufficient to determine the bit string c.[2] Thisis a linear improvement over the best classical algorithm,which requires N queries. The BV algorithm was devel-oped to help separate a class of problems that can besolved in polynomial time on a quantum computer withbounded errors (BQP) from its classical counterpart. For

a problem to belong to BQP, a single quantum query suc-cess probability needs to be greater than 2/3.[5]

We compile the BV algorithm into our native gateset, comprised of single-qubit rotations and two-qubitXX-gates. Optimization during compilation reduces thenumber of needed gates compared to naively translatingthe textbook circuit from CNOT gates into rotations andXX-gates. The compilation exploits the full connectivityof our qubits, since we do not need SWAP operations.The implementation of BV requires a single-qubit ancillaand a register of N qubits. There are 2N possible bitstrings, therefore for our 10-qubit register there are 1024possible oracle implementations. We measured each im-plementation 500 times, conditioned upon on the mea-sured ancilla state, and plot the output distribution inFigure 3b. Each oracle implementation has, dependingon the unknown bit string c, between zero and ten two-qubit gates between the ancilla and the qubit register,corresponding to the number of ones in the binary repre-sentation of the unknown bit string. The process matrixthat maps the encoded oracle to the measured outputstate is nearly diagonal, resulting in a highly-peaked dis-tribution at the encoded oracle. For our system, the aver-age overlap between output state and unknown bit stringis 78% (Figure 3c), where 87.8% of oracle implementa-

5

FIG. 5. Comparison of data to noise model. (a,c) show, respectively, the first 128 by 128 section of Bernstein-Vazirani (BV)and Hidden Shift (HS) results for comparison with the result of a minimal error model including bit-flip errors and detectionmis-identification (b,d). Crosstalk errors produce different patterns of errors between the BV and HS due to the structure ofthe two algorithms.

tions achieve the 2/3 success criteria defined by BQP.Conditioning the output on the ancilla state results ina 5.1 percentage point increase in the raw success prob-ability of 73% to 78% and an 14.5 percentage point in-crease in the fraction of oracle implementations above theBQP threshold from 73.3% to 87.8%. The average over-lap in Figure 3c decreases with the number of two-qubitgates needed in the oracle. The off-diagonal componentsof the process matrix show errors since these should allhave zero population. In Figure 3b, the dominant erroris single-qubit bit-flips from |1〉 to |0〉 during the mea-

surement process, which appear as faint diagonals in thelower left quadrant of the figure (see Methods). However,even for the oracle implementation where we have thelowest success probability, the next-most-probable stateis still 4 times less likely than the correct string.

The HS algorithm consists of two N-bit to N-bit func-tion oracles f and g, which are the same up to a shift by ahidden bit string s, such that g(x) = f(x+s). The goal isto determine the hidden shift s by querying the oracles.In our implementation [7] of the HS algorithm, the oraclesare inner product or “bent” functions f =

∑i x2i−1x2i

6

and g = f(x+ s), where x is the input and xi is the i-thbit of x (an example is shown in (Figure 4a). Classically

it can be shown that determining the shift s requires√

2N

queries where N is the length of the bit string s. On aquantum computer, in principle, the shift can be read outin a single query.[6, 7] In contrast to the BV algorithm,the quantum implementation of the HS algorithm showsan exponential reduction in the number of queries to theoracle compared to a classical computer.[7]

As with the BV algorithm, we compile the HS algo-rithm into our native gates. There are 2N available or-acle implementations corresponding to the 2N possiblehidden bit strings s. We execute all 1024 possible im-plementations on our 10-qubit register (Figure 4b). Thecorrect output state is the state corresponding to thehidden shift. The average overlap between the outputstate and s was 35% (Figure 4c), and of the 1024 oracles,1017 had most likely output states corresponding to theshift. The success probability for HS is lower and moreuniform than that of BV because all of the oracles havethe same number of two-qubit gates (10) and many moresingle-qubit gates (25-40). Every oracle implementationin HS has at least as many gates as the most challengingBV oracle implementation and therefore is more diffi-cult. Given our average single- and two-qubit fidelities,we would not expect to surpass the BQP threshold forthe HS oracles. However, the successful determination ofthe shift was achieved much more frequently than if wesampled a classical distribution where the success prob-ability would have been 0.1%.

In summary, we have constructed the largest pro-grammable quantum computer to date that is capableof performing algorithms. We have used a trapped ionquantum computer to perform the largest quantum im-plementations of the Bernstein-Vazirani and Hidden Shiftalgorithms. Using a 10-qubit register, we implement all1024 possible oracles for each algorithm. We exceed theBQP threshold for 87.8% of the oracle implementationsin the Bernstein-Vazirani algorithm, an application de-signed to define this complexity class. We also demon-strate 35% overlap between the measured and expectedoutput states in the implementation of the Hidden Shift

algorithm, which is a more demanding application dueto its higher gate count and exponential speed up overits classical analog. The success of both algorithms isa result of high-fidelity native gates and efficient gatecompilation and compression in the fully-connected iontrap system. The demonstration of these two canoni-cal algorithms is a starting point for benchmarking anyquantum computer. Computing real problems on largersystems with more qubits will require even more gates inthe future with even higher quality, and similar standardalgorithms to those demonstrated here will likely play acrucial role in benchmarking quantum computers in thefuture.

METHODS

The dominant error in our data is single-qubit bit flipsdriven by addressing crosstalk. In the BV algorithm,these errors show up predominantly as qubits that arein state |0〉 when they should be in state |1〉 in the out-put state. We model this as 3% probability of incor-rectly preparing one qubit in the |0〉 state, which we ap-ply twice: first to model errors that occur in the imple-mentation of the oracle, and then to model errors in thesingle-qubit gates that occur before readout. We estimatethe 3% probability of incorrectly preparing one qubit byfitting by eye the relative amplitude of error states. Fi-nally, detection mis-identification (measuring the state ofone ion as other than it should be) is modeled at 0.2%probability on the resulting state, the measured SPAMinfidelity measured with microwaves (Table I). 1/16 ofour measurement results for the BV algorithm are shownin Figure 5a for comparison with the result of this errormodel (Figure 5b).

In the HS algorithm, the crosstalk errors show up asqubits that are measured to be in the wrong state (either|0〉 or |1〉). We model this as a 1% error of incorrectlypreparing one qubit in either state, which we apply 5times, followed 0.2% detection mis-identification. 1/16of our measurement results for the the HS algorithm areshown in Figure 5c for comparison with the result of thiserror model (Figure 5d).

[1] Shor, P. Polynomial-time algorithms for prime factor-ization and discrete logarithms on a quantum computer.SIAM J. Comput. 26, 14841509 (1997).

[2] Grover, L. K. Quantum computers can search arbitrarilylarge databases by a single query. Phys. Rev. Lett. 79,4709–4712 (1997).

[3] Trabesinger, A. Quantum simulation. Nature Physics 8,263 (2012).

[4] Farhi, E., Goldstone, J. & Gutmann, S. A quantumapproximate optimization algorithm. arXiv:1411.4028(2014).

[5] Bernstein, E. & Vazirani, U. Quantum complexity the-ory. SIAM J. Comput. 26, 1411–1473 (1997).

[6] van Dam, W., Hallgren, S. & Ip, L. Quantum algorithmsfor some hidden shift problems. SIAM J. Comput. 36,763–778 (2006).

[7] Rotteler, M. Quantum algorithms for highly non-linearboolean functions. Proceedings of the 21st Annual ACM-SIAM Symposium on Discrete Algorithms (SODA10)448457 (2010).

[8] Barends, R. et al. Superconducting quantum circuits atthe surface code threshold for fault tolerance. Nature508, 500–503 (2014).

[9] Monz, T. et al. Realization of a scalable shor algorithm.Science 351, 1068–1070 (2016).

[10] Debnath, S. et al. Demonstration of a small pro-

7

grammable quantum computer with atomic qubits. Na-ture 536, 63–66 (2016).

[11] Roy, T. et al. A programmable three-qubit su-perconducting processor with all-to-all connectivity.arXiv:1809.00668 (2018).

[12] Erhard, A. et al. Characterizing large-scale quantumcomputers via cycle benchmarking. arXiv:1902.08543(2019).

[13] Chow, J. M. et al. Universal quantum gate set ap-proaching fault-tolerant thresholds with superconductingqubits. Phys. Rev. Lett. 109 (2012).

[14] Devitt, S. J. Performing quantum computing experi-ments in the cloud. Phys. Rev. A 94, 032329 (2016).

[15] Pokharel, B., Anand, N., Fortman, B. & Lidar, D. A.Demonstration of fidelity improvement using dynamicaldecoupling with superconducting qubits. Phys. Rev. Lett.121, 220502 (2018).

[16] Hong, S. S. et al. Demonstration of a parametrically-activated entangling gate protected from flux noise.arXiv:1901.08035 (2019).

[17] Nam, Y. et al. Ground-state energy estimation of thewater molecule on a trapped ion quantum computer.arXiv:1902.10171 (2019).

[18] Harty, T. P. et al. High-fidelity preparation, gates, mem-ory, and readout of a trapped-ion quantum bit. Phys.Rev. Lett. 113, 220501 (2014).

[19] Gaebler, J. P. et al. High-fidelity universal gate set for9Be

+ion qubits. Phys. Rev. Lett. 117, 060505 (2016).

[20] Ballance, C. J., Harty, T. P., Linke, N. M., Sepiol, M. A.& Lucas, D. M. High-fidelity quantum logic gates us-ing trapped-ion hyperfine qubits. Phys. Rev. Lett. 117,060504 (2016).

[21] Crain, S. et al. High-speed, low-crosstalk detection of

a trapped 171Yb+

ion ancilla qubit using superconduct-ing nanowire single photon detectors. arXiv:1902.04059(2019).

[22] Landsman, K. A. et al. Verified quantum informationscrambling. Nature 567, 61–65 (2019).

[23] Zhang, J. et al. Observation of a many-body dynami-cal phase transition with a 53-qubit quantum simulator.Nature 551, 601–604 (2017).

[24] Bohnet, J. G. et al. Quantum spin dynamics and entan-glement generation with hundreds of trapped ions. Sci-ence 352, 1297–1301 (2016).

[25] Linke, N. M. et al. Experimental comparison of two quan-tum computing architectures. Proceedings of the NationalAcademy of Sciences 114, 3305–3310 (2017).

[26] Murali, P. et al. Full-stack, real-system quantum com-puter studies: Technology comparisons and architecturalinsights. ISCA Submission 1300 (2019).

[27] Maunz, P. L. W. High optical access trap 2.0. (2016).URL https://www.osti.gov/servlets/purl/1237003.

[28] Hayes, D. et al. Entanglement of atomic qubits using an

optical frequency comb. Phys. Rev. Lett. 104, 140501(2010).

[29] Sørensen, A. & Mølmer, K. Quantum computation withions in thermal motion. Phys. Rev. Lett. 82, 1971–1974(1999).

[30] Choi, T. et al. Optimal quantum control of multimodecouplings between trapped ion qubits for scalable entan-glement. Phys. Rev. Lett. 112, 190502 (2014).

[31] Wu, Y., Wang, S.-T. & Duan, L.-M. Noise analysis forhigh-fidelity quantum entangling gates in an anharmoniclinear paul trap. Phys. Rev. A 97, 062325 (2018).

[32] Knill, E. et al. Randomized benchmarking of quantumgates. Phys. Rev. A 77, 012307 (2008).

[33] Fallek, S. D. et al. Transport implementation of the bern-stein–vazirani algorithm with ion qubits. New Journal ofPhysics 18, 083030 (2016).

ACKNOWLEDGEMENTS

The authors would like to thank David Moehring forhis guidance in the construction of this apparatus, andthe EURIQA team at the University of Maryland andDuke University for sharing their designs and for usefulconversations.

AUTHOR CONTRIBUTIONS

Experimental data collected and analyzed by K.W.,K.M.B and S.D.; circuit compilation and gate design byY.N., S.D. and N.G.; the apparatus was designed andbuilt by K.W., K.M.B, J.M.A., S.D., JS.C., N.C.P., M.C.,C.C., K.M.H, J.M., J.D.WC., S.A., J.A., P.S., M.W.,A.D., A.B., S.M.K., V.C., M.K., C.M. and J.K.; K.M.B.and K.W. prepared the manuscript, with input from allauthors.

COMPETING INTERESTS

The authors declare no competing interests.

CORRESPONDENCE

Correspondence and requests for materials should beaddressed to Kenneth Wright (email: [email protected]).

8

Ion Gate Fidelity SPAM from RB SPAM from Microwave0 99.57(5) 99.31(9) 99.82(4)1 99.62(6) 99.1(1) 99.77(5)2 99.18(7) 99.3(1) 99.78(5)3 99.25(9) 99.6(2) 99.78(5)4 99.40(9) 99.3(2) 99.84(4)5 99.46(3) 99.32(7) 99.77(5)6 99.48(3) 99.27(6) 99.82(4)7 99.55(4) 99.40(8) 99.83(4)8 99.59(3) 98.94(6) 99.80(4)9 99.64(2) 99.35(4) 99.79(5)10 99.32(6) 99.3(1) 99.79(5)

TABLE I. Single-qubit randomized benchmarking (RB) re-sults and microwave SPAM results expressed in percentage(%). To determine single-qubit fidelities for each qubit weapply laser pulses to perform randomized benchmarking forπ/2 gates, using π gates to randomize the computational axes.The data are fit to a power law as described in the text. Theaverage single-qubit fidelity is 99.5%. We can obtain SPAMerrors from either the RB results or from a microwave pulse.For microwaves we tune the frequency of the microwave to thequbit splitting and the pulse time is set to drive a spin flipfrom |0〉 → |1〉, where the fidelity of detecting the |1〉 state isthe measured SPAM fidelity. The average SPAM fidelity fromRB is 99.3% and with microwave based operations the aver-age SPAM fidelity is 99.80%. The uncertainties for the RBresults are errors from the fit to a power law (see text) and theuncertainties for microwaves are statistical errors on a bino-

mial distribution,√

P|1〉(1−P|1〉)

nexpt, set by the photon counting

statistics.

1 2 3 4 5 6 7 8 9 10Ion 0

Ion 1

98.5+0.1−0.3 97.7+0.4

−0.5 98.5+0.1−0.3 97.2+0.4

−0.5 98.5+0.1−0.3 96.9+0.5

−0.5 97.2+0.3−0.5 98.7+0.4

−0.5 95.5+0.4−0.6 97.1+0.1

−0.3 0

97.7+0.4−0.6 98.9+0.1

−0.3 98.2+0.1−0.3 97.4+0.1

−0.3 97.8+0.1−0.3 98.1+0.1

−0.3 98.4+0.1−0.3 97.7+0.3

−0.5 97.9+0.1−0.3 1

98.0+0.2−0.3 97.5+0.3

−0.4 96.5+0.5−0.6 98.4+0.1

−0.3 98.0+0.1−0.3 97.2+0.3

−0.5 97.3+0.1−0.3 96.0+0.6

−0.6 2

96.4+0.4−0.5 97.4+0.1

−0.3 97.1+0.4−0.5 98.9+0.1

−0.3 96.0+0.3−0.5 98.0+0.1

−0.3 97.7+0.1−0.3 3

98.6+0.3−0.6 97.3+0.4

−0.4 97.3+0.5−0.5 98.3+0.4

−0.5 97.8+0.1−0.3 96.5+0.5

−0.6 4

96.5+0.4−0.6 97.1+0.3

−0.5 98.4+0.3−0.4 95.1+0.5

−0.7 96.7+0.5−0.6 5

96.2+0.4−0.6 97.2+0.3

−0.6 98.1+0.4−0.5 98.2+0.4

−0.5 6

97.3+0.4−0.6 98.5+0.3

−0.3 97.3+0.4−0.6 7

96.7+0.4−0.5 97.0+0.3

−0.6 8

97.5+0.4−0.5 9

TABLE II. Raw fidelity of native two-qubit gates expressed in percentage (%). For each qubit pair, we perform the gateand measure the joint populations of pair qubits as a function of analysis pulse phase angle to determine the parity contrastof the created Bell state. The resulting parity and joint-population are determined using maximum likelihood estimation toextract the fidelities enumerated above. The uncertainties are the 1σ confidence interval determined from maximum likelihoodestimation. The average fidelity is 97.5% with a minimum and maximum fidelity of 95.1% and 98.9% respectively.