Litteken Uchicago 0330D 16801

Download as pdf or txt
Download as pdf or txt
You are on page 1of 174

THE UNIVERSITY OF CHICAGO

ADAPTING COMPILATION STRATEGIES FOR ARCHITECTURE-SPECIFIC


FEATURES FOR QUANTUM COMPUTING

A DISSERTATION SUBMITTED TO
THE FACULTY OF THE DIVISION OF THE PHYSICAL SCIENCES
IN CANDIDACY FOR THE DEGREE OF
DOCTOR OF PHILOSOPHY

DEPARTMENT OF COMPUTER SCIENCE

BY
ANDREW LITTEKEN

CHICAGO, ILLINOIS
JUNE 2023
Copyright c 2023 by Andrew Litteken
All Rights Reserved
TABLE OF CONTENTS

LIST OF FIGURES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . v

LIST OF TABLES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xvii

ACKNOWLEDGMENTS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xviii

ABSTRACT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xix

1 INTRODUCTION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

2 GENERAL COMPILATION STRATEGIES FOR QUANTUM ARCHITECTURES 6


2.1 Background: Quantum Programs . . . . . . . . . . . . . . . . . . . . . . . . 6
2.2 Background: Quantum Hardware . . . . . . . . . . . . . . . . . . . . . . . . 8
2.3 The Quantum Compilation Problem . . . . . . . . . . . . . . . . . . . . . . 10
2.3.1 Decomposition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.3.2 Mapping for Quantum Architectures . . . . . . . . . . . . . . . . . . 12
2.3.3 Routing for Quantum Architectures . . . . . . . . . . . . . . . . . . . 13

3 NEUTRAL ATOM ARCHITECTURES: ADAPTING TO FLEXIBLE OPERATIONS


AND FRAGILE QUBITS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
3.1 Background: Neutral Atoms . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
3.2 Compiling for Neutral Atom Architectures . . . . . . . . . . . . . . . . . . . 18
3.2.1 Adapting Mapping and Routing . . . . . . . . . . . . . . . . . . . . . 18
3.2.2 Experimental Evaluation . . . . . . . . . . . . . . . . . . . . . . . . . 20
3.2.3 Evaluation Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
3.3 Error on Neutral Atom Devices . . . . . . . . . . . . . . . . . . . . . . . . . 28
3.4 Adapting to Atom Loss . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
3.4.1 Recovery Strategies . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
3.4.2 Experimental Evaluation . . . . . . . . . . . . . . . . . . . . . . . . . 37
3.5 Designing for Atom Loss . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
3.5.1 Constraining Methods . . . . . . . . . . . . . . . . . . . . . . . . . . 43
3.5.2 Evaluation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
3.5.3 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
3.6 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55

4 HIGHER-RADIX ARCHITECTURES: TAKING ADVANTAGE OF VARIABLY COSTED


OPERATIONS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
4.2 Background: Higher-Radix Computation . . . . . . . . . . . . . . . . . . . . 59
4.3 Background: Quantum Optimal Control . . . . . . . . . . . . . . . . . . . . 61
4.4 Routing Intermediate Qudit Circuits . . . . . . . . . . . . . . . . . . . . . . 62
4.4.1 Higher Radix Communication Circuits . . . . . . . . . . . . . . . . . 63
iii
4.4.2 Synthesizing Gate for Higher Radix Computation with Quantum Op-
timal Control . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
4.4.3 Adapting Compilation to Higher Radix Devices . . . . . . . . . . . . 66
4.4.4 Evaluation Benchmarks . . . . . . . . . . . . . . . . . . . . . . . . . 67
4.4.5 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
4.4.6 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
4.5 Qompress: Compiling Quantum Circuits Using Partial and Mixed-Radix Op-
erations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
4.5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
4.5.2 Compression and Gate Set . . . . . . . . . . . . . . . . . . . . . . . . 76
4.5.3 Synthesizing Mixed-Radix and Full-Ququart Gates . . . . . . . . . . 79
4.5.4 Experimental Demonstration of Single-Ququart Control . . . . . . . . 83
4.5.5 Adapting Compilation for Qubits-on-Ququarts . . . . . . . . . . . . . 85
4.5.6 Qubit Compression Strategies . . . . . . . . . . . . . . . . . . . . . . 88
4.5.7 Evaluation Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
4.5.8 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97
4.5.9 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
4.6 The Quantum Waltz: Exploiting Mixed-Radix and Full-Ququart Operations
for Native Three-Qubit Gates . . . . . . . . . . . . . . . . . . . . . . . . . . 106
4.6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
4.6.2 Three Qubit Gates on Ququarts . . . . . . . . . . . . . . . . . . . . . 109
4.6.3 Adapting Qubits on Ququarts to Three-Qubit Gates . . . . . . . . . 114
4.6.4 Using Three-Qubit Gates . . . . . . . . . . . . . . . . . . . . . . . . . 115
4.6.5 Evaluation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117
4.6.6 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122
4.6.7 Alternate Strategies . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130
4.6.8 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131

5 CONCLUSION AND FUTURE WORK . . . . . . . . . . . . . . . . . . . . . . . 133

6 REFERENCES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137

iv
LIST OF FIGURES

1.1 A representation of a classical compilation pipeline, where multiple device inde-


pendent optimizations are performed before passing the program off to a more
device specific compilation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.2 A representation of a classically inspired quantum compilation pipeline, where
multiple device independent optimizations are performed before passing the pro-
gram off to a more device specific compilation. . . . . . . . . . . . . . . . . . . . 3
1.3 A representation of a quantum compilation pipeline that uses a similar algorithmic
backing for each optimization step, but can be altered, using the specific details
of the device, for better compilation results. . . . . . . . . . . . . . . . . . . . . 4

2.1 The decomposition of the three-qubit Toffoli gate into one- and two-qubit gates. 7
2.2 A Bernstein-Vazirani example circuit. . . . . . . . . . . . . . . . . . . . . . . . . 7
2.3 An example of an architecture with six qubits, with connections such that it
forms a hexagon. A four-qubit Bernstein Vazirani has been mapped onto this
architecture. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.4 A routed Bernstein-Vazirani example circuit. . . . . . . . . . . . . . . . . . . . . 10

3.1 Examples of interactions on a neutral atom device. (a) Interactions of various


distances are permitted up to a maximum. Gates can occur in parallel if their
zones do not intersect. The interaction marked with green checks can occur
in parallel with the middle interaction. (b) The maximum interaction distance
specifies which physical qubits can interact. Compiler strategies suited for this
variable distance are needed for neutral atom architectures. (c) Neutral atom
systems are prone to sporadic atom loss. Efficient adaptation to this loss reduces
computation overhead. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16

v
3.2 An example of how SWAP paths of the same length can result in different inter-
action distances from other qubits on the architecture and how not all SWAPs
will have the same effect on locality as on hardware where the adjacency graph
is the interaction graph. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
3.3 Post compilation gate count across benchmarks. On the left are percent savings
over the distance 1 baseline averaged over program sizes up to 100 qubits. Each
color is a max interaction distance. Noticeably, there is less additional improve-
ment as the MID increases, indicating most benefit is gained for smaller distances.
On the right is a sample benchmark (holds in general) with many program sizes
compiled for the whole range of MIDs. As the program size increases, larger MIDs
show benefits before flattening off. . . . . . . . . . . . . . . . . . . . . . . . . . . 23
3.4 Post Compilation depth across all benchmarks. On the left, the reduction in
depth over the distance 1 baseline. Each bar is the average over all benchmark
sizes. On the right we see a similar drop off in post-compilation depth for the
QFT-Adder. We’ve chosen this specific benchmark to highlight the effect of
restriction zones. Here we show a subset of all sizes run. Depth initially drops
but for larger interaction distances some of this benefit is lost. We expect this to
be more dramatic for even larger programs. . . . . . . . . . . . . . . . . . . . . 25

vi
3.5 The induced restriction zone from interaction distance increases serialization. In
the prior results this is hard to discern because compared to low interaction
distance the amount of gate savings translates to depth reduction. Here we
compare benchmarks compiled with our restriction zone and a program with
no restriction zone, to mimic an ideal, highly parallel execution. The existence
of a restriction zone has the most effect on programs which are parallel to begin
with. On the right we directly compare this effect on the QAOA benchmark; solid
line is compiled with realistic restriction zone and dashed is ideal. The separation
between the corresponding lines signifies the effect of the restriction zone. . . . . 26
3.6 Compiling programs directly to three qubit gates reduces both gate count and
depth. Here we highlight a serial and parallel application written to three qubit
gates. Here dashed lines are compiled to two qubit gates decomposing all Toffoli
gates before mapping and routing. Solid lines compile with native Toffoli gates.
With native implementation of three qubit gates we obtain huge reductions in
both depth and gate count for both benchmarks. . . . . . . . . . . . . . . . . . 27
3.7 Program success rate as a function of two-qubit error rate. Because current NA
error rates are lagging behind competitive technologies we scan over a range of
two-qubit error rates for each of the benchmarks all on 50 qubit programs (49 for
CNU) with max interaction distance of 3. Examining pairs of solid and dashed
lines we can compare NA to SC. In the limit of very low two qubit error rate,
systems can support error correction. Both SC and NA systems scale at roughly
the same rate (slope of the line) but the NA system diverges from the completely
random outcome at higher error, allowing us to run programs on the hardware
much sooner. The further to the lower right a point is, the higher fidelity circuit
it can perform with worse error rate. . . . . . . . . . . . . . . . . . . . . . . . . 31

vii
3.8 Another way to examine the data of Figure 3.7 is to ask, given a desired program
success rate, what the required two qubit error rate is. Here we sweep again over
two qubit error rates and record the maximum program size to run with success
probability greater than 2/3. Again, examining pairs of solid and dashed lines
we can compare NA to SC. With the reduced gate counts and depth we expect
to be able to run larger programs sooner. . . . . . . . . . . . . . . . . . . . . . . 32
3.9 Examples of two different atom loss coping strategies. (a) shows the initial con-
figuration of three qubits, with the spare qubits in a light grey, and in use qubits
black. (b) Represents how the atoms are shifted into the spare qubits to accommo-
date a lost atom under the virtual remapping strategy. Notice that the interaction
is no longer within interaction distance 1. (c) Demonstrates how rerouting strate-
gies swaps a qubit to a valid interaction configuration, then returns it. Numbers
indicate the order of swaps. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
3.10 Atom loss as a percentage of total device size which can be sustained before a
reload of the array is needed. Each program is 30 qubits on a 100-qubit device.
As the interaction distance increases most strategies can sustain more atom loss.
Strategies like full recompilation can sustain large numbers of atom loss but as
we will see are expensive computationally. Fast, hardware solutions or hybrid
solutions can sustain fewer numbers of holes but have lower overhead. We show
two representative benchmarks parallel vs. serial. “im” represents the interaction
graph methods. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37

viii
3.11 For strategies which modify the program such as recompilation or rerouting strate-
gies, additional gates could be added leading to a lower overall success rate. Here
we trace the success rate of our three program modifying strategies. The full
recompilation strategy (circles) is a rough upper bound which best accounts for
holes as they appear able to move the entire program to a more appropriate lo-
cation and route best. The gap between strategies on the same MID gets smaller
as the MID gets larger. Here we’ve chosen the two-qubit error rate corresponding
to approximate 0.6 success rate to begin with (based on Figure 3.8) in order to
best demonstrate the change in shot success probability over a range of atom loss. 40
3.12 Strategies that are able to reduce the number of reloads necessary greatly reduce
the overhead time when running circuits. Here we show the overhead time for all
strategies except recompilation. The proportion of time dedicated to reloading
is shown by the dominant color in each bar, followed by fluorescence in red, and
recompilation in black. Any strategy whose overhead exceeds that of always
reloading, such as full recompilation, should not be considered. . . . . . . . . . . 41
3.13 A timeline of 20 successful shots for Compile Small and Reroute with reload
time of 0.3 s and fluorescing time of 6 ms. A majority of the overhead time is
contributed by the reload time and fluorescence, indicating that the duration and
count of these actions is crucial to overall runtime. . . . . . . . . . . . . . . . . 42
3.14 Sensitivity to the rate of atom loss for the balanced Compile Small and Reroute
strategy. In prior experiments we used a fixed rate of 2% atom loss. For larger
systems this rate could be worse and in the future we might expect this rate to
be much better. For each interaction distance we see as the rate of atom loss gets
better we can run many more trials before we must perform a reload and reset.
Some error bars don’t show on the log axis. . . . . . . . . . . . . . . . . . . . . 43

ix
3.15 Examples of four different atom loss coping strategies. (a) Shows the initial
configuration of qubits using the original compiler that makes use of the entirety
of the architecture. (b) shows the result of using the hardware graph to shift the
qubits away from a newly lost atom. (c) shows how the interaction graph could
make a different decision, taking the shortest interaction path to an open qubit.
(d) demonstrates how the architecture can be divided such that we have multiple
opportunities to start in a section with fewer lost qubits. (e) shows how a wear
leveling approach would appear after one remapping of the circuit. (f) shows
how these divided sections can be used to exploit some level of parallelism in the
architecture. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
3.16 Decreases in success rate for different mitigation strategies for 30 qubit circuits.
Each color is a different strategy, solid lines are circuits with interaction distance
3, and dash lines are interaction distance 5. Run over 50 trials, each error bar
represents one standard deviation from the mean. We see decreases for all strate-
gies, with full recompilation able to maintain the highest probability of success
and relocation better approximating recompilation than our baseline strategy. . 47
3.17 Decreases in success rate for different mitigation strategies for 90 qubit circuits.
We exclude QAOA from this test as the success rates are not meaningful at
current error rates. Each error bar represents one standard deviation from the
mean over 50 trials. We see the same patterns, with the exception that interaction
remapping is often able to outperform other strategies. . . . . . . . . . . . . . . 47
3.18 The average number of shots per reload 10, 20 and 30 qubit benchmarks at
interaction distance four. Error bars indicate one standard deviation from the
mean. Each color is a different recovery strategy. We find that there is significant
improvement from our relocation strategies (red and green), greatly exceeding
the baseline number of shots, more closely matching recompilation. . . . . . . . 49

x
3.19 The average number of shots per reload 50, 70 and 90 qubit benchmarks at
interaction distance four. Error bars indicate one standard deviation from the
mean. Each color is a different recovery strategy. We exclude QAOA since it is
not a practical circuit at this size. . . . . . . . . . . . . . . . . . . . . . . . . . . 49
3.20 The normalized advantage of the relocation strategy to rerouting and interaction
distance four. Each color represents a different circuit, and the advantage of
relocation decreases at a factor inversely proportional to the size of the circuit. . 50
3.21 Overhead times for different recovery strategies at different maximum interaction
distances. The major color in each bar represents time dedicated to reloading.
As follows from the significant advantage seen previously, relocation is a very
efficient strategy. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
3.22 Decreases in success rate for different mitigation strategies for parallel 30 qubit
circuits. Run over 50 trials, each error bar represents one standard deviation
from the mean. In general, we see that for higher interaction distances, increased
parallelism does not heavily affect the rate of decrease in probability of success
rate. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
3.23 The average number of shots per reload for 10, 20 and 30 qubit benchmarks at
interaction distance four at different levels of parallelism. Error bars indicate one
standard deviation from the mean. Each color is a level of parallelism. Parallel
indicates that as much of the architecture is used as possible, and relocation
means that only one instance is run at a time. . . . . . . . . . . . . . . . . . . . 54

4.1 The decomposition of a qutrit based SWAP gate. . . . . . . . . . . . . . . . . . 64


4.2 A representation of the dynamic map used for mapping and routing when routing
quantum devices in higher, mixed radix configurations. Each edge is weighted
with the different time it will take to perform a SWAP along that edge. . . . . . 67

xi
4.3 A logarithmic depth Generalized Toffoli circuit using qubits with 5 controls and
4 ancilla. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
4.4 A logarithmic depth Generalized Toffoli Circuit using qutrits with 7 controls. . . 68
4.5 A logarithmic depth Generalized Toffoli Circuit using ququarts with 7 controls. . 69
4.6 A logarithmic depth Generalized Toffoli Circuit using ququints with 10 controls. 70
4.7 The depth of the Generalized Toffoli gates at different levels of qudits before
communication gates are added to the circuit. . . . . . . . . . . . . . . . . . . . 71
4.8 The depth of the Generalized Toffoli gates at different levels of qudits after com-
munication gates are added to the circuit. . . . . . . . . . . . . . . . . . . . . . 71
4.9 The circuit duration of the Generalized Toffoli at different controls for different
qudit levels. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
4.10 The space time product, or number of qudits used by duration, at different con-
trols for different qudit levels. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
4.11 Pairs of qubits can be compressed in four-dimensional ququarts and interact with
each other internally or through partial operations, enabling novel compilation
techniques and space reduction. . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
4.12 Two qubits q0 and q1 can be encoded into a ququart (blue oval) and interact
with each other internally (green), with a bare qubit q outside (yellow), or with
encoded qubits in a different ququart (red). CX arrows point from control qubit
to target qubit. Along each CX link, corresponding SWAP gates are defined with
the same superscripts/subscripts (not shown). . . . . . . . . . . . . . . . . . . . 78
4.13 Exemplary state evolutions of two CX gates between (a) two bare qubits and (b)
a bare qubit and an encoded qubit. (a) The control qubit q0 is in state |1i, hence
the state of the target qubit q1 is flipped. (b) The encoded qubit q0 inside the
ququart controls the CX gate targeting the bare qubit q. The ququart state |3i
corresponds to the two-qubit state |q0 q1 i = |11i, so the state of q is flipped. . . . 81

xii
4.14 Interleaved Randomized Benchmarking for an optimal control H ⊗ H pulse on
a superconducting transmon ququart following our qubit encoding. We use two-
qubit Clifford sequences of gate depth up to 100 and average each data point over
10 samples. Error bars show the standard deviation of the mean but they are
smaller than the mean markers. Red: Standard two-qubit Randomized Bench-
marking to estimate the average Clifford gate fidelity to be FRB ≈ 95.8%. Blue:
Interleaving the H ⊗ H pulse between the RB Cliffords yields a combined per-
operation fidelity of FIRB ≈ 92.1%, resulting in an H ⊗ H fidelity FHH ≈ 96.0%. 84
4.15 An example of an exhaustive search on a cylinder-based interaction graph (a),
using a (b) critical path focused selection, and a (c) selection strategy that allows
for any pair to be selected. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
4.16 (a), (b) Sample generalized Toffoli gate with corresponding interaction graph. (c),
(d) Cuccaro adder with interaction graph, M and N are three qubit gate blocks.
Some circuits, such as the CNU circuit and Cuccaro adder have clusters of qubits
that interact with one another. These clusters can be identified by finding the
cycles in the interaction graph. We have identified the cycles here with colored
edges. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
4.17 Examples of graph-based circuit interaction graphs. . . . . . . . . . . . . . . . . 97
4.18 Expected Gate Probability of Success for each benchmark. Each color represents
a different compilation strategy, where the black line is the exhaustive solution
developed previously, and is the goal. FQ is the previous baseline for generalized
ququart computation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
4.19 The distribution of gate types for different pairing strategies for a 30 qubit Torus
QAOA circuit. Each color represents a different “style” of gates. In particular the
darker blue represents CX gates between two ququarts, and the red represents an
internal CX gate within a single ququart. . . . . . . . . . . . . . . . . . . . . . . 100

xiii
4.20 Gate Expected Probability of Success as the qubit gate error rate increases and
the ququart error gate rate stays constant. The black line represents the crossover
point where the error of the qubit only compilation is greater than ququart com-
pilation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
4.21 Expected Coherence Probability of Success for each benchmark. Each color rep-
resents a different compilation strategy, where the black line is the exhaustive
solution, and in theory ideal, developed previously. EC line stops short for com-
putational reasons, requiring many more classical resources. FQ is the previous
baseline for generalized ququart computation. . . . . . . . . . . . . . . . . . . . 102
4.22 Expected Coherence Probability of Success for Cuccaro and Torus QAOA with
10x better ququart and qubit T1 times. . . . . . . . . . . . . . . . . . . . . . . . 103
4.23 Expected Coherence Probability of Success for several 25 qubit benchmarks with
10x better ququart and qubit T1 times as the ququart time increases T1 from 1/3
the qubit T1 time to the entire qubit T1 time. Each dashed line represents the
point where the coherence success rate no longer outweighs the gate success rate
gains for the corresponding circuit. . . . . . . . . . . . . . . . . . . . . . . . . . 104
4.24 Ranges of gate based probability of success for CNU and Cylinder QAOA on
three different architectural topologies. This is the combined set of ratios of
improvement for circuits sizes 5 to 40. . . . . . . . . . . . . . . . . . . . . . . . 105
4.25 A comparison of executing a Toffoli gate on a three-qubit-only system versus a
Toffoli gate between a ququart and qubit in a mixed-radix system. In a qubit-
only system, we must use a decomposition that uses eight two-qubit gates that
can be reduced to one two-qudit gate that has a shorter duration. . . . . . . . . 107

xiv
4.26 Visualization of connectivity advantages in qubit-ququart systems. Encoding
qubits in ququarts (lightblue) enables triangle connectivity between triples of
qubits, where two of which are encoded in the same ququart and one appears
either in a bare qubit or encoded in a neighboring ququart. . . . . . . . . . . . . 110
4.27 Visualization comparing the evolution of a |3i-controlled X gate in a mixed-radix
environment for a CCX gate in (a) and a CX gate in (b). . . . . . . . . . . . . . 111
4.28 Examples of two-control and two-target three-qubit gates in a mixed radix envi-
ronment. a) The Toffoli with two controls q0 and q1 . b) A configuration where
both controls are encoded in the ququart and the target is mapped to a qubit.
c) A configuration where the controls are split across the qubit and the ququart
and the target is encoded in the ququart. d) The CSWAP gate with two targets
q1 and q2 . e) A configuration where both targets are encoded in the ququart and
the control is mapped to the qubit. f) A configuration where the targets are split
across the qubit and the ququart and the control is encoded in the ququart. . . 113
4.29 Different decompositions for the Toffoli Gate. a) is the base Toffoli circuit. b)
is Toffoli circuit with a swapped second control and target from the original. By
surrounding the control and the target with Hadamards, we perform the same
operation. c) The Toffoli gate constructed from a CCZ gate which can be used
as a Toffoli by surrounding the target with Hadamard gates. (d) The Toffoli gate
constructed from an iToffoli gate, which requires a controlled S † gate in addition
to Hadamard gates. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114

xv
4.30 Simulated results for QRAM, Generalized Toffoli, Cuccaro Adder and Select Cir-
cuit from 5 to 21 qubits with different mixed-radix and full-ququart compilation
strategies. The mixed-radix strategies do not have complete error bars due to the
requirement to simulate a four-level system for every qubit which would require
more than 86 GB of memory per circuit in our simulation framework. The final
graph is the average fidelity improvement for each compilation method over the
qubit-only compilation method as the size of the circuit increases. . . . . . . . . 123
4.31 EPS statistics for the generalized Toffoli circuit. We show the gate and coherence
EPS on the left and the product EPS on the right. . . . . . . . . . . . . . . . . 124
4.32 The results of several sensitivity studies. (a) Changes in CCZ compilation strate-
gies’ fidelities as gate error ququarts increases. (b) Changes in CCZ compilation
strategies’ fidelities as coherence error for the |2i and |3i level states changes. (c)
Differences in fidelities between mixed-radix and full-ququart compilation strate-
gies as the distribution of CX gates to CCX gates in a circuit changes. In all
graphs The black line represents the qubit-only compilation method. Below this
point mixed-radix or full-ququart methods are more error prone than using only
qubits. Please note the different scaling on the y-axis. . . . . . . . . . . . . . . . 127
4.33 The results of the different mixed-radix and full-ququart compilation strategies
on random mappings of only qubit-encoding, or only ququart encoding. . . . . . 129

xvi
LIST OF TABLES

3.1 Full Runtimes with Two Instances of Parallelism . . . . . . . . . . . . . . . . . . 55

4.1 Times used for various gates across different levels of qudits. An asterisk indicates
an interpolated value. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
4.2 Gate durations for one and two-qubit gates synthesized in the qubit-only, mixed-
radix and full-ququart environments. . . . . . . . . . . . . . . . . . . . . . . . . 81
4.3 Mixed-Radix and Full-Ququart Three-Qubit Gate Durations . . . . . . . . . . . 110

xvii
ACKNOWLEDGMENTS

I want to first thank my advisor Fred Chong for his guidance, flexibility and opportunities he
has provided me with over the course of my PhD. I’d also like my committee Hank Hoffmann
and Robert Rand who have given their time to advise me on my dissertation and previous
projects. I’d like to also thank all of my collaborators - Jonathan Baker, Lennart Maximillian
Siefert, Jason Chadwick, Natalia Nottingham, and Casey Duckering.
I would also like to thank all of those who have encouraged, supported and offered their
advice as I’ve worked towards my PhD. Through all of my studies, I have had many incredible
teachers, professors, colleagues and friends who prepared me for whatever came next. And, I
would like to thank all of my family, who have always been on my side and made it possible
to be where I am today.

xviii
ABSTRACT

Quantum Computing is in an era of growing pains. As quantum algorithms become more


tuned to the computational capabilities of near-term quantum devices, compilation tech-
niques that fully utilize the architecture while avoiding potential pitfalls are a necessity.
Many different types of quantum devices are being developed each with its own advantages
and unique features. It is unclear which architecture will be dominant, and many are in
active development. Some problems, such as limited connectivity, imperfect gate execution,
and lower coherence times are shared across architectures, and mapping, routing and com-
munication algorithms have been developed to circumvent and avoid these issues. But, each
architecture has its own set of specific problems to overcome. It is not simply the case that
a single compilation pipeline or set of algorithms will be able to make the best use of each
of these architectures. This work proposes an exploration into adapting a general quantum
compiler algorithm to several different architectures.
This thesis will focus on two main architectures, Neutral Atom and Superconducting qu-
dit devices. Neutral Atom architectures have additional flexibility in routing qubits beyond
nearest neighbor connectivity, while incurring some serialization cost due to increasingly
large “areas of restriction”. But, we are still able to take advantage of these more flexible
interactions through careful mapping and routing. Additionally, we can craft strategies that
make use of these features to circumvent Neutral Atom Architectures’ biggest execution
downfall: atom loss. Many quantum architectures can be naturally extended to extra com-
putational states beyond the traditional |0i and |1i used in classical computing and most
quantum computation. These extra states can be used to reduce the number of operations,
but at the cost of slower operations and decreased stability of the computational units. Once
again, we can adapt an existing base algorithm to mix-and-match qubit-only, mixed-radix,
and higher-radix operations to develop efficient programs that make use of less resources and
increase the chances of successful quantum computation.

xix
CHAPTER 1

INTRODUCTION

Quantum computing has seen a renewed explosion in development, investment and interest
in the past decade as new technologies have been built and architectures with more qubits
have been constructed such as those from IBM [40], Google [5, 56], Rigetti [94], IonQ [23] and
many more. New machines have seen marked success over prior designs, and new applications
have been developed to take advantage of the current state of quantum computing such as
QAOA [35] and VQE [112], beyond the flagship algorithms such as Grover’s Search [46]
and Shor’s factoring algorithm [103]. There are ongoing collaborations between scientific
and technology-based communities as they continually push to develop this fundamentally
different form of computation.
In classical computing, there is a well-defined strategy for general purpose computation.
While there are certainly variations in certain processor architectures and instruction set
differences, there has been consolidation on the Von-Neumann based architectures. This has
allowed for the development of device-independent optimizations at a higher, intermediate
representation level [66], where compilers have been able to take advantage of the similarities
of a shared model to reduce program metrics such as runtime, memory usage, and code size.
After device specific optimization has been performed, the program is compiled for a specific
processor, using optimizations designed for those platforms, such as register allocation, in-
structions selection and specific assembly language optimizations for those architectures. It is
a very sequential system due to the general principles underlying most computing platforms
following the model demonstrated in Figure 1.1.
This remains true for many accelerators, such as GPUs and FPGAs. Programs may
have to be written in certain dialects specific for these devices, but they still have a general
strategy for computation. That is, strategies that work on one device, will work on another.
There can be device-independent optimizations, before using device-specific features that
1
Figure 1.1: A representation of a classical compilation pipeline, where multiple device inde-
pendent optimizations are performed before passing the program off to a more device specific
compilation.

take advantage of the design decisions for a specific piece of hardware.


At first glance, it may seem that quantum computers should be able to take advantage
of the accelerator model. Each uses the same general framework for quantum computation
but has different strategies to achieve the same goal. However, the current era of quantum
computing, commonly referred to as the Noisy Intermediate-Scale Quantum Computing
(NISQ) era [90], is defined by error-prone gates, unstable qubits and low qubit counts, and
the best path towards fault-tolerant quantum computing is still unclear. There is active
development among many different styles of architectures such as superconducting qubits,
trapped ions, neutral atoms and photonic computing. Each has a different set of trade-
offs, whether that be ease of control, fidelity of operations, qubit connectivity, number of
qubits, or qubit lifetime. There are some device independent optimizations that can be
performed, such as gate cancellation and resynthesis, but can only reduce gate count and
circuit duration at a high level. To be useful in the current era, each architectural style uses
different optimization strategies throughout the entirety of the stack in order to be used
most effectively.
Most quantum compilation frameworks such as tket [105], Qiskit [4], Cirq [31], include
some amount of device independent optimizations to process a quantum circuit, translating
it from the general circuit model to a new circuit that fits the connectivity and native
gate constraints of the architecture in a compilation flow detailed in Figure 1.2. Usually,
these optimization passes are parameterized by the details of the machine such as the qubit
2
Figure 1.2: A representation of a classically inspired quantum compilation pipeline, where
multiple device independent optimizations are performed before passing the program off to
a more device specific compilation.

connectivity and the native gate set, but the optimization algorithms themselves are device
independent. These strategies fail to consider the benefits and drawbacks of the architecture
itself. While these strategies reduce gate count and circuit duration to some degree, they may
actively work against the architecture in other cases. One example is the need for a specific
router that schedules qubit communication to avoid crosstalk [79]. Put simply, general
purpose optimizations only get us so far. However, it is not practical to write routers and
optimizations for each architectural-style. The different features across architectures with the
same underlying technology become too varied to redevelop optimizations for each system
that may be relatively similar, but the details are somewhat different. This is especially true
since the foundations of each algorithm will likely be very similar to one another across each
architecture.
We can instead adapt general algorithms for quantum computing to be specific to hard-
ware circuits are compiled for, instead of inventing new technologies whole cloth. Instead of
having device-independent optimizations feed into a device specific pipeline, we can share a
common optimization algorithm “shell” that is adapted to the specific, unique and efficient
features of an architecture. This structure is visualized in Figure 1.3. While this work does
not propose a compiler framework that successfully achieves this goal, it describes differ-
ent compilation algorithms and how they can be adapted for several different architectural
features.
We first describe the basics of quantum computing. We start with the general strategy for
3
Figure 1.3: A representation of a quantum compilation pipeline that uses a similar algorith-
mic backing for each optimization step, but can be altered, using the specific details of the
device, for better compilation results.

computation and how it can achieve the expected computational speed-ups. We also describe
the circuit model that is usually used during compilation to describe quantum programs and
is used when developing optimizations at the circuit level. We follow this with a description
of a general compiler algorithms developed previously and is used in several other works, but
usually with a focus on a specific architecture or avoiding a specific kind of error [10, 77].
Second, we explore Neutral Atom architectures, an emerging architecture with unique
features, such as long-distance interactions, and the ability to perform native multi-qubit
gates. But the use of these features can induce serialization of gate execution causing circuit
slowdowns. The unique features of neutral atoms have significant opportunities for compi-
lation to take advantage of. Additionally, they also uncover a solution to a major detriment
of this kind of architecture, atom loss. An individual atom represents a single qubit in the
system, and if an atom is lost, it could require an entire reloading of the architecture, a
time intensive process. The unique features of neutral atoms allow for runtime and compiler
opportunities to recover failed shots in software, preventing the need for a full reset of the
device.
Third, we adapt our techniques for three different instances of higher radix qudits. That
is, quantum devices that can access the states beyond the traditional |0i and |1i states that
are used for qubits. Higher radix devices are more error prone, requiring more calibration
and longer gate times to accurately alter the state. This presents an interesting trade-off
for compilation. The extra computational space provides a reduction in physical devices

4
required to perform computation. In theory, with access to four states, it would be possible
to represent two times the information than an architecture with the same number of devices,
but with only access to two levels per device. Finding a trade-off between the extra coherence
error and circuit duration versus fewer gates and saved space is a difficult compiler problem
to solve. In the first application, we stay within the bounds of the native gate sets translated
from qubit computation for higher radix qudits, extending work from [42] where qutrits
were used for gate and circuit depth reduction. The second and third applications cover a
compression strategy where we take advantage of the fact that two qubits can be encoded into
the four levels of a ququart quantum device. This significantly increases qubit connectivity,
and the speed of some gates between two qubits encoded in the same ququart. However,
qudits have lower lifetimes, and the speed of gates across ququarts are increased. We adapt
compiler algorithms, and alter the native gate set to dynamically establish the best trade-off
between these different features and drawbacks. Additionally, the development of partial-
ququart operations that do not require the decoding into two qubits from a single ququart
greatly expands the viability of qubit-on-ququart computation. Further, in the third qudit
section, the development of three-qubit gates that have much faster execution time can
improve specific applications. And, through the observation that the configuration of these
three-qubit gates affects the duration of the gates, we can develop compiler algorithms that
take advantage of this fact to further improve ququart based computation.
These varied architectures demonstrate the need for more adaptable compiler frameworks
as in an era of varied technologies and error. Fundamentally, it is important to develop
quantum compiler pipelines built from common pieces that can be strategically customized
to fit the unique features of a quantum architecture.

5
CHAPTER 2

GENERAL COMPILATION STRATEGIES FOR QUANTUM

ARCHITECTURES

This chapter introduces the basics of how quantum programs are constructed and the con-
straints we will be working within when compiling quantum programs to quantum hardware.

2.1 Background: Quantum Programs

The fundamental unit of quantum computing is the quantum bit, or qubit, which exists as a
linear superposition between the |0i and the |1i states. Most quantum programs manipulate
a register of N quantum bits where the state of the qubits is given as a linear superposition
of the 2N basis vectors. This state evolves by the application of operations or gates. For
example, the single qubit X gate transforms |ψi = α |0i + β |1i into X |ψi = β |0i + α |1i.
Some gates, such as the Hadamard gate, push the qubit into a state of superposition, existing
in the |0i and |1i state at the same time. For example, the Hadamard gate transforms
|ψi = |0i into H |ψi = √1 |0i + √1 |1i. Half the time, we will measure this qubit to be in
2 2
the |0i state, and half the time we will measure it to be in the |1i state.
Gates can also operate on multiple qubits at a time. For example, the CNOT gate
applies an X gate to the target qubit if and only if the control qubit is in the |1i state.
This CX gate transforms the bitstring |ψi = α |00i + β |01i + γ |10i + δ |11i into CX |ψi =
α |00i + β |01i + δ |10i + γ |11i. These multiqubit operators induce entanglement between
the qubit states. The measured state of one qubit will influence the measured state of qubits
it is entangled with. Together, superposition and entanglement are central to the expected
power of quantum computation and to solving problems that are intractable on classical
computers.
While most quantum compilation focuses on manipulating the one- and two-qubit gates,

6
• • • • T

• = • • T T†

H T† T T† T H

Figure 2.1: The decomposition of the three-qubit Toffoli gate into one- and two-qubit gates.
|Ψ1 i H • H

|Ψ2 i H • H

|Ψ3 i H • H

|Ψ4 i X H H
Figure 2.2: A Bernstein-Vazirani example circuit.

an operation can operate across any number of qubits. An example of this is the three-
qubit Toffoli gate, which changes the |0i and |1i position of a target qubit if both of the
control qubits are in the |1i state. These N −qubit operators can be decomposed into a set
of universal one- and two-qubit gates that represent the same set of operations. However,
this can often insert many extra gates into a circuit as replicating a gate exactly can be an
expensive process and is an open problem [86]. An example of a decomposition of the Toffoli
gate can be found in Figure 2.1.
Most quantum programs for gate-based quantum computation are expressed in the quan-
tum circuit model. In this model, each qubit is represented by a wire in a diagram, and each
gate is presented by a box or circle on a qubit with a progression of operations from left to
right. A multi-qubit gate is represented by a box that touches multiple qubits. Generally,
when we have a controlled operation, like a CX gate, the qubit that controls the target is
represented by a filled in dot. A small Bernstein-Vazirani circuit is found in Figure 2.2.
There are many more aspects to quantum computing, such as error correction, but this
foundation is all that is necessary for this thesis. For a complete introduction to quantum
computing see [83].

7
2.2 Background: Quantum Hardware

Quantum architectures are in a time of rapid development and severe constraints. In this era,
there are several limitations facing quantum hardware, and different architectural techniques,
such as superconducting qubits, ion traps, and neutral atoms have different trade-offs.
The first of these constraints is imperfect gates. Each of the listed architectures have error
prone gates and cannot perform operations perfectly. For superconducting architectures one-
qubit gates can be performed with 99.9% fidelity, and two-qubits gates with 99.4% fidelity
[2]. In some cases, this may even be variable depending on which qubits are interacting.
This is similar for ion traps with 99.99936% and 99.99987% [19]. Neutral Atoms, which are
earlier in development, have the lowest with 99.6% and 96.5% [67]. Due to these error prone
gates, programs cannot contain too many operations. As more operations are added, the
probability of a successful circuit decreases as well as each additional operation could fail.
When we begin compiling circuits to hardware, we must be conscious of this fact and add
as few gates as possible and remove as many redundant gates as we can.
A second constraint is gate durations and qubit coherence time. Current qubits are
fragile and can only maintain a non-ground energy state for a certain period of time before
it decoheres, and the quantum state collapses. In most cases, this will result in the wrong
measurement. Similar to how more gates increases the probability of error, the longer a
quantum program runs, the more likely a qubit is to decohere. Where superconducting
qubits have relatively high gate success rates, it has the shortest coherence times of 121.1 µs
[2], due in part to the difficulty of manufacturing high fidelity superconducting chips [82].
On the other hand neutral atoms and ion traps have much longer coherence times on the
order of minutes [34, 115]. Consequently, the speed of gates is much more important for
superconducting devices as compared to ion traps and neutral atoms which use single atoms
to represent qubits, using lasers and optical tweezers to keep each in position.
Another significant issue facing quantum computers is interacting qubits with one another

8
Figure 2.3: An example of an architecture with six qubits, with connections such that it
forms a hexagon. A four-qubit Bernstein Vazirani has been mapped onto this architecture.

on a device with limited connectivity. That is, when not all qubits can interact with any other
qubit. Currently quantum computers are constructed of up to hundreds of qubits. Even at
this scale, it is not possible to interact any one qubit with any other qubit. For example,
see Figure 2.3, where we have a hexagon connectivity. However, we can define connections
between qubits, and use a special operation called a SWAP gate that exchanges the quantum
data between two physical qubits. This allows us to move qubits across the architecture in a
process called communication. The process of adapting the circuit in 2.2 to fit this topology
can be seen in Figure 2.4 with the initial mapping in 2.3. In superconducting devices, there
is very low connectivity as the resonators between qubits must be defined on a physical
chip. On the other hand, ion traps and neutral atoms have longer interaction distances, and
somewhat higher connectivity. As circuit size grows, so will the need for communication.
Adding extra gates incurs error from both the gate success and the coherence of qubits.
Ensuring that as little communication is used as possible is important when running actual
circuits on quantum hardware.

9
|Ψ1 i H • × × • H

|Ψ2 i H • H

|Ψ3 i H × × H

|Ψ4 i X H H
Figure 2.4: A routed Bernstein-Vazirani example circuit.

There are many other issues facing quantum computation such as crosstalk [76] and
efficient decompositions of to native gate sets [114]. But, all together, these restrictions
demonstrate a need to decrease the use of quantum resources when circuits are run. This
mainly includes reducing qubit lifetime and the number of gates used.
When we execute quantum circuits, we perform many runs of the same circuit. This is
done to build a meaningful distribution of results that can be drawn from. The entire circuit
is run, and measured, before being reset. This can be a time intensive process, and can incur
significant overhead if the reset process is significant.

2.3 The Quantum Compilation Problem

A compiler for a quantum program needs to be able to mold a circuit to fit the constraints
of an architecture while reducing the amount of quantum resources required. There have
been a number of optimizations developed that take advantage of commutation relationships,
gate cancellations and circuit synthesis to reduce gate count and circuit depth. These work
well for optimizing the algorithmic level, but do not necessarily work to fit a circuit to an
architecture. Instead, this work focuses on changing a circuit to fit the constraints of the
architecture, mainly matching connectivity in the most efficient way possible, and choosing
the most efficient decompositions of circuits.
The first step is mapping, where the program qubits must be assigned to hardware qubits
with the goal of minimizing the distance between interacting qubits over the course of the
program. As noted, most hardware only supports interactions between a limited set of qubits.
10
Qubits mapped too far from each other must be moved within interaction distance by in-
serting SWAPs or other communication operations before performing the original operation.
This communication is often very expensive and every extra gate needed for communication
contributes to the overall error rate of the final program. It is common for the mapping
and routing steps to occur in tandem as routing changes the mapping of the qubits over the
course of the program. Finally, scheduling consists of deciding when to execute which gates
and is usually dictated by factors such as run time or crosstalk where we may delay gates to
avoid crosstalk effects but possibly increase runtime.
The basic mapping and routing algorithms in this work are based heavily on previous
work from [77, 10]. There have been many different explorations into developing efficient
mapping and routing heuristics [118, 110, 119, 27, 51]. Several of these strategies were
developed with specific issues in mind. For example, Murali et. al. focused on crosstalk
mitigation in [79], and avoiding particularly noisy connections in [77]. The mapping and
routing compilation strategies laid out in this section focus on more general strategies that
reduce the number of SWAPs, which is applicable for any machine, rather than reducing
other sources of errors.

2.3.1 Decomposition

An important step in compilation is retargeting the gates used in the circuit to a gate set
native to the hardware. Different architectures natively support different sets of gates. As
long as the native gate set is universal, it is possible to fully translate any other unitary gate
into said native gate set. In many cases, this is sufficiently represented by a single one-qubit
gate that can be reparameterized by rotations around each axis and one two-qubit entangler
such as the CX gate [83]. This is the strategy used in this work, with some exceptions to
use native three-qubit entangling gates. Different native gate sets have varying entangling
capabilities, and there have been studies exploring how to best decompose circuits based on

11
the two or greater-qubit gates available within the context of compilation [86, 33].

2.3.2 Mapping for Quantum Architectures

We represent the underlying hardware topology as a graph, where nodes are hardware qubits
and edges are between qubits which can interact on the device. In this model, each edge is
weighted the same way. That is, a gate performed along this edge has the same probability
of success.
For most quantum programs, the entire control flow is known at compile time making
optimal solutions for mapping and routing possible but exponentially difficult to find. Looka-
head strategies base mapping and routing on the sum of weighted future interactions with
operations further into the future weighted less. The entire circuit is mapped and routed
in steps. At each step, we consider the weighted interaction graph where nodes are program
qubits and edges between nodes are weighted by the lookahead function:

X
w(u, v) = f (t − tc )
t∈o(u,v)

where w(u, v) is the weight between program qubits u and v, t is a layer of the program, and
tc is the current layer, i.e. the frontier of the program DAG. When considering a multiqubit
gate we add this weighting function between all pairs of qubits in the gate. The function f (t)
is an arbitrary, usually monotonically decreasing, function. For example, f (t) = 1/(1 + t) or
f (t) = e−t are usually used as a lookahead function.
For the initial mapping, we begin by placing the qubits with the greatest interaction
weight in the weighted interaction graph. We place these qubits adjacent in the center of
the device based on the hardware graph. For every subsequent logical qubit in this graph,
we consider all possible assignments to hardware qubits adjacent to the hardware locations

12
of already mapped logical qubits and choose the best based on a score:

X
s(u, h) = d(h, ϕ(v)) × w(u, v)
mapped v

where h is the potential hardware location, and ϕ is the mapping from program qubits to
hardware qubits. d is the physical distance between qubits on the hardware graph. The
goal is to place qubits which interact frequently close to each other in order to avoid extra
SWAPs during routing. We choose the hardware location h which minimizes this score. We
place qubits ordered by their weight to those previously mapped, greatest first. This process
continues until all the qubits are mapped.

2.3.3 Routing for Quantum Architectures

For routing and scheduling, we proceed through the circuit layer by layer, considering op-
erations in the frontier as potential gates to execute. We can only execute gates where the
qubits are adjacent together on the hardware interaction graph, we need to have a strategy
that can move qubits closer together without ruining the locality of the current mapping of
the qubits and future operation.
We use a metric called disruption. This balances finding the shortest path between two
qubits, without disrupting future operations. This leads to the following scoring function
for each neighboring physical qubit to the qubits involved in the operations that is strictly
closer to the other qubit in the operation:

X
D(u, j) = [d(ϕ(u), ϕ(v)) − d(j, ϕ(v))] × w(u, v)+
v

[d(j, ϕ(v)) − d(ϕ(u), ϕ(v))] × w(ϕ−1 (j), v)

13
In this case, j is the candidate swap location u is one logical qubit in the operation, and v
is the other. We perform this scoring for all qubits in the operation and pick the SWAP that
maximizes this function. In this function, moving further away from future interactions or
displacing the qubit in position j by moving it far from its future interactions is penalized.

14
CHAPTER 3

NEUTRAL ATOM ARCHITECTURES: ADAPTING TO

FLEXIBLE OPERATIONS AND FRAGILE QUBITS

In the past several years, many leading gate-based quantum computing technologies such
as trapped ions and superconducting qubits, have managed to build small-scale systems
containing on the order of tens of qubits [120, 91, 56, 107]. However, each of these systems
have unique scalability challenges [61, 21]. For example, IBM devices have continued to grow
in size while error rates have remained high, larger than what is needed for quantum error
correction [45]. Trapped ion machines, despite many promising results, have fundamental
challenges in controllability [78]. It is unclear whether any of these platforms in present form
will be capable of executing large-scale quantum computation needed for algorithms with
quantum speedup like Grover’s Search [102] or Shor’s Prime Factorization [46].
An alternative approach is to consider emerging quantum technologies and new archi-
tectures. In this chapter, we explore hardware composed of arrays of individually trapped,
ultra-cold neutral atoms which have shown great promise and have unique properties which
make them appealing from a software standpoint [98]. These properties include potential
for high-fidelity quantum gates, indistinguishable qubits that enable scaling to many qubits,
long-range qubit interactions that approximate global (or generally high) connectivity, and
the potential to perform multiqubit (≥ 3 operands) operations without expensive decompo-
sitions to native gates [67].
Neutral-atom (NA) architectures also face unique challenges. Long-range interactions
induce zones of restriction around the operating qubits which prevent simultaneous opera-
tions on qubits in these zones. Most importantly, the atoms in a neutral atom device can be
lost via random processes during and between computation. In the worst case, the compiled
program no longer fits on the now sparser grid of qubits, requiring a reload of the entire array
every cycle. This is a costly operation to repeat for thousands of trials. Coping with this loss
15
(a) Restriction Zone (b) Max Distance 3 (c) Atom Loss

Figure 3.1: Examples of interactions on a neutral atom device. (a) Interactions of various
distances are permitted up to a maximum. Gates can occur in parallel if their zones do
not intersect. The interaction marked with green checks can occur in parallel with the
middle interaction. (b) The maximum interaction distance specifies which physical qubits
can interact. Compiler strategies suited for this variable distance are needed for neutral
atom architectures. (c) Neutral atom systems are prone to sporadic atom loss. Efficient
adaptation to this loss reduces computation overhead.

in a time-efficient manner is important to minimizing the run time of input programs while
not dramatically increasing either gate count or depth, both of which will reduce program
success rate. In Figure 3.1 we show a small piece of a neutral atom system with many gates
of various sizes and distances being executed in parallel. Restriction zones are highlighted
and importantly no pair of gates have intersecting restriction zones. When atoms are lost
during computation, rather than a uniform grid we have a much sparser graph and qubits
will be further apart on average. A key to the success of a neutral atom system is resilience
to loss of atoms, avoiding expensive reloads.
In this chapter, we first explore the trade-offs in neutral atom architectures to assess
both current viability and near future prospects. We extend current compilation methods,
particularly mapping and routing, to directly account for the unique NA constraints like
long-range interactions, areas of restriction, and native implementation of multiqubit gates.
We then propose several coping strategies at the hardware and software level to adapt to
loss of atoms during program execution and we evaluate the tradeoff of execution time and
resilience to atom loss versus program success rate. We then extend the more basic atom loss
16
recovery strategies to make more full use of an architecture and further reduce the execution
time while more effectively maintaining program success rate.

3.1 Background: Neutral Atoms

We first briefly introduce some background on the underlying neutral atom technology. A
nice introduction can be found in [48]. Rather than being fabricated on a physical chip,
atoms in a NA system are trapped via reconfigurable, optical tweezer arrays. These atoms
can be arranged in one, two, or even three dimensions [34, 12, 59, 13]. In this chapter
we consider 2D configurations of these atoms, but arbitrary arrangements of atoms are
possible. Historically, one of the major difficulties with scalable neutral atom systems was
the probabilistic nature of atom trapping, but this challenge has since been overcome and
defect-free arrays of more than 100 atoms [84] have been demonstrated, and more recently,
dual species arrays of more than 512 atoms have been demonstrated as well [104]. However,
while neutral atom devices are able to trap and interact with several hundreds of qubits, the
error rates of demonstrated gates is much lower than that of superconducting architectures.
One-qubit gates have a success rate of 99%, as compared to 99.9%, and two-qubit gates have
a success rate of 96.5%, as compared to 99%.
The single atom qubit states can be manipulated using Raman transitions which im-
plement single qubit gates. In order to execute gates between qubits, atoms are optically
coupled to highly excited Rydberg states leading to a strong dipole-dipole interaction be-
tween the atoms [54]. These Rydberg interactions enable multiple atoms at once to interact
strongly and are used to realize multiqubit gates with 87.5% fidelity [67], something that is
not very easily achieved in other devices.
This method of physical quantum computation lends itself to unique capabilities beyond
arbitrary configurations. The main benefit is that due to the long-range of these interactions,
gates between qubits which are not directly adjacent in the atom array are feasible. However,

17
these longer distance interactions induce a zone of restriction as a function of the distance.
As the distance between qubits in an interaction increases, so does the area of restrictions
around both qubits in the interaction. This zone of restriction constrains which qubits can
be acted on at the same time. Two gates can only occur in parallel if their restriction zones
do not overlap.
The strategy has its own unique downsides as well. The weaker optical tweezers can “lose
hold” on the atoms, which are subsequently lost from the array. When this occurs, that
particular run of the circuit is invalid, and the entire array must be reset via a detection
and reloading process. More details of this process will be discussed in Section 3.4 Since the
loading of qubits into the array is relatively slow, on the order of one second, compared to
program execution which usually takes milliseconds, this is a barrier to scalability for neutral
atom devices.

3.2 Compiling for Neutral Atom Architectures

3.2.1 Adapting Mapping and Routing

The general compilation strategies laid out in 2.3 could easily be used on a neutral atom
device. If we only used adjacent atom connectivity on the hardware graph, we would likely
not run into the issue of conflicting parallel operations. However, this would not be taking full
advantage of the long-distance capabilities of a neutral atom device. This section specifically
adjusts the general mapping and routing algorithms laid out previously.
In a neutral atom device, we are not constricted by physical connections on a fabricated
chip as we are on superconducting devices. Instead, we are constrained by a maximum
interaction distance, or dmax , which is the furthest physical distance, dp (u, v) that two qubits
u, v can be from each other and still interact with one another. Now, our interaction graph
of hardware qubits contains an edge between nodes u, v if d(u, v) ≤ dmax . This breaks

18
Figure 3.2: An example of how SWAP paths of the same length can result in different
interaction distances from other qubits on the architecture and how not all SWAPs will have
the same effect on locality as on hardware where the adjacency graph is the interaction
graph.

the original assumptions of our interaction graph introduced previously, where a connection
implied physical adjacency and intractability. Here, the physical distance could be much
longer, affecting the area of restriction with parallel operations, even though the interaction
cost is the same. One choice of SWAP can be significantly different to another due to the
increased range of operations, as seen in Figure 3.2 where the qubits could be moved further
away from neighboring qubits, even if the SWAP paths are the same length.
For mapping, we can simply use the new interaction-graph in place of the original adja-
cency based interaction graph. The logic in this case remains the same. However, it is not
necessarily the case for the router. Here, we must reconcile the difference of physical distance
versus the interaction distance. We still explore the neighbors of the current physical qubit in
the interaction graph that move the qubits strictly closer, but we score each candidate SWAP
differently. Due to the increased number of paths between qubits, in place of the original
distance function, we use a normalized hardware difference: dnorm (p, q) = bdh (p, q)/dmax c
where dh . This let’s estimate the number of SWAPs required to interact two qubits with
one another. We also add an additional variable to our maximization function by adding the
distance between the current hardware qubit and the candidate SWAP qubit and the dis-

19
ruption function: S(q, p) = dnorm (q, p) + D(q, p). This weights the distance moved towards
the target, as well as considering the improvement, or detriment, to locality that the change
causes.
This consideration does not cover the issue of serialization due to the restriction zones
induced by longer interaction distances. Due to the non-local effects of different interactions,
this needs to be considered differently than the general algorithm. In general, the compiler
attempts to use an “As Soon As Possible” scheduling strategy. This strategy first ensures
that any operations that were deferred previously are scheduled first. We then examine the
frontier of potentially executable operations and schedule as many in parallel as possible.
That is, as many operations that do not overlap with other areas of restriction and the qubits
are within range of one another. We then perform a round of SWAP operations that do not
conflict with one another that move the qubits closer.
As neutral atom technology is also able to perform native three-qubit gates, we can
make a simple change that generalizes these algorithms adaptable to many qubits. When
determining which candidate SWAP qubits to check, we only explore those that reduce the
sum of the distances to the other qubits. This ensures that we consistently move the locality
of the qubits to one another.

3.2.2 Experimental Evaluation

We now explore how these compilations changes affect the resulting circuits, and how well
we can balance the benefits of Neutral Atom architectures.

Benchmarks

For this work, we have chosen a set of quantum programs which are parametrized, the input
size can be specified, to allow us to study how the advantages and disadvantages of a NA
system change as the program size increases. Specifically, we study Bernstein-Vazirani [14],

20
a common quantum benchmark, with the all 1s oracle to maximize gates, Cuccaro Adder
[28], a ripple carry adder with no parallelism, the CNU gate [11], a logarithmic depth and
highly parallel decomposition of a very common subcircuit, QFT Adder [97], a circuit with
two QFT components and a highly parallel addition component, and QAOA for MAX-CUT
[35], a promising near-term algorithm, on random graphs with a fixed edge density of 0.1.

Experimental Setup

For most experiments, we compile our benchmarks, with sizes up to 100, on a 10 × 10


NA device. We have a fixed radius of restriction but vary max interaction distance from
1 (emulating superconducting systems) up to the maximum needed for global connectivity
(here hypot(9, 9) ≈ 13). In relevant benchmarks we compile with decomposed multiqubit
gates and without. All experiments were run on a machine using Python 3.7 [113], Intel(R)
Xeon(R) Silver 4110 2.10GHz, 132 GB of RAM, on Ubuntu 16.04 LTS. All plot error bars
show ±1 standard deviation.

3.2.3 Evaluation Results

In this section, we explore promising architectural advantages provided by the neutral atom
technology. We examine long-range interactions where atoms distant on the device can
interact similar to a device with high connectivity. However, the cost of this longer-range
interaction is higher serialization due to the proportional increase in restricted area. Second,
we explore the native execution of multiqubit gates on the NA platform. Since neutral atom
technology is still in its early stages, it can be unfair to compare expected program success
rates from current gate error rates and coherence times. We analyze common metrics, gate
count and depth, which are good predictors of a program’s success rate if executed.
Trapped ion and superconducting architectures currently support qubit interaction only
between adjacent qubits. In SC systems this usually corresponds to a 2D grid or some other

21
sparse connectivity, where each qubit is able to interact with a small number of qubits.
One of the important promises of trapped ions is all-to-all connectivity where each qubit can
interact freely with any other qubit in the same trap. However, each trap is currently limited
by the number of ions it can support and expensive interactions across different traps.
In NA architectures, the connectivity lies somewhere between these two extremes. The
atoms, while often arranged in a 2D grid, have all-to-all connectivity beyond immediate
neighbors, i.e. within a fixed radius. This radius is dictated by the capabilities of the hard-
ware and can theoretically reach as large as the device. However, current demonstrations
have been more limited, for example up to distance 4. In this work, our experiments an-
alyze the full sweep of interaction distances to understand the importance of long-range
interactions to optimizing program success rate predictors.
Long-range interactions in NA are not free, we define an area of restriction imposed by
interacting qubits at a distance d from each other, f (d). Specifically, given this interaction
distance between qubits of the set Q all other qubits q 6∈ Q with distance less than f (d) to
any of the interacting qubits cannot be operated on in parallel. Furthermore, suppose we
have two operations to be performed in parallel. These two operations can only execute in
parallel if their areas of restriction do not overlap. For experiments in this work we explore
the function f (d) = d/2. Intuitively, as this function becomes more restrictive, i.e. the areas
surrounding the interacting qubits get larger, fewer total operations will be executable in
parallel, affecting the total execution time of the program.
Long-range interactions are important for reducing the total number of gates required
for execution on devices with relatively limited connectivity. Limited connectivity requires
compilers to add in many extra SWAP operations. The lower the connectivity, the greater
the average distance between qubits on the device therefore more SWAPs are required to
execute multiqubit gates between arbitrary sets of qubits. In Figure 3.3, we explore the gate
counts of compiled programs for various sizes over a range of maximum interaction distances

22
Gate Count Savings from Interaction Distance BV Gate Count
2 3 4 5 8 13
100% 1,200

post-compilation gate count


99
reduction in gate count

1,000
80% 87
800 75
60% 63
51
600 39
40% 27
400 15
3
20% 200

0% 0
BV CNU Cuccaro QFT-Adder QAOA 1 2 3 4 5 6 7 8 9 10 11 12 13
maximum interaction distance

Figure 3.3: Post compilation gate count across benchmarks. On the left are percent savings
over the distance 1 baseline averaged over program sizes up to 100 qubits. Each color is a max
interaction distance. Noticeably, there is less additional improvement as the MID increases,
indicating most benefit is gained for smaller distances. On the right is a sample benchmark
(holds in general) with many program sizes compiled for the whole range of MIDs. As the
program size increases, larger MIDs show benefits before flattening off.

up to the largest possible distance supported on the device. In each of these experiments,
all programs are compiled to 1 and 2 qubit gates only. Intuitively, we might assume having
a larger maximum interaction distance will necessarily be better than a smaller one since
it emulates global connectivity therefore not requiring any additional SWAP operations. In
general, we find the most benefit in the first few improvements in max interaction distance
with more relative gain for larger programs. The reduction in gate count is due solely to a
reduction in total SWAPs.
Importantly, the benefit obtained from increasing max interaction distance tapers off
with vanishing benefit. The rightmost points in these figures correspond to an interaction
distance the full width of the device, providing all-to-all connectivity. At this distance no
additional SWAP gates are required, so this is the minimum possible number of gates to
execute the input program. This distance is not required to obtain the minimum (or near
the minimum). In fact, a smaller interaction distance is sufficient. This is promising in cases
where large interaction distances cannot be obtained and hardware engineers can focus on
building higher fidelity short to mid-range interactions. For larger devices, the curves will be
similar, however, requiring increasingly larger interaction distances to obtain the minimum.

23
The shape of the curve will be more elongated, related directly to the average distance
between qubits.
A similar trend exists for circuit depth as seen in Figure 3.4. As interaction distance in-
creases the depth tends to decrease, with the most benefit found in larger programs. Again,
the rate of benefit declines quickly. We expected that as the interaction distance increased,
the depth would decrease initially, then increase again due to restriction zones proportional
to the interaction distance. As the maximum allowed distance increases, the average size of
these zones will increase, limiting parallelism. However, there are several important factors
diminishing the presence of this effect. First, SWAPs are a dominant cost in both gate count
and depth, often occurring on the critical path. Therefore, reducing the need for communi-
cation typically corresponds to a decrease in depth. Second, many quantum programs are
not especially parallel and often do not contain many other gates which need to be executed
at the same time, limiting the potential for conflicting restriction zones. In our set of bench-
marks, the circuits with high initial parallelism like CNU and QFT-Adder (a long stretch of
parallel gates in the middle) do show increases in depth with increased interaction size but
are not especially dramatic. In cases where gate error is dominant over coherence times, the
reduction in gate count far outweighs the induced cost of greater depth or run time.
This isn’t to say there is no cost from the presence of a restriction zone. In Figure
3.5 we analyze the relative cost of the restriction zones. In this set of experiments the
program is compiled with the same maximum interaction distance. In the ideal case it is
compiled with no restriction zones, resembling other architectures which permit simultaneous
interactions on any set of mutually disjoint pairs. These two circuits have the same number
of gates, including SWAPs. When no parallelism is lost, either from the original circuit or
from parallelized communication, these lines are close. A large gap indicates the increased
interaction distance causes serialization of gates. One additional side effect, which we do not
model here due to complexity of simulation, is the effect of crosstalk. By limiting which qubits

24
Depth Savings from Interaction Distance QFT-Adder Depth
2 3 4 5 8 13
100% 4,000

post-compilation depth
80% 66
reduction in depth

3,000
58
50
60% 42
2,000 34
26
40%
18
1,000 10
20%

0% 0
BV CNU Cuccaro QFT-Adder QAOA 1 2 3 4 5 6 7 8 9 10 11 12 13
maximum interaction distance

Figure 3.4: Post Compilation depth across all benchmarks. On the left, the reduction in
depth over the distance 1 baseline. Each bar is the average over all benchmark sizes. On the
right we see a similar drop off in post-compilation depth for the QFT-Adder. We’ve chosen
this specific benchmark to highlight the effect of restriction zones. Here we show a subset of
all sizes run. Depth initially drops but for larger interaction distances some of this benefit
is lost. We expect this to be more dramatic for even larger programs.

can interact in parallel we can effectively minimize the effects of crosstalk implicitly. This
can be made more explicit by artificially extending the restriction zone to reduce crosstalk
error by increasing serialization.
Long-range interactions are not the only unique property of NA architectures. One
of the important promises of NA hardware is the ability to interact multiple qubits and
execute complex instructions natively. For example, gates like the three qubit Toffoli could
be executed in a single step. This is important for several reasons.
First, it doesn’t require expensive decompositions to one- and two-qubit gates. Gates
like the generalized Toffoli have expensive decompositions, transforming compact complex
instructions into long strings of gates before SWAPs or other communication is added. The
base, 3 qubit Toffoli itself requires 6 two qubit gates and interactions between every pair
of qubits. If all these Toffoli gates could be executed natively without decomposition this
saves up to 6x in gate count alone. Toffoli gates are fairly common in quantum algorithms
extended from classical algorithms like arithmetic since they simulate logical ANDs and ORs.
If even larger gates are supported, this improvement will be even larger.
Second, efficient decomposition of multiqubit gates often requires large numbers of extra

25
Depth Increase due to Gate Serialization QAOA Depth
350% 2 3 4 5 8 13
500
300%

post-compilation depth
250% 400 50
increase in depth

50
40
200% 300 40
30
150% 30
200
20
100% 20
100
50%

0% 0
BV CNU Cuccaro QFT-Adder QAOA 1 2 3 4 5 6 7 8 9 10 11 12 13
maximum interaction distance

Figure 3.5: The induced restriction zone from interaction distance increases serialization.
In the prior results this is hard to discern because compared to low interaction distance
the amount of gate savings translates to depth reduction. Here we compare benchmarks
compiled with our restriction zone and a program with no restriction zone, to mimic an
ideal, highly parallel execution. The existence of a restriction zone has the most effect on
programs which are parallel to begin with. On the right we directly compare this effect on
the QAOA benchmark; solid line is compiled with realistic restriction zone and dashed is
ideal. The separation between the corresponding lines signifies the effect of the restriction
zone.

ancilla qubits. For example, in our CNU benchmark we use the logarithmic depth decomposi-
tion which requires O(n) ancilla, where n is the number of controls. When complex gates are
executable natively, additional qubits are typically not needed, reducing space requirements
for efficient implementation of gates.
In the NA architecture, execution of these gates does come with some constraints. For
example, to execute a 3 qubit Toffoli gate, each interacting qubit needs to be less than the
maximum interacting distance to every other interacting qubit. Therefore, with only an
interaction distance of 1 it is impossible to execute these gates and instead they must be
decomposed and for larger gates more qubits will need to be brought into proximity. While
not explored explicitly in this work, larger control gates will require increasingly larger
interaction distances. In general, the more qubits interacting, the larger the restriction zone,
increasing serialization if the qubits are too spread out.
Our set of benchmarks contains two circuits written explicitly in terms of Toffoli gates:
CNU and Cuccaro. In Figure 3.6 we analyze the effect of native implementation of these gates
rather than decomposition. The benefit is substantial in both cases requiring many fewer
26
CNU Gate Count Cuccaro Gate Count CNU Depth Cuccaro Depth
2,500 3,000
1,600
400
post-compilation gate count

2,000 2,500 1,400

post-compilation depth
91 94 91 1,200 94
2,000 300
1,500 91 94 91 1,000 94
59 54 59 54
59 1,500 54 59 800 54
200
1,000 19 14 19 14
19 1,000 14 19 600 14

100 400
500
500
200
0 0 0 0
1 5 10 13 1 5 10 13 1 5 10 13 1 5 10 13
maximum interaction distance maximum interaction distance maximum interaction distance maximum interaction distance

Figure 3.6: Compiling programs directly to three qubit gates reduces both gate count and
depth. Here we highlight a serial and parallel application written to three qubit gates. Here
dashed lines are compiled to two qubit gates decomposing all Toffoli gates before mapping and
routing. Solid lines compile with native Toffoli gates. With native implementation of three
qubit gates we obtain huge reductions in both depth and gate count for both benchmarks.

gates across all maximum interaction distances. While these gates have been demonstrated,
their fidelity is much lower than the demonstrated fidelity of two qubit gates. However, a
simple estimation given by the product of the gate errors in the decomposition shows the
fidelity of the Toffoli gate is greater than that of the decomposition. We give a more precise
analysis of this effect in the next section.
Both long-range interactions and native implementation of multiqubit gates prove to
be very advantageous, though the benefit is tempered by a distance-dependent area of re-
striction which serializes communication and computation. The importance of these effects
is input dependent. Programs written without multiqubit gates cannot take advantage of
native implementation. Programs which are inherently serial are less affected by large re-
striction zones at long interaction distances. One of the most important observations is
that excessively long interaction distances are not required and most benefit is obtained in
the first few increases. However, as the input program size increases for larger hardware,
we expect more benefit to be gained from long interaction distances. This trend is evident
here where small programs have almost no benefit from increasing distance 2 to 3 but large
27
programs nearing the device size see much more.

3.3 Error on Neutral Atom Devices

In the previous section we explored the effect on several key circuit parameters like gate
count, depth, and parallelism. These metrics are often good indicators for the success rate
of programs on near and intermediate term quantum devices where gate error is relatively
high and coherence times relatively low. In the case where gate errors and coherence times
are uniform across the device and comparable between technologies, these parameters are
sufficient for determining the advantage of one architecture over another. However, current
quantum technology is still in development with some more mature than others. For example,
superconducting systems and trapped ion devices have a several year head start over recently
emerging neutral atoms.
Consequently, physical properties like gate errors and coherence times are lagging a few
years behind their counterparts. It is critical however to evaluate new technologies early
and often to determine practical advantages at the systems level and to understand if the
unique properties offered by a new technology are able to catapult the co-designed archi-
tecture ahead of its competitors. In this section, we evaluate the predicted success rate
of programs compiled to a uniform piece of hardware with currently demonstrated error
rates and coherence times. It is important to note the gate fidelities and T1 times used as
a starting point are often measured from small systems as no large scale NA architecture
has been engineered to date. The average error, or T1 , across the hardware may have vari-
ance, as demonstrated in other publicly available technologies, though neutral atoms promise
uniformity and indistinguishability in their qubits, similar to trapped ions.
Simulating a general quantum system incurs exponential cost with the size of the system.
It is impractical to model all sources of errors during computation and simplified models are
typically used to predict program success rate. Here we compute the probability a program

28
will succeed to be the probability that no gate errors happen times the probability that
no decoherence errors occur. If pgate,i is the probability an i-qubit gate succeeds and ni is
Q ni
the number of i-qubit gates then the probability no gate error occurs is given by i pgate,i .
Here we consider circuits with up to i = 3. For neutral atoms, we consider two different
coherence times for the ground state and excited state i.e. T1,g , T1,e and T2,g , T2,e where
the ground state coherence times are often much longer than excited state coherence times.
Qubits exist in the excited state when they are participating in multiqubit interactions only.
The probability coherence errors occur is given as e−∆g /T1,g −∆g /T2,g −∆e /T1,e −∆e /T2,e where
∆g , ∆e are the durations spent in the ground and excited states, respectively. Often, gate
fidelities already include the effects of T1 and T2 , i.e. pgate,i includes coherence error. There-
fore, we will consider the probability of no coherence error as e−∆g /T1,g −∆g /T2,g only. These
simplifications serve as an upper bound approximation on the success rate of a program.
In this section we compare against superconducting systems, specifically, using error
values available via IBM for their Rome device, accessed on 11/19/2020. While we directly
compare using the same simulation techniques, we want to emphasize the purpose of these
figures is to indicate the value gained from decreasing gate count and reducing depth relative
to other available technology and to suggest that these improvements help overcome current
gate error in neutral atom technology. These are not meant to suggest neutral atoms in their
current stages are superior to superconducting qubits.
Our simulation results are across a large sweep of error rates from an order of magni-
tude worse to many orders of magnitude better, where error rates are expected to progress
in order to make error correction feasible. The point is to evaluate different technologies
with comparable error rates, how much is saved by architectural differences rather than the
current status of hardware error rates, especially when neutral atoms are years behind in
development.
In Figure 3.7 we analyze the potential of NA architectures on three representative bench-

29
marks. Here we sweep across various physical error rates and extract the predicted error
rate with those parameters; lower is better. In both CNU and Cuccaro we permit 3 qubit
gates while the others contain only one- and two- qubit gates. The superconducting curves
correspond to similar simulations using error rates and coherence times provided by IBM.
At lower physical error rates, we expect all architectures to perform well, at virtually no
program error rate. At this limit, the hardware is below the threshold for error correction.
On the other hand, in the limit of high physical error rates, we expect no program to succeed
with any likelihood and to produce random results. For near- and intermediate-term quan-
tum computation, the regions between the limits are most important and the divergence
from this all-noise outcome determines how quickly a device becomes viable for practical
computation. For comparable error rates between superconducting and NA architectures,
we see great advantage obtained via long-range interactions and native multiqubit gates,
diverging more quickly than the limited connectivity SC devices.
Alternatively, we might ask what physical error rates are needed to run programs of a
given size with probability of success above some threshold. In Figure 3.8, we consider this
question for a threshold success rate of 2/3. Here we sweep across physical error rates and
compute the largest possible program of each benchmark we can successfully execute. There
are two interpretations. First, for a fixed physical error rate we can determine what size
program is likely to be successfully executed. Alternatively, suppose we have a program of
a given size we want to execute, we can then decide what physical error rate we need to run
that program successfully. For a fixed error rate, we find we can execute a larger program
or, equivalently, require worse physical error rates than a superconducting system to run a
desired program.

30
Success Rate Comparison

100
sample error rate

BV
CNU
10−1 Cuccaro
QFT-Adder
QAOA
SC device
NA device
Current SC
10−5 10−4 10−3 10−2 10−1
two-qubit gate error

Figure 3.7: Program success rate as a function of two-qubit error rate. Because current NA
error rates are lagging behind competitive technologies we scan over a range of two-qubit
error rates for each of the benchmarks all on 50 qubit programs (49 for CNU) with max
interaction distance of 3. Examining pairs of solid and dashed lines we can compare NA to
SC. In the limit of very low two qubit error rate, systems can support error correction. Both
SC and NA systems scale at roughly the same rate (slope of the line) but the NA system
diverges from the completely random outcome at higher error, allowing us to run programs
on the hardware much sooner. The further to the lower right a point is, the higher fidelity
circuit it can perform with worse error rate.

31
Program Size Comparison BV
100 CNU
Cuccaro
QFT-Adder
QAOA
80 SC device
NA device
largest runnable size

Current SC
60

40

20

1
10−5 10−4 10−3 10−2 10−1
two-qubit gate error

Figure 3.8: Another way to examine the data of Figure 3.7 is to ask, given a desired program
success rate, what the required two qubit error rate is. Here we sweep again over two qubit
error rates and record the maximum program size to run with success probability greater
than 2/3. Again, examining pairs of solid and dashed lines we can compare NA to SC. With
the reduced gate counts and depth we expect to be able to run larger programs sooner.

32
3.4 Adapting to Atom Loss

So far, we’ve focused primarily on properties of a neutral atom system that are usually
advantageous and have analyzed that while there are tradeoffs, the gates and depth saved
drastically outweighs any cost. However, neutral atom systems are not without limitation.
The atoms in a NA system are trapped using optical dipole traps such as optical tweezers.
While this technique offers great flexibility in array geometries and the prospect of scaling
to large atom counts, the trapping potential per atom is weak when compared to trapped
ion architectures. Consequently, neutral atoms can be lost more easily during or between
computations forming a sparser grid. Fortunately, this loss can be detected via fluorescence
imaging and remedied with various software and hardware approaches with different overhead
costs analyzed here.
Atom loss has critical implications on program execution. For near-term computation, we
run programs hundreds or thousands of times to obtain a distribution of answers. Consider
running a program, after each trial we evaluate whether atoms have been lost. If an atom
used by the computation has been lost, then we have an incomplete answer. Upon a loss, we
have no way of knowing if it occurred in computation and must disregard the run from our
distribution and perform another shot. Furthermore, a program compiled for the original
grid of qubits may no longer be executable. The program must either be recompiled for the
now sparser grid of qubits, the array of atoms can be reloaded, or the compiled circuit can
be adapted to the atom loss. The first two solutions are costly in terms of overhead time.
The third may provide a faster alternative, and provide an opportunity to perform more
executions in the same time while maintaining a valid program.
We model atom loss from two processes. The first is based on vacuum limited lifetime
where there is a finite chance a background atom collides with the qubit atom held in the
optical tweezers displacing the qubit. We approximate this to occur with probability 0.0068
over the course of a program and is uniform across all qubits [26]. Loss during readout

33
is much more likely. In some systems readout occurs by ejecting atoms which are not in
a given state, resulting in about 50% atom loss every cycle [38]. This model is extremely
destructive and coping strategies are only effective if the program is much smaller than the
total size of the hardware. Alternative, potentially lossless techniques, have been proposed
for measurement, but are not perfect with loss approximately 2% [65] uniformly across all
measured atoms. Detecting atom loss is done via fluorescence which takes on the order of
6ms.
We lay out several coping mechanisms for atom loss and examine their effectiveness in
terms of overheads such as the time to perform array loads, qubit fluorescence, and potential
recompilation. In each of these experiments we assume the input program is smaller than the
total number of hardware qubits in the original grid, otherwise any loss of an atom requires
a reload. These additional unused qubits after initial compilation are considered spares,
borrowing from classical work on DRAM sparing [81]. Currently, no program that executes
with reasonably high success will use the entire grid and these spares will come at no cost.
Potentially, as many atom losses can be sustained as the number of spares. However, an
array reload is always possible if there is an unlucky set of holes or no more spares. Below we
detail various strategies. These strategies work to adapt the resulting circuit from mapping
and routing, using the features of neutral atom architectures to adjust the normal runtime
strategy of a quantum circuit.

3.4.1 Recovery Strategies

We have several different strategies we can pursue to recover from an incurred atom loss.
The first set of strategies are all performed on the originally compiled circuit, it does not try
to influence the original compilation.

• Always Reload. Every time an atom loss is detected for a qubit used by the compiled
program, we reload the entire array. This naive strategy is efficient when array reloads
34
qc qb qa qc qb qa qc qb qa

1 76 5 4

spare in use
2 3
(a) Circuit Before Loss (b) Virtual Remapping (c) Reroute

Figure 3.9: Examples of two different atom loss coping strategies. (a) shows the initial
configuration of three qubits, with the spare qubits in a light grey, and in use qubits black.
(b) Represents how the atoms are shifted into the spare qubits to accommodate a lost
atom under the virtual remapping strategy. Notice that the interaction is no longer within
interaction distance 1. (c) Demonstrates how rerouting strategies swaps a qubit to a valid
interaction configuration, then returns it. Numbers indicate the order of swaps.

are fast since only a single compilation step is needed.

• Always Full Recompile. When an interfering atom loss is detected, we update the
hardware topology accordingly and recompile the input program. This fails when the
topology becomes disconnected, requiring a reload.

• Virtual Remapping. NA architectures support long-range interactions up to a maxi-


mum interaction distance. Therefore, shifts in the qubit placement are only detrimental
to execution when the required qubit interactions exceed this distance. For this strat-
egy, we start with a virtual mapping of physical qubits to physical qubits where each
qubit maps to itself. When an atom is lost, we check if it is used in the program. If so,
we adjust the virtual mapping by shifting the qubits in a column or row of the qubit
in the cardinal direction with the most unused qubits starting from the lost atom to
the edge of the device. This process is shown in Figure 3.9b. In this figure, addressing
qb would now point to the original location of qc , and addressing qc would point to
the qubit to the left of qc ’s original location. If there are no spare qubits, we perform
a full reload. Otherwise, we execute the gates in order according to the mapping. If
two qubits which must interact are now too far apart, we reload. This strategy is
efficient in terms of overhead since this virtual remapping can be done on the order of

35
40 ns in hardware [29] via a lookup table. However, this strategy can be inefficient in
the number of reloads required since it is easy to exceed the interaction distance. We
later explore how many atom losses can be sustained before reload which allows us to
estimate how many reloads will be required on average.

• Minor Rerouting. Here we perform the same shifting strategy as in Virtual Remapping
to occupy available spares. However, rather than failing when distance between the
remapped qubits exceeds the maximum interaction distance, we attempt to find a
path over the usable qubits and insert SWAP gates along this path. To simplify
computation, we SWAP the qubits on the found path, execute the desired gate, then
reverse the process to maintain the expected mapping. The rerouting is shown in
Figure 3.9c. Too many additional SWAPs are detrimental to program success. We
may force a reload if the expected success rate drops, for example by half, from the
original program’s expected success rate.

• Compile to Smaller Than Max Interaction Distance. For most programs there are
diminishing returns to compiling to larger maximum interaction distances but can be
sensitive to atom loss. In this strategy we compile to an interaction distance less than
the max so when qubits get shifted away from each other it will take more shifts to
exceed the true maximum distance. The overhead is the same as Virtual Remapping
but the compiled program could be less efficient than one compiled to the true max
distance.

• Compile Small and Minor Reroute. This strategy is based on Compile to Smaller Than
Max Interaction Distance but performs the same rerouting strategy as Minor Rerouting
with similar overhead costs.

Each of the above strategies can be augmented by moving beyond the hardware adjacency
graphs. Moving along the adjacency graph can maintain locality of nearby qubits, but it
36
Max Hole Tolerance (CNU) Max Hole Tolerance (Cuccaro)
80%
virtual remapping reroute compile small c. small+reroute recompile im reroute c. small+im reroute

60%
# holes/device size

40%

20%

0%
2 3 4 5 6 2 3 4 5 6
maximum interaction distance maximum interaction distance

Figure 3.10: Atom loss as a percentage of total device size which can be sustained before
a reload of the array is needed. Each program is 30 qubits on a 100-qubit device. As the
interaction distance increases most strategies can sustain more atom loss. Strategies like
full recompilation can sustain large numbers of atom loss but as we will see are expensive
computationally. Fast, hardware solutions or hybrid solutions can sustain fewer numbers of
holes but have lower overhead. We show two representative benchmarks parallel vs. serial.
“im” represents the interaction graph methods.

perturbs many qubits as we perform different shifting strategies. We can instead shift the
qubits along the shortest path to an available qubit on the interaction graph. This moves
fewer qubits and can be more flexible. The interaction-based strategies are not bounded by
the “walls” of the architecture. A demonstration of this is shown in Figure 3.15c.

3.4.2 Experimental Evaluation

Excluding the first approach of reloading with any interfering atom loss, we examine how
many losses can be sustained without exceeding the constraints of the architecture in size,
dimension, or needed interaction distances. Figure 3.10 shows the maximum number of holes
supported by the different strategies for a 30 qubit Cuccaro adder and a 29 qubit CNU. The
entries for compile small and compile small + reroute are compiled to one less than the
maximum interaction distance. We do not compile to interaction distance 1, so we do not
have entries for these strategies at interaction distance 2.

37
Maximum Resistance to Loss

As would be expected, recompile is able to support the most lost atoms since the only
failure cases are disconnected hardware topology or fewer atoms than required qubits. In
fact, since our example circuits use 30% and 29% of the hardware, once the interaction
distance overcomes any disconnected pieces, recompiling can sustain 70% atom loss and is
the ideal case for how many atoms can be recovered by a given coping mechanism. The non-
rerouting strategies, while a fast solution since no paths need be found, offer limited atom
loss recovery since it is easy to move outside the maximum interaction distance. The simple
virtual remapping is only able to support a small amount of atom loss but does increase as
the max distance increases. As predicted, compiling to a smaller interaction distance does
enable more resilience to atom loss since more movement can be tolerated before exceeding
the maximum interaction distance. Both rerouting strategies have the disconnected topology
failure case but reach the additional failure case of not having space in any direction to shift
the qubits in the event of atom loss. This does not occur as often for the non-routing
mechanisms since we hit the barrier of exceeding the interaction distance first. Due to the
lack of space to move the qubits both of the rerouting strategies for hardware-based shifting
are only able to sustain 50% atom loss at higher interaction distances. Surprisingly, we
find that the interaction graph based shifting and rerouting strategies are comparable to
recompiling from the start. This is due to the fact that we are not limited by the boundaries
of the architecture in these strategies, the shifting strategy moves the qubits whereever
needed until either we reach the disconnected graph, or run out of space for the required
qubits. These are the same failure conditions of the recompilation case, so is able to recover
the same amount of lost atoms.

38
Success Rate

However, these different strategies add varying numbers of extra SWAPs to handle atom loss,
lowering the success rate. As more atoms are lost, more SWAPs are needed, and the rate
decreases as seen in Figure 3.11 for Cuccaro and CNU with the rerouting and recompiling
strategies. With current error rates, success is very low for 30 qubit circuits. To better
demonstrate how the success rate changes, we use lower error rates so that, without atom
loss, about 2/3s of shots succeed. For any strategy, as the interaction distance increases,
fewer SWAPs are needed so the shot success stays higher. Since the recompilation strategy is
able to schedule and map qubits with full knowledge of the current state, including missing
atoms, it can achieve the best routing and success rate out of all the atom loss strategies.
All of the rerouting strategies have lower rates since, as seen previously, they tend to add
more SWAPs per atom loss as compared to recompilation. This is even more pronounced
for the interaction graph based solutions, since the qubits are moved further away from
their original positions. But, since compiling to a smaller MID before rerouting means the
interaction distance is exceeded less often, these strategies require fewer SWAPs, boosting
its rate over simply rerouting.
However, when examining the decrease in success rate for interaction graph based shifts,
we find that the potential to move qubits much further away from their original positions is
detrimental to success rate as well. We are not just increasing the depth of the circuit due to
longer range interactions, extra communication is required, increasing the number of gates,
and decreasing the probability of success more than the hardware based solutions.

Overhead Time

Taking this into account, we examine the estimated overhead time of each strategy for 500
runs of a given circuit. We use a 2% chance of atom loss for a measured qubit, and a 0.68%
chance of atom loss due to atom collision in a vacuum. The overhead times for CNU are seen

39
Shot Success Rate Drop (CNU) Shot Success Rate Drop (Cuccaro)
reroute, MID: 2 3 5 c. small+im reroute, MID: 3 5 c. small+reroute, MID: 3 5
0.8 im reroute, MID: 2 3 5 recompile, MID: 2 3 5

0.6
0.7

0.6 0.5
estimated shot success

0.5
0.4

0.4
0.3

0.3

0.2
0.2

0.1
0.1

0 0
0 2 4 6 8 10 12 14 16 18 20 0 2 4 6 8 10 12 14 16 18 20
number of holes number of holes

Figure 3.11: For strategies which modify the program such as recompilation or rerouting
strategies, additional gates could be added leading to a lower overall success rate. Here we
trace the success rate of our three program modifying strategies. The full recompilation
strategy (circles) is a rough upper bound which best accounts for holes as they appear able
to move the entire program to a more appropriate location and route best. The gap between
strategies on the same MID gets smaller as the MID gets larger. Here we’ve chosen the
two-qubit error rate corresponding to approximate 0.6 success rate to begin with (based on
Figure 3.8) in order to best demonstrate the change in shot success probability over a range
of atom loss.

40
Overhead Time for 500 Shots (CNU)
260 always reload compile small c. small+reroute c. small+im reroute
virtual remapping reroute im reroute
240
220
200
180
overhead time (s)

160
140
120
100
80
60
40
20
0
2 3 4 5 6
maximum interaction distance

Figure 3.12: Strategies that are able to reduce the number of reloads necessary greatly
reduce the overhead time when running circuits. Here we show the overhead time for all
strategies except recompilation. The proportion of time dedicated to reloading is shown by
the dominant color in each bar, followed by fluorescence in red, and recompilation in black.
Any strategy whose overhead exceeds that of always reloading, such as full recompilation,
should not be considered.

in Figure 3.12. For any rerouting strategy that requires extra SWAPs, a reload is forced once
the number of added swaps would decrease the success rate by 50%. For a 96.5% successful
two-qubit gate, this would be six SWAPs.
Recompilation is not shown in Figure 3.12 as software compilation exceeds the array
reload time, and the overhead time is larger than simply reloading. Other strategies are
always more time efficient than reloading all of the atoms. Additionally, since compiling
to a small size requires fewer fixes with swaps, the overhead time tends to be smaller at
lower interaction distances. Essentially, the fewer times the strategy requires a reload the
faster the overhead time is. Subsequently, we see our interaction model rerouting strategies
performing less well as compared to the basic hardware based methods as the additional

41
(c) Scaled for clarity
35µs 6ms 20+61µs

(b) Atom-loss resilent execution


0.57 0.71

(a) Entire Trace


0 time (s) 1.92

compile run circuit fluorescence circuit fixup reload atoms

Figure 3.13: A timeline of 20 successful shots for Compile Small and Reroute with reload
time of 0.3 s and fluorescing time of 6 ms. A majority of the overhead time is contributed
by the reload time and fluorescence, indicating that the duration and count of these actions
is crucial to overall runtime.

swaps require more reloading more often. Additionally, as interaction distances increase, the
overhead time of each strategy converges. A sample timeline of 20 successful shots using
compile small and reroute can be seen in Figure 3.13. After initial compilation, reloading
takes a majority of the time, so any ability to reduce the number of reloads vastly reduces
the overall run time.

Scalability

Compile small + reroute is an efficient way to improve loss resilience, we next examine
the sensitivity of the successful shot count before a reload to the rate of atom loss for this
strategy. Figure 3.14 shows how the number of successful shots changes as the rate of atom
loss changes. A 10x improvement offers an expected 10x improvement in the number of
successful shots before a reload must occur. This is because the rate of atom loss decreases
as technology improves, reducing the number of reloads, improving overhead time.

42
Successful Shot Rate Sensitivity (CNU)

102

# successful shots before reload


101

100

10−1

10−2 6
5
10−3 4
3
10−1 100 101
atom loss rate improvement factor

Figure 3.14: Sensitivity to the rate of atom loss for the balanced Compile Small and Reroute
strategy. In prior experiments we used a fixed rate of 2% atom loss. For larger systems this
rate could be worse and in the future we might expect this rate to be much better. For each
interaction distance we see as the rate of atom loss gets better we can run many more trials
before we must perform a reload and reset. Some error bars don’t show on the log axis.

3.5 Designing for Atom Loss

The previous strategies for atom loss only acted on the originally compiled circuit. In the
original compilation the mapping and routing strategies had access to the entirety of the
architecture. Here, we constrain the mapping and routing problem to specific sections of
the architecture in order to further increase the usefulness of the device before performing a
reload of the neutral atom array.

3.5.1 Constraining Methods

• Focused Use and Migration. In the original compiler, a circuit could be mapped and
routed onto any qubit in the architecture. This gives global scope, and finds good
mapping and routing solutions, but can lead to less dense mapping across the architec-
ture. Instead, we define a bounding box big enough to hold the circuit, seen in Figure
3.15d, where for a six qubit circuit, we define a 2 by 3 bounding box. This can be done

43
b a b c a b c a1 b 1 c 1
a b a d a b d e f d e f d1 e1 f1
c d f c f c f
e e e d a2 b 2 c 2
d2 e 2 f 2
(a) Initial Mapping (b) Hardware (c) Interaction (d) Focused Use (e) Focused Use (f) Full Parallelism
Compilation Remapping Remapping Bounding Post Initial Mapping
Boxes Relocation

Figure 3.15: Examples of four different atom loss coping strategies. (a) Shows the initial
configuration of qubits using the original compiler that makes use of the entirety of the
architecture. (b) shows the result of using the hardware graph to shift the qubits away
from a newly lost atom. (c) shows how the interaction graph could make a different decision,
taking the shortest interaction path to an open qubit. (d) demonstrates how the architecture
can be divided such that we have multiple opportunities to start in a section with fewer lost
qubits. (e) shows how a wear leveling approach would appear after one remapping of the
circuit. (f) shows how these divided sections can be used to exploit some level of parallelism
in the architecture.

in two ways: a loose or tight configuration. The former defines both s by the ceiling
of the square root of the number of qubits. The latter defines one dimension by the
square root, and the second by the number of qubits divided by the square root. This
box is then tiled across the architecture. In the event that the bounding box does not
neatly fit, some overlap is allowed. The circuit is mapped and routed entirely within
one of these sections. When atoms are lost, we continue using any recovery method
previously defined. When the recovery method fails, or the estimated probability of
success falls below a certain threshold, rather than resetting the array, we directly
remap the circuit to a new section of the architecture, accounting for any previously
lost atoms. The process after one relocation is shown in Figure 3.15e. Since the atoms
in the new section have not been used for computation, fewer will have been lost. We
repeat this process until each section has been visited once. Then, the entire array is
reloaded before restarting the process.

• Parallel Executions. The non-overlapping sections defined for Focused Use can be
treated as if each is its own architecture. Instead of shifting the mapped circuit from
one defined region to the next, we can create a single aggregate circuit [30] that con-

44
tains multiple instances of the smaller circuit. This is similar to [93] where multiple
variational circuits have been mapped onto a single architecture to improve their per-
formance. We map the circuit onto the architecture multiple times, and run multiple
shots of the individual circuit in one run of the aggregate circuit. Then any individual
circuits that lost no computational atoms can be treated as a successful shot. If an
atom is lost, we can use any recovery methods previously explored until it fails. At
this point, the array is reloaded.

• Focused Use + Partial Parallelism. Rather than filling every section with qubits, we
only fill some of the sections. Then, when the recovery strategy fails, we use a different
set of the predefined sections. We are exploiting some amount of parallelism, but are
giving more resources for the more focused recovery strategies to make use of. This
process continues until each section has been used at least once. The entire array is
then reloaded.

3.5.2 Evaluation

We use the same two benchmarks used in the previous atom loss section, with the addition
of two more near term algorithms variational algorithms. These are the QAOA circuit,
described previously and the VQE circuit. VQE attempts to find the minimum eigenvalue
of a wave function encoded in parameterized rotation gates and entangled qubits [112]. It is
then run several times while tuning the rotation gates. We test one of these iterations. We
use linear entanglement, meaning that the first qubit targets the second, the second then
targets the third, and so on. This lends itself to very low density, highly serialized circuit.
We study these circuits at multiple sizes to understand how our recovery strategies scale as
well.
We also use the same atom loss simulation parameters including the 10 by 10 array of
qubits, and the same rates of loss of atoms from both an imperfect vacuum and measurement.
45
3.5.3 Results

We already understand how our original compiling-to-smaller interaction distance with vir-
tual remapping and rerouting fares best among our recovery with respect to recompilation
in maintaining fidelity and best in overhead time, and we will use it as our baseline strategy
to beat without new relocation and parallelism based strategies.
While shifting along the interaction graph provides benefits for larger circuits, we find
that focused use, and relocation to unmapped sections of the architectures can improve
probability of success across circuit size, Figures 3.16 and 3.17. Initially, as atoms are lost,
this strategy maintains the same probability of success as our baseline. This is expected
since we are mapping into a specific section of the architecture, we can still generate a similar
mapping and routing. However, after several atoms have been lost, we see an increase in
the average probability of success indicating a relocation to a new part of the architecture.
Since there have been fewer atoms lost in this new section there is less adjustment, increasing
the probability of success. Focused use and relocation is able to more closely adhere to the
probabilities of success achieved via recompilation, making it a more effective strategy from
this perspective. While we do not achieve strictly better probabilities of success in the 90
qubit circuits, we are able to maintain the same probabilities of success. We do not see the
same benefits since the tiles in this circuit will need to be quite large, at least 9 qubits by
10 qubits. This requires significant overlap and relocation will not have as great an effect.
Additionally, this strategy does not make use of the interaction graph remapping, and does
not benefit from the increased flexibility.
It should also be noted that we do not see a significant difference between the loose and
tight bounding boxes. Both follow the same trend line in each of the benchmarks, indicating
that the circuit is mapped and routed similarly in each case.
We should also note that remapping along the hardware array often fails for larger circuits.
For the 90 qubit circuits, Figure 3.17, we fail to find significant average probabilities of

46
Figure 3.16: Decreases in success rate for different mitigation strategies for 30 qubit circuits.
Each color is a different strategy, solid lines are circuits with interaction distance 3, and
dash lines are interaction distance 5. Run over 50 trials, each error bar represents one
standard deviation from the mean. We see decreases for all strategies, with full recompilation
able to maintain the highest probability of success and relocation better approximating
recompilation than our baseline strategy.

Figure 3.17: Decreases in success rate for different mitigation strategies for 90 qubit circuits.
We exclude QAOA from this test as the success rates are not meaningful at current error
rates. Each error bar represents one standard deviation from the mean over 50 trials. We see
the same patterns, with the exception that interaction remapping is often able to outperform
other strategies.

47
success for several circuits after the loss of more than one or two atoms when using our
original strategies. They do not provide enough flexibility to find available atoms at this
size, which is why some strategies cannot be seen on the graph. However, when we shift
along the interaction graph, we are able to achieve much higher fidelities. Particularly for
the Cuccaro Adder, we see that for a maximum interaction distance of 5, the interaction
model based approach achieves much higher probabilities of success. The non-recompilation
counterparts are not able to sustain a minimal amount of atom loss at this size of circuit.

Successful Shots before Reload

Another metric to consider is how many successful shots can be completed prior to reloading
the entire array. Every shot that can be performed without an additional reload reduces the
overall time to run a circuit. We analyze the average successful shots per reload cycle for
each mitigation method for each of our benchmarks.
There is a much more pronounced effect on average shots per reload when utilizing the
new relocation strategy over the rerouting strategy for both large and small circuits, Figures
3.19 and 3.18. We find a much higher rate of average shots per reload for relocation based
strategies. Once the array has been divided into several different sections, when the circuit
is relocated to a new section it will be mostly free of atom loss. Any atoms lost will be due
to computational atoms being remapped into that space, or the rare event of atoms lost to
environmental factors and will be much lower. All of which is conducive to much higher
probabilities of success and reduces reloads.
As the circuit size increases, we see diminishing returns of the relocation strategy. As
the circuit size increases, the number of distinct sections that can be laid out on the array
without overlap decreases. We do not have as many opportunities to perform a pseudo-reload
as the circuit uses more qubits. We can even quantify this relationship. In Figure 3.20 we
see the relationship between the increase in the number of shots against the number of times

48
Figure 3.18: The average number of shots per reload 10, 20 and 30 qubit benchmarks at
interaction distance four. Error bars indicate one standard deviation from the mean. Each
color is a different recovery strategy. We find that there is significant improvement from our
relocation strategies (red and green), greatly exceeding the baseline number of shots, more
closely matching recompilation.

Figure 3.19: The average number of shots per reload 50, 70 and 90 qubit benchmarks at
interaction distance four. Error bars indicate one standard deviation from the mean. Each
color is a different recovery strategy. We exclude QAOA since it is not a practical circuit at
this size.

49
Figure 3.20: The normalized advantage of the relocation strategy to rerouting and interaction
distance four. Each color represents a different circuit, and the advantage of relocation
decreases at a factor inversely proportional to the size of the circuit.

the circuit can be fit onto the architecture. For each circuit, we follow the same pattern.
The advantage in the number of shots of relocation to rerouting is roughly proportional to
the size of the circuit. For a 30 qubit circuit, this is up to 3.5x improvement. For a 10 qubit
circuit, this is up to a 8x improvement in average shots per reload. Empirical results do not
match the exact ratio since the bounding box tiles do not always fit neatly onto the array.
Additionally, these bounding boxes are often larger than the circuit, causing the advantage
to be lower.
While average shots per reload cycle is a good indicator for a strategy’s potential to
recover efficiently, it does not take into account the overhead time to determine the best
course of action. A strategy is only viable if it is faster than reloading the array. If this
cannot be achieved, it is more effective to reload the atoms as it will give us the highest
probability of success. Full recompilation cannot be considered for this reason. Previous
50
work found large gains over reloading via the compiling to a smaller interaction distance and
rerouting. We will be comparing against this strategy to determine effectiveness.
In Figure 3.21, we examine the effects of different mitigation strategies on the overhead
time to execute 500 shots for a 30 qubit Generalized Toffoli circuit. While these strategies
are dependent on the length of the circuit, it will scale similarly for each strategy. We
see significant reduction in time dedicated to reloading the entire array for our relocation
strategies and increases for interaction graph based strategies. However, the increases in
calculating the solutions for relocation do not outweigh the overhead time saved by relocation.
For interaction distance four, our relocation strategy outperforms the basic rerouting strategy
by 55% in reload times, and 45% overall.
As the maximum interaction distance increases, the margin of the overall time for each
strategy decreases. As the maximum interaction distance increases, we do not need to add
communication to recover from an incompatible circuit. This reduces the number of reloads
required for each recovery method. The time dedicated to fluorescence does not decrease, as
each shot, successful or unsuccessful, still requires a fluorescence.

Effects of Parallelism

The benefits of relocation could be improved upon by using multiple non-overlapping sections
of the device at the same time. In doing so, we use as much of the architecture as possible.
For example, using 90 qubits to run three concurrent 30 qubit circuits. Since we are using
almost all of the qubits, we will use the interaction graph based remapping as it has proven
to be more effective in situations where most of the architecture has been filled. We examine
the average probability of success for three concurrent 30 qubit circuits in Figure 3.22. This
strategy can only withstand 10 lost qubits at the most since we can only recover as many
qubits as those that have been unmapped. Any increased serialization from running circuits
in parallel does not substantially impact the decrease in probability of success as compared

51
Figure 3.21: Overhead times for different recovery strategies at different maximum inter-
action distances. The major color in each bar represents time dedicated to reloading. As
follows from the significant advantage seen previously, relocation is a very efficient strategy.

52
Figure 3.22: Decreases in success rate for different mitigation strategies for parallel 30 qubit
circuits. Run over 50 trials, each error bar represents one standard deviation from the mean.
In general, we see that for higher interaction distances, increased parallelism does not heavily
affect the rate of decrease in probability of success rate.

to previous mitigation strategies.


We also analyze the effects of this strategy on the number of average shots per reload
cycle in Figure 3.23. By packing as many instances of circuits as possible into the array,
there is no space to shift the qubits causing the recovery strategy to fail quickly and requiring
a reload. This frequency will increase the overhead time, indicating that full parallel usage
of a neutral atom architecture is not viable at current atom loss rates.

Parallelism with Relocation

Without extra atoms to recover from atom loss, full parallelism is not effective. For smaller
circuits, we do not need to fill the entire architecture. By choosing to only use a portion
of the previously defined bounding boxes, we can take advantage of partial parallelism. By
reducing parallelism, we have the opportunity to make use of relocation as well, potentially
adding to the overhead time gains we have already found. In Figure 3.23 we examine multiple
levels of parallelism for 10, 20 and 30 qubit circuits. The percentage of parallelism indicates
what percentage of the architecture we are using at a time, 30% parallelism means that three
instances of a 10 qubit circuit are being run at a time, with 3 different areas to relocate these

53
Figure 3.23: The average number of shots per reload for 10, 20 and 30 qubit benchmarks at
interaction distance four at different levels of parallelism. Error bars indicate one standard
deviation from the mean. Each color is a level of parallelism. Parallel indicates that as much
of the architecture is used as possible, and relocation means that only one instance is run at
a time.

54
Benchmark Baseline Relocation 2 Parallel % Fluorescence
Time (s) Time (s) Time (s) Decrease
Cuccaro-10 10.91 4.67 3.13 51.22
Cuccaro-20 13.82 9.58 8.24 54.02
Cuccaro-30 26.19 17.05 14.80 40.71
CNU-10 7.98 4.59 2.82 50.60
CNU-20 24.09 8.68 10.50 53.12
CNU-30 30.61 16.58 15.15 35.80
QAOA-10 5.83 4.41 2.66 51.05
QAOA-20 14.83 7.54 7.49 52.70
QAOA-30 42.67 16.43 18.68 35.48
Linear VQE-10 6.26 4.16 2.42 50.83
Linear VQE-20 8.77 6.61 5.01 53.75
Linear VQE-30 15.93 9.79 9.56 36.12

Table 3.1: Full Runtimes with Two Instances of Parallelism

three 10 qubit circuits. We can achieve similar shots per reload to the relocation strategy
when running up to four instances of a 10 qubit circuit, and up to two instances of a 20
qubit circuit or 30 qubit circuit.
Since setting up parallelism occurs during the initial compilation phase, this strategy will
realize the same gains as relocation in terms of reduction of overhead time due to reload time.
However, by running multiple circuits at the same time, we are executing several shots per
fluorescence cycle. We will have successfully reduced the overhead time from fluorescence
by a factor of the number of circuits run at the same time. This can be seen in Table 3.1
for two parallel instances for 10, 20 and 30 qubit circuits. Put together with the gains from
relocation, we see total reductions in overhead time for each of our benchmarks of up to 70%
for a 10 qubit circuit, 60% for a 20 qubit circuit, and 50% for a 30 qubit circuit, seen in
Table 3.1.

3.6 Discussion

Reducing the overhead of running compiled quantum programs is critical to successfully ex-
ecuting useful near- and intermediate term quantum algorithms. Neutral atoms have many
55
attractive properties: long-range interactions, native implementation of multiqubit gates,
and ease of scalability. These advantages reduce gate counts and depths dramatically by in-
creasing the relative connectivity of the underlying hardware. While long-range interactions
induce larger restriction zones which inhibit some parallelism, the amount of gate and depth
savings far outweighs this cost.
The dominant cost in NA systems is atom loss. Weaker trapping potential and destruc-
tive measurement leads to the loss of atoms as computation is performed. We explore various
strategies including the extremes of full recompilation and always reloading. Full recompi-
lation is able to sustain high atom loss but is slow when thousands of trials are needed. But
reloading is also slow and is the dominant hardware cost. Our reroute and compile small
strategies balance atom loss resilience and shot success rate to save computation time. As
resources become more constrained, we require more flexible techniques. By exploiting the
unique feature of long distance interactions, we can make use of a small number of atoms
to improve how many successful shots can be achieved before reloading the circuit via the
interaction graph. This flexibility gives an escape valve for more edge case circuits.
However, we do not only focus on adjusting the initial mapping and routing. By changing
the initial mapping and routing while focusing on effectively using all of an architecture
and more actively avoiding lost atoms, we can further reduce overhead time. Through the
creation of tiled bounding boxes, dividing the array into several small architectures, and
fully exhausting each of these sections before moving into a new section without atom loss
overhead time due to reloading the array is reduced by a factor of the number of times the
circuit fits on the architecture. This can all be done while keeping the probability of success
high and without significant increases in the time to adapt the circuit to the lost atoms.
Finally, we also use these different sections to exploit parallelism. While using every
available atom on a device is not effective due to space constraints, using less than 100%
of the architecture combined with relocation achieves the same number of shots per reload

56
while also reducing the number of runs of the array required. This reduces the overhead
time dedicated to florescence, further improving our ability to quickly run repeated shots on
the neutral atom device.
Overall, the unique advantages of neutral atom systems, long-distance interactions and
multiqubit operations, dramatically reduce communication and depth overheads which trans-
lates into lower error rate requirements to obtain successful programs, showing potential for
the advancement of neutral atom architectures. Like their competitors, there are funda-
mental drawbacks of a NA system; here we’ve highlighted the problem of atom loss. This
probabilistic loss is inherent in the trapping process itself. But, we demonstrate that soft-
ware solutions can effectively mitigate the problems due to atom loss. This is critical for
the overall development of the platform: by solving fundamental problems at the systems
level, hardware developers can focus on solving and optimizing other problems, such as gate
errors, and focus on the process of co-design which can accelerate the advancement of the
hardware tremendously.

57
CHAPTER 4

HIGHER-RADIX ARCHITECTURES: TAKING ADVANTAGE

OF VARIABLY COSTED OPERATIONS

4.1 Introduction

As was covered in previous sections, a typical quantum architecture is constructed of binary


qubits, which have two distinct states representing 0 and 1. However, this binary abstraction
is not the whole picture. Some quantum hardware, such as superconducting qubits [25] and
trapped ions[95], have natural access to energy levels beyond the 0 and the 1 states that can
be used as extra computation space. And, some algorithms have been designed specifically
to take advantage of higher-radix computing [109, 53].
One potential option is the use of intermediate qudits [42], which rely on a mixed radix
strategy where a smaller number of qudits are used and are used for a shorter amount of
time. While specialized in its use, the expected advantages are strong, for example in the
generalized Toffoli decomposition (also the subject of this work) using qutrits (three level
quantum systems) temporarily enables a logarithmic depth (approximately circuit duration)
decomposition with linear number of two-qudit gates and requires no additional ancilla space,
scratch bits. The best known qubit-only circuits can obtain logarithmic depth with linear
gate counts [9]. These circuits require linear amounts of additional space which makes
efficient implementations infeasible on error-limited devices and bounding the program size
on available hardware.
Another uses the concept of four-level ququarts to compress the data of two qubits into a
single physical ququart device without the loss of information. This is a more general way to
save computational space. However, this strategy has been avoided due to several drawbacks
that scale with qudit dimension: quadratically increasing logic gate execution time [68], re-
duced coherence time, and increased difficulty of experimentally applying gates [52, 15].
58
These are significant disadvantages for a NISQ device that already has limited connectiv-
ity, short qubit lifetimes, and high gate error. However, if these issues can be effectively
mitigated, through specialized gates and compilation strategies, this general compression
method can effectively double the available computational space on a quantum computer for
any arbitrary circuit.
This chapter explores both methods of utilizing higher radix quantum devices and how we
need to adapt our mapping and routing methods to handle these cases. We will particularly
focus on the trade off of gate count for circuit duration, and whether we reduce communi-
cation costs through clever mapping and routing to avoid increases in circuit duration such
that the error due to decohering qubits outweighs error reduction in other areas.

4.2 Background: Higher-Radix Computation

Classically, the basic computation unit is the bit which takes the value of either 0 or 1.
In the quantum setting, we often consider the quantum bit (qubit) the most fundamental
unit. For many quantum technologies, such as superconducting qubits and trapped ions,
the implementation naturally has access to infinitely many discrete levels which can be
truncated to any dimension d giving qudits which exist as linear superpositions of d levels
as |ψi = α0 |0i + α1 |1i + ... + αd−1 |d − 1i where if we choose d = 2 we recover the qubit.
In general, qudit computation is not asymptotically better than qubit computation – both
schemes can universally express quantum computation, and full translation from one radix
to another affords only constant advantages [87].
Use of additional logical levels for quantum computation is not new. There have been
two primary considerations for their use. First, full translations from qubits to qudits, usu-
ally of small dimension d, have been proposed. For example, translation has been used to
implement arithmetic or Shor’s algorithm [18, 17], which require an expanded gate set to
implement generalized ternary gates. Second, temporary use of additional logical states has

59
been shown to be useful in specific cases. In these cases, programs begin and end entirely
as qubits, but during computation temporarily access additional logical levels. These works
focus on a small set of applications such as the generalized Toffoli gate and adders [8, 117]
and rely on hand optimization to extract benefit from these states without using too many
gates or spending too much time occupying these states. This temporary use strategy has
been generalized as compression [7] to generate ancilla to speed up specialized subcircuits.
These prior works operate in the gate model only, assuming that gates on different dimension
qudits are effectively the same - they can be executed with equivalent fidelity and execute in
a similar amount of time. They also generally ignore architectural connectivity constraints
and the inherent increased cost to communicate qudits at long distances. While gate rep-
resentations are useful, they omit crucial system level details, such as gate duration and
pulse implementation fidelity, which determine the viability of mixed radix computation in
general. Prior work has shown a worst-case quadratic increase in gate duration [68] for
higher-dimensional gates, meaning in practice we must be extremely careful about how we
use qudits.
For a variety of reasons, such as an increasing number of error channels which become
harder to control, using large numbers of states is often impractical and instead we should
carefully choose the computing radix based on our target applications and available hardware.
For example, measurement (the process of collapsing a quantum state to a classical value)
of high level systems is often challenging for trapped ions and if possible we should try to
measure only qubits. Additionally, characterizing higher energy levels is difficult as they are
more prone to noise and suffer from lower coherence times, at a rate of around T1 /(d − 1),
where d is the qudit dimension and T1 is the coherence time for a qubit [15].
As mentioned in 2.1, to manipulate quantum states we apply gates, which can be repre-
sented as unitary matrices. For the most part, hardware supports at most gates on 1 or 2
inputs and all larger gates must be synthesized directly from smaller ones. For higher radix

60
computation, in 4.4 we consider a basis set which consists of the generalized versions of the
X+k gates and the Xi,j gates which are classical permutations of the basis elements. The
first behaves by shifting every basis element’s coefficient +k modulo the dimension of the
system. The second behaves by swapping the coefficients of the i-th and j-th states and
leaving all other coefficients the same [80]. We also consider the controlled versions of these
gates. In some decompositions we will use Toffoli-like gates on higher dimensions, which
have constant depth decompositions into 1 and 2 qudit gates.
However, we can expand this gate set to be any set of operations that interact with the
|2i and |3i state. In 4.4 we will use these gate sets to develop higher-radix circuits that use
these states in an intermediate capacity, expanding on the work in [8]. In 4.5.1 and 4.6 we
will develop a gate set that is functionally the basis of a two-level system, but operates on a
higher-radix ququart based architecture.

4.3 Background: Quantum Optimal Control

Gates on a device are implemented by applying hardware-specific control fields fk (t) to


the qudit(s) involved in the operation. In superconducting architectures, control fields are
analog microwave pulses. Quantum optimal control searches for a control sequence that best
replicates the effect of a target logic gate. Different algorithms and toolboxes have been
designed for this purpose [58, 106, 88, 47]. In this work, we use the open-source optimal
control software Juqbox [89, 88] to find the shortest-duration control pulse sequence that
reaches a specified fidelity for each logic gate of interest. Juqbox optimizes the control fields
fk (t) to minimize an objective function J[fk ] = 1 − F [fk ] + L[fk ] consisting mainly of the
gate fidelity
1 n o2

F [fk ] = 2 Tr UT [fk ] V (4.1)
h

61
between the target unitary V and the applied transformation UT [fk ], where h is the Hilbert
space dimension of the logical subspace. This is achieved by repeatedly solving the Schrödinger
equation and adjusting the control fields in every iteration to minimize J. The full Hilbert
space of the optimization typically includes additional guard states to capture the influence
of higher energy levels present in superconducting systems [62, 88]. As these guard states are
not part of the logical subspace, their populations are penalized with a leakage term L[fk ]
in the objective function.

4.4 Routing Intermediate Qudit Circuits

While the advantage conferred from clever circuit design is clear from prior exploration of
intermediate qudits, prior works omit two critical considerations from their constructions.
First, these circuits are not compiled to any real device. While some devices promise all-to-all
connectivity, most devices have limited connections between devices meaning programs must
have movement operations inserted during compilation. This introduces large amounts of
communication overhead, on average, which must be minimized [51]. Furthermore, swapping
the states of different dimensional qudits requires tailored communication operations which
get more expensive as the dimension of the inputs increases. Second, qubit gates and qudit
gates are not created equal. From a circuit point of view, they take up one unit of time
each, but when implemented on device this may not be the case. To execute gates on
hardware, analog pulse sequences must be generated, for example via the optimal control
process described earlier. The duration of these pulse sequences determines the length of a
given gate. In some cases, converging to a high quality pulse sequence of minimal length is
difficult [88, 89].
In this section, we consider a straight-forward compilation of intermediate-qudit circuits
to connectivity-limited superconducting devices. To do so, we introduce swap gates designed
explicitly to communicate qudits of different dimensions, for example qubits with ququarts

62
(4 level systems). Second, we consider realistic gate times obtained via optimal control
for superconducting based quantum architectures to better study expected circuit durations
beyond circuit depth. Together, we use these compilation tools to study one representative
implementation of the generalized Toffoli using different levels of intermediate qudits. In
particular, this section explores

• Detailed efficient decompositions of SWAP gates between qudits of different dimensions

• A simple compilation pipeline transforming intermediate qudit circuits into ones which
obey connectivity-limited hardware constraints

• An introduction of intermediate ququart and ququint (5 level system) implementations


of the generalized Toffoli, a critical circuit component for many larger algorithms

• An evaluation of the near-term benefits conferred by using intermediate qudits using


realistic gate times for superconducting devices.

4.4.1 Higher Radix Communication Circuits

Current hardware has limited connectivity and only qudits which are adjacent on hardware
may interact. Communication operations, called SWAPs, are required to move qudit states
around the device. Here, we present a generalized decomposition of the SWAP gate taking
in qudits of any dimension.
The decomposition of the qudit SWAP gate is straightforward and follows from the qubit-
qubit swap. The key idea is to consider size 2 subsets of the basis elements. For qubits, this
amounts to the subset {0, 1} for which the decomposition is three 1 controlled X01 flips. For
qutrit SWAPs we now have three possible subsets {0, 1}, {0, 2} and {1, 2} and then perform
partial SWAPs as if the qudits exist only in the subspace spanned by the elements of the
subset. Here, we would do three 1 controlled X01 flips followed by three 2 controlled X02

63
1 X01 1 2 X02 2 2 X12 2

X01 1 X01 X02 2 X02 X12 2 X12

Figure 4.1: The decomposition of a qutrit based SWAP gate.

flips and then three 2 controlled X12 flips. The control value can be either of the elements
of the subset.
As the dimension increases, a SWAP is decomposed into O(n2 ) 2-cycles where n is the
dimension of both qudits. For communication between qudits of different dimensions the
same strategy applies but the scaling is O(nm) where n, m are the dimensions of the two
input qudits. There have been other approaches for generalized SWAP gates by using more
contrived basis elements [111, 116], but we find these to be intuitive and match the expected
asymptotic scaling that optimal control predicts for gate durations. Additionally, this version
of the SWAP gate will handle mixed radix inputs. That is, if one device is in a different radix
from the other, the same set of gates for the higher radix SWAP can be used as would be used
if they were both higher radix devices. In Figure 4.1 we show a swap gate decomposition for
two qutrits.

4.4.2 Synthesizing Gate for Higher Radix Computation with Quantum

Optimal Control

By using optimal control analog pulse sequences can be produced for a given unitary. Typ-
ically, durations are chosen beforehand and the goal is to produce a high fidelity (quality)
implementation of the gate within the allotted duration. Finding optimal pulse durations is
challenging, and is expected to scale quadratically with the dimension of the input unitary.
We consider gate times obtained via optimal control to accurately account for communi-
cation time costs and total circuit duration. While hardware supports a limited gate set,
optimal control procedures can permit us to synthesize any unitary. But, we limit this to
a small gate set on bounded numbers of qudits to limit classical overhead. Instead, we can
64
Qudit Levels Interaction Time (ns) Swap Time (ns)

1 30 -
2 50 -
3 50 -
4 50 -
0, 1* 150 600
0, 2 500 1200
0, 3 500 1500
0, 4 600 1800
1, 1 500 900
1, 2 500 1200
1, 3 500 1500
1, 4∗ 600 1800
2, 2∗ 675 2950
2, 3 850 5000
2, 4∗ 1025 7050
3, 3 850 5000
3, 4∗ 1025 7050
4, 4∗ 1200 7500

Table 4.1: Times used for various gates across different levels of qudits. An asterisk indicates
an interpolated value.

interpolate to predict other gate durations or produce pulses for gates in the decomposition.
As it is not the topic of this work, please see [88, 89] for more information. Gate times for
this work are listed in Table 4.1. The times for qubit-qubit interactions and swaps, and
single qubit interactions are known from gate times from devices such as IBM’s hardware.
Times for qubit-ququart, ququart-ququart, and single ququart interactions and swaps were
found via optimal control. Gate durations for qutrits and ququints were found via a linear
interpolation between these points. The exact times for any gate also depends on the exact
Hamiltonian used to model the underlying device or other physical restrictions. The num-
bers obtained here are from a standard superconducting Hamiltonian, as would be found for
IBM’s hardware [2].
As seen in table 4.1, as the radix increases, so does the time required to perform an oper-
ation on each qudit. As the highest energy level increases, the pulses required to achieved the

65
desired result becomes more complex as it needs to satisfy more constraints and transitions,
elongating the necessary pulse duration. There are several examples of how this could be
achieved physically [63]. These increased radix devices also have higher rates of decoherence,
but if the circuit depth and communication this will be balanced by a shorter circuit dura-
tion. While this section does not focus on other architectures, the presented circuits would
be valid, but would require a new set of control experiments [92]. However, we would expect
similar scaling on these architectures due to the higher level of control required.

4.4.3 Adapting Compilation to Higher Radix Devices

Since many qudits will be in higher-level states, we must use qudit SWAP gates to effectively
route the circuit on a device. While the same challenges from the generic mapping and
routing algorithm still apply, these qudit SWAPs present a new wrinkle. Not all SWAPs
have the same time costs, so one SWAP could be more detrimental in terms of circuit duration
than two SWAPs between lower radix qudits. Additionally, since higher radix qudits have
lower coherence times, longer circuit durations are detrimental to the success rate of circuit
execution.
To adapt our generic compiler algorithm to this structure, we develop a dynamic distance
function that uses the time to perform a SWAP instead of using simple distance in the edge
weight. This is done by maintaining a collection of time values in the routing algorithm and
mapping between each qudit and its maximum possible radix at a given time. We can use
these two values to create a weighted connectivity graph, as seen in Figure 4.2. We then
perform the disruption calculation as we did in the original algorithm with this new distance
function, only examining qubits that move closer together.

66
Figure 4.2: A representation of the dynamic map used for mapping and routing when routing
quantum devices in higher, mixed radix configurations. Each edge is weighted with the
different time it will take to perform a SWAP along that edge.

4.4.4 Evaluation Benchmarks

We focus on the Generalized Toffoli, or N-Controlled X gate as our benchmark. This circuit
is easily generalized to higher level quantum systems with similar constructions at each level.
At the qubit level, this gate can be implemented with any number of ancilla qubits [9]. But,
to achieve this circuit in logarithmic depth we need to use n − 2 ancilla for n number of
controls, significantly reducing the amount of usable space available to execute a quantum
circuit. An example of this circuit for 5 controls with 3 ancilla qubits is in Figure 4.3.
An ancilla-free log depth version of the generalized Toffoli has also been developed for
qutrit capable quantum systems [42]. By using intermediate qutrits, this circuit uses 1 and 2
controlled +1 gates to reduce the number of qubits used and number of gates used. The set
of controls are treated as nodes in a tree, recursively stepping down the child nodes in the
tree to create the circuit. The child nodes target their parent node with a controlled qudit

67
• • 1 1
• • +1 2 2 −1
• •
1 1
• •
+1 2 −1
• •
• • 1 1
• +1 2 2 −1
• 1 1

Figure 4.3: A logarithmic depth Gener- Figure 4.4: A logarithmic depth Gener-
alized Toffoli circuit using qubits with 5 alized Toffoli Circuit using qutrits with 7
controls and 4 ancilla. controls.

Toffoli gate which could increment or decrease the state of a qubit rather than performing
an X gate. In the case of the leaf nodes, we use a 1-controlled qudit Toffoli gate. Then, any
internal nodes use a 2 control since their states will have been increased from 1 to 2 if the
leaf qubits were in the 1 state. Finally, the qubit representing the root of the tree performs
a 2 controlled X gate on the target qubit. We then reverse the controlled qudit Toffoli gates
to return the input qutrit to their original, non-elevated state. An example of this circuit
with 7 controls is seen in Figure 4.4.
We expand on this idea by constructing generalized Toffoli circuits that take advantage
of even higher level quantum systems. We use the same tree-like framework developed
for qutrit circuits, but can use the additional computational space afforded to us by these
systems. Specifically, we use this extra space to remove the use of the gates involving three
qudits, replacing them with two gates that use only two qudits each. While this may seem
like an increase in gate count, the three count qudit gates is decomposed into many more
one and two qudit gates. For qudit Toffoli gates, six two-qudit gates and eight single-qudit
gates, which is significantly higher. In the ququart case, we use each leaf control to target
its parent qudit with a +1 gate. If the controls are in the 1 state, the target will end in
the 3 state after being targeted by two controlled +1 gates. Then the internal controls are
68
1 1

+1 +1 3 3 −1 −1

1 1

+1 +1 3 −1 −1

1 1

+1 +1 3 3 −1 −1

1 1

Figure 4.5: A logarithmic depth Generalized Toffoli Circuit using ququarts with 7 controls.

3 controlled +1 gates. Similarly, the final controlled X gate becomes a 3-controlled X gate.
An example of this can be seen in Figure 4.5.
A similar strategy can be used for any n level system. We split create the control tree into
a tree with n − 2 children rather than the original 2 children in the qutrit and ququart case.
An example of this expansion into a generalized Toffoli gate for 5 level qudits (ququints) can
be seen in Figure 4.6.

4.4.5 Results

In this work we observe the changes in gate count, depth, duration, and space time product
of a given circuit as we adjust our benchmark to utilize the higher radix levels of a quantum
system, up to ququint level devices for the generalized Toffoli gate. Both gate count and
depth of a circuit on their own are helpful metrics to compare the runtime of two different
quantum circuits. However, as mentioned, different gates have different lengths, and may
use more qubits through ancilla. So, it is necessary to examine the duration of the circuit
as well. The product of the duration and number of qubits is also used as the space-time
product. Minimizing this metric indicates an increase efficiency of the compiled circuit on a

69
1 1
+1 +1 +1 3 3 −1 −1 −1
1 1
1 1
+1 +1 +1 3 −1 −1 −1
1 1
+1 +1 +1 3 3 −1 −1 −1
1 1
1 1
1 1
+1 +1 +1 3 3 −1 −1 −1
1 1
1 1

Figure 4.6: A logarithmic depth Generalized Toffoli Circuit using ququints with 10 controls.

given device.
When measuring the time from these circuits, we construct a directed graph with nodes
representing the operations, and edges representing dependencies based on qudits used. We
are able to label these edges with the time it will take based on the current levels of the
qudits at that time. This allows us to find the longest path through the graph, giving us
the duration of the critical path, and the duration of the overall circuit. We examine each
circuit on a 12 qudit by 12 qudit grid architecture. This provides a middle ground between
the structure of current quantum architectures, and devices with higher connectivity. It also
is a large enough architecture to examine the scalability of a circuit as they grow in size.

Benefits of Qutrits

Previous examinations of the intermediate qutrit generalized Toffoli gate did not examine
how it may be affected by routing the circuit on an limited connectivity architecture. In
Figure 4.7, we see that the depth of the log-depth qutrit generalized Toffoli gate is greater

70
Depth of Circuit, Before Routing Depth of Circuit, After Routing
Log Qubit Log Qutrit Log Ququart Log Ququint
150 2,000
Depth of Circuit

100
1,000
50

0 0
0 20 40 60 0 20 40 60
Number of Controls Number of Controls

Figure 4.7: The depth of the Generalized Figure 4.8: The depth of the Generalized
Toffoli gates at different levels of qudits Toffoli gates at different levels of qudits
before communication gates are added to after communication gates are added to
the circuit. the circuit.

than the depth of the qubit generalized Toffoli gate, despite the fact that it requires more
Toffoli gates to implement the qubit gate. Once SWAP gates are integrated into the circuit
via the qudit compilation framework, in Figure 4.8, the depth of the qutrit generalized
Toffoli gate is significantly lower than the qubit generalized Toffoli gate. However, referring
to Table 4.1, the higher the level of the operations, the longer the times. Upon measuring
the duration of a circuit after inserting SWAPs, seen in Figure 4.9 we see that at smaller
numbers of controls, qutrit circuit duration is much lower than qubit circuit duration, but
as the number of controls increases, these values reverse. Where the qutrit circuit was half
the duration of the qubit circuit from 5 to 15 controls, they are 1.5 times the duration from
35 controls onward. The extra time required by the SWAP gates exceeds the benefits gained
from using fewer gates.
However, the construction of the qutrit circuit does not require any ancilla. To take
this into account, we examine the change in space time product of the qutrit versus qubit
generalized Toffoli gates in Figure 4.10. As the qutrit circuit does not require extra qubits,
it can achieve a much lower space time product. This is especially evident when the qutrit
circuit has a low duration, starting at a ratio of 3 to 1 from qubit to qutrit space time ratio.
As the number of controls increases, this ratio converges to 1.7 to 1 space time ratio from
71
Circuit Duration Space Time Product
·105 ·107
Circuit Duration (ns) Log Qubit Log Qutrit Log Ququart Log Ququint

Space Time Product


1.5 1.5

1 1

0.5 0.5

0 0
0 20 40 60 0 20 40 60
Number of Controls Number of Controls

Figure 4.9: The circuit duration of the Figure 4.10: The space time product, or
Generalized Toffoli at different controls number of qudits used by duration, at dif-
for different qudit levels. ferent controls for different qudit levels.

qubits to qutrits, which matches the expected constant increase in computational space from
qubits to qutrits, log2 (3) = 1.58.

Benefits of Ququarts over Qutrits

Moving from the qutrit circuit to the ququart circuit does not reduce the number of the
qudits. However, since we do not need to decompose any three qudit gates, the circuit
requires fewer gates overall. The difference in depth before and after inserting SWAPs can
be seen in Figure 4.7 and Figure 4.8. After routing, the depth of the qutrit circuit is 5 to
6 times greater than the ququart circuit, following from the ququart circuit using two two
qudit gates, with a depth of two, rather than a Toffoli gate, which uses eight single qudit
gates and six two qudit gates and has a depth of 11.
Taking the duration of these ququart gates into account in Figure 4.9, the reduction
in gate count and depth generally outweighs the increase in time to perform the qudit
interactions. Generally, the ququart circuit time will approach the qutrit duration when the
number of controls approaches a power of two. Certain numbers of controls are not conducive
to the tree-like structure, and will result in an excess number of gates. When we examine the
space time product, ququarts more efficiently utilizes the architecture than both qubits and
qutrits. In fact, the qubit to ququart space time product ratio is an average of 2.3 after the
72
initial few increases in controls, which is inline with the expected increase of computational
space from qubits to ququarts, log2 (4) = 2. From qutrits to ququarts the expected increase
would be log3 (4) = 1.9, and we find the ratio in space time product from qutrits to ququarts
to be 1.43, an increase in the efficiency of the use of the device.

Diminishing Benefits of Higher Levels

When we move into higher level systems, specifically levels 5, 6 and 7, we do not see the
benefits of using fewer qubits. Since these generalized Toffoli gates have similar constructions
to the ququart gate, and do not have the benefits of gate reduction, and do not see significant
decreases in the number of gates or depth of the circuit. As seen in Figure 4.7 and Figure
4.8, the ququint depth is the same as the ququart depth before and after insertion SWAPs
for routing. Interactions at the ququint level take longer than at the ququart level. In Figure
4.9 we see that the time for a ququint circuit to execute is generally greater than the time
for a ququart circuit to execute. This translates to an increased space time product when
compared to ququarts. The benefits afforded by the computation space at this level of qudit
do not outweigh the cost of the complexity and operations required to reach this level over
ququart circuits.

4.4.6 Discussion

This section described a strategy for qudit communication through efficient decomposition
of SWAPs, and developed a straightforward compilation pipeline for qudit circuits, using
realistic gate times derived from optimal control. Further, we were able to introduce a
generalized Toffoli gate that uses intermediate ququarts, ququints, following a structure that
can be generalized to any number of qudit levels built on strategies developed for intermediate
qutrits. Finally, we compiled these circuits for potential architectures to evaluate the benefits
at each level of qudit through examination in the change in gate count, depth, execution

73
time, and space time product.
Increasing the highest possible level of a qudit offers opportunities to improve the uti-
lization of the architecture. From qubits to qutrits we see significant decreases in space time
product to due reduction in ancilla use, even with increases in time of both interactions and
SWAP gates. This trend continues from qutrits to ququarts, where the extra computational
space allows us to use two two qudit gates rather than a three qudit gate, reducing duration
of the circuit, and the space time product of the circuit as well and improving the use of
the architecture. These advantages show that there are gains to be found by implementing
these higher dimensional systems and integrating qudit compilation strategies to improve
the usage of near term quantum devices.
By adapting the decomposition strategies with more specific communication gates, and
adjusting our general purpose algorithm to be time aware, and not just connectivity distance
aware, we can efficiently use higher-radix operations significantly reducing opportunities for
gate error while maintaining similar circuit duration. Since gate error is a much greater
source of error than coherence error in the long term, this is a clear case of how careful
compilation for specific architectural features is able to improve circuit success rate.

4.5 Qompress: Compiling Quantum Circuits Using Partial and

Mixed-Radix Operations

4.5.1 Introduction

Higher qudit states can be used in a more general way, compared to intermediate qudits
circuits, in order to save computational space. The information of two qubits can be fully
encoded in a single ququart [7], storing two logical units’ worth of information in a single
computational unit of higher radix. However, this strategy has been avoided due to several
drawbacks that scale with qudit dimension: quadratically increasing logic gate execution

74
Figure 4.11: Pairs of qubits can be compressed in four-dimensional ququarts and interact with
each other internally or through partial operations, enabling novel compilation techniques
and space reduction.

time [68], reduced coherence time, and increased difficulty of experimentally applying gates
[52, 15]. These are significant disadvantages for a NISQ device that already has limited con-
nectivity, short qubit lifetimes, and high gate error. However, if these issues can be further
mitigated, this general compression method can effectively double the available computa-
tional space on a quantum computer for any arbitrary circuit.
Previous work has attempted temporary ququart encoding to harness the potential for
fast "internal" gates between the two encoded qubits, but had to decode for any operation
outside of a specific ququart. In this work, we use quantum optimal control to synthesize
high-fidelity, low duration pulse implementations of specific set of gates, that can interact
two qubits in the same ququart, a qubit outside of a ququart with a qubit encoded in a
ququart, or two different qubits encoded in two different ququarts while leaving others as
bare qubits. Each of these gates have varying speeds, and are shown in Figure 4.11.
A mixed-radix paradigm where qubits are directly manipulated within an encoded ququart
adds a degree of flexibility not found in the prior work. Inter-ququart operations no longer
incur the extra cost of encoding and decoding. Motivated by these new gates, we explore
strategies to efficiently generalize partial ququart compression of a circuit to be used with

75
any qubit-based circuit. The partial qubit-ququart and mixed-radix operations inform the
design of a compilation pipeline designed to take advantage of the flexibility of our gate set
while minimizing the error due to decoherence.
This section proposes several different strategies to find the ideal compression to make the
best use of the newly-found ququart compilation methods. Specifically, this section adapts
the target gate set, pulse generation, mapping and routing sections of the compiler by:

• Developing a library of high-fidelity mixed qubit-ququart operations for partially com-


pressed qubit systems via quantum optimal control for a realistic device Hamiltonian.

• Demonstrating viability of ququart operations via pulses generated optimal control on


hardware not previously designed for ququart pulses.

• Detailing a compiler pipeline that takes advantage of partial ququart operations for
mixed-radix computing.

• Evaluating several compression strategies to selectively encode qubits in ququarts using


the new compiler pipeline.

• Demonstrating simulated increases in gate fidelity of more than 50% over standard
qubit-only compilation across various quantum architectures, as well as up to 2x in-
creased qubit capacity, while accounting for the increased coherence errors associated
with operation on ququarts.

4.5.2 Compression and Gate Set

We can redefine the higher-radix gate set to more natively represent traditional qubit opera-
tions. To do this, we follow a qubit to ququart compression scheme inspired by [7]. Encoding
the state of two qubits |q0 q1 i into a single ququart state promotes one of the qubits q0 to a
ququart while transforming the other qubit q1 to an ancilla in the ground state |0i. While

76
this is referred to as a compression, we do not lose any data. It is the full representation of
two qubits, in a single physical unit. We define the encoding gate ENC as





 |0i |0i → |0i |0i




 |0i |1i → |1i |0i

ENC = (4.2)




 |1i |0i → |2i |0i




 |1i |1i → |3i |0i

This gate defines the encoding scheme; for example, the |2i ququart state represents the |10i
qubit-qubit state, and we can perform logical operations on the ququart states as if they are
the states of two qubits. We note that, since we do not expect q0 or q1 to be in a ququart
state prior to the encoding operation, the extension of ENC to a full ququart-ququart unitary
gate is arbitrary.
Additionally, to measure a ququart, we simply use the mapping in reverse to determine
the state of the two encoded qubits from the state of the ququart. In a physically realized
system, we are not able to measure one encoded qubit without simultaneously measuring
the other, unless the qubits are first decoded.

Gate Set

Under this encoding scheme, qubit-equivalent operations can be derived for ququarts that
encode two qubits, some of which are shown in Figure 4.12.
For example, X gates acting on the encoded qubit state |q0 q1 i correspond to the ququart
operators X0 ∼ X ⊗ 1 (which switches the population of |0i with |2i, as well as |1i with |3i)
and X1 ∼ 1 ⊗ X. These gates can also be executed in parallel by applying a ququart gate
X0,1 ∼ X ⊗ X.
Any two-qubit gate acting on qubits encoded in the same ququart can be expressed

77
Figure 4.12: Two qubits q0 and q1 can be encoded into a ququart (blue oval) and interact with
each other internally (green), with a bare qubit q outside (yellow), or with encoded qubits in
a different ququart (red). CX arrows point from control qubit to target qubit. Along each
CX link, corresponding SWAP gates are defined with the same superscripts/subscripts (not
shown).

as a single-ququart gate. For example, an internal SWAP gate (SWAPin ) corresponds to a


simple X12 operation that exchanges populations of the ququart |1i and |2i states. Similarly,
we can define internal CX gates {CX0 , CX1 } between the two encoded qubits. On some
hardwares, these operations can be done much faster (and with higher fidelity) than the
corresponding two-qubit operations. However, in order to compile quantum circuits for a
mixed-radix quantum computer using both qubits and ququarts, two-qubit gates between a
bare qubit and an encoded qubit as well as between encoded qubits in different ququarts are
necessary. To this end we extend our gate set by a variety of partial CX- and SWAP-like
gates acting on various qubit pairs as shown in Figure 4.12. We define four partial CX
gates {CXq0 , CXq1 , CX0q , CX1q } and two partial SWAP gates {SWAPq0 , SWAPq1 } that
realize the respective two-qubit gate between a bare qubit and an encoded qubit. Similarly,
four partial CX gates {CX00 , CX01 , CX10 , CX11 } and three partial SWAP gates {SWAP00 ,
SWAP01 , SWAP11 }. We focus only on unique gates, SWAP01 and SWAP10 are equivalent.
can manipulate the states of encoded qubits in different ququarts. We further include the
full ququart-ququart SWAP4 gate to enable routing of ququarts.

78
4.5.3 Synthesizing Mixed-Radix and Full-Ququart Gates

Qudit logic gates are typically assumed to require long durations to implement, due to
the increased complexity of the Hilbert spaces involved [68]. This is a problem for several
reasons: First, longer-duration gates lead to overall longer-duration circuits; second, higher
qudit states have shorter decoherence times [15], amplifying the negative effect of these longer
circuits. However, by explicitly synthesizing control pulses for higher-radix gates, we find
gate durations that scale efficiently enough to provide overall circuit benefits when combined
with a tailored compiler.
We obtain durations of gates in our system by explicitly optimizing high-fidelity control
pulses for superconducting hardware. We model two-qudit subsystems with two weakly
coupled, anharmonic transmons [62] with total Hamiltonian

2  
X † ξk † † 
† †

H(t) = ωk ak ak + ak ak ak ak + J a1 a2 + a2 a1
2
k=1
(4.3)
2  

X
+ fk (t) ak + ak ,
k=1

where the first two terms comprise the drift Hamiltonian and the last describes the effect of
the control fields fk (t) applied to the transmons. We choose realistic physical parameters
inspired by [100]: The 0-1 transition frequencies of the transmons are ω1 /2π = 4.914 GHz
and ω2 /2π = 5.114 GHz, and both transmons have the same anharmonicity ξ1 /2π = ξ2 /2π =
−330 MHz. They are effectively coupled with J/2π = 3.8 MHz. For single-qudit gates we
reduce this model to k = 1 and remove the J coupling term. We restrict the maximum
amplitude of the control fields to fmax = 45 MHz to constrain potential leakage into guard
states.

79
Optimizing Pulses

We aim to find control pulses of shortest duration for each gate. As real quantum systems
are limited by noise effects such as decoherence, this allows us to investigate the expected
cost of using these gates in quantum circuits. Juqbox only allows pulse optimization for a
fixed time interval [0, T ], therefore we minimize pulse durations T by applying the technique
from [99], which involves iterative re-optimization with previous pulse results. This method
aims to find the shortest-duration pulse that can execute the desired gate above a minimum
fidelity target. In this work, our target fidelity for single-qudit gates is F = 0.999 and for
two-qudit gates F = 0.99.
Figure 4.13 visualizes equivalent state evolutions in two different CX gates: the standard
CX2 between two bare qubits and the partial CX0q controlled by an encoded qubit and
targeting an unencoded qubit. In both cases, the control qubit is in state |1i (the ququart
state |3i is equivalent to the encoded qubit state |11i), causing the target state to flip from
|0i to |1i. An important observation is the difference in complexity of the state dynamics
for the two gates. CX0q operates on twice as many states in the logical subspace as CX2 .
This suggests that the difficulty of finding high-fidelity control pulses increases with Hilbert
space dimension and explains the variation in gate durations we observe when repeatedly
optimizing for complex two-qudit gates.
The closed system considered does not account for the full dynamics of a real quantum
device. We have not specifically optimized these pulses under a more detailed model due
to the increased computational cost of these optimizations, especially for the large Hilbert
spaces involved in two-qudit operations.

Qudit Gate Durations

We observe several interesting relationships across gate types. Using an internal CNOT or
SWAP instead of its corresponding qubit-qubit operation gives a significant speedup. Addi-

80
(a) CX2 state evolution (b) CX0q state evolution

Figure 4.13: Exemplary state evolutions of two CX gates between (a) two bare qubits and
(b) a bare qubit and an encoded qubit. (a) The control qubit q0 is in state |1i, hence the
state of the target qubit q1 is flipped. (b) The encoded qubit q0 inside the ququart controls
the CX gate targeting the bare qubit q. The ququart state |3i corresponds to the two-qubit
state |q0 q1 i = |11i, so the state of q is flipped.

Table 4.2: Gate durations for one and two-qubit gates synthesized in the qubit-only, mixed-
radix and full-ququart environments.

(a) Qudit (b) Qubit Only


X 35 X0 87 CX2 251
X1 66 X0,1 86 CZ2 236
CX0 83 CX1 84 SWAP2 504
SWAPin 78
(c) Mixed-Radix (d) Full-Ququart
CX0q 560 CXq0 880 CX00 544 CX01 544
CX1q 632 CXq1 812 CX10 700 CX11 700
CZq0 384 CZq1 404 CZ00 392 CZ01 488
SWAPq0 680 SWAPq1 792 CZ11 776 SWAP00 916
ENC 608 SWAP01 892 SWAP11 964

81
tionally, a SWAP between a bare qubit and an encoded qubit (680 or 792 ns) is significantly
faster than a SWAP between two encoded qubits (892-964 ns). We emphasize that these
relationships are specific to this Hamiltonian, and another system may have different trade-
offs in gate durations. Our goal is to design a compiler that can adapt to different sets of
gate durations to determine favorable ququart encodings, without explicitly depending on a
specific relationship such as the internal CNOT advantage shown here.

Technical Barriers

The main technical barrier to a physical implementation of this scheme is accurate control of
quantum logical units at higher energy levels, not the ability to access them. There are several
documented machines that use qutrits [15, 49, 22, 39] and demonstrations of ququart devices
[24]; although calibration of these devices is more time-consuming, it is not a fundamental
limitation. In superconducting systems, as the energy level increases, increased charge noise
causes phase errors and faster state decay, making control more difficult [62, 52, 15].
There is precedent for use of more than two energy levels, e.g. as protection or guard
states or to improve measurement and expedite qubit-level gates [62, 64, 41, 70], indicating
that it is possible to reliably use higher-energy states. However, at present, we are not
aware of experiments using highly-optimized pulses for these energy levels, which would be
a potential challenge for the results shown in this work.
As access to calibrated devices with high-dimensional qudits is extremely limited, our
methods are currently difficult to physically verify, and the efficiency of controlling these
energy states remains an open question. The overhead and difficulty of controlling higher
energy states are potential issues and may change the benefits found in the work. We attempt
to be as realistic as possible by designing this scheme with a transmon Hamiltonian based
on IBM’s real hardware, similar to that which has experimentally demonstrated support for
qutrit operations [39]. Quantum optimal control has experimentally been proven successful

82
in synthesizing accurate pulses within a certain error threshold for qubit-only gates on similar
devices [3], so we are optimistic that the scheme and compilation strategy laid out in this
work will be applicable. Additionally, in an ongoing collaboration with the Schuster group
[1], we are working to validate our methodology and have to date demonstrated some single
ququart gates with high fidelity. While the control pulses that we obtained in this section do
not correspond to a real, fully characterized device, we again emphasize that the compilation
techniques discussed in the following section are independent of the specific control pulses
used, and will adapt to gate durations and error rates different than obtained here. We
intend this work to act as a proof-of-concept to demonstrate the potential value of accessing
higher-energy qudit states and as a motivating factor to improve ququart characterization
and control.

4.5.4 Experimental Demonstration of Single-Ququart Control

Driven by advantages found in theoretical studies [87], experimental researchers have ex-
plored the implementation of these higher-dimensional systems, leading to realizations of
qutrit devices which manipulate the third energy level [39, 50, 96, 44, 75, 121]. These works
show that including higher levels is possible although challenging due to higher susceptibility
to noise and lower coherence times.
Motivated by the findings for ququart-specific applications we have been studying control
of four energy levels in experiment on a physical device. We extend the capabilities of
one qudit of the superconducting transmon device presented in [removed for anonymity] to
include the fourth state. We implement two-qubit Randomized Benchmarking (RB) [72] on
this single ququart following our encoding scheme. RB is a common method to characterize
the average Clifford gate fidelity. This is achieved by executing Clifford circuits of varying
depth, which perform the identity operation in the ideal case, and measuring the probabilities
of the system returning to the ground state (survival probabilities). The fidelity can be

83
Figure 4.14: Interleaved Randomized Benchmarking for an optimal control H ⊗ H pulse on a
superconducting transmon ququart following our qubit encoding. We use two-qubit Clifford
sequences of gate depth up to 100 and average each data point over 10 samples. Error
bars show the standard deviation of the mean but they are smaller than the mean markers.
Red: Standard two-qubit Randomized Benchmarking to estimate the average Clifford gate
fidelity to be FRB ≈ 95.8%. Blue: Interleaving the H ⊗ H pulse between the RB Cliffords
yields a combined per-operation fidelity of FIRB ≈ 92.1%, resulting in an H ⊗ H fidelity
FHH ≈ 96.0%.

84
extracted from exponential regression. The RB circuits are generated using Qiskit [4].
We additionally implement Interleaved Randomized Benchmarking (IRB) [73] to specif-
ically find the fidelity of the single-ququart gate H ⊗ H, which performs a Hadamard gate
on each encoded qubit in parallel, used by the compiler below. The gate control pulse is
designed using similar optimal control methods as discussed in Section 4.5.3 adapted to this
experimental device.
Results from this work are shown in Fig. 4.14. We find an average Clifford gate fidelity of
FRB ≈ 95.8% from normal RB while interleaving with the H ⊗ H gate yields FIRB ≈ 92.1%
fidelity per operation. From that the specific gate fidelity, FHH ≈ 96.0%, can be extracted.
This first study shows that ququarts can be realized in experiment and optimal control yields
high-quality pulses to manipulate their state. At the time of this writing we are not aware of
any comparable demonstration. We are convinced that the fidelities can be improved with
more carefully engineered ququart devices and more sophisticated pulse design methods.

4.5.5 Adapting Compilation for Qubits-on-Ququarts

Compiling a circuit for an architecture which supports full encoding of qubits into ququarts
is not fundamentally different from compiling in a qubit-only architecture. However, we must
account for higher connectivity and varying efficiency of operations between different pairs
of qubits based on their configuration. This includes a new library of communication gates,
including the encoding gate, mixed-radix gates, and cross-ququart gates. We extend the
original compiler algorithms to account for these new variables and fully utilize the unique
set of gates to minimize space requirements while increasing expected circuit success rate
and ideally minimizing increased circuit durations.

85
Representing a Ququart Architecture

Each quantum unit in the system can either be treated as a bare qubit only accessing the
lowest two energy levels, or a full ququart which can access the lowest four. We represent
this mixed-radix architecture via two graphs. One represents the overall topology of the
architecture, where each node represents a single quantum logic unit and edges represent
allowed communication between these units. The second represents the logical qubits we
will be mapping onto from our original circuit. In this graph, each physical unit is treated as
if it is a ququart, and is expanded into two qubit nodes that are connected to one another,
called an interaction graph. Both qubit nodes in the expanded ququart are connected to
each qubit node contained in adjacent ququarts. For example, if a ququart was connected
to n other ququarts, each encoded qubit is connected to 2n + 1 other encoded qubits. In
the secondary qubit graph, there are 2V nodes and 4E + V edges, where V and E are the
number of physical logic units and number of interacting pairs of units, respectively. This
expansion is shown in Figure 4.12. Edges are annotated with a set of execution times and
error rates for CX gates and SWAP gates, which differ depending on whether either unit is
a bare qubit or encoded within a ququart.

Extending Mapping, Routing, and Scheduling

In our proposed mixed-radix system, not all communication is equivalent, or even similar,
in cost. A CX gate between a qubit and ququart is lower fidelity and has a longer duration
than a CX gate between encoded qubits inside the same ququart. Therefore, our compilation
schemes aim to reduce error and circuit duration by taking advantage of gates unique to
ququart systems.
The goal of mapping is still to place frequently-interacting qubits close to one another
on the device. Since the ququart architecture is modeled as a set of qubits with different
connections, we can easily extend mapping strategies designed for qubit-only architectures to

86
ququart-based architectures. First, we find an interaction weight between each pair of qubits
in the original circuit. We use the original weighting function to determine the interaction
between all of the logical qubits. We find the center-most ququart in the architecture and
place this qubit in the first encoding position. For each unmapped qubit, we compute the
greatest sum of weights to the already placed qubits and score each potential placement by
how strongly it interacts with mapped qubits and its distance from them.
In our mixed-radix system, we only consider the second encoding location in ququarts
if the first encoding location has already been mapped to. In the original mapping func-
tion, or when the mixed-radix system is in a qubit only configuration, distance can be a
simple shortest path calculation, with edges weighted by the fidelity of the connection. We
do the same for ququarts, but with dynamic weighting based on the current encoding of
qubits. We represent the current duration of gate g at connection (i, j) as T (i, j, g) and
the fidelity as F (i, j, g). So, for a gate at a given connection, the probability of success is
S(i, j, g) = F (i, j, g)e−T (i,j,g)/T1,i e−T (i,j,g)/T1,j , where T1 is the decoherence time for a qubit
or a ququart, whichever is being used. This is a common metric of success for a quantum
gate as used in [108], and is similar to that used in Section 3.3. The aggregate probability
of success for a given path Pn of length n is then modeled as:

S(P ) = − log(S(Pn−1 , Pn , CX))


n−2
X (4.4)
+ [− log(S(Pi , Pi+1 )), SWAP))]
i=1

This calculation allows us to account for errors introduced by both the decoherence time
and the number of gates. We repeat this process iteratively until every program qubit has
been mapped to a hardware location represented in the extended ququart graph.
In our system, qubits are still tracked individually even if they are both stored in the
same physical object. Therefore, routing for a mixed-radix device is still performed at the

87
qubit level. Any qubit routing strategy, such as lookahead strategies [119, 55, 4], can be
translated directly to encoded qubits in ququarts through routing based on the logical qubit
architecture graph rather than the ququart-level graph. The main goal is to disrupt the
current mapping as little as possible as qubits that need to interact are moved closer to one
another. We choose the candidate location which disrupts the current state of the circuit
the least and moves the qubits closer together based on the probability of success laid out
in Equation (4.4).
However, we do place some constraints on qubit movement. First, we do not encode
new ququarts during routing. Second, we avoid swapping “through” ququarts when possible.
Ququart operations remain expensive, and pairings determined during mapping are often
beneficial to reducing execution costs. Both of these will incur extra costs. Additionally,
by placing these restrictions on the compiler, we ensure that there are fewer changes in the
maximum energy level of the computational units, we can cache the calculated distances,
significantly reducing the amount of classical computation required.
Circuits using fully encoded ququarts require more serialization than an analogous qubit-
only circuit. If, at a given time, each qubit in a fully encoded ququart is involved in a CX
operation, we can no longer execute both gates in parallel and they must be sequenced one
after the other. To break this tie and avoid bottlenecks, we select whichever operation is
on the longest execution path of remaining qubits. This can also be an issue for two single-
qubit gates targeting two qubits in the same encoded ququart. In this case, we combine
both operations into one single-ququart gate, as executing one gate acting on a full ququart
is less error prone than executing two single-qubit gates.

4.5.6 Qubit Compression Strategies

Poor initial mapping can incur high communication costs later on. Additionally, ququart
gates are inherently more time-intensive and can incur serialization, increasing circuit du-

88
ration and error due to decoherence. Simply extending qubit mapping strategies to the
expanded ququart-based graphs will not take any of these factors into account, or take full
advantage of the enhanced connectivity of fully encoded ququarts. In this section, we explore
various compression strategies to better account for these features.
The true effect of individual compression circuit duration or fidelity cannot be reliably
predicted without recompiling the circuit with a prescribed compression. Further, finding
the optimal set of compressions is computationally difficult on its own since there are expo-
nentially many possible subsets. To get a sense for an upper bound on circuit quality, we
explore an exhaustive, but iterative and greedy, search of qubit compressions.
At each step, we recompile the input circuit with every pair of qubits compressed and
choose the option that maximizes the circuit fidelity. Searching every pair, while more
complete, is computationally demanding. Circuit duration is defined by the length of the
critical path and therefore it may be more advantageous to explore only compressions which
affect the critical path. When choosing compressions we prioritize in order compressions
which affect: 1) qubits in non-communication gates on the critical path, 2) qubits off the
critical path that insert communication operations along the critical path, and 3) qubits off
the critical path. In group order, we choose the best compression, if any, and repeat. We
compare this ordered selection to an unordered selection in Figure 4.15.
As can be seen in Figure 4.15, it is not the case that the compression of any two qubits
will be advantageous to the combined gate success rate or error due to decoherence. There
are two classes of advantageous compression. Some compressions are beneficial due to high
interactivity between the two encoded qubits, winning from the fast internal gates. An
example is the pairing of qubits 6 and 11 in the circuit shown in Figure 4.15(a). These qubits
interact very often, so compression increases the probability of success of CX gate by turning
it into a single-qudit gate. The second compression type wins by reducing communication,
such as the pairing of qubits 4 and 14 in the cylinder graph based circuit. These qubits

89
(a)

(b)

(c)

Figure 4.15: An example of an exhaustive search on a cylinder-based interaction graph (a),


using a (b) critical path focused selection, and a (c) selection strategy that allows for any
pair to be selected.

90
(a) (b) (c) (d)

Figure 4.16: (a), (b) Sample generalized Toffoli gate with corresponding interaction graph.
(c), (d) Cuccaro adder with interaction graph, M and N are three qubit gate blocks. Some
circuits, such as the CNU circuit and Cuccaro adder have clusters of qubits that interact
with one another. These clusters can be identified by finding the cycles in the interaction
graph. We have identified the cycles here with colored edges.

rarely interact, but this reduces the diameter of the circuit, decreasing required SWAP
gates. Critical path prioritization and an unordered search of the entire space result in
similar success rate gains, but different compressions but depends on the circuit structure.
Examining all potential compressions is quadratic in the number of qubits in the circuit,
on top of the relatively high complexities required for mapping and routing. Using the
insights found through this method, we build strategies that are able to approximate these
benefits.

Extended Qubit Mapping (EQM)

EQM relies entirely on the interactions between the pairs of qubits, and follows the algorithm
in Section 4.5.5. There is no explicit pair selection in this method. The qubits are filled
greedily, based on the highest weight to the already placed qubits. This strategy clusters
qubits that interact often closely together, but will likely preemptively encode two qubits as
it sees immediate benefits based on those already placed, focusing on identifying interaction-
based compressions. But, it does have classical advantages: EQM is wrapped into the
mapping step, and the only additional complexity is due to the expanded qudit graph.

91
Ring Based (RB)

While frequency of interaction is a good metric for identifying locality, it fails to account
for more holistic interaction structures. Figure 4.16 shows that for certain circuits, such
as the generalized Toffoli gate and the Cuccaro adder, there is a regular, triangle-based
structure in the interaction graph. We can use compressions of these triangles to transform
the interaction graph into a line. Looking at Figure 4.16(b), this means compressing pairs 0
and 1, 2 and 6, and 3 and 4. A linear interaction graph is a much more favorable structure
to mapping. In fact, for certain configurations, it can be mapped and routed without any
SWAPs on any architecture.
We can generalize this triangle compression to any size cycle. If we can find beneficial
compressions within a given set of cycles, we may be able to further transform circuit inter-
actions for more effective mappings. It is computationally expensive to find all cycles in an
undirected graph, therefore for each qubit, we find the minimum size cycle which includes
it. This ensures that each qubit is contained in at least one cycle without finding all cycles.
We find the minimum cycle length from this set of cycles, and use it to bound the maximum
identifiable cycle size.
After we identify cycles, we find the qubits with the fewest interactions outside their
cycle and test compressions with each other qubit in the cycle. For each compression, we
collect information such as the number of shared neighbors between qubits, the weight of
interaction between the compressed qubits, how often they are interacted on simultaneously,
and how many cycles this pair appears in.
We prioritize compressions that maximize the number of internal interactions and the
number of connections to other qubits while minimizing the time the encoded qubits are used
at the same time, to make full use of both the higher fidelity internal ququart interactions
and the new partial ququart operations without introducing serialization. We pick the best
compression based on these metrics. The qubit nodes are removed from the interaction

92
graph and replaced with a single node representing the pair, with edges connected to each
interacting qubit or pair. Following these adjustments to the interaction graph, we recollect
interaction statistics and repeat for other pairings until no other beneficial compressions can
be made.

Average Weight per Edge (AWE)

Another method, Average Weight per Edge, makes compressions to maximize the average
edge weight of the interaction graph. This method takes advantage of shared interactions to
increase potential locality and reduce communication.
We examine each pair of qubits and select the pair that maximizes the average weight
per edge once the two qubit node is collapsed into a single node. We repeat until no more
compressions can be made, or until there is no pairing that would increase the average weight
per edge.

Progressive Pairing (PP)

The final strategy is an extension of EQM, which is more computationally intensive but
collects more information. We start by mapping the circuit to a qubit-only architecture.
This provides a full picture of how the circuit can be laid out on device, compared to the
incremental version provided by EQM. We again examine each potential pairing and compute
the estimated fidelity with and without the compression based on changes in distance between
interacting pairs without remapping and rerouting. In this scheme there are two choices
when compressing two qubits A and B, either A is first or B is first. We select compression
adjustment that is expected to increase the fidelity the most.
After selecting the pair that has the greatest increase in estimated overall fidelity, we
remap and re-evaluate the circuit with the given pair and repeat the process. We continue
this process until we cannot find any pair that will reduce the estimated overall fidelity. While

93
this method scales quadratically, we avoid mapping and routing for each possible pairing by
only determining distances. While less scalable than the other methods, Progressive Pairing
attempts to replicate the exhaustive search with much less classical overhead.

4.5.7 Evaluation Methods

Architectures

One of the biggest limitations on a quantum architecture is simply the number of qubits
available for execution, which constrains many currently feasible programs. Therefore, for
different strategies across circuit sizes, we test circuits on regular architectures that are
just large enough for the circuit in question. We use a rectangle grid-based mesh as our

architecture, with dimensions of d ne × d√nne where n is the number of qubits in the circuit.
Each qubit is connected to four other qubits directly adjacent to it except along the border
of the architecture. This construction allows for testing the scalability of our methods and
test performance in the edge cases of high architectural usage.
We also examine our methods on the 65-qubit IBM Ithaca heavy-hex topology [2] and a
65-qubit ring architecture. This gives us a sense of how well our methods work on devices
with more constrained connectivity.

Gate and Coherence Times Statistics We have shown gate times found via optimal
control in Section 4.5.3. We use the target fidelities for these gates as their success rates.
Single-qudit gates (regardless of Hilbert dimension) are optimized to a success rate of 99.9%
and two-qudit gates to 99% when simulated. The product of the success rate of every gate
in the circuit is the Expected Probability of Success (EPS) with respect to gate execution.
To account for increasing noise in higher dimensional systems we consider their coherence
times (T1 ), after which the qubits are unable to maintain their elevated energy state. For a
d-level system, the coherence time is estimated to be 1/d−1 of the T1 time for the original

94
two-level system [15]. Critically, because gate durations get longer with corresponding lower
T1 times we capture the increasing susceptibility to error in a mixed-radix system. We use
a 163.5 µs qubit T1 time in our architecture, which gives a 54.5 µs worst case T1 time when
we are in a ququart state. [2]. The product of the probabilities of no decoherence for each
qubit, e−tqb /T1qb −tqd /T1qd , is the EPS with respect to coherence time. Here tqb is time spent
in the qubit state, and tqd is the time spent in the ququart state.
The product of these two statistics gives the overall EPS for the entirety of the circuit,
and gives an upper bound for the amount of error found in the circuit. It is the worst case for
both gate fidelity and coherence time as it assumes each qubit will be measured and used for
the entire duration of the circuit, and uses the worst case coherence time ratio from qubits
to qudits. However, it does not take into account the effects of more complicated errors such
as crosstalk.

Baselines

We compare against two different baseline strategies, representing the extremes of compila-
tion for qubits on a ququart-supported device.

Qubit-Only Compilation One extreme is to never encode any ququarts. In this case,
we use EQM but never allow exploration of the second encoded position. These strategies
are comparable to other compilation pipelines described in [77].

Full Ququart Pairing with Encoding and Decoding (FQ) There has been discussion
of the potential of fast internal ququart operations and use ququarts for general qubit circuit
compilation [7]. However, without partial ququart operations, which were assumed to be as
expensive as any multi-qudit operation, any external operation required decoding the two
qubits, performing an operation, and re-encoding. A drawback of this strategy is that extra
space is always required, since a decoding operation requires an ancilla bit to decode into.
95
We are unaware of an automatic compilation pipeline constructed for this strategy. We
use an EQM strategy which allocates extra decoding space next to fully encoded ququarts
as the mapping strategy. For routing, we use a similar strategy, but only at the qudit level.
This means that we only use full qudit SWAP gates to ensure two qudits are adjacent to one
another. Additionally, we must route the empty decoding ancilla next to the neighboring
qubits so that we may perform decompression.

Benchmarking

We use a number of different types of circuits to explore many use cases to test the viability of
each compression strategy. One set of circuits we explore have localized groups of qubits that
interact together. They have no potentially random operations that disrupt the grouping of
operations. These include the Cuccaro Adder [28], Generalized Toffoli (CNU) [11], Quantum
RAM (QRAM) [43] and Bernstein Vazirani (BV) [14].
We are also interested in certain graph-based interaction structures. We use a QAOA
construction [35] that accepts a graph where each node represents a qubit and an edge
represents an interaction. For each edge, in a random order, we perform a CX, a Z gate,
and another CX gate. These circuits are not necessarily used in practice, but allow for
the examination of particular interaction structures which may be relevant in optimization
problems. We examine a random graph with edge density 30%, a cylindrical graph shown in
Figure 4.17(a), a similar torus structure (Figure 4.17(b)) and a binary welded tree (Figure
4.17(c)).
This set of circuits were selected to be representative of different interaction graphs,
and to explore the effects on mostly parallel circuits (Generalized Toffoli) versus mostly
serial circuits (Cuccaro Adder and QRAM). Each of these circuits is tested across a range
of different sizes to examine how these different strategies scale as the size of the circuit
increases.

96
(a) Cylinder Graph
(c) Binary Welded Tree
(b) Torus Graph

Figure 4.17: Examples of graph-based circuit interaction graphs.

These compilers were constructed on top of the Qiskit library, version 0.18.3 [4], on
Python 3.9 [113]. The benchmarks were run on a machine with Intel(R) Xeon(R) Silver
4110 2.10GHz, 132 GB of RAM, on Ubuntu 16.04 LTS.

4.5.8 Results

Two benefits of encoding into ququarts are the decrease in required gates due to reduced
communication requirements and the transformation of some two-qubit gates into much
faster single-ququart gates. In Figure 4.18, we examine how the gate expected probability
of success (EPS) compares to qubit-only compilation for each compression strategy on the
same architecture.
Primarily, FQ (orange line, baseline) is consistently worse than our qubit-only baseline.
This is expected. Every out-of-pair operation requires more communication, plus the decode
and encode steps. As a result, we actually see increases in the number of gates, as well as a
decrease in the overall circuit gate fidelity.
Certain compression strategies are more advantageous than others on different circuit
constructions. In the more regularly structured circuits, namely CNU and Cuccarro adder,
we see the greatest gains in circuit fidelity due to gates from EQM (blue) and RB (red)
strategies, with improvements over 50% for both. In fact, these gains match or exceed the

97
Figure 4.18: Expected Gate Probability of Success for each benchmark. Each color represents
a different compilation strategy, where the black line is the exhaustive solution developed
previously, and is the goal. FQ is the previous baseline for generalized ququart computation.

EC (black, ideal) case, which requires much more classical computation and is impractical
for even moderately sized programs. These circuits have focused interaction on varying
sets of qubits over time, as seen in Figure 4.16. The circuit interaction graphs are very
easy to flatten into a line by compressing qubits within cycles, eliminating the need for
communication. Then only CX operations are needed.
However, the RB strategy is less consistent for BV and QRAM circuits. For BV, this
makes sense; there are no cycles to examine in the interaction graph, so no compressions
are made. But, QRAM has many cycles. In this case, because the cycles share edges rather
than nodes, a compression on one cycle may adversely affect others. Predicting interactions
between the cycles is more computationally difficult and was not explored. However, EQM
can more closely mimic the interaction graph’s connectivity on the architecture, reducing
the number of SWAP gates and increasing internal CX interactions.
In the graph-based circuits, where each edge is weighted similarly, no method clearly wins
over the other strategies and we only find up to 20% improvements in gate success rate. The
most consistent performer is still EQM, which almost never drops below the corresponding

98
qubit compilation success rate. On the other hand, strategies that prioritize communication,
such as AWE and PP, are much more inconsistent. These strategies group qubits together
that reduce the distance to interact with more qubits. While this might seem to decrease
communication, it does not. Instead, qubits are constantly shifted into positions where
there is not much locality to exploit. That is, highly-interacting qubits are not necessarily
placed close to one another, increasing the chance that communication will be required.
These methods significantly reduce the number of internal ququart operations and increase
serialization of communication due to both qubits in a pair being required independently
for communication. This is shown in Figure 4.19. The overall numbers of gates are similar
between EQM and AWE, but EQM uses significantly more internal CX gates (red bars).
AWE and PP tend to use more SWAP gates (grey, green, and cyan bars) and many more
partial CX operations (orange and blue bars). EQM enables more success rate improvement
by prioritizing higher fidelity multi-qubit operations rather than prioritizing communication
reduction.
While communication reduction is important, it is difficult to predict prior to compres-
sions being made and recompilation, and it is more crucial to find the compressions that
will increase internal CX count early in the circuit with a general sense of global scope, and
leave the router to dynamically adapt to changing communication costs.

Sensitivity to Better Qubit Error It may not be the case that ququart and qubit gates
have identical error. As mentioned previously, ququart gates are more difficult to control and
may have lower fidelity in a real system. In Figure 4.20, we demonstrate how strategies react
to higher fidelity qubit-only gates for the Cuccaro circuit and for the cylinder QAOA. The
strategies maintain their relationship to each other, but, as expected, see diminishing returns
as qubit error improves. However, there is some variability in the pairing methods as the
qubit error rate improves, as they attempt to preference qubit operations as they become less
error prone. Even in these cases, the underlying architecture may be space limited meaning
99
Figure 4.19: The distribution of gate types for different pairing strategies for a 30 qubit
Torus QAOA circuit. Each color represents a different “style” of gates. In particular the
darker blue represents CX gates between two ququarts, and the red represents an internal
CX gate within a single ququart.

100
Figure 4.20: Gate Expected Probability of Success as the qubit gate error rate increases and
the ququart error gate rate stays constant. The black line represents the crossover point
where the error of the qubit only compilation is greater than ququart compilation.

we require larger hardware than required by a mixed-radix strategy.

Error Due to Circuit Duration Ququart compressions have a downside. Each com-
pression induces serialization and requires the use of the longer partial ququart gate times.
In Figure 4.21, we explore how each compression strategy affects the error due to increasing
circuit duration. As the circuit duration increases, a qudit is more likely to decohere due to
the significantly worse T1 time. The probability of an entire system maintaining coherence
for a mixed-radix circuit is described as the product of e−tqb /T1qb −tqd /T1qd for each qubit. tqb
is the time spent in the qubit state and tqd is the time spent in the ququart state, and T1x is
the T1 coherence time for the qudit in the noted state. We first notice that we significantly
improve upon the time incurred by FQ; all other compression strategies are able to more
effectively mitigate circuit duration increases. Partial SWAP operations are faster than full
101
Figure 4.21: Expected Coherence Probability of Success for each benchmark. Each color
represents a different compilation strategy, where the black line is the exhaustive solution,
and in theory ideal, developed previously. EC line stops short for computational reasons,
requiring many more classical resources. FQ is the previous baseline for generalized ququart
computation.

qudit SWAP operations and prevent extraneous communication. We also find that EQM-
based compression generally leads to the best coherence probability of success. Furthermore
we observe that, in some cases, the highest gate probability of success does not always match
the best coherence probability of success. This is due to the fact that some single-qubit and
multi-qubit gates can no longer be performed in parallel due to compression, significantly
contributing to the circuit duration. We also notice that we are able to mostly match the
duration found through exhaustive search of critical path success rates.
However, a total success rate is the success rate via gate fidelity product times the
coherence time success rate, and at current T1 times (listed in Section 4.5.7) decoherence
error outweighs the benefits of success rate increases. We examine the coherence probability
of success for a Cuccaro circuit and Torus QAOA in Figure 4.22, with a 10x better T1
time for both qubits and ququarts. While the margin between qubit and ququart circuits
improves, it will still outweigh the gate success.
However, the 1 : 3 ratio of T1 times used in this work is a worst case scenario. Two
102
Figure 4.22: Expected Coherence Probability of Success for Cuccaro and Torus QAOA with
10x better ququart and qubit T1 times.

qubits encoded in a ququart may not be in the maximum |3i state at all times, and will not
be subject to the same loss in T1 time at all points. Additionally, physically realized qutrit
devices have seen T1 relationships that are better than the expected 1/2 reduction [15] and
can be specifically designed to enhance the expected decay rates for different energy levels.
Using the circuit durations found here, we plot the change in success rate due to circuit
duration as the ratio of T1 time changes in Figure 4.23. As the T1 for ququarts increases,
we find that there is a point, the dashed lines, before the T1 times are equal where the total
success rate is improved with ququarts. In these instances, ququarts would be expected to
perform better than qubits. While it is difficult to avoid an increase in circuit duration,
with enough gains through internal CX gates and SWAP count reduction, encoding qubits
into ququarts has the ability to extend what can be computed on a device and perform with
better expected probability of success.

103
Figure 4.23: Expected Coherence Probability of Success for several 25 qubit benchmarks
with 10x better ququart and qubit T1 times as the ququart time increases T1 from 1/3 the
qubit T1 time to the entire qubit T1 time. Each dashed line represents the point where the
coherence success rate no longer outweighs the gate success rate gains for the corresponding
circuit.

104
Figure 4.24: Ranges of gate based probability of success for CNU and Cylinder QAOA on
three different architectural topologies. This is the combined set of ratios of improvement
for circuits sizes 5 to 40.

Computation on Lower Connectivity Most quantum architectures, especially super-


conducting devices, have much lower connectivity than the grid architecture we assumed. As
described in Section 4.5.5, we also test our methods on the IBM Ithaca topology and a sim-
ilarly sized ring-based topology. We describe the range of critical path fidelity improvement
for some circuits in Figure 4.24. The patterns shown are consistent across all benchmarks.
We do not see any significant difference in performance between architectures. This is ex-
pected. We use the similar routing steps for both qubits and ququarts, the main difference
being that the connectivity is slightly expanded for the ququart routing. Our methods can
successfully adapt to different structures with similar effects for each.

4.5.9 Discussion

Architectural size is a huge limitation on useful computation on quantum computers, and


many architectural models for quantum architectures have access to higher level states. These
higher level states can be used to compress quantum information. In the case of ququarts,
two qubits can be fit into a single physical unit in a process we call compression. There are
difficulties in working with these more complex objects such as shorter coherence times, and
105
longer gate times work against each other to make ququart compilation challenging.
In this section, we adjust several different aspects of the quantum compilation pipeline.
We explicitly realize a gate set for mixed qubit-ququart systems including partial operations,
without the need to encode and decode ququarts. We also demonstrate the viability of this
encoding scheme and gate set via the execution of a H ⊗ H gate on real superconducting
hardware. With a gate set designed specifically for ququart computation and communication,
we design and evaluate a compiler to mitigate time-intensive ququart-ququart interactions
and the increased gate-based communication required to interact two ququarts in past mod-
els. Our compiler also exploits higher qubit connectivity and reduces the count of needed
physical units to execute programs requiring more qubits than available. By adapting our
standard compilation strategies to take advantage of the benefits of higher radix computa-
tion, we are able to mitigate this cost and explore the best way to use of faster operations,
such as internal CX gates, enabled by encoding multiple qubits in the same physical unit.
We develop several strategies that emphasize the need to prioritize these fast interactions
and exploit the locality of certain circuit structures for better gate expected probability of
success, and lower increases in circuit duration.

4.6 The Quantum Waltz: Exploiting Mixed-Radix and

Full-Ququart Operations for Native Three-Qubit Gates

4.6.1 Introduction

As we’ve seen, success of quantum algorithms greatly depends on how many error-prone gates
are used and the total program duration. In most currently competitive quantum systems,
e.g. superconducting systems, trapped ions, and neutral atoms, gates which act on many
qubits simultaneously (≥ 3 operands) must be decomposed, contributing to both increased
gate counts and increased circuit depth. In this work, we focus primarily on superconducting

106
Figure 4.25: A comparison of executing a Toffoli gate on a three-qubit-only system versus a
Toffoli gate between a ququart and qubit in a mixed-radix system. In a qubit-only system,
we must use a decomposition that uses eight two-qubit gates that can be reduced to one
two-qudit gate that has a shorter duration.

systems which also suffer from limited connectivity between devices, further exacerbating the
decomposition problem.
The significant increase in the number of operations required to perform a calculation
strains the device in multiple ways. Many circuits include gates, such as the Toffoli gate, to
perform reversible arithmetic calculations; thus, three-qubit operations are common across
implementations of quantum algorithms [7, 46, 28, 43]. Finding implementations of these
gates without having to reduce them to more elementary gates could be expedient to saving
the valuable computational resources of the device.
We’ve explored several uses of higher-radix computation, intermediate-qudit computa-
tion and qubit compression into ququarts with both the upsides of reduced communication
and higher fidelity CX gates on compressed qubits with the downsides of lower coherence
times. In this section, we observe the simple fact that one ququart is equivalent to exactly
two qubits and therefore the information of two individual qubit devices can instead occupy
a single device which has access to four logical states [7]. This has significant advantages
on the relative connectivity of the qubit information. By performing this compression, we
can access three qubits worth of information by only interacting two physical devices. This

107
type of mixed-radix gate (four-level system interacting with an adjacent two-level system) is
equivalent to performing a three-qubit gate. Similarly, we can consider two adjacent ququarts
which allows us to perform interactions on up to four qubits worth of information by con-
trolling only two physical devices; we call these full-ququart gates. This simple observation
could remove the need to perform expensive decompositions of three- or four-qubit gates,
as visualized in Figure 4.25. We are primarily focused on common three-qubit interactions
since they appear more commonly in real applications, unlike four-qubit gates.
We examine using ququarts to dynamically encode and decode gates to perform na-
tive three-qubit gates on ququarts on a simulated superconducting device in a compilation
pipeline called the Quantum Waltz, a dance done in three-four time. In particular, the major
contributions of this section are:

• A library of two- and three-qubit gates performed on mixed-radix and full-ququart


schemes, synthesized via quantum optimal control for a realistic Hamiltonian for a
superconducting architecture.

• Identifying specific relationships between the controls and targets of three-qubit gates
that allow for more efficient execution of mixed-radix and full-ququart gates.

• Detailing a compiler that choreographs three-qubit gates into particular configurations


on ququarts for better performance by managing the controls and targets of three qubit
gates.

• Simulating a wide variety of benchmarks compiled with different strategies.

• Demonstrating how three-qubit gates on ququarts can achieve a 2x improvement in


simulation in a mixed-radix environment and up to 3x fidelity improvements in a full
ququart environment, as well as insights into the right situations to implement these
gates.

108
4.6.2 Three Qubit Gates on Ququarts

We use the same compression strategies developed in Section 4.5.2, and use the same pulse
generation techniques as well. The gates described here are an extension to a native gate
set for ququarts to avoid decomposition of three-qubit gates to longer strings of one- and
two-qubit gates.

Connectivity Advantage

The set of two qubit gates laid out in Section 4.5.2 are enough to universally perform general
qubit computation on ququarts [71, 83], but simply compiling to two-qubit gates would not
take advantage of the flexibility of this abstraction. When we think about a ququart as two
encoded qubits, we virtually increase the connectivity between qubits on our device. An ex-
ample of this is seen in Figure 4.26. Each of the encoded qubits in a ququart is connected to
an adjacent qubit, or both of the encoded qubits in an adjacent ququart. As highlighted by
each of the different colors in Figure 4.26, this creates many triangle subgraph relationships
between encoded qubits that would not exist otherwise. This is significant, as triangle sub-
graphs are uncommon in current hardware due to the increased probability of crosstalk as the
density of interaction increases and frequencies overlap [76, 32]. But, triangle-based interac-
tions are common in many different circuits that use three-qubit gates to perform arithmetic
calculations. Here, we increase the number of virtual connections without increasing num-
ber of physical connections to create four virtual triangles, and four interactions between
encoded qubits, something that isn’t easy on current hardware topologies.
The triangle relationships existing across two physical units gives rise to another impor-
tant phenomenon. It is not fundamentally harder to interact three or four qubits worth of
information than two qubits worth with a single operation. These gates are equivalent to
either mixed-radix or full-ququart gates. For example, if we have a fully encoded ququart
next to a bare qubit and perform a Toffoli gate targeting the qubit, it is equivalent to a

109
Figure 4.26: Visualization of connectivity advantages in qubit-ququart systems. Encoding
qubits in ququarts (lightblue) enables triangle connectivity between triples of qubits, where
two of which are encoded in the same ququart and one appears either in a bare qubit or
encoded in a neighboring ququart.

|3i-controlled X gate on the qubit. This is computationally simpler in terms of number of


state transitions, than the several |1i- and |3i-controlled X that would be required in the
decomposition and can be seen in the state evolutions in Figure 4.27. This gate implemen-
tation gives superconducting qubits more natural access to the native multi-qubit gates.It
also avoids decompositions that add a significant number of extra gates.

Pulse Generation

Table 4.3: Mixed-Radix and Full-Ququart Three-Qubit Gate Durations

(a) Mixed-Radix (b) Full-Ququart


CCXq01 619 CCX01,0 536 CCX01,1 552
CCX 1q0 697 CCX0,01 785 CCX0,10 785
CCX 01q 412 CCX1,10 785 CCX1,01 680
CCZ01q 264 CCZ01,0 232 CCZ01,1 310
CSWAP 01q 684 CSWAP01,0 680 CSWAP01,1 744
CSWAP 10q 762 CSWAP10,0 758 CSWAP10,1 822
CSWAP q01 444 CSWAP0,01 510 CSWAP1,01 432

Multi-control Gates Just as a native two-qubit gate on one physical ququart offered
significant improvement in both fidelity and execution time as compared to the standard
110
(a) CCX01q state evolution (b) CX0q state evolution

Figure 4.27: Visualization comparing the evolution of a |3i-controlled X gate in a mixed-


radix environment for a CCX gate in (a) and a CX gate in (b).

two-qubit gate on two physical qubits, native three-qubit gates on two physical units have
the potential to offer a significant improvement in gate fidelity and execution time. In
Table 4.3, we show pulse durations for different orientations of the three-qubit Toffoli gate
in both mixed-radix and full-ququart configurations. These gates were synthesized using
the same fidelity targets and pulse generation techniques as the two-qubit gates, therefore
resulting in an overall higher fidelity than if decomposed with many gates of the same target
fidelity. After synthesizing the different configurations of Toffoli gates, we find that there
is a substantial difference in the gate duration depending on which qubits are controls and
which is the target.
Consider the mixed-radix example where both control qubits are encoded in the same
ququart, and the target qubit is in the bare qubit, or CCX01q , seen in Figure 4.28(b). This
configuration is about two-thirds the time of the CCX0q1 , seen in Figure 4.28(c), where the
control qubits are split across the bare qubit and the ququart, while the target is encoded
in the ququart as well. The reason for this difference is twofold. The first follows from the
two-qubit only gates. Gates which use the ququart as a control and the qubit as a target

111
are generally faster. In this case, the pulse only needs to induce population changes between
the |0i and |1i state of the qubit, rather than between the transitions between the pair |0i
and |1i, and |2i and |3i of the ququart. The second is that the entire ququart acts as the
control, only changing the state of the bare qubit if the ququart is in the |3i state. In the
split-control case, the ququart must control on both the |2i and |3i state.
The same concept of separation of controls and targets follows for the full-ququart Toffoli
gates as well. Regardless of whether the target qubit is in the first or second encoding of the
ququart, CCX0,10 or CCX0,11 , it is substantially faster to keep the controls encoded in the
same ququart with the target encoded in a separate ququart.

Target-Independent Gates Separating the controls and targets into different devices
yields more efficient gate execution; however, compiling circuits to conform to this configu-
ration is unnecessarily constraining. Instead, we consider a situation where all multi-qudit
gates are target-independent and only affect the global state when all three qubits are in
|1i. For example, the Toffoli gate, or CCX, is locally equivalent to CCZ which is target-
independent, as seen in Figure 4.29.
When pulses are synthesized, CCZ is much more efficient as seen in Table 4.3, remark-
ably on par with the speed of the qubit only gates. In addition, we only need to define three
configurations: CCZq,01 , CCZ0,01 , CCZ1,01 , as opposed to the nine possible CCX configura-
tions, reducing computational overhead. We postulate the short duration of these gates is
because CCZ only changes the phase of the entire three-qubit state and does not change the
population of any state. A population change for the Toffoli gate is more complicated. This
makes the CCZ (and generally target-independent gates) a valuable tool when compiling
three-qubit gates. In practice, this removes the need to distinguish targets from controls.

Multi-target Gates We also consider gates that use one control qubit to affect the state
of some number of other qubits, for example the CSWAP. With our methods we synthesize

112
Figure 4.28: Examples of two-control and two-target three-qubit gates in a mixed radix
environment. a) The Toffoli with two controls q0 and q1 . b) A configuration where both
controls are encoded in the ququart and the target is mapped to a qubit. c) A configuration
where the controls are split across the qubit and the ququart and the target is encoded in
the ququart. d) The CSWAP gate with two targets q1 and q2 . e) A configuration where both
targets are encoded in the ququart and the control is mapped to the qubit. f) A configuration
where the targets are split across the qubit and the ququart and the control is encoded in
the ququart.

113
• • •
• = H X H = •
X H • H H • H

Figure 4.29: Different decompositions for the Toffoli Gate. a) is the base Toffoli circuit. b) is
Toffoli circuit with a swapped second control and target from the original. By surrounding
the control and the target with Hadamards, we perform the same operation. c) The Toffoli
gate constructed from a CCZ gate which can be used as a Toffoli by surrounding the target
with Hadamard gates. (d) The Toffoli gate constructed from an iToffoli gate, which requires
a controlled S † gate in addition to Hadamard gates.

gates to the same fidelity targets as before and show their times in Table 4.3. As expected,
we find benefits when separating the control qubits from the target qubits as depicted in
Figure 4.28(e) versus Figure 4.28(f). When both targets are encoded the same ququart,
we limit the possible state changes to be between |1i and |2i in that ququart, rather than
needing an expensive mixed-radix or full-ququart based qubit SWAP gate.
In this work we show these gates improve the execution of quantum circuits compared
to traditional binary-only strategies. While we are able to perform any configuration of
three qubit gates directly in mixed-radix or full-ququarts scenarios, we take care to use best
configurations to minimize time in the less stable |2i or |3i states.

4.6.3 Adapting Qubits on Ququarts to Three-Qubit Gates

To execute three-qubit gates (like Toffoli or CCZ), circuit qubits must be routed into a
connected subgraph of the interaction graph, e.g. for CCZ(q0 , q1 , q2 ) requires q0 ∼ q1 and
q1 ∼ q2 but it is not guaranteed that q0 ∼ q2 , where ∼ defines adjacency. We need to adapt
our original strategy to fit within a framework that can best take advantage of routing three-
qubit gates together, when they have varying cost based on their configuration between two
physical devices.

114
4.6.4 Using Three-Qubit Gates

Qubit-Only In a qubit-only regime we cannot efficiently execute a native three-qubit gate.


In these cases, we use a decomposition into eight CX operations [101]. This decomposition
has the flexibility of being target-independent from a compilation standpoint.
Alternatively, we can use a directly-optimized three-qubit pulse sequence. QOC software
failed to find a solution for a direct CCZ operation, so we synthesize a pulse implementing the
iToffoli gate and use the decomposition shown in Figure 4.29d inspired by [60]. However, as
seen in the decomposition, the iToffoli gate requires an controlled-S † , and in most topologies
as extra SWAP to perform the controlled-S † .
This is an expensive compilation, as it requires eight two-qubit gates, and 14 one-qubit
gates. But it does not require using the less stable |2i and |3i state. Other, cheaper decom-
positions of CCX and CCZ require complete connectivity between the operands.

Intermediate Mixed-Radix We also permit temporary use of the higher energy levels to
perform an operation. By performing an encoding gate (ENC) followed by the three-qubit
gate and a final decode (ENC† ) operation, we get temporary access to the full connectivity
needed on a shared physical connection to perform fast three-qubit gates.
As noted, the compiler should opt to encode qubits of similar type, i.e. both controls
together or both targets together. Let U (q0 , q1 , q2 ) be the operation with q0 , q1 the same
(either both controls or both targets). In some cases, encoding is simple because the routing
strategy (prior work) results in A ∼ B as in 4.28(b). However, it may fail to do this by
default and we may have q0 ∼ q2 and q1 ∼ q2 but not q0 ∼ q1 and in 4.28(c).
In this case, we have three options to compile to a favorable configuration. First, we could
enforce the ideal relationship through additional gates by adding an additional SWAP(q0 ,
q2 ) (or equivalently SWAP(q1 , q2 )). Second, in the special case where U = Toffoli we can
change which pair is the same type with Hadamard gates as in Figure 4.29 to use the most

115
efficient implementation; we call this re-targeting. Third, if U permits, we transform U into
U 0 so that q0 , q1 , q2 are all the same type; for example we transform CCX to CCZ so each
operand is a “control.” While the additional re-targeting or transformation gates add both
error and duration, they enable the shortest duration version of U to be used for an overall
net increase in fidelity. In this work, we consider the special cases of U ∈ {CCV |V ∈ SU (2)},
i.e the set of locally equivalent gates to CCX. We leave the generalized case to future work
in circuit synthesis. For the multi-target situation there is no easy fix and SWAPs must be
inserted.

Full Ququart Many of the compilation strategies laid out for mixed-radix three-qubit
gates can be used in full-ququart compilation as well. However, some configurations are
extremely unfavorable (minus the target-independent CCZ) for the execution of both two-
qubit and three-qubit gates as seen in Table 4.3. To remedy this, we can perform an internal
SWAP to exchange the encoding positions first.
The router by default, described previously, does not distinguish control or target. How-
ever, when executing three-qubit gates, we ensure only qubits of the same type are encoded
if it does not require an extra swap operation.
In a compiler, any of these techniques could be put together to decide the best decom-
position for a Toffoli based on the current orientation and placement of the qubits in the
mixed-radix or full-ququart system.

Adapting Mapping and Routing

We only need to adapt the mapping and routing algorithms described in Section 4.5.5 where
we considered the fidelity differences as a distance metric between different routing paths.
In this case, we need to ensure that we are able to generalize to three-qubit gates.
The first step is to decompose the operations in the circuit to native gates supported by

116
the device. As our compiler will need to handle the native execution of three-qubit gates,
we decompose to the CX, CCX, CCZ or CSWAP along with a parameterized single-qubit
rotation gate based on which gate set and architecture we are studying. This set is enough
to support universality, and any gate can be decomposed to some combination of these gates.
Our mapping strategy does not change, and we still only weigh each three-qubit gate
as if it is one operation. When routing, we track the circuit qubits on the interaction
graph and use SWAP gates until the interacting qubits are adjacent. We still attempt
to disrupt advantageous qubit layouts as little as possible by using adaptive weights that
change as operations are scheduled based on [10]. This strategy attempts to keep qubits
interacting in the near future close to one another. To generalize to three-qubit based
routing we modify the cost function to a sum over all qubits in the operation, C(i) =
P −1
j∈Qo /i D(i, ϕ (n))(d(ϕ(i), ϕ(j)) − d(n, ϕ(j)) where Qo is now a set of all operands.

4.6.5 Evaluation

Baselines

We compare against two strategies. The first is a compilation that routes the circuit with
three-qubit gates, before decomposing them to one- and two-qubit gates only. This is in
line with current practices for most compilation pipelines. The second baseline does not
decompose to these smaller gates. Instead, the iToffoli-based decomposition directly on
qubits similar to [60]. This is a more challenging gate to synthesize as discussed previously.
For simulation, this gate has a 99% fidelity and with 912 ns duration. The CS † gate has a
99% fidelity and a 126 ns duration.

117
Circuits

We primarily study benchmarks which already include three-qubit gates in their formulation
and which can be parameterized by number of qubits to study scalability of our methods
We examine five circuits with different constructions. The first is the Generalized Toffoli
(CNU) circuit [9], which flips the state of a target qubit if all the controls are one. This
circuit uses exclusively Toffoli gate based decomposition and is highly parallel. The Cuccaro
Adder [28] is nearly entirely serialized using 2n + 2 qubits with a mix of three-, two- and
single-qubit gates to add two n-bit numbers. Third is a QRAM circuit which uses primarily
CSWAP gates to retrieve data from or move data into a set of qubits [43]. The fourth is a
Select circuit, which is a preparation mechanism used in Quantum Phase Estimation (QPE)
[69]. It performs a particular Pauli operation on n qubits for each potential 2m states of
m index qubits [6]. For our case, the choice of Pauli string does not affect compilation.
To keep the fidelity of circuit simulation within comparable bounds, we only select on two
random values rather than all of the potential 2m values the index qubits could be in. The
fifth is a purely synthetic circuit to study relative strength of our architecture on potential
distributions of CX versus CCX gates.

Hardware Topology and Error

For this section, we consider the same underlying hardware topology for each comparison
point - a 2D mesh. This type of grid architecture has relative density on the upper end of
realized superconducting connectivity graphs, reflective of Google’s Sycamore chip [5].
This is a rather optimistic connectivity structure, and not fundamentally different from
the quantum heavy-hex processors designed by IBM [40]. Here we consider a grid design

with dimensions d ne × d√nne with connections to each qudit adjacent to it on the device.
An important feature of a quantum architecture is the coherence time, which is the limit
for how long a qubit is able to maintain its state. We use a realistic T1 time from an IBM

118
device of 163.45µs [2]. Higher energy levels decohere more quickly. In theory, each state
decays at a rate of 1/k where k is the energy level. We therefore use 81.73µs and 54.15µs
as the T1 times for the |2i and |3i states.

Circuit Estimation

In addition to circuit simulation, we use two metrics to estimate the fidelity of a circuit
without simulation. These will give us insight into how compiled circuits may perform by
comparing direct simulation to our estimated fidelity.
The first is the product of all of the gate success rates in the circuit, called the gate
expected probability of success (gate EPS). Since there are multiple classes of multi-qubit
gates, some of which have higher fidelity than others, we use the product of these success
rates.
The second is more complicated, as it attempts to capture the chance that the state of
a qudit will collapse over the course of the circuit. We model this as an exponential decay
where the probability of no decoherence is 3k=1 exp(k ∗ tk /T1 ) where tk is the time the
Q

qubit spends in state k. When we construct the circuit we keep track of how long each qudit
exists in the |1i state as the maximum state, and how long it exists in the maximum |3i
state and calculate the probability of not decohering over the course of the execution for
each qudit. The product of the expected success of each qudit is the EPS due to coherence
for the entire circuit. When multiplied by the gate EPS, we have the EPS for the entire
circuit solely through estimation.

Circuit Simulation

We want to go further than estimation. Since access to devices that can use ququarts at this
scale are limited, we must use simulation. In general, simulation of quantum systems in the
presence of noise with a classical computer is exponentially difficult since for a general pure

119
quantum state containing n qudits requires dn complex numbers and for non-pure states
(which is acquired in noisy quantum systems) requires the representation of a density matrix
ρ of size dn × dn . In either case, at each step of the simulation we repeatedly perform matrix
multiplication with dn ×dn matrices for each time step in the circuit. While the more generic
simulation method is more precise, it is not feasible outside of high performance computing
settings and even in this case it is bounded and requires specific initial conditions [122].
A slightly more scalable alternate approach is the so-called trajectory method which
requires us to store only a dn sized state-vector [20]. The trajectory method is a stochastic
approach and instead of maintaining the entire density matrix we begin with a random
pure state, and apply gates as normal but inject randomly drawn coherent and non-coherent
errors into the system at each step and the resulting quantum state remains pure after
renormalization. In the limit, the trajectory method converges the same results of full
density matrix simulation but has some space savings which allows us to simulate a few
extra qudits. This work simulates circuits of up to 24 qubits, which can be represented by
12 ququarts using this method.
For this work, for each circuit of interest, we generate at least 1000 random quantum
states and for each we simulate once and compute the average fidelity over all trials and all
random states. We emphasize the need for the random quantum states rather than beginning
with random classical states (e.g. bitstrings) even if many of our circuits are classical in
nature because quantum errors do not always have noticeable effects on classical inputs (e.g.
a Z-type Pauli error will have no effect on the measurement outcome of a classical bitstring).
In the past, prior work on simulation of qudit systems neglects the realistic duration
differences between gates which results in drastically different usage patterns and simply
injecting errors on a moment-to-moment basis can skew results. For example, in this work
our direct-to-pulse compilation of CCX and CCZ gates have significantly different execution
times. In a naive execution of the trajectory method, circuits with each of these perform

120
similarly despite this execution time difference because the total number of moments is
nearly the same. We modify the trajectory method simulation slightly to account for this
difference. Rather than inserting many idle gates during each time step using the longest
gate time included in that time step, before each gate, we insert one idle gate using the exact
time that qudit has been idle. This is a more accurate representation of how long, when,
and from which state these qudits could be decohering.

Noise Model for Qudit Systems

For qubits we consider both symmetric depolarizing and amplitude damping errors. There
are four possible single-bit channels: no error (I), bit flip errors (X), phase flip errors (Z)
and bit and phase flip errors (Y = ZX). In simulation each error channel is drawn with
probability p/3. Two-qubit errors are given as the product of single-qubit errors, e.g. X ⊗ X
for a bit flip on both interacting qubits; there are 16 possible channels of this type so each
error occurs with probability p/15 and no error (I ⊗ I) occurs with probability 1 − 15p.
For a general qudit system, we consider a generalized form of these errors. The “bit-flip"
type gates become X+1mod d and the “phase-flip” errors become
 2 n o
Zd = diag(1, exp{ω}, exp ω , ... exp ω d−1 ) where ω j is the j − th root of unity. The
d−1
product of {I, X+1mod d , ...X+1mod 2 d−1
d } and {I, Zd , Zd , ..., Zd } is a basis for all d × d Pauli
matrices which allows us to construct a general symmetric qudit depolarizing channel. This
explains the expected increase in error for using qudit systems: For a two-qubit gate the
chance of no error is 1 − 15p while for a ququart this chance diminishes to 1 − 255p let alone
possible differences in p [74].
Amplitude damping for qubits can be described as non-unitary transformations on the
√ √
quantum state with operators K0 = diag(1, 1 − λ1 ) and K1 = λ1 e0,1 . Here ei,j refers to
a matrix with all 0’s except for a 1 in the i-th row and j-th column and is of appropriate
√ √ p
dimension. In the general qudit case we have K0 = diag(1, 1 − λ1 , 1 − λ2 , ... 1 − λd ),

121
√ p
K1 = λ1 e0,1 , ... Kd = λd e0,d−1 . Since we primarily focus on a superconducting system
in this study we take λm = 1 − exp{−m∆t/T1 } where ∆t is the idling duration and T1 is
the coherence time of the qubit [57].
In this work we are also concerned with the manipulation of mixed-radix systems.When
drawing an error for such a system, for example a qubit-ququart interaction, we consider
only relevant errors for the respective participant. For instance, a two-qudit error is drawn
from P2 ⊗ P4 and not from P4 ⊗ P4 (where Pd is the set of d-dimensional Paulis, exactly the
set of potential errors described above). Similarly, for two-qubit gates on encoded qubits, we
consider only single ququart errors since gates on encoded systems are equivalent to single-
ququart gates. Our model does not account for several real problems - for example crosstalk
and population leakage - as these are challenging to incorporate into simulation. These are
problems that both qubits and qudits experience, though there is an expectation that leaking
in particular is worse for higher dimension d.

4.6.6 Results

When we are able to perform native implementations of three-qubit gates via ququarts, we
significantly reduce the number of gates that need to be executed, reducing failure due to the
use of many error prone gates. However, the reduction in time to execute these gates is not
enough to overcome the reduced coherence time of higher radix states. In Figure 4.30, we
examine the simulation fidelities for three-qubit compilation strategies across different sized
circuits using Toffoli gate based decompositions. Each point represents the average fidelity
of 1000+ different initial states run once, each with randomly inserted error. The error bars
are the standard error, which is the standard deviation of all the trials divided by the square
root of the number of trials. The mixed-radix compilation schemes stop at 12 qubits due to
memory based computational limitations. While mixed-radix circuits start in an all-qubit
state, we must model them as if they are entirely on ququarts, since to accurately inject error

122
Figure 4.30: Simulated results for QRAM, Generalized Toffoli, Cuccaro Adder and Select
Circuit from 5 to 21 qubits with different mixed-radix and full-ququart compilation strate-
gies. The mixed-radix strategies do not have complete error bars due to the requirement
to simulate a four-level system for every qubit which would require more than 86 GB of
memory per circuit in our simulation framework. The final graph is the average fidelity
improvement for each compilation method over the qubit-only compilation method as the
size of the circuit increases.

into the system we must be able to model the higher levels at all times. This restricts the
number of physical devices we are able to simulate in this scheme to 12 simulable ququarts.
The first thing we notice is that all of our mixed-radix and full-ququart compilation
strategies are able to exceed the fidelity of our baseline qubit-only compilation scheme.
From a pure gate error perspective, this should not be unexpected, each of these schemes
greatly reduces the number of gates required to execute the same operation. Figure 4.31
demonstrates how the EPS for gate error is substantially improved by using three two-
ququart gates gates or one two-ququart gate for full-ququart computation rather than eight
two-qubit gates for mixed-radix computation. And, as hoped, the simulation finds that using
the |2i and |3i state, and the idle time potentially spent in these states, does not outweigh
the benefits of the using fewer gates and the shorter circuit duration. The shorter duration
of the gates is counteracts the increased decoherence rate of the ququarts. Figure 4.31 also
demonstrates the same point. The coherence EPS between all of the mixed-radix strategies
and the qubit-only baseline are nearly the same, and is actually improved for full-ququart
strategies. The general trend of our simulation results is mirrored in Figure 4.31 where the
total EPS of the circuit is shown. As the EPS trends match what we find in our simulated

123
Figure 4.31: EPS statistics for the generalized Toffoli circuit. We show the gate and coherence
EPS on the left and the product EPS on the right.

results, we are able to infer that the scaling of simulation will match the scaling of the
EPS results, which we can calculate when simulation becomes intractable. While we only
show examples of the generalized Toffoli circuit, the results are similar for other circuits.
However, we note that the mixed-radix strategies only marginally outperform or match
the simulated results of the qubit-only iToffoli based strategy. This makes sense when we
examine the iToffoli based decomposition. We must insert an extra SWAP gate to perform
the corrective Controlled-S gate, resulting in a similar number of gates, and duration, for
both decompositions. Additionally, this SWAP results in extra corrective SWAPs later on.
The extra communication disrupts the layout of the circuit further than was intended, and
can require extra gates.
Digging into the difference between the strategies, we find that the mixed-radix strategies
are all relatively similar to one another, with some additional separation as the size of the
circuit increases. We first compare the mixed-radix Hadamard corrected CCX gates, shown

124
in light blue, to the mixed-radix gates without this correction, shown in pink. While there is
some cancellation between the single qubit gates, the extra serialization and marginal gate
error of the correction gates is a drawback to using this correction strategy based on the
simulated results. The reduction in time from the better configuration of the CCX gate is
not always enough to overcome these additional costs. If we instead use CCZ decomposition,
shown in green, we consistently achieve the same, or better, fidelity, especially as the size
of the circuit increases and the reduction from CCZ gates is more pronounced. In these
cases, the benefits found by using shorter target-independent gates from the start rather
than retargeting improves the fidelity of the circuit in this mixed-radix regime. In Figure
4.30d we find that the mixed radix gates achieve 2x better fidelities for circuit size 12, which
is a significant improvement over qubit-only computation. This alone would be an important
optimization for three-qubit gate based circuits.
We also find that the ququart compilation scheme, shown in grey, has an even higher
fidelity improvement, up to 3x reductions as seen in Figure 4.30d. The reasoning behind this
is two-fold. The first is that we no longer need to encode and decode gates before each three-
qubit gate in this scheme. As seen, gate reduction is important, and this reduces the number
of gates. The second is the reduction of communication. With the higher connectivity
afforded at all times, we further reduce qubit communication required to perform certain
gates. Both of these factors add to the reduction in time, keeping the full-ququart based
circuits under the coherence limits and maintaining higher circuit fidelity overall. At this
level we are able to further reduce the overall circuit time by using faster, target-independent
CZ gates in place of CX.
It is worth noting the points where full-ququart compilation does not outperform mixed-
radix compilation to the same degree. This is particularly noticeable in the QRAM circuit.
There are more than double the CX gates as Toffolis in this circuit. As a result, the serializa-
tion induced by ququarts with slower two-qubit gates reduces the effectiveness of ququarts.

125
This will be explored later in a sensitivity study. Additionally, these benchmarks are only
kernels of computations that could be used within the context of larger circuits. In such
cases, we will not have the benefit of a perfect mapping to start. This would not affect the
improvement in fidelity from the qubit only to the mixed-radix strategies, but the effort to
encode the qubits into a full-ququart regime before execution may outweigh the benefits of
ququart compilation.

Special Gate Case Study: CSWAP

As detailed in Section 4.6.2 we could instead decompose to a different-three qubit gate in


the circuit. In the case of QRAM, this is the CSWAP gate. In Figure 4.32a, we explore the
differences in fidelity when we use CSWAP gates alongside the original results using CCZ
gates. A CSWAP can be constructed from two CX gates and one CCX gate, but cannot be
re-targeted in the same way. Regardless, in the mixed-radix state, by orienting the CSWAP
such that the targets are separate from the controls when possible, we see improvements
over the CCZ decomposition. In fact it is able to beat the full-ququart CCZ compilation in
some cases because of the reduced number of CX gates. While we can always attempt to
encode the qubits favorably in a mixed-radix environment, this is not as natural a change
when compiling for ququarts, and could lead to bad configurations if we solely focus on the
disruption of qubits on ququarts. If we focus on the CSWAP in a full-ququart regime, the
basic version shown in blue, and instead use the strategy that places the targets in the same
ququart, shown in bright pink, we find even more improvement to our full-ququart encoding
regime. This further indicates the importance of using the best decomposition possible by
separating the targets and the controls of certain gates.

126
Figure 4.32: The results of several sensitivity studies. (a) Changes in CCZ compilation strate-
gies’ fidelities as gate error ququarts increases. (b) Changes in CCZ compilation strategies’
fidelities as coherence error for the |2i and |3i level states changes. (c) Differences in fidelities
between mixed-radix and full-ququart compilation strategies as the distribution of CX gates
to CCX gates in a circuit changes. In all graphs The black line represents the qubit-only
compilation method. Below this point mixed-radix or full-ququart methods are more error
prone than using only qubits. Please note the different scaling on the y-axis.

Sensitivity to Ququart Gate Error Rate

While we synthesized our gates using a realistic hamiltonian, it is still more difficult to
physically realize gates that access higher energy levels. In Figure 4.32b we explore how
the simulated fidelity changes as the error on ququart and mixed-radix gate increases for an
11-qubit Cuccaro Adder. Both strategies see a very fast drop off as the gate error increases,
crossing over the qubit-only baseline fidelity when the ququart error rate is between two and
four times worse that qubit gates for mixed-radix compilation (97% fidelity), and between
four and six times worse for full-ququart compilation (94%). We also find that the iToffoli
strategy outperforms the full-ququart strategy at three times worse ququart gates than qubit
gates as well. While these are still high fidelity targets, it does indicate that we do not need
our three-qubit gates to have the same fidelity as our two-qubit gates for these strategies to
be successful.

127
Sensitivity to Ququart Coherence Error Rate

In this work we selected the expected theoretical decrease in coherence time as the grounding
for most of our simulation experiments. However, physical realizations don’t always meet
reality. Accessing higher energy levels may prove to be more costly in terms of coherence time
due to lack of control, or it may be less of an issue as has been found by some testbeds when
accessing the qutrit state [22]. In Figure 4.32c we demonstrate the effects of changing the rate
that the |2i and |3i levels decohere for an 12-qubit QRAM circuit. The main detail to note
is that as the rate increases, the distance between mixed-radix and full-ququart fidelities
decreases until mixed-radix becomes higher fidelity. Mixed-radix gates do not spend as
much time in the higher level states, so as machines are developed and these levels are more
unstable, it may be better to avoid using full ququart encodings for larger circuits.

Ratio of Three Qubit Gates to Two Qubit Gates

It may not always be the case that the number of three qubit gates greatly exceeds the number
of two qubit gates. Future applications may have a higher mix of two-qubit gates to three-
qubit gates, or may only require a few three qubit gates to perform the desired operation.
In Figure 4.32d, we example how the fidelity of different mixes of two-qubit to three-qubit
gates is affected by compilation using a full-ququart strategy versus a mixed-radix strategy
for an 11-qubit circuit. As the ratio of two-qubit to three-qubit gates increases, it becomes
less and less profitable to use a full ququart encoding. At 80% CX gates it becomes more
profitable to remain in the mixed-radix regime. Using CX gates on ququarts requires more
serialization, since we cannot perform two separate operations on qubits encoded in the same
ququart. This increases the time, and we start seeing the effects of reduced coherence times.
This changes the calculus about when mixed-radix is better than full-ququart compilation.
In cases where we don’t need to use as many three qubit gates, it does not make as much
sense to use ququarts for the entirety of the circuit. While this indicates that quantum

128
Figure 4.33: The results of the different mixed-radix and full-ququart compilation strategies
on random mappings of only qubit-encoding, or only ququart encoding.

circuits that only use two-qubit gates do not benefit from this encoding scheme, we can use
resynthesis tools [123] to automatically insert three-qubit gates into the circuit, such as in
[85]. However, resynthesis can introduce additional error as a perfect direct translation is
not always possible and is better explored in a future work. We also include the iToffoli
strategy in this analysis as well. We find that it matches the mixed-radix strategy, further
solidifying that these strategies have similar performance characteristics.

Random Mappings Most of the benchmarks used in this evaluation are “kernels” of
computation, and are intended to be used as smaller portions of larger circuits. As a result,
the assumption that the compiler will have the opportunity to generate a well mapped
circuit from the outset is not necessarily true. We explore the sensitivity of the different
strategies to random mappings of qubits. In particular, we explore three different setups
in 4.33 for the Generalized Toffoli circuit. The first starts with all data qubits in random
qubits, without encodings. Then we performing routing, and use mixed-radix CCZ gates

129
to perform the operations. The second starts with the same random mapping, but instead,
contracts the mapping such that all qubits are encoded into ququarts. We then perform all
computation on this encoding, before returning the qubits to qubit-only encodings. The third
starts in a random ququart encoded mapping, before performing the computation solely on
ququarts. In this study, we find that mixed-radix and full-ququart compilations are relatively
similar to one another. The encoding and decoding in the mixed-radix case, is a slightly less
expensive set of operations than encoding to ququarts and performing all communication in
this encoding. And, we find that a random mapping on ququarts does not necessarily hinder
efficient execution of the circuit. As explored in the previous sensitivity study, it may not
always be the case that circuits have high percentages of three-qubit gates, the mixed-radix
strategy will be more advantageous in many cases when a good initial mapping cannot be
generated.

4.6.7 Alternate Strategies

While this work is the first we are aware of to explore compilation of higher-radix qudits
for three qubit gates via encoding specifically for superconducting qubits, there have been
studies using existing superconducting qubit technology to execute native three qubit gates.
In particular, Kim et al. [60] and Gokhale et al. [43] have explored driving two connections
between three qubits in a line to perform three qubit gates in a superconducting architecture.
These studies found varying success in doing so. Gokhale developed a technique to execute
two CX gates in parallel in a single CXX gate. These gates did not find improved fidelity,
but achieved similar goals of faster parallel gate execution than serial execution as described
in this work. It should be noted that this was done without explicit calibration for this sort
of operation.
On the other hand, Kim developed a very high fidelity, 98.2%, iToffoli gate between three
superconducting qubits. In this case, both controls induce a state change in a center qubit by

130
driving the connections between the qubits. This gate is performed very quickly with a gate
duration of 392 ns. While a meaningful result, it is difficult to compare this work to our own
as the device has significantly different Hamiltonian than this work. Additionally, this took
significant calibration between each of the three qubits, which can be an expensive process.
Requiring calibration between each set of three qubits on a device is much more expensive
than calibrating each pair of connected devices as would be required by our scheme and is
already done on current hardware. This chapter is the basis for our iToffoli baseline, which
we found to be similar in performance to the mixed-radix strategy. However, the computa-
tional complexity of generating these pulses is much higher, requiring additional optimizable
controls and a larger simulated Hilbert space (when taking into account the simulation of
additional “guard” energy levels). Additionally, the mixed-radix scheme presented here only
requires calibration between each pair of qudits, which is similar to the processes already in
use on quantum computers. On the other hand, the iToffoli scheme would require calibration
between each trio of qubits, which would add a significant additional overhead.
There have also been a few physical realizations of the iToffoli gate that use iSWAP
gate, a CPHASE gate, and a reverse iSWAP gate using the |2i state to change the state
of a qubit in [50, 36]. Galda et. al. [39] explore using qutrits on IBM’s Jakarta device
to implement a Toffoli gate with 78% fidelity. This is similar to our work, using the more
accessible |2i state to perform a three-qubit gate. These are conceptually similar to the
encode, mixed-radix Toffoli, and decode scheme that was laid out in this work. We believe
that a machine specifically designed with qudits in mind could enable much higher fidelities
for similar experiments.

4.6.8 Discussion

This section takes advantage of increased connectivity, and interaction potential when we
have encoded qubits into a four-level system. Encoding qubits in this way allows for the

131
interaction of three to four qubits across a single physical connection, and we synthesize a
library or efficient three-qubit gates via optimal control that take advantage of this virtual
connectivity and are much faster and higher fidelity than performing the decomposition of
a three-qubit gate. We also notice that specific configurations of three-qubit gates, those
that encode similar qubit actions in a gate within the same ququart, produce more efficient
results. We then use these gates to develop compilation strategies, the quantum waltz, that
use the most efficient configurations of three-qubit gates on mixed-radix and full-ququart
systems to produce circuits that achieve 2x and 3x better simulated fidelities in mixed-radix
and full-ququart environments, respectively compared to two-qubit based strategies. We
also demonstrate that ququart-based gates are a viable alternative to iToffoli based three-
qubit pulse strategies with potential practical upsides. Despite difficulties of accessing and
performing operations on higher level states, this efficient implementation of three-qubit
gates provides worthwhile benefits for quantum computation.
Mixed-radix and full-ququart implementations of three-qubit gates makes ququart com-
putation an invaluable piece of the quantum computing repertoire. It is more flexible than
previous hand optimized circuits to improve circuit execution via higher radix devices, does
not require that the entire circuit use quaternary-based logic, and can be selectively applied
to certain sections of larger circuits via the encoding scheme. Realized implementations
of these gates provide a massive opportunity to improve near-term execution of quantum
circuits and expand the capabilities of quantum computers.

132
CHAPTER 5

CONCLUSION AND FUTURE WORK

The fundamental aspect of this work is that the core of an algorithm for general purpose
quantum computing can be adapted for specific architectures with considerable improvement
to the results of quantum programs. There are some shared errors between quantum archi-
tectures, mainly gate errors and decoherence errors and compilation goals such as low gate
count and circuit duration are also shared. But these architecture-specific changes allow for
the use of the unique features of architectures to further reduce these metrics and in some
cases, working around their flaws. A general mapping, routing or scheduling algorithm, such
as those included in most quantum frameworks, applied without alteration to a quantum
program to be run on a specific device will fail to produce the most successful version of that
program.
In the context of Neutral Atom devices, we adapt mapping and routing strategies to
continue to honor a more dense interaction graph, but understand that some SWAPs could
be much more disruptive to locality than a general communication algorithm. To get around
this issue, we adapt our disruption method to take into account a physical distance, not just
a topological distance, between qubits. This gave our router a better sense of how a given
mapping would be affected by a certain SWAP path. Even further, a traditional routing
algorithm would prefer to use the entirety of a machine, and hit the bounds of maximum
interaction distances. This is a detriment to Neutral Atom devices. Atom loss can make this
infeasible. But by adapting the algorithm to be more constrained, a use a “smaller” device in
both size and maximum interaction distance, Neutral Atoms save significant overhead time,
increasing utilization of the device.
Higher-radix qudit systems have a unique feature of non-static interaction graphs. A
SWAP can be more error prone, or have a longer duration, breaking the initial assumption
that all SWAPs have the same error cost to a circuit. Further, this is dependent on the current
133
maximum energy level of the qudit at the time. Adapting to this strategy meant keeping
track of the current state of the qudits, and updating the interaction graph accordingly in
both the qudit interaction graph, and the expanded qubit interaction graph in the cases
where two qubits were encoded into a single ququart. Additionally, with several different
sets of generated pulses to achieve the computation of qubits-on-ququarts, we drastically
improve our ability to accommodate three-qubit gates which significantly improves both the
circuit duration and the error from gate execution. This circumvents the increased error from
reduced coherence time of higher-radix qudits, and provides real benefits for using qudits.
The explorations detailed in this work are a small portion of the many different archi-
tectures in development in quantum computing. IBM has a road map, expecting processors
with many thousands of qubits [40]. This will not be achievable on a single chip, and we
will likely require changes to our routing algorithms that can account for more error prone,
and expensive interconnects [37]. Neutral Atom devices can perform physical movement of
qubits, not just data movement as was explored today [16]. In the short term, it may be vi-
able to have tailored compiler toolchains for each type of hardware, but as we move towards
larger quantum hardware, the compiler pipeline will need to be able to adapt.
With increases in circuit size on the horizon, there are several avenues of future work
to been explored. The first, put simply, is the issue of scalability. Hardware will continue
to increase in size in both number of qubits and connectivity, and circuits will make use of
these qubits. Many of the compilation algorithms use global information, rather than local
decision making. These strategies can scale poorly in the classical sense. While it can be
argued that heavy optimization is not an issue for quantum computation since any error
reduction can significantly improve the program, this isn’t the most practical approach for
cloud providers attempting to efficiently run circuits from many varied users. Other future
work involves mapping and routing circuits effectively across different sections of a quantum
devices, using the more error prone interconnects mentioned earlier in an effective way and

134
without accumulating error. Ideally, quantum computing and hardware will also be moving
into a more fault-tolerant, potentially error corrected era. The role for mapping, routing and
pulse control operations is more nebulous in this regime. Exploring compilation for different
gate sets tailored to these error correction requirements is a worthwhile exploration, as are
interesting ways to use the unique features of a device, such as long distance interaction.
Another direction focuses more on compilation strategies for larger circuits. Many of the
benchmarks used in this work are computational kernels that are used as a smaller piece
of a larger circuit. Yet, they are used as demonstrations of mapping of routing algorithms.
In the near future, it may be the case that different portions of a circuit will have vastly
different interaction patterns. The mapping and routing strategies listed in this work have
some resiliency to this concept through the use of a lookahead pattern, but their effectiveness
as a “mid-circuit” strategy has not be explored. It is worth developing heuristics for how
a circuit can be subdivided into different interaction patterns, and developing algorithms
that perform mid-circuit remapping, or change their routing strategy entirely to fit the new
interaction strategy between qubits.
A third path is developing an understanding of how different optimizations affect one
another. Many quantum compilation frameworks tout a suite of different optimizations, often
times with a number of different preset lists of optimizations that claim to have certain effects
on a circuit that use a general intuition of what should be best for the circuit in terms of depth
and gate count. But, it would be valuable to have an automated intuition for which kinds
of optimizations should come before others. An exploration into how different optimization
orders help or hurt a circuit could be invaluable to developing future optimizations and adding
them to existing pipelines. Additionally, a method for information sharing across a compiler
pipeline such that optimizations do not conflict with one another could be instrumental as
optimizations for quantum architectures increase in number.
As we develop new algorithms and compiler optimizations to execute quantum programs,

135
it is important to reason about how it can be adaptable to specific features, without requiring
an upheaval of the entire process. In this work, while the core of each compilation pipeline
was relatively similar, it did not natively fit within the bounds of a particular framework,
they were rewritten from the ground up to inherit the native features of the architecture
itself. Most today’s frameworks are only able to support the basic assumptions laid out in
the original algorithm as it is the most device-independent way to perform an optimization.
As we push into larger, and more varied quantum systems, software developers and
quantum compiler engineers need to be cognizant about the fact that not all devices can be
assumed to be the same. Although they may have shared features, it is not enough to base
a general algorithm on. This is especially true when it will be applied to several different
architectures. We must work to develop systems that can be informed of particular features
earlier in compilation so that different architectures can be used to their full potential.

136
REFERENCES

[1] David Schuster Lab.

[2] IBM Quantum Devices.

[3] M. Abdelhafez, B. Baker, A. Gyenis, P. Mundada, A. A. Houck, D. Schuster, and


J. Koch. Universal gates for protected superconducting qubits using optimal control.
Phys. Rev. A, 101(2):022321, Feb. 2020.

[4] M. S. ANIS, Abby-Mitchell, H. Abraham, AduOffei, R. Agarwal, G. Agliardi,


M. Aharoni, V. Ajith, I. Y. Akhalwaya, G. Aleksandrowicz, T. Alexander, M. Amy,
S. Anagolum, Anthony-Gandon, E. Arbel, A. Asfaw, A. Athalye, A. Avkhadiev,
C. Azaustre, P. BHOLE, A. Banerjee, S. Banerjee, W. Bang, A. Bansal, P. Bark-
outsos, A. Barnawal, G. Barron, G. S. Barron, L. Bello, Y. Ben-Haim, M. C. Bennett,
D. Bevenius, D. Bhatnagar, P. Bhatnagar, A. Bhobe, P. Bianchini, L. S. Bishop,
C. Blank, S. Bolos, S. Bopardikar, S. Bosch, S. Brandhofer, Brandon, S. Bravyi,
N. Bronn, Bryce-Fuller, D. Bucher, A. Burov, F. Cabrera, P. Calpin, L. Capelluto,
J. Carballo, G. Carrascal, A. Carriker, I. Carvalho, A. Chen, C.-F. Chen, E. Chen, J. C.
Chen, R. Chen, F. Chevallier, K. Chinda, R. Cholarajan, J. M. Chow, S. Churchill,
CisterMoke, C. Claus, C. Clauss, C. Clothier, R. Cocking, R. Cocuzzo, J. Connor,
F. Correa, Z. Crockett, A. J. Cross, A. W. Cross, S. Cross, J. Cruz-Benito, C. Culver,
A. D. Córcoles-Gonzales, N. D, S. Dague, T. E. Dandachi, A. N. Dangwal, J. Daniel,
M. Daniels, M. Dartiailh, A. R. Davila, F. Debouni, A. Dekusar, A. Deshmukh,
M. Deshpande, D. Ding, J. Doi, E. M. Dow, P. Downing, E. Drechsler, E. Dumitrescu,
K. Dumon, I. Duran, K. EL-Safty, E. Eastman, G. Eberle, A. Ebrahimi, P. Een-
debak, D. Egger, ElePT, Emilio, A. Espiricueta, M. Everitt, D. Facoetti, Farida,
P. M. Fernández, S. Ferracin, D. Ferrari, A. H. Ferrera, R. Fouilland, A. Frisch,
A. Fuhrer, B. Fuller, M. GEORGE, J. Gacon, B. G. Gago, C. Gambella, J. M. Gam-
137
betta, A. Gammanpila, L. Garcia, T. Garg, S. Garion, J. R. Garrison, J. Garrison,
T. Gates, H. Georgiev, L. Gil, A. Gilliam, A. Giridharan, Glen, J. Gomez-Mosquera,
Gonzalo, S. d. l. P. González, J. Gorzinski, I. Gould, D. Greenberg, D. Grinko,
W. Guan, D. Guijo, J. A. Gunnels, H. Gupta, N. Gupta, J. M. Günther, M. Haglund,
I. Haide, I. Hamamura, O. C. Hamido, F. Harkins, K. Hartman, A. Hasan, V. Havlicek,
J. Hellmers, \. Herok, S. Hillmich, H. Horii, C. Howington, S. Hu, W. Hu, C.-H. Huang,
J. Huang, R. Huisman, H. Imai, T. Imamichi, K. Ishizaki, Ishwor, R. Iten, T. Itoko,
A. Ivrii, A. Javadi, A. Javadi-Abhari, W. Javed, Q. Jianhua, M. Jivrajani, K. Johns,
S. Johnstun, Jonathan-Shoemaker, JosDenmark, JoshDumo, J. Judge, T. Kachmann,
A. Kale, N. Kanazawa, J. Kane, Kang-Bae, A. Kapila, A. Karazeev, P. Kassebaum,
T. Kehrer, J. Kelso, S. Kelso, H. v. Kemenade, V. Khanderao, S. King, Y. Kobayashi,
Kovi11Day, A. Kovyrshin, R. Krishnakumar, P. Krishnamurthy, V. Krishnan, K. Kr-
sulich, P. Kumkar, G. Kus, R. LaRose, E. Lacal, R. Lambert, H. Landa, J. Lapeyre,
J. Latone, S. Lawrence, C. Lee, G. Li, T. J. Liang, J. Lishman, D. Liu, P. Liu,
Lolcroc, A. K. M, L. Madden, Y. Maeng, S. Maheshkar, K. Majmudar, A. Maly-
shev, M. E. Mandouh, J. Manela, Manjula, J. Marecek, M. Marques, K. Marwaha,
D. Maslov, P. Maszota, D. Mathews, A. Matsuo, F. Mazhandu, D. McClure, M. McE-
laney, C. McGarry, D. McKay, D. McPherson, S. Meesala, D. Meirom, C. Mendell,
T. Metcalfe, M. Mevissen, A. Meyer, A. Mezzacapo, R. Midha, D. Miller, H. Miller,
Z. Minev, A. Mitchell, N. Moll, A. Montanez, G. Monteiro, M. D. Mooring, R. Morales,
N. Moran, D. Morcuende, S. Mostafa, M. Motta, R. Moyard, P. Murali, D. Murata,
J. Müggenburg, T. NEMOZ, D. Nadlinger, K. Nakanishi, G. Nannicini, P. Nation,
E. Navarro, Y. Naveh, S. W. Neagle, P. Neuweiler, A. Ngoueya, T. Nguyen, J. Nican-
der, Nick-Singstock, P. Niroula, H. Norlen, NuoWenLei, L. J. O’Riordan, O. Ogun-
bayo, P. Ollitrault, T. Onodera, R. Otaolea, S. Oud, D. Padilha, H. Paik, S. Pal,
Y. Pang, A. Panigrahi, V. R. Pascuzzi, S. Perriello, E. Peterson, A. Phan, K. Pilch,

138
F. Piro, M. Pistoia, C. Piveteau, J. Plewa, P. Pocreau, A. Pozas-Kerstjens, R. Pracht,
M. Prokop, V. Prutyanov, S. Puri, D. Puzzuoli, Pythonix, J. Pérez, Quant02, Quintiii,
R. I. Rahman, A. Raja, R. Rajeev, I. Rajput, N. Ramagiri, A. Rao, R. Raymond,
O. Reardon-Smith, R. M.-C. Redondo, M. Reuter, J. Rice, M. Riedemann, Rietesh,
D. Risinger, P. Rivero, M. L. Rocca, D. M. Rodríguez, RohithKarur, B. Rosand,
M. Rossmannek, M. Ryu, T. SAPV, N. R. C. Sa, A. Saha, A. A. Saki, S. Sanand,
M. Sandberg, H. Sandesara, R. Sapra, H. Sargsyan, A. Sarkar, N. Sathaye, N. Savola,
B. Schmitt, C. Schnabel, Z. Schoenfeld, T. L. Scholten, E. Schoute, M. Schulter-
brandt, J. Schwarm, J. Seaward, Sergi, I. F. Sertage, K. Setia, F. Shah, N. Shammah,
W. Shanks, R. Sharma, Y. Shi, J. Shoemaker, A. Silva, A. Simonetto, D. Singh,
D. Singh, P. Singh, P. Singkanipa, Y. Siraichi, Siri, J. Sistos, I. Sitdikov, S. Sivarajah,
Slavikmew, M. B. Sletfjerding, J. A. Smolin, M. Soeken, I. O. Sokolov, I. Sokolov,
V. P. Soloviev, SooluThomas, Starfish, D. Steenken, M. Stypulkoski, A. Suau, S. Sun,
K. J. Sung, M. Suwama, O. S\textbackslashlowik, H. Takahashi, T. Takawale, I. Tav-
ernelli, C. Taylor, P. Taylour, S. Thomas, K. Tian, M. Tillet, M. Tod, M. Tomasik,
C. Tornow, E. d. l. Torre, J. L. S. Toural, K. Trabing, M. Treinish, D. Trenev, Tr-
ishaPe, F. Truger, G. Tsilimigkounakis, D. Tulsi, D. Tuna, W. Turner, Y. Vaknin,
C. R. Valcarce, F. Varchon, A. Vartak, A. C. Vazquez, P. Vijaywargiya, V. Vil-
lar, B. Vishnu, D. Vogt-Lee, C. Vuillot, J. Weaver, J. Weidenfeller, R. Wieczorek,
J. A. Wildstrom, J. Wilson, E. Winston, WinterSoldier, J. J. Woehr, S. Woerner,
R. Woo, C. J. Wood, R. Wood, S. Wood, J. Wootton, M. Wright, L. Xing, J. YU,
B. Yang, U. Yang, J. Yao, D. Yeralin, R. Yonekura, D. Yonge-Mallo, R. Yoshida,
R. Young, J. Yu, L. Yu, Yuma-Nakamura, C. Zachow, L. Zdanski, H. Zhang, I. Zi-
daru, B. Zimmermann, C. Zoufal, aeddins-ibm, alexzhang13, b63, bartek-bartlomiej,
bcamorrison, brandhsn, chetmurthy, deeplokhande, dekel.meirom, dime10, dlasecki,
ehchen, ewinston, fanizzamarco, fs1132429, gadial, galeinston, georgezhou20, georgios-

139
ts, gruu, hhorii, hhyap, hykavitha, itoko, jeppevinkel, jessica-angel7, jezerjojo14, jliu45,
johannesgreiner, jscott2, klinvill, krutik2966, ma5x, michelle4654, msuwama, nico-lgrs,
nrhawkins, ntgiwsvp, ordmoj, s. pahwa, pritamsinha2304, rithikaadiga, ryancocuzzo,
saktar-unr, saswati-qiskit, septembrr, sethmerkel, sg495, shaashwat, smturro2, stern-
parky, strickroman, tigerjack, tsura-crisaldo, upsideon, vadebayo49, welien, willhbang,
wmurphy-collabstar, yang.luh, and M. Čepulkovskis. Qiskit: An Open-source Frame-
work for Quantum Computing, 2021.

[5] F. Arute, K. Arya, R. Babbush, D. Bacon, J. C. Bardin, R. Barends, R. Biswas,


S. Boixo, F. G. S. L. Brandao, D. A. Buell, B. Burkett, Y. Chen, Z. Chen, B. Chiaro,
R. Collins, W. Courtney, A. Dunsworth, E. Farhi, B. Foxen, A. Fowler, C. Gidney,
M. Giustina, R. Graff, K. Guerin, S. Habegger, M. P. Harrigan, M. J. Hartmann, A. Ho,
M. Hoffmann, T. Huang, T. S. Humble, S. V. Isakov, E. Jeffrey, Z. Jiang, D. Kafri,
K. Kechedzhi, J. Kelly, P. V. Klimov, S. Knysh, A. Korotkov, F. Kostritsa, D. Land-
huis, M. Lindmark, E. Lucero, D. Lyakh, S. Mandrà, J. R. McClean, M. McEwen,
A. Megrant, X. Mi, K. Michielsen, M. Mohseni, J. Mutus, O. Naaman, M. Neeley,
C. Neill, M. Y. Niu, E. Ostby, A. Petukhov, J. C. Platt, C. Quintana, E. G. Rieffel,
P. Roushan, N. C. Rubin, D. Sank, K. J. Satzinger, V. Smelyanskiy, K. J. Sung, M. D.
Trevithick, A. Vainsencher, B. Villalonga, T. White, Z. J. Yao, P. Yeh, A. Zalcman,
H. Neven, and J. M. Martinis. Quantum supremacy using a programmable supercon-
ducting processor. Nature, 574(7779):505–510, Oct. 2019.

[6] R. Babbush, C. Gidney, D. W. Berry, N. Wiebe, J. McClean, A. Paler, A. Fowler, and


H. Neven. Encoding Electronic Spectra in Quantum Circuits with Linear T Complex-
ity. Physical Review X, 8(4), Oct. 2018.

[7] J. M. Baker, C. Duckering, and F. T. Chong. Efficient quantum circuit decompositions

140
via intermediate qudits. In 2020 IEEE 50th International Symposium on Multiple-
Valued Logic (ISMVL), pages 303–308. IEEE, 2020.

[8] J. M. Baker, C. Duckering, P. Gokhale, N. C. Brown, K. R. Brown, and F. T. Chong.


Improved quantum circuits via intermediate qutrits. ACM Transactions on Quantum
Computing, 1(1):1–25, 2020.

[9] J. M. Baker, C. Duckering, A. Hoover, and F. T. Chong. Decomposing Quantum


Generalized Toffoli with an Arbitrary Number of Ancilla, 2019. _eprint: 1904.01671.

[10] J. M. Baker, C. Duckering, A. Hoover, and F. T. Chong. Time-sliced quantum circuit


partitioning for modular architectures. In Proceedings of the 17th ACM International
Conference on Computing Frontiers. ACM, May 2020.

[11] A. Barenco, C. H. Bennett, R. Cleve, D. P. DiVincenzo, N. Margolus, P. Shor,


T. Sleator, J. A. Smolin, and H. Weinfurter. Elementary gates for quantum com-
putation. Physical review A, 52(5):3457, 1995. Publisher: APS.

[12] D. Barredo, S. De Léséleuc, V. Lienhard, T. Lahaye, and A. Browaeys. An atom-


by-atom assembler of defect-free arbitrary two-dimensional atomic arrays. Science,
354(6315):1021–1023, 2016. Publisher: American Association for the Advancement of
Science.

[13] D. Barredo, V. Lienhard, S. De Leseleuc, T. Lahaye, and A. Browaeys. Synthetic


three-dimensional atomic structures assembled atom by atom. Nature, 561(7721):79–
82, 2018. Publisher: Nature Publishing Group.

[14] E. Bernstein and U. Vazirani. Quantum complexity theory. SIAM Journal on com-
puting, 26(5):1411–1473, 1997. Publisher: SIAM.

[15] M. S. Blok, V. V. Ramasesh, T. Schuster, K. O’Brien, J. M. Kreikebaum, D. Dahlen,

141
A. Morvan, B. Yoshida, N. Y. Yao, and I. Siddiqi. Quantum Information Scrambling
in a Superconducting Qutrit Processor. Physical Review X, 11(2):021010, Apr. 2021.

[16] D. Bluvstein, H. Levine, G. Semeghini, T. T. Wang, S. Ebadi, M. Kalinowski,


A. Keesling, N. Maskara, H. Pichler, M. Greiner, V. Vuletić, and M. D. Lukin. A
quantum processor based on coherent transport of entangled atom arrays. Nature,
604(7906):451–456, Apr. 2022.

[17] A. Bocharov, S. X. Cui, M. Roetteler, and K. M. Svore. Improved Quantum Ternary


Arithmetics. June 2016.

[18] A. Bocharov, M. Roetteler, and K. M. Svore. Factoring with qutrits: Shor’s algorithm
on ternary and metaplectic quantum architectures. Physical Review A, 96(1):012306,
July 2017.

[19] N. C. Brown and K. R. Brown. Comparing Zeeman qubits to hyperfine qubits in the
context of the surface code: $^{174}$Yb$^{+}$ and $^{171}$Yb$^{+}$. Physical
Review A, 97(5):052301, May 2018. arXiv:1803.02545 [quant-ph].

[20] T. A. Brun. A simple model of quantum trajectories. American Journal of Physics,


70(7):719–737, 2002.

[21] C. D. Bruzewicz, J. Chiaverini, R. McConnell, and J. M. Sage. Trapped-ion quantum


computing: Progress and challenges. Applied Physics Reviews, 6(2):021314, 2019.
Publisher: AIP Publishing LLC.

[22] A. Cervera-Lierta, M. Krenn, A. n. Aspuru-Guzik, and A. Galda. Experimental High-


Dimensional Greenberger-Horne-Zeilinger Entanglement with Superconducting Trans-
mon Qutrits. Physical Review Applied, 17(2), Feb. 2022.

[23] P. Chapman. Scaling IonQ’s Quantum Computers: The Roadmap, 2020.

142
[24] Y. Chi, J. Huang, Z. Zhang, J. Mao, Z. Zhou, X. Chen, C. Zhai, J. Bao, T. Dai,
H. Yuan, M. Zhang, D. Dai, B. Tang, Y. Yang, Z. Li, Y. Ding, L. K. Oxenløwe, M. G.
Thompson, J. L. O’Brien, Y. Li, Q. Gong, and J. Wang. A programmable qudit-based
quantum processor. Nature Communications, 13(1):1166, Dec. 2022.

[25] S. U. Cho, M.-H. Bae, K. Kang, and N. Kim. Tunable superconducting qudit mediated
by microwave photons. AIP Advances, 5(8):087186, Aug. 2015.

[26] J. P. Covey, I. S. Madjarov, A. Cooper, and M. Endres. 2000-Times Repeated Imaging


of Strontium Atoms in Clock-Magic Tweezer Arrays. Phys. Rev. Lett., 122(17):173201,
May 2019. Publisher: American Physical Society.

[27] A. Cowtan, S. Dilkes, R. Duncan, A. Krajenbrink, W. Simmons, and S. Sivarajah. On


the qubit routing problem. arXiv preprint arXiv:1902.08091, 2019.

[28] S. A. Cuccaro, T. G. Draper, S. A. Kutin, and D. P. Moulton. A new quantum


ripple-carry addition circuit. arXiv preprint quant-ph/0410184, 2004.

[29] V. Cuppu, B. Jacob, B. Davis, and T. Mudge. A performance comparison of contem-


porary DRAM architectures. In Proceedings of the 26th International Symposium on
Computer Architecture (Cat. No.99CB36367), pages 222–233, 1999.

[30] P. Das, S. S. Tannu, P. J. Nair, and M. Qureshi. A Case for Multi-Programming


Quantum Computers. In Proceedings of the 52nd Annual IEEE/ACM International
Symposium on Microarchitecture, MICRO ’52, pages 291–303, New York, NY, USA,
2019. Association for Computing Machinery. event-place: Columbus, OH, USA.

[31] C. Developers. Cirq, Dec. 2022.

[32] Y. Ding, P. Gokhale, S. F. Lin, R. Rines, T. Propson, and F. T. Chong. Systematic


Crosstalk Mitigation for Superconducting Qubits via Frequency-Aware Compilation.

143
In 2020 53rd Annual IEEE/ACM International Symposium on Microarchitecture (MI-
CRO). IEEE, Oct. 2020.

[33] C. Duckering, J. M. Baker, A. Litteken, and F. T. Chong. Orchestrated trios: compiling


for efficient communication in Quantum programs with 3-Qubit gates. In Proceedings
of the 26th ACM International Conference on Architectural Support for Programming
Languages and Operating Systems. ACM, Apr. 2021.

[34] M. Endres, H. Bernien, A. Keesling, H. Levine, E. R. Anschuetz, A. Krajenbrink,


C. Senko, V. Vuletic, M. Greiner, and M. D. Lukin. Atom-by-atom assembly of defect-
free one-dimensional cold atom arrays. Science, 354(6315):1024–1027, 2016. Publisher:
American Association for the Advancement of Science.

[35] E. Farhi, J. Goldstone, and S. Gutmann. A Quantum Approximate Optimization


Algorithm, 2014.

[36] A. Fedorov, L. Steffen, M. Baur, M. P. d. Silva, and A. Wallraff. Implementation of a


Toffoli gate with superconducting circuits. Nature, 481(7380):170–172, Dec. 2011.

[37] B. Foxen, J. Y. Mutus, E. Lucero, R. Graff, A. Megrant, Y. Chen, C. Quintana,


B. Burkett, J. Kelly, E. Jeffrey, Y. Yang, A. Yu, K. Arya, R. Barends, Z. Chen,
B. Chiaro, A. Dunsworth, A. Fowler, C. Gidney, M. Giustina, T. Huang, P. Klimov,
M. Neeley, C. Neill, P. Roushan, D. Sank, A. Vainsencher, J. Wenner, T. C. White, and
J. M. Martinis. Qubit compatible superconducting interconnects. Quantum Science
and Technology, 3(1):014005, Nov. 2017. Publisher: IOP Publishing.

[38] A. Fuhrmanek, R. Bourgain, Y. R. P. Sortais, and A. Browaeys. Free-Space Lossless


State Detection of a Single Trapped Atom. Phys. Rev. Lett., 106(13):133003, Mar.
2011. Publisher: American Physical Society.

144
[39] A. Galda, M. Cubeddu, N. Kanazawa, P. Narang, and N. Earnest-Noble. Implementing
a Ternary Decomposition of the Toffoli Gate on Fixed-Frequency Transmon Qutrits,
Sept. 2021.

[40] J. Gambetta. Expanding the IBM Quantum roadmap to anticipate the future of
quantum-centric supercomputing, 2022.

[41] K. Geerlings, Z. Leghtas, I. M. Pop, S. Shankar, L. Frunzio, R. J. Schoelkopf, M. Mir-


rahimi, and M. H. Devoret. Demonstrating a Driven Reset Protocol for a Supercon-
ducting Qubit. Phys. Rev. Lett., 110(12):120501, Mar. 2013.

[42] P. Gokhale, J. M. Baker, C. Duckering, N. C. Brown, K. R. Brown, and F. T. Chong.


Asymptotic improvements to quantum circuits via qutrits. In Proceedings of the 46th
International Symposium on Computer Architecture. ACM, June 2019.

[43] P. Gokhale, S. Koretsky, S. Huang, S. Majumder, A. Drucker, K. R. Brown, and F. T.


Chong. Quantum Fan-out: Circuit Optimizations and Technology Modeling, 2020.

[44] N. Goss, A. Morvan, B. Marinelli, B. K. Mitchell, L. B. Nguyen, R. K. Naik, L. Chen,


C. Jünger, J. M. Kreikebaum, D. I. Santiago, J. J. Wallman, and I. Siddiqi. High-
Fidelity Qutrit Entangling Gates for Superconducting Circuits, June 2022.

[45] D. Gottesman. An introduction to quantum error correction and fault-tolerant quan-


tum computation. In Quantum information science and its contributions to math-
ematics, Proceedings of Symposia in Applied Mathematics, volume 68, pages 13–58,
2010.

[46] L. K. Grover. A fast quantum mechanical algorithm for database search. In Proceedings
of the Twenty-Eighth Annual ACM Symposium on Theory of Computing, pages 212–
219, 1996. _eprint: quant-ph/9605043.

145
[47] S. Günther, N. A. Petersson, and J. L. DuBois. Quantum Optimal Control for Pure-
State Preparation Using One Initial State. arXiv:2106.09148 [quant-ph], Aug. 2021.

[48] L. Henriet, L. Beguin, A. Signoles, T. Lahaye, A. Browaeys, G.-O. Reymond, and


C. Jurczak. Quantum computing with neutral atoms. Quantum, 4:327, Sept. 2020.
Publisher: Verein zur Forderung des Open Access Publizierens in den Quantenwis-
senschaften.

[49] A. Hill. Beyond Qubits: Unlocking the Third State in Quantum Processors, Dec. 2021.
Publication Title: Rigetti Computing.

[50] A. D. Hill, M. J. Hodson, N. Didier, and M. J. Reagor. Realization of arbitrary


doubly-controlled quantum phase gates, 2021.

[51] Y. Hirata, M. Nakanishi, S. Yamashita, and Y. Nakashima. An efficient conversion of


quantum circuits to a linear nearest neighbor architecture. Quantum Information and
Computation, 11(1):142, 2011.

[52] A. A. Houck, J. Koch, M. H. Devoret, S. M. Girvin, and R. J. Schoelkopf. Life after


charge noise: recent results with transmon qubits. Quantum Information Processing,
8(2-3):105–115, June 2009.

[53] S. S. Ivanov, H. S. Tonchev, and N. V. Vitanov. Time-efficient implementation of


quantum search with qudits. Physical Review A, 85(6):062321, June 2012.

[54] D. Jaksch, J. Cirac, P. Zoller, S. Rolston, R. Côté, and M. Lukin. Fast quantum gates
for neutral atoms. Physical Review Letters, 85(10):2208, 2000. Publisher: APS.

[55] S. Jandura. Improving a Quantum Compiler, Sept. 2018.

[56] J. Kelly. A Preview of Bristlecone, Google’s New Quantum Processor, Mar. 2018.
Publication Title: Google AI Blog.
146
[57] N. Khammassi, I. Ashraf, X. Fu, C. Almudever, and K. Bertels. QX: A high-
performance quantum computer simulation platform. In Design, Automation & Test
in Europe Conference & Exhibition (DATE), 2017, pages 464–469, 2017.

[58] N. Khaneja, T. Reiss, C. Kehlet, T. Schulte-Herbrüggen, and S. J. Glaser. Optimal


control of coupled spin dynamics: design of NMR pulse sequences by gradient ascent
algorithms. Journal of Magnetic Resonance, 172(2):296–305, Feb. 2005.

[59] H. Kim, W. Lee, H.-g. Lee, H. Jo, Y. Song, and J. Ahn. In situ single-atom array
synthesis using dynamic holographic optical tweezers. Nature communications, 7(1):1–
8, 2016. Publisher: Nature Publishing Group.

[60] Y. Kim, A. Morvan, L. B. Nguyen, R. K. Naik, C. Jünger, L. Chen, J. M. Kreikebaum,


D. I. Santiago, and I. Siddiqi. High-fidelity three-qubit iToffoli gate for fixed-frequency
superconducting qubits. Nature Physics, 18(7):783–788, May 2022.

[61] M. Kjaergaard, M. E. Schwartz, J. Braumüller, P. Krantz, J. I.-J. Wang, S. Gustavsson,


and W. D. Oliver. Superconducting qubits: Current state of play. Annual Review of
Condensed Matter Physics, 11:369–395, 2020. Publisher: Annual Reviews.

[62] J. Koch, T. M. Yu, J. Gambetta, A. A. Houck, D. I. Schuster, J. Majer, A. Blais, M. H.


Devoret, S. M. Girvin, and R. J. Schoelkopf. Charge-insensitive qubit design derived
from the Cooper pair box. Phys. Rev. A, 76(4):042319, Oct. 2007.

[63] P. Krantz, M. Kjaergaard, F. Yan, T. P. Orlando, S. Gustavsson, and W. D. Oliver.


A quantum engineer’s guide to superconducting qubits. Applied Physics Reviews,
6(2):021318, June 2019. Publisher: AIP Publishing.

[64] R. Krishna, V. Makwana, and A. P. Suresh. A Generalization of Bernstein-Vazirani


Algorithm to Qudit Systems, 2016.

147
[65] M. Kwon, M. F. Ebert, T. G. Walker, and M. Saffman. Parallel Low-Loss Measurement
of Multiple Atomic Qubits. Phys. Rev. Lett., 119(18):180504, Oct. 2017. Publisher:
American Physical Society.

[66] C. Lattner and V. Adve. LLVM: A Compilation Framework for Lifelong Program
Analysis and Transformation. In CGO, pages 75–88, San Jose, CA, USA, Mar. 2004.

[67] H. Levine, A. Keesling, G. Semeghini, A. Omran, T. T. Wang, S. Ebadi, H. Bernien,


M. Greiner, V. Vuletić, H. Pichler, and others. Parallel implementation of high-fidelity
multiqubit gates with neutral atoms. Physical review letters, 123(17):170503, 2019.
Publisher: APS.

[68] S. Lloyd and R. Maity. Efficient implementation of unitary transformations.


arXiv:1901.03431 [quant-ph], Jan. 2019.

[69] G. H. Low and I. L. Chuang. Hamiltonian Simulation by Qubitization. Quantum,


3:163, July 2019.

[70] E. Lucero, M. Hofheinz, M. Ansmann, R. C. Bialczak, N. Katz, M. Neeley, A. D.


O’Connell, H. Wang, A. N. Cleland, and J. M. Martinis. High-Fidelity Gates in a
Single Josephson Qubit. Phys. Rev. Lett., 100(24):247001, June 2008.

[71] M. Luo and X. Wang. Universal quantum computation with qudits. Science China
Physics, Mechanics & Astronomy, 57(9):1712–1717, Sept. 2014.

[72] E. Magesan, J. M. Gambetta, and J. Emerson. Scalable and Robust Randomized


Benchmarking of Quantum Processes. Physical Review Letters, 106(18):180504, May
2011.

[73] E. Magesan, J. M. Gambetta, B. R. Johnson, C. A. Ryan, J. M. Chow, S. T. Merkel,


M. P. da Silva, G. A. Keefe, M. B. Rothwell, T. A. Ohki, M. B. Ketchen, and M. Stef-

148
fen. Efficient Measurement of Quantum Gate Error by Interleaved Randomized Bench-
marking. Physical Review Letters, 109(8):080505, Aug. 2012.

[74] D. Miller, T. Holz, H. Kampermann, and D. Bruß. Propagation of generalized Pauli


errors in qudit Clifford circuits. Physical Review A, 98(5), Nov. 2018.

[75] A. Morvan, V. V. Ramasesh, M. S. Blok, J. M. Kreikebaum, K. O’Brien, L. Chen, B. K.


Mitchell, R. K. Naik, D. I. Santiago, and I. Siddiqi. Qutrit Randomized Benchmarking.
Physical Review Letters, 126(21):210504, May 2021.

[76] P. Mundada, G. Zhang, T. Hazard, and A. Houck. Suppression of Qubit Crosstalk in


a Tunable Coupling Superconducting Circuit. Physical Review Applied, 12(5), Nov.
2019. Publisher: American Physical Society (APS).

[77] P. Murali, J. M. Baker, A. J. Abhari, F. T. Chong, and M. Martonosi. Noise-Adaptive


Compiler Mappings for Noisy Intermediate-Scale Quantum Computers, 2019.

[78] P. Murali, D. M. Debroy, K. R. Brown, and M. Martonosi. Architecting Noisy


Intermediate-Scale Trapped Ion Quantum Computers. In Proceedings of the
ACM/IEEE 47th Annual International Symposium on Computer Architecture, pages
529–542. IEEE Press, 2020.

[79] P. Murali, D. C. McKay, M. Martonosi, and A. Javadi-Abhari. Software Mitigation


of Crosstalk on Noisy Intermediate-Scale Quantum Computers. In Proceedings of the
Twenty-Fifth International Conference on Architectural Support for Programming Lan-
guages and Operating Systems, pages 1001–1016, Mar. 2020. arXiv:2001.02826 [quant-
ph].

[80] A. Muthukrishnan and C. R. Stroud. Multivalued logic gates for quantum computation.
Phys. Rev. A, 62(5):052309, Oct. 2000. Publisher: American Physical Society.

149
[81] P. J. Nair. Architectural Techniques to Enable Reliable and Scalable Memory Systems.
arXiv preprint arXiv:1704.03991, 2017.

[82] A. Nersisyan, S. Poletto, N. Alidoust, R. Manenti, R. Renzas, C.-V. Bui, K. Vu,


T. Whyland, Y. Mohan, E. A. Sete, S. Stanwyck, A. Bestwick, and M. Reagor. Man-
ufacturing low dissipation superconducting quantum processors, 2019.

[83] M. A. Nielsen and I. L. Chuang. Quantum Computation and Quantum Information:


10th Anniversary Edition. Cambridge University Press, New York, NY, USA, 10th
edition, 2011.

[84] D. Ohl de Mello, D. Schäffner, J. Werkmann, T. Preuschoff, L. Kohfahl, M. Schlosser,


and G. Birkl. Defect-Free Assembly of 2D Clusters of More Than 100 Single-Atom
Quantum Systems. Physical Review Letters, 122(20), May 2019. Publisher: American
Physical Society (APS).

[85] T. Patel, D. Silver, and D. Tiwari. Geyser: A Compilation Framework for Quantum
Computing with Neutral Atoms. In Proceedings of the 49th Annual International
Symposium on Computer Architecture, ISCA ’22, pages 383–395, New York, NY, USA,
2022. Association for Computing Machinery.

[86] T. Patel, E. Younis, C. Iancu, W. de Jong, and D. Tiwari. QUEST: Systematically


Approximating Quantum Circuits for Higher Output Fidelity. In Proceedings of the
27th ACM International Conference on Architectural Support for Programming Lan-
guages and Operating Systems, ASPLOS ’22, pages 514–528, New York, NY, USA,
2022. Association for Computing Machinery. event-place: Lausanne, Switzerland.

[87] A. Pavlidis and E. Floratos. Arithmetic Circuits for Multilevel Qudits Based on Quan-
tum Fourier Transform. Physical Review A, 103(3):032417, Mar. 2021.

150
[88] N. A. Petersson and F. Garcia. Optimal Control of Closed Quantum Systems via
B-Splines with Carrier Waves, 2021. _eprint: 2106.14310.

[89] N. A. Petersson, F. M. Garcia, A. E. Copeland, Y. L. Rydin, and J. L. DuBois. Discrete


Adjoints for Accurate Numerical Optimization with Application to Quantum Control,
2020. _eprint: 2001.01013.

[90] J. Preskill. Quantum Computing in the NISQ era and beyond. Quantum, 2:79, Aug.
2018.

[91] I. B. M. Quantum. IBM Unveils Breakthrough 127-Qubit Quantum Processor. Publi-


cation Title: IBM Newsroom.

[92] J. Randall, S. Weidt, E. D. Standing, K. Lake, S. C. Webster, D. F. Murgia, T. Nav-


ickas, K. Roth, and W. K. Hensinger. Efficient preparation and detection of microwave
dressed-state qubits and qutrits with trapped ions. Physical Review A, 91(1), Jan. 2015.
Publisher: American Physical Society (APS).

[93] S. Resch, A. Gutierrez, J. S. Huh, S. Bharadwaj, Y. Eckert, G. Loh, M. Oskin, and


S. Tannu. Accelerating Variational Quantum Algorithms Using Circuit Concurrency,
2021.

[94] Rigetti. Rigetti Computing Announces Commercial Availability of 80-Qubit Aspen-M


System and Results of CLOPS Speed Tests, Feb. 2022.

[95] M. Ringbauer, M. Meth, L. Postler, R. Stricker, R. Blatt, P. Schindler, and


T. Monz. A universal qudit quantum processor with trapped ions. Technical Report
arXiv:2109.06903, arXiv, Sept. 2021.

[96] T. Roy, Z. Li, E. Kapit, and D. I. Schuster. Realization of two-qutrit quantum algo-
rithms on a programmable superconducting processor, Nov. 2022.

151
[97] L. Ruiz-Perez and J. C. Garcia-Escartin. Quantum arithmetic with the quantum
Fourier transform. Quantum Information Processing, 16(6):152, 2017. Publisher:
Springer.

[98] M. Saffman. Quantum computing with atomic qubits and Rydberg interactions:
progress and challenges. Journal of Physics B: Atomic, Molecular and Optical Physics,
49(20):202001, 2016. Publisher: IOP Publishing.

[99] L. M. Seifert, J. Chadwick, A. Litteken, F. T. Chong, and J. M. Baker. Time-Efficient


Qudit Gates through Incremental Pulse Re-seeding, 2022.

[100] S. Sheldon, E. Magesan, J. M. Chow, and J. M. Gambetta. Procedure for systematically


tuning up cross-talk in the cross-resonance gate. Physical Review A, 93(6):060302, June
2016.

[101] V. V. Shende and I. L. Markov. On the CNOT-cost of TOFFOLI gates. 2008.

[102] P. W. Shor. Polynomial-Time Algorithms for Prime Factorization and Discrete Log-
arithms on a Quantum Computer. SIAM Journal on Computing, 1995. _eprint:
arXiv:quant-ph/9508027 Published: SIAM J.Sci.Statist.Comput. 26 (1997) 1484.

[103] P. W. Shor. Polynomial-Time Algorithms for Prime Factorization and Discrete Loga-
rithms on a Quantum Computer. SIAM Journal on Computing, 26(5):1484–1509, Oct.
1997.

[104] K. Singh, S. Anand, A. Pocklington, J. T. Kemp, and H. Bernien. A dual-element, two-


dimensional atom array with continuous-mode operation, 2021. _eprint: 2110.05515.

[105] S. Sivarajah, S. Dilkes, A. Cowtan, W. Simmons, A. Edgington, and R. Duncan.


t|ket>: a retargetable compiler for NISQ devices. Quantum Science and Technology,
6(1):014003, Jan. 2021.

152
[106] S. E. Sklarz and D. J. Tannor. Loading a Bose-Einstein condensate onto
an optical lattice: An application of optimal control theory to the nonlinear
Schr\textbackslash"odinger equation. Physical Review A, 66(5):053619, Nov. 2002.

[107] R. S. Smith, M. J. Curtis, and W. J. Zeng. A practical quantum instruction set


architecture. arXiv preprint arXiv:1608.03355, 2016.

[108] S. Stein, N. Wiebe, Y. Ding, P. Bo, K. Kowalski, N. Baker, J. Ang, and A. Li. EQC:
ensembled quantum computing for variational quantum algorithms. In Proceedings
of the 49th Annual International Symposium on Computer Architecture, pages 59–71,
New York New York, June 2022. ACM.

[109] A. Taheri Monfared, M. Haghparast, and K. Datta. Quaternary Quantum/Reversible


Half-Adder, Full-Adder, Parallel Adder and Parallel Adder/Subtractor Circuits. Inter-
national Journal of Theoretical Physics, 58(7):2184–2199, July 2019.

[110] S. S. Tannu and M. Qureshi. Ensemble of diverse mappings: Improving reliability of


quantum computers by orchestrating dissimilar mistakes. In Proceedings of the 52nd
Annual IEEE/ACM International Symposium on Microarchitecture, pages 253–265,
2019.

[111] S. Terzi and A. Gençten. Development of some two and three qutrit quan-
tum logic gates. AIP Conference Proceedings, 2178(1):030027, 2019. _eprint:
https://aip.scitation.org/doi/pdf/10.1063/1.5135425.

[112] J. Tilly, H. Chen, S. Cao, D. Picozzi, K. Setia, Y. Li, E. Grant, L. Wossnig, I. Rungger,
G. H. Booth, and J. Tennyson. The Variational Quantum Eigensolver: a review of
methods and best practices, 2021.

[113] G. Van Rossum and F. L. Drake. Python 3 Reference Manual. CreateSpace, Scotts
Valley, CA, 2009.
153
[114] J. J. Vartiainen, M. Möttönen, and M. M. Salomaa. Efficient Decomposition of Quan-
tum Gates. Physical Review Letters, 92(17), Apr. 2004. Publisher: American Physical
Society (APS).

[115] P. Wang, C.-Y. Luan, M. Qiao, M. Um, J. Zhang, Y. Wang, X. Yuan, M. Gu, J. Zhang,
and K. Kim. Single ion qubit with estimated coherence time exceeding one hour. Nature
Communications, 12(1):233, Jan. 2021.

[116] Y. Wang, Z. Hu, B. C. Sanders, and S. Kais. Qudits and High-Dimensional Quantum
Computing. Frontiers in Physics, 8:479, 2020.

[117] Y. Wang and M. Perkowski. Improved complexity of quantum oracles for ternary
grover algorithm for graph coloring. In 2011 41st IEEE International Symposium on
Multiple-Valued Logic, pages 294–301. IEEE, 2011.

[118] R. Wille, L. Burgholzer, and A. Zulehner. Mapping quantum circuits to IBM QX


architectures using the minimal number of SWAP and H operations. In 2019 56th
ACM/IEEE Design Automation Conference (DAC), pages 1–6. IEEE, 2019.

[119] R. Wille, O. Keszocze, M. Walter, P. Rohrs, A. Chattopadhyay, and R. Drechsler.


Look-ahead schemes for nearest neighbor optimization of 1D and 2D quantum circuits.
In 2016 21st Asia and South Pacific design automation conference (ASP-DAC), pages
292–297. IEEE, 2016.

[120] K. Wright, K. Beck, S. Debnath, J. Amini, Y. Nam, N. Grzesiak, J.-S. Chen, N. Pisenti,
M. Chmielewski, C. Collins, and others. Benchmarking an 11-qubit quantum computer.
Nature communications, 10(1):1–6, 2019. Publisher: Nature Publishing Group.

[121] X. Wu, S. L. Tomarken, N. A. Petersson, L. A. Martinez, Y. J. Rosen, and J. L.


DuBois. High-fidelity software-defined quantum logic on a superconducting qudit.
Physical Review Letters, 125(17):170502, Oct. 2020.
154
[122] X.-C. Wu, S. Di, E. M. Dasgupta, F. Cappello, H. Finkel, Y. Alexeev, and F. T. Chong.
Full-State Quantum Circuit Simulation by Using Data Compression. In Proceedings of
the International Conference for High Performance Computing, Networking, Storage
and Analysis, SC ’19, New York, NY, USA, 2019. Association for Computing Machin-
ery.

[123] E. Younis, C. C. Iancu, W. Lavrijsen, M. Davis, E. Smith, and USDOE. Berkeley


Quantum Synthesis Toolkit (BQSKit) v1, Apr. 2021.

155

You might also like