Superconducting Qubit Readout Pulse Optimization Using Deep Reinforcement Learning

Download as pdf or txt
Download as pdf or txt
You are on page 1of 106

Superconducting Qubit Readout Pulse Optimization

Using Deep Reinforcement Learning


by
Cole R. Hoffer
Submitted to the Department of Electrical Engineering and Computer
Science
in partial fulfillment of the requirements for the degree of
Master of Engineering in Electrical Engineering and Computer Science
at the
MASSACHUSETTS INSTITUTE OF TECHNOLOGY
February 2021

c Massachusetts Institute of Technology 2021. All rights reserved.

Author . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Department of Electrical Engineering and Computer Science
February, 2021

Certified by . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
William D. Oliver
Associate Professor of Electrical Engineering and Computer Science
Thesis Supervisor

Accepted by . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Katrina LaCurts
Chair, Master of Engineering Thesis Committee
2
Superconducting Qubit Readout Pulse Optimization
Using Deep Reinforcement Learning
by
Cole R. Hoffer

Submitted to the Department of Electrical Engineering and Computer Science


on February, 2021, in partial fulfillment of the
requirements for the degree of
Master of Engineering in Electrical Engineering and Computer Science

Abstract
Quantum computers promise to solve specific problems significantly faster than classi-
cal computers. Superconducting quantum processors comprising more than 50 qubits
can achieve quantum supremacy, the ability to outperform existing classical comput-
ers for particular problems. However, to build a useful quantum computer, the quan-
tum processor’s constituent components such as their control and readout must be
very well-calibrated. Qubit-state readout of contemporary superconducting quantum
processors is a significant error source. In an efficient, frequency-multiplexed readout
of multiple qubits, effects such as drive cross-talk increase the complexity of optimal
readout pulse shapes, requiring computationally intensive methods to discover high-
fidelity pulse shapes. In this thesis, we extend existing readout optimization methods
to work in multi-qubit environments and present a new pulse shaping optimization
module using deep reinforcement learning. Compared to conventional readout meth-
ods in a simulated environment, we are able to reduce required readout pulse lengths
by over 63% in single-qubit environments and by over 57% in multi-qubit environ-
ments. In addition to discussing how the deep reinforcement learning pulse shaping
module will be used in experimental contexts, we also evaluate the future generalized
use of deep reinforcement learning methods in quantum computing.

Thesis Supervisor: William D. Oliver


Title: Associate Professor of Electrical Engineering and Computer Science

3
4
Acknowledgments
I want to start by thanking everyone in the EQuS group that supported me during
my MEng. Importantly, I would like to thank my advisor, Professor Will Oliver, for
taking a chance and providing me the opportunity to work in this fantastic group. The
research was both challenging and fulfilling, and the knowledge and skills I gained this
year will stick with me for a long time. Next, I would also like to send a huge thank
you to Ben Lienhard, who served as a fantastic mentor when I first started at the
group and became a great collaborator as the year went on. In no way would I have
gotten as far as I did without all of his help and expertise. I would also like to thank
Antti for being a great resource in helping scope this project and its experimental
realization. Another shout out must go to Billy Woltz, one of my best friends at MIT
and the one who helped me get the position in EQuS! Not only did he help me find
this opportunity, but he was also a constant resource throughout my time here and
was always a fabulous friend. Finally, I want to thank everyone else in the lab. While
I was only on campus in the lab until March (thanks 2020!), everybody in the group
was always supportive and helpful.
Next, I want to thank all those who have supported me this year and during my
undergraduate years at MIT. First and foremost, my mom and dad are responsible
for any success I’ve found here at MIT, and I am eternally grateful for their love and
support. Second, I want to thank all of the close friends I’ve made at MIT. Their
support, quirkiness, and late-night Domino’s pizza breaks during my years at MIT
were a major reason I was able to successfully pursue my academic goals. Finally, it’s
imperative that I especially thank Molly, whose support for me over the past thirteen
years has been nothing short of amazing.

5
6
Contents

1 Introduction 19

2 Physics Background 23
2.1 Quantum Computing Introduction . . . . . . . . . . . . . . . . . . . 23
2.2 Superconducting Quantum Computing . . . . . . . . . . . . . . . . . 24
2.2.1 Linear LC Circuits and Quantum Harmonic Oscillators . . . . 24
2.2.2 Josephson Junctions . . . . . . . . . . . . . . . . . . . . . . . 25
2.2.3 The Cooper Pair Box . . . . . . . . . . . . . . . . . . . . . . . 26
2.2.4 The Transmon Qubit . . . . . . . . . . . . . . . . . . . . . . . 28

3 Superconducting Qubit Readout 31


3.1 Cavity Quantum Electrodynamics . . . . . . . . . . . . . . . . . . . . 32
3.2 Circuit QED and Dispersive Readout . . . . . . . . . . . . . . . . . . 33
3.3 Multiplexed Readout . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
3.4 Computational Simulation . . . . . . . . . . . . . . . . . . . . . . . . 37
3.4.1 Cavity Bloch Equation Formulation . . . . . . . . . . . . . . . 37
3.4.2 Simulation System Parameters . . . . . . . . . . . . . . . . . . 39

4 Machine Learning Background 41


4.1 Machine Learning . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
4.2 Reinforcement Learning . . . . . . . . . . . . . . . . . . . . . . . . . 42
4.2.1 Markov Decision Processes . . . . . . . . . . . . . . . . . . . . 44
4.2.2 Bellman Equation . . . . . . . . . . . . . . . . . . . . . . . . . 44

7
4.2.3 Value Iteration . . . . . . . . . . . . . . . . . . . . . . . . . . 45
4.2.4 Policy Iteration . . . . . . . . . . . . . . . . . . . . . . . . . . 46
4.3 Deep Learning and Neural Networks . . . . . . . . . . . . . . . . . . 47
4.3.1 Feed-Forward Neural Networks . . . . . . . . . . . . . . . . . 47
4.3.2 Cost Functions and Regularization . . . . . . . . . . . . . . . 50
4.3.3 Training Neural Networks . . . . . . . . . . . . . . . . . . . . 51
4.4 Deep Reinforcement Learning . . . . . . . . . . . . . . . . . . . . . . 52

5 Methodology 55
5.1 Conventional Approach . . . . . . . . . . . . . . . . . . . . . . . . . . 55
5.2 CLEAR Pulse . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
5.2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
5.2.2 Scaling to Multi-Qubit Systems . . . . . . . . . . . . . . . . . 56
5.2.3 Bayesian Optimization . . . . . . . . . . . . . . . . . . . . . . 57
5.3 Deep Reinforcement Learning Generated Pulses . . . . . . . . . . . . 60
5.3.1 Deep Q-Networks . . . . . . . . . . . . . . . . . . . . . . . . . 60
5.3.2 Proximal Policy Optimization . . . . . . . . . . . . . . . . . . 63
5.3.3 Other Approaches Considered . . . . . . . . . . . . . . . . . . 66

6 Resonator Reset 69
6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
6.2 Optimization Setup . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
6.3 Single Qubit Simulation . . . . . . . . . . . . . . . . . . . . . . . . . 71
6.3.1 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
6.3.2 Comparing DQN and PPO Optimizations . . . . . . . . . . . 73
6.4 5 Qubit Simulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
6.4.1 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
6.4.2 Flexibility in Generated Pulse Shapes . . . . . . . . . . . . . . 77
6.5 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77

8
7 Resonator Photon Injection 79
7.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
7.2 Optimization Setup . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
7.3 Single Qubit Simulation Results . . . . . . . . . . . . . . . . . . . . . 82
7.4 5 Qubit Simulation Results . . . . . . . . . . . . . . . . . . . . . . . . 84
7.4.1 Injection Results . . . . . . . . . . . . . . . . . . . . . . . . . 84
7.4.2 Entire Readout Pulse Results . . . . . . . . . . . . . . . . . . 86
7.5 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88

8 Experimental Setup 89
8.1 Measurement Setup . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
8.2 Labber Pulse Optimization Driver . . . . . . . . . . . . . . . . . . . . 90
8.3 Resonator Reset Experiment Design . . . . . . . . . . . . . . . . . . . 91
8.4 Resonator Photon Injection Experiment Design . . . . . . . . . . . . 91

9 Conclusion and Future Work 95


9.1 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
9.2 Future Work Possibilities . . . . . . . . . . . . . . . . . . . . . . . . . 97
9.2.1 Additional Applications . . . . . . . . . . . . . . . . . . . . . 97
9.2.2 Pulse Length Optimization . . . . . . . . . . . . . . . . . . . . 98
9.2.3 Continuous Experimental Optimization . . . . . . . . . . . . . 99

9
10
List of Figures

2-1 Energy potential for a quantum harmonic oscillator (QHO) (a) and a
transmon (b). In the QHO, all energy levels have equal spacing ℎ̄𝜔. In
the transmon, the non-linear Josephson inductance (Eq. 2.11) creates
non-equal spacing between each energy level, allowing the first two lev-
els to be individually addressable and create the qubit’s computational
subspace. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25

2-2 Circuit diagrams for a Cooper pair box qubit (a) and a transmon qubit
(b) where the crossed square is the symbol for the Josephson junction
circuit element. The transmon builds upon the Cooper pair box ar-
chitecture by including a large shunt capacitor to reduce sensitivity to
charge noise. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27

2-3 The charge dispersion energies for the 3 first three qubit levels from
Eq. 2.12. As the ratio 𝐸𝐽 /𝐸𝐶 increases, the levels exponentially flatten
and enter the transmon regime. . . . . . . . . . . . . . . . . . . . . . 28

3-1 General cartoon of single qubit dispersive readout. First, a readout


resonator’s frequency is shifted by the underlying qubit state. This
shifted frequency can be probed by a readout pulse, whose reflected
magnitude is effects based on that resonator frequency shift. These ef-
fects are articulated through the post processing of the readout signal
output, where the qubit state can be determined by using discrimina-
tion techniques on the readout signal in the IQ-plane (b). . . . . . . . 32

11
3-2 General representation of a cavity quantum electrodynamic system.
Here a two level atom travels through the cavity in time 𝑡transit and
decays at a rate 𝛾. While traveling through the cavity, the atom with
the photons in the cavity with coupling strength 𝑔 and the photons in
the cavity leak out of the cavity at a rate 𝜅. . . . . . . . . . . . . . . 34

3-3 (a) Circuit schematic of a 5 qubit multiplexed readout system, where


each qubit is coupled to an individual resonator. The resonators are
capacitively coupled to a common drive line. (b) The associated trans-
mitted power for each qubit in the excited and ground states as a
measure of the resonator frequency. . . . . . . . . . . . . . . . . . . . 37

4-1 General flow of a Markov decision processes. Given a current state (𝑠)
and reward (𝑅) at time step 𝑡, an agent chooses an action that is fed
into the environment which produces the next state and reward at for
time step 𝑡 + 1, continuing the cycle until reaching a terminal state. . 44

4-2 (a) Basic feed-forward neural network architecture with an input layer
of 3 nodes, 2 hidden layers with a width of 4 nodes each and single
output layer node. (b) The value of each node in a neural network not
in the input layer is determined by finding the sum of a bias term and
the dot product between the previous layer’s node values and the inter-
layer weight values. This computed value is then fed into a non-linear
activation function. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48

4-3 Common activation functions and their derivatives found in FNNs,


including the linear, ReLU and sigmoid activation functions. . . . . . 50

5-1 Arbitrary CLEAR and rectangular pulse shapes. Two amplitude seg-
ments at the beginning of the pulse help inject photons into the res-
onator quickly, allowing the resonator to reach a steady photon count
in a shorter time. Two amplitude segments are also added to the end
of the pulse to actively reset the resonator to the vacuum state. . . . 57

12
5-2 Wall-clock time needed per optimization step when optimizing differ-
ent numbers of pulse segments. Regardless of the number of segments
being optimized, the length of each optimization step generally in-
creases as the optimization step number increases as we expect from the
Bayesian optimization algorithm. Additionally, increasing the number
of amplitude segments needing to be optimized also results in requir-
ing more optimization time per step throughout the entire optimization
process. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59

5-3 A simple Deep Q-Network architecture where we are trying to construct


a single 4 segment pulse with 2 possible amplitude options. (a) A state
is provided as the input for a deep fully-connected neural network. The
neural network’s outputs approximate the Q-value for each possible
action according to the state input. (b) The action with the maximum
associated Q-value is selected, appended to pulse, and sent to interact
with the simulated or physical environment. (c) The state-action pair’s
actual Q-value is calculated in this environment and is saved in an
experience replay buffer. (d) A batch of expected and actual Q-values
with their state-action pairs are used to train the neural network. (e)
The new state generated by the environment is passed back to the
input of the neural network, restarting the process. . . . . . . . . . . 62

5-4 A simple Proximal Policy Optimization architecture where we are try-


ing to build two 2-segment pulses where any bounded continuous ampli-
tude can be chosen. We use two deep fully connected neural networks.
The first is the policy network that aims to approximate an optimal
policy (the best pulse shape). The second is the value network other
that takes in the predicted pulse from the policy network and predicts
the value of that pulse in the environment. . . . . . . . . . . . . . . . 66

13
6-1 (a) Generated CLEAR, DQN and PPO pulse shapes for single qubit
readout resonator reset. In this case, the conventional rectangular read-
out pulse passively waits for the photons to decay from the resonator,
thus the pulse amplitude is 0. (b) Photon counts over time due to the
generated pulse shapes. . . . . . . . . . . . . . . . . . . . . . . . . . . 72

6-2 2 dimensional representation of 25 dimensional pulses generated by


DQN and PPO using t-SNE. With t-SNE, similar high dimensional
data points are situated close together in the 2 dimensional represen-
tation. On the top row, each marker (pulse) is colored based on its
reward, or number of photons remaining while on the bottom row each
marker is colored based on the iteration it was made in its optimization
cycle. Here we show every 50th pulse generated. . . . . . . . . . . . . 74

6-3 (a) Generated CLEAR, DQN, and PPO pulse shapes for single qubit
readout resonator photon injection and stabilization. In this case, the
conventional rectangular readout pulse passively waits for the photons
to decay from the resonator, thus no pulse amplitude is applied. (b)
Photon counts over time caused from the resulting generated pulse
shapes. (c) Distribution of 25 generated CLEAR reset pulses. . . . . 76

14
7-1 A basic rectangular readout pulse at a normalized drive power 𝑃norm =
4 (a) and the resulting photon counts for the resonators in a 5-qubit
system (b). There are three main sections of the readout pulse that we
consider. The first section is the photon injection optimization range
at the beginning of the pulse, which we actively aim to minimize the
required length of. Second, the middle section is a constant amplitude

section that is driven at 𝑃 and is equal in length to the preceding
pulse section that is being length optimized. Finally the last segment is
reset optimization section that was discussed in Chapter 6. As we seen
in (b), the photon counts from a rectangular pulse have an oscillating
evolution where the photon counts eventually stabilize based on the
normalized drive power 𝑃norm . . . . . . . . . . . . . . . . . . . . . . . 80

7-2 (a) Generated CLEAR, DQN, and PPO pulse shapes for single qubit
readout resonator photon injection and stabilization. In this case, the
conventional approach is a single amplitude rectangular pulse. (b)
Photon counts over time as a result of applying the generated pulse
shapes. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83

7-3 Generated pulse shapes for photon injection and stabilization using a
rectangular pulse, (a) CLEAR pulse and (b) PPO pulse. In this setup,
the amplitudes of the first half of the pulse are optimized while the

second half of the pulse is at a constant amplitude 𝑃 . . . . . . . . . 85

7-4 Resonator photon counts for the full readout pulse (photon injection
and reset) using rectangular pulses (left), optimized CLEAR pulses
(center), and pulses generated using PPO (right). . . . . . . . . . . . 86

15
7-5 IQ-plane cavity trajectories per resonator for each type of generated
full readout pulse (rectangular, CLEAR, PPO). Trajectories with red
markers are associated with a system where all 5 qubits are in the
excited state, and vice versa for with ground state trajectories with
blue markers. Markers are spaced 20ns apart to highlight the speed in
which cavity states are reached. Each IQ-place also has a dashed grey
circle that represents a resonator state with 4 photons (stabilization
target), and a smaller dotted grey circle that represents a resonator
state with 0.10 photons (reset target). . . . . . . . . . . . . . . . . . . 87

8-1 (a) Simplified setup for qubit control and multiplexed readout for a
superconducting multi qubit system. On the left, pulse shapes for
qubit control and readout are created, combined, and sent to the 5-
qubit chip. On the right, after passing through the chip, the readout
signal returns to room temperature after passing through a chain of
amplifiers, allowing us to acquire measurement data. (b) A magnified
view of the 5-qubit chip in the super cooled mixing chamber. The chip’s
qubit’s are connected to a single drive line in a multiplexed fashion,
allowing us to control and readout all 5 qubits with a single multiplexed
signal. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
8-2 Sample Ramsey experiment pulse sequence. Here, a single Ramsey
experiment begins by applying a set of 𝜋-pulses to initialize the qubit
states. This is followed by a rectangular readout pulse that has an
optimized reset component at the end to empty the resonator as quickly
as possible. Following the readout pulse, we apply two sets of 𝜋/2-
pulses, spaced apart by a time 𝑡𝑅 . These 𝜋/2-pulses rotate the qubit-
state dependent Bloch vector on the equator, and then back to the
z-axis. During that time between the two 𝜋/2-pulses, various amounts
of detuning occurs which affects how vector returns back to it’s original
vector position following the second 𝜋/2 pulse. . . . . . . . . . . . . . 92

16
List of Tables

3.1 System parameters used in our simulations. Parameter values for the
5-qubit qubit chip based off a lab experiment focused on qubit-state
discrimination. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39

9.1 Final readout pulse lengths for the various conventional and optimiza-
tion methods in this research. The single-qubit environment includes
optimized CLEAR pulses and pulses generated with PPO and DQN.
In the multi-qubit environment, we only compare optimized CLEAR
pulses and pulses generated with PPO. . . . . . . . . . . . . . . . . . 97

17
18
Chapter 1

Introduction

Quantum computers promise to solve specific problems significantly faster than classi-
cal computers. Superconducting quantum processors comprising more than 50 qubits
can achieve quantum supremacy, with the ability to outperform existing classical
computers for particular problems [1, 2, 3]. To realize a useful quantum computer,
however, the quantum processor’s constituent components, their control, and their
readout must be very well-calibrated.
A critical component in any quantum computing system is the mechanism used
to read out the system’s qubit states. For superconducting qubit systems, we look to
perform quantum non-demolition measurements by using a method called dispersive
readout, where each qubit is entangled with photons in a linear readout resonator [4].
The qubit shifts the resonator’s frequency in response to the qubit state, which allows
us to readout the qubit state when that response is probed. This same technique
scales to multiplexed readout, where we can control multiple qubit-resonator pairs
simultaneously. However, scaling to multiplexed readout also introduces sources of
cross-talk that can make optimizing ideal pulse shapes much more difficult, which we
tackle here with deep reinforcement learning methods.
The goals of optimizing readout pulse shapes are to both quickly populate and
stabilize the resonator states, as well as to actively reset the resonators as quickly as
possible. First, quickly populating and stabilizing the resonator states allows us to
shorten the time needed to probe the resonator’s dispersive shift. Although it is often

19
the case that having more fine-tuned pulse shape control allows us to shorten the
required pulse length compared to a single segment rectangular pulse, the complexity
of the optimal pulse shapes requires powerful optimization techniques. The second
goal in optimizing readout pulse shapes is to actively reset the resonators following
a measurement. The conventional method is to simply turn off the readout pulses
and allow the photons to naturally decay from the resonator. This is a slow and
passive process, which often requires waiting multiple cavity decay times and limits
the performance of important tasks such as quantum error correction [5]. As with
populating and stabilizing the resonator with photons, we can actively shorten the
time required to reset the resonator states using an optimized, fine-tuned pulse shape.

General improvements have been realized by engineering readout pulse shapes that
populate and stabilize the readout resonator as well as actively reset it. A currently
used example is the Cavity Level Excitation and Reset (CLEAR) pulse, a rectangular
pulse that includes constant amplitude pulses before and after the rectangular pulse
to quicken this excitation and reset process [6]. The CLEAR pulse method actively
depopulated the resonator around two cavity time constants faster than the conven-
tional approach. While the CLEAR pulse helps quicken the readout process, it also
raises the question of how the CLEAR pulse can be scaled to multiplexed readout
systems, as well as if more fine-tuned control of the readout pulses can further reduce
the required readout pulse lengths.

These questions guide the focus of this thesis, which applies deep reinforcement
techniques for fine-tuned optimal control of readout pulse shapes. Reinforcement
learning is a subset of machine learning where an agent learns to maximize a reward
by choosing the appropriate action at a given state [7]. Rather than comparing
predictions with labeled data sets like in supervised learning applications (e.g., image
classification), reinforcement learning algorithms “learn” through the process of their
agents directly interacting with the environment. Additionally, reinforcement learning
algorithms do not need to know how the underlying system functions. Rather, they
simply need to be able to compute the reward that a specific state-action pair will
produce. The growth of reinforcement learning research has been shown through

20
success in complex games such as chess and Go, as well as robot planning [8, 9, 10].

Optimization environments with a relatively small number of state-action pairs


like Tic-Tac-Toe can be solved through a variety of tabular methods where we attempt
to learn the expected reward for all state-action pairs [7]. However, while tabular re-
inforcement learning techniques provide a strong foundation for problems like optimal
control, those techniques quickly become impractical with increased dimensionality
and precision. Tabular solution methods require us to approximate rewards for every
state-action pair in the environment, a number that grows exponentially as the state
or action set increases. Approximate solution methods address this issue by replacing
the state-action pair table with a deep neural network that approximates the expected
reward for executing an action at a particular state. These new deep reinforcement
algorithms have greatly expanded the size and complexity of problems that can be
addressed with reinforcement learning. These algorithms have begun to be used on
problems in quantum optimal control, such as unitarily steering a quantum system
towards a desired dynamic [11].

Critical to the overall success of this research is having the resources to explore and
tune different reinforcement algorithms on actual quantum processors. However, we
do not have unlimited access to our laboratory’s quantum processors and thus need
to initially rely on robust simulation software. The implementation realities of deep
reinforcement learning add to the necessity of beginning the work with a simulator.
First, deep reinforcement learning algorithms are notoriously resource-intensive to
train and tune. Initially relying on a simulator allows us to take advantage of resources
like cloud computing and specialized GPUs that immensely improve training speeds
and algorithm iterations. Second, training these deep reinforcement algorithms on
the simulator provides a training “head start” before beginning work on the actual
quantum computer. We will constrain our simulator to the limitations of the physical
devices we will be using as much as possible (e.g., limiting the set of pulse amplitudes
provided by the arbitrary waveform generator). While the simulator is not able to
account for all effects that may be experienced by the lab’s quantum computer, it
will allow the algorithms to train purely on the base-level readout physics. Thus, the

21
majority of training on the quantum computer will be fine-tuning, saving a significant
amount of training time.
For the structure of this thesis, we will start with an introduction to the building
blocks of superconducting qubits in Chapter 2. In Chapter 3, we will discuss the
basics of cavity QED and its counterpart circuit QED , which will also allow us
to introduce the physics behind superconducting qubit readout. From there, we
will transition to an introduction to machine learning and reinforcement learning in
Chapter 4. We can then go into depth about the various deep reinforcement learning
algorithms we use to optimize optimal multiplexed readout pulse shapes in Chapter
5. These algorithms will drive our simulation results both with optimizing pulses
that populate and stabilize readout resonators to reduce the time needed to perform
qubit measurements (Chapter 7), as well as optimizing pulses that actively reset
those resonators following a measurement (Chapter 6). Finally in Chapter 8, we will
describe the experimental designs that are being conducted based off the previous
simulation results.

22
Chapter 2

Physics Background

2.1 Quantum Computing Introduction

While a classical bit is represented as a binary digit with values 0 and 1, a qubit can
be in a superposition of both 0 and 1. We write the quantum state of a single qubit
as a wave function
|Ψ⟩ = 𝛼 |0⟩ + 𝛽 |1⟩ , (2.1)

where |𝛼|2 + |𝛽|2 = 1. If our quantum system has 𝑁 qubits, there will be a resulting
complex superposition of all 2𝑁 permutations, where the system’s entire state can be
defined by the wave function

|Ψ⟩ = 𝑐1 |0..00⟩ + 𝑐2 |0..01⟩ + ... + 𝑐2𝑁 |1..11⟩ . (2.2)

Due to the ability to store exponentially more information than its classical counter-
part, and to apply quantum transformations on all states simultaneously, quantum
processors can tackle many difficult problems that their classical counterparts cannot.
These include, but are not limited too, factoring large numbers exponentially quicker
using Shor’s algorithm, as well as simulating complex quantum systems that can be
used in applications like chemistry and physics [12, 13, 14]. In this thesis, we focus
on superconducting qubits as the foundation for building these quantum processors.

23
2.2 Superconducting Quantum Computing

2.2.1 Linear LC Circuits and Quantum Harmonic Oscillators

To understand superconducting qubits, we can start from the basics with a classical
linear 𝐿𝐶 resonant circuit. This 𝐿𝐶 circuit serves as an electrical counterpart to a
simple mechanical harmonic oscillator. We first start by defining Φ as the branch
flux on inductor 𝐼 and by defining 𝑄 as the branch charge on the capacitor crossing
voltage 𝑉 . This leads to the energy terms of the capacitor and inductor respectively
being defined as
𝑄2
𝒯𝐶 = , (2.3)
2𝐶

Φ2
𝒰𝐿 = . (2.4)
2𝐿
Using the Legendre transformation (the difference between the kinetic and potential
energy terms), the classical Hamiltonian is further defined as

𝑄2 Φ2
𝐻= + . (2.5)
2𝐶 2𝐿

ˆ and Φ̂ to
In order to enter the quantum regime, we use the quantum operators 𝑄
ˆ = Φ̂𝑄
replace 𝑄 and Φ, where [Φ̂, 𝑄] ˆ −𝑄
ˆ Φ̂ = 𝑖ℎ̄. Naturally, this leads to the resulting

quantum Hamiltonian being defined as

𝑄ˆ2 Φ̂2
𝐻= + . (2.6)
2𝐶 2𝐿

This Hamiltonian is equivalent to the Hamiltonian of a quantum harmonic os-


cillator (QHO), which describes a particle in a one-dimensional quadratic potential.
We define ℎ̄ as the reduced Planck’s constant where ℎ̄ = ℎ/2𝜋, 𝜑 as the reduced flux
or phase across an inductor, and finally 𝑎 as the annihilation operator and 𝑎† as the

creation operator. Setting 𝜔𝑟 = 1/ 𝐿𝐶 as the systems resonant frequency, the QHO

24
Figure 2-1: Energy potential for a quantum harmonic oscillator (QHO) (a) and a
transmon (b). In the QHO, all energy levels have equal spacing ℎ̄𝜔. In the transmon,
the non-linear Josephson inductance (Eq. 2.11) creates non-equal spacing between
each energy level, allowing the first two levels to be individually addressable and
create the qubit’s computational subspace.

Hamiltonian in the second quantization form is defined as


(︂ )︂
† 1
𝐻 = ℎ̄𝜔𝑟 𝑎 𝑎+ . (2.7)
2

A major limitation of quantum harmonic oscillators in quantum computing is


that energy levels are equally spaced at ℎ̄𝜔𝑟 (Fig. 2-1). To build a qubit, we need
to be able to address two of these energy states exclusively to build a computational
subspace. More specifically, we need 𝜔0→1 to differ from 𝜔1→2 to address these states
as part of computational subspace. The work to address this issue is introduced in
the forthcoming sections.

2.2.2 Josephson Junctions

To fix the problem of having quantized energy levels being evenly spaced, we intro-
duce a non-linearity, or anharmonicity, into our system using Josephson junctions. A

25
Josephson junction is formed between two superconducting electrodes that are sep-
arated by an insulating barrier, in which Cooper pairs can coherently tunnel across
that insulating barrier [15]. Two relations define the dynamics of Josephson junctions.
The first Josephson relation defines the current flowing through the junction as

𝐼 = 𝐼𝑐 sin(𝜑(𝑡)), (2.8)

where 𝐼𝑐 is defined as the critical current of the junction (the maximum supported
current that can be sustained by the junction) and 𝜑(𝑡) is defined as the phase differ-
ence across the junction. That phase difference evolves in the presence of a potential
𝑉 such that
ℎ̄ 𝑑𝜑
𝑉 = . (2.9)
2𝑒 𝑑𝑡

As Φ0 = 2𝑒
is the superconducting magnetic flux quantum, we can use the two
relations to see that
Φ0 𝑑𝐼
𝑉 = . (2.10)
𝐼𝑐 cos(𝜑) 𝑑𝑡
Combined with Faraday’s Law 𝑉 = −𝐿 𝑑𝐼
𝑑𝑡
, this allows us to define the non-linear
Josephson inductance as
Φ0
𝐿𝐽 = . (2.11)
2𝜋𝐼𝑐 cos(𝜑)
Having the inductance be non-linear results in a non-linear oscillator, thus allowing
the Josephson junction (and combination of junctions) to serve as the backbone of
building superconducting qubits. We will now introduce two of those qubits, the
Cooper pair box, and the proceeding transmon qubit.

2.2.3 The Cooper Pair Box

One of the simplest superconducting qubits is the Cooper pair box (or charge qubit),
which creates a superconducting island that is connected to a superconducting reser-
voir using the previously described Josephson junction (Fig. 2-2) [16]. The Cooper
pair box qubit’s states are dependent on the presence of additional Cooper pairs in
the superconducting island.

26
Figure 2-2: Circuit diagrams for a Cooper pair box qubit (a) and a transmon qubit
(b) where the crossed square is the symbol for the Josephson junction circuit element.
The transmon builds upon the Cooper pair box architecture by including a large shunt
capacitor to reduce sensitivity to charge noise.

The Cooper pair box Hamiltonian is defined as

𝑛 − 𝑛𝑔 )2 − 𝐸𝐽 cos 𝜑,
𝐻 = 4𝐸𝐶 (ˆ (2.12)

where 𝑛
ˆ is the Cooper pair number operator representing the number of excess Cooper
pairs on one of the islands, 𝑛𝑔 is the effective offset charge, and the reduced flux 𝜑 is
ˆ [17]. Here, 𝐸𝐽 is the Josephson energy and 𝐸𝐶 = 𝑒2 /2𝐶Σ where
a conjugate pair to 𝑛
𝐶Σ = 𝐶𝑆 + 𝐶𝐽 is the total capacitance including both the shunt capacitance and the
capacitance of the junction. We can also write the Hamiltonian in the charge basis
as
𝐸𝐽 ∑︁
𝑛 − 𝑛𝑔 )2 −
𝐻 = 4𝐸𝐶 (ˆ (|𝑛⟩ ⟨𝑛 + 1| + |𝑛 + 1⟩ ⟨𝑛|). (2.13)
2 𝑛

The Cooper pair box is operated in a regime where 𝐸𝐽 ≪ 4𝐸𝐶 so the eigenstates
are number-like. A major issue with the Cooper pair box operating as a qubit is its
sensitivity to 1/𝑓 charge noise [18]. One mitigation approach is to operate the system
at “sweet spots” where the charge dispersion slope is zero at 𝑛𝑔 ± 0.5 [19]. Operating
at the sweet spots allows one to reduce the system to a two-level qubit, as states
with the |𝑛⟩ and |𝑛 + 1⟩ Cooper pairs form avoided level crossings (Fig. 2-3). These
are also the points with the largest anharmonicity, making them advantageous for

27
Figure 2-3: The charge dispersion energies for the 3 first three qubit levels from
Eq. 2.12. As the ratio 𝐸𝐽 /𝐸𝐶 increases, the levels exponentially flatten and enter the
transmon regime.

assigning the ground and first excited states as the computational subspace |0⟩ and |1⟩.
Unfortunately, while this approach can help raise the coherence times significantly,
the location(s) of the sweet spot can change frequently due to system fluctuations,
requiring the gate voltage to be reset often.

2.2.4 The Transmon Qubit

The transmon qubit extends the Cooper pair box by including a large shunt capacitor
that allows the qubit to reduce its sensitivity to charge noise (Fig. 2-2) [20]. While
the Hamiltonian describing the transmon is equal to that of the Cooper pair box
(Eq. 2.12), the addition of a large shunt capacitor allows the transmon to operate in
a regime where the 𝐸𝐽 /𝐸𝐶 ratio significantly exceeds that of a Cooper pair box. By
increasing that ratio (𝐸𝐽 /𝐸𝐶 = 50), the charge dispersion decreases exponentially,
leading to a qubit transition frequency that is very insensitive to charge noise as
depicted in Fig. 2-3(c).

28
The transmon frequency is that of an anharmonic LC oscillator, which can be
calculated similarly to the QHO except that the transition frequency is now

1 √︀
𝜔𝑞 = √ ≈ 8𝐸𝐽 𝐸𝐶 . (2.14)
𝐿𝐽 𝐶Σ

We often want to be able to adjust the frequency of a transmon, which we can


accomplish by pairing junctions in parallel in what is known as a DC SQUID (super-
conducting quantum interference device). By applying an external magnetic flux Φext
through the SQUID, we can tune the effective critical current and thus the qubit’s
Josephson energy. Having these junctions be symmetric introduces flux noise that can
affect the transmon qubit frequencies, thus we design these SQUID loops to be asym-
metric [21]. Looking at our original transmon Hamiltonian in Eq. 2.12, we update
our Josephson energy term 𝐸𝐽 such that
√︃ (︂ )︂ (︂ )︂
𝜋Φext 𝜋Φext
𝐸𝐽′ = 𝐸𝐽Σ cos2 2
+ 𝑑2 sin , (2.15)
Φ0 Φ0

where 𝐸𝐽Σ = 𝐸𝐽1 + 𝐸𝐽2 , we define 𝑑 as the junction asymmetry parameter 𝑑 =


((𝐸𝐽2 /𝐸𝐽1 − 1)/((𝐸𝐽2 /𝐸𝐽1 + 1). The resulting transmon Hamiltonian is now

𝑛 − 𝑛𝑔 )2 − 𝐸𝐽′ cos 𝜑,
𝐻 = 4𝐸𝐶 (ˆ (2.16)

and the transmon qubit frequency is now tunable with respect to Φext

√︀
𝜔𝑞 (Φext ) = 8𝐸𝐽′ 𝐸𝐶 . (2.17)

Altogether, while the anharmonicity (the difference between the 0th and 1st state
separation and the 1st and 2nd state separation) is reduced in the Transmon, we can
now specifically address the two lowest levels as a qubit (Fig. 2-1). Additionally, the
asymmetric transmon is both tunable and resistant to charge noise allowing for higher
coherences throughout the system.

29
30
Chapter 3

Superconducting Qubit Readout

One of the critical components of any quantum computing system is the ability to
perform fast and high fidelity qubit state readout. We specifically look at quantum
non-demolition readout (QND), where the outcome of the measurement is not affected
by the readout process. The general readout process is illustrated in Fig. 3-1, where
applying a rectangular pulse at the resonator frequency allows us to probe the qubit
state-dependent dispersive shift in the resonator. That dispersive shift is picked up by
the readout signal after it interacts with the resonator and following of post processing
effort, that signal data can be post-processed and used to discriminate qubit states in
what is known as the IQ-plane. Here, the IQ-plane refers to the real and an imaginary
part, or an “in-phase” (I) and a “quadrature” (Q) part of a modulating signal that
changes the carrier’s frequency slightly.
The specific goal of this research is to optimize the control pulse shapes of the
readout process in terms of populating the readout resonators more quickly to probe
that dispersive shift and actively resetting the resonators so future operations are not
affected and can be performed efficiently. The actual post-processing of the readout
signal and subsequent discrimination of qubit states is outside the scope of this work
as it is another active area of research in improving readout [22].
We will begin this chapter by first covering how individual atoms interact with pho-
tons in a cavity through cavity quantum electrodynamics (QED). We then transition
to circuit QED, an analog of cavity QED where we now use a microwave-frequency

31
Figure 3-1: General cartoon of single qubit dispersive readout. First, a readout
resonator’s frequency is shifted by the underlying qubit state. This shifted frequency
can be probed by a readout pulse, whose reflected magnitude is effects based on that
resonator frequency shift. These effects are articulated through the post processing
of the readout signal output, where the qubit state can be determined by using
discrimination techniques on the readout signal in the IQ-plane (b).

waveguide transmission line resonator coupled to superconducting qubits that allow


us to perform quantum non-demolition readout. Finally, we scale from single-qubit
readout to multiplexed dispersive readout of multiple qubits.

3.1 Cavity Quantum Electrodynamics


In cavity quantum electrodynamics (QED), we look at the interactions between atoms
and harmonic oscillator excitations (specifically focusing on microwave photons here).
As shown in Fig. 3-2, by placing a two-level atom in a cavity, the coherent interactions
between the atom and photons passing through the atoms can be assessed using
the photons leaking from the cavity. This system can be described by the Jaynes-
Cummings Hamiltonian

(︂ )︂
† 1 𝜔𝑎
𝐻𝐽𝐶 = 𝜔𝑟 𝑎 𝑎 + + 𝜎𝑧 + 𝑔(𝑎𝜎+ + 𝑎† 𝜎− ), (3.1)
2 2
where 𝜔𝑎 is the atomic transition frequency, 𝜔𝑟 is the cavity/resonator frequency,
and 𝜎𝑧 is the Pauli-z operator. The final term 𝑔(𝑎𝜎+ + 𝑎† 𝜎− ) describes the dipole
interaction of the atom absorbing a photon from or emitting a photon to the cavity

32
at the atom-cavity coupling strength 𝑔, where 𝜎+ = |0⟩ ⟨1| and 𝜎− = |1⟩ ⟨0|.
In cavity QED, we work in what is known as the strong coupling regime, where
the atom-cavity coupling strength 𝑔 ≫ 𝜅 and 𝑔 ≫ 𝛾, where 𝜅 and 𝛾 are the cavities
and atom decay rates. In the strong coupling regime, if the cavity and atom are at
the same frequency the states |0, ↑⟩ (cavity in the ground state, atom in an excited
state) and |1, ↓⟩ (cavity in an excited state, atom in a ground state) oscillate at the
frequency 𝑔/2𝜋. Being in the 𝑛-excitation state, the energy difference between the

states is 2𝑔 𝑛 which creates an anharmonicity (𝛿) and thus the ability to function
as a qubit.
We also often aim to operate cavity QED in what is known as the dispersive regime,
where the atom and resonator detuning is large such that |∆| = |𝜔𝑎 − 𝜔𝑟 | ≫ 𝑔. By
expanding to the second-order in 𝑔 and making the unitary transformation

{︁ 𝑔 }︁
𝑈 = exp (𝑎𝜎+ + 𝑎† 𝜎− ) , (3.2)

the updated Jaynes-Cummings Hamiltonian in the dispersive regime can be approx-


imated as
𝑔2
(︂ )︂ (︂ )︂
† † 1 𝜔𝑎
𝐻JC,disp ≈ 𝑈 𝐻𝐽𝐶 𝑈 = 𝜔𝑟 + 𝜎𝑧 𝑎 𝑎+ + 𝜎𝑧 . (3.3)
∆ 2 2
𝑔2
Here, the cavity frequency is dispersively shifted by Δ
depending on the atom state
𝜎𝑧 . Here, instead of using the Pauli-z matrix, we can just approximate 𝜎𝑧 as being
𝜎𝑧 = −1 if the atom is in the excited state or 𝜎𝑧 = 1 if the atom is in the ground
state.

3.2 Circuit QED and Dispersive Readout

The cavity QED formulation can also be applied to superconducting circuits in what
is known as circuit QED [23, 24]. The “cavity” in circuit QED is now a microwave-
frequency coplanar transmission line resonator. Superconducting qubits are placed
at a voltage anti-node in the resonator with qubit-resonator interaction strength 𝑔.
We again operate in both the strong coupling regime and the dispersive regime as

33
Figure 3-2: General representation of a cavity quantum electrodynamic system. Here
a two level atom travels through the cavity in time 𝑡transit and decays at a rate 𝛾.
While traveling through the cavity, the atom with the photons in the cavity with
coupling strength 𝑔 and the photons in the cavity leak out of the cavity at a rate 𝜅.

the qubit state-dependent frequency shift can be probed to perform QND readout.
The modified Jaynes-Cummings Hamiltonian operating in the dispersive regime for
circuit QED is defined as
(︂ [︂ ]︂)︂
† † 1 𝜎𝑧
𝐻JC,disp ≈ 𝜔𝑟 𝑎 𝑎 + 𝜔𝑞 + 2𝜒 𝑎 𝑎 + (3.4)
2 2

where the dispersive shift is represented by 𝜒 = 𝑔 2 /∆ and the qubit transition fre-
quency is 𝜔𝑞 [25]. Additionally, to include the effect of introducing a drive of frequency
𝜔𝑑 on the resonator, we define a drive Hamiltonian

𝐻D = 𝜖(𝑡)(𝑎† 𝑒−𝑖𝜔𝑑 + 𝑎𝑒𝑖𝜔𝑑 ), (3.5)

where 𝜖(𝑡) is the drive amplitude at time 𝑡 and 𝜔𝑑 is the drive frequency [26]. To
choose appropriate drives, we can set 𝜔𝑑 = 𝜔𝑟 , which allows for the average number
of photons in a drive to be independent of the underlying qubit state in the absence
of any non-linear corrections. One such non-linearity we include in our simulator is
the Kerr non-linearity, a behavior that negatively shifts the cavity frequency from
the measurement frequency when in the ground state (and vice-versa for the excited

34
state). The Kerr non-linearity and its subsequent Hamiltonian are defined as

2𝑔 4 𝛿(3𝜔𝑞4 + 2𝜔𝑞2 𝜔𝑟2 + 3𝜔𝑟4 )


𝐾= , (3.6)
(𝜔𝑞2 − 𝜔𝑟2 )4

𝐻K = 𝐾(𝑎† 𝑎)2 . (3.7)

Finally, to incorporate cavity dampening as a function of the cavity photon decay rate
𝜅, we can combine our previously defined Hamiltonians and use the master equation

𝐻 = 𝐻JC, disp + 𝐻D + 𝐻K (3.8)

ˆ
𝜌˙ = −𝑖[𝐻, 𝜌] + 𝜅𝒟[𝑎]𝜌. (3.9)

Based on these calculations and parameters, there are a couple of important values
we can keep track of throughout our work. First, we can define the critical photon
number as
∆2
𝑛𝑐𝑟𝑖𝑡 = 2. (3.10)
4𝑔
The critical photon number a threshold. Exceeding that quantity of photons in a
resonator will result in the dispersive Hamiltonian defined in Eq. 3.4 no longer be a
valid approximation. For our simulator, we use these quantities as an upper bound
for how many photons can populate a resonator when generating optimized readout
pulses. Another important parameter we can calculate is the 𝑃1ph , the steady drive
power required to inject one photon into the cavity at a steady state. We can get a
close approximation of 𝑃1ph by finding the steady-state solution of 𝐻JC,disp + 𝐻D ,

𝑃1ph
1 = 𝑎† 𝑎 = . (3.11)
𝜒2 + (𝜅/2)2

Once we have 𝑃1ph , we can drive the resonator for a sufficiently long time in order
√︀
to inject 𝑃norm photons, where the drive amplitude 𝜖 = 𝑃norm 𝑃1ph and 𝑃norm is the
ratio between the applied power and 𝑃1ph .

35
3.3 Multiplexed Readout

One of the key challenges for scaling up superconducting quantum processors to


hundreds or thousands of qubits is how to scale the control and readout infrastructure
as the number of qubits increases. It is infeasible to have individual amplification
chains for every qubit as we develop chips high number of qubits. Instead, we look to
implement a frequency-multiplexed design, where multiple qubit-resonator pairs are
controlled on a single amplification chain. To individually control qubits, we space
their respective resonator operating frequencies, often in the range of 50-100 MHz.
The more qubits we want on a single drive line, the closer together those frequency
spacing’s are going to be, amplifying the need for optimized pulse shapes that can
handle various forms of cross-talk.

The important next question is how do we control these multiplexed qubits si-
multaneously? This can be accomplished by using what is known as frequency-
multiplexing. For a single pulse, separate real and imaginary parts of a complex
signal, 𝐴𝐼 (𝑡) and 𝐴𝑄 (𝑡), are inputted into an IQ-mixer with local oscillator frequency
𝜔LO . The output of the mixer and our readout signal is

𝐴in = 𝐴𝐼 (𝑡) cos (𝜔LO 𝑡) − 𝐴𝑄 (𝑡) sin (𝜔LO 𝑡) = Re (𝐴𝐼 (𝑡) + 𝑖𝐴𝑄 (𝑡))𝑒𝑖𝜔LO 𝑡 .
{︀ }︀
(3.12)

To apply multiple readout tones simultaneously, we simply calculate the overall read-
out signal as the sum of readout tones that are set at different frequencies. Our
frequency-multiplexed readout signal is now defined as

∑︁
Re 𝐴𝑗 (𝑡)𝑒𝑖𝜔RO,j 𝑡 ,
{︀ }︀
𝐴in = (3.13)
𝑗

where 𝐴𝑗 = 𝐴I,j − 𝑖𝐴Q,j and 𝜔d,j is the drive frequency for qubit 𝑗 such that 𝜔RO,j =
𝜔LO + 𝜔d,j .

36
Figure 3-3: (a) Circuit schematic of a 5 qubit multiplexed readout system, where each
qubit is coupled to an individual resonator. The resonators are capacitively coupled
to a common drive line. (b) The associated transmitted power for each qubit in the
excited and ground states as a measure of the resonator frequency.

3.4 Computational Simulation


Because this research is focused on exploring machine learning approaches to optimize
pulse shapes, it’s important to have robust simulation software that allows us to test
and iterate on many models and training hyper-parameter sets. In the previous
sections, we described a single qubit dispersive readout through the master equation.
Master equations such as in Eq. 3.9 can be solved with existing software libraries like
QuTip [27]. However, these master equation solvers are particularly slow, often taking
up to a minute of wall-clock time just to simulate a simple rectangular readout pulse.
Obviously, this is way too slow for machine learning algorithms that needs many
simulated pulses to find optimal pulse shapes, so we look to convert the master to a
Bloch equation formulation.

3.4.1 Cavity Bloch Equation Formulation

The cavity Bloch equation formulation allows us to describe the evolutionary behav-
ior of the system in terms of change over time [28]. Specifically, we can focus on the
change over time for specific operators, like the photon count operator 𝑎† 𝑎 and the
cavity state operator 𝑎. Even more importantly, we found that they offer an impor-

37
tant computational speedup compared to existing master equation and Monte-Carlo
solvers in software libraries like QuTiP, enabling us to run more accurate and more
frequent simulations [27].
We first begin with defining the cavity Bloch equation for the evolution of an
arbitrary operator 𝐴 as

𝑑𝑡 ⟨𝐴⟩ = 𝑇 𝑟{𝐴𝜌} ˆ
˙ = −𝑖⟨[𝐴, 𝐻]𝜌⟩ + 𝜅⟨𝒟[𝑎]𝐴⟩. (3.14)

Additionally, the dissipator term can be rewritten as

ˆ 1
𝒟[𝑎]𝐴 = (2𝑎† 𝐴𝑎 − 𝑎† 𝑎𝐴 − 𝐴𝑎† 𝑎). (3.15)
2

Now with the general formulation we can find the cavity Bloch equation for the
evolution of the cavity state operator 𝑎 and photon count operator 𝑎† 𝑎 using the
Hamiltonian from Eq. 3.8. For the cavity state operator 𝑎,

ˆ
𝑑𝑡 ⟨𝑎⟩ = −𝑖⟨[𝑎, 𝐻]𝜌⟩ + 𝜅⟨𝒟[𝑎]𝑎⟩
𝜅 (3.16)
= −𝑖∆rm − 𝑖𝜒⟨𝑎𝜎𝑧 ⟩ + 𝑖𝜖(𝑡) − 𝑖𝐾⟨𝑎⟩ − 𝑖𝐾⟨𝑎† 𝑎𝑎⟩ − ⟨𝑎⟩.
2

Here ∆rm = |𝜔𝑟 − 𝜔𝑑 |. In a single qubit system with only one pulse, ∆rm = 0 as we
drive the pulse at the resonator frequency. For the photon count operator 𝑎† 𝑎 we find
the cavity Bloch equation to be,

ˆ
𝑑𝑡 ⟨𝑎† 𝑎⟩ = −𝑖⟨[𝑎† 𝑎, 𝐻]𝜌⟩ + 𝜅⟨𝒟[𝑎]𝑎 †
𝑎⟩
(3.17)
= −2𝜖(𝑡) Im{⟨𝑎⟩} − 𝜅⟨𝑎† 𝑎⟩.

Next, we look to extend these formulations to multiplexed readout environments.


This extension to having multiple pulses on the same drive leads to various forms
of cross-talk. The most significantly impacting source of cross-talk the control or
readout pulse for one qubit-resonator pair having an impact on neighboring qubit-
resonator pairs. Additionally, we also look at coupling effects between neighboring
qubits and resonators. For the cavity state operator 𝑎 and photon number operator

38
𝑎† 𝑎, we update the cavity Bloch equations for each resonator 𝑖 to be defined as

∑︁
𝑑𝑡 ⟨(𝑎)𝑖 ⟩ = − 𝑖 (∆rm(𝑖,𝑗) − 𝜖𝑗 (𝑡))
𝑗∈𝑅
(3.18)
∑︁
† 𝜅𝑖
−𝑖 (𝜒(𝑖,𝑗) ⟨𝑎𝑖 𝜎𝑧( 𝑗) ⟩ + 𝐾(𝑖,𝑗) ⟨𝑎𝑖 ⟩ + 𝐾(𝑖,𝑗) ⟨(𝑎 𝑎)𝑖 𝑎𝑖 ⟩) − ⟨𝑎𝑖 ⟩,
𝑗∈𝑄
2

𝑑𝑡 ⟨(𝑎† 𝑎)𝑖 ⟩ = −2𝜖𝑖 (𝑡) Im{⟨𝑎𝑖 ⟩} − 𝜅⟨(𝑎† 𝑎)𝑖 ⟩. (3.19)


∑︀
Here, the effects of control cross-talk are encapsulated in −𝑖 𝑗∈𝑅 (∆rm(𝑖,𝑗) − 𝜖𝑗 (𝑡)), as
the absolute difference between the drive and resonator frequencies will not be equal
to zero for drives targeting other qubit-resonator pairs. Additionally, the effects of
the qubit 𝑗 interacting with resonator 𝑖 are picked up in the dispersive shifts and
Kerr non-linearities −𝑖 𝑗∈𝑄 (𝜒(𝑖,𝑗) ⟨𝑎𝑖 𝜎𝑧( 𝑗) ⟩ + 𝐾(𝑖,𝑗) ⟨𝑎𝑖 ⟩ + 𝐾(𝑖,𝑗) ⟨(𝑎† 𝑎)𝑖 𝑎𝑖 ⟩). For prac-
∑︀

tical purposes in our simulations, we reduce the set of cross-talk interactions only to
nearest neighbors, as beyond that the cross-talk contributions become negligible.

3.4.2 Simulation System Parameters

For our computational simulations, we use the 5-qubit chip system parameters from a
recent lab experiment using the same chip we will perform our own real experiments
on, for which the data can be found in Tab. 3.1 below [22].

Qubit Resonator
𝜔𝑞 /2𝜋 𝑇1 𝑇2 𝑇2𝐸 𝜔𝑟 /2𝜋 𝜅/2𝜋 𝑔/2𝜋 𝑃1ph 𝑡𝑘
𝜒/𝜅 𝑛crit
(GHz) (𝜇𝑠) (𝜇𝑠) (𝜇𝑠) (GHz) (MHz) (MHz) (MHz) (𝑛𝑠)
1 5.092 39.2 0.5 5.1 7.062 5.35 0.16 102.9 36.1 14.65 186.9
2 4.404 6.4 0.5 2.9 7.102 5.63 0.07 81.7 14.1 19.07 177.6
3 5.000 21.3 0.2 3.8 7.152 6.62 0.12 103.6 35.9 12.93 151.1
4 4.560 11.8 0.4 3.3 7.197 7.43 0.06 89.9 21.7 15.47 134.6
5 5.165 21.0 5.7 31.9 7.254 12.00 0.07 105 63.9 6.93 83.3

Table 3.1: System parameters used in our simulations. Parameter values for the
5-qubit qubit chip based off a lab experiment focused on qubit-state discrimination.

39
40
Chapter 4

Machine Learning Background

4.1 Machine Learning

Machine learning is a field of algorithms that aim to automatically learn patterns


from data. Building a model from a set of training data, machine learning algorithms
can learn generalizable parameters that describe that data and then apply those
parameters to a hidden test dataset. Algorithms that fall under the umbrella of
machine learning range from simple linear regression to 175 billion parameter deep
natural language processors [29]. There are three main families of machine learning
problems.

Supervised Learning

In supervised learning, we train models using a fully labeled dataset to perform pre-
diction on a set of unlabeled test data. This often means that supervised learning
models require large amounts of diverse training data that is trained on repetitively.
Common examples include regression, image classification, and time series prediction.

Unsupervised Learning

In unsupervised learning, the goal of the models is to find hidden structures of unla-
belled data, often in terms of relating individual data points to each other. Common

41
use cases include clustering, auto-encoding, and anomaly detection. Because unsu-
pervised learning environments do not have data labels, researchers often have to add
extra steps to interpret results, such assigning classes to learned clusters.

Reinforcement Learning

The final family of machine learning algorithms is reinforcement learning. In rein-


forcement learning, we aim to have an agent learn how to map situations to actions to
maximize an external reward signal. Importantly, this agent is forced to interact with
an environment to explore new state-action pairs and potential rewards. The agent
interacting with an environment is an important difference compared to supervised
and unsupervised learning. Here we do not have a large prelabeled training dataset,
instead, new training inputs are directly observed from the environment.
The focus of this chapter and the rest of the thesis will be on reinforcement learn-
ing algorithms and their building blocks. Specifically, we will look at reinforcement
learning algorithms backed by deep neural networks, giving us greater ability to op-
timize policies in large complex environments.

4.2 Reinforcement Learning


Reinforcement learning is a subset of machine learning where an agent aims to maxi-
mize a reward by choosing the appropriate action at a given state. Rather than com-
paring predictions with labeled datasets like in supervised learning, reinforcement
learning algorithms “learn” through the process of their agents directly interacting
with the environment. Additionally, reinforcement algorithms do not need to know
how the underlying system functions. Rather, they simply need to be able to ap-
proximate the reward for a state-action pair. The growth of reinforcement learning
research and applications have been shown through success in complex games such
as Chess and Go [8, 9], as well as robot automation and planning [10].
To best describe a reinforcement learning problem, we need to define four main
components:

42
∙ The environment defines how our agent can operate. Specifically, the possible
states an agent can be in, the possible actions an agent can take from a state,
the transition model from one state to another, and the termination conditions.
We often work in contexts where an agent has a limited view of the environment,
often just the state they are in and the available actions from that state.

∙ The policy, which is what is directly or indirectly “learned”, governs how agents
select actions in an environment given a current state. The policy may be as
simple as a lookup table or predefined action logic. More often in complex
environments, an optimal policy is optimized through approximation models
like neural networks.

∙ A reward signal defines the agent’s goal in a reinforcement learning problem.


Upon arriving in a new state, the environment provides a reward for taking the
previous action, and thus the goal of the agent is to maximize the total reward
while acting in that environment. Goals can vary in magnitude and frequency.
In some problems like learning how to play Go, the reward signal is space, simply
being 1 at the end of an episode if you win or 0 if you lose [8]. Conversely,
some environments might provide frequent rewards to enable multiple policy
considerations, such as having a small negative reward at each step to train an
agent to take as few steps as possible to reach a goal.

∙ While the reward signal gives an agent the immediate reward for arriving at
a state, it is also important for an agent to have an idea of the future value
of arriving at a state in an environment. We encode this information in what
is known as a value function, where the value is the amount of total reward
an agent can expect given its current state and learned policy. A good value
function is especially important in environments that might have many local
maxima in terms of rewards, as an agent can better plan to find global max-
ima. Learning a good value function is obviously much harder than taking in
an immediate reward, and highlights the necessity for an agent to thoroughly
explore environments that are sufficiently large and/or complex.

43
Figure 4-1: General flow of a Markov decision processes. Given a current state (𝑠) and
reward (𝑅) at time step 𝑡, an agent chooses an action that is fed into the environment
which produces the next state and reward at for time step 𝑡 + 1, continuing the cycle
until reaching a terminal state.

4.2.1 Markov Decision Processes

Reinforcement learning can be mathematically formalized in the form of finite Markov


decision processes (MDP). At each time 𝑡, the black-box environment gives an agent
its current state 𝑠𝑡 and reward for arriving at that state 𝑅𝑡 . Given that information
and the information, it’s learned about the environment, the agent then selects an
action 𝑎𝑡 that is sent to the environment to be performed. The cycle then continues
with the agent returning the next state and reward 𝑠𝑡+1 and 𝑅𝑡+1 and continues until
the agent reaches a terminal state in the environment. In a general MDP formulation,
every transition is probabilistic. If we select an action 𝑎 from state 𝑠, the next state
𝑠′ and reward 𝑅 come from a probability distribution 𝑃 . Additionally, any step 𝑡 is
only dependent on the previous step 𝑡 − 1.

4.2.2 Bellman Equation

As discussed, a key component in any agent’s success is its ability to encode the value
of particular states and actions in the environment. In the general MDP framework,
the value function is often encoded in what is known as the Bellman equation [30].

44
The Bellman equation is defined as

∑︁
𝑉 (𝑠) = max(𝑅(𝑠, 𝑎) + 𝛾 𝑃 (𝑠′ |𝑠, 𝑎)𝑉 (𝑠′ )), (4.1)
𝑎
𝑠′

where the value of state 𝑠 is determined by the available action 𝑎 that maximizes
both immediate reward 𝑅 and future value discounted by factor 𝛾. From this equa-
tion, we can find the optimal value function with value iteration and infer an optimal
policy (Section 2.3) or directly optimize an optimal policy using policy iteration (Sec-
tion 2.4).

4.2.3 Value Iteration

The first method we will use to find optimal policies in the MDP problems is value
iteration. Here, we aim to learn the best value function 𝑉 (𝑠), which can then be used
to find a deterministic policy 𝜋 ′ using the following algorithm

Algorithm 1 Value Iteration [7]


Randomly initialize 𝑉 (𝑠) for all 𝑠 ∈ 𝑆.
Initialize ∆ > 0 as a small threshold to determine if the value function has con-
verged.
loop
𝛿←0
for 𝑠 ∈ 𝑆 do
𝑣 ← 𝑉 (𝑠)
𝑉 (𝑠) ← max𝑎 (𝑅(𝑠, 𝑎) + 𝛾 𝑠′ 𝑃 (𝑠′ |𝑠, 𝑎)𝑉 (𝑠′ ))
∑︀
𝛿 ← 𝑚𝑎𝑥(𝛿, |𝑣 − 𝑉 (𝑠)|
until ∆ > 𝛿
𝜋 ′ (𝑠) = argmax𝑎 (𝑅(𝑠, 𝑎) + 𝛾 𝑠′ 𝑃 (𝑠′ |𝑠, 𝑎)𝑉 (𝑠′ ))
∑︀

Here, an agent will get better at optimizing the value function as it gains more
experience. This usually means that we need to visit many states multiple times to
fully explore our environment. One key benefit though is that knowing an optimal
value function allows one to infer an optimal policy simply by tapping into that value
function.

45
4.2.4 Policy Iteration

An alternative to value iteration is to directly optimize for an optimal policy using


what is known as policy iteration, with the general algorithm below.

Algorithm 2 Policy Iteration [7]


Randomly initialize 𝑉 (𝑠) and 𝜋(𝑠) ∈ 𝐴 for all 𝑠 ∈ 𝑆.
Initialize ∆ > 0 as a small threshold to determine value function has convergence.

(a) Policy Evaluation


loop
𝛿←0
for 𝑠 ∈ 𝑆 do
𝑣 ← 𝑉 (𝑠)
𝑉 (𝑠) ← max𝑎 (𝑅(𝑠, 𝜋(𝑠)) + 𝛾 𝑠′ 𝑃 (𝑠′ |𝑠, 𝜋(𝑠))𝑉 (𝑠′ ))
∑︀
𝛿 ← 𝑚𝑎𝑥(𝛿, |𝑣 − 𝑉 (𝑠)|
until ∆ > 𝛿

(b) Policy Improvement


policy-stable ← true
for 𝑠 ∈ 𝑆 do
old-action ← 𝜋(𝑠)
𝜋(𝑠) ← argmax𝑎 (𝑅(𝑠, 𝑎) + 𝛾 𝑠′ 𝑃 (𝑠′ |𝑠, 𝑎)𝑉 (𝑠′ ))
∑︀
if old-action ̸= 𝜋(𝑠) then
policy-stable ← false

if policy-stable = false then


Go to (a)

The differences between policy iteration and value iteration are subtle but im-
portant. In policy iteration, we select actions from a current policy compared to
computing the value of every available action and post-selecting an action. We are
also always optimizing an optimal policy in policy optimization, instead of only find-
ing an optimal policy a single time in the end once we’ve found an optimal value
function. Policy iteration approaches are especially helpful in environments that
have continuous action spaces, where it would be infeasible to compute a value for
every state-action pair. However, value iteration approaches can be advantageous in
discrete action space environments as it can be quicker to optimize [7]. We use one
of each approach in our DRL algorithms found in Chapter 5 Section 3.

46
4.3 Deep Learning and Neural Networks

The driving force behind many of the advances in machine learning over the past
decade comes from neural networks and what is known as “deep learning”. A neural
network is a biologically-inspired network graph consisting of layers of connected
nodes. The value of each node is the sum of a bias term and the product of each
connected node from a previous layer and edge weight that connects the two nodes.
That node value is then fed through a non-linear function and becomes an input for
the next layer of nodes. The values of the bias term and edge weights are trainable
parameters that are updated based on the performance goal of the problem at hand.
The ideas behind neural networks have been around since the 1940s, with research
focus waning on and off during the 20th-century [31]. The explosion in research
that has fueled so many advancements in fields such as computer vision and robotic
control has come from what we know as deep learning. With computational and
memory advancements, we can now store and train networks that have many hidden
layers (“depth”). By increasing depth, we increase the number of learnable non-linear
transformations and thus allow these networks to outperform many existing systems.
A helpful comparison is looking at an 𝑛-dimensional linear binary classifier, which can
be viewed as a neural network that only has an 𝑛 node input layer and single node
output layer. If our dataset is not linearly separable in the beginning, we might look
to perform some manual data transformation before optimizing a line of separation.
With neural networks, we can use any number of hidden layers to allow our system to
learn what the ideal non-linear data transformations are, instead of relying on human
subject experts and trial and error.

4.3.1 Feed-Forward Neural Networks

In this thesis, we focus on the most basic neural network architecture called a feed-
forward neural network (FNN), or multi-layered perceptron (MLP). As we see in Fig
4-2(a), the FNN is comprised of layers of connected neurons that map in an input
vector to an output value/vector. As shown in Fig. 4-2(b), the value of any node 𝑖 in

47
Figure 4-2: (a) Basic feed-forward neural network architecture with an input layer of
3 nodes, 2 hidden layers with a width of 4 nodes each and single output layer node.
(b) The value of each node in a neural network not in the input layer is determined
by finding the sum of a bias term and the dot product between the previous layer’s
node values and the inter-layer weight values. This computed value is then fed into
a non-linear activation function.

layer 𝑙 ∈ 𝐿 in our neural network can be defined as


(︃ )︃
∑︁
𝑎𝑙𝑖 = 𝜎 𝑙
𝑤(𝑖,𝑗) 𝑎𝑙−1
𝑗 + 𝑏𝑙𝑖 = 𝜎(𝑧𝑖𝑙 ), (4.2)
𝑗

𝑙
where 𝑤(𝑖,𝑗) is the weight between nodes 𝑖 and 𝑗 between layers 𝑙 − 1 and 𝑙, 𝑏𝑙𝑖 is
the bias weight, and 𝜎 is an activation function (see 4-3). These activation functions
are critical to neural networks because they introduce non-linearities into our neural
networks, providing a relatively efficient means for the networks to identify complex
data patterns. Here we describe the most common activation functions used in FNNs.

Linear

The linear activation function (Fig. 4-3) simply returns the node value as it was passed
into the activation function. We almost always do not use linear activation functions
in the hidden layers as they do not introduce non-linearity to the network. However,
they are commonly found in output layers that require continuous values such as in
value prediction contexts and regression contexts. The linear activation function and
its derivative are given by
𝜎(𝑧) = 𝑧 (4.3)

48
𝜎 ′ (𝑧) = 1 (4.4)

ReLU

The rectified linear unit activation function (ReLU) (Fig. 4-3) is typically the default
activation function used in the hidden layers of FNNs. The ReLU function is favored
heavily because while it introduced non-linearity, it is still very similar to that of a
linear transformation, and thus can benefit from being a good generalizer and from
easy to optimize with gradient-based methods. The ReLU activation function and its
derivative are given by
𝜎(𝑧) = max(0, 𝑧) (4.5)

⎨1, if 𝑧 ≥ 0


𝜎 (𝑧) = (4.6)
⎩0, if 𝑧 < 0

Sigmoid

Sigmoid activation functions (Fig. 4-3) are commonly used in the output layer of
neural networks when we are performing binary classifications. Typically, we find
that the sigmoid output layer is a single node, however, some problems do have
multiple output nodes when performing multiple binary classifications in a single
forward pass. The sigmoid activation function and its derivative are given by

1
𝜎(𝑧) = (4.7)
1 + 𝑒−𝑧
(︂ )︂ (︂ )︂
′ 1 1
𝜎 (𝑧) = 1− (4.8)
1 + 𝑒−𝑧 1 + 𝑒−𝑧

Softmax

The softmax activation function is similar to the sigmoid activation in that it maps
the output node value to a probability, but now working in a setting where our
output layer has multiple nodes. Softmax is widely used for multi-class prediction
and reinforcement learning applications where we can map output nodes to available

49
Figure 4-3: Common activation functions and their derivatives found in FNNs, in-
cluding the linear, ReLU and sigmoid activation functions.

actions. The softmax activation function and its derivative are given by

𝑒𝑧
𝜎(𝑧) = ∑︀ (4.9)
𝑖 𝑒𝑖

𝑒𝑧 𝑒𝑧
(︂ )︂ (︂ )︂

𝜎 (𝑧) = ∑︀ 𝑖
1 − ∑︀ 𝑖 (4.10)
𝑖𝑒 𝑖𝑒

4.3.2 Cost Functions and Regularization

The most essential component of any machine learning system is how we train it. The
first step in training a deep neural network is to define a loss function. If we follow
the architecture in Fig. 4-2(a) and state the single output node has a linear activation
function, we can consider the network to be making a continuous prediction. In that
case, a common cost function is the mean squared error over 𝑛 examples,

𝑛
1 ∑︁
𝐶(𝜃) = (𝑦𝑖 − 𝑦ˆ𝑖 (𝜃))2 . (4.11)
𝑛 𝑖=1

Like any optimization technique, neural networks are susceptible to overfitting,


especially because neural networks tend to have many more trainable parameters.
There are several potential regularization techniques we can use to prevent overfitting.

50
One example is to directly include a regularizer in the cost function as

𝑛
1 ∑︁
𝐶(𝜃) = (𝑦𝑖 − 𝑦ˆ𝑖 (𝜃))2 + Ω(𝜃), (4.12)
𝑛 𝑖=1

where Ω(𝜃) (L2 regularizer) is the squared sum of all model weights. Applying this
regularization terms penalizes the model for updating to large individual weight val-
ues, which can be a major cause of overfitting. Another regularization technique we
use is batch normalization [32]. Here, instead of feeding our neural network one input
at a time, we create a minibatch of 𝑛 inputs and normalize the input values in respect
to that batch.

4.3.3 Training Neural Networks

Now that we have defined a neural network architecture and cost function, we can
move to actually train the network. We update the network weights using a gra-
dient descent-like algorithm that computes the cost functions derivative concerning
all trainable weights in the network. Since deep neural network weights can easily
number in the tens of millions in some applications, it’s impractical to compute ev-
ery single gradient individually. Instead, we use a method called backpropagation
(“backprop”) that allows us to take advantage of the network’s layer structure and
generate these gradients by essentially just recursively applying the chain rule for
partial differentiation [33].
A particular weight or bias in a neural network is updated such that

𝜕𝐶
𝑤 := 𝑤 − 𝜂 , (4.13)
𝜕𝑤

𝜕𝐶
𝑏 := 𝑏 − 𝜂 , (4.14)
𝜕𝑏
where 𝐶 is our cost function and 𝜂 is the learning rate for the training optimization
𝜕𝐶 𝜕𝐶
loop. To calculate 𝜕𝑤
and 𝜕𝑏
for a particular layer, we rely on using the chain rule
and four key equations.

51
1. For our first key equation, we use our cost function to define the error of a node
𝑖 in any layer 𝑙 as the change in the cost function concerning the node value 𝑧𝑖𝑙
such that
𝜕𝐶 𝜕𝐶
∆𝑙𝑖 = 𝑙
= 𝐿 𝜎 ′ (𝑧𝑖𝑙 ). (4.15)
𝜕𝑧𝑖 𝜕𝑎𝑖

2. Similar to the first equation, we can also define the error of a node 𝑖 in any layer
𝑙 as the change in the cost function concerning the bias term 𝑏𝑙𝑖 . Here we can
set 𝜕𝑏𝑙𝑖 /𝜕𝑧𝑖𝑙 = 1 and
𝜕𝐶 𝜕𝐶 𝜕𝑏𝑙𝑖 𝜕𝐶
∆𝑙𝑖 = 𝑙
= 𝑙 𝑙
= 𝑙. (4.16)
𝜕𝑧𝑖 𝜕𝑏𝑖 𝜕𝑧𝑖 𝜕𝑏𝑖

3. Working recursively, any error from a node in layer 𝑙 only depends on node
errors in layer 𝑙 + 1. Thus we can use the chain rule to find

𝑙+1
𝜕𝐶 ∑︁ 𝜕𝐶 𝜕𝑧𝑗 ∑︁
𝑙+1 ′ 𝑙
∆𝑙𝑖 = 𝑙 = 𝑙+1 𝑙
= ∆𝑙+1
𝑗 𝑤(𝑖,𝑗) 𝜎 (𝑧𝑖 ). (4.17)
𝜕𝑧𝑖 𝑗
𝜕𝑧 𝑗 𝜕𝑧 𝑖 𝑗

4. Finally, we can differentiate the cost functions concerning the weights of the
network between layers 𝑖 and 𝑗 with

𝜕𝐶 𝜕𝐶 𝜕𝑧𝑖𝑙
𝑙
= = ∆𝑙𝑖 𝑎𝑙−1
𝑗 . (4.18)
𝜕𝑤(𝑖,𝑗) 𝜕𝑧𝑖𝑙 𝜕𝑤(𝑖,𝑗)
𝑙

Using these four equations, we start by passing through an input into our network
and calculating all 𝑎𝑙𝑖 and 𝑧𝑖𝑙 values. Next, we can directly calculate the output layer
𝐿 error using Eq. 4.15. With this value, we can move back across the network and
calculate all recursively calculate all subsequent errors. Finally, the values of all errors
𝜕𝐶 𝜕𝐶
∆𝑙𝑖 are used to find all 𝜕𝑤
and 𝜕𝑏
and update our network parameters.

4.4 Deep Reinforcement Learning


Now that we have an understanding of reinforcement learning and deep neural net-
works, we look to combine the two ideas with deep reinforcement. As we see in
reinforcement algorithms like value iteration (Algo. 1) and policy iteration (Algo. 2),

52
finding ideal agent policies often requires an extremely high number of episodes of
training in an environment to converge on a policy. With large and complex environ-
ments, it is impractical to try to approach these problems iteratively. There are too
many possible states and actions to store value functions in memory and there are
too many possible states to explore to ensure an optimal policy. To remedy this issue,
as pulse shaping is an extremely large and complex environment, we look to utilize
deep neural networks as part of our reinforcement learning algorithms. Typically, this
means using those neural networks as approximators for a value function or policy
distribution. Instead of storing every state’s value function result in memory, we can
use a neural network to approximate the future value of that state and then train
the neural network on the real value that is seen later. Similarly on the policy side,
instead of constantly updating a policy model, we can use a neural network to ap-
proximate the distribution of models that are trained based on the generated policy’s
performance. We will explore two of these deep reinforcement algorithms, Deep Q
Networks, and Proximal Policy Optimization in Chapter 5.

53
54
Chapter 5

Methodology

5.1 Conventional Approach

As we’ve described earlier in Chapter 3, the conventional approach to dispersive


readout pulse shaping is to use a sufficiently long rectangular pulse. Because the res-
onator’s frequency is shifted by the qubit state, that rectangular pulse can probe that
frequency shift and thus allow us to discriminate the underlying qubit states. How-
ever, this process is slow. First, a rectangular pulse sent to the resonator frequency
often needs to be fairly long (multiple microseconds) to populate and stabilize the res-
onator. Second, once we’ve finished probing, it’s important to get the resonators back
to the vacuum state. With the rectangular pulse method, we simply stop applying
the pulse and passively allow the photons to naturally decay out of the resonator.

Here we look at two approaches to generate readout pulse shapes. First, we


reimplement the existing Cavity Level Excitation and Reset (CLEAR) pulse [34] and
extend it to be optimizable in a multiplexed readout environment. Second, we use
two deep reinforcement learning techniques to generate readout pulse shapes with
much finer amount of control compared to other available methods.

55
5.2 CLEAR Pulse

5.2.1 Introduction

In circuit QED devices, a large research focus has surrounded improving readout speed
and fidelity (qubit-state classification accuracy in this case), including the develop-
ment and improvement of Purcell filters and quantum-limited amplifiers [35, 36, 37].
From a pulse shaping optimization perspective, the most successful approach has
been found through the Cavity Level Excitation and Reset (CLEAR) pulse [34]. The
CLEAR pulse relies on a set of two short segments attached to the front and end of
a typical rectangular pulse that helps populate and depopulate the resonator with
photons quicker than conventional methods (Fig. 5-1). For a single qubit, the two
amplitude values of the photon injection and reset components can be empirically
optimized depending on the target photon population number 𝑛
¯ of its subsequent
drive power 𝑃norm = 𝑃/𝑃1ph where 𝑃1ph is the drive power required for a steady-state
photon population of 𝑛
¯ = 1. The pulse amplitudes corresponding to different photon
target values need to be optimized separately because higher drive powers can force
the cavity outside of the linear regime where the steady-state photon number 𝑛0 is
now dependent on the prepared qubit state.

5.2.2 Scaling to Multi-Qubit Systems

One of the key benefits of the CLEAR pulse for single-qubit readout is the ease of opti-
mization, as only two amplitude segments need to be tuned at a given time, which can
be achieved through a simple empirical optimization. When scaling to 𝑛-qubit multi-
plexed readout, this optimization becomes a 2𝑛-parameter optimization for both the
injection and reset portions of the pulse, thus requiring a more complex optimization
protocol that can handle the higher parameter number and more complex optimiza-
tion space due to the various forms of cross-talk introduced in multiplexed readout
(Chapter 3 Section 3).

56
Figure 5-1: Arbitrary CLEAR and rectangular pulse shapes. Two amplitude segments
at the beginning of the pulse help inject photons into the resonator quickly, allowing
the resonator to reach a steady photon count in a shorter time. Two amplitude
segments are also added to the end of the pulse to actively reset the resonator to the
vacuum state.

5.2.3 Bayesian Optimization

We handle this increase in optimization complexity with a technique called Bayesian


optimization. Bayesian optimization is an effective black-box optimization scheme
that attempts to build a probabilistic model of an underlying function that makes it
simple to both identify good parameters to sample from (explore) and good parameter
candidates to maximize reward (exploit). Once we have an optimized model for our
underlying black-box environment, selecting an optimal parameter set is simply

𝑥* = argmax 𝑓 (𝑥). (5.1)


𝑥∈𝒳

To build this underlying model, we use Gaussian Processes (GPs), which are a
probability distribution over a collection of possible functions fit to a set of points.
A Gaussian process can be defined as a normal distribution with a mean vector 𝜇
and a covariance matrix Σ, where the expected value of the underlying function is
also equal to 𝜇. Additionally, the variance of each test point 𝑖 can be found along

57
the diagonal of the covariance matrix Σ. As we see, increasing the number of initial
and optimization points increases the dimension of 𝜇 and Σ, often resulting in the
optimization algorithm progressively getting slower over time (Fig. 5-2).

A key component to Gaussian processes is defining a kernel that builds our co-
variance matrix Σ. In this work, we use the common Radial Basis Function (RBF)
kernel where

||𝑥 − 𝑥′ ||2
{︂ }︂

𝐾(𝑥, 𝑥 ) = exp , (5.2)
2𝜎 2
with 𝜎 being a tunable parameter that increases or decreases the variance.

For our optimizer’s acquisition function, we use a technique called GP Upper Con-
fidence Bound (GP-UCB) to compute the ideal next sample [38]. GP-UCB exploits
the upper confidence boundary on the underlying Gaussian posterior such that we
can simply choose the maximum 𝑥 in

𝛼UCB (𝑥) = 𝜇(𝑥) − 𝜅𝜎(𝑥). (5.3)

Here, 𝜅 is a tunable parameter that adjusts how much our acquisition function fa-
vors exploring potential high reward points. With complex underlying optimization
spaces, it’s often advantageous to have higher 𝜅 to be able to explore as much of the
environment as possible.

Algorithm 3 Bayesian Optimization


Place a Gaussian process prior on 𝑓
Observe 𝑓 at 𝑥0 initial uniform random points
Update posterior probability distribution on 𝑓 based on initial points
for 𝑖 = 1, 2, ... in 𝑁 do
Select 𝑥𝑖 as the maximizer of the chosen acquisition function.
Calculate 𝑦𝑖 = 𝑓 (𝑥𝑖 )
Update posterior probability distribution on 𝑓 based on all points.
Return largest observed 𝑓 (𝑥)

58
Figure 5-2: Wall-clock time needed per optimization step when optimizing different
numbers of pulse segments. Regardless of the number of segments being optimized,
the length of each optimization step generally increases as the optimization step num-
ber increases as we expect from the Bayesian optimization algorithm. Additionally,
increasing the number of amplitude segments needing to be optimized also results in
requiring more optimization time per step throughout the entire optimization process.

Scaling

One of the disadvantages of Bayesian Optimization, and why it is not used to gen-
erate pulses that contain more than two segments, is its poor scaling. Specifically,
every update to the posterior probability distribution requires using every parameter-
observation pair, including the initial random points. More complex environments
will likely require more optimization steps to find a suitable set of parameters. This
problem is exaggerated when working in a multi-qubit system where each additional
amplitude segment creates 5 additional parameters needing to be optimized, and thus
creates an even more complex optimization environment. The effects of increasing
the number of parameters in a Bayesian Optimizer can be seen in Fig. 5-2.

59
5.3 Deep Reinforcement Learning Generated Pulses

5.3.1 Deep Q-Networks

Q-Learning

The first deep reinforcement learning algorithm we implement and experiment with
is the Deep Q-Network (DQN) algorithm [39]. DQN’s are an extension of the pop-
ular Q-Learning algorithm that allows us to explore larger state and action spaces,
as well as take advantage of modern deep neural network techniques and hardware
optimizations [40].
In tabular Q-Learning settings, we aim to find an optimal policy by creating a
Q-table that holds a value for each state-action pair. Each table entry signifies the
current value of a pair by acting in that given state, as well as the agent’s future value
by making decisions from that new state-action pair. More specifically, the Q-table
is updated with a variation of the Bellman equation (Eq. 5.4), where the state-action
pair Q-value of 𝑄new (𝑠𝑡 , 𝑎𝑡 ) is updated by summing the old value with optimal next
step in the table. Using value iteration and enough training iterations, optimal values
for the Q-table can be eventually computed. When the table is complete, an agent
can simply find an optimal solution by performing table look-ups and selecting the
action that produces the largest Q-value at each state.

𝑄𝑛𝑒𝑤 (𝑠𝑡 , 𝑎𝑡 ) = 𝑄(𝑠𝑡 , 𝑎𝑡 ) + 𝛼 * (Reward(𝑠𝑡 , 𝑎𝑡 ) + 𝛾 max 𝑄(𝑠𝑡+1 , 𝑎)) (5.4)


𝑎

Exploration vs. Exploitation

One of the key challenges in any reinforcement learning is striking the correct balance
between exploration and exploitation. Particularly, when we have large and complex
environments, we must be able to explore different parts of that environment so we
do not end up simply optimizing to a local maximum based on a random starting
point, a problem that other optimization schemes in quantum control like GRAPE
encounter [41]. In DQN, we actually handle this with a simple approach called 𝜖-

60
greedy action selection. Here, we choose a random action with probability 𝜖 or an
action generated by our training policy with probability 1 − 𝜖. A common strategy is
to start with a high 𝜖 value at the beginning of the training that eventually descends
to a value very close to zero, allowing the algorithm to “explore” in the beginning and
gradually “exploit” as training continues.

Deep Q-Networks

A major issue with tabular Q-Learning approaches is the reliance on storing the
Q-table in memory, as well as the algorithm’s necessity to explore many state-action
pairs. This becomes infeasible for environments with large and precise state and action
spaces, such as the case with multi-qubit pulse shaping. Deep Q-Networks replace
the Q-table with a deep neural network that takes in a state input and outputs the
approximated Q-values for each action (Fig. 5-3). To train the network, we store
past network predictions in a replay buffer that is periodically sampled uniformly or
by prioritizing samples with higher rewards [42]. The underlying neural network is
trained by minimizing the loss function,

𝐿𝑖 (𝜃𝑖 ) = 𝐸ˆ(𝑠𝑖 ,𝑎𝑖 ,𝑟𝑖 ,𝑠𝑖+1 )∼𝒟 [(𝑦𝑖 − 𝑄(𝑠𝑖 , 𝑎𝑖 ; 𝜃𝑖 ))2 ] (5.5)

For pulse shaping, we define our state as being the partially constructed pulse
shape and our action space being the discrete set of all available amplitudes. Because
DQN’s select one action at a time, the pulse shape is built sequentially, where 𝑅 = 0
automatically for any pulse incomplete pulse shape. While this creates to a sparse
reward signal, the eventual value of good pulse shapes can be adequately encoded
by using a high discount rate (𝛾 > 0.99). This algorithm requires us to have a
discrete action space as the network output predicts a Q-value for each possible action.
Discretization can be helpful in that the physical implementation using an AWG also
has a discrete set of amplitudes, and thus the model can mirror that action space.
Unfortunately, a large action set size also makes the neural network approximator
more difficult and time-consuming to train due to a high number of output nodes.

61
Figure 5-3: A simple Deep Q-Network architecture where we are trying to construct
a single 4 segment pulse with 2 possible amplitude options. (a) A state is provided
as the input for a deep fully-connected neural network. The neural network’s outputs
approximate the Q-value for each possible action according to the state input. (b) The
action with the maximum associated Q-value is selected, appended to pulse, and sent
to interact with the simulated or physical environment. (c) The state-action pair’s
actual Q-value is calculated in this environment and is saved in an experience replay
buffer. (d) A batch of expected and actual Q-values with their state-action pairs are
used to train the neural network. (e) The new state generated by the environment is
passed back to the input of the neural network, restarting the process.

Algorithm 4 DQN’s with Experience Replay [39]


Create an experience replay buffer 𝒟 of size 𝑁
Initialize Q-value function neural network with random weights
for 𝑒𝑝𝑖𝑠𝑜𝑑𝑒 = 1...𝑀 do
for time step 𝑡 = 1...𝑇 do
Select random action with probability 𝜖,
or select action 𝑎𝑡 = argmax𝑎 𝑄(𝑠𝑡 , 𝑎; 𝜃).
Perform action 𝑎𝑡 in environment and observe reward 𝑟𝑡 and next state 𝑠𝑡+1
Store transition (𝑠𝑡 , 𝑎𝑡 , 𝑟𝑡 , 𝑠𝑡+1 ) in 𝒟.
Uniformly samples batch of transitions from 𝒟.
For each sample 𝑗 in batch of transitions, set 𝑦𝑗 = 𝑟𝑗 + 𝛾 max𝑎 𝑄(𝑠𝑗+1 , 𝑎; 𝜃),
where all future 𝑄 values are equal to 0 at terminal states.
Perform gradient descent to minimize loss and update weights.

62
Scaling

For scaling to multi-qubit pulse engineering, we looked at a few for modifying the DQN
approach. The key underlying challenge is that for any time step 𝑡, the algorithm now
needs to select 𝑛 pulse amplitudes where 𝑛 is the number of qubits in the system.
One approach that has been used for some video game environments is to create
every combination of paired actions to become the new action space [39]. In pulse
shaping that would mean the action space size would now be 𝑎𝑛 where 𝑎 is the number
of discrete amplitude values available to be chosen. Even if we restrict that set to
210 amplitude values, that still results in an output layer with 250 nodes, which is
too large for any system both in terms of training that large of a network and even
storing it in memory. Another approach we looked at was to have separate networks
for each pulse. By sharing other pulses state information in each other’s input layer,
we could train the networks simultaneously and but keep the size of each network
similar to that of a single pulse environment. Unfortunately, this approach was unable
to converge in its training in a reasonable time-frame that could even be practical for
real-world applications.

5.3.2 Proximal Policy Optimization

Proximal Policy Optimization (PPO) is a state of the art deep reinforcement learning
algorithm designed to directly optimize policy objectives [43]. Here, we use deep
neural networks to learn an optimal distribution of policies for a given objective. The
policy distribution is sampled several times and the resulting data is used to generate a
better distribution of policies for the next step. PPO can generate entire multiplexed
pulse shapes in a single forward pass through the policy network, as each output node
is now a continuous variable that maps to a particular pulse segment. This is a huge
advantage compared to DQN where we had to cycle through the network 𝑛 times for
an 𝑛-segment long pulse. Additionally, because we work in the continuous output
regime, we can allow for much finer control of the pulse amplitudes without blowing
up our neural network size as we saw with DQNs.

63
Policy Gradients

PPO belongs to a class of reinforcement learning algorithms called policy gradient


methods, where the goal of the algorithm is to directly learn optimal policies, rather
than infer them from optimal value functions like in Deep Q-Networks. The use of
a value function may still be useful, as we will see in the PPO loss function, but it
is not required for policy gradient algorithms. The loss of a general policy gradient
algorithm is defined as

𝐿PG (𝑠, 𝑎, 𝜃) = 𝐸ˆ log 𝜋𝜃 (𝑎|𝑠)𝐴ˆ ,


[︀ ]︀
(5.6)

where the advantage function 𝐴ˆ is the absolute difference between the actual value
and expected value that we approximate ahead of time. Additionally, log 𝜋𝜃 (𝑎|𝑠) are
the log probabilities of taking each action. To minimize loss, we need to both be
able to accurately predict the value of selecting action 𝑎 and also be confident in that
action prediction.

Actor-Critic

PPO utilizes what is known as an actor-critic approach, in which an “Actor” agent


uses the current state to approximate an optimal continuous action, while the “Critic”
agent approximates the expected value for the new state-action pair. In this context,
our value approximator network is used in the context of the advantage function
ˆ where being able to predict the correct value minimizes 𝐴.
𝐴, ˆ Actor-critic methods

compare well to single action-value methods like DQN because they require less com-
putation to operate in a much larger action space and because the separation into
two processes increases overall stability [7]. Similar to DQN, we will use deep neural
network approximators for the Actor and Critic due to our large state and action set
size.

64
PPO

We can extend the policy gradient and actor-critic approaches to form the PPO
algorithm. In PPO, instead of keeping a buffer of experience that we sample from to
train our networks, we simply train on the most recent example and then discard that
experience. This makes the algorithm much more sample efficient, but also susceptible
to taking disadvantageously large step sizes that can force our policy distribution away
from an optimal set of policies. The loss function of PPO becomes

𝐿PPO (𝜃) = 𝐸ˆ min(𝑟(𝜃)𝐴,


ˆ clip(𝑟(𝜃), 1 − 𝜖, 1 + 𝜖)𝐴)
ˆ ,
[︀ ]︀
(5.7)

𝜋𝜃 (𝑎|𝑠)
𝑟(𝜃) = . (5.8)
𝜋𝜃,old (𝑎|𝑠)
To remedy the issue of losing a good distribution of policies, there are two main
alterations PPO includes compared to general policy gradients. First, we introduce
𝜋𝜃 (𝑎|𝑠)
the idea of “trust policy regions” with by using 𝑟(𝜃) = 𝜋𝜃,old (𝑎|𝑠)
[44]. Here, instead
of sampling new pulses based on a blind training update step, we actually sample
pulses from the previous policy distribution. This allows us to see which particular
pulses are having more success at a more constrained scale and train our network
more intelligently. To further restrain our network update size, we also introduce a
clipping term, which allows us to train our network based on samples taken in a range
that can be expanded/reduced by the tunable parameter 𝜖.

Algorithm 5 PPO for Pulse Generating Training [43]


Place a Gaussian process prior on 𝑓
Observe 𝑓 at 𝑥0 initial uniform random points
Update posterior probability distribution on 𝑓 based on initial points
for 𝑖 = 1, 2, ... in 𝑁 do
Use the existing policy 𝜃𝑜𝑙𝑑 to generate 𝑁 pulses.
Compute the advantage estimates 𝐴ˆ1 , ..., 𝐴ˆ𝑁 for the generated pulses.
Optimize 𝐿𝑐𝑙𝑖𝑝 with respect to 𝜃, using 𝐾 epochs and a minibatch size 𝑀 ≪ 𝑁
Update 𝜃𝑜𝑙𝑑 ← 𝜃

65
Figure 5-4: A simple Proximal Policy Optimization architecture where we are trying
to build two 2-segment pulses where any bounded continuous amplitude can be cho-
sen. We use two deep fully connected neural networks. The first is the policy network
that aims to approximate an optimal policy (the best pulse shape). The second is the
value network other that takes in the predicted pulse from the policy network and
predicts the value of that pulse in the environment.

5.3.3 Other Approaches Considered

In addition to the DQN and PPO deep reinforcement learning approaches, we also
briefly experimented with two other deep reinforcement learning algorithms in our
simulated environments.

Deep Deterministic Policy Gradient

The first additional algorithm we looked at was the Deep Deterministic Policy Gradi-
ent (DDPG) algorithm, a model-free algorithm that can learn optimal policies in high
dimensional and continuous action spaces [45]. A key difference between DDPG and
PPO is that DDPG is an “off-policy” approach where similarly to DQN’s, it collects a

66
running history of pulses and their rewards, which is then sampled for training. PPO
differs in that it doesn’t store a running history. Instead, each generated pulse and
the subsequent reward is used to train the model once and then is discarded, making
PPO more sample efficient.

AlphaZero

AlphaZero is a relatively new approach that has gained fame for outperforming mas-
ter human players in games like Chess and Go without human game play samples
for training [9]. While this project will not allow us to take advantage of self-play
applications that powered its gameplay success, the method still provides a great tool
for optimal control applications. In fact, AlphaZero was the critical algorithm behind
the most consequential work in combining deep reinforcement learning and quantum
optimal control [11] thus far.
AlphaZero employs a deep neural network in conjunction with deep look-ahead
in a guided tree search, which allows for a predictive hidden-variable approximation
of the state-action pairs. This search, a Monte Carlo Tree Search (MCTS), enables
the algorithm to look many steps ahead in the environment, while DQN and DDPG
are limited to a single-step look ahead. Additionally, the MCTS method allows for
more efficient exploration of state-action pair space as it is guided probabilistically.
DQN and DDPG on the other hand inefficiently rely on random action selection for
the initial segments of training to explore the state-action pair space. While this
MCTS approach does have these advantages, it also ended up requiring many more
simulations per model update than that of any other approach. Because the training
lacked efficiency, the training was overall too slow for practical implementation.

67
68
Chapter 6

Resonator Reset

6.1 Introduction

The first area of the multiplexed superconducting qubit readout we look at is resonator
reset. More specifically, we ask how quickly can photons be removed from the readout
resonators following a qubit state measurement. Fast readout resonator reset is critical
as the conventional method of simply waiting for photons to decay from the resonator
is time-consuming and often prohibitive to tasks like quantum error correction that
requires quick operations following a measurement [5].

Typically, readout resonator reset is realized by simply waiting a fixed time long
enough to ensure a majority of photons from all resonators have naturally decayed.
The expected photon number in a resonator 𝑟 over time is simply 𝑛𝑟 = 𝑒−𝜅𝑡 𝑃norm [34],
where 𝑃norm is the ratio between the applied power and the power required to inject 1
photon into a given resonator. In this chapter, we aim to optimize the end of a readout
pulse to reduce the time required to remove photons from all readout resonators. We
require the optimized reset pulse component to be unconditional on the underlying
qubit states. We will compare the conventional method of letting photons naturally
decay to that of the optimized CLEAR ring down pulse segments and to that of a
deep reinforcement learning backed active reset pulse.

69
6.2 Optimization Setup

To optimize ideal pulse shapes, we must first define our optimization space and setup.
Since we are only concerned with resonator reset in this chapter, we can define our
black box reward signal as simply being the negative sum of photons remaining in the
resonators. Similar to the work that first identified the CLEAR pulse, we set a target
photon number of 0.10 as a lower boundary per resonator to account for any amount
of coherent noise and thermal excitation that we may experience in experiments. The
reward function for the rest of readout resonators 𝑅 is defined below as:


⎨0, if ∀𝑟 ∈ 𝑅, 𝑛𝑟 ≤ 0.10

Reward = (6.1)
⎩− ∑︀

𝑛𝑟 , otherwise
𝑟∈𝑅

For populating the resonator(s) before reset, we apply a sufficiently long rectangu-

lar pulse at a drive power 𝑃 where 𝑃 = 𝑃norm 𝑃1ph , in order to inject 𝑃norm photons
into the resonator(s). For all forthcoming simulated results, we set 𝑃norm = 4. In our
simulations, that rectangular pulse length is 3 microseconds to ensure stable photon
counts in our 5 qubit setup. This is longer than we may really need in an experimental
environment, but in simulation, it ensures the photon counts are fully stabilized.

For our optimized reset pulse shapes, a total of 210 amplitudes values between
√ √
− 𝑃 and 𝑃 can be selected by the optimization methods. For methods that pro-
duce amplitudes in a continuous space like CLEAR and PPO, the outputted contin-
uous amplitude is simply mapped to the closest available discrete amplitude. For all
pulses used in the simulation, we convolve the generated pulse shapes with a Gaussian
kernel (𝜎 = 5) to encompass the effects of a non-ideal step response [46, 47].

For CLEAR pulses generated using Bayesian optimization, we use 250 random
initial points and up to 2,500 optimization points. For PPO generated pulses, we
allow up to 51,200 training episodes, as each episode is much quicker compared to
Bayesian optimization.

70
6.3 Single Qubit Simulation

6.3.1 Results

We start with a single qubit resonator reset setup, comparing natural photon decay
following a rectangular pulse, a CLEAR pulse, and pulses generating using DQN
and PPO. For these single qubit simulation results, we use the qubit and resonator
parameters from Qubit 1 in Tab. 3.1 and ignore the effects of all other qubits and
resonators. Here we find passive natural decay requires us to wait for 700 ns for the
photon count to reach the target population count.

The optimized pulse shapes can be seen in Fig. 6-1(a). As expected, with the
absence of other drives and qubit resonator pairs, the ideal pulse shape mirrors the
CLEAR pulse. The first half of each pulse is largely at the most negative amplitude
in the available range, while the second of the pulse is largely at the most positive
available amplitude. This pattern agrees with previous work in reset and injection
that has found that overshooting a particular target (reset or photon inject), followed
by a correction is often faster then the typical rectangular pulse followed by turning
off the pulse [34, 48, 49].

The DQN pulses and PPO pulses are 250 ns long, while the CLEAR pulse is 260 ns.
This constitutes a 64.3% and 62.9% time decrease respectively, or around 2.41 and
2.35 photon decay times. The extra 10 ns is an artifact of the CLEAR pulse requiring
equal length amplitude segments, and cutting that CLEAR pulse 10 ns short would
still allow us to reach our reset target. We see the evidence of this in Fig. 6-1(b)
where the time-dependent photon counts for the CLEAR pulse and the reinforcement
learning-backed methods are practically identical. Additionally, these photon counts
show how the initial negative amplitude segment decreases quickly but overshoots,
but the second positive amplitude compensates and gets the photon count to zero.

71
Figure 6-1: (a) Generated CLEAR, DQN and PPO pulse shapes for single qubit
readout resonator reset. In this case, the conventional rectangular readout pulse
passively waits for the photons to decay from the resonator, thus the pulse amplitude
is 0. (b) Photon counts over time due to the generated pulse shapes.

72
6.3.2 Comparing DQN and PPO Optimizations

While the pulses generated by DQN and PPO are largely identical and allow us to
reach our target photon count in equal time, there are some stark differences in the
training cycles of the two methods. Particularly, we see that DQN training cycles are
much slower and much more inefficient compared to PPO training cycles. For our
reported 250 ns reset pulses, it took DQN about 79,000 training episodes (or about 2
million training steps) to find an optimal pulse. PPO on the other hand only required
about 4,000 single-step training episodes, proving to be a much more sample-efficient
approach.

These differences are illustrated in Fig. 6-2. Here we represent the different
pulses generated using t-SNE, a method that reduces high dimensional data to a
two-dimensional space where similar high dimensional points are spatially closer to-
gether in the reduced dimensional space [50]. We first can look at the top row of
Fig. 6-2, where a marker’s, or pulse shape’s, color is determined by the number of
remaining photons. Here the DQN algorithm (top-left) needs to explore a wide range
of possible pulse shapes, often producing similar pulse shapes many times. This isn’t
surprising as the DQN algorithm employs a greedy exploration strategy that forces
the algorithm to choose random amplitudes at a decreasing frequency. Unfortunately,
this is the smallest amount of exploration and training episodes we found that could
produce this optimal pulse shape. The PPO algorithm (top-right) on the other hand
does not require random exploration and is much more sample efficient, where it
spends about half of its training episodes fine-tuning good pulse shapes compared to
exploring poor-performing pulse shapes.

In addition to the neural network scaling issues discussed in Chapter 5 Section


3.1, DQN provides no advantage in pulse length minimization compared to PPO (it
actually performs worse in Chapter 7 Section 3’s single qubit injection setup) and is
considerably slower to train. Thus, we do not try to employ DQN for future multi-
qubit environments and work solely with CLEAR pulses and PPO generated pulses.

73
Figure 6-2: 2 dimensional representation of 25 dimensional pulses generated by DQN
and PPO using t-SNE. With t-SNE, similar high dimensional data points are situated
close together in the 2 dimensional representation. On the top row, each marker
(pulse) is colored based on its reward, or number of photons remaining while on
the bottom row each marker is colored based on the iteration it was made in its
optimization cycle. Here we show every 50th pulse generated.

74
6.4 5 Qubit Simulation

6.4.1 Results

Following our single-qubit readout reset pulse optimization, we expand our work to
generate optimal pulse shapes for a simulated environment with 5 qubits (Tab. 3.1).
We again use the conventional passive reset setup as our base model and look to
improve upon it using a CLEAR pulse backed with Bayesian optimization, as well
as a PPO generated pulse. As discussed in Chapter 3 Section 3 and 5, expanding to
a multiplexed setup introduces various forms of cross talk that can make ideal pulse
shapes more complex and harder to optimize.
For our 5 qubit multiplexed readout setup, a passive reset setup requires waiting
for 700 ns for all resonators to have photon counts below our target of a photon number
of 0.10. We find that resonator reset pulses generated by PPO are minimized to a
length of 380 ns. For CLEAR pulses backed by Bayesian optimization, the minimum
pulse length is 460 ns. Compared to the 700 ns required for the conventional passive
photon decay method, this constitutes a 45.7% (∼ 2.18*𝑡𝑘,𝑎𝑣𝑔 ) time decrease for PPO
and a 34.3% (∼ 1.64 * 𝑡𝑘,𝑎𝑣𝑔 ) time decrease for CLEAR respectfully.
The pulse shapes generated by the various methods can be seen in Fig. 6-3(a).
First, looking at the generated CLEAR pulses, we see that they are following a similar
pattern as in the single-qubit reset optimization where we have a negative amplitude
followed by a positive one. However in this multi-qubit environment, the amplitudes
no longer are at the most negative and positive values of the available amplitude
range. This can be attributed to two facts. First, in the multi-qubit setup all reset
pulses are the same length, largely dependent on the individual resonator that needs
the most time to reset. This means that the pulses for the other resonators need to
find pulses for a specific pulse length that isn’t necessarily its minimum, and thus
amplitudes can be smaller in magnitude (see Section 4.2 for more details). Second,
the multi-qubit environment also introduces cross-talk from off resonant pulses and
from coupling effects between resonators and non-paired qubits, which both affect
what the optimal pulse shapes.

75
Figure 6-3: (a) Generated CLEAR, DQN, and PPO pulse shapes for single qubit
readout resonator photon injection and stabilization. In this case, the conventional
rectangular readout pulse passively waits for the photons to decay from the resonator,
thus no pulse amplitude is applied. (b) Photon counts over time caused from the
resulting generated pulse shapes. (c) Distribution of 25 generated CLEAR reset
pulses.

76
Moving to the PPO generated pulses, we see that these pulses are more complex
compared to CLEAR pulses. The PPO generated pulse amplitudes gravitate to the
highest and lowest available amplitudes, fluctuating between the two at various fre-
quencies. We can also observe the photon count overtime per resonator with the
various reset pulses applied in Fig. 6-3(b). Here we see that similar to a passive reset,
the CLEAR pulse photon counts mostly tend to strictly decrease over time. PPO
pulse photon counts on the other hand fluctuate between decreasing and increasing
frequently, likely an artifact of certain resonator states being easier to remove large
amounts of photons from, even if that requires slightly increasing the photon count
to get to those states.

6.4.2 Flexibility in Generated Pulse Shapes

One common occurrence we start seeing in multi-qubit pulse shaping optimizations


(in the context of both reset and injection) is the effect of the resonator’s photon
decay time 𝑡𝑘 on the optimized pulse shapes. More specifically, for resonators with
lower photon decay times, the range of pulses that can fulfill the reset goal expands as
shown in Fig. 6-3(c). This isn’t too surprising given our system has resonators with
varying 𝜅 and 𝑡𝑘 values, and those with lower 𝑡𝑘 will naturally decay quicker. Because
we require all 5 pulses to be equal in length in our optimization setup, this means
some individual pulses are not necessarily optimal (particularly pulses for resonators
4 and 5) as they could potentially be quicker compared to the other pulses. However,
the overall system is still at its minimum required pulse length.

6.5 Summary
In this chapter, we compared optimized active readout resonator reset pulse shapes
compared to passively turning off a readout reset pulse and letting photons naturally
decay from the resonators. In a single-qubit environment, we found deep reinforce-
ment learning methods, DQN and PPO, produced largely identical pulse shapes to
the 2 amplitude-segment CLEAR reset pulse. These optimized reset pulses were 250-

77
260 ns long compared to passive reset which required 700 ns. The single-qubit reset
optimization cycles revealed that DQN compared unfavorably to PPO as it required
much more training and optimization space exploration compared to PPO to generate
pulse shapes of the same length. Next, we looked at multi-qubit readout resonator
reset. Here we found that PPO generated more complex pulse shapes compared to
the CLEAR reset pulse shapes. The PPO generated pulse shapes also outperformed
the CLEAR reset pulse shapes, requiring 380 ns compared to 460 ns. Both optimized
pulses were substantial improvements to the passive photon decay method that re-
quired 700 ns.

78
Chapter 7

Resonator Photon Injection

7.1 Introduction

In the second step of multiplexed superconducting qubit readout, we look at “photon


injection,” or the process of populating and stabilizing the readout resonators with
photons. Here, we are concerned about generating a readout pulse shape that can
allow us to probe the dispersive shifts as quickly as possible. The faster photons can
be injected and stabilized in the resonators, the faster we can probe the dispersive
shift. With the conventional rectangular pulse, we typically see the resonator photon
counts oscillate and reach a steady-state based upon the normalized drive power 𝑃norm
and any cross-talk interactions that may be encountered in a multiplexed readout
system. Fig. 7-1(b) shows the photon counts of our simulated five-qubit system with
𝑃norm = 4. As with resonator reset pulse shaping, resonators with lower photon decay
times (e.g., resonator 5) typically reach a steady photon count quicker than other
resonators. Here we will still enforce that all pulse shapes for a given single- or multi-
qubit system are equal length to each other and aim to optimize for pulse shapes that
shorten overall the time needed to steady those photon counts as much as possible.

79
Figure 7-1: A basic rectangular readout pulse at a normalized drive power 𝑃norm = 4
(a) and the resulting photon counts for the resonators in a 5-qubit system (b). There
are three main sections of the readout pulse that we consider. The first section is the
photon injection optimization range at the beginning of the pulse, which we actively
aim to minimize the required length √ of. Second, the middle section is a constant
amplitude section that is driven at 𝑃 and is equal in length to the preceding pulse
section that is being length optimized. Finally the last segment is reset optimization
section that was discussed in Chapter 6. As we seen in (b), the photon counts from
a rectangular pulse have an oscillating evolution where the photon counts eventually
stabilize based on the normalized drive power 𝑃norm .

7.2 Optimization Setup

Similar to our resonator reset optimization setup, we require the optimized pulses
to be unconditional on the underlying qubit states. We still allow a set of 1,024
amplitudes available to be selected, however, the minimum and maximum amplitudes

are now 0 and 2 𝑃 respectfully. One major concern we now have with resonator
photon injection, especially in simulated environments, is ensuring we do not exceed
the critical photon numbers that cause the dispersive Jaynes-Cummings Hamiltonian
approximation (Eq. 3.4) to break down. To combat this in a simulated environment,
we introduce a tunable penalty term Φ that is set to a large value (> 100) to avoid
pulses that inject too many photons in the resonators. The penalty is applied if any
resonator’s critical photon number is exceeded during the readout pulse.

80
An additional difference between the photon injection and reset pulses is that we
do not optimize the entire photon injection pulse. This follows the CLEAR pulse
setup, where we only optimize the beginning of the photon injection pulse to speed
up the photon injections and stabilization [34]. For simplicity, we have the length of
the optimized portion and length of the unoptimized portion of the photon injection
pulse be equal as illustrated in Fig. 7-1(a).

In simulation, we cannot directly calculate how well a readout pulse shape will
enable us to discriminate qubit states. Instead, we aim to find pulses that inject
a stable number of photons determined by the normalized drive power. Reaching
a stable photon count allows us to probe the separation in the resonator states (as
depicted in Fig 7-5) based on the underlying qubit state. The conventional single-
amplitude rectangular pulse already does this in the real world, but it is a slow process
to get the photon counts and resonator states to stabilize. Thus to find optimal
pulses, we set the reward for the simulated setup as the sum of the absolute difference
between a resonator’s photon count (𝑛𝑟 ) and its stable photon count determined by
the normalized drive power (𝑛ˆ𝑟 ). For a pulse of length 𝐿, we look for a stable photon
count for the last 𝐿′ nanoseconds of that pulse. Because minor oscillations can still
occur, we define a stable photon count as one that is within 0.10 photons for the
entire time window of 𝐿′ nanoseconds. If this condition is not met, the reward is any
penalty accrued plus the negative sum of the absolute differences between the photon
counts and the stable photon counts for the last 𝐿′ nanoseconds. The full reward
function for the set of resonators 𝑅 is defined below as:


⎨−Φ, if ∀𝑟 ∈ 𝑅, ∀𝑙 ∈ [𝐿 − 𝐿′ , 𝐿], |𝑛ˆ𝑟 − 𝑛𝑙𝑟 | ≤ 0.1


Reward = 𝐿
|𝑛ˆ𝑟 − 𝑛𝑙𝑟 |, otherwise.
∑︀ ∑︀
⎩−Φ −


𝑟∈𝑅 𝑙=𝐿−𝐿′
(7.1)

As in our reset pulse optimization, for CLEAR pulses generating using Bayesian
optimization, we use 250 random initial points and up to 2,500 optimization points.
Then again for PPO generated pulses, we allow up to 51,200 training episodes.

81
7.3 Single Qubit Simulation Results

We begin with optimizing single qubit system pulse shapes, comparing the conven-
tional single-amplitude rectangular pulse to optimized CLEAR, DQN, and PPO gen-
erated pulses. Like with our resonator reset simulations, we use the qubit and res-
onator parameters from Qubit 1 in Tab. 3.1 and ignore all other qubits and resonators
to create a single-qubit system. We set the stable photon count length, 𝐿′ , to 100ns
to minimize the required pulse length as much as possible.

Following optimization training cycles, we see the optimized pulses allow us to


substantially decrease the necessary length required to reach a stable photon count
(Fig. 7-2). The conventional single amplitude rectangular readout pulse requires
1470ns to reach a stable photon count. In our CLEAR and PPO pulse, we can
minimize the required pulse length to 550ns, a 62.6% reduction (around 4.92 photon
decay times). For DQN, the shortest pulse we were able to optimize for was 600ns, a
59.2% reduction (around 4.65 photon decay times).

In terms of the pulse shapes generated by the optimization methods, we see that
similarly to the reset pulses, the DQN and PPO pules seem to follow a similar pattern
to the CLEAR pulse. Again, we largely see that the pulses optimize to the highest
available amplitude in the first half of the optimizable range, while the second half
amplitude is the lowest available amplitude. This again follows the idea that ideal
way to shape these single-qubit readout pulses is to overshoot our target initially and
follow up with compensating amplitude.

It is unclear why the DQN was unable to find optimal pulse shapes at lower pulse
lengths. It is worth noting that the DQN pulse struggles to mirror the generated
CLEAR pulse like the PPO generated pulse largely can. Similar to the discussion
in Chapter 6 Section 3.2, the DQN pulse again requires an incredibly larger number
of training episodes compared to PPO. Thus, paired with its reduced performance
in shortening the optimized pulse length, we again do not use DQN when scaling to
multi-qubit environments.

82
Figure 7-2: (a) Generated CLEAR, DQN, and PPO pulse shapes for single qubit
readout resonator photon injection and stabilization. In this case, the conventional
approach is a single amplitude rectangular pulse. (b) Photon counts over time as a
result of applying the generated pulse shapes.

83
7.4 5 Qubit Simulation Results

7.4.1 Injection Results

For our 5-qubit multiplexed readout simulation, we use the system parameters from
Tab. 3.1 and compare the conventional rectangular pulse approach, a CLEAR pulse
optimized with Bayesian optimization, and finally a deep reinforcement learning ap-
proach using PPO. Similar to the single-qubit setup, We set the stable photon count
length, 𝐿′ , to 100ns.

The conventional set of single-amplitude rectangular readout pulses requires a


pulse length of 1610ns to have all five resonator photon counts reach a steady target.
Using a CLEAR pulse optimized with Bayesian optimization, we reduce the required
pulse length to 780ns, a 51.6% reduction. Like we saw in multi-qubit reset, pulse
shapes optimized using PPO produced either further pulse length reduction and only
require a 600ns pulse length, a 62.7% reduction compared to a rectangular pulse.

The optimized CLEAR and PPO pulse shapes can be found in Fig. 7-3. With the
optimized CLEAR pulse shapes, we see a similar pattern to what we saw with the
multi-qubit CLEAR reset pulses. We continue to see a high amplitude segment fol-
lowed by a lower amplitude segment, but the magnitudes of both amplitude segments
are mostly not at the minimum and maximum of the available amplitude range for the
various reasons discussed in Chapter 6 Section 4.1. With PPO generated pulses, we
again see a similar pattern to the multi-qubit reset environment, where the optimized
amplitudes largely stay at the amplitude range boundaries, and switch back and forth
at various frequencies. What seems to be happening with the PPO generated pulse
shapes here and in the multi-qubit reset environment is that it’s still following the
idea that we can overshoot our target and then compensate with the latter parts of
the pulse shape. With these PPO generated pulse shapes, it seems to be doing that
multiple times in the same pulse shape, allowing it to reduce the required pulse length
compared to the multi-qubit CLEAR pulses.

84
Figure 7-3: Generated pulse shapes for photon injection and stabilization using a
rectangular pulse, (a) CLEAR pulse and (b) PPO pulse. In this setup, the amplitudes
of the first half of the
√ pulse are optimized while the second half of the pulse is at a
constant amplitude 𝑃 .

85
Figure 7-4: Resonator photon counts for the full readout pulse (photon injection and
reset) using rectangular pulses (left), optimized CLEAR pulses (center), and pulses
generated using PPO (right).

7.4.2 Entire Readout Pulse Results

Now, that we have optimized both readout reset and photon injection pulses, we can
combine the two to generate the full readout pulse shapes. To recap, the conven-
tional rectangular readout pulse followed by natural photon decay requires 2310ns,
an optimized CLEAR pulse requires 1440ns, and an optimized PPO generated pulse
requires 980ns.

The time-dependent photon counts of each applied pulse are illustrated in Fig. 7-4.
All methods seem to rely on rapidly injecting photons and backing off before stabi-
lizing with some of the resonators (1 and 3 in particular). While these photon spikes
do occur when injecting the resonators with photons, it’s important to remember we
purposely optimize for pulses that do not allow for resonators to have their critical
photon number exceeded.

More informative perhaps is observing the IQ-plane trajectories of the resonator


states in Fig. 7-5. Here, we see that cavity states are separated based on the underly-
ing qubit states. In our optimized CLEAR and PPO pulse shapes, we can approach
those separated cavity states much quicker as shown by the increase in the marker
separation, as well as take more direct trajectories to those steady states. While both
the CLEAR pulse and PPO generated pulse IQ-planes highlight how much quicker
the resonator state’s are changing, we do also see that typically, PPO generated pulses

86
Figure 7-5: IQ-plane cavity trajectories per resonator for each type of generated
full readout pulse (rectangular, CLEAR, PPO). Trajectories with red markers are
associated with a system where all 5 qubits are in the excited state, and vice versa
for with ground state trajectories with blue markers. Markers are spaced 20ns apart
to highlight the speed in which cavity states are reached. Each IQ-place also has
a dashed grey circle that represents a resonator state with 4 photons (stabilization
target), and a smaller dotted grey circle that represents a resonator state with 0.10
photons (reset target).

seem to be taking more direct paths then those from the CLEAR pulses (especially
resonators 2 and 4).

One IQ-plan trajectory that stands out is for the PPO pulse in resonator 5
(bottom-right of Fig. 7-5). Here we see that unlike most other trajectories, this
one does not have a very direct and efficient trajectory. The reasoning behind this is
the same as we saw with the different CLEAR pulses being found for CLEAR reset
pulses in Fig. 6-3(c). Because resonator 5 has a much lower photon decay time than
the other resonators, it can naturally stabilize its photon count and naturally reset
with conventional methods much quicker than the other resonators. When optimizing
multiple pulses with the same length, this means there is likely to be a large number
of possible pulses, many not entirely efficient, that allow us to reach our goal states.

87
One could try to optimize different lengths for each of the individual pulses, but the
total length of the multiplexed readout pulse would likely still be the same as it’s
already dependent on the resonators with higher photon decay times.

7.5 Summary
In this chapter, we compared optimized readout pulse shapes for injecting photons into
readout resonators as quickly as possible. We compared these optimized pulse shapes
to the conventional method of using a single-amplitude rectangular pulse shape. In the
single-qubit environment, we saw substantial pulse length reductions in the optimized
pulse shapes compared to the rectangular pulse that required 1470 ns. The CLEAR
and PPO generated pulse shapes, which were basically identical, produced 550 ns
long pulse shapes while DQN produced a 600 ns pulse shape. Again, similar to the
optimized reset pulses, DQN required a much more intense training cycle compared
to PPO, even though in this case it produced a longer pulse shape. In the multi-qubit
environment, we again dropped DQN and optimized a CLEAR and PPO generated
pulse shape. Here, we again saw substantial pulse length reductions. While the
rectangular pulse needed 1610 ns, the CLEAR pulse required 780 ns and the PPO
generated pulse only required 600 ns.

88
Chapter 8

Experimental Setup

In this chapter, we will briefly discuss how the actual experiment will be set up. First,
we will discuss the actual cryogenic hardware setup that will be used to conduct
these pulse shaping optimization experiments. Then we will discuss the experimental
designs for both optimizing ideal resonator reset pulses and readout resonator photon
injection pulses on the actual hardware.

8.1 Measurement Setup

As discussed before, these experiments and simulations are performed using a super-
conducting 5 qubit multiplexed chip. In the experimental measurement setup, the
qubit chip resides in the mixing chamber of a cryogenic dilution refrigerator that
allows us to lower temperatures to around 10 mK. Qubit control and readout pulses
are programmed and sent to an arbitrary waveform generator (AWG), whose output
signal is up-scaled to the various qubit transition frequencies using an IQ-mixer and
local oscillator. The control and readout pulses are then combined and carry on to
the 5-qubit chip.
Once the readout signal acquires the qubit-state dependent phase shifts, we need
to be able to measure that output signal. From the chip, as is seen in Fig. 8-1(a),
a series of amplifiers boost the readout signal to various dilution refrigerator tem-
perature stages. Once the signal reaches room temperature, the signal is once again

89
Figure 8-1: (a) Simplified setup for qubit control and multiplexed readout for a
superconducting multi qubit system. On the left, pulse shapes for qubit control
and readout are created, combined, and sent to the 5-qubit chip. On the right,
after passing through the chip, the readout signal returns to room temperature after
passing through a chain of amplifiers, allowing us to acquire measurement data. (b)
A magnified view of the 5-qubit chip in the super cooled mixing chamber. The chip’s
qubit’s are connected to a single drive line in a multiplexed fashion, allowing us to
control and readout all 5 qubits with a single multiplexed signal.

amplified and then sent in a heterodyne detector that eventually allows the signal’s IQ
components to be digitized and used to calculate the values used in our experimental
reward functions.

8.2 Labber Pulse Optimization Driver

In our experimental setup, control and readout pulses are programmed in the Labber
software program that interfaces with multiple arbitrary waveform generators [51]. To
be able to perform readout pulse optimization in the various experiments, we created
a custom Labber driver to construct an online optimization loop. This driver both
sends in custom pulse shapes generated by the optimization software, as well as takes
in the outputted readout data that forms our reward spaces described in the chapter’s
upcoming sections. Importantly, the optimization software is the same as what is used
in the simulated environment, as the overall software package is modular such that
different environments, both simulated and experimental, can be easily switched in
and out.

90
8.3 Resonator Reset Experiment Design

Like in our simulation environment we can optimize experimental resonator reset


pulses by relying on calculating the number of photons in the resonator’s following
that optimized reset pulse. In the experimental setting, we use what is known as
Ramsey experiment, which allows us to extract the number of remaining photons in a
cavity following a pulse [52, 53]. As seen in Fig. 8-2, we apply a pulse sequence where
we first use a control pulse to set various qubit states, followed by a readout pulse
with or without an optimized reset portion. Following this, we apply a 𝜋/2-pulse,
wait time 𝑡𝑅 , and apply a second 𝜋/2-pulse. We perform this experiment for a range
of 𝑡𝑅 times. With this collected data, we can fit the following function,

1
𝑆(𝑡𝑅 ) = [1 − Im{exp{−(Γ2 + 𝑖∆)𝑡𝑅 + 𝑖(𝜑0 − 2𝑛0 𝜒𝜏 )}}] , (8.1)
2

where 𝜏 = (1 − 𝑒−𝜅+2𝜒𝑖 )/(𝜅 + 2𝜒𝑖), Γ2 = 1/𝑇2𝐸 , and 𝑛0 is the number of photons


in the resonator at the start of the Ramsey experiment [34]. Given that we have all
parameter values already from first characterizing our multi-qubit chip, we can fit the
acquired trace data and solve for 𝜑0 and 𝑛0 in order to calculate the resonator photon
count following a readout reset pulse.

8.4 Resonator Photon Injection Experiment Design

Following pulse shape optimization for readout resonator reset pulse shapes, we can
then look to optimize experimental pulse shapes for injecting photons into the readout
resonators. Here, we no longer need to worry about measuring actual photon numbers.
Instead, we care about maximizing the qubit-state assignment fidelity ℱassignment while
minimizing the readout pulse length.

The fidelity of qubit-state assignment is described as the probability of assigning


a qubit state (ground or excited) correctly based upon how that qubit is explicitly

91
Figure 8-2: Sample Ramsey experiment pulse sequence. Here, a single Ramsey ex-
periment begins by applying a set of 𝜋-pulses to initialize the qubit states. This is
followed by a rectangular readout pulse that has an optimized reset component at
the end to empty the resonator as quickly as possible. Following the readout pulse,
we apply two sets of 𝜋/2-pulses, spaced apart by a time 𝑡𝑅 . These 𝜋/2-pulses rotate
the qubit-state dependent Bloch vector on the equator, and then back to the z-axis.
During that time between the two 𝜋/2-pulses, various amounts of detuning occurs
which affects how vector returns back to it’s original vector position following the
second 𝜋/2 pulse.

prepared. For a single qubit 𝑖, the qubit-state assignment fidelity is

ℱ𝑖 = 1 − [𝑃 (0𝑖 |𝜋𝑖 ) + 𝑃 (1𝑖 |∅𝑖 )]/2, (8.2)

where ∅𝑖 denotes preparing the qubit in the ground state and 𝜋𝑖 is describing putting
the qubit in the excited state, while 0𝑖 and 1𝑖 are the predicted qubit states. To
expand this measure of fidelity to a multi-qubit system, we take the geometric mean
of all the individual qubit-state assignment fidelities. In our 5-qubit state setup this,
the overall qubit-state assignment fidelity is

ℱassignment = (ℱ1 ℱ2 ℱ3 ℱ4 ℱ5 )1/5 . (8.3)

To obtain an optimized readout pulse’s qubit-state assignment fidelity, we can


simply follow the steps required to obtain that fidelity using a conventional mul-
tiplexed rectangular pulse. First, we send a readout pulse to the multi-qubit chip

92
that probes the various qubit-state dependent resonator phase shifts. Upon passing
through numerous amplifiers and reaching room temperature, the readout signal is
then fed into a heterodyne detector. In the heterodyne detector, the readout signals
I and Q components are digitized, providing the data source that can subsequently
be used to discriminate the underlying qubit states.
There are a few approaches we can take to performing qubit-state discrimination.
A standard approach is to use a matched filter where we apply an integration kernel to
the outputted readout signals. This integration kernel reduces the readout signals to
a single value that is dependent on the signal phase, for which a simple discriminator
can be used to determine the qubit state [54]. Other methods employed include
support vector machine’s (SVM’s) that classify the demodulated readout signal, as
well as deep neural networks that classify raw readout signal data [55, 22]. Regardless
of the method, we start by training/fitting a discriminator using the conventional
rectangular readout pulse first. Then we apply the trained/fitted discriminator to
readout signal outputs created generated by the optimized readout pulse shapes,
trying to maximize ℱassignment .

93
94
Chapter 9

Conclusion and Future Work

9.1 Conclusion
In this research thesis, we have demonstrated that deep reinforcement learning meth-
ods can be employed to optimize multiplexed superconducting qubit readout pulses
(Table 9.1). We looked at a variety of optimization methods to improve upon the
conventional approach of applying a single-amplitude rectangular pulse followed by
a passive wait for photons to naturally decay from the resonator. These methods
included combining Bayesian optimization with the Cavity-Level Excitation and Re-
set (CLEAR) pulse, as well as two deep reinforcement learning methods, Deep Q-
Networks (DQN) and Proximal Policy Optimization (PPO).
First, we examined and expanded upon the CLEAR pulse, a pulse shaping tech-
nique that adds two tunable amplitude segments to the beginning and to the end
of a typical rectangular pulse. These segments help inject photons into the readout
resonator quicker as well as actively reset those resonator’s following a measurement’s
conclusion. Because the original research regarding the CLEAR pulse only looked at
optimizing a single-qubit system, and thus a single readout pulse, we expanded upon
that research to support multiplexed readout by introducing Bayesian optimization.
This allowed us to tune amplitude segments pairs for all five pulses simultaneously
and in a data-efficient manner. In single-qubit readout, the optimized CLEAR pulses
provided a 62.7% pulse length decrease compared to the conventional rectangular

95
readout pulse. In our multi-qubit readout simulation system, the CLEAR pulses
provided a 46.3% pulse length decrease.
Next, we added more fine-tuned control using deep reinforcement learning (DRL).
In particular, we looked at two methods. Deep Q-Networks (DQN) uses deep neu-
ral networks to predict the value of selecting a specific amplitude from a discrete
space. With these value approximations, we can construct pulse shapes one ampli-
tude segment at a time. The second DRL method employed was Proximal Policy
Optimization (PPO), a method that directly predicts the single best amplitude from
a continuous range using deep neural networks. In our case, PPO was constructed
such that it could predict an entire set of readout pulses in a single neural network
pass, saving valuable training time.
For single-qubit readout, DQN generated pulses that reduced the required pulse
length by 60.8%, while PPO generated pulse shapes reduced the required pulse length
required by 63.1%. In general, we do see that the CLEAR pulse shapes and the PPO
generated pulse shapes are very similar in the single-qubit environment. This isn’t
too surprising as the CLEAR pulse should be optimal in an environment free of noise
and cross-talk from other pulses and qubit-resonator pairs. Additionally, we found
that DQN training proved to be both slower than PPO and more sample inefficient.
Paired with its reduced final performance, we did not employ DQN for the multi-
qubit simulation environment. In that multi-qubit environment, PPO reduced the
required pulse length by 57.6% compared to the rectangular pulse. Compared to the
optimized CLEAR pulse, the PPO generated readout pulse provided an additional
21.0% pulse length reduction.
Finally, following the success of our readout pulse shaping optimization in a sim-
ulated multi-qubit readout environment, we discussed the experimental setup and
designs that will be used to perform the real-world experiments by the group in the
future.
Altogether, the successful results in the simulated environment show that state-of-
the-art machine learning methods can be used for optimal control in superconducting
quantum environments. Additionally, this work provides a framework for using deep

96
reinforcement learning, or any other to be discovered optimization technique, in a
variety of simulated and experimental settings.

Pulse Shaping Optimization Method


Rectangular + Decay CLEAR PPO DQN
Single-Qubit Reset 700 ns 260 ns 250 ns 250 ns
Single-Qubit Injection 1470 ns 550 ns 550 ns 600 ns
Single-Qubit Complete Pulse 2170 ns 810 ns 800 ns 850 ns
Multi-Qubit Reset 700 ns 460 ns 380 ns -
Multi-Qubit Injection 1610 ns 780 ns 600 ns -
Multi-Qubit Complete Pulse 2310 ns 1240 ns 980 ns -

Table 9.1: Final readout pulse lengths for the various conventional and optimization
methods in this research. The single-qubit environment includes optimized CLEAR
pulses and pulses generated with PPO and DQN. In the multi-qubit environment, we
only compare optimized CLEAR pulses and pulses generated with PPO.

9.2 Future Work Possibilities

9.2.1 Additional Applications

The first broad set of possibilities for future work is using the software package to
generate optimal pulse shapes for different use cases, both in simulated environments
and experimental setups. Outside of multiplexed readout, a couple of areas of using
deep reinforcement learning have been explored in simulated environments already.
The most substantial work involves using AlphaZero to generate optimal single-qubit
gate pulse shapes [11]. However the AlphaZero approach in that research suffered from
similar issues that we found in non-PPO algorithms, namely that they were too slow
training-wise and that existing optimization methods could sometimes outperform
them. Based on what we found in this thesis, it would seem that a deep policy gradient
algorithm like PPO could be more successful in a context like single- and two-qubit
gate optimal pulse shaping. Importantly, our work provides a path toward being able

97
to implement these DRL optimal control approaches in experimental settings. To the
best of our knowledge, this would be the first experimental implementation of using
DRL in a quantum optimal control setting.
It’s important to note that software created for this research is not restricted
to only working with the readout simulation environment we describe. Instead, it
is a modular system designed to allow users to swap in different optimization tech-
niques (CLEAR, PPO, DQN, etc.), as well as different simulation parameter values
and entirely different environments. For those different simulation environments, the
software package only requires a user to accept pulse shapes in an array-like format
and then output a scalar reward signal. All simulation logic to get from a pulse shape
to that reward signal is entirely up to an end-user to implement, making the software
package extremely extendable. The same holds true for using the software system
in experimental environments. Instead of implementing simulation software, one can
send the generated pulse shapes to an instrument interface like Labber and implement
a method to convert measurement data to a scalar reward, all while not needing to
change any of the optimization software.

9.2.2 Pulse Length Optimization

One of the largest opportunities for extending the current deep reinforcement learning
implementation is finding a way to include optimizing the length of the generated
pulse in addition to the actual pulse shape. Currently, we have to set a specific,
locked-in pulse length, as well as the length of each pulse segment, at the start of
each new optimization cycle. If that optimization proves to be successful, we reduce
the pulse length and keep this cycle going until we find a pulse length that can’t
be shortened any further. Obviously, having to run multiple optimization cycles is
a bottleneck especially for deep learning applications where training can take long
amounts of time.
The goal of future work from a reinforcement perspective would be to include
finding the optimal pulse length in addition to the optimal pulse amplitudes. To do
this, we could create set pulse segment lengths, and look to include a “stop” segment

98
as part of the available action set in our reinforcement learning environment. One
could even have a setup where individual pulses could be stopped at different times,
eliminating strange individual pulses associated with resonators that have shorter
photon decay times (see Fig. 6-3(c) and resonator 5 in Fig. 7-5). A reward signal
would need to be carefully designed to incorporate both minimizing the pulse length
and reaching the desired resonator state. While this would most likely increase the
overall training time needed to perform a single optimization cycle, it would eliminate
the need to run multiple optimization cycles and thus should reduce the overall time
needed to find optimal readout pulses.

9.2.3 Continuous Experimental Optimization

Another opportunity for practical use of deep reinforcement learning in supercon-


ducting quantum computing systems is with recalibration efforts. It is common for
systems to need frequent recalibration, whether it be changing 𝑇1 times, drift in the
control fields used to drive logic gates or various sources of noise in the system or from
the instruments controlling the system [56, 57, 58]. This means that if we use PPO
to optimize an ideal set of readout pulses, those pulses may become less effective over
time.
PPO provides an intriguing possibility to try and compensate for those systems
changes in real-time. Once an optimal readout pulse is generated, one could theo-
retically still allow PPO to continuously optimize an ideal readout pulse shape. As
discussed in Chapter 5 Section 3.2, when PPO is updating its neural network weights,
it generates a set of policies (pulse shapes) from the existing policy distribution, then
trains off the best performing new policies. If we were to perform this step when-
ever our pulse performance drops below a certain threshold, we could potentially
“follow” the changes in the system as they are occurring, rather than shutting down
an experiment and recalibrating from scratch.

99
100
Bibliography

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


Oliver, “A quantum engineer’s guide to superconducting qubits”, Appl. Phys.
Rev., 6(2), 2019.

[2] R. Babbush et al. F. Arute, K. Arya, “Quantum supremacy using a pro-


grammable superconducting processor”, Nature, 574(7779):505–510, 2019.

[3] J. Preskill, “Quantum Computing in the NISQ era and beyond”, Quantum, 2:79,
2018.

[4] Alexandre Blais, Ren-Shou Huang, Andreas Wallraff, S. M. Girvin, and R. J.


Schoelkopf, “Cavity quantum electrodynamics for superconducting electrical cir-
cuits: An architecture for quantum computation”, Phys. Rev. A, 69:062320,
2004.

[5] Austin G. Fowler, Matteo Mariantoni, John M. Martinis, and Andrew N. Cle-
land, “Surface codes: Towards practical large-scale quantum computation”, Phys.
Rev. A, 86:032324, 2012.

[6] D.T. McClure, Hanhee Paik, L.S. Bishop, M. Steffen, Jerry M. Chow, and Jay M.
Gambetta, “Rapid driven reset of a qubit readout resonator”, Physical Review
Applied, 5(1), 2016.

[7] Richard S. Sutton and Andrew G. Barto, Reinforcement Learning: An Introduc-


tion, A Bradford Book, Cambridge, MA, USA, 2018.

[8] David Silver, Aja Huang, Christopher J. Maddison, Arthur Guez, Laurent Sifre,
George van den Driessche, Julian Schrittwieser, Ioannis Antonoglou, Veda Pan-
neershelvam, Marc Lanctot, Sander Dieleman, Dominik Grewe, John Nham,
Nal Kalchbrenner, Ilya Sutskever, Timothy Lillicrap, Madeleine Leach, Koray
Kavukcuoglu, Thore Graepel, and Demis Hassabis, “Mastering the game of go
with deep neural networks and tree search”, Nature, 529:484–503, 2016.

[9] David Silver, Thomas Hubert, Julian Schrittwieser, Ioannis Antonoglou,


Matthew Lai, Arthur Guez, Marc Lanctot, Laurent Sifre, Dharshan Kumaran,
Thore Graepel, Timothy P. Lillicrap, Karen Simonyan, and Demis Hassabis,
“Mastering chess and shogi by self-play with a general reinforcement learning
algorithm”, ArXiv, abs/1712.01815, 2017.

101
[10] S. Gu, E. Holly, T. Lillicrap, and S. Levine, “Deep reinforcement learning for
robotic manipulation with asynchronous off-policy updates”, In 2017 IEEE In-
ternational Conference on Robotics and Automation (ICRA), pages 3389–3396,
2017.

[11] J. J. Sorensen M. Dalgaard, F. Motzoi and J. Sherson, “Global optimization of


quantum dynamics with AlphaZero deep exploration”, npj Quantum Inf., 6:6,
2020.

[12] Peter W. Shor, “Polynomial-time algorithms for prime factorization and dis-
crete logarithms on a quantum computer”, SIAM Journal on Computing,
26(5):1484–1509, 1997.

[13] Richard P. Feynman, Simulating Physics with Computers, page 133–153, Perseus
Books, USA, 1999.

[14] I. M. Georgescu, S. Ashhab, and Franco Nori, “Quantum simulation”, Rev. Mod.
Phys., 86:153–185, 2014.

[15] Michael Tinkham, Introduction to Superconductivity, Dover Publications, 2


edition, 2004.

[16] V. Bouchiat, D. Vion, P. Joyez, D. Esteve, and M. H. Devoret, “Quantum


Coherence with a Single Cooper Pair”, Physica Scripta Volume T, 76:165–170,
1998.

[17] Yuriy Makhlin, Gerd Schön, and Alexander Shnirman, “Quantum-state engineer-
ing with josephson-junction devices”, Reviews of Modern Physics, 73(2):357–400,
2001.

[18] G. Ithier, E. Collin, P. Joyez, P. J. Meeson, D. Vion, D. Esteve, F. Chiarello,


A. Shnirman, Y. Makhlin, J. Schriefl, and et al., “Decoherence in a supercon-
ducting quantum bit circuit”, Physical Review B, 72(13), 2005.

[19] D. Vion, “Manipulating the quantum state of an electrical circuit”, Science,


296(5569):886–889, 2002.

[20] Jens Koch, Terri M. Yu, Jay Gambetta, A. A. Houck, D. I. Schuster, J. Majer,
Alexandre Blais, M. H. Devoret, S. M. Girvin, and R. J. Schoelkopf, “Charge-
insensitive qubit design derived from the cooper pair box”, Phys. Rev. A,
76:042319, 2007.

[21] M.D. Hutchings, J.B. Hertzberg, Y. Liu, N.T. Bronn, G.A. Keefe, Markus Brink,
Jerry M. Chow, and B.L.T. Plourde, “Tunable superconducting qubits with flux-
independent coherence”, Physical Review Applied, 8(4), 2017.

[22] B. Lienhard, A. Vepsäläinen, L. C. G. Govia, C. R. Hoffer, J. Qiu, D. Riste,


M. Ware, D. Kim, R. Winik, A. Melville, B. Niedzielski, J. Yoder, G. J. Ribeill,
T. A. Ohki, H. K. Krovi, T. P. Orlando, S. Gustavsson, and W. D. Oliver, “Deep

102
neural network discrimination of multiplexed superconducting qubit states”, Un-
published, 2021.

[23] Alexandre Blais, Ren Shou Huang, Andreas Wallraff, S. M. Girvin, and R. J.
Schoelkopf, “Cavity quantum electrodynamics for superconducting electrical cir-
cuits: An architecture for quantum computation”, Phys. Rev. A - At. Mol. Opt.
Phys., 69(6):1–14, 2004.

[24] A. Wallraff, D. I. Schuster, A. Blais, L. Frunzio, R.-S. Huang, J. Majer, S. Ku-


mar, S. M. Girvin, and R. J. Schoelkopf, “Strong coupling of a single photon
to a superconducting qubit using circuit quantum electrodynamics”, Nature,
431(7005):162–167, 2004.

[25] Alexandre Blais, Ren-Shou Huang, Andreas Wallraff, S. M. Girvin, and R. J.


Schoelkopf, “Cavity quantum electrodynamics for superconducting electrical cir-
cuits: An architecture for quantum computation”, Physical Review A, 69(6),
2004.

[26] J Dalibard, J.-M Raimond, and Jean Zinn-Justin, Systemes fondamentaux en


optique quantique = Fundamental systems in quantum optics / edite par J. Dal-
ibard, J.-M. Raimond et J. Zinn-Justin., North-Holland, Amsterdam, 1992.

[27] J.R. Johansson, P.D. Nation, and Franco Nori, “Qutip: An open-source python
framework for the dynamics of open quantum systems”, Computer Physics Com-
munications, 183(8):1760–1772, 2012.

[28] F. Bloch, “Nuclear induction”, Phys. Rev., 70:460–474, 1946.

[29] Tom B. Brown, Benjamin Mann, Nick Ryder, Melanie Subbiah, Jared Kaplan,
Prafulla Dhariwal, Arvind Neelakantan, Pranav Shyam, Girish Sastry, Amanda
Askell, Sandhini Agarwal, Ariel Herbert-Voss, Gretchen Krueger, Tom Henighan,
Rewon Child, Aditya Ramesh, Daniel M. Ziegler, Jeffrey Wu, Clemens Winter,
Christopher Hesse, Mark Chen, Eric Sigler, Mateusz Litwin, Scott Gray, Ben-
jamin Chess, Jack Clark, Christopher Berner, Sam McCandlish, Alec Radford,
Ilya Sutskever, and Dario Amodei, “Language models are few-shot learners”,
2020.

[30] Richard Bellman, “The theory of dynamic programming”, Bull. Amer. Math.
Soc., 60(6):503–515, 1954.

[31] Warren S. McCulloch and Walter Pitts, A Logical Calculus of the Ideas Immanent
in Nervous Activity, page 15–27, MIT Press, Cambridge, MA, USA, 1988.

[32] Sergey Ioffe and Christian Szegedy, “Batch normalization: Accelerating deep
network training by reducing internal covariate shift”, 2015.

[33] David E. Rumelhart and David Zipser, Feature Discovery by Competitive Learn-
ing, page 205–242, Ablex Publishing Corp., USA, 1988.

103
[34] D. T. McClure, Hanhee Paik, L. S. Bishop, M. Steffen, Jerry M. Chow, and
Jay M. Gambetta, “Rapid driven reset of a qubit readout resonator”, Phys. Rev.
Applied, 5:011001, 2016.
[35] Johannes Heinsoo, Christian Kraglund Andersen, Ants Remm, Sebastian Krin-
ner, Theodore Walter, Yves Salathé, Simone Gasparinetti, Jean Claude Besse,
Anton Potočnik, Andreas Wallraff, and Christopher Eichler, “Rapid High-fidelity
Multiplexed Readout of Superconducting Qubits”, Phys. Rev. Appl., 10(3):1–14,
2018.
[36] N T Bronn, B Abdo, K Inoue, S Lekuch, A D Córcoles, J B Hertzberg, M Takita,
L S Bishop, J M Gambetta, and J M Chow, “Fast, high-fidelity readout of
multiple qubits”, Journal of Physics: Conference Series, 834:012003, 2017.
[37] C. Macklin, K. O’Brien, D. Hover, M. E. Schwartz, V. Bolkhovsky, X. Zhang,
W. D. Oliver, and I. Siddiqi, “A near–quantum-limited Josephson traveling-wave
parametric amplifier”, Science, 350(6258):307–310, 2015.
[38] Niranjan Srinivas, Andreas Krause, Sham M. Kakade, and Matthias W. Seeger,
“Gaussian process bandits without regret: An experimental design approach”,
CoRR, abs/0912.3995, 2009.
[39] Volodymyr Mnih, Koray Kavukcuoglu, David Silver, Andrei A. Rusu, Joel Ve-
ness, Marc G. Bellemare, Alex Graves, Martin Riedmiller, Andreas K. Fid-
jeland, Georg Ostrovski, Stig Petersen, Charles Beattie, Amir Sadik, Ioannis
Antonoglou, Helen King, Dharshan Kumaran, Daan Wierstra, Shane Legg, and
Demis Hassabis, “Human-level control through deep reinforcement learning”,
Nature, 518(7540):529–533, 2015.
[40] Christopher J. C. H. Watkins and Peter Dayan, “Q-learning”, Machine Learning,
8(3):279–292, 1992.
[41] Navin Khaneja, Timo Reiss, Cindie Kehlet, Thomas Schulte-Herbrüggen, and
Steffen 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, 2005.
[42] Tom Schaul, John Quan, Ioannis Antonoglou, and David Silver, “Prioritized
experience replay”, CoRR, abs/1511.05952, 2015.
[43] John Schulman, Filip Wolski, Prafulla Dhariwal, Alec Radford, and Oleg Klimov,
“Proximal policy optimization algorithms”, 2017.
[44] John Schulman, Sergey Levine, Philipp Moritz, Michael I. Jordan, and Pieter
Abbeel, “Trust region policy optimization”, 2017.
[45] Timothy P. Lillicrap, Jonathan J. Hunt, Alexander Pritzel, Nicolas Manfred Otto
Heess, Tom Erez, Yuval Tassa, David Silver, and Daan Wierstra, “Continuous
control with deep reinforcement learning”, CoRR, abs/1509.02971, 2015.

104
[46] F. Motzoi, J. M. Gambetta, S. T. Merkel, and F. K. Wilhelm, “Optimal control
methods for rapidly time-varying hamiltonians”, Physical Review A, 84(2), 2011.

[47] Samuel Boutin, Christian Kraglund Andersen, Jayameenakshi Venkatraman, An-


drew J. Ferris, and Alexandre Blais, “Resonator reset in circuit QED by optimal
control for large open quantum systems”, Phys. Rev. A - At. Mol. Opt. Phys.,
96:1–11, 2017.

[48] Yanbing Liu, Srikanth J. Srinivasan, D. Hover, Shaojiang Zhu, R. McDermott,


and A. A. Houck, “High fidelity readout of a transmon qubit using a super-
conducting low-inductance undulatory galvanometer microwave amplifier”, New
Journal of Physics, 16(11):113008, 2014.

[49] Evan Jeffrey, Daniel Sank, J. Y. Mutus, T. C. White, J. Kelly, R. Barends,


Y. Chen, Z. Chen, B. Chiaro, A. Dunsworth, A. Megrant, P. J. J. O’Malley,
C. Neill, P. Roushan, A. Vainsencher, J. Wenner, A. N. Cleland, and John M.
Martinis, “Fast accurate state measurement with superconducting qubits”, Phys.
Rev. Lett., 112:190504, 2014.

[50] Laurens van der Maaten and Geoffrey Hinton, “Visualizing data using t-sne”,
Journal of Machine Learning Research, 9(86):2579–2605, 2008.

[51] Labber, “Labber is a general-purpose instrument control and lab au-


tomation software package, with particular focus on quantum appli-
cations.”, https://www.keysight.com/us/en/products/software/application-
sw/labber-software.html.

[52] Norman F. Ramsey, “A molecular beam resonance method with separated oscil-
lating fields”, Phys. Rev., 78:695–699, 1950.

[53] E. L. Hahn, “Spin echoes”, Phys. Rev., 80:580–594, 1950.

[54] Colm A. Ryan, Blake R. Johnson, Jay M. Gambetta, Jerry M. Chow, Marcus P.
Da Silva, Oliver E. Dial, and Thomas A. Ohki, “Tomography via correlation
of noisy measurement records”, Phys. Rev. A - At. Mol. Opt. Phys., 91(2):1–7,
2015.

[55] Easwar Magesan, Jay M. Gambetta, A.D. Còrcoles, and Jerry M. Chow, “Ma-
chine Learning for Discriminating Quantum Measurement Trajectories and Im-
proving Readout”, Phys. Rev. Lett., 114:200501, 2015.

[56] P. V. Klimov, J. Kelly, Z. Chen, M. Neeley, A. Megrant, B. Burkett, R. Barends,


K. Arya, B. Chiaro, Yu Chen, A. Dunsworth, A. Fowler, B. Foxen, C. Gidney,
M. Giustina, R. Graff, T. Huang, E. Jeffrey, Erik Lucero, J. Y. Mutus, O. Naa-
man, C. Neill, C. Quintana, P. Roushan, Daniel Sank, A. Vainsencher, J. Wenner,
T. C. White, S. Boixo, R. Babbush, V. N. Smelyanskiy, H. Neven, and John M.
Martinis, “Fluctuations of energy-relaxation times in superconducting qubits”,
Phys. Rev. Lett., 121:090502, 2018.

105
[57] M. A. Fogarty, M. Veldhorst, R. Harper, C. H. Yang, S. D. Bartlett, S. T. Flam-
mia, and A. S. Dzurak, “Nonexponential fidelity decay in randomized bench-
marking with low-frequency noise”, Physical Review A, 92(2), 2015.

[58] Jonathan J. Burnett, Andreas Bengtsson, Marco Scigliuzzo, David Niepce, Ma-
rina Kudra, Per Delsing, and Jonas Bylander, “Decoherence benchmarking of
superconducting qubits”, npj Quantum Information, 5(1), 2019.

106

You might also like

pFad - Phonifier reborn

Pfad - The Proxy pFad of © 2024 Garber Painting. All rights reserved.

Note: This service is not intended for secure transactions such as banking, social media, email, or purchasing. Use at your own risk. We assume no liability whatsoever for broken pages.


Alternative Proxies:

Alternative Proxy

pFad Proxy

pFad v3 Proxy

pFad v4 Proxy