Skip to content

Latest commit

 

History

History
848 lines (575 loc) · 29.2 KB

intro.rst

File metadata and controls

848 lines (575 loc) · 29.2 KB

Introduction to Quantum Computing

With every breakthrough in science there is the potential for new technology. For over twenty years, researchers have done inspiring work in quantum mechanics, transforming it from a theory for understanding nature into a fundamentally new way to engineer computing technology. This field, quantum computing, is beautifully interdisciplinary, and impactful in two major ways:

  1. It reorients the relationship between physics and computer science. Physics does not just place restrictions on what computers we can design, it also grants new power and inspiration.
  2. It can simulate nature at its most fundamental level, allowing us to solve deep problems in quantum chemistry, materials discovery, and more.

Quantum computing has come a long way, and in the next few years there will be significant breakthroughs in the field. To get here, however, we have needed to change our intuition for computation in many ways. As with other paradigms --- such as object-oriented programming, functional programming, distributed programming, or any of the other marvelous ways of thinking that have been expressed in code over the years --- even the basic tenants of quantum computing opens up vast new potential for computation.

However, unlike other paradigms, quantum computing goes further. It requires an extension of classical probability theory. This extension, and the core of quantum computing, can be formulated in terms of linear algebra. Therefore, we begin our investigation into quantum computing with linear algebra and probability.

From Bit to Qubit

Probabilistic Bits as Vector Spaces

From an operational perspective, a bit is described by the results of measurements performed on it. Let the possible results of measuring a bit (0 or 1) be represented by orthonormal basis vectors \(\vec{0}\) and \(\vec{1}\). We will call these vectors outcomes. These outcomes span a two-dimensional vector space that represents a probabilistic bit. A probabilistic bit can be represented as a vector

$$\vec{v} = a\,\vec{0} + b\,\vec{1},$$

where \(a\) represents the probability of the bit being 0 and \(b\) represents the probability of the bit being 1. This clearly also requires that \(a+b=1\). In this picture the system (the probabilistic bit) is a two-dimensional real vector space and a state of a system is a particular vector in that vector space.

import numpy as np
import matplotlib.pyplot as plt


outcome_0 = np.array([1.0, 0.0])
outcome_1 = np.array([0.0, 1.0])
a = 0.75
b = 0.25

prob_bit = a * outcome_0 + b * outcome_1

X, Y = prob_bit
plt.figure()
ax = plt.gca()
ax.quiver(X, Y, angles='xy', scale_units='xy', scale=1)
ax.set_xlim([0, 1])
ax.set_ylim([0, 1])
plt.draw()
plt.show()

image

Given some state vector, like the one plotted above, we can find the probabilities associated with each outcome by projecting the vector onto the basis outcomes. This gives us the following rule:

$$\begin{aligned} \operatorname{Pr}(0) = \vec{v}^T.\vec{0} = a \\ \operatorname{Pr}(1) = \vec{v}^T.\vec{1} = b, \end{aligned}$$

where Pr(0) and Pr(1) are the probabilities of the 0 and 1 outcomes respectively.

Dirac Notation

Physicists have introduced a convenient notation for the vector transposes and dot products we used in the previous example. This notation, called Dirac notation in honor of the great theoretical physicist Paul Dirac, allows us to define

$$\begin{aligned} \vec{v} = \vert v\rangle \\ \vec{v}^T = \langle v \vert \\ \vec{u}^T \cdot \vec{v} = \langle u \vert v \rangle \end{aligned}$$

Thus, we can rewrite our "measurement rule" in this notation as

$$\begin{aligned} Pr(0) = \langle v \vert 0 \rangle = a \\ Pr(1) = \langle v\vert 1 \rangle = b \end{aligned}$$

We will use this notation throughout the rest of this introduction.

Multiple Probabilistic Bits

This vector space interpretation of a single probabilistic bit can be straightforwardly extended to multiple bits. Let us take two coins as an example (labelled 0 and 1 instead of H and T since we are programmers). Their states can be represented as

$$\begin{aligned} |\,u\rangle = \frac{1}{2}|\,0_u\rangle + \frac{1}{2}|\,1_u\rangle \\\ \end{aligned}$$

,0_vrangle + frac{1}{2}|,1_vrangle,

where \(1_u\) represents the outcome 1 on coin \(u\). The combined system of the two coins has four possible outcomes \(\{ 0_u0_v,\;0_u1_v,\;1_u0_v,\;1_u1_v \}\) that are the basis states of a larger four-dimensional vector space. The rule for constructing a combined state is to take the tensor product of individual states, e.g.

$$|\,u\rangle\otimes|\,v\rangle = \frac{1}{4}|\,0_u0_v\rangle+\frac{1}{4}|\,0_u1_v\rangle+\frac{1}{4}|\,1_u0_v\rangle+\frac{1}{4}|\,1_u1_v\rangle.$$

Then, the combined space is simply the space spanned by the tensor products of all pairs of basis vectors of the two smaller spaces.

Similarly, the combined state for \(n\) such probabilistic bits is a vector of size \(2^n\) and is given by \(\bigotimes_{i=0}^{n-1}|\,v_i\rangle\). We will talk more about these larger spaces in the quantum case, but it is important to note that not all composite states can be written as tensor products of sub-states (e.g. consider the state \(\frac{1}{2}\,1_u1_v\rangle\)). The most general composite state of \(n\) probabilistic bits can be written as \(\sum_{j=0}^{2^n - 1} a_{j} (\bigotimes_{i=0}^{n-1}|\,b_{ij}\rangle\)) where each \(b_{ij} \in \{0, 1\}\) and \(a_j \in \mathbb{R}\), i.e. as a linear combination (with real coefficients) of tensor products of basis states. Note that this still gives us \(2^n\) possible states.

Qubits

Quantum mechanics rewrites these rules to some extent. A quantum bit, called a qubit, is the quantum analog of a bit in that it has two outcomes when it is measured. Similar to the previous section, a qubit can also be represented in a vector space, but with complex coefficients instead of real ones. A qubit system is a two-dimensional complex vector space, and the state of a qubit is a complex vector in that space. Again we will define a basis of outcomes \(\{|\,0\rangle, |\,1\rangle\}\) and let a generic qubit state be written as


α| 0⟩ + β| 1⟩.

Since these coefficients can be imaginary, they cannot be simply interpreted as probabilities of their associated outcomes. Instead we rewrite the rule for outcomes in the following manner:

$$\begin{aligned} \operatorname{Pr}(0) = |\langle v\,|\,0 \rangle|^2 = |\alpha|^2 \\ \operatorname{Pr}(1) = |\langle v\,|\,1 \rangle|^2 = |\beta|^2, \end{aligned}$$

and as long as \(|\alpha|^2 + |\beta|^2 = 1\) we are able to recover acceptable probabilities for outcomes based on our new complex vector.

This switch to complex vectors means that rather than representing a state vector in a plane, we instead represent the vector on a sphere (called the Bloch sphere in quantum mechanics literature). From this perspective the quantum state corresponding to an outcome of 0 is represented by:

image

Notice that the two axes in the horizontal plane have been labeled \(x\) and \(y\), implying that \(z\) is the vertical axis (not labeled). Physicists use the convention that a qubit's \(\{|\,0\rangle, |\,1\rangle\}\) states are the positive and negative unit vectors along the z axis, respectively. These axes will be useful later in this document.

Multiple qubits are represented in precisely the same way, by taking linear combinations (with complex coefficients, now) of tensor products of basis states. Thus \(n\) qubits have \(2^n\) possible states.

An Important Distinction

An important distinction between the probabilistic case described above and the quantum case is that probabilistic states may just mask out ignorance. For example a coin is physically only 0 or 1 and the probabilistic view merely represents our ignorance about which it actually is. This is not the case in quantum mechanics. Assuming events occuring at a distance from one another cannot instantaneously influence each other, the quantum states ---as far as we know --- cannot mask any underlying state. This is what people mean when they say that there is no local hidden variable theory for quantum mechanics. These probabilistic quantum states are as real as it gets: they don't just describe our knowledge of the quantum system, they describe the physical reality of the system.

Some Code

Let us take a look at some code in pyQuil to see how these quantum states play out. We will dive deeper into quantum operations and pyQuil in the following sections. Note that in order to run these examples you will need to install pyQuil and download the QVM and Compiler. Each of the code snippets below will be immediately followed by its output.

# Imports for pyQuil (ignore for now)
import numpy as np
from pyquil.quil import Program
from pyquil.api import WavefunctionSimulator


# create a WavefunctionSimulator object
wavefunction_simulator = WavefunctionSimulator()

# pyQuil is based around operations (or gates) so we will start with the most
# basic one: the identity operation, called I. I takes one argument, the index
# of the qubit that it should be applied to.
from pyquil.gates import I

# Make a quantum program that allocates one qubit (qubit #0) and does nothing to it
p = Program(I(0))

# Quantum states are called wavefunctions for historical reasons.
# We can run this basic program on our connection to the simulator.
# This call will return the state of our qubits after we run program p.
# This api call returns a tuple, but we'll ignore the second value for now.
wavefunction = wavefunction_simulator.wavefunction(p)

# wavefunction is a Wavefunction object that stores a quantum state as a list of amplitudes
alpha, beta = wavefunction

print("Our qubit is in the state alpha={} and beta={}".format(alpha, beta))
print("The probability of measuring the qubit in outcome 0 is {}".format(abs(alpha)**2))
print("The probability of measuring the qubit in outcome 1 is {}".format(abs(beta)**2))

Our qubit is in the state alpha=(1+0j) and beta=0j The probability of measuring the qubit in outcome 0 is 1.0 The probability of measuring the qubit in outcome 1 is 0.0

Applying an operation to our qubit affects the probability of each outcome.

# We can import the qubit "flip" operation, called X, and see what it does.
# We will learn more about this operation in the next section.
from pyquil.gates import X

p = Program(X(0))

wavefunc = wavefunction_simulator.wavefunction(p)
alpha, beta = wavefunc

print("Our qubit is in the state alpha={} and beta={}".format(alpha, beta))
print("The probability of measuring the qubit in outcome 0 is {}".format(abs(alpha)**2))
print("The probability of measuring the qubit in outcome 1 is {}".format(abs(beta)**2))

Our qubit is in the state alpha=0j and beta=(1+0j) The probability of measuring the qubit in outcome 0 is 0.0 The probability of measuring the qubit in outcome 1 is 1.0

In this case we have flipped the probability of outcome 0 into the probability of outcome 1 for our qubit. We can also investigate what happens to the state of multiple qubits. We'd expect the state of multiple qubits to grow exponentially in size, as their vectors are tensored together.

# Multiple qubits also produce the expected scaling of the state.
p = Program(I(0), I(1))
wavefunction = wavefunction_simulator.wavefunction(p)
print("The quantum state is of dimension:", len(wavefunction.amplitudes))

p = Program(I(0), I(1), I(2), I(3))
wavefunction = wavefunction_simulator.wavefunction(p)
print("The quantum state is of dimension:", len(wavefunction.amplitudes))

p = Program()
for x in range(10):
    p += I(x)
wavefunction = wavefunction_simulator.wavefunction(p)
print("The quantum state is of dimension:", len(wavefunction.amplitudes))

The quantum state is of dimension: 4 The quantum state is of dimension: 16 The quantum state is of dimension: 1024

Let's look at the actual value for the state of two qubits combined. The resulting dictionary of this method contains outcomes as keys and the probabilities of those outcomes as values.

# wavefunction(Program) returns a coefficient array that corresponds to outcomes in the following order
wavefunction = wavefunction_simulator.wavefunction(Program(I(0), I(1)))
print(wavefunction.get_outcome_probs())

{'00': 1.0, '01': 0.0, '10': 0.0, '11': 0.0}

Qubit Operations

In the previous section we introduced our first two operations: the I (or Identity) operation and the X (or NOT) operation. In this section we will get into some more details on what these operations are.

Quantum states are complex vectors on the Bloch sphere, and quantum operations are matrices with two properties:

  1. They are reversible.
  2. When applied to a state vector on the Bloch sphere, the resulting vector is also on the Bloch sphere.

Matrices that satisfy these two properties are called unitary matrices. Such matrices have the characteristic property that their complex conjugate transpose is equal to their inverse, a property directly linked to the requirement that the probabilities of measuring qubits in any of the allowed states must sum to 1. Applying an operation to a quantum state is the same as multiplying a vector by one of these matrices. Such an operation is called a gate.

Since individual qubits are two-dimensional vectors, operations on individual qubits are 2x2 matrices. The identity matrix leaves the state vector unchanged:

$$\begin{aligned} I = \left(\begin{matrix} 1 & 0\\\ 0 & 1 \end{matrix}\right) \end{aligned}$$

so the program that applies this operation to the zero state is just

$$I\,|\,0\rangle = \left(\begin{matrix}$$

1 & 0\ 0 & 1 end{matrix}right)left(begin{matrix} 1 \ 0 end{matrix}right) = left(begin{matrix} 1 \ 0 end{matrix}right) = |,0rangle

p = Program(I(0))
print(wavefunction_simulator.wavefunction(p))

(1+0j)|0>

Pauli Operators

Let's revisit the X gate introduced above. It is one of three important single-qubit gates, called the Pauli operators:

$$\begin{aligned} X = \left(\begin{matrix} 0 & 1\\\ 1 & 0 \end{matrix}\right) \qquad Y = \left(\begin{matrix} 0 & -i\\\ i & 0 \end{matrix}\right) \qquad Z = \left(\begin{matrix} 1 & 0\\\ 0 & -1 \end{matrix}\right) \end{aligned}$$

from pyquil.gates import X, Y, Z


p = Program(X(0))
wavefunction = wavefunction_simulator.wavefunction(p)
print("X|0> = ", wavefunction)
print("The outcome probabilities are", wavefunction.get_outcome_probs())
print("This looks like a bit flip.\n")

p = Program(Y(0))
wavefunction = wavefunction_simulator.wavefunction(p)
print("Y|0> = ", wavefunction)
print("The outcome probabilities are", wavefunction.get_outcome_probs())
print("This also looks like a bit flip.\n")

p = Program(Z(0))
wavefunction = wavefunction_simulator.wavefunction(p)
print("Z|0> = ", wavefunction)
print("The outcome probabilities are", wavefunction.get_outcome_probs())
print("This state looks unchanged.")

X1> The outcome probabilities are {'0': 0.0, '1': 1.0} This looks like a bit flip.

Y1> The outcome probabilities are {'0': 0.0, '1': 1.0} This also looks like a bit flip.

Z0> The outcome probabilities are {'0': 1.0, '1': 0.0} This state looks unchanged.

The Pauli matrices have a visual interpretation: they perform 180-degree rotations of qubit state vectors on the Bloch sphere. They operate about their respective axes as shown in the Bloch sphere depicted above. For example, the X gate performs a 180-degree rotation about the \(x\) axis. This explains the results of our code above: for a state vector initially in the +\(z\) direction, both X and Y gates will rotate it to -\(z\), and the Z gate will leave it unchanged.

However, notice that while the X and Y gates produce the same outcome probabilities, they actually produce different states. These states are not distinguished if they are measured immediately, but they produce different results in larger programs.

Quantum programs are built by applying successive gate operations:

# Composing qubit operations is the same as multiplying matrices sequentially
p = Program(X(0), Y(0), Z(0))
wavefunction = wavefunction_simulator.wavefunction(p)

print("ZYX|0> = ", wavefunction)
print("With outcome probabilities\n", wavefunction.get_outcome_probs())

ZYX|0> = [ 0.-1.j 0.+0.j] With outcome probabilities {'0': 1.0, '1': 0.0}

Multi-Qubit Operations

Operations can also be applied to composite states of multiple qubits. One common example is the controlled-NOT or CNOT gate that works on two qubits. Its matrix form is:

$$\begin{aligned} CNOT = \left(\begin{matrix} 1 & 0 & 0 & 0 \\\ 0 & 1 & 0 & 0 \\\ 0 & 0 & 0 & 1 \\\ 0 & 0 & 1 & 0 \\\ \end{matrix}\right) \end{aligned}$$

Let's take a look at how we could use a CNOT gate in pyQuil.

from pyquil.gates import CNOT


p = Program(CNOT(0, 1))
wavefunction = wavefunction_simulator.wavefunction(p)
print("CNOT|00> = ", wavefunction)
print("With outcome probabilities\n", wavefunction.get_outcome_probs(), "\n")

p = Program(X(0), CNOT(0, 1))
wavefunction = wavefunction_simulator.wavefunction(p)
print("CNOT|01> = ", wavefunction)
print("With outcome probabilities\n", wavefunction.get_outcome_probs(), "\n")

p = Program(X(1), CNOT(0, 1))
wavefunction = wavefunction_simulator.wavefunction(p)
print("CNOT|10> = ", wavefunction)
print("With outcome probabilities\n", wavefunction.get_outcome_probs(), "\n")

p = Program(X(0), X(1), CNOT(0, 1))
wavefunction = wavefunction_simulator.wavefunction(p)
print("CNOT|11> = ", wavefunction)
print("With outcome probabilities\n", wavefunction.get_outcome_probs(), "\n")

CNOT00> With outcome probabilities {'00': 1.0, '01': 0.0, '10': 0.0, '11': 0.0}

CNOT11> With outcome probabilities {'00': 0.0, '01': 0.0, '10': 0.0, '11': 1.0}

CNOT10> With outcome probabilities {'00': 0.0, '01': 0.0, '10': 1.0, '11': 0.0}

CNOT01> With outcome probabilities {'00': 0.0, '01': 1.0, '10': 0.0, '11': 0.0}

The CNOT gate does what its name implies: the state of the second qubit is flipped (negated) if and only if the state of the first qubit is 1 (true).

Another two-qubit gate example is the SWAP gate, which swaps the \( |01\rangle \) and \(|10\rangle \) states:

$$\begin{aligned} SWAP = \left(\begin{matrix} 1 & 0 & 0 & 0 \\\ 0 & 0 & 1 & 0 \\\ 0 & 1 & 0 & 0 \\\ 0 & 0 & 0 & 1 \\\ \end{matrix}\right) \end{aligned}$$

from pyquil.gates import SWAP


p = Program(X(0), SWAP(0,1))
wavefunction = wavefunction_simulator.wavefunction(p)

print("SWAP|01> = ", wavefunction)
print("With outcome probabilities\n", wavefunction.get_outcome_probs())

SWAP10> With outcome probabilities {'00': 0.0, '01': 0.0, '10': 1.0, '11': 0.0}

In summary, quantum computing operations are composed of a series of complex matrices applied to complex vectors. These matrices must be unitary (meaning that their complex conjugate transpose is equal to their inverse) because the overall probability of all outcomes must always sum to one.

The Quantum Abstract Machine

We now have enough background to introduce the programming model that underlies Quil. This is a hybrid quantum-classical model in which \(N\) qubits interact with \(M\) classical bits:

image

These qubits and classical bits come with a defined gate set, e.g. which gate operations can be applied to which qubits. Different kinds of quantum computing hardware place different limitations on what gates can be applied, and the fixed gate set represents these limitations.

Full details on the Quantum Abstract Machine and Quil can be found in the Quil whitepaper.

The next section on measurements will describe the interaction between the classical and quantum parts of a Quantum Abstract Machine (QAM).

Qubit Measurements

Measurements have two effects:

  1. They project the state vector onto one of the basic outcomes
  2. (optional) They store the outcome of the measurement in a classical bit.

Here's a simple example:

# Create a program that stores the outcome of measuring qubit #0 into classical register [0]
p = Program()
classical_register = p.declare('ro', 'BIT', 1)
p += Program(I(0)).measure(0, classical_register[0])

Up until this point we have used the quantum simulator to cheat a little bit --- we have actually looked at the wavefunction that comes back. However, on real quantum hardware, we are unable to directly look at the wavefunction. Instead we only have access to the classical bits that are affected by measurements. This functionality is emulated by :pyQuantumComputer.run. Note that the run command is to be applied on the compiled version of the program.

from pyquil import get_qc


qc = get_qc('9q-square-qvm')
print (qc.run(qc.compile(p)))

[[0]]

We see that the classical register reports a value of zero. However, if we had flipped the qubit before measurement then we obtain:

p = Program()
classical_register = p.declare('ro', 'BIT', 1)
p += Program(X(0))   # Flip the qubit
p.measure(0, classical_register[0])   # Measure the qubit

print (qc.run(qc.compile(p)))

[[1]]

These measurements are deterministic, e.g. if we make them multiple times then we always get the same outcome:

p = Program()
classical_register = p.declare('ro', 'BIT', 1)
p += Program(X(0))   # Flip the qubit
p.measure(0, classical_register[0])   # Measure the qubit

trials = 10
p.wrap_in_numshots_loop(shots=trials)

print (qc.run(qc.compile(p)))

[[1], [1], [1], [1], [1], [1], [1], [1], [1], [1]]

Classical/Quantum Interaction

However this is not the case in general --- measurements can affect the quantum state as well. In fact, measurements act like projections onto the outcome basis states. To show how this works, we first introduce a new single-qubit gate, the Hadamard gate. The matrix form of the Hadamard gate is:

$$\begin{aligned} H = \frac{1}{\sqrt{2}}\left(\begin{matrix} 1 & 1\\\ 1 & -1 \end{matrix}\right) \end{aligned}$$

The following pyQuil code shows how we can use the Hadamard gate:

from pyquil.gates import H


# The Hadamard produces what is called a superposition state
coin_program = Program(H(0))
wavefunction = wavefunction_simulator.wavefunction(coin_program)

print("H|0> = ", wavefunction)
print("With outcome probabilities\n", wavefunction.get_outcome_probs())

H0> + (0.7071067812+0j)|1> With outcome probabilities {'0': 0.49999999999999989, '1': 0.49999999999999989}

A qubit in this state will be measured half of the time in the \( |0\rangle \) state, and half of the time in the \( |1\rangle \) state. In a sense, this qubit truly is a random variable representing a coin. In fact, there are many wavefunctions that will give this same operational outcome. There is a continuous family of states of the form

$$\frac{1}{\sqrt{2}}\left(|\,0\rangle + e^{i\theta}|\,1\rangle\right)$$

that represent the outcomes of an unbiased coin. Being able to work with all of these different new states is part of what gives quantum computing extra power over regular bits.

p = Program()
ro = p.declare('ro', 'BIT', 1)

p += Program(H(0)).measure(0, ro[0])

# Measure qubit #0 a number of times
p.wrap_in_numshots_loop(shots=10)

# We see probabilistic results of about half 1's and half 0's
print (qc.run(qc.compile(p)))

[[0], [1], [1], [0], [1], [0], [0], [1], [0], [0]]

pyQuil allows us to look at the wavefunction after a measurement as well:

coin_program = Program(H(0))
print ("Before measurement: H|0> = ", wavefunction_simulator.wavefunction(coin_program), "\n")

ro = coin_program.declare('ro', 'BIT', 1)
coin_program.measure(0, ro[0])
for _ in range(5):
    print ("After measurement: ", wavefunction_simulator.wavefunction(coin_program))

Before measurement: H0> + (0.7071067812+0j)|1>

After measurement: (1+0j)1> After measurement: (1+0j)1> After measurement: (1+0j)|1>

We can clearly see that measurement has an effect on the quantum state independent of what is stored classically. We begin in a state that has a 50-50 probability of being \( |0\rangle \) or \( |1\rangle \). After measurement, the state changes into being entirely in \( |0\rangle \) or entirely in \( |1\rangle \) according to which outcome was obtained. This is the phenomenon referred to as the collapse of the wavefunction. Mathematically, the wavefunction is being projected onto the vector of the obtained outcome and subsequently rescaled to unit norm.

# This happens with bigger systems too, as can be seen with this program,
# which prepares something called a Bell state (a special kind of "entangled state")
bell_program = Program(H(0), CNOT(0, 1))
wavefunction = wavefunction_simulator.wavefunction(bell_program)
print("Before measurement: Bell state = ", wavefunction, "\n")

classical_regs = bell_program.declare('ro', 'BIT', 2)
bell_program.measure(0, classical_regs[0]).measure(1, classical_regs[1])

for _ in range(5):
    wavefunction = wavefunction_simulator.wavefunction(bell_program)
    print("After measurement: ", wavefunction.get_outcome_probs())

Before measurement: Bell state = (0.7071067812+0j)11>

After measurement: {'00': 0.0, '01': 0.0, '10': 0.0, '11': 1.0} After measurement: {'00': 0.0, '01': 0.0, '10': 0.0, '11': 1.0} After measurement: {'00': 0.0, '01': 0.0, '10': 0.0, '11': 1.0} After measurement: {'00': 0.0, '01': 0.0, '10': 0.0, '11': 1.0} After measurement: {'00': 0.0, '01': 0.0, '10': 0.0, '11': 1.0}

The above program prepares entanglement because, even though there are random outcomes, after every measurement both qubits are in the same state. They are either both \( |0\rangle \) or both \( |1\rangle \). This special kind of correlation is part of what makes quantum mechanics so unique and powerful.

Classical Control

There are also ways of introducing classical control of quantum programs. For example, we can use the state of classical bits to determine what quantum operations to run.

true_branch = Program(X(7)) # if branch
false_branch = Program(I(7)) # else branch

# Branch on ro[1]
p = Program()
ro = p.declare('ro', 'BIT', 8)
p += Program(X(0)).measure(0, ro[1]).if_then(ro[1], true_branch, false_branch)

# Measure qubit #7 into ro[7]
p.measure(7, ro[7])

# Run and check register [7]
print (qc.run(qc.compile(p)))

[[1 1]]

The second [1] here means that qubit 7 was indeed flipped.

image

Example: The Probabilistic Halting Problem

A fun example is to create a program that has an exponentially increasing chance of halting, but that may run forever!

p = Program()
ro = p.declare('ro', 'BIT', 1)
inside_loop = Program(H(0)).measure(0, ro[0])
p.inst(X(0)).while_do(ro[0], inside_loop)

qc = get_qc('9q-square-qvm')
print (qc.run(qc.compile(p)))

[[0]]

image

Next Steps

We hope that you have enjoyed your whirlwind tour of quantum computing. You are now ready to check out the Installation and Getting Started guide!

If you would like to learn more, Nielsen and Chuang's Quantum Computation and Quantum Information is a particularly excellent resource for newcomers to the field.

If you're interested in learning about the software behind quantum computing, take a look at our blog posts on The Quantum Software Challenge.