Learning Home Catalog Composer Lab
Learning
Home Catalog Composer Lab Return to course
Learning
Fundamentals of quantum algorithms
Quantum query algorithms Quantum algorithmic foundations Phase-estimation and factoring Grover's algorithm

Quantum query algorithms

Introduction

This course investigates computational advantages offered by quantum information. That is, we will think about what we can do with quantum computers, and the advantages they can offer over ordinary classical computers. To be specific, our focus will be on what we can do with a single quantum computer — as opposed to a distributed setting where multiple quantum computers interact over a network, for instance. (There are, in fact, quantum advantages to be found in distributed settings, where communication and cryptography come into play, but this topic is outside of the scope of this unit.)

We will begin with a natural question: What are the advantages that a quantum computer might potentially offer?

The first potential advantage, which is paramount among all others, is that quantum computers might provide faster solutions to some computational problems. Time is a truly precious resource — and it is this potential, that quantum computers may allow us to solve certain computational problems that classical computers are too slow to solve, that has driven quantum computing research for the past few decades.

There are other computational resources besides time that can be considered. The amount computer memory required to perform computations — usually referred to as the space required for computations — is one alternative that is often studied. As it turns out, however, quantum computers have a limited potential to offer advantages in space usage over classical computers. Computer memory is also relatively inexpensive and, unlike time, can be reused. For these reasons, time is of greater concern, and will be our primary focus.

One thing that quantum computers cannot do is to provide computational solutions to problems that classical computers cannot solve — irrespective of the resources required — such as the famous halting problem formulated by Alan Turing in the 1930s. Quantum computers can be simulated by classical computers, so any computational problem that can be solved by a quantum computer can also be solved by a classical computer, though it might take the classical computer much, much longer to find a solution.

While the time required to solve problems is our main concern, we will deviate slightly from this focus for the purposes of this first lesson. What we will do is to formulate a simple algorithmic framework — known as the query model — and explore the advantages that quantum computers offer within this framework.

The query model of computation is like a petri dish for quantum algorithmic ideas. It is rigid and unnatural, in the sense that it does not accurately represent the sorts of computational problems we generally care about in practice. Nevertheless, it has proved to be incredibly useful as a tool for developing quantum algorithmic techniques, including ones that power the most well-known quantum algorithms (such as Shor's factoring algorithm). It also happens to be a very useful framework for explaining these techniques.

After introducing the query model, we will discuss the very first quantum algorithm discovered, which is Deutsch's algorithm, along with an extension of Deutsch's algorithm known as the Deutsch-Jozsa algorithm. These algorithms demonstrate quantifiable advantages of quantum over classical computers, and in fact the Deutsch-Jozsa algorithm can be used to solve multiple computational problems in the query model framework. We will then discuss a related quantum algorithm — known as Simon's algorithm — which, for reasons that will be explained, offers a more robust and satisfying advantage of quantum over classical computations.

The query model of computation

When we model computations in mathematical terms, we typically have in mind the sort of process represented by the following figure, where information is provided as input, a computation takes place, and output is produced.

Illustration of a standard computation.

It is true that the computers we use today continuously receive input and produce output, essentially interacting with both us and with other computers in a way not reflected by the figure. Here, however, the intention is not to represent the ongoing operation of computers, but rather to create a simple abstraction of computation, focusing on isolated computational tasks.

For example, the input might encode a number, a vector, a matrix, a graph, a description of a molecule, or something more complicated, while the output encodes a solution to the computational task we have in mind. The key point is that the input is provided to the computation, usually in the form of a binary string, with no part of it being hidden.

High-level description

In the query model of computation, on the other hand, the entire input is not provided to the computation. Rather, the input is made available in the form of a function, which the computation accesses by making queries. Alternatively, we may view that computations in the query model have to bits (or segments of bits) of the input.

Illustration of a computation in the query model.

We often refer to the input as being provided by an oracle or black box in the context of the query model. Both terms suggest that a complete description of the input is hidden from the computation, with the only way to access it being to ask questions. It is as if we're consulting the Oracle at Delphi about the input: she won't tell us everything she knows, she only answers specific questions. The term black box makes sense especially when we think about the input as being represented by a function: we cannot look inside the function and understand how it works, we can only evaluate it on arguments we select.

We're going to be working exclusively with binary strings throughout the lesson, so let's write Σ={0,1}\Sigma = \{0,1\} to refer to the binary alphabet throughout for the sake of brevity. We'll be thinking about different computational problems, with some simple examples described shortly, but for all of them the input will be represented by a function taking the form

f:ΣnΣmf:\Sigma^n \rightarrow \Sigma^m

for two positive integers nn and m.m. Naturally, we could choose a different name in place of f,f, but we'll stick with ff throughout the lesson.

To say that a computation makes a query means that some string xΣnx \in \Sigma^n is selected and then the string f(x)Σmf(x)\in\Sigma^m is made available. The precise way that this works for quantum algorithms will be discussed shortly — we need to make sure that this is possible to do with a unitary quantum operation allowing queries to be made in superposition — but for now we can think about it intuitively at a high level.

Finally, the way that we will measure efficiency of query algorithms is simple: we'll count the number of queries to the input they require. This is related to the time required to perform a computation, but it's not exactly the same because we're ignoring the time required for operations other than the queries, and we're treating the queries as if they each have unit cost. (We can also take the operations besides the queries into account, and that is sometimes done — but restricting our attention just to the number of queries helps to keep things simple.)

Examples of query problems

Here are a few simple examples of query problems.

  • OR. The input function takes the form f:ΣnΣf:\Sigma^n \rightarrow \Sigma (so m=1m=1 for this problem). The task is to output 11 if there exists a string xΣnx\in\Sigma^n for which f(x)=1,f(x) = 1, and to output 00 if there is no such string. If we think about the function ff as representing a sequence of 2n2^n bits to which we have random access, the problem is to compute the OR of these bits.

  • Parity. The input function again takes the form f:ΣnΣ.f:\Sigma^n \rightarrow \Sigma. The task is to determine whether the number of strings xΣnx\in\Sigma^n for which f(x)=1f(x) = 1 is even or odd. To be precise, the required output is 00 if the set {xΣn:f(x)=1}\{x\in\Sigma^n : f(x) = 1\} has an even number of elements and 11 if it has an odd number of elements. If we think about the function ff as representing a sequence of 2n2^n bits to which we have random access, the problem is to compute the parity (or XOR) of these bits.

  • Minimum. The input function takes the form f:ΣnΣmf:\Sigma^n \rightarrow \Sigma^m for any choices of positive integers nn and m.m. The required output is the string y{f(x):xΣn}y \in \{f(x) : x\in\Sigma^n\} that comes first in the lexicographic (i.e., dictionary) ordering of Σm.\Sigma^m. If we think about the function ff as representing a sequence of 2n2^n integers encoded as strings of length mm in binary notation to which we have random access, the problem is to compute the minimum of these integers.

Sometimes we also consider query problems where we have a promise on the input. What this means is that we're given some sort of guarantee on the input, and we're not responsible for what happens when this guarantee is not met. Another way to describe this type of problem is that some input functions (the ones for which the promise is not satisfied) are considered as "don't care" inputs. No requirements at all are placed on algorithms when they're given "don't care" inputs.

Here's one example of a problem with a promise:

  • Unique search. The input function takes the form f:ΣnΣ,f:\Sigma^n \rightarrow \Sigma, and we are promised that there is exactly one string zΣnz\in\Sigma^n for which f(z)=1,f(z) = 1, with f(x)=0f(x) = 0 for all strings xz.x\neq z. The task is to find this unique string z.z.

All four of the examples just described are natural in the sense that they're easy to describe and we can imagine a variety of situations or contexts in which they might arise — but some query problems of interest aren't like this at all. In the study of the query model we sometimes come up with very complicated and highly contrived problems where it's difficult to imagine that anyone would ever actually want to solve them in practice. This doesn't mean the problems aren't interesting, it's just part of the study of the model to look for extremes that reveal potential advantages of quantum computing. Sometimes things that seems contrived or unnatural can provide unexpected clues or inspire new ideas. Shor's quantum algorithm for factoring, which was directly inspired by Simon's algorithm, is a great example.

Query gates

When we're working with circuit models of computation, queries are made by special gates called query gates. The simplest way to do this for classical Boolean circuits is to define query gates that compute the input function ff directly, as the following figure suggests.

A classical query gate.

When we create a Boolean circuit for a query problem, the input function ff is accessed through these gates, and the number of queries that the circuit makes is simply the number of query gates that appear in the circuit. In this case the wires of the circuit are initialized to fixed values, which should be considered as part of the algorithm rather than as part of the input.

For example, here is a Boolean circuit with classical query gates that solves the parity problem described above for a function of the form f:ΣΣf:\Sigma\rightarrow\Sigma:

Classical query algorithm for parity.

This algorithm makes two queries because there are two query gates. The way it works is that the function ff is queried on the two possible inputs, 00 and 1,1, and the results are plugged into the Boolean circuit from Lesson 3 that computes the XOR.

For quantum circuits, this definition of query gates doesn't work very well because we'll get non-unitary gates for some functions f,f, and we won't be able to apply them to quantum states. What we do instead is to define unitary query gates that operate as this figure suggests on standard basis states:

A unitary query gate.

Here our assumption is that xΣnx\in\Sigma^n and yΣmy\in\Sigma^m are arbitrary strings. The notation yf(x)y\oplus f(x) refers to the bitwise exclusive OR of strings (which have length mm in this case). For example, 001101=100.001 \oplus 101 = 100.

Intuitively speaking, what the gate UfU_f does (for any chosen function ff) is to echo the top input string xx and XOR the function value f(x)f(x) onto the bottom input string y.y. This is a unitary operation for every function f.f. To be more precise, as a matrix UfU_f is always a permutation matrix, meaning a matrix with a single 11 in each row and each column, with all other entries being 0.0. Applying a permutation matrix to a vector simply shuffles the entries of the vector (hence the term permutation matrix), and therefore does not change that vector's Euclidean norm — revealing that permutation matrices are always unitary.

Notice that when we analyze query algorithms by simply counting the number of queries that a query algorithm makes, we're completely ignoring the difficulty of physically constructing the query gates (for both the classical and quantum versions just described). That might seem unreasonable — but we must keep in mind that we're not trying to describe practical computing or fully account for the resources required. Rather, we're just defining a theoretical model that helps to shed light on the potential advantages of quantum computing. We will have more to say about this point in the lesson following this one when we turn our attention to a more standard model of computation where inputs are given explicitly to circuits as binary strings.

Deutsch's algorithm

Deutsch's algorithm solves the parity problem (described above) for the special case that n=1.n = 1. In the context of quantum computing this problem is sometimes referred to as Deutsch's problem, and we will follow that nomenclature in this lesson — but really it's just the simplest possible nontrivial version of the parity problem.

To be precise, the input is represented by a function f:ΣΣf:\Sigma \rightarrow \Sigma from one bit to one bit. As we observed in Lesson 1, there are 4 such functions:

af1(a)0010af2(a)0011af3(a)0110af4(a)0111\rule[-10mm]{0mm}{15mm} \begin{array}{c|c} a & f_1(a)\\ \hline 0 & 0\\ 1 & 0 \end{array} \qquad \begin{array}{c|c} a & f_2(a)\\ \hline 0 & 0\\ 1 & 1 \end{array} \qquad \begin{array}{c|c} a & f_3(a)\\ \hline 0 & 1\\ 1 & 0 \end{array} \qquad \begin{array}{c|c} a & f_4(a)\\ \hline 0 & 1\\ 1 & 1 \end{array}

The first and last of these functions are constant and the middle two are balanced, meaning that the two possible output values for the function occur the same number of times as we range over the inputs. Deutsch's problem is to determine which of these two categories the input function belongs to: constant or balanced.

Deutsch's problem
Input: a function f:{0,1}{0,1}f:\{0,1\}\rightarrow\{0,1\}
Output: 00 if ff is constant, 11 if ff is balanced

If we view the input function ff in Deutsch's problem as representing random access to a string, we're thinking about a two-bit string: f(0)f(1).f(0)f(1).

functionstringf100f201f310f411\begin{array}{cc} \mathsf{function} & \mathsf{string}\\ \hline f_1 & 00 \\ f_2 & 01 \\ f_3 & 10 \\ f_4 & 11 \end{array}

When viewed in this way, Deutsch's problem is to compute the parity (or, equivalently, the exclusive-OR) of the two bits.

Any classical query algorithm for this problem must query both bits, f(0)f(0) and f(1).f(1). If we learn that f(1)=1,f(1) = 1, for instance, the answer could still be 00 (in case f(0)=1f(0) = 1) or 11 (in case f(0)=0f(0) = 0). Every other case is similar: knowing just one of two bits doesn't provide any information at all about their parity. So, the Boolean circuit described in the previous section is the best we can do in terms of the number of queries required to solve this problem.

Quantum circuit description

Deutsch's algorithm solves Deutsch's problem using a single query, therefore providing a quantifiable advantage of quantum over classical computations. This may be a modest advantage — one query as opposed to two — but we have to start somewhere. Scientific advancements often have seemingly humble origins.

Here is a quantum circuit that describes Deutsch's algorithm:

Deutsch's algorithm

Analysis

To analyze Deutsch's algorithm, we will trace through the action of the circuit above and identify the states of the qubits at the times suggested by this figure:

States during Deutsch's algorithm

The initial state is 10,\vert 1\rangle \vert 0 \rangle, and the two Hadamard operations on the left-hand side of the circuit transform this state to

π1=+=12(01)0+12(01)1.\vert \pi_1 \rangle = \vert - \rangle \vert + \rangle = \frac{1}{2} \bigl( \vert 0\rangle - \vert 1\rangle \bigr) \vert 0\rangle + \frac{1}{2} \bigl( \vert 0\rangle - \vert 1\rangle \bigr) \vert 1\rangle.

Notice here that we've only partially expanded out the expression of this state (by expanding +\vert + \rangle but not \vert - \rangle). There is nothing a priori that tells us that we should express the state in this way, but it turns out to be convenient for the analysis.

Next, the UfU_f gate is performed. According to the definition of the UfU_f gate, the value of the function ff for the classical state of the top/rightmost qubit is XORed onto the bottom/leftmost qubit, which transforms π1\vert \pi_1\rangle into the state

π2=12(0f(0)1f(0))0+12(0f(1)1f(1))1.\vert \pi_2 \rangle = \frac{1}{2} \bigl( \vert 0 \oplus f(0) \rangle - \vert 1 \oplus f(0) \rangle \bigr) \vert 0 \rangle + \frac{1}{2} \bigl( \vert 0 \oplus f(1) \rangle - \vert 1 \oplus f(1) \rangle \bigr) \vert 1 \rangle.

We can simplify this expression by observing that the formula

0a1a=(1)a(01)\vert 0 \oplus a\rangle - \vert 1 \oplus a\rangle = (-1)^a \bigl( \vert 0\rangle - \vert 1\rangle \bigr)

works for both possible values aΣ.a\in\Sigma. More explicitly, the two cases are as follows.

0010=01=(1)0(01)0111=10=(1)1(01)\begin{aligned} \vert 0 \oplus 0\rangle - \vert 1 \oplus 0\rangle & = \vert 0 \rangle - \vert 1 \rangle = (-1)^0 \bigl( \vert 0\rangle - \vert 1\rangle \bigr)\\ \vert 0 \oplus 1\rangle - \vert 1 \oplus 1\rangle & = \vert 1 \rangle - \vert 0\rangle = (-1)^1 \bigl( \vert 0\rangle - \vert 1\rangle \bigr) \end{aligned}

Thus, we can alternatively express π2\vert\pi_2\rangle like this:

π2=12(1)f(0)(01)0+12(1)f(1)(01)1=((1)f(0)0+(1)f(1)12).\begin{aligned} \vert\pi_2\rangle & = \frac{1}{2} (-1)^{f(0)} \bigl( \vert 0 \rangle - \vert 1 \rangle \bigr) \vert 0 \rangle + \frac{1}{2} (-1)^{f(1)} \bigl( \vert 0 \rangle - \vert 1 \rangle \bigr) \vert 1 \rangle \\ & = \vert - \rangle \biggl( \frac{(-1)^{f(0)} \vert 0\rangle + (-1)^{f(1)} \vert 1\rangle}{\sqrt{2}}\biggr). \end{aligned}

Something interesting has happened here! Although the action of the UfU_f gate on standard basis states leaves the top/rightmost qubit alone and XORs the function value onto the bottom/leftmost qubit, here we see that the state of the top/rightmost qubit has changed (in general) while the state of the bottom/leftmost qubit remains the same — specifically being in the \vert - \rangle state before and after the UfU_f gate is performed. This phenomenon is known as the phase kickback, and we will have more to say about it shortly.

With one final simplification, which is to pull the factor of (1)f(0)(-1)^{f(0)} outside of the sum, we obtain this expression of the state π2\vert\pi_2\rangle:

π2=(1)f(0)(0+(1)f(0)f(1)12)={(1)f(0)+if f(0)f(1)=0(1)f(0)if f(0)f(1)=1.\begin{aligned} \vert\pi_2\rangle & = (-1)^{f(0)} \vert - \rangle \biggl( \frac{\vert 0\rangle + (-1)^{f(0) \oplus f(1)} \vert 1\rangle}{\sqrt{2}}\biggr) \\ & = \begin{cases} (-1)^{f(0)} \vert - \rangle \vert + \rangle & \text{if $f(0) \oplus f(1) = 0$}\\[1mm] (-1)^{f(0)} \vert - \rangle \vert - \rangle & \text{if $f(0) \oplus f(1) = 1$}. \end{cases} \end{aligned}

Notice that in this expression, we have f(0)f(1)f(0) \oplus f(1) in the exponent of 1-1 rather than f(1)f(0),f(1) - f(0), but we obtain the same value after exponentiating either way. This is because the value (1)k(-1)^k for any integer kk depends only on whether kk is even or odd.

Applying the final Hadamard gate to the top qubit leaves us with the state

π3={(1)f(0)0if f(0)f(1)=0(1)f(0)1if f(0)f(1)=1,\vert \pi_3 \rangle = \begin{cases} (-1)^{f(0)} \vert - \rangle \vert 0 \rangle & \text{if $f(0) \oplus f(1) = 0$}\\[1mm] (-1)^{f(0)} \vert - \rangle \vert 1 \rangle & \text{if $f(0) \oplus f(1) = 1$}, \end{cases}

which leads to the correct outcome with probability 11 when the right/topmost qubit is measured.

Further remarks on the phase kickback

Before moving on, let's look at the analysis above from a slightly different angle that sheds some light on the phase kickback phenomenon.

First, notice the following formula works for all choices of bits b,cΣ.b,c\in\Sigma.

bc=Xcb\vert b \oplus c\rangle = X^c \vert b \rangle

This can be verified by checking it for the two possible values c=0c = 0 and c=1c = 1:

b0=b=1b=X0bb1=¬b=Xb=X1b.\begin{aligned} \vert b \oplus 0 \rangle & = \vert b\rangle = \mathbb{1} \vert b \rangle = X^0 \vert b \rangle\\ \vert b \oplus 1 \rangle & = \vert \neg b\rangle = X \vert b \rangle = X^1 \vert b \rangle. \end{aligned}

Using this formula, we see that

Uf(ba)=bf(a)a=(Xf(a)b)aU_f \bigl(\vert b\rangle \vert a \rangle\bigr) = \vert b \oplus f(a) \rangle \vert a \rangle = \bigl(X^{f(a)}\vert b \rangle\bigr) \vert a \rangle

for every choice of bits a,bΣ.a,b\in\Sigma. Because this formula is true for b=0b=0 and b=1,b=1, we see by linearity that

Uf(ψa)=(Xf(a)ψ)aU_f \bigl( \vert \psi \rangle \vert a \rangle \bigr) = \bigl(X^{f(a)}\vert \psi \rangle\bigr) \vert a \rangle

for all qubit state vectors ψ,\vert \psi\rangle, and therefore

Uf(a)=(Xf(a))a=(1)f(a)a.U_f \bigl( \vert - \rangle \vert a \rangle \bigr) = \bigl(X^{f(a)} \vert - \rangle \bigr) \vert a \rangle = (-1)^{f(a)} \vert - \rangle \vert a \rangle.

The key that makes this work is that X=.X\vert - \rangle = - \vert - \rangle. In mathematical terms, the vector \vert - \rangle is an eigenvector of the matrix XX having eigenvalue 1.-1. (We'll discuss eigenvectors and eigenvalues in greater detail in Lesson 7, where the phase kickback phenomenon is generalized to other unitary operations.)

Keeping in mind that scalars float freely through tensor products, we find an alternative way of reasoning how the operation UfU_f transforms π1\vert \pi_1\rangle into π2\vert \pi_2\rangle in the analysis above:

π2=Uf(+)=12Uf(0)+12Uf(1)=((1)f(0)0+(1)f(1)12).\begin{aligned} \vert \pi_2 \rangle & = U_f \bigl( \vert - \rangle \vert + \rangle \bigr)\\ & = \frac{1}{\sqrt{2}} U_f \bigl(\vert - \rangle \vert 0\rangle \bigr) + \frac{1}{\sqrt{2}} U_f \bigl(\vert - \rangle \vert 1\rangle \bigr)\\ & = \vert - \rangle \biggl( \frac{(-1)^{f(0)} \vert 0\rangle + (-1)^{f(1)} \vert 1\rangle}{\sqrt{2}}\biggr). \end{aligned}

Implementation

First we'll define a quantum circuit that implements a query gate for one of the four functions f1,f_1, f2,f_2, f3,f_3, or f4f_4 from one bit to one bit described above (and in Lesson 1). As was discussed previously, the implementation of query gates is not really a part of Deutsch's algorithm itself — here we're essentially just showing one way to prepare the input (in the form of a circuit implementation of a query gate).

Authenticate to run code cells
Reset Copy to clipboard

No output produced

We can see what each circuit looks like using the draw method.

Authenticate to run code cells
Reset Copy to clipboard

Output:

Next we will create the actual quantum circuit for Deutsch's algorithm, substituting the query gate with a quantum circuit implementation given as an argument. Shortly we'll plug in one of the four circuits defined by the function deutsch_function we defined earlier. Barriers are included to show the visual separation between the query gate implementation and the rest of the circuit, but they aren't necessary and can safely be removed.

Authenticate to run code cells
Reset Copy to clipboard

No output produced

Again we can see what the circuit looks like using the draw method.

Authenticate to run code cells
Reset Copy to clipboard

Output:

Finally, we'll create a function that runs the circuit just defined one time and outputs the appropriate result: "constant" or "balanced."

Authenticate to run code cells
Reset Copy to clipboard

No output produced

The following code cell runs Deutsch's algorithm on any one of the four functions defined above.

Authenticate to run code cells
Reset Copy to clipboard

Output:

'balanced'

The Deutsch-Jozsa algorithm

Deutsch's algorithm provides an example of a quantum algorithm for a query problem that outperforms classical algorithms, but the advantage is quite modest: one query versus two. The Deutsch-Josza algorithm extends this advantage, and it can in fact be used to solve a couple of different query problems.

Here is a quantum circuit description of the Deutsch-Jozsa algorithm. (There may also be a classical post-processing step depending on the specific problem being solved.)

Deutsch-Jozsa algorithm

Of course, we haven't actually discussed what problems this algorithm solves; that is done in the subsections that follow.

The Deutsch-Jozsa problem

We will begin with the query problem for which the Deutsch-Josza algorithm was originally intended, which is known as the Deutsch-Jozsa problem.

For this problem, the input function takes the form f:ΣnΣf:\Sigma^n \rightarrow \Sigma for an arbitrary positive integer n.n. Like Deutsch's problem, the task is to output 00 if ff is constant and 11 if ff is balanced (which again means that the number of input strings on which the function takes the value 00 is equal to the number of input strings on which the function takes the value 11).

Notice that when nn is larger than 1,1, there are functions of the form f:ΣnΣf:\Sigma^n \rightarrow \Sigma that are neither constant nor balanced. If n=2,n = 2, for instance, the function defined as

f(00)=0f(01)=0f(10)=0f(11)=1\begin{aligned} f(00) & = 0 \\ f(01) & = 0 \\ f(10) & = 0 \\ f(11) & = 1 \end{aligned}

falls into neither of these two categories. For the Deutsch-Jozsa problem, we simply don't worry about functions like this — they are considered to be "don't care" inputs. That is, for this problem we have a promise that ff is either constant or balanced.

Deutsch-Jozsa problem
Input: a function f:{0,1}n{0,1}f:\{0,1\}^n\rightarrow\{0,1\}
Promise: ff is either constant or balanced
Output: 00 if ff is constant, 11 if ff is balanced

The Deutsch-Jozsa algorithm, with its single query, solves this problem in the following sense: if every one of the nn measurement outcomes is 0,0, then the function ff is constant; and otherwise, if at least one of the measurement outcomes is 1,1, then the function ff is balanced. Another way to say this is that there is a classical post-processing step, which is to compute the OR of the measurement outcomes.

Algorithm analysis

To analyze the performance of the Deutsch-Jozsa algorithm for the Deutsch-Jozsa problem, it's helpful to begin by thinking about the action of a single layer of Hadamard gates.

A Hadamard operation can be expressed as a matrix in the usual way,

H=(12121212),H = \begin{pmatrix} \frac{1}{\sqrt{2}} & \frac{1}{\sqrt{2}} \\[2mm] \frac{1}{\sqrt{2}} & -\frac{1}{\sqrt{2}} \end{pmatrix},

but we can also express this operation in terms of its action on standard basis states:

H0=120+121H1=120121.\begin{aligned} H \vert 0\rangle & = \frac{1}{\sqrt{2}} \vert 0 \rangle + \frac{1}{\sqrt{2}} \vert 1 \rangle\\[1mm] H \vert 1\rangle & = \frac{1}{\sqrt{2}} \vert 0 \rangle - \frac{1}{\sqrt{2}} \vert 1 \rangle. \end{aligned}

These two equations can be combined into a single formula,

Ha=120+12(1)a1=12b{0,1}(1)abb,H \vert a \rangle = \frac{1}{\sqrt{2}} \vert 0 \rangle + \frac{1}{\sqrt{2}} (-1)^a \vert 1 \rangle = \frac{1}{\sqrt{2}} \sum_{b\in\{0,1\}} (-1)^{ab} \vert b\rangle,

which is true for both choices of aΣ.a\in\Sigma.

Now suppose that instead of just a single qubit we have nn qubits, and we perform a Hadamard operation on each. The combined operation on the nn qubits is described by the tensor product HHH\otimes \cdots \otimes H (nn times), which we write as HnH^{\otimes n} for succinctness and clarity. Using the formula from above, followed by expanding and then simplifying, we can express the action of this combined operation on the classical states of nn qubits like this:

Hnxn1x1x0=(Hxn1)(Hx0)=(12yn1Σ(1)xn1yn1yn1)(12y0Σ(1)x0y0y0)=12nyn1y0Σn(1)xn1yn1++x0y0yn1y0\begin{aligned} & H^{\otimes n} \vert x_{n-1} \cdots x_1 x_0 \rangle \\ & \qquad = \bigl(H \vert x_{n-1} \rangle \bigr) \otimes \cdots \otimes \bigl(H \vert x_{0} \rangle \bigr) \\ & \qquad = \Biggl( \frac{1}{\sqrt{2}} \sum_{y_{n-1}\in\Sigma} (-1)^{x_{n-1} y_{n-1}} \vert y_{n-1} \rangle \Biggr) \otimes \cdots \otimes \Biggl( \frac{1}{\sqrt{2}} \sum_{y_{0}\in\Sigma} (-1)^{x_{0} y_{0}} \vert y_{0} \rangle \Biggr) \\ & \qquad = \frac{1}{\sqrt{2^n}} \sum_{y_{n-1}\cdots y_0 \in \Sigma^n} (-1)^{x_{n-1}y_{n-1} + \cdots + x_0 y_0} \vert y_{n-1} \cdots y_0 \rangle \end{aligned}

Here, by the way, we're writing binary strings of length nn as xn1x0x_{n-1}\cdots x_0 and yn1y0,y_{n-1}\cdots y_0, following the convention for numbering the individual bits used in Qiskit.

This formula provides us with a useful tool for analyzing the quantum circuit above. After the first layer of Hadamard gates is performed, the state of the n+1n+1 qubits (including the leftmost/bottom qubit, which is treated separately from the rest) is

(H1)(Hn00)=12nxn1x0Σnxn1x0.\bigl( H \vert 1 \rangle \bigr) \bigl( H^{\otimes n} \vert 0 \cdots 0 \rangle \bigr) = \vert - \rangle \otimes \frac{1}{\sqrt{2^n}} \sum_{x_{n-1}\cdots x_0 \in \Sigma^n} \vert x_{n-1} \cdots x_0 \rangle.

When the UfU_f operation is performed, this state is transformed into

12nxn1x0Σn(1)f(xn1x0)xn1x0\vert - \rangle \otimes \frac{1}{\sqrt{2^n}} \sum_{x_{n-1}\cdots x_0 \in \Sigma^n} (-1)^{f(x_{n-1}\cdots x_0)} \vert x_{n-1} \cdots x_0 \rangle

through exactly the same phase kick-back phenomenon that we saw in the analysis of Deutsch's algorithm.

Then the second layer of Hadamard gates is performed, which (by the same formula as above) transforms this state to

12nxn1x0Σnyn1y0Σn(1)f(xn1x0)+xn1yn1++x0y0yn1y0.\vert - \rangle \otimes \frac{1}{2^n} \sum_{x_{n-1}\cdots x_0 \in \Sigma^n} \sum_{y_{n-1}\cdots y_0 \in \Sigma^n} (-1)^{f(x_{n-1}\cdots x_0) + x_{n-1}y_{n-1} + \cdots + x_0 y_0} \vert y_{n-1} \cdots y_0 \rangle.

This state may appear to be a bit complicated, and without knowing more about the function ff there is little that can be concluded about the probabilities to obtain different measurement outcomes. But fortunately, all we need to know is the probability that every one of the measurement outcomes is 0,0, because that's the probability that the algorithm concludes that ff is constant. This probability is

12nxn1x0Σn(1)f(xn1x0)2={1if f is constant0if f is balanced.\Biggl\vert \frac{1}{2^n} \sum_{x_{n-1}\cdots x_0 \in \Sigma^n} (-1)^{f(x_{n-1}\cdots x_0)} \Biggr\vert^2 = \begin{cases} 1 & \textsf{if $f$ is constant}\\[1mm] 0 & \textsf{if $f$ is balanced}. \end{cases}

In greater detail, if ff is constant, then either f(xn1x0)=0f(x_{n-1}\cdots x_0) = 0 for every string xn1x0,x_{n-1}\cdots x_0, in which case the value of the sum is 2n,2^n, or f(xn1x0)=1f(x_{n-1}\cdots x_0) = 1 for every string xn1x0,x_{n-1}\cdots x_0, in which case the value of the sum is 2n.-2^n. Dividing by 2n2^n and taking the square of the absolute value yields 1.1. If, on the other hand, ff is balanced, then ff takes the value 00 on half of the strings xn1x0x_{n-1}\cdots x_0 and the value 11 on the other half, so the +1+1 terms and 1-1 terms in the sum cancel and we're left with the value 0.0.

So, we conclude that the algorithm operates correctly in the two cases that ff is constant and ff is balanced. (If ff is neither constant nor balanced, then all bets are off and we can't say too much — although there will always be a nonzero probability to obtain the outcome 11 for at least one of the measurements when ff is not constant.)

Classical difficulty

The Deutsch-Jozsa algorithm works 100% of the time, always giving us the correct answer when the promise is met, and requires a single query. How does this compare with classical query algorithms for the Deutsch-Jozsa problem?

Any deterministic classical algorithm for the Deutsch-Jozsa problem must make lots of queries — exponentially many, in fact. To be precise, 2n1+12^{n-1} + 1 queries are required in the worst case. This is because if a deterministic algorithm queries ff on 2n12^{n-1} different input strings and obtains the same value every time, it still doesn't know for sure if the function is constant or balanced — both answers are still possible depending on the output of ff on the strings that weren't queried. So, we have a very significant advantage of quantum over classical algorithms in this regard.

However, there is a catch, which is that probabilistic classical algorithms can solve the Deutsch-Jozsa problem with very high probability using just a few queries. In particular, if we simply choose some inputs to ff randomly and evaluate ff on those strings, it would be very unlikely for the output values to all be the same when ff is balanced. To be specific, if we choose kk input strings x1,,xkΣnx^1,\ldots,x^k \in \Sigma^n uniformly at random, evaluate f(x1),,f(xk)f(x^1),\ldots,f(x^k) and answer 00 if they're all the same and 11 if not, then we'll always be correct when ff is constant and wrong in the case that ff is balanced with probability just 2k+1.2^{-k + 1}. So, if we take k=11,k = 11, for instance, this algorithm will answer correctly with probability greater than 99.999.9%.

So, for this reason we do still have a rather modest advantage of quantum over classical algorithms — but it is nevertheless a quantifiable advantage that represents an improvement over Deutsch's algorithm.

Implementation

To implement the Deutsch-Jozsa algorithm in Qiskit, we'll start by generating a quantum circuit that implements a query operation for a randomly selected function that satisfies the promise: with 50% chance the function is constant, and with 50% change the function is balanced. For each possibility, the function is selected uniformly from the possibilities.

The argument to dj_function is the number of input bits of the function.

Authenticate to run code cells
Reset Copy to clipboard

No output produced

We can show the quantum circuit implementation of the query gate using the draw method as usual.

Authenticate to run code cells
Reset Copy to clipboard

Output:

Next we define a function that creates the Deutsch-Jozsa circuit, taking a quantum circuit implementation of a query gate as an argument.

Authenticate to run code cells
Reset Copy to clipboard

No output produced

Finally, a function that runs the Deutsch-Jozsa circuit once is defined.

Authenticate to run code cells
Reset Copy to clipboard

No output produced

We can test our implementation by choosing a function randomly, displaying the quantum circuit implementation of a query gate for this function, and then running the Deutsch-Jozsa algorithm on that function.

Authenticate to run code cells
Reset Copy to clipboard

Output:

'balanced'

The Bernstein-Vazirani problem

Next we will discuss a problem known as the Bernstein-Vazirani problem. It is also called the Fourier sampling problem, although there are more general formulations of this problem that also go by that name.

In order to describe this problem, it will be helpful to introduce some notation. For two binary strings x=xn1x0x = x_{n-1} \cdots x_0 and y=yn1y0y = y_{n-1}\cdots y_0 of length n,n, we define

xy=xn1yn1x0y0.x \cdot y = x_{n-1} y_{n-1} \oplus \cdots \oplus x_0 y_0.

We'll refer to this operation as the binary dot product. An alternative way to define it is as follows.

xy={1if xn1yn1++x0y0 is odd0if xn1yn1++x0y0 is even.x \cdot y = \begin{cases} 1 & \textsf{if $x_{n-1} y_{n-1} + \cdots + x_0 y_0$ is odd}\\[0.5mm] 0 & \textsf{if $x_{n-1} y_{n-1} + \cdots + x_0 y_0$ is even.} \end{cases}

Notice that this is a symmetric operation, meaning that the result doesn't change if we swap xx and y,y, so we're free to do that whenever it's convenient.

One way to think about the binary dot product xyx \cdot y is that it equals the parity of those bits of xx in positions where the string yy has a 1,1, which is equivalent to the parity of those bits of yy in positions where the string xx has a 1.1.

With this notation in hand we can now define the Bernstein-Vazirani problem.

Bernstein-Vazirani problem
Input: a function f:{0,1}n{0,1}f:\{0,1\}^n\rightarrow\{0,1\}
Promise: there exists a binary string s=sn1s0s = s_{n-1} \cdots s_0 for which f(x)=sxf(x) = s\cdot x for all xΣnx\in\Sigma^n
Output: the string ss

We don't actually need a new quantum algorithm for this problem, the Deutsch-Jozsa algorithm (not including the post-processing step of computing the OR of the measurement outcomes) solves it. In the interest of clarity, let's refer to the quantum circuit above (without the classical post-processing step) as the Deutsch-Jozsa circuit.

Algorithm analysis

To analyze the Deutsch-Jozsa circuit for the Bernstein-Vazirani problem, we'll begin with a quick observation. Using the binary dot product, we can alternatively describe the action of nn Hadamard gates on the standard basis states of nn qubits as follows.

Hnx=12nyΣn(1)xyyH^{\otimes n} \vert x \rangle = \frac{1}{\sqrt{2^n}} \sum_{y\in\Sigma^n} (-1)^{x\cdot y} \vert y\rangle

Similar to what we saw when analyzing Deutsch's algorithm, this is because the value (1)k(-1)^k for any integer kk depends only on whether kk is even or odd.

Turning to the circuit, after the first layer of Hadamard gates is performed, the state of the n+1n+1 qubits is

12nxΣnx.\vert - \rangle \otimes \frac{1}{\sqrt{2^n}} \sum_{x \in \Sigma^n} \vert x \rangle.

The query gate is then performed, which (through the phase kickback phenomenon) transforms the state to

12nxΣn(1)f(x)x.\vert - \rangle \otimes \frac{1}{\sqrt{2^n}} \sum_{x \in \Sigma^n} (-1)^{f(x)} \vert x \rangle.

Using the formula above for the action of the second layer of Hadamard gates, the state becomes

12nxΣnyΣn(1)f(x)+xyy.\vert - \rangle \otimes \frac{1}{2^n} \sum_{x \in \Sigma^n} \sum_{y \in \Sigma^n} (-1)^{f(x) + x \cdot y} \vert y \rangle.

We can now make some simplifications to this state, focusing on the exponent of 1-1 inside the sum. We're promised that f(x)=sxf(x) = s\cdot x for some string s=sn1s0,s = s_{n-1} \cdots s_0, so we can express the state as

12nxΣnyΣn(1)sx+xyy.\vert - \rangle \otimes \frac{1}{2^n} \sum_{x \in \Sigma^n} \sum_{y \in \Sigma^n} (-1)^{s\cdot x + x \cdot y} \vert y \rangle.

Because sxs\cdot x and xyx\cdot y are binary values, we can replace the addition with the exclusive-OR — again because the only thing that matters for an integer in the exponent of 1-1 is whether it is even or odd. Making use of the symmetry of the binary dot product, we obtain this expression for the state:

12nxΣnyΣn(1)(sx)(yx)y.\vert - \rangle \otimes \frac{1}{2^n} \sum_{x \in \Sigma^n} \sum_{y \in \Sigma^n} (-1)^{(s\cdot x) \oplus (y \cdot x)} \vert y \rangle.

(Parentheses have been added for clarity, although they aren't really necessary: it is conventional to treat the binary dot product as having higher precedence than the exclusive-OR. An easy way to remember this is that the binary dot product looks like multiplication and the exclusive-OR looks like addition.)

At this point we will make use of the following formula.

(sx)(yx)=(sy)x(s\cdot x) \oplus (y \cdot x) = (s \oplus y) \cdot x

We can obtain the formula through a similar formula for bits,

(ac)(bc)=(ab)c,(a c) \oplus (b c) = (a \oplus b) c,

together with an expansion of the binary dot product and bitwise exclusive-OR:

(sx)(yx)=(sn1xn1)(s0x0)(yn1xn1)(y0x0)=(sn1yn1)xn1(s0y0)x0=(sy)x\begin{aligned} (s\cdot x) \oplus (y \cdot x) & = (s_{n-1} x_{n-1}) \oplus \cdots \oplus (s_{0} x_{0}) \oplus (y_{n-1} x_{n-1}) \oplus \cdots \oplus (y_{0} x_{0}) \\ & = (s_{n-1} \oplus y_{n-1}) x_{n-1} \oplus \cdots \oplus (s_{0} \oplus y_{0}) x_{0} \\ & = (s \oplus y) \cdot x \end{aligned}

This allows us to express the state of the circuit immediately prior to the measurements like this:

12nxΣnyΣn(1)(sy)xy.\vert - \rangle \otimes \frac{1}{2^n} \sum_{x \in \Sigma^n} \sum_{y \in \Sigma^n} (-1)^{(s\oplus y)\cdot x} \vert y \rangle.

The final step is to make use of yet another formula, which works for every binary string z=zn1z0.z = z_{n-1}\cdots z_0.

12nxΣn(1)zx={1if z=0n0if z0n\frac{1}{2^n} \sum_{x \in \Sigma^n} (-1)^{z \cdot x} = \begin{cases} 1 & \text{if $z = 0^n$}\\ 0 & \text{if $z\neq 0^n$} \end{cases}

(Here we're using a simple notation for strings that we'll use several more times in the lesson: 0n0^n is the all-zero string of length n.n.)

A simple way to argue that this formula works is to consider the two cases separately. If z=0n,z = 0^n, then zx=0z\cdot x = 0 for every string xΣn,x\in\Sigma^n, so the value of each term in the sum is 1,1, and we obtain 11 by summing and dividing by 2n.2^n. On the other hand, if any one of the bits of zz is equal to 1,1, then the binary dot product zxz\cdot x is equal to 00 for exactly half of the possible choices for xΣnx\in\Sigma^n and 11 for the other half — because the value of the binary dot product zxz\cdot x flips (from 00 to 11 or from 11 to 00) if we flip the bit of xx in any position where zz has a 1.1.

If we now apply this formula to simplify the state of the circuit prior to the measurements, we obtain

12nxΣnyΣn(1)(sy)xy=s,\vert - \rangle \otimes \frac{1}{2^n} \sum_{x \in \Sigma^n} \sum_{y \in \Sigma^n} (-1)^{(s\oplus y)\cdot x} \vert y \rangle = \vert - \rangle \otimes \vert s \rangle,

owing to the fact that sy=0ns\oplus y = 0^n if and only if y=s.y = s.

Thus, the measurements reveal precisely the string ss we're looking for.

Classical difficulty

While the Deutsch-Jozsa circuit solves the Bernstein-Vazirani problem with a single query, any classical query algorithm must make at least nn queries to solve the problem. This can be reasoned through a so-called information theoretic argument. Each classical query reveals a single bit of information about the solution, and there are nn bits of information that need to be uncovered.

It is, in fact, possible to solve the Bernstein-Vazirani problem classically by querying the function on each of the nn strings having a single 1,1, in each possible position, and 00 for all other bits, which reveals the bits of ss one at a time.

So, the advantage of quantum over classical algorithms for this problem is 11 query versus nn queries.

Implementation

We've already implemented the Deutsch-Jozsa circuit above, and here we will make use of it to solve the Bernstein-Vazirani problem. First we'll define a function that implements a query gate for the Bernstein-Vazirani problem given any binary string s.s.

Authenticate to run code cells
Reset Copy to clipboard

Output:

Now we can run the Deutsch-Jozsa circuit on the function. Note that the compile_circuit function was defined in a previous code cell, which must be run prior to the cell that follows.

Authenticate to run code cells
Reset Copy to clipboard

Output:

'1001'

Remark on nomenclature

In the context of the Bernstein-Vazirani problem, it is common that the Deutsch-Jozsa algorithm is referred to as the "Bernstein-Vazirani algorithm." This is slightly misleading, because the algorithm is the Deutsch-Jozsa algorithm, as Bernstein and Vazirani made clear in their work.

What Bernstein and Vazirani did after showing that the Deutsch-Jozsa algorithm solves the Bernstein-Vazirani problem (as it is stated above) was to define a much more complicated problem, known as the recursive Fourier sampling problem. This is a highly contrived problem where solutions to different instances of the problem effectively unlock new levels of the problem arranged in a tree-like structure. The Bernstein-Vazirani problem described earlier is essentially just the base case of this more complicated problem.

This more complicated problem was the first known example of a query problem where quantum algorithms have a so-called super-polynomial advantage over probabilistic algorithms, thereby surpassing the advantage of quantum over classical offered by the Deutsch-Jozsa algorithm. Intuitively speaking, the recursive version of the problem effectively amplifies the 11 versus nn advantage of quantum algorithms to something much larger. Arguably, the most difficult part of this analysis is to show that classical query algorithms cannot solve the problem without making lots and lots of queries. This is actually quite typical — it can be very difficult to rule out creative approaches that allow classical query algorithms to solve problems efficiently.

Simon's problem, and the algorithm for it described in the next section, does provide a much simpler example of a super-polynomial (and, in fact, exponential) advantage of quantum over classical algorithms, and for this reason the recursive Fourier sampling problem is less often discussed. It is, nevertheless, a very interesting computational problem from the viewpoint of computational complexity theory.

Simon's algorithm

Simon's algorithm is a quantum query algorithm for a problem known as Simon's problem. This is a promise problem with a flavor similar to the Deutsch-Jozsa and Bernstein-Vazirani problems, but the specifics are different. Simon's algorithm is significant because it provides an exponential advantage of quantum over classical (including probabilistic) algorithms, and the technique it uses inspired Peter Shor's discovery of an efficient quantum algorithm for factoring integers.

Simon's problem

The input function for Simon's problem takes the form

f:ΣnΣmf:\Sigma^n \rightarrow \Sigma^m

for positive integers nn and m.m. We could restrict our attention to the case m=nm = n in the interest of simplicity, but there's not much to be gained in making this assumption — Simon's algorithm and its analysis are basically the same either way.

Simon's problem
Input: a function f:ΣnΣmf:\Sigma^n \rightarrow \Sigma^m
Promise: there exists a string sΣns\in\Sigma^n such that [f(x)=f(y)][(x=y)(xs=y)][f(x) = f(y)] \Leftrightarrow [(x = y) \vee (x \oplus s = y)] for all x,yΣnx,y\in\Sigma^n
Output: the string ss

We'll unpack the promise to better understand what says momentarily, but before doing this let's be clear that most functions don't satisfy this promise; it requires that ff has a very special structure. It's also important to note that if the promise is met, there can only be one string ss that works. So, there is always a unique correct answer to the problem for functions that satisfy the promise.

To better understand the promise, it is helpful to consider two main cases: the first case is that ss is the all-zero string 0n,0^n, and the second case is that ss is not the all-zero string.

  • Case 1: s=0n.s=0^n. If ss is the all-zero string, then we can simplify the if and only if statement in the promise so that it reads [f(x)=f(y)][x=y].[f(x) = f(y)] \Leftrightarrow [x = y]. Because this if and only if statement must be true for all strings x,yΣn,x,y\in\Sigma^n, we see that saying that the promise is met for the string s=0ns = 0^n is equivalent to saying that ff is a one-to-one function.

  • Case 2: s0n.s\neq 0^n. If ss is not the all-zero string, then the promise implies that ff is two-to-one, meaning that for every possible output string of f,f, there are exactly two input strings that cause ff to output that string — and moreover these two input strings must take the form xx and xsx \oplus s for some string x.x.

Here's an example of a function taking the form f:Σ3Σ5f:\Sigma^3 \rightarrow \Sigma^5 that satisfies the promise for the string s=011.s = 011.

f(000)=10011f(001)=00101f(010)=00101f(011)=10011f(100)=11010f(101)=00001f(110)=00001f(111)=11010\begin{aligned} f(000) & = 10011 \\ f(001) & = 00101 \\ f(010) & = 00101 \\ f(011) & = 10011 \\ f(100) & = 11010 \\ f(101) & = 00001 \\ f(110) & = 00001 \\ f(111) & = 11010 \end{aligned}

There are 88 different input strings and 44 different output strings, each of which occurs twice — so this is a two-to-one function. Moreover, for any two different input strings that produce the same output string, we see that the bitwise XOR of these two input strings is equal to 011,011, which is equivalent to saying that either one of them equals the other XORed with s.s.

Notice that the only thing that matters about the actual output strings is whether they're the same or different for different choices of input strings. For instance, in the example above, there are four strings (10011,10011, 00101,00101, 00001,00001, and 1101011010) that appear as outputs of f.f. We could replace these four strings with different strings, so long as they're all distinct, and the correct solution s=011s = 011 would not change.

Quantum circuit description

Here is a quantum circuit diagram representing Simon's algorithm.

Simon's algorithm

To be clear, there are nn qubits on the top that are acted upon by Hadamard gates and mm qubits on the bottom that go directly into the query gate. It looks very similar to the algorithms we've already discussed in the lesson, but this time there's no phase kickback.

To solve Simon's problem using this circuit will actually require several independent runs of this circuit followed by a classical post-processing step, which will be described later after the behavior of the circuit is analyzed.

Analysis

The analysis of Simon's algorithm begins along similar lines to the Deutsch-Jozsa algortithm. After the first layer of Hadamard gates is performed on the top nn qubits, the state becomes

12nxΣn0mx.\frac{1}{\sqrt{2^n}} \sum_{x\in\Sigma^n} \vert 0^m \rangle \vert x\rangle.

When the UfU_f is performed, the output of the function ff is XORed onto the all-zero state of the bottom mm qubits, leaving the n+mn+m qubits in the state

12nxΣnf(x)x.\frac{1}{\sqrt{2^n}} \sum_{x\in\Sigma^n} \vert f(x) \rangle \vert x\rangle.

When the second layer of Hadamard gates is performed, we obtain the following state by using the same formula for the action of a layer of Hadamard gates as before.

12nxΣnyΣn(1)yxf(x)y\frac{1}{2^n} \sum_{x\in\Sigma^n} \sum_{y\in\Sigma^n} (-1)^{y\cdot x} \vert f(x) \rangle \vert y\rangle

At this point the analysis diverges from the ones for the previous algorithms in this lesson. We're interested in the probability for the measurements to result in each possible string yΣn.y\in\Sigma^n. Using the rule discussed in Lesson 2, the probability p(y)p(y) to obtain the string yy is equal to

p(y)=12nxΣn(1)yxf(x)2p(y) = \left\|\frac{1}{2^n} \sum_{x\in\Sigma^n} (-1)^{y\cdot x} \vert f(x) \rangle \right\|^2

To get a better handle on these probabilities, we'll need just a bit more notation and terminology. First, the range of the function ff is the set containing all of its output strings.

range(f)={f(x):xΣn}\operatorname{range}(f) = \{ f(x) : x\in \Sigma^n \}

Second, for each string zrange(f),z\in\operatorname{range}(f), we express the set of all input strings that cause the function to evaluate to this output string zz as f1({z}).f^{-1}(\{z\}).

f1({z})={xΣn:f(x)=z}f^{-1}(\{z\}) = \{ x\in\Sigma^n : f(x) = z \}

(This notation should not be confused with the inverse of the function f.f. Here we don't necessarily have that ff is invertible. We also see that the argument on the left-hand side is the set {z}\{z\} rather than the element z,z, which is the clue we need to avoid confusion. The set f1({z})f^{-1}(\{z\}) is known as the preimage of {z}\{z\} under f.f. We can define the preimage under ff of any set in place of {z}\{z\} in an analogous way — it's the set of all elements that ff maps to that set.)

Using this notation, we can split up the sum in our expression for the probabilities above to obtain

p(y)=12nzrange(f)(xf1({z})(1)yx)z2.p(y) = \left\| \frac{1}{2^n} \sum_{z\in\operatorname{range}(f)} \Biggl(\sum_{x\in f^{-1}(\{z\})} (-1)^{y\cdot x}\Biggr) \vert z \rangle \right\|^2.

Every string xΣnx\in\Sigma^n is represented exactly once by the two summations — we're basically just putting these strings into separate buckets depending on which output string z=f(x)z = f(x) they produce when we evaluate the function f,f, and then summing separately over all the buckets.

We can now evaluate the Euclidean norm squared to obtain

p(y)=122nzrange(f)xf1({z})(1)yx2.p(y) = \frac{1}{2^{2n}} \sum_{z\in\operatorname{range}(f)} \left\vert \sum_{x\in f^{-1}(\{z\})} (-1)^{y\cdot x} \right\vert^2.

To simplify these probabilities further, let's take a look at the value

xf1({z})(1)yx2(1)\Biggl\vert \sum_{x\in f^{-1}(\{z\})} (-1)^{y\cdot x} \Biggr\vert^2 \tag{1}

for an arbitrary selection of zrange(f).z\in\operatorname{range}(f).

If it happens to be the case that s=0n,s = 0^n, then ff is a one-to-one function and there's always just a single element xf1({z}),x\in f^{-1}(\{z\}), for every zrange(f).z\in\operatorname{range}(f). The value of the expression (1)(1) is 11 in this case.

If, on the other hand, s0n,s\neq 0^n, then there are exactly two strings in the set f1({z}).f^{-1}(\{z\}). To be precise, if we choose wf1({z})w\in f^{-1}(\{z\}) to be any one of these two strings, then the other string must be wsw \oplus s by the promise in Simon's problem. Using this observation we can simply (1)(1) as follows.

xf1({z})(1)yx2=(1)yw+(1)y(ws)2=(1)yw(1+(1)ys)2=1+(1)ys2={4ys=00ys=1\begin{aligned} \Biggl\vert \sum_{x\in f^{-1}(\{z\})} (-1)^{y\cdot x} \Biggr\vert^2 & = \Bigl\vert (-1)^{y\cdot w} + (-1)^{y\cdot (w\oplus s)} \Bigr\vert^2 \\ & = \Bigl\vert (-1)^{y\cdot w} \Bigl(1 + (-1)^{y\cdot s}\Bigr) \Bigr\vert^2 \\ & = \Bigl\vert 1 + (-1)^{y\cdot s} \Bigr\vert^2 \\ & = \begin{cases} 4 & y \cdot s = 0\\ 0 & y \cdot s = 1 \end{cases} \end{aligned}

So, it turns out that the value (1)(1) is independent of the specific choice of zrange(f)z\in\operatorname{range}(f) in both cases.

We can now finish off the analysis by looking at the same two cases as before separately.

  • Case 1: s=0n.s = 0^n. In this case the function ff is one-to-one, so there are 2n2^n strings zrange(f),z\in\operatorname{range}(f), and we obtain

    p(y)=122n2n=12n.p(y) = \frac{1}{2^{2n}} \cdot 2^n = \frac{1}{2^n}.

    In words, the measurements result in a string yΣny\in\Sigma^n chosen uniformly at random.

  • Case 2: s0n.s \neq 0^n. In this case ff is two-to-one, so there are 2n12^{n-1} elements in range(f).\operatorname{range}(f). Using the formula from above we conclude that the probability to measure each yΣny\in\Sigma^n is

    p(y)=122nzrange(f)xf1({z})(1)yx2={12n1ys=00ys=1p(y) = \frac{1}{2^{2n}} \sum_{z\in\operatorname{range}(f)} \Biggl\vert \sum_{x\in f^{-1}(\{z\})} (-1)^{y\cdot x} \Biggr\vert^2 = \begin{cases} \frac{1}{2^{n-1}} & y \cdot s = 0\\ 0 & y \cdot s = 1 \end{cases}

    In words, we obtain a string chosen uniformly at random from the set {yΣn:ys=0},\{y\in\Sigma^n : y \cdot s = 0\}, which contains 2n12^{n-1} strings. (Because s0n,s\neq 0^n, exactly half of the binary strings of length nn have binary dot product 11 with ss and the other have binary dot product 00 with s,s, as we already observed in the analysis of the Deutsch-Jozsa algorithm for the Bernstein-Vazirani problem.)

Classical post-processing

We now know what the probabilities are for the possible measurement outcomes when we run the quantum circuit for Simon's algorithm. Is this enough information to determine ss?

The answer is yes, provided that we're willing to repeat the process several times and accept that it could fail with some probability (which we can make very small by running the circuit enough times). The essential idea is that each execution of the circuit provides us with statistical evidence concerning s,s, and we can use that evidence find ss with very high probability if we run the circuit sufficiently many times.

Let's suppose that we run the circuit independently kk times, for k=n+10.k = n + 10. There's nothing special about this particular number of iterations — we could take kk to be larger (or smaller) depending on the probability of failure we're willing to tolerate, as we will see. Choosing k=n+10k = n + 10 will ensure that we have greater than a 99.999.9% chance to recover s.s.

By running the circuit kk times, we obtain strings y1,...,ykΣn.y^1,...,y^{k} \in \Sigma^n. To be clear, the superscripts here are part of the names of these strings, not exponents or indexes to their bits, so we have

y1=yn11y01y2=yn12y02    yk=yn1ky0k\begin{aligned} y^1 & = y^1_{n-1} \cdots y^1_{0}\\ y^2 & = y^2_{n-1} \cdots y^2_{0}\\ & \;\; \vdots\\ y^{k} & = y^{k}_{n-1} \cdots y^{k}_{0} \end{aligned}

We now form a matrix MM having kk rows and nn columns by inserting the bits of these string as follows.

M=(yn11y01yn12y02yn1ky0k)M = \begin{pmatrix} y^1_{n-1} & \cdots & y^1_{0}\\ y^2_{n-1} & \cdots & y^2_{0}\\ \vdots & \ddots & \vdots \\ y^{k}_{n-1} & \cdots & y^{k}_{0} \end{pmatrix}

Now, we don't know what ss is at this point — our goal is to find this string. But imagine for a moment that we do know the string s,s, and we form a column vector vv from the bits of the string s=sn1s0s = s_{n-1} \cdots s_0 as follows.

v=(sn1s0)v = \begin{pmatrix} s_{n-1}\\ \vdots\\ s_0 \end{pmatrix}

If we perform the matrix-vector multiplication MvM v modulo 22 — meaning that we performed the multiplication as usual and then take the remainder of the entries of the result after dividing by 22 — we obtain this:

Mv=(y1sy2syks)=(000).M v = \begin{pmatrix} y^1 \cdot s\\ y^2 \cdot s\\ \vdots\\ y^{k} \cdot s \end{pmatrix} = \begin{pmatrix} 0\\ 0\\ \vdots\\ 0 \end{pmatrix}.

That is, treated as a column vector vv as just described, the string ss will always be an element of the null space of the matrix M,M, provided that we do the arithmetic modulo 2.2. This is true in both the case that s=0ns = 0^n and s0n.s\neq 0^n. (The all-zero vector is always in the null space of M,M, regardless of how MM is chosen — and the all-zero vector will be joined by the vector vv whose entries are the bits of ss in case s0n.s\neq 0^n.)

Using linear algebra, it is possible to efficiently calculate a description of the null space of M.M. Specifically, it can be done using Gaussian elimination, which works the same way when arithmetic is done modulo 22 as it does with real or complex numbers.

The question that remains is whether there will be any other vectors in the null space of MM besides the ones corresponding to 0n0^n and s.s.

The answer is that this is very unlikely when we choose k=n+10.k = n + 10. To be precise, in both the case that s=0ns = 0^n and s0n,s\neq 0^n, with probability greater than 12101 - 2^{-10} the vectors in the null space of MM will correspond precisely to the set {0n,s}.\{0^n, s\}. So, we have greater than a 99.999.9% chance to determine ss from the null space of M,M, as claimed when we chose k.k.

If we replace k=n+10k = n + 10 with k=n+rk = n + r for any choice of a positive integer r,r, the probability of success will be at least 12r,1 - 2^{-r}, so the probability of failure can very easily be made extremely small. For instance, if we take k=2n,k = 2n, the probability of failure becomes exponentially small in n.n.

Classical difficulty

How many queries does a classical query algorithm need to solve Simon's problem?

The answer is: a lot, in general.

There are different precise statements that can be made about the classical difficulty of this problem, and here's just one of them. If we have any probabilistic query algorithm, and that algorithm makes fewer than 2n/2112^{n/2 - 1} - 1 queries, which is a number of queries that's exponential in n,n, then that algorithm will fail to solve Simon's problem with probability at least 1/2.1/2.

Sometimes, proving impossibility results like this can be very challenging, but this one isn't too difficult to prove through an elementary probabilistic analysis. Here we will just briefly examine the intuition behind this analysis.

We're trying to find the hidden string s,s, but so long as we don't query the function on two strings having the same output value, we'll get very limited information about s. Intuitively speaking, all we learn is that the hidden string ss is not the exclusive-OR of any two distinct strings we've queried. And if we query fewer than 2n/2112^{n/2 - 1} - 1 strings, there will still be a lot of choices of ss that we haven't ruled out. This is not a formal proof, but this is the basic idea.

So, in summary, Simon's algorithm provides us with a striking advantage of quantum over classical algorithms within the query model. In particular, Simon's algorithm solves Simon's problem with a number of queries that's linear in the number of input bits nn of our function, whereas any classical algorithm, even if it's probabilistic, needs to make a number of queries that's exponential in nn in order to solve Simon's problem with a reasonable probability of success.

Implementation

To implement Simon's algorithm in Qiskit, we'll use the fact that we can convert unitary matrices into gates in Qiskit using the .unitary method. Specifically, we'll use this methodology to define a query gate for a randomly chosen function satisfying Simon's problem for a given string s.s.

Authenticate to run code cells
Reset Copy to clipboard

No output produced

Next we'll define a function that runs the circuit in Simon's problem kk times and reports the results.

Authenticate to run code cells
Reset Copy to clipboard

No output produced

The following code cell illustrates how the function works when we plug in the query gate constructed above. Feel free to try different arguments, but keep in mind that the cost of the simulation we've built is exponential in the number of qubits we require — so don't make the string ss too long if you don't want to wait!

Authenticate to run code cells
Reset Copy to clipboard

Output:

['11011',
 '11100',
 '11111',
 '01110',
 '00111',
 '11100',
 '10101',
 '01110',
 '01101',
 '00100',
 '11111',
 '01110']

To do the post-processing, we can make use of the galois package, which has a built-in function for computing the null space modulo 2.

Authenticate to run code cells
Reset Copy to clipboard

Output:

And finally we can try it out.

Authenticate to run code cells
Reset Copy to clipboard

Output:

Measurement results:

['11101',
 '00111',
 '01011',
 '11010',
 '11101',
 '01111',
 '00100',
 '10110',
 '01100',
 '00111',
 '01111',
 '10110',
 '00111',
 '00000',
 '00111']
Null space:

GF([[1, 0, 0, 1, 1]], order=2)
Guess for hidden string s:

'10011'

Was this page helpful?