Datasets:
Tasks:
Text Generation
Modalities:
Text
Sub-tasks:
language-modeling
Languages:
English
Size:
100K - 1M
License:
\chapter{Quantum circuits} | |
Now that we've discussed qubits, we can talk about how to use them in circuits. | |
The key change --- and the reason that quantum circuits can do things that | |
classical circuits cannot --- is the fact that we are allowing | |
linear combinations of $0$ and $1$. | |
\section{Classical logic gates} | |
In classical logic, we build circuits which take in some bits for input, | |
and output some more bits for input. | |
These circuits are built out of individual logic gates. | |
For example, the \vocab{AND gate} can be pictured as follows. | |
\[ | |
\Qcircuit @C=0.8em @R=.7em { | |
\lstick{0} &\qw & \multigate{1}{\textsc{and}} & \rstick{0} \qw \\ | |
\lstick{0} &\qw & \ghost{\textsc{and}} & | |
} | |
\hspace{4.5em} | |
\Qcircuit @C=0.8em @R=.7em { | |
\lstick{0} &\qw & \multigate{1}{\textsc{and}} & \rstick{0} \qw \\ | |
\lstick{1} &\qw & \ghost{\textsc{and}} & | |
} | |
\hspace{4.5em} | |
\Qcircuit @C=0.8em @R=.7em { | |
\lstick{1} &\qw & \multigate{1}{\textsc{and}} & \rstick{0} \qw \\ | |
\lstick{0} &\qw & \ghost{\textsc{and}} & | |
} | |
\hspace{4.5em} | |
\Qcircuit @C=0.8em @R=.7em { | |
\lstick{1} &\qw & \multigate{1}{\textsc{and}} & \rstick{1} \qw \\ | |
\lstick{1} &\qw & \ghost{\textsc{and}} & | |
} | |
\] | |
One can also represent the AND gate using the ``truth table'': | |
\[ | |
\begin{array}{|cc|c|} | |
\hline | |
A & B & A \text{ and } B \\ \hline | |
0 & 0 & 0 \\ | |
0 & 1 & 0 \\ | |
1 & 0 & 0 \\ | |
1 & 1 & 1 \\ | |
\hline | |
\end{array} | |
\] | |
Similarly, we have the \vocab{OR gate} and the \vocab{NOT gate}: | |
\[ | |
\begin{array}{|cc|c|} | |
\hline | |
A & B & A \text{ or } B \\ \hline | |
0 & 0 & 0 \\ | |
0 & 1 & 1 \\ | |
1 & 0 & 1 \\ | |
1 & 1 & 1 \\ | |
\hline | |
\end{array} | |
\qquad | |
\begin{array}{|c|c|} | |
\hline | |
A & \text{not } A \\ \hline | |
0 & 1 \\ | |
1 & 0 \\ | |
\hline | |
\end{array} | |
\] | |
We also have a so-called \vocab{COPY gate}, which duplicates a bit. | |
\[ | |
\Qcircuit @C=0.8em @R=.7em { | |
\lstick{0} & \qw & \multigate{1}{\textsc{copy}} & \rstick{0} \qw & \\ | |
&& & \rstick{0} \qw & | |
} | |
\qquad\qquad | |
\Qcircuit @C=0.8em @R=.7em { | |
\lstick{1} & \qw & \multigate{1}{\textsc{copy}} & \rstick{1} \qw & \\ | |
&& & \rstick{1} \qw & | |
} | |
\] | |
Of course, the first theorem you learn about these gates is that: | |
\begin{theorem} | |
[AND, OR, NOT, COPY are universal] | |
The set of four gates AND, OR, NOT, COPY is universal in the sense that | |
any boolean function $f : \{0,1\}^n \to \{0,1\}$ | |
can be implemented as a circuit using only these gates. | |
\end{theorem} | |
\begin{proof} | |
Somewhat silly: we essentially write down a circuit that OR's across | |
all input strings in $f\pre(1)$. | |
For example, suppose we have $n=3$ and want to simulate the function | |
$f(abc)$ with $f(011) = f(110) = 1$ and $0$ otherwise. | |
Then the corresponding Boolean expression for $f$ is simply | |
\[ | |
f(abc) = | |
\left[ \text{(not $a$) and $b$ and $c$} \right] | |
\text{ or } | |
\left[ \text{$a$ and $b$ and (not $c$)} \right]. | |
\] | |
Clearly, one can do the same for any other $f$, | |
and implement this logic into a circuit. | |
\end{proof} | |
\begin{remark} | |
Since | |
$x \text{ and } y = \text{not } ( (\text{not $x$}) \text{ or } (\text{not $y$}))$, | |
it follows that in fact, we can dispense with the AND gate. | |
\end{remark} | |
\section{Reversible classical logic} | |
\prototype{CNOT gate, Toffoli gate.} | |
For the purposes of quantum mechanics, this is not enough. | |
To carry through the analogy we in fact need gates that are \vocab{reversible}, | |
meaning the gates are bijections from the input space to the output space. | |
In particular, such gates must take the same number of input and output gates. | |
\begin{example}[Reversible gates] | |
\listhack | |
\begin{enumerate}[(a)] | |
\ii None of the gates AND, OR, COPY are reversible for dimension reasons. | |
\ii The NOT gate, however, is reversible: | |
it is a bijection $\{0,1\} \to \{0,1\}$. | |
\end{enumerate} | |
\end{example} | |
\begin{example} | |
[The CNOT gate] | |
The controlled-NOT gate, or the \vocab{CNOT} gate, | |
is a reversible $2$-bit gate with the following truth table. | |
\[ | |
\begin{array}{|rr|rr|} | |
\hline | |
\multicolumn{2}{|c|}{\text{In}} & \multicolumn{2}{|c|}{\text{Out}} \\ | |
\hline | |
0 & 0 & 0 & 0 \\ | |
1 & 0 & 1 & 1 \\ | |
0 & 1 & 0 & 1 \\ | |
1 & 1 & 1 & 0 \\ \hline | |
\end{array} | |
\] | |
In other words, this gate XOR's the first bit to the second bit, | |
while leaving the first bit unchanged. | |
It is depicted as follows. | |
\[ | |
\Qcircuit @C=1em @R=.7em { | |
\lstick{x} & \ctrl{1} & \rstick{x} \qw \\ | |
\lstick{y} & \targ & \rstick{x+y \mod 2} \qw | |
} | |
\] | |
The first dot is called the ``control'', | |
while the $\oplus$ is the ``negation'' operation: | |
the first bit controls whether the second bit gets flipped or not. | |
Thus, a typical application might be as follows. | |
\[ | |
\Qcircuit @C=1em @R=.7em { | |
\lstick{1} & \ctrl{1} & \rstick{1} \qw \\ | |
\lstick{0} & \targ & \rstick{1} \qw | |
} | |
\] | |
\end{example} | |
So, NOT and CNOT are the only nontrivial reversible gates on two bits. | |
We now need a different definition of universal for our reversible gates. | |
\begin{definition} | |
A set of reversible gates can \vocab{simulate} a Boolean function $f(x_1 \dots x_n)$, | |
if one can implement a circuit which takes | |
\begin{itemize} | |
\ii As input, $x_1 \dots x_n$ plus some fixed bits set to $0$ or $1$, | |
called \vocab{ancilla bits}\footnote{% | |
The English word ``ancilla'' means ``maid''.}. | |
\ii As output, the input bits $x_1, \dots, x_n$, | |
the output bit $f(x_1, \dots, x_n)$, | |
and possibly some extra bits (called \vocab{garbage bits}). | |
\end{itemize} | |
The gate(s) are \vocab{universal} if they can simulate any Boolean function. | |
\end{definition} | |
For example, the CNOT gate can simulate the NOT gate, | |
using a single ancilla bit $1$, | |
according to the following circuit. | |
\[ | |
\Qcircuit @C=1em @R=.7em { | |
\lstick{x} & \ctrl{1} & \rstick{x} \qw \\ | |
\lstick{1} & \targ & \rstick{\text{not } x} \qw | |
} | |
\] | |
Unfortunately, it is not universal. | |
\begin{proposition} | |
[CNOT $\not\Rightarrow$ AND] | |
The CNOT gate cannot simulate the boolean function ``$x \text{ and } y$''. | |
\end{proposition} | |
\begin{proof}[Sketch of Proof] | |
One can see that any function simulated using only CNOT gates | |
must be of the form | |
\[ a_1 x_1 + a_2 x_2 + \dots + a_n x_n \pmod 2 \] | |
because CNOT is the map $(x,y) \mapsto (x, x+y)$. | |
Thus, even with ancilla bits, we can only create functions | |
of the form $ax+by+c \pmod 2$ for fixed $a$, $b$, $c$. | |
The AND gate is not of this form. | |
\end{proof} | |
So, we need at least a three-qubit gate. | |
The most commonly used one is: | |
\begin{definition} | |
The three-bit \vocab{Toffoli gate}, also called the CCNOT gate, is given by | |
\[ | |
\Qcircuit @C=1em @R=.7em { | |
\lstick{x} & \ctrl{1} & \rstick{x} \qw \\ | |
\lstick{y} & \ctrl{1} & \rstick{y} \qw \\ | |
\lstick{z} & \targ & \rstick{z + xy \pmod 2} \qw | |
} | |
\] | |
So the Toffoli has two controls, and toggles the last bit if and only if | |
both of the control bits are $1$. | |
\end{definition} | |
This replacement is sufficient. | |
\begin{theorem} | |
[Toffoli gate is universal] | |
The Toffoli gate is universal. | |
\end{theorem} | |
\begin{proof} | |
We will show it can \emph{reversibly} simulate | |
AND, NOT, hence OR, | |
which we know is enough to show universality. | |
(We don't need COPY because of reversibility.) | |
For the AND gate, we draw the circuit | |
\[ | |
\Qcircuit @C=1em @R=.7em { | |
\lstick{x} & \ctrl{1} & \rstick{x} \qw \\ | |
\lstick{y} & \ctrl{1} & \rstick{y} \qw \\ | |
\lstick{0} & \targ & \rstick{x \text{ and } y} \qw | |
} | |
\] | |
with one ancilla bit, and no garbage bits. | |
For the NOT gate, we use two ancilla $1$ bits and one garbage bit: | |
\[ | |
\Qcircuit @C=1em @R=.7em { | |
\lstick{1} & \ctrl{1} & \rstick{1} \qw \\ | |
\lstick{z} & \ctrl{1} & \rstick{z} \qw \\ | |
\lstick{1} & \targ & \rstick{\text{not } z} \qw | |
} | |
\] | |
This completes the proof. | |
\end{proof} | |
Hence, in theory we can create any classical circuit we desire | |
using the Toffoli gate alone. | |
Of course, this could require exponentially many gates for even the | |
simplest of functions. | |
Fortunately, this is NO BIG DEAL because I'm a math major, | |
and having $2^n$ gates is a problem best left for the CS majors. | |
\section{Quantum logic gates} | |
In quantum mechanics, since we can have \emph{linear combinations} of basis | |
elements, our logic gates will instead consist of \emph{linear maps}. | |
Moreover, in quantum computation, gates are always reversible, | |
which was why we took the time in the previous section to show | |
that we can still simulate any function when restricted to reversible gates | |
(e.g.\ using the Toffoli gate). | |
First, some linear algebra: | |
\begin{definition} | |
Let $V$ be a finite dimensional inner product space. | |
Then for a map $U : V \to V$, the following are equivalent: | |
\begin{itemize} | |
\ii $\left< U(x), U(y) \right> = \left< x,y \right>$ for $x,y \in V$. | |
\ii $U^\dagger$ is the inverse of $U$. | |
\ii $\norm{x} = \norm{U(x)}$ for $x \in V$. | |
\end{itemize} | |
The map $U$ is called \vocab{unitary} | |
if it satisfies these equivalent conditions. | |
\end{definition} | |
Then | |
\begin{moral} | |
Quantum logic gates are unitary matrices. | |
\end{moral} | |
In particular, unlike the classical situation, | |
quantum gates are always reversible | |
(and hence they always take the same number of input and output bits). | |
For example, consider the CNOT gate. | |
Its quantum analog should be a unitary map $\UCNOT : H \to H$, | |
where $H = \CC^{\oplus 2} \otimes \CC^{\oplus 2}$, | |
given on basis elements by | |
\[ | |
\UCNOT(\ket{00}) = \ket{00}, \quad | |
\UCNOT(\ket{01}) = \ket{01} | |
\] | |
\[ | |
\UCNOT(\ket{10}) = \ket{11}, \quad | |
\UCNOT(\ket{11}) = \ket{10}. | |
\] | |
So pictorially, the quantum CNOT gate is given by | |
\[ | |
\Qcircuit @C=0.8em @R=.7em { | |
\lstick{\ket0} & \ctrl{1} & \rstick{\ket0} \qw \\ | |
\lstick{\ket0} & \targ & \rstick{\ket0} \qw \\ | |
} | |
\hspace{6em} | |
\Qcircuit @C=0.8em @R=.7em { | |
\lstick{\ket0} & \ctrl{1} & \rstick{\ket0} \qw \\ | |
\lstick{\ket1} & \targ & \rstick{\ket1} \qw \\ | |
} | |
\hspace{6em} | |
\Qcircuit @C=0.8em @R=.7em { | |
\lstick{\ket1} & \ctrl{1} & \rstick{\ket1} \qw \\ | |
\lstick{\ket0} & \targ & \rstick{\ket1} \qw \\ | |
} | |
\hspace{6em} | |
\Qcircuit @C=0.8em @R=.7em { | |
\lstick{\ket1} & \ctrl{1} & \rstick{\ket1} \qw \\ | |
\lstick{\ket1} & \targ & \rstick{\ket0} \qw \\ | |
} | |
\] | |
OK, so what? | |
The whole point of quantum mechanics is that we allow linear | |
qubits to be in linear combinations of $\ket0$ and $\ket1$, | |
too, and this will produce interesting results. | |
For example, let's take $\xdown = \frac{1}{\sqrt2} (\ket0-\ket1)$ | |
and plug it into the top, with $\ket 1$ on the bottom, and see what happens: | |
\[ | |
\UCNOT \left( \xdown \otimes \ket1 \right) | |
= \UCNOT \left( \frac{1}{\sqrt2} (\ket{01}-\ket{11}) \right) | |
= \frac{1}{\sqrt2} \left( \ket{01}-\ket{10} \right) | |
= \ket{\Psi_-} | |
\] | |
which is the fully entangled \emph{singlet state}! Picture: | |
\[ | |
\Qcircuit @C=0.8em @R=.7em { | |
\lstick{\xdown} & \ctrl{1} & \rstick{\ket{\Psi_-}} \qw \\ | |
\lstick{\ket1} & \targ & \rstick{} \qw \\ | |
} | |
\] | |
Thus, when we input mixed states into our quantum gates, | |
the outputs are often entangled states, | |
even when the original inputs are not entangled. | |
\begin{example} | |
[More examples of quantum gates] | |
\listhack | |
\begin{enumerate}[(a)] | |
\ii Every reversible classical gate that we encountered before | |
has a quantum analog obtained in the same way as CNOT: | |
by specifying the values on basis elements. | |
For example, there is a quantum Tofolli gate which | |
for example sends | |
\[ | |
\Qcircuit @C=0.8em @R=.7em { | |
\lstick{\ket1} & \ctrl{1} & \rstick{\ket1} \qw \\ | |
\lstick{\ket1} & \ctrl{1} & \rstick{\ket1} \qw \\ | |
\lstick{\ket0} & \targ & \rstick{\ket1} \qw. | |
} | |
\] | |
\ii The \vocab{Hadamard gate} on one qubit is a rotation given by | |
\[ | |
\begin{bmatrix} | |
\frac{1}{\sqrt2} & \frac{1}{\sqrt2} \\ | |
\frac{1}{\sqrt2} & -\frac{1}{\sqrt2} | |
\end{bmatrix}. | |
\] | |
Thus, it sends $\ket0$ to $\xup$ and $\ket1$ to $\xdown$. | |
Note that the Hadamard gate is its own inverse. | |
It is depicted by an ``$H$'' box. | |
\[ | |
\Qcircuit @C=0.8em @R=.7em { | |
\lstick{\ket0} & \gate{H} & \rstick{\xup} \qw | |
} | |
\] | |
\ii More generally, if $U$ is a $2 \times 2$ unitary matrix | |
(i.e.\ a map $\CC^{\oplus 2} \to \CC^{\oplus 2}$) then | |
there is \vocab{$U$-rotation gate} similar to the previous one, | |
which applies $U$ to the input. | |
\[ | |
\Qcircuit @C=0.8em @R=.7em { | |
\lstick{\ket\psi} & \gate{U} & \rstick{U\ket\psi} \qw | |
} | |
\] | |
For example, the classical NOT gate is represented by $U = \sigma_x$. | |
\ii A \vocab{controlled $U$-rotation gate} generalizes the CNOT gate. | |
Let $U : \CC^{\oplus 2} \to \CC^{\oplus 2}$ be a rotation gate, | |
and let $H = \CC^{\oplus 2} \otimes \CC^{\oplus 2}$ be a $2$-qubit space. | |
Then the controlled $U$ gate has the following circuit diagrams. | |
\[ | |
\Qcircuit @C=0.8em @R=.7em { | |
\lstick{\ket0} & \ctrl{1} & \rstick{\ket0} \qw \\ | |
\lstick{\ket\psi} & \gate{U} & \rstick{\ket\psi} \qw | |
} | |
\hspace{8em} | |
\Qcircuit @C=0.8em @R=.7em { | |
\lstick{\ket1} & \ctrl{1} & \rstick{\ket1} \qw \\ | |
\lstick{\ket\psi} & \gate{U} & \rstick{U\ket\psi} \qw | |
} | |
\] | |
Thus, $U$ is applied when the controlling bit is $1$, | |
and CNOT is the special case $U = \sigma_x$. As before, | |
we get interesting behavior if the control is mixed. | |
\end{enumerate} | |
\end{example} | |
And now, some more counterintuitive quantum behavior. | |
Suppose we try to use CNOT as a copy, with truth table. | |
\[ | |
\begin{array}{|rr|rr|} | |
\hline | |
\multicolumn{2}{|c|}{\text{In}} & \multicolumn{2}{|c|}{\text{Out}} \\ | |
\hline | |
0 & 0 & 0 & 0 \\ | |
1 & 0 & 1 & 1 \\ | |
0 & 1 & 0 & 1 \\ | |
1 & 1 & 1 & 0 \\ \hline | |
\end{array} | |
\] | |
The point of this gate is to be used with a garbage $0$ at the bottom | |
to try and simulate a ``copy'' operation. | |
So indeed, one can check that | |
\[ | |
\Qcircuit @C=0.8em @R=.7em { | |
\lstick{\ket0} & \multigate{1}{U} & \rstick{\ket0} \qw \\ | |
\lstick{\ket0} & \ghost{U} & \rstick{\ket0} \qw | |
} | |
\hspace{8em} | |
\Qcircuit @C=0.8em @R=.7em { | |
\lstick{\ket1} & \multigate{1}{U} & \rstick{\ket1} \qw \\ | |
\lstick{\ket0} & \ghost{U} & \rstick{\ket1} \qw | |
} | |
\] | |
Thus we can copy $\ket0$ and $\ket1$. | |
But as we've already seen if we input $\xdown \otimes \ket0$ into $U$, | |
we end up with the entangled state $\ket{\Psi_-}$ | |
which is decisively \emph{not} the $\xdown \otimes \xdown$ we wanted. | |
And in fact, the so-called \vocab{no-cloning theorem} implies | |
that it's impossible to duplicate an arbitrary $\ket\psi$; | |
the best we can do is copy specific orthogonal states as in the classical case. | |
See also \Cref{prob:baby_no_clone}. | |
\section{Deutsch-Jozsa algorithm} | |
The Deutsch-Jozsa algorithm is the first example of a nontrivial | |
quantum algorithm which cannot be performed classically: | |
it is a ``proof of concept'' that would later inspire Grover's search algorithm | |
and Shor's factoring algorithm. | |
The problem is as follows: we're given a function $f : \{0,1\}^n \to \{0,1\}$, | |
and promised that the function $f$ is either | |
\begin{itemize} | |
\ii A constant function, or | |
\ii A balanced function, meaning that exactly half the inputs map to | |
$0$ and half the inputs map to $1$. | |
\end{itemize} | |
The function $f$ is given in the form of a reversible black box $U_f$ which | |
is the control of a NOT gate, so it can be represented as the circuit diagram | |
\[ | |
\Qcircuit @C=1em @R=0.7em { | |
\lstick{\ket{x_1x_2 \dots x_n}} & /^n \qw & \multigate{1}{U_f} & | |
\rstick{\ket{x_1x_2\dots x_n}} \qw \\ | |
\lstick{\ket{y}} & \qw & \ghost{U_f} & \rstick{\ket{y+f(x) \mod 2}}\qw \\ | |
} | |
\] | |
i.e.\ if $f(x_1, \dots, x_n) = 0$ then the gate does nothing, | |
otherwise the gate flips the $y$ bit at the bottom. | |
The slash with the $n$ indicates that the top of the input really consists | |
of $n$ qubits, not just the one qubit drawn, | |
and so the black box $U_f$ is a map on $n+1$ qubits. | |
The problem is to determine, | |
with as few calls to the black box $U_f$ as possible, | |
whether $f$ is balanced or constant. | |
\begin{ques} | |
Classically, show that in the worst case we may need | |
up to $2^{n-1}+1$ calls to the function $f$ to answer the question. | |
\end{ques} | |
So with only classical tools, it would take $O(2^n)$ queries to determine | |
whether $f$ is balanced or constant. | |
However, | |
\begin{theorem} | |
[Deutsch-Jozsa] | |
The Deutsch-Jozsa problem can be determined in a quantum circuit | |
with only a single call to the black box. | |
\end{theorem} | |
\begin{proof} | |
For concreteness, we do the case $n=1$ explicitly; | |
the general case is contained in \Cref{prob:deutsch_jozsa}. | |
We claim that the necessary circuit is | |
\[ | |
\Qcircuit @C=1em @R=0.7em { | |
\lstick{\ket0} & \gate{H} & \multigate{1}{U_f} & \gate{H} & \meter \qw \\ | |
\lstick{\ket1} & \gate{H} & \ghost{U_f} & \qw & \\ | |
} | |
\] | |
Here the $H$'s are Hadamard gates, | |
and the meter at the end of the rightmost wire | |
indicates that we make a measurement along the usual $\ket0$, $\ket1$ basis. | |
This is not a typo! Even though classically the top wire is just | |
a repeat of the input information, | |
we are about to see that it's the top we want to measure. | |
Note that after the two Hadamard operations, the state we get is | |
\begin{align*} | |
\ket{01} &\xmapsto{H^{\otimes 2}} | |
\left( \frac{1}{\sqrt2}(\ket0+\ket1) \right) | |
\otimes | |
\left( \frac{1}{\sqrt2}(\ket0-\ket1) \right) \\ | |
&= | |
\half \Big( \ket0\otimes\big(\ket0-\ket1\big) | |
\; + \; \ket1\otimes\big(\ket0-\ket1\big) \Big). | |
\end{align*} | |
So after applying $U_f$, we obtain | |
\[ | |
\half\Big( | |
\ket0\otimes\big(\ket{0+f(0)}-\ket{1+f(0)}\big) | |
\; + \; \ket1\otimes\big(\ket{0+f(1)}-\ket{1+f(1)}\big) | |
\Big) | |
\] | |
where the modulo $2$ has been left implicit. | |
Now, observe that the effect of going from | |
$\ket0-\ket1$ to $\ket{0+f(x)}-\ket{1+f(x)}$ is merely | |
to either keep the state the same (if $f(x)=0$) | |
or to negate it (if $f(x)=1$). | |
So we can simplify and factor to get | |
\[ | |
\half | |
\left( (-1)^{f(0)}\ket0 + (-1)^{f(1)}\ket1 \right) | |
\otimes | |
\left( \ket0-\ket1 \right). | |
\] | |
Thus, the picture so far is: | |
\[ | |
\Qcircuit @C=1em @R=0.7em { | |
\lstick{\ket0} & \gate{H} & \multigate{1}{U_f} & | |
\rstick{\frac{1}{\sqrt2}% | |
\Big((-1)^{f(0)}\ket0+(-1)^{f(1)}\ket1\Big)} \qw \\ | |
\lstick{\ket1} & \gate{H} & \ghost{U_f} & | |
\rstick{\frac{1}{\sqrt2}(\ket0-\ket1)} \qw | |
} | |
\] | |
In particular, the resulting state is not entangled, | |
and we can simply discard the last qubit (!). | |
Now observe: | |
\begin{itemize} | |
\ii If $f$ is constant, then the upper-most state is $\pm\xup$. | |
\ii If $f$ is balanced, then the upper-most state is $\pm\xdown$. | |
\end{itemize} | |
So simply doing a measurement along $\sigma_x$ will give us the answer. | |
Equivalently, perform another $H$ gate | |
(so that $H\xup = \ket0$, $H\xdown = \ket1$) | |
and measuring along $\sigma_z$ in the usual $\ket0$, $\ket1$ basis. | |
Thus for $n=1$ we only need a single call to the oracle. | |
\end{proof} | |
\section\problemhead | |
\begin{problem}[Fredkin gate] | |
The \vocab{Fredkin gate} (also called the controlled swap, or CSWAP gate) | |
is the three-bit gate with the following truth table: | |
\[ | |
\begin{array}{|rrr|rrr|} | |
\hline | |
\multicolumn{3}{|c|}{\text{In}} & \multicolumn{3}{|c|}{\text{Out}} \\ | |
\hline | |
0 & 0 & 0 & 0 & 0 & 0 \\ | |
0 & 0 & 1 & 0 & 0 & 1 \\ | |
0 & 1 & 0 & 0 & 1 & 0 \\ | |
0 & 1 & 1 & 0 & 1 & 1 \\ | |
1 & 0 & 0 & 1 & 0 & 0 \\ | |
1 & 0 & 1 & 1 & 1 & 0 \\ | |
1 & 1 & 0 & 1 & 0 & 1 \\ | |
1 & 1 & 1 & 1 & 1 & 1 \\\hline | |
\end{array} | |
\] | |
Thus the gate swaps the last two input bits whenever the first bit is $1$. | |
Show that this gate is also reversible and universal. | |
\begin{hint} | |
One way is to create CCNOT using a few Fredkin gates. | |
\end{hint} | |
\begin{sol} | |
To show the Fredkin gate is universal | |
it suffices to reversibly create a CCNOT gate with it. | |
We write the system | |
\begin{align*} | |
(z,\neg z,-) &= \opname{Fred}(z,1,0) \\ | |
(x,a,-) &= \opname{Fred}(x,1,0) \\ | |
(y,b,-) &= \opname{Fred}(y,a,0) \\ | |
(-,c,-) &= \opname{Fred}(b,0,1) \\ | |
(-,d,-) &= \opname{Fred}(c, z, \neg z). | |
\end{align*} | |
Direct computation shows that $d = z+xy\pmod 2$. | |
\end{sol} | |
\end{problem} | |
\begin{problem} | |
[Baby no-cloning theorem] | |
\label{prob:baby_no_clone} | |
Show that there is no unitary map $U$ on two qubits | |
which sends $U(\ket\psi \otimes \ket0) = \ket\psi \otimes \ket\psi$ | |
for any qubit $\ket\psi$, i.e.\ | |
the following circuit diagram is impossible. | |
\[ | |
\Qcircuit @C=0.8em @R=.7em { | |
\lstick{\ket\psi} & \multigate{1}{U} & \rstick{\ket\psi} \qw \\ | |
\lstick{\ket0} & \ghost{U} & \rstick{\ket\psi} \qw | |
} | |
\] | |
\begin{hint} | |
Plug in $\ket\psi=\ket0$, $\ket\psi=\ket1$, $\ket\psi = \xup$ | |
and derive a contradiction. | |
\end{hint} | |
\end{problem} | |
\begin{problem}[Deutsch-Jozsa] | |
\label{prob:deutsch_jozsa} | |
Given the black box $U_f$ described in the Deutsch-Jozsa algorithm, | |
consider the following circuit. | |
\[ | |
\Qcircuit @C=1em @R=0.7em { | |
\lstick{\ket{0\dots0}} & /^n \qw & \gate{H^{\otimes n}} & \multigate{1}{U_f} | |
& \gate{H^{\otimes n}} & \meter \qw \\ | |
\lstick{\ket1} & \qw & \gate{H} & \ghost{U_f} & \qw & \\ | |
} | |
\] | |
That is, take $n$ copies of $\ket 0$, apply the Hadamard rotation to all of them, | |
apply $U_f$, reverse the Hadamard to all $n$ input bits | |
(again discarding the last bit), then measure all $n$ bits | |
in the $\ket0$/$\ket1$ basis (as in \Cref{ex:simult_measurement}). | |
Show that the probability of measuring $\ket{0\dots0}$ | |
is $1$ if $f$ is constant and $0$ if $f$ is balanced. | |
\begin{hint} | |
First show that the box sends | |
$\ket{x_1} \otimes \dots \otimes \ket{x_m} \otimes \xdown$ | |
to $(-1)^{f(x_1, \dots, x_m)} | |
(\ket{x_1} \otimes \dots \otimes \ket{x_m} \otimes \xdown)$. | |
\end{hint} | |
\begin{sol} | |
Put $\xdown = \frac{1}{\sqrt2} (\ket0-\ket1)$. | |
Then we have that $U_f$ sends | |
\[ | |
\ket{x_1} \dots \ket{x_m} \ket 0 | |
- \ket{x_1} \dots \ket{x_m} \ket 1 | |
\xmapsto{U_f} | |
\pm \ket{x_1} \dots \ket{x_m} \ket 0 | |
\mp \ket{x_1} \dots \ket{x_m} \ket 1 | |
\] | |
the sign being $+$, $-$ exactly when $f(x_1, \dots, x_m) = 1$. | |
Now, upon inputting $\ket0 \dots \ket0 \ket1$, we find that $H^{\otimes m+1}$ maps it to | |
\[ 2^{-n/2} \sum_{x_1, \dots, x_n} \ket{x_1} \dots \ket{x_n} \xdown. \] | |
Then the image under $U_f$ is | |
\[ 2^{-n/2} \sum_{x_1, \dots, x_n} (-1)^{f(x_1, \dots, x_n)} \ket{x_1} \dots \ket{x_n} \xdown. \] | |
We now discard the last qubit, leaving us with | |
\[ 2^{-n/2} \sum_{x_1, \dots, x_n} (-1)^{f(x_1, \dots, x_n)} \ket{x_1} \dots \ket{x_n}. \] | |
Applying $H^{\otimes m}$ to this, we get | |
\[ 2^{-n/2} \sum_{x_1, \dots, x_n} (-1)^{f(x_1, \dots, x_n)} | |
\cdot | |
\left( | |
2^{-n/2} | |
\sum_{y_1, \dots, y_n} | |
(-1)^{x_1 y_1 + \dots + x_n y_n} | |
\ket{y_1} \ket{y_2} \dots \ket{y_n} | |
\right) | |
\] | |
since $H\ket0 = \frac{1}{\sqrt2}(\ket0+\ket1)$ | |
while $H\ket1 = \frac{1}{\sqrt2}(\ket0-\ket1)$, | |
so minus signs arise exactly if $x_i = 0$ and $y_i = 0$ simultaneously, | |
hence the term $(-1)^{x_1 y_1 + \dots + x_n y_n}$. | |
Swapping the order of summation, we get | |
\[ | |
2^{-n} | |
\sum_{y_1, \dots, y_n} | |
C(y_1, \dots, y_n) | |
\ket{y_1} \ket{y_2} \dots \ket{y_n} | |
\] | |
where $C_{y_1, \dots, y_n} | |
= \sum_{x_1, \dots, x_n} (-1)^{f(x_1, \dots, x_n) | |
+x_1 y_1 + \dots + x_n y_n}$. | |
Now, we finally consider two cases. | |
\begin{itemize} | |
\ii If $f$ is the constant function, then we find that | |
\[ | |
C(y_1, \dots, y_n) = | |
\begin{cases} | |
\pm 1 & y_1 = \dots = y_n = 0 \\ | |
0 & \text{otherwise}. | |
\end{cases} | |
\] | |
To see this, note that the result is clear for $y_1 = \dots = y_n = 0$; | |
otherwise, if WLOG $y_1 = 1$, then the terms for $x_1 = 0$ exactly cancel | |
the terms for $x_1 = 0$, pair by pair. | |
Thus in this state, the measurements all result in $\ket0 \dots \ket0$. | |
\ii On the other hand if $f$ is balanced, we derive that | |
\[ C(0, \dots, 0) = 0. \] | |
Thus \emph{no} measurements result in $\ket 0 \dots \ket 0$. | |
\end{itemize} | |
In this way, we can tell whether $f$ is balanced or not. | |
\end{sol} | |
\end{problem} | |
\begin{dproblem}[Barenco et al, 1995; arXiv:quant-ph/9503016v1] | |
Let | |
\[ | |
P = \begin{bmatrix} 1 & 0 \\ 0& i \end{bmatrix} | |
\qquad | |
Q = \frac{1}{\sqrt2}\begin{bmatrix} 1 & -i \\ -i & 1 \end{bmatrix} | |
\] | |
Verify that the quantum Toffoli gate can be implemented | |
using just controlled rotations via the circuit | |
\[ | |
\Qcircuit @R=1em @C=0.7em { | |
\lstick{\ket{x_1}} & \qw & \ctrl{2} & \ctrl{1} & \ctrl{1} & \qw & \ctrl{1} & \qw \\ | |
\lstick{\ket{x_2}} & \ctrl{1} & \qw & \gate{P} & \targ & \ctrl{1} & \targ & \qw \\ | |
\lstick{\ket{x_3}} & \gate{Q} & \gate{Q} & \qw & \qw & \gate{Q^\dagger} & \qw & \qw | |
} | |
\] | |
This was a big surprise to researchers when discovered, | |
because classical reversible logic requires three-bit gates (e.g. Toffoli, Fredkind). | |
\begin{hint} | |
This is direct computation. | |
\end{hint} | |
\end{dproblem} | |