Quantum Information
1.Introduction
2.Quantum mechanics essentials
3.Measurement and uncertainty
4.Qubits and the Bloch sphere
4.1.Qubits
4.2.Inside the Bloch sphere
4.3.Time-evolution of a qubit
4.4.Pauli Matrices
4.5.Problems
5.Bipartite systems
6.Entanglement applications
7.Information theory
8.Changelog
9.Bibliography
\(\newcommand{\Hint}[1]{{\flushleft{\bf Hint:}} #1}\)
\(\newcommand{\Tr}{\mathrm{Tr}}\)
\(\renewcommand{\d}{\mathrm{d}}\)
\(\renewcommand{\vec}[1]{\mathbf{#1}}\)
\(\newcommand{\cvec}[1]{\left( \begin{array}{c} #1 \end{array} \right)}\)
\(\newcommand{\ket}[1]{\left\vert #1 \right\rangle}\)
\(\newcommand{\bra}[1]{\left\langle #1 \right\vert}\)
\(\newcommand{\ipop}[3]{\left\langle #1 \left\vert #2 \vphantom{#1 #3} \right\vert #3 \right\rangle}\)
\(\newcommand{\ip}[2]{\left\langle #1 \left\vert #2 \vphantom{#1} \right\rangle \right.}\)
\(\newcommand{\ev}[1]{\left\langle #1 \right\rangle}\)
\(\newcommand{\com}[2]{\left[ \hat{#1}, \hat{#2} \right]}\)
\(\newcommand{\acom}[2]{\left\{ \hat{#1}, \hat{#2} \right\}}\)
\(\newcommand{\norm}[1]{\vert\vert \ket{#1} \vert \vert}\)
4.Qubits and the Bloch sphere
4.1.Qubits
7
In this chapter we want to understand the smallest dimensional (yet extremely interesting) quantum system. Because of the equivalence relation \(\ket{\psi} \sim c \ket{\psi}\), a one-dimensional Hilbert space is trivial since it only describes a single state. Therefore the smallest non-trivial system has dimension two, and in QI we refer to such a system as a qubit. A standard orthonormal basis is labelled \(\{ \ket{0}, \ket{1} \}\) (sometimes you will also find it written as \(\{\ket{\uparrow},\ket{\downarrow}\}\) denoting the two states of a spin 1/2 particle with spin “up” or “down”) so we see the analogy with a classical bit which can take either value \(0\) or \(1\) hence we usually call the basis \(\{ \ket{0}, \ket{1} \}\) the computational basis. The difference for the qubit is that the state can be a linear combination of \(\ket{0}\) and \(\ket{1}\). The qubit is important since many general features can be understand in this simple Hilbert space. Also, larger systems can often be described a being built from several qubits – this is used in quantum information and especially in quantum computation.
So let us try and characterise the most general pure state of the qubit Hilbert space \(\mathcal{H}_q =\mbox{span} \{ \ket{0}, \ket{1} \}\). Now if we have a qubit system the most general pure state can be written as a linear combination of the two basis vectors
\[ \ket{\psi} = a \ket{0}+b\ket{1}\,, \]
with \(a,b\in\mathbb{C}\) and \(|a|^2+|b|^2=1\) so that the state is already normalised \(\ip{\psi}{\psi}=1\).
The most general qubit state is labelled by two complex numbers, subject to a norm constraint and the fact that the overall phase is irrelevant. This leaves two real parameters, which parametrise a sphere.
Up to an irrelevant overall phase we can assume that \(a\in\mathbb{R}\), hence any normalised pure state can be written in the form
\[ \ket{\psi} = \cos\left(\frac{\theta}{2}\right)\ket{0} + e^{i\phi}\sin\left(\frac{\theta}{2}\right)\ket{1} \]
where \(0 \le \theta \le \pi\) and \(0 \le \phi \lt{} 2\pi\). Note that the normalisation condition \(|a|^2+|b|^2=1\) now becomes \(\cos\left(\frac{\theta}{2}\right)^2+ \vert e^{i\phi}\sin\left(\frac{\theta}{2}\right)\vert^2=1\).
This new coordinates \((\theta,\phi)\) are just a convenient parametrisation of \(|a|^2+|b|^2=1\) with \(a\in\mathbb{R}\) in terms of angles which correspond to the angles in spherical polar coordinates. Indeed the equation \(|a|^2+|b|^2=1\) with \(a\in\mathbb{R}\) can be rewritten as \(a^2+ (\mbox{Re}\,b)^2+(\mbox{Im}\, b)^2 =1\) which is indeed the equation describing a 2-dimensional sphere, what we call \(S^2\). Therefore we have a one-to-one mapping between pure qubits states and points on a sphere (say of radius \(1\).) This is called the Bloch Sphere.
Now any point on the Bloch sphere can also be labelled by its position vector, called the Bloch vector\(\vec{r}=\left(\begin{matrix}x\\y\\z\end{matrix}\right)\) which has \(|\vec{r}| =\sqrt{x^2+y^2+z^2}= 1\). The dictionary between Cartesian coordinates and polar coordinates is very simple
\[ x = \sin(\theta) \cos(\phi)\,,\quad y=\sin(\theta) \sin(\phi)\,,\quad z= \cos(\theta)\,, \]
\(\theta\) is the polar angle measured from the \(z\)-axis, while \(\phi\) is the azimuthal angle measured around the \(z\)-axis as in Figure bloch.
Figure 4.1: Graphic representation of the Bloch sphere. The states \(\ket{0},\ket{1}\) are on the \(z\)-axis, the state \(\ket{\pm} = (\ket{0}\pm\ket{1})/\sqrt{2}\) are on the \(x\)-axis, and the states \(\ket{L}= (\ket{0}+i\ket{1})/\sqrt{2}\,,\,\ket{R}= (\ket{0}-i\ket{1})/\sqrt{2}\) are on the \(y\)-axis.
There are six special states in the qubit system, which are eigenvectors of three special operators (to be discussed later).
There are six special states on the Bloch sphere whose cartesian coordinates are simply \(\left(\begin{matrix}\pm1\\0\\0\end{matrix}\right),\left(\begin{matrix}0\\\pm1\\0\end{matrix}\right),\left(\begin{matrix}0\\0\\ \pm1\end{matrix}\right)\) given by
\begin{align*} \vec{r}&= \left(\begin{matrix}1\\0\\0\end{matrix}\right)\,,\qquad (\theta,\phi) = (\pi/2,0)\,,\\ &\qquad\rightarrow\quad\ket{+}=\frac{1}{\sqrt{2}} (\ket{0}+ \ket{1})\to\frac{1}{\sqrt{2}} \left(\begin{matrix}1 \\ 1\end{matrix}\right)\,,\\[1ex] \vec{r}&= \left(\begin{matrix}-1\\0\\0\end{matrix}\right)\,,\qquad (\theta,\phi) = (\pi/2,\pi)\,,\\ &\qquad\rightarrow\quad\ket{-}=\frac{1}{\sqrt{2}} (\ket{0}- \ket{1})\to\frac{1}{\sqrt{2}} \left(\begin{matrix}1 \\ -1\end{matrix}\right)\,,\\[1ex] \vec{r}&= \left(\begin{matrix}0\\1\\0\end{matrix}\right)\,,\qquad (\theta,\phi) = (\pi/2,\pi/2)\,,\\ &\qquad\rightarrow\quad \ket{L}=\frac{1}{\sqrt{2}} (\ket{0}+i \ket{1})\to\frac{1}{\sqrt{2}} \left(\begin{matrix}1 \\ i\end{matrix}\right)\,,\\[1ex] \vec{r}&= \left(\begin{matrix}0\\-1\\0\end{matrix}\right)\,,\qquad (\theta,\phi) = (\pi/2,3\pi/2)\,,\\ &\qquad\rightarrow\quad\ket{R}=\frac{1}{\sqrt{2}} (\ket{0}- i\ket{1})\to\frac{1}{\sqrt{2}} \left(\begin{matrix}1 \\ -i \end{matrix}\right)\,,\\[1ex] \vec{r}& \left(\begin{matrix}0\\0\\1\end{matrix}\right)\,,\qquad (\theta,\phi) = (0, \cdot)\,,\\ &\qquad\rightarrow\quad\ket{0}\to\left(\begin{matrix}1 \\ 0\end{matrix}\right) \,,\\[1ex] \vec{r}&= \left(\begin{matrix}0\\0\\-1\end{matrix}\right)\,,\qquad (\theta,\phi) = (\pi ,\cdot)\,,\\ &\qquad\rightarrow\quad\ket{1}\to\left(\begin{matrix}0 \\1\end{matrix}\right)\,. \end{align*}
These states are special as they lie at the intersection of the Bloch sphere with the cardinal axis and we will see later on they will each have well defined measurements for three special observables for the qubit system.
Show that besides the computational basis \(\{\ket{0},\ket{1}\}\), the qubit Hilbert space \(\mathcal{H}_q\) can be described in terms of the orthonormal basis \(\{\ket{+},\ket{-}\}\) or by the orthonormal basis \(\{\ket{L},\ket{R}\}\).
4.2.Inside the Bloch sphere
8
We want to rewrite now the density matrix for a general pure state written using polar angles to cartesian coordinates. To this end we start with a generic pure state \( \ket{\psi} = \cos\left(\frac{\theta}{2}\right)\ket{0} + e^{i\phi}\sin\left(\frac{\theta}{2}\right)\ket{1} \) and we first rewrite it using as above the standard basis representation for the two basis ket vector \(\ket{0}\to \left(\begin{matrix}1 \\ 0 \end{matrix}\right)\,,\,\ket{1}\to \left(\begin{matrix}0 \\ 1 \end{matrix}\right)\) hence we have
\begin{align*} &\ket{\psi}\to\left(\begin{matrix}\cos\left(\frac{\theta}{2}\right) \\ e^{i\phi}\sin\left(\frac{\theta}{2}\right) \end{matrix}\right) \,,\\ \,\hat{\rho} &=\ket{\psi}\bra{\psi} \to \rho = \left(\begin{matrix}\cos\left(\frac{\theta}{2}\right) \\ e^{i\phi}\sin\left(\frac{\theta}{2}\right) \end{matrix}\right) \left(\begin{matrix}\cos\left(\frac{\theta}{2}\right) & e^{-i\phi}\sin\left(\frac{\theta}{2}\right)\end{matrix}\right)\\ &\rho = \frac{1}{2} \left(\begin{matrix}1+\cos(\theta) & e^{-i\phi}\sin(\theta)\\ e^{+i\phi}\sin(\theta) &1-\cos(\theta) \end{matrix}\right)=\frac{1}{2} \left(\begin{matrix}1+z& x-i y \\ x+ i y &1-z\end{matrix}\right)\,, \end{align*}
where in the last line we used some simple trigonometric identities and finally the dictionary spherical \(\leftrightarrow\) cartesian coords.
It can be seen quite easily now that the density matrix can be written in terms of the Bloch vector \(\vec{r}=\left(\begin{matrix}x\\y\\z\end{matrix}\right)\) as \(\rho = \frac{1}{2}\left( \mathbb{I}_2+ \vec{r}\cdot\vec{\sigma} \right)\), where we introduce the three matrices \(\sigma_i\) called the Pauli \(\sigma\)-matrices
\[ \sigma_1 = \left(\begin{matrix}0&1\\ 1 & 0 \end{matrix}\right)\,,\quad \sigma_2 = \left(\begin{matrix}0&-i\\ i & 0 \end{matrix}\right)\,,\quad \sigma_3 = \left(\begin{matrix}1&0\\ 0 & -1 \end{matrix}\right)\,, \]
A state with given Bloch vector can be mapped to a density matrix using the Pauli matrices.
so \(\vec{r}\cdot\vec{\sigma} = r_1 \sigma_1 +r_2 \sigma_2 +r_3 \sigma_3\).
As the density matrix is linear in the Bloch vector, mixed states also have Bloch vector given in terms of a linear combination of the Bloch vectors of the states in the ensemble. Since the coefficients are probabilities, it is easy to see that the Bloch vector for a mixed state must be the position vector of a point inside the Bloch sphere.
Let us check this in detail. Suppose we have the ensemble \(\{(p_i,\rho_i)\}\) for some probabilities \(p_i\) and pure states \(\rho_i\) defined by vectors \(\vec{r}_i\) on the Bloch sphere, i.e. \(| \vec{r}_i|^2=1\). The density matrix for this mixed state is given by
\begin{equation} \begin{aligned} \rho &= \sum_i p_i \rho_i = \sum_i p_i \frac{\left( \mathbb{I}_2 + \vec{r}_i \cdot\vec{\sigma} \right)}{2} \\ &= \frac{1}{2}\left( \mathbb{I}_2 + \vec{r}\cdot\vec{\sigma} \right)\,, \end{aligned} \end{equation}
where we defined the vector \(\vec{r} = \sum_i p_i \vec{r}_i\). As said above, since the density matrix is linear in the Bloch vector, mixed states as well can be represented in terms of a Bloch vector \(\vec{r}\). Let us check now where this vector lies:
\begin{equation} \begin{aligned} | \vec{r} |^2 &= | \sum_i p_i \vec{r}_i |^2 = \sum_{ij} p_i p_j \,\vec{r}_i \cdot \vec{r}_j\\ &\leq \sum_{ij} p_i p_j |\vec{r}_i| | \vec{r}_j |\leq \sum_{ij} p_i p_j \leq 1\,, \end{aligned} \end{equation}
where we used the real Cauchy Schwartz inequality. Note that equality holds if and only if all \(\vec{r}_i\) are collinear hence all the density matrices \(\rho_i\) of the ensemble are given by the same density matrix, hence we do not really have a mixed state but rather a pure one.
For genuine mixed states the inequality is strict and so mixed states are defined by point inside the Bloch sphere, i.e. \(| \vec{r} | \lt{}1\).
The Bloch sphere gives us a nice geometrical interpretation for the full set of states of the qubit system: points on the sphere define pure states, points inside the sphere define mixed states.
Remember the previous test to distinguish whether a density matrix corresponds to a pure or a mixed state by considering whether \(\Tr(\rho^2)=1\) or \(\lt{}1\) respectively. The condition \(\Tr(\rho^2)\lt{}1\) for \(\rho\) to correspond to a mixed state has now for a qubit a geometric interpretation:
\[ \Tr(\rho^2) = \frac{1}{2}\left( 1 + |\vec{r}|^2 \right) \le 1\,, \]
mixed states are inside the Bloch sphere \(|\vec{r}|\lt{}1\) , pure state are on the Bloch sphere \(|\vec{r}|=1\).
4.3.Time-evolution of a qubit
Unitary transformations of a qubit are represented as rotations of the Bloch sphere about the origin. This illustrates that unitary transformations cannot transform pure states to mixed states, but also we see that not all mixed states are related in this way. Indeed \(\Tr(\rho^2) = (1 + |\vec{r}|^2)/2\) is invariant under unitary transformations, and is a measure of how mixed a state is, with \(\Tr(\rho^2) = 1\) for pure states to \(\Tr(\rho^2) = 1/2\) for the “most mixed” state corresponding to the origin, i.e. \(\rho = \mathbb{I}_2/2\). Of course, measurements are not unitary transformations and they act as projectors hence they can transform any state to a pure state.
Show that the quantity \(\Tr(\hat{\rho}^2) \) is invariant under time evolution.
Solution: We simply need to remember that a density matrix \(\hat{\rho}\) transforms under time evolution has \( \hat{\rho} \to \hat{U} \hat{\rho}\hat{U}^\dagger\) hence
\[\hat{\rho}^2 \to \hat{U} \hat{\rho}\hat{U}^\dagger \hat{U} \hat{\rho}\hat{U}^\dagger =\hat{U} \hat{\rho}^2\hat{U}^\dagger\]
using the fact that time evolution is a unitary operation \(\hat{U}^\dagger \hat{U} =\hat{I}\). We then conclude that \( \Tr(\hat{U} \hat{\rho}^2\hat{U}^\dagger)= \Tr(\hat{\rho}^2) \) using cyclicity of the trace. Since \(\Tr(\hat{\rho}^2) \) measures how mixed a state is (we will quantify better later on) it is then clear that time evolution cannot possibly change that.
Let \(\vec{r}_1\) and \(\vec{r}_2\) denote two distinct points on the Bloch sphere, i.e. \(|\vec{r}_1|= |\vec{r}_1|=1\) and \(\vec{r}_1\neq \vec{r}_2\), and consider the ensemble \(\{ (p,\vec{r}_1),(1-p,\vec{r}_2)\}\) with \(0 \leq p \leq 1\). The density matrix corresponding to this ensemble is given by
\[ \rho = \rho_1 +\rho_2 = \frac{1}{2}(\mathbb{I}_2 + \vec{r}\cdot \sigma)\,, \]
with \(\vec{r} = p \,\vec{r}_1+(1-p)\vec{r}_2\). Note that geometrically the Bloch vector \(\vec{r}\) lies on the line between the points \(\vec{r}_1\) and \(\vec{r}_2\) and since \(0 \leq p \leq 1\) the point \(\vec{r}\) on this line will always fall within the Bloch sphere.
The state \(\rho\) will be pure if and only if \(p=1\), in which case \(\rho=\rho_1\), or \(p=0\), in which case \(\rho=\rho_2\). Note in particular that mixing can never produce a state farther from the origin then the farthest initial state. Furthermore once we have chosen a mixed state, i.e. a point inside the Bloch sphere, we can find an infinite number of ways to write it as an ensemble of two pure states! We just need to consider any line passing through this point which will intersect the Bloch sphere in two points corresponding to the two pure states of which this mixed state is an ensemble of.
In particular the “most” mixed of the qubit states is \(\rho = \mathbb{I}_2 / 2\) which correspond of an ensemble of any two antipodal points, for example the North and South poles \(\ket{0}, \ket{1}\) states (but any other two antipodal points on the sphere will produce the same), each one of them with probability \(50\%\).
For a qubit system it is simple to define a “distance” between states given by the geometric distance of the relative positions using the Bloch sphere, i.e. \(| \vec{r}_1 - \vec{r}_2|\) if the two states are described by the Bloch vectors \(\vec{r}_1\,,\,\vec{r}_2\). This turns out to be equal to what is called the trace distance between the two states
\[ D(\hat{\rho}_1,\hat{\rho}_2) = \frac{1}{2}\Tr \vert \hat{\rho}_1 -\hat{\rho}_2 \vert\,, \]
where we need to give the definition of the operator \(| \hat{A} |\).
The operator \(| \hat{A} |\) is defined to be the positive operator
\[ \vert \hat{A} \vert = \sqrt{ \hat{A}^\dagger \hat{A} }\,. \]
Note (check) that the operator \(\hat{A}^\dagger \hat{A}\) is both hermitian and positive hence its square root is well defined. For this, you just need to chose a basis that diagonalises \(\hat{A}^\dagger \hat{A}\) with non-negative eigenvalues (because the operator is hermitian and positive); in this basis the operator \(\sqrt{ \hat{A}^\dagger \hat{A} }\) will then be diagonal with eigenvalues given by the square root of the eigenvalues of \(\hat{A}^\dagger \hat{A}\).
In practical terms if the operator \( \hat{A} \) is hermitian with eigenvalues \(a_i\) then \( \hat{A}^\dagger \hat{A} =\hat{A}^2 \) and we have that
\[ \Tr \vert \hat{A} \vert = \sum_i \vert a_i\vert\,. \]
With this definition for \(\Tr \vert \hat{A} \vert\) we can then easily compute the trace distance between two qubit states
\begin{equation} \begin{aligned} D(\hat{\rho}_1,\hat{\rho}_2) &= \frac{1}{2}\Tr \vert \hat{\rho}_1 -\hat{\rho}_2 \vert =\frac{1}{4} \Tr \vert (\vec{r}_1 -\vec{r}_2)\cdot \sigma \vert \\ & =\frac{1}{2}\vert \vec{r}_1 -\vec{r}_2\vert\,, \end{aligned} \end{equation}
and indeed as anticipated the trace distance for qubit states correspond precisely to the geometric distance in the Bloch sphere representation.
The “trace distance” between two states described by a density matrix corresponds to the geometric distance between the two states in the Bloch sphere representation.
We can rewrite the trace distance between two states as
\begin{equation} \begin{aligned} D(\hat{\rho}_1,\hat{\rho}_2) &=\frac{1}{2}\Tr \vert \hat{\rho}_1 -\hat{\rho}_2 \vert \\ &= \frac{1}{2} \Tr \sqrt{ (\hat{\rho}_1 -\hat{\rho}_2)^\dagger (\hat{\rho}_1 -\hat{\rho}_2) } \\ &= \frac{1}{2} \Tr \sqrt{ (\hat{\rho}_1 -\hat{\rho}_2)^2 } = \frac{1}{2}\sum_i \vert \lambda_i\vert\,, \end{aligned} \end{equation}
where \(\lambda_i\) are the eigenvalues of the hermitian but not necessarily positive operator \(\hat{\rho}_1 -\hat{\rho}_2\). This defines a metric on the space of density matrices, i.e. something with the following properties
Note that there are many notions of “distance” when discussing quantum mechanical states. The trace distance is a possible meaningful notion of distance between states in that it tells us how distinguishable with measurements two states are, i.e. the closer they are the less “distinguishable” they are by simply measuring observables. Another such notion is what is called fidelity although we will not cover this.
Compute the trace distance between the qubit states
\begin{align*} \hat{\rho}_1 &= \frac{1}{4} \ket{0}\bra{0} + \frac{3}{4} \ket{1}\bra{1} \rightarrow \rho_1 =\left( \begin{matrix} \frac{1}{4} & 0\\ 0& \frac{3}{4}\end{matrix}\right)\,,\\ \hat{\rho}_2 &= \frac{2}{3} \ket{0}\bra{0} + \frac{1}{3} \ket{1}\bra{1} \rightarrow\rho_2=\left( \begin{matrix} \frac{2}{3} & 0\\ 0& \frac{1}{3}\end{matrix}\right)\,. \end{align*}
If we compute the matrix \(\rho_1-\rho_2\) we obtain
\begin{equation} \begin{aligned} \rho_1-\rho_2 &=\left( \begin{matrix} \frac{-5}{12} & 0\\ 0& \frac{5}{12}\end{matrix}\right)\\ \Rightarrow D(\hat{\rho}_1,\hat{\rho}_2)&=\frac{1}{2}\Tr \vert \rho_1 -\rho_2 \vert \\ &= \frac{1}{2} \left( \left\vert \frac{-5}{12} \right\vert +\left\vert \frac{5}{12}\right\vert \right)= \frac{5}{12}\,. \end{aligned} \end{equation}
We could have also computed the two Bloch vectors associated to the two density matrices \(\vec{r}_1 = (0,0,-1/2)^T\) and \(\vec{r}_2 = (0,0,1/3)^T\) and in fact we have
\[ D(\hat{\rho}_1,\hat{\rho}_2) = \frac{1}{2} \vert \vec{r}_1 - \vec{r}_2 \vert = \frac{1}{2} \left\vert \left( \begin{matrix}0 \\ 0 \\ -\frac{5}{6} \end{matrix}\right) \right\vert = \frac{5}{12} \,. \]
4.4.Pauli Matrices
We will summarise here some key properties of the so-called Pauli matrices, which were introduced above when we constructed the density matrix corresponding to a generic qubit state. The following can be checked easily.
\begin{align*} \sigma_1^\dagger&= \sigma_1\,,\qquad\sigma_2^\dagger =\sigma_2\,,\qquad\sigma_3^\dagger=\sigma_3\,,\\ \mbox{Tr}(\sigma_1)&=\mbox{Tr}(\sigma_2)=\mbox{Tr}(\sigma_3)=0\,,\\ [\sigma_i,\sigma_j] & =\sigma_i\sigma_j-\sigma_j\sigma_i= 2 i\, \epsilon_{ijk}\, \sigma_k\,,\\ \{\sigma_i,\sigma_j\} & = \sigma_i\sigma_j+\sigma_j\sigma_i = 2\delta_{ij}\mathbb{I}_2\,,\\ \sigma_i \sigma_j &= \delta_{ij} \mathbb{I}_2 + i\,\epsilon_{ijk}\sigma_k\,, \end{align*}
where \(\delta_{ij}\) is the Kronecker delta and \(\epsilon_{ijk}\) the Levi-Civita tensor.
So the Pauli matrices are Hermitian, traceless matrices, they actually form a basis for the vector space of \(2\times 2\) hermitian, traceless matrices (Ex. check that this is a vector space over the real numbers).
If we define the operators
\begin{align*} X& = \frac{1}{2}(\mathbb{I}_2- \sigma_1) = \frac{1}{2}\left(\begin{matrix}1& -1\\ -1 & 1\end{matrix}\right)\,,\\ Y &= \frac{1}{2}(\mathbb{I}_2- \sigma_2)= \frac{1}{2}\left(\begin{matrix}1& i\\ -i & 1\end{matrix}\right)\,,\\ Z &= \frac{1}{2}(\mathbb{I}_2- \sigma_3)= \frac{1}{2}\left(\begin{matrix}0& 0\\ 0 & 2\end{matrix}\right)\,,\qquad \end{align*}
we can easily see that the six “special” states defined above are precisely the eigenvectors of these three operators with eigenvalues either \(0\) or \(1\), i.e.
\begin{align*} X \ket{+} = 0 \ket{+}\,,\qquad &X \ket{-} = 1 \ket{-}\,,\\ Y \ket{L} = 0 \ket{L}\,,\qquad&Y \ket{R} = 1 \ket{R}\,,\\ Z \ket{0} = 0 \ket{0}\,,\qquad&Z \ket{1} = 1 \ket{1}\,. \end{align*}
Finally another important property of the Pauli matrices is that their exponential are unitary matrices
\begin{align*} e^{ i \alpha \sigma_1} &= \left(\begin{matrix}\cos( \alpha) & i \sin(\alpha)\\ i \sin(\alpha) & \cos(\alpha) \end{matrix}\right)= \cos(\alpha) \mathbb{I}_2 +i \sin(\alpha) \sigma_1\,,\\ e^{ i \alpha \sigma_2} &= \left(\begin{matrix}\cos( \alpha) & \sin(\alpha)\\ - \sin(\alpha) & \cos(\alpha) \end{matrix}\right)= \cos(\alpha) \mathbb{I}_2 +i \sin(\alpha) \sigma_2\,,\\ e^{ i \alpha \sigma_3} &= \left(\begin{matrix}e^{i \alpha} & 0 \\0 &e^{- i \alpha} \end{matrix}\right)= \cos(\alpha) \mathbb{I}_2 +i \sin(\alpha) \sigma_3\,, \end{align*}
where \(\alpha \in \mathbb{R}\).
Compute these exponentials and check that these are indeed unitary matrices.
More in general we can consider the unitary transformation
\[ U_\alpha(\vec{n}) = e^{i \alpha \vec{n}\cdot \sigma} = \cos(\alpha) \mathbb{I}_2 +i \sin(\alpha) \vec{n}\cdot \sigma\,, \]
where again \(\alpha \in \mathbb{R}\) and \(\vec{n}\in\mathbb{R}^3\) is a \(3\)-dimensional unit vector, i.e. \(| \vec{n}|^2=1\).
This is a unitary transformation so it is a perfectly valid time evolution operator. If we now act on the density matrix \(\rho = \frac{1}{2}\left( \mathbb{I}_2 + \vec{r}\cdot \sigma \right)\) of a state defined by the vector \(\vec{r}\) on the Bloch sphere we can see that
\[ U_\alpha(\vec{n}) \rho \,U_\alpha(\vec{n})^\dagger = \frac{1}{2}\left[ \mathbb{I}_2 + ( R_\alpha(\vec{n}) \vec{r}) \cdot \sigma\right]\,, \]
where \(R_\alpha(\vec{n})\) is the \(3\times 3\) orthogonal matrix corresponding to a rotation of an angle \(2\alpha\) around the axis defined by the unit vector \(\vec{n}\) acting on the three dimensional vector \(\vec{r}\). This means that if we prepare our qubit say in the state \(\ket{0}\) with Bloch vector \(\vec{r}=(0,0,1)^T\) by performing the unitary time evolution \(U_\alpha(\vec{n})\) we can transform this state in any other state on the Bloch sphere with Bloch vector \(\vec{r}'\) by just choosing some specific \(\alpha,\vec{n}\) such that the corresponding rotation of the Bloch vector \(R_\alpha(\vec{n})(0,0,1)^T =\vec{r}'\). This will be extremely useful later on when discussing applications of entanglement.
4.5.Problems
  1. An arbitrary qubit density matrix (for a mixed or pure state) can be written
    \[ \hat{\rho} = \frac{1}{2} \left( I + \vec{r} \cdot \vec{\sigma} \right) \]
    where \(I\) is the \(2 \times 2\) identity matrix, \(\sigma_i\) are the three Pauli sigma-matrices and the real ‘Bloch’ vector \(\vec{r}\) has length \(|\vec{r}| \le 1\).
    1. Suppose \(\hat{\rho}_1\) and \(\hat{\rho}_2\) are density matrices for the pure states \(\ket{\psi_1}\) and \(\ket{\psi_2}\) respectively. What condition on \(\hat{\rho}_1\hat{\rho}_2\) is equivalent to the statement that \(\ket{\psi_1}\) is orthogonal to \(\ket{\psi_2}\)?
    2. Express the condition for \(\hat{\rho}_1\hat{\rho}_2\) in part (a) (and now allowing pure or mixed states) in terms of conditions on the Bloch vectors \(\vec{r}_1\) and \(\vec{r}_2\) defining the two density matrices. How are two orthogonal qubit states represented on the Bloch sphere?
    Solution:
    1. Note that if \(\ket{\psi_1}\) is orthogonal to \(\ket{\psi_2}\) then \(\ip{\psi_1}{\psi_2} = 0\) so
      \[ \hat{\rho}_1 \hat{\rho}_2 = \ket{\psi_1} \bra{\psi_1} \ket{\psi_2} \bra{\psi_2} = (\ip{\psi_1}{\psi_2}) \ket{\psi_1} \bra{\psi_2} = 0 . \]
      Conversely if \(\hat{\rho}_1 \hat{\rho}_2 = 0\) then
      \[ 0 = \Tr \left( \hat{\rho}_1 \hat{\rho}_2 \right) = (\ip{\psi_1}{\psi_2}) \Tr \left( \ket{\psi_1} \bra{\psi_2} \right) = (\ip{\psi_1}{\psi_2}) (\ip{\psi_2}{\psi_1}) = \left\vert \ip{\psi_1}{\psi_2} \right\vert^2 . \]
      However, this means that \(\ip{\psi_1}{\psi_2} = 0\).
      Therefore we have the following:
      \[ \hat{\rho}_1 \hat{\rho}_2 = 0 \iff \ip{\psi_1}{\psi_2} = 0 \iff \ip{\psi_2}{\psi_1} = 0 \iff \hat{\rho}_2 \hat{\rho}_1 = 0 . \]
    2. Using the relation between the density matrix and the Bloch vector we have
      \[ 4\hat{\rho}_1 \hat{\rho}_2 = \left( I + \vec{r}_1 \cdot \vec{\sigma} \right) \left( I + \vec{r}_2 \cdot \vec{\sigma} \right) = I + (\vec{r}_1 + \vec{r}_2) \cdot \vec{\sigma} + (\vec{r}_1 \cdot \vec{r}_2)I + i (\vec{r}_1 \times \vec{r}_2) \cdot \vec{\sigma} \]
      where the last two terms come from using an identity for the \(\sigma\)-matrices in terms of their commutation (\([\sigma_i, \sigma_j] = 2i \epsilon_{ijk}\sigma_k\)) and anti-commutation (\(\{\hat{\sigma_i}, \hat{\sigma_j}\} \equiv \hat{\sigma_i}\hat{\sigma_j} + \hat{\sigma_j}\hat{\sigma_i} = 2\delta_{ij}I\)) relations.
      Since the identity matrix and the three \(\sigma\)-matrices are linearly independent, the product of density matrices can only vanish if
      \begin{eqnarray*} 1 + \vec{r}_1 \cdot \vec{r}_2 & = & 0 \\ \vec{r}_1 + \vec{r}_2 + i \vec{r}_1 \times \vec{r}_2 & = & \vec{0} \end{eqnarray*}
      Recall that the Bloch vectors cannot have magnitude greater than one, i.e. they must give points on (for pure states) or inside (for mixed states) the Bloch sphere which has radius one. With this restriction the first condition is equivalent to \(\vec{r}_2 = -\vec{r}_1\) and \(|\vec{r}_1| = 1\). I.e. the two states must be pure states and must sit on antipodal points of the Bloch sphere. Clearly the second condition is then satisfied. For a single qubit we see that a mixed state cannot be orthogonal to any other state.
      We could alternatively take the orthogonality condition to be \(\Tr(\hat{\rho}_1 \hat{\rho}_2) = 0\) and since the \(\sigma\)-matrices are traceless we would only get the first constraint. However, note that when \(\vec{r}_2 = -\vec{r}_1\) the second condition above is automatically satisfied, so these conditions for orthogonality are equivalent.
  2. Consider a qubit system with standard basis \(\left\{ \ket{0}, \ket{1} \right\}\). Define the following states:
    \begin{eqnarray*} \ket{+} = \frac{1}{\sqrt{2}} \left( \ket{0} + \ket{1} \right) & , & \ket{-} = \frac{1}{\sqrt{2}} \left( \ket{0} - \ket{1} \right) \\ \ket{L} = \frac{1}{\sqrt{2}} \left( \ket{0} + i\ket{1} \right) & , & \ket{R} = \frac{1}{\sqrt{2}} \left( \ket{0} - i\ket{1} \right) \end{eqnarray*}
    1. Find the density matrices for each of the pure states \(\ket{0}, \ket{1}, \ket{+}, \ket{-}, \ket{L}, \ket{R}\).
    2. Find the density matrices for each of the following mixed states:
      1. \(\ket{0}\) with probability \(\frac{1}{2}\), \(\ket{1}\) with probability \(\frac{1}{4}\), \(\ket{+}\) with probability \(\frac{1}{4}\).
      2. \(\ket{0}\) with probability \(\frac{1}{2}\), \(\ket{1}\) with probability \(\frac{1}{2}\).
      3. \(\ket{+}\) with probability \(\frac{1}{2}\), \(\ket{-}\) with probability \(\frac{1}{2}\).
      4. \(\ket{L}\) with probability \(\frac{1}{2}\), \(\ket{R}\) with probability \(\frac{1}{2}\).
    3. Using the Bloch sphere, sketch the regions which can be described as an ensemble (with arbitrary probabilities whose total is \(1\)) of the following pure states:
      1. \(\ket{0}\) and \(\ket{1}\)
      2. \(\ket{+}\) and \(\ket{-}\)
      3. \(\ket{L}\) and \(\ket{R}\)
      4. \(\ket{1}\) and \(\ket{L}\)
      5. \(\ket{0}\), \(\ket{+}\) and \(\ket{R}\)
    Solution:
    1. For any pure state \(\ket{\psi}\) the density operator is \(\hat{\rho} = \ket{\psi}\bra{\psi}\). For a qubit system we can always write the operator \(\hat{\rho}\) using the standard basis as
      \[ \hat{\rho} = \rho_{00}\ket{0}\bra{0} + \rho_{01}\ket{0}\bra{1} + \rho_{10}\ket{1}\bra{0} + \rho_{11}\ket{1}\bra{1} . \]
      Obviously this can be represented in matrix form as
      \[ \rho = \left( \begin{array}{cc} \rho_{00} & \rho_{01} \\ \rho_{10} & \rho_{11} \end{array} \right) \]
      You can either calculate in Dirac notation or immediately write the states as vectors \(u\) and then calculate \(\rho = u u^{\dagger}\). E.g. for \(\ket{L}\) either note that \(\bra{L} = \frac{1}{\sqrt{2}}(\bra{0} - i\bra{1})\) or just calculate
      \[ \rho_L = \left(\frac{1}{\sqrt{2}}\right)^2 \cvec{1 \\ i} (\begin{array}{cc} 1 & -i \end{array}) = \frac{1}{2} \left( \begin{array}{cc} 1 & -i \\ i & 1 \end{array} \right) . \]
      Similarly for the other cases you should find
      \begin{eqnarray*} \rho_0 & = & \left( \begin{array}{cc} 1 & 0 \\ 0 & 0 \end{array} \right) \\ \rho_1 & = & \left( \begin{array}{cc} 0 & 0 \\ 0 & 1 \end{array} \right) \\ \rho_+ & = & \frac{1}{2} \left( \begin{array}{cc} 1 & 1 \\ 1 & 1 \end{array} \right) \\ \rho_- & = & \frac{1}{2} \left( \begin{array}{cc} 1 & -1 \\ -1 & 1 \end{array} \right) \\ \rho_R & = & \frac{1}{2} \left( \begin{array}{cc} 1 & i \\ -i & 1 \end{array} \right) . \end{eqnarray*}
    2. For a mixed state the density matrix is just the linear combination of the density matrices for each component of the ensemble, weighted by the probability. For the four cases in this question we have density matrices
      \begin{eqnarray*} \frac{1}{2}\rho_0 + \frac{1}{4}\rho_1 + \frac{1}{4}\rho_+ & = & \frac{1}{8} \left( \begin{array}{cc} 5 & 1 \\ 1 & 3 \end{array} \right) \\ \frac{1}{2}\rho_0 + \frac{1}{2}\rho_1 & = & \frac{1}{2} \left( \begin{array}{cc} 1 & 0 \\ 0 & 1 \end{array} \right) \\ \frac{1}{2}\rho_+ + \frac{1}{2}\rho_- & = & \frac{1}{2} \left( \begin{array}{cc} 1 & 0 \\ 0 & 1 \end{array} \right) \\ \frac{1}{2}\rho_L + \frac{1}{2}\rho_R & = & \frac{1}{2} \left( \begin{array}{cc} 1 & 0 \\ 0 & 1 \end{array} \right) . \end{eqnarray*}
    3. Either recall for lectures, or check by extracting the Bloch vector from the density matrix, where each of these pure states sits on the Bloch sphere. You should find that \(\ket{0}\) and \(\ket{1}\) sit on the positive and negative \(z\)-axis respectively. Similarly for \(\ket{+}\) and \(\ket{-}\) on the \(x\)-axis and \(\ket{L}\) and \(\ket{R}\) on the \(y\)-axis. Of course, being pure states, all are distance one from the origin.
      Now a mixed state of \(\ket{0}\) and \(\ket{1}\) will have density matrix given by a linear combination of \(\rho_0\) and \(\rho_1\), but the coefficients are probabilities adding up to one so the density matrix must be of the form
      \[ \rho = (1 - \lambda) \rho_0 + \lambda \rho_1 = \rho_0 + \lambda (\rho_1 - \rho_0) \]
      where \(0 \le \lambda \le 1\). But this is just the parametric equation for the straight line between the points on the Bloch sphere given by \(\rho_0\) and \(\rho_1\), i.e. in this case the points on the \(z\)-axis with \(z=1\) and \(z=-1\). The limited range of \(\lambda\) means that we only consider the line segment with endpoints given by \(\rho_0\) and \(\rho_1\).
      Similarly we can describe the other mixed states of two pure states. Note that as for \(\ket{0}\) and \(\ket{1}\), the lines for \(\ket{+}\) and \(\ket{-}\) and for \(\ket{L}\) and \(\ket{R}\) go through the origin. Indeed the origin is the point corresponding to equal weighting in each case, and this explains the corresponding results in the previous part.
      For the final case we have a mixture of 3 pure states. Then the linear combination gives the parametric equation for the plane containing the three points on the Bloch sphere. Again the coefficients are probabilities totally one, and this constraint defines the triangle with the three points giving the vertices.
      In general a mixed state ensemble with arbitrary probabilities will be represented by the polyhedron, and its interior, defined by the vertices given by the pure states in the ensemble. In special cases where the vertices are co-planar or co-linear the shape will degenerate to a polygon or a line.
  3. Suppose a single qubit system is in the state \(\ket{0}\).
    1. What are the possible outcomes, and associated probabilities, of a measurement of the observable \(\sigma_3\)?
    2. If instead \(\sigma_1\) is measured, what are the possible outcomes and probabilities?
    3. Now suppose \(\sigma_1\) is measured first, then \(\sigma_3\) is measured and then \(\sigma_1\) is measured again. Describe the possible outcomes and probabilities at each stage. Is there any relation between the results of the two measurements of \(\sigma_1\)?
    Solution:
    1. The outcome would be \(1\) with probability \(1\) since \(\ket{0}\) is an eigenstate of \(\sigma_3\) with eigenvalue \(1\).
    2. The eigenstates of \(\sigma_1\) are \(\ket{\pm} = (\ket{0} \pm \ket{1})/\sqrt{2}\) with eigenvalues \(\pm 1\). Since \(\ket{0} = (\ket{+} + \ket{-})/\sqrt{2}\), either outcome \(\pm 1\) is possible with probability \(1/2\).
    3. The first measurement of \(\sigma_1\) will produce result \(\pm 1\) and change the state to \(\ket{\pm}\) with equal probability. Since the states \(\ket{\pm} = (\ket{0} \pm \ket{1})/\sqrt{2}\), the measurement of \(\sigma_3\) will give result \(\pm 1\) and change the state to \(\ket{0}\) or \(\ket{1}\) with equal probability, and without any correlation to the earlier measurement of \(\sigma_1\). Since these states are \((\ket{+} \pm \ket{-})/\sqrt{2}\) the second measurement of \(\sigma_1\) will give result \(\pm 1\) and produce the state \(\ket{\pm}\) with equal probability. The final result does not depend in any way on the previous measurements, in particular it is independent of the result of the first measurement of \(\sigma_1\).
Ask a question about the highlighted paragraph(s):
Your question will be raised anonymously and the answer will appear here at some point.
Long-tap anywhere to ask a question (or see the reply). Right-click anywhere to ask a question (or see the reply).