Chapter 1Free to read

Two-level quantum systems


Published Copyright © IOP Publishing Ltd 2022
Pages 1-1 to 1-14

Download ePub chapter

You need an eReader or compatible software to experience the benefits of the ePub3 file format.

Download complete PDF book, the ePub book or the Kindle book

Export citation and abstract

BibTeX RIS

Share this chapter

978-0-7503-3963-6

Abstract

Chapter 1 summarizes concepts and rules of vectors and matrices used in this book, and fundamental knowledge of quantum mechanics. They are applied throughout this book.

This article is available under the terms of the IOP-Standard Books License

All rights reserved. No part of this publication may be reproduced, stored in a retrieval system or transmitted in any form or by any means, electronic, mechanical, photocopying, recording or otherwise, without the prior permission of the publisher, or as expressly permitted by law or under terms agreed with the appropriate rights organization. Multiple copying is permitted in accordance with the terms of licences issued by the Copyright Licensing Agency, the Copyright Clearance Centre and other reproduction rights organizations.

Permission to make use of IOP Publishing content other than as set out above may be sought at permissions@ioppublishing.org.

Shinil Cho has asserted his right to be identified as the author of this work in accordance with sections 77 and 78 of the Copyright, Designs and Patents Act 1988.

We need a basic knowledge of linear algebra (vectors and matrices) throughout this book. Because readers of this book may be unfamiliar with vectors and matrices, in the first half of this chapter we will briefly describe them. In the second half, rotations of spins and the coordinates, projection operators to represent observations, and entanglement and superposition of quantum states will be discussed. Mathematics of photon-based qbit is also given.

1.1. Vectors and matrices

1.1.1. Calculation rules of vectors and matrices

Two-dimensional vector

In this book, we use Dirac's vector notation used in quantum physics. Consider an arbitrary two-dimensional vector ∣$\mathit{\unicode[Book Antiqua]{x76}}$> as shown in figure 1.1 in the ket vector format: $| \mathit{\unicode[Book Antiqua]{x76}}\rangle =\left[\begin{array}{c}{\mathit{\unicode[Book Antiqua]{x76}}}_{1}\\ {\mathit{\unicode[Book Antiqua]{x76}}}_{2}\end{array}\right]$ where $\mathit{\unicode[Book Antiqua]{x76}}$ 1 and $\mathit{\unicode[Book Antiqua]{x76}}$ 2 are the x- and y-components, respectively. For a Euclidean space, the coefficients are real values.

Figure 1.1.

Figure 1.1. Two-dimensional vector space.

Standard image High-resolution image

Using a set of unit vectors, $| {e}_{1}=\left[\begin{array}{c}1\\ 0\end{array}\right]$ and $| {e}_{2}\rangle =\left[\begin{array}{c}0\\ 1\end{array}\right]$, the vector ∣$\mathit{\unicode[Book Antiqua]{x76}}$> can be expressed as

Equation (1.1)

Define the bra vector: $\langle u| =\left[{u}_{1}{u}_{2}\right]$, and the inner (scalar) product of two vectors, <u∣ and ∣$\mathit{\unicode[Book Antiqua]{x76}}$>, can be defined as

Equation (1.2)

Using the inner product, the orthonormal property of the unit vectors can be expressed as

Equation (1.3)

Also, the definiton of the outer product of the two vectors is given by

Equation (1.4)

Matrices as operators on vectors

Because of the orthonormal property of the unit vectors, the vector components, $\mathit{\unicode[Book Antiqua]{x76}}$ 1 and $\mathit{\unicode[Book Antiqua]{x76}}$ 2, can be given by the inner products of the unit vector and the vector ∣$\mathit{\unicode[Book Antiqua]{x76}}$>:

Equation (1.5)

Therefore,

Equation (1.6)

Here, the outer products of the unit vectors, $| e{}_{1}\rangle \langle {e}_{1}| $ and $| e{}_{2}\rangle \langle {e}_{2}| $, are called the projection operators:

Equation (1.7)

Notice ${\hat{P}}_{1}| \mathit{\unicode[Book Antiqua]{x76}}\rangle ={\mathit{\unicode[Book Antiqua]{x76}}}_{1}| {e}_{1}\rangle $, ${\hat{P}}_{2}| \mathit{\unicode[Book Antiqua]{x76}}\rangle ={\mathit{\unicode[Book Antiqua]{x76}}}_{2}| {e}_{2}\rangle $, ${\hat{P}}_{1}| \mathit{\unicode[Book Antiqua]{x76}}\rangle +{\hat{P}}_{2}| \mathit{\unicode[Book Antiqua]{x76}}\rangle =| \mathit{\unicode[Book Antiqua]{x76}}\rangle $, and thus

Equation (1.8)

where $\hat{I}$ is the 2×2 unit matrix.

Below is a summary of the addition and multiplication of matrices. Suppose

the basic matrix calculation rules are:

  • (1)  
    Equation (1.9)
  • (2)  
    Addition:
    Equation (1.10)
  • (3)  
    Multiplication:

Equation (1.11)

Rotation and translation of a vector can be performed by applying corresponding matrices. For example, as shown in figure 1.1, rotating a vector ∣$\mathit{\unicode[Book Antiqua]{x76}}$> by angle θ to create a new vector, ∣$\mathit{\unicode[Book Antiqua]{x76}}$ '>, is expressed by ∣$\mathit{\unicode[Book Antiqua]{x76}}$ '>=R(θ)∣$\mathit{\unicode[Book Antiqua]{x76}}$>. The rotational matrix, R(θ), can be obtained as follows. Let ${\mathit{\unicode[Book Antiqua]{x76}}}_{1}=\mathit{\unicode[Book Antiqua]{x76}}\,\cos \,\alpha $ and ${\mathit{\unicode[Book Antiqua]{x76}}}_{2}=\mathit{\unicode[Book Antiqua]{x76}}\,\sin \,\alpha $. The components of the rotated vector are given by

and

In the matrix representation (equation (1.11)), the above equations can be expressed as

and the rotational matrix is, thus, given by

Equation (1.12)

Notice that rotation of the two-dimensional coordinate system (x, y) by angle θ to another coordinate system, (X, Y), is mathematically equivalent to the vector rotation by angle –θ. Thus,

Equation (1.13)

1.1.2. Combining two different vector spaces—direct product

Consider two different two-dimensional vector spaces where each vector space has its own set of unit vectors or an orthonormal basis. Suppose a vector in each vector space is respectively given by

A vector in the coupled vector space is expressed by the direct product of the two vectors:

Equation (1.14)

The coupled vector space becomes four-dimensional, and the coupled orthonormal basis is given by

Equation (1.15)

Notice that the direct product has the following distribution rule:

Equation (1.16)

where the subscripts, A and B, indicate two different vector spaces. Notice that we often write $| u\rangle \otimes | \mathit{\unicode[Book Antiqua]{x76}}\rangle =| u\rangle | \mathit{\unicode[Book Antiqua]{x76}}\rangle $ for short. It is important not to change the order of vector for the direct product.

We also define the direct product of two matrices, $A=\left[\begin{array}{cc}{a}_{11} & {a}_{12}\\ {a}_{21} & {a}_{22}\end{array}\right]$ and $B=\left[\begin{array}{cc}{b}_{11} & {b}_{12}\\ {b}_{21} & {b}_{22}\end{array}\right]$, as

Equation (1.17)

We will use the direct products when we discuss universal operators in chapter 2.

1.2. Foundation of quantum mechanics

1.2.1. General properties of quantum states

Here is a summary of properties that quantum systems must satisfy [1]. Whenever we create a quantum algorithm, it must satisfy these conditions.

  • (1)  
    The quantum states of a system can be described by a single-valued continuous complex-valued function, $| \psi \rangle $, and the physics observables, A, including energy, position, momentum, angular momentum, spins, and particle creation/annihilation, can be expressed as a mathematical operator, $\hat{A}$. Measurement values, {εk ; k=1, 2, ...}, of the observable A satisfy the equation, $\hat{A}| {\psi }_{k}\rangle ={\varepsilon }_{k}| {\psi }_{k}\rangle $, where εk is the eigen value, and $| \psi {}_{k}\rangle $ is eigen function of the equation. When the observable is Hamiltonian (energy), the equation is called the Schrödinger equation, $\hat{H}| {\psi }_{k}\rangle ={\varepsilon }_{k}| {\psi }_{k}\rangle $.
  • (2)  
    A single measurement of the physical observable A yields one of the possible eigen values {εk }. The equation itself does not determine the eigen functions and the eigen values without a boundary condition.
  • (3)  
    A quantum state can be normalized, and its operators are linear:
    • (i)  
      $\langle \psi | \psi \rangle =1$ where $\langle \psi | ={\left(| \psi \rangle \right)}^{\dagger }$ is the conjugate transpose of ∣ψ>, and
    • (ii)  
      $\hat{A}\left(a| {\psi }_{1}\rangle +b| {\psi }_{2}\rangle \right)=a\hat{A}| {\psi }_{1}\rangle +b\hat{A}| {\psi }_{2}\rangle $ where the coefficients a and b are constants of complex values.
  • (4)  
    Because the physical observables are real values, the eigen values must be real numbers. Therefore, the operator $\hat{A}$ must be Hermitian, i.e., ${\hat{A}}^{\dagger }=\hat{A}$.Proof: Suppose $\hat{A}| {\psi }_{k}\rangle ={\varepsilon }_{k}| {\psi }_{k}\rangle $, and the eigen value is real, i.e., ${\varepsilon }_{k}={\varepsilon }_{k}^{* }$. $\langle {\psi }_{k}| \hat{A}| {\psi }_{k}\rangle =\langle {\psi }_{k}| {\varepsilon }_{k}| {\psi }_{k}\rangle ={\varepsilon }_{k}\langle {\psi }_{k}| {\psi }_{k}\rangle ={\varepsilon }_{k},{\rm{a}}{\rm{n}}{\rm{d}}$ ${\left(\langle {\psi }_{k}| \hat{A}| {\psi }_{k}\rangle \right)}^{\dagger }=\langle {\psi }_{k}| {\hat{A}}^{\dagger }| {\psi }_{k}\rangle ={\varepsilon }_{k}^{* }$.Thus, if the eigenvalue is a real number, ${\hat{A}}^{\dagger }=\hat{A}$.
  • (5)  
    The eigen functions are orthogonal, i.e., $\langle {\psi }_{i}| {\psi }_{j}\rangle =0$ if $i\ne j$.Proof. Suppose $\hat{A}| {\psi }_{i}\rangle ={\varepsilon }_{i}| {\psi }_{i}\rangle $ and $\hat{A}| {\psi }_{j}\rangle ={\varepsilon }_{j}| {\psi }_{j}\rangle $ where $\hat{A}$ is Hermitian.Notice ${\left(\hat{A}| {\psi }_{i}\rangle \right)}^{\dagger }=\langle {\hat{A}}^{\dagger }{\psi }_{i}| =\langle \hat{A}{\psi }_{i}| $ because $\hat{A}$ is Hermitian. Also notice that because the eigen values are real, ${\left(\hat{A}| {\psi }_{i}\rangle \right)}^{\dagger }={\left({\varepsilon }_{i}| {\psi }_{i}\rangle \right)}^{\dagger }={\varepsilon }_{i}\langle {\psi }_{i}| $. Therefore, $\langle {\hat{A}}^{\dagger }{\psi }_{i}| {\psi }_{j}\rangle =\langle {\psi }_{i}| \hat{A}{\psi }_{j}\rangle ={\varepsilon }_{j}\langle {\psi }_{i}| {\psi }_{j}\rangle ,{\rm{a}}{\rm{n}}{\rm{d}}$ $\langle {\hat{A}}^{\dagger }{\psi }_{i}| {\psi }_{j}\rangle -\langle {\psi }_{i}| \hat{A}{\psi }_{j}\rangle =\left({\varepsilon }_{i}-{\varepsilon }_{j}\right)\langle {\psi }_{i}| {\psi }_{j}\rangle =0$. Thus, $\langle {\psi }_{i}| {\psi }_{j}\rangle =0$ unless ${\varepsilon }_{i}={\varepsilon }_{j}$.
  • (6)  
    The complete set of eigen functions {$| \psi {}_{k}\rangle $, k=1, 2, 3, ....} forms an orthonormal basis. In other words, any quantum state of the given system can be expressed by the superposition of eigen functions: $| \psi \rangle =\displaystyle \displaystyle \sum _{k}{c}_{k}| {\psi }_{k}\rangle $ where $\langle {\psi }_{j}| {\psi }_{k}\rangle ={\delta }_{jk}$ and ${c}_{k}=\langle {\psi }_{k}| \psi \rangle $.
  • (7)  
    Suppose a coordinate system for observation is changed to another coordinated system, a complete set of the eigen functions {$| \psi {}_{k}\rangle $, k=1, 2, 3, ....} is transformed to another set of eigen functions {$| \varphi {}_{k}\rangle $, k=1, 2, 3, ....}, i.e., $| \varphi \rangle =\hat{U}| \psi \rangle $ where $\hat{U}$ is the operator for this transformation. Then, $| \varphi {\rangle }_{j}=\displaystyle \displaystyle \sum _{k=1}{u}_{jk}| \psi {\rangle }_{k}$, and $\langle {\varphi }_{i}| {\varphi }_{j}\rangle =\displaystyle \displaystyle \sum _{k,m}{({u}_{ik})}^{\dagger }{u}_{mj}\langle {\psi }_{k}| {\psi }_{m}\rangle $. Since both sets of eigen functions are orthonormal, $\langle {\varphi }_{i}| {\varphi }_{j}\rangle ={\delta }_{i,j}=\displaystyle \displaystyle \sum _{k,m}{({u}_{ik})}^{\dagger }{u}_{mj}{\delta }_{k,m}$.Note: An operator, which satisfies $\displaystyle \displaystyle \sum _{k}{({u}_{ik})}^{\dagger }{u}_{kj}={\delta }_{i,j}$, is called the unitary operator, and ${\hat{U}}^{\dagger }\hat{U}=\hat{I}$, i.e., ${\hat{U}}^{\dagger }={\hat{U}}^{-1}$. Operators that change quantum states must be unitary.
  • (8)  
    The expectation value of the observable A is given by $\langle A\rangle =\langle \psi | \hat{A}| \psi \rangle $.In particular, $\langle A\rangle =\langle {\psi }_{k}| \hat{A}| {\psi }_{k}\rangle ={\varepsilon }_{k}$ if $| \psi \rangle =| {\psi }_{k}\rangle $.
  • (9)  
    A projection operator is given by ${\hat{P}}_{k}=| {\psi }_{k}\rangle \langle {\psi }_{k}| $.If $| \psi \rangle =\displaystyle \sum _{k}{c}_{k}| {\psi }_{k}\rangle $, then ${\hat{P}}_{k}| \psi \rangle =\displaystyle \sum _{j}{c}_{j}| {\psi }_{k}\rangle \langle {\psi }_{k}| {\psi }_{j}\rangle =\displaystyle \sum _{j}{c}_{j}| {\psi }_{k}\rangle {\delta }_{kj}={c}_{k}| {\psi }_{k}\rangle $.The probability of observing the eigen state $| \psi {}_{k}\rangle $ is
  • (10) Matrix representation of operator $\hat{U}$.

Let $| \psi \rangle =\displaystyle \displaystyle \sum _{j}{c}_{j}| {\psi }_{k}\rangle $ where ${c}_{j}=\langle {\psi }_{j}| \psi \rangle $, and $| \psi ^{\prime} \rangle =\hat{U}| \psi \rangle =\displaystyle \sum _{j}{c}_{j}\hat{U}| {\psi }_{j}\rangle $.

If $| \psi ^{\prime} \rangle =\displaystyle \displaystyle \sum _{j}c{^{\prime} }_{j}| {\psi }_{k}\rangle $ where $c{^{\prime} }_{j}=\langle {\psi }_{j}| \psi ^{\prime} \rangle $, then $| \psi ^{\prime} \rangle $ becomes

$| \psi ^{\prime} \rangle =\displaystyle \sum _{i}c^{\prime} i| {\psi }_{i}\rangle =\displaystyle \sum _{i}{c}_{i}\hat{U}| {\psi }_{i}\rangle $, and thus,

Define $\langle {\psi }_{i}| \hat{U}| {\psi }_{j}\rangle ={U}_{ij}$, and we obtain $c^{\prime} {}_{i}=\displaystyle \sum _{j}{U}_{ij}{c}_{j}$. Therefore, the transform

$| \psi \rangle \to | \psi ^{\prime} \rangle =\hat{U}| \psi \rangle $ can be represented by a linear transform of {ci}: $c{^{\prime} }_{i}=\displaystyle \sum _{j}{U}_{ij}{c}_{j}$.

1.3. Quantum state vectors

Without rigorous mathematics, we describe quantum states in an analogy of the two-dimensional vectors described above. Here, we focus on two-level systems to describe a quantum bit (qbit). Readers may wonder whether the term, quantum bit, should be written as 'qubit' or 'qbit.' Refer to an interesting article on this subject. 1

1.3.1. Two-level quantum state vector: qbit

An electron is known to have two possible spin states. It can be described on a two-dimensional complex vector space. When we measure the spin state along the z-axis or the 'vertical' axis, which is the default direction of external magnetic field, we observe the spin up (↑) or down (↓) state. We assign the following vectors ∣0> and ∣1> to the spin states:

Equation (1.18)

The spin state vectors form an orthonormal basis of the vector space, {∣0>, ∣1>} where $\langle 0| 0\rangle =\langle 1| 1\rangle =1$, and $\langle 0| 1\rangle =\langle 1| 0\rangle =0$. An arbitrary spin state can be given by $| \psi \rangle =a| 0\rangle +b| 1\rangle $ where the coefficients a and b are complex value constants, and the spin state is normalized, $\left|\langle \psi | \psi \rangle \right|=| a{| }^{2}+| b{| }^{2}=1$ because either spin state will be observed by a measurement. The spin state $| \psi \rangle $ is called a qbit in the quantum computation/information. The complex coefficients, a and b, can be expressed as $a=\,\cos \,\left(\displaystyle \frac{\theta }{2}\right)$ and $b={e}^{i\varphi }\,\sin \,\left(\displaystyle \frac{\theta }{2}\right)$, using phase angles, φ and θ, which satisfy the normalization condition, ∣a2 + ∣b2=1. A Bloch sphere, shown in figure 1.2, is often used to visualize the qbit dynamics although we will not use it in this book.

Figure 1.2.

Figure 1.2. Bloch sphere.

Standard image High-resolution image

1.3.2. Projection operators for spin states

If one observes the spin state, $| \psi \rangle =a| 0\rangle +b| 1\rangle $, along the z-axis, the measurement yields either spin up or down. Observation of the spin state can be interpreted as the application of the projection operators. From equation (1.7), the projection operators for the spin state described by the orthonormal basis, {∣0>, ∣1>}, are given

Equation (1.19)

The probability of observing the spin up state along the z-axis, ∣a2, can be calculated from a measurement

Equation (1.20)

and the probability of observing the spin down state along the z-axis, ∣b2, can be calculated from a measurement

Equation (1.21)

Once a measurement is conducted on a quantum system, the quantum state of the system is collapsed, and a successive measurement yields the same state as the previous measurement. This means

1.3.3. Time evolution of spin states

To obtain the time evolution and rotation of spins, we use exponential operators. We define the form of exponential operators as

Equation (1.22)

where u is a complex number. Suppose the wave function is given by $| \psi (0)\rangle $ at t = 0, the wave function satisfies the time dependent Schrödinger equation:

where $\hat{H}$ is the Hamiltonian of the system. From the Schrödinger equation, the time evolution of the wave function is given by

Equation (1.23)

Proof.

1.3.4. Rotation of spin states

Define the rotational operator, ${\hat{R}}_{j}(\delta \theta )$ of an infinitesimal angle $\delta \vec{\theta }$ about the rotational axis j, such that ${\hat{R}}_{j}(\delta \vec{\theta })| \psi (\vec{r})\rangle =| \psi ^{\prime} (\vec{r})\rangle $ on the wave function $| \psi (\vec{r})\rangle $. Because the position of the vector after it is rotated forward by an infinitesimal angle $\delta \vec{\theta }$ is given by $\vec{r}^{\prime} =\vec{r}+\delta \vec{\theta }\times \vec{r}$ using the vector product (figure 1.3), we obtain

and thus,

Equation (1.24)

where $\vec{p}=-i\hslash \nabla $ is the linear momentum, and $\vec{L}=\vec{r}\times \vec{p}$ is the angular momentum.

Figure 1.3.

Figure 1.3. Infinitesimal rotation of the position vector.

Standard image High-resolution image

The infinitesimal angle is defined as $\delta \vec{\theta }=\mathop{\mathrm{lim}}\limits_{N\to \infty }\,\vec{\theta }/N$, and we obtain

Equation (1.25)

The spin rotational operators about the x, the y, and the z axes are, respectively, given by replacing the angular momentum of each axis with Pauli's spin matrix of the corresponding axis:

Equation (1.26)

where j = x, y, or z, and

${\hat{\sigma }}_{x}=\left[\begin{array}{cc}0 & 1\\ 1 & 0\end{array}\right]$, ${\hat{\sigma }}_{y}=\left[\begin{array}{cc}0 & -\,i\\ i & 0\end{array}\right]$, ${\hat{\sigma }}_{z}=\left[\begin{array}{cc}1 & 0\\ 0 & -\,1\end{array}\right]$, and $\hat{I}=\left[\begin{array}{cc}1 & 0\\ 0 & 1\end{array}\right]$.

Thus,

Equation (1.27)

Equation (1.28)

Equation (1.29)

where we have omitted the suffix of the angles. For basic concepts of spin states, refer to a textbook on quantum mechanics [1].

1.3.5. Rotation of a spin observation coordinate frame

In order to observe a spin state, an external magnetic field is applied to the spin. If the external magnet is along the z-axis, we observe either ∣0> or ∣1> as we discussed above (equations (1.19) and (1.20)). Now, if we rotate the external magnetic field by angle θ about one of the coordinate axes, we rotate the observation coordinate frame. Thus, if we rotate the z-axis by angle θ on the zx-plane, i.e., if we rotate the 'vertical' spin-measurement coordinate frame about the y-axis by angle θ, we obtain a new 'tilted' measurement coordinate frame where the rotated orthonormal basis is given by:

Equation (1.30)

and

In particular, if θ=π/2, the direction of the external magnetic field becomes 'horizontal,' or

Equation (1.31)

Recall that if we observe the spin state, $| \psi \rangle =a| 0\rangle +b| 1\rangle $, along the 'vertical' coordinates frame where the external magnetic field is along the z-axis, an observed state is either ∣0> or ∣1> with the corresponding probability ∣a2 or ∣b2. If we observe the spin state using the 'horizontal' measurement frame where the external magnetic field is along the x-axis, the spin state to be observed can be determined by applying the projection operators, ${\hat{P}}_{| 0^{\prime} \rangle }=| 0^{\prime} \rangle \langle 0^{\prime} | $ and ${\hat{P}}_{| 1^{\prime} \rangle }=| 1^{\prime} \rangle \langle 1^{\prime} | $ to the basis ∣0> and ∣1>:

Equation (1.32)

and

Thus, the spin state becomes $| \psi \rangle =\displaystyle \frac{1}{\sqrt{2}}(a| 0^{\prime} \rangle +b| 1^{\prime} \rangle )$ in the 'horizontal' measurement frame, and we observe ∣0'> or ∣1'>, and the probability of observing ∣0'> is ∣a2/2 and the probability of observing ∣1'> spin is ∣b2/2 along the horizontal magnetic field. This is a useful description when we discuss Bell's inequality where three different observation frames are involved (section 5.1).

1.4. Non-cloning principle for qbit

One of the distinct characteristics of a qbit is that it cannot be copied because an act of observation to copy a qbit, $| \psi \rangle =a| 0\rangle +b| 1\rangle $, changes the quantum state to either $| 0\rangle $ or $| 1\rangle $, and any successive measurement yields the same state as the first measurement. Let us restate this non-cloning principle.

Suppose we could define a 'copy' operator, $\hat{C}$, such that $\hat{C}| \psi \rangle =| \psi \rangle | \psi \rangle $ where $| \psi \rangle =a| 0\rangle +b| 1\rangle $. Then, by the definition of the copy operator,

On the other hand,

Equation (1.33)

Therefore, unless a=0 or b=0, we cannot define the copy operator. If a=0, then b2 = b and thus b=1; and if b=0, a2=a and a=1. These conditions indicate the pure quantum state $| \psi \rangle =| 0\rangle $ or $| \psi \rangle =| 1\rangle $ in which there is no quantum fluctuation. The non-cloning principle plays an important role in detecting data tapping while transmitting a secret code using the quantum teleportation (section 6.3).

1.5. Quantum entanglement

1.5.1. What is entanglement?

Suppose a set of two electron spins, spin A and spin B, are coupled, the net spin S is 1 or 0. Such an example is two electrons in a hydrogen molecule. As shown in table 1.1, if S = 1, then Sz = +1, 0, or −1 whereas, if S=0, then Sz = 0 only.

Table 1.1.  Two coupled spins.

Net spin S Sz Quantum state
S = 1 +1 $| {0}_{A}{0}_{B}\rangle $
0 $\displaystyle \frac{1}{\sqrt{2}}\left(| {0}_{A}{1}_{B}\rangle +| {0}_{A}{1}_{B}\rangle \right)$
−1 $| {1}_{A}{1}_{B}\rangle $
S = 0 0 $\displaystyle \frac{1}{\sqrt{2}}\left(| {0}_{A}{1}_{B}\rangle -| {1}_{A}{0}_{B}\rangle \right)$

The quantum states of S=1 and Sz = 0 or S= 0 and Sz = 0 are called the entangled states. Suppose we conduct an observation of spin A when S = 0. Using the projection operator, the spin state becomes

Equation (1.34)

That is, if spin A is up (↑), then spin B must be down (↓). Observing the spin A will automatically determine the spin state B. It seems to be trivial, and this is also true for the classical physics. However, recall that unless we observe one of them, both spins A and B have equal probabilities of spin up and down!

What is important to perceive is that, in classical physics, all spin states are pre-determined regardless of whether we observe them or not. Therefore, the spin states A and B are already determined with the condition of the total spin value. On the other hand, in quantum physics, unless we make observation of spin A, spin A can be both up and down with equal probabilities, and spin B can also be both up and down with equal probabilities. This argument is very much like the quantum logic of Schrödinger's cat.

According to quantum mechanics, an act of observation of the spin A state instantly determines the spin B state, depending on the outcome of the spin state A. Remember that the spin A state is undermined before observation. This is why we call it the 'spooky' behavior of quantum entanglement. Furthermore, the entanglement is non-localizing, i.e., even if the spins are physically separated (without observation) by an astronomical distance after making the coupled state, they are still entangled!

The quantum entanglement is such a strange and unique quantum behavior, and it is one of the key properties that make quantum computation and information theory distinct from the classical theory. One of its astonishing examples is quantum teleportation, which will be discussed in chapter 5.

1.5.2. Superposition and entanglement

Imagine that two classical sinusoidal waves of frequencies, sin(ω1 t) and sin(ω2 t) are superposed, the resultant wave is given by

There are two possible frequencies after superposition of two classical waves. The quantum superposition is fundamentally different from the superposition of n-classical waves which linearly gives only n different states. The number of states exponentially increases in the quantum superposition.

A qbit $| \psi \rangle =a| 0\rangle +b| 1\rangle $ is a superposition of ∣0> and ∣1> states. The possible states from superposition of two qbits are given by the direct product defined by equation (1.14)

Equation (1.35)

Now, the number of possible states become four. If n-qbits are superposed,

Equation (1.36)

This represents a superposed 2n -state from $| \mathrm{00\cdots 0}\rangle $ to $| \mathrm{11\cdots 1}\rangle $ with equal weight. Computation using superposed qbits is superior to the classical binary bit calculation (quantum supremacy) because we can perform computation on all the superposed states at once.

It must be noted that the quantum entanglement is not superposition of quantum states. If two qbits are entangled, they cannot be expressed as the direct product of two qbits. In other words, they cannot be separated as two independent quantum events. In chapter 3, we show quantum gates that superpose, entangle, and detangle qbits using conditional gates.

1.6. Another example of qbit

Other than spins, another example of qbit is linearly polarized photons. Using a linear polarizer, we can obtain the horizontally polarized photon (↔) and the vertically polarized photon (↕) as we define $| \leftrightarrow \rangle =| 0\rangle =\left[\begin{array}{c}1\\ 0\end{array}\right]$ and $| \updownarrow \rangle =| 1\rangle =\left[\begin{array}{c}0\\ 1\end{array}\right]$, respectively.

A photon with an arbitrary linear polarization can be expressed as $| \psi \rangle =a| \updownarrow \rangle +b| \leftrightarrow \rangle $ because of the orthonormal condition:

Equation (1.37)

For photon qbits, the projection operators represent the observation of light using a polarizer whose polarization axis is either in the horizontal or in the vertical direction.

Equation (1.38)

and

Equation (1.39)

Similar to the electron spin, the probability of observing the horizontally polarized photon is given by ∣a2 from the measurement

${\hat{P}}_{0}| \psi \rangle =| 0\rangle \langle 0| \psi \rangle =| 0\rangle \langle 0| \left(a| 0\rangle +b| 1\rangle \right)=a| 0\rangle =a\left[\begin{array}{c}1\\ 0\end{array}\right]$, and the probability of observing the vertically polarized photon is given by ∣b2 from the measurement

We use photon based qbits in chapter 6, Quantum Cryptograph.

References

  • [1]Liboff R L 2003 For a general description of quantum mechanics, refer to, for example Introductory Quantum Mechanics  (Reading, MA: Addison Wesley) 

Export references: BibTeX RIS

Footnotes