\( \newcommand{\blah}{blah-blah-blah} \newcommand{\eqb}[1]{\begin{eqnarray*}#1\end{eqnarray*}} \newcommand{\eqbn}[1]{\begin{eqnarray}#1\end{eqnarray}} \newcommand{\bb}[1]{\mathbf{#1}} \newcommand{\mat}[1]{\begin{bmatrix}#1\end{bmatrix}} \newcommand{\nchoose}[2]{\left(\begin{array}{c} #1 \\ #2 \end{array}\right)} \newcommand{\defn}{\stackrel{\vartriangle}{=}} \newcommand{\rvectwo}[2]{\left(\begin{array}{c} #1 \\ #2 \end{array}\right)} \newcommand{\rvecthree}[3]{\left(\begin{array}{r} #1 \\ #2\\ #3\end{array}\right)} \newcommand{\rvecdots}[3]{\left(\begin{array}{r} #1 \\ #2\\ \vdots\\ #3\end{array}\right)} \newcommand{\vectwo}[2]{\left[\begin{array}{r} #1\\#2\end{array}\right]} \newcommand{\vecthree}[3]{\left[\begin{array}{r} #1 \\ #2\\ #3\end{array}\right]} \newcommand{\vecfour}[4]{\left[\begin{array}{r} #1 \\ #2\\ #3\\ #4\end{array}\right]} \newcommand{\vecdots}[3]{\left[\begin{array}{r} #1 \\ #2\\ \vdots\\ #3\end{array}\right]} \newcommand{\eql}{\;\; = \;\;} \definecolor{dkblue}{RGB}{0,0,120} \definecolor{dkred}{RGB}{120,0,0} \definecolor{dkgreen}{RGB}{0,120,0} \newcommand{\bigsp}{\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;} \newcommand{\plss}{\;\;+\;\;} \newcommand{\miss}{\;\;-\;\;} \newcommand{\implies}{\Rightarrow\;\;\;\;\;\;\;\;\;\;\;\;} \newcommand{\prob}[1]{\mbox{Pr}\left[ #1 \right]} \newcommand{\exval}[1]{\mbox{E}\left[ #1 \right]} \newcommand{\variance}[1]{\mbox{Var}\left[ #1 \right]} \newcommand{\kt}[1]{\left\vert #1 \right\rangle} \newcommand{\br}[1]{\left\langle #1 \right\vert} \newcommand{\bkt}[2]{\left\langle #1 \middle\vert #2 \right\rangle} \newcommand{\inr}[2]{\left\langle #1 \middle\vert #2 \right\rangle} \newcommand{\inrs}[2]{\left\langle #1 \: \middle\vert\: #2 \right\rangle} \newcommand{\inrh}[2]{ \left\langle \vphantom{\huge x} #1 \: \middle\vert \: #2 \right\rangle } \newcommand{\swich}[3]{\left\langle #1 \middle\vert #2 \middle\vert #3\right\rangle} \newcommand{\swichs}[3]{\left\langle #1 \:\middle\vert \: #2 \: \middle\vert \: #3\right\rangle} \newcommand{\swichh}[3]{\left\langle \vphantom{\huge x} #1 \;\middle\vert \; #2 \; \middle\vert \; #3\right\rangle} \newcommand{\otr}[2]{\left\vert #1 \right\rangle\!\left\langle #2 \right\vert} \newcommand{\otrh}[2]{\left\vert \vphantom{\huge x} #1 \right\rangle\!\left\langle \vphantom{\huge x} #2 \right\vert} \newcommand{\pss}{\large\psi} \newcommand{\re}{\mbox{Re }} \newcommand{\im}{\mbox{Im }} \newcommand{\mag}[1]{\left\vert #1 \right\vert} \newcommand{\magsq}[1]{{\left\vert #1 \right\vert}^2} \newcommand{\magsqh}[1]{{\left\vert \vphantom{\huge x} #1 \right\vert}^2} \newcommand{\isqt}[1]{\frac{#1}{\sqrt{2}}} \newcommand{\mbx}[1]{\;\;\;\;\;\;\;\;{\scriptsize \color{Gray}{\mbox{ #1}}}} \newcommand{\ksi}{\kt{\psi}} \newcommand{\parenh}[1]{\left( \vphantom{\huge x} #1 \right)} \newcommand{\parenl}[1]{\left(\vphantom{\LARGE x} #1 \,\right)} \newcommand{\khi}{\kt{\phi}} \newcommand{\cnot}{C_{\scriptsize NOT}} \newcommand{\setl}[1]{\left\{\vphantom{\LARGE x} #1 \,\right\}} \newcommand{\smm}[1]{\mbox{\( #1 \)}} \newcommand{\cz}{C_{\scriptsize Z}} \newcommand{\ccnot}{CC_{\scriptsize NOT}} \newcommand{\ccz}{CC_{\scriptsize Z}} \newcommand{\mbxx}[1]{{\scriptsize \color{Gray}{\mbox{ #1}}}} \newcommand{\sml}[1]{{\scriptsize #1}} \newcommand{\smm}[1]{\mbox{\( #1 \)}} \newcommand{\smml}[1]{ {\scriptsize{\mbox{\(#1\)}} } } \newcommand{\smb}[1]{ {\tiny{\mbox{\(#1\)}} } } \newcommand{\exps}[1]{\exp ({\scriptsize #1} )} \newcommand{\isqts}[1]{\smm{\frac{#1}{\sqrt{2}}}} \newcommand{\halfsm}{\smm{\frac{1}{2}}} \newcommand{\smf}[2]{\mbox{\( \frac{#1}{#2} \)}} \)


Module 9: Toward the continuum: What is quantum mechanics?

Module-9 solved problems

 


Module objectives

 

The main goals of this module:

 


9.1   Discrete quantum mechanics

 

Let's start by casting quantum computing in terms of a discrete formulation of quantum mechanics:

  • What is the system under consideration?
        \(\rhd\) In our case: a group of qubits

  • The system at any time is in some state represented by a vector:
        \(\rhd\) Example: \(\ksi\) = the current state of n qubits

  • A state can change in only one of two ways:
    • A unitary operation
    • A measurement.

  • A unitary operation is represented by a unitary matrix.

  • A measurement is represented either by:
    • A collection of projectors, one per subspace.
    • Or a Hermitian that packages the projectors.

  • In the standard model, a collection of unitary operators \(U_1, U_2,\ldots U_k\) are applied to a starting state \(\kt{\psi_0}\): $$\eqb{ \kt{\psi_1} & \eql & U_1 \kt{\psi_0} \\ \kt{\psi_2} & \eql & U_2 \kt{\psi_1} \\ & \vdots & \\ \kt{\psi_k} & \eql & U_k \kt{\psi_{k-1}} \\ }$$ And then, measurement is performed on the end state \(\kt{\psi_k}\):

  • Pictorially:

  • The only additional feature of the theory not obvious above is entanglement and the effect of measurement on an entangled vector:
    • For example, suppose a 1-qubit (Alice's) measurement is performed on $$ \ksi \eql \isqts{1} \parenl{ \kt{00} + \kt{11} } $$
    • The 1-qubit measurement is really a 2-qubit measurement using two projectors, \(\otr{0}{0} \otimes I\) and \(\otr{1}{1} \otimes I\).
    • This will result in one of two 2-qubit states: \(\kt{00}\) or \(\kt{11}\)
    • The result is that further measurement on Bob's side is completely determined.

  • In some sense, this is a complete theory of discrete quantum mechanics:
        \(\rhd\) There is no other (discrete) phenomenon to describe.

  • However, it is not enough to describe everything else in nature, just the abstract qubits in quantum computing and their behavior.
 

But even a supposedly complete theory may have holes:

  • For example, how is one to explain the instantaneous effect of Alice's measurement on Bob's qubit?
 


9.2   The EPR paradox, hidden variables and Bell's theorem

 

In 1935, Einstein, Podolsky and Rosen (EPR) wrote a landmark critique of quantum theory:

  • They said any valid theory must have two features:
    1. Realism. Any theory that perfectly predicts some physically observable quantity must have a variable in it that tracks or computes that quantity.
    2. Locality. Nature acts locally and physical influences are subject to the speed of light limitation.

  • But the Alice-Bob measurement of an entangled pair shows that:
    • After Alice's measurement, quantum theory predicts Bob's outcome exactly.
    • This is instantaneous no matter the distance between Alice and Bob.
    • Thus, quantum theory's realism is decidedly nonlocal.

  • Therefore, according to EPR, quantum theory is incomplete because it does not have corresponding variables that exhibit locality in this case.

  • For example, is it possible that Alice's and Bob's qubits carry in them some hidden state?
    • Perhaps science is not advanced enough to see or probe the hidden state.
    • And yet the hidden state determines the outcomes.
    • After all, the history is science is rife with examples where some underlying scale-dependent quantity was unknown to some fairly successful theories
          \(\rhd\) Example: Newtonian mechanics (not accurate when scaled to high velocity)

  • This seeming contradiction in the theory itself was referred to (historically) as the EPR paradox.
        \(\rhd\) It is no longer considered a paradox today.
 

Let's now suppose there exist local hidden variables, that travel with the qubits:

  • We will show that any theory with such variables will predict outcomes different from quantum theory.

  • Because quantum theory's predictions match experimental results, any such theory must be rejected.

  • The reasoning is delicate and requires many steps, which we will undertake in the next few sections.

  • The findings about hidden variables have been generalized into a collection of results called Bell's Inequalities:
    • Each inequality is something that must be satisfied by a local (hidden-variable) theory.
    • But quantum theory is shown to violate such inequalities.

  • The inequalities are named for John Bell, the physicist who derived the first of them and proved E, P, and R were wrong.
 


9.3   Hidden variables, part I: measuring polarization

 

First, let's recall the use of the sandwich for projectors:

  • For any \(\ksi\) and projector \(P\), we know that $$ \mbox{Probability of seeing outcome \(P\ksi\)} \eql \magsq{P\ksi} $$

  • An alternative way to calculate this probability is to write it as a projector sandwich: $$ \mbox{Probability of seeing outcome \(P\ksi\)} \eql \swich{\psi}{P}{\psi} $$ (See Module 2).

  • Doing so enables exploiting Dirac-notation inner-products, as we'll see when we work out the details.
 

Now consider this set up:

  • A device generates EPR photon-pairs whose polarizations are entangled in the first Bell state: $$ \ksi \eql \isqts{1} \parenl{ \kt{00} + \kt{11} } $$ where
    • \(\kt{0}\) represents vertical polarization
    • \(\kt{1}\) represents horizontal

  • Alice measures first, using a polarizer oriented at angle \(\theta_1\).

  • Bob later measures with polarizer at \(\theta_2\) from the vertical.

  • Qualitatively, what are the outcomes?
    • For Alice, her photon either passes (P) or gets absorbed (A).
    • Same for Bob.

  • We will be interested in these two cases:
    • Agreement:
      • If Alice's photon passes, so does Bob's (P-P)
      • If Alice's photon is absorbed, so is Bob's (A-A).
    • Disagreement: different results on their photons. (P-A, A-P).
 

As a first step, let's consider how a single photon behaves when a polarizer is oriented at an angle:

  • Suppose this basis is represented by $$\eqb{ \kt{v} & \equiv & \;\; \mbox{Pass through} \\ \kt{v^\perp} & \equiv & \;\; \mbox{Absorb} \\ }$$

  • For the photon, we'll use $$\eqb{ \kt{0} & \equiv & \;\; \mbox{Vertically polarized} \\ \kt{1} & \equiv & \;\; \mbox{Horizontally polarized} \\ }$$

  • Consider a vertically polarized photon (in state \(\kt{0}\)) arriving at a \(\theta\)-filter.

  • It will be convenient to express the filter's \(\kt{v},\kt{v^\perp}\) basis vectors in terms of the standard basis \(\kt{0},\kt{1}\).

  • When \(\theta=0\), we already know that $$\eqb{ \kt{v} & \eql & \kt{0} & \mbx{Vertical photons pass through with probability 1} \\ \kt{v^\perp} & \eql & \kt{1} & \mbx{They get absorbed with probability 0} }$$

  • Even though \(\kt{v}, \kt{v^\perp}\) are technically complex vectors, the particular ones so far are real-valued, which will allow us to draw (using real coordinates):

  • Let's work through measurement in the \(\theta=0\) case:
    • Here, \(\kt{v}=\kt{0},\kt{v^\perp}=\kt{1}\).
    • Now express the incoming photon \(\kt{0}\) in terms of the \(\theta=0\) filter's basis: $$ \mbox{photon } \;\; \kt{0} \eql 1. \kt{0} + 0 \kt{1} \;\;\;\;\; \mbox{Measurement basis on right side} $$
    • Then, the outcome state is \(\kt{v}=\kt{0}\) with probability 1.
    • Thus, for vertical photons, $$\eqb{ \mbox{With filter} \;\; \kt{v} = \kt{0} & \Rightarrow & \mbox{Pr[pass through]=1} \\ \mbox{With filter} \;\; \kt{v^\perp} = \kt{1} & \Rightarrow & \mbox{Pr[absorb]=0} \\ }$$

  • Similarly, if \(\theta=\frac{\pi}{2}\) (horizontal filter), then one can depict \(\kt{v}, \kt{v^\perp}\) as:

    In this case, the filter basis is: $$\eqb{ \kt{v} & \eql & \kt{1} \\ \kt{v^\perp} & \eql & \kt{1} \\ }$$ And so, for vertical photons: $$\eqb{ \mbox{With filter} \;\; \kt{v} = \kt{1} & \Rightarrow & \mbox{Pr[pass through]=0} \\ \mbox{With filter} \;\; \kt{v^\perp} = \kt{0} & \Rightarrow & \mbox{Pr[absorb]=1} \\ }$$

  • One expects a high probability of pass-through for angles near \(\theta=0\), and low probability for angles near \(\theta=\frac{\pi}{2}\).

  • So, for a generic \(\theta\), we can use the (real-coordinate) geometry to intuit \(\kt{v}, \kt{v^\perp}\) in terms of \(\kt{0}, \kt{1}\):

    From the geometry: $$\eqb{ \kt{v} & \eql & \cos\theta\kt{0} + \sin\theta\kt{1} & \equiv & \mbox{Pass through} \\ \kt{v^\perp} & \eql & \sin\theta\kt{0} - \cos\theta\kt{1} & \equiv & \mbox{Absorb} \\ }$$

  • Aside: $$ \kt{v^\perp} \eql -\sin\theta\kt{0} + \cos\theta\kt{1} $$ would work just as well.

  • Let's confirm orthonormality by recalling the general case: $$ \kt{v} \eql \alpha\kt{0} + \beta\kt{1} $$ Then if $$ \kt{v^\perp} \eql \beta^*\kt{0} - \alpha^*\kt{1} $$ both \(\kt{v}\) and \(\kt{v^\perp}\) are an orthonormal basis because $$ \inr{v^\perp}{v} \eql \inrh{ \beta^*\kt{0} - \alpha^*\kt{1} }{ \alpha\kt{0} + \beta\kt{1} } \eql \beta\alpha - \alpha\beta \eql 0 $$ and their lengths are 1.

  • Thus, we see that with $$\eqb{ \kt{v} & \eql & \cos\theta\kt{0} + \sin\theta\kt{1}\\ \kt{v^\perp} & \eql & \sin\theta\kt{0} - \cos\theta\kt{1} }$$ we get $$ \inr{v^\perp}{v} \eql \inrh{ \sin\theta\kt{0} - \cos\theta\kt{1} }{ \cos\theta\kt{0} + \sin\theta\kt{1} } \eql \sin\theta\cos\theta - \cos\theta\sin\theta \eql 0 $$ And their lengths are $$ \cos^2\theta + \sin^2\theta \eql 1 $$

  • Let's work through with \(\theta = \frac{\pi}{4}\):
    • Then, $$\eqb{ \kt{v} & \eql & \cos\theta\kt{0} + \sin\theta\kt{1} & \eql & \isqts{1} \kt{0} + \isqts{1} \kt{1} & \eql & \kt{+}\\ \kt{v^\perp} & \eql & \sin\theta\kt{0} - \cos\theta\kt{1} & \eql & \isqts{1} \kt{0} - \isqts{1} \kt{1} & \eql & \kt{-}\\ }$$
    • This is the Hadamard basis, which we've seen has probability \(\frac{1}{2}\) for letting a vertically polarized photon pass.

  • So far so good, but we need to establish that the intuitive approach above is actually correct.
 

Let's work out the the probability of vertical-photon pass-through for general \(\theta\):

  • Much of the delicate arguments that follow measurement will depend on this:
    • We're going to identify the probability that their measurements have the same pass-or-absorb outcomes.
    • That is, when do Alice and Bob see the same outcome: both see "pass" or both see "absorb"?
    • We will need to tease this out of measurement.
    We'll follow our projector approach to measurement.

  • The starting point is the measurement basis when the filter is at an angle \(\theta\): $$\eqb{ \kt{v} & \eql & \cos\theta\kt{0} + \sin\theta\kt{1}\\ \kt{v^\perp} & \eql & \sin\theta\kt{0} - \cos\theta\kt{1} \\ }$$

  • Define the projector $$\eqb{ P_v & \eql & \otr{v}{v} \\ & \eql & \otrh{ \cos\theta\kt{0} + \sin\theta\kt{1} }{ \cos\theta\br{0} + \sin\theta\br{1} } \\ & \eql & \cos^2\theta \otr{0}{0} + \sin\theta\cos\theta \otr{0}{1} + \sin\theta\cos\theta \otr{1}{0} + \sin^2 \theta \otr{1}{1} }$$

  • Now apply the projector to the incoming \(\kt{0}\) photon: $$\eqb{ P_v \kt{0} & \eql & \parenl{ \cos^2\theta \otr{0}{0} + \sin\theta\cos\theta \otr{0}{1} + \sin\theta\cos\theta \otr{1}{0} + \sin^2\theta \otr{1}{1} } \, \kt{0} \\ & \eql & \cos^2\theta \kt{0} + \sin\theta\cos\theta \kt{1} }$$

  • Thus, the probability of pass through is $$ \swich{0}{P_v}{0} \eql \cos^2\theta $$ which if plotted looks like:

    This agrees with experiment, validating the theory.

 

In-Class Exercise 1: Calculate \(\magsq{P_v\kt{0}}\) directly and show that it leads to the same result.
 

Now let's go back to the entangled Alice-Bob set up:

  • Since they each use a different angle, we'll distinguish the two measurement bases.

  • Alice's basis is: $$\eqb{ \kt{v_1} & \eql & \cos\theta_1\kt{0} + \sin\theta_1\kt{1} & \mbx{Pass} \\ \kt{v_1^\perp} & \eql & \sin\theta_1\kt{0} - \cos\theta_1\kt{1} & \mbx{Absorb} \\ }$$

  • And Bob's is: $$\eqb{ \kt{v_2} & \eql & \cos\theta_2\kt{0} + \sin\theta_2\kt{1} & \mbx{Pass} \\ \kt{v_2^\perp} & \eql & \sin\theta_2\kt{0} - \cos\theta_2\kt{1} & \mbx{Absorb} \\ }$$

  • Now let's apply projective measurement, starting with Alice:
    • Alice's measurement has projectors $$\eqb{ P_{v_1} & \eql & \otr{v_1}{v_1} & \mbx{Pass} \\ P_{v_1^\perp} & \eql & \otr{v_1^\perp}{v_1^\perp} & \mbx{Absorb} \\ }$$
    • When applied, the resulting 2-qubit measurement projectors are \(P_{v_1} \otimes I\) and \(P_{v_1^\perp} \otimes I\).

  • Similarly, Bob's measurement is described by the two projectors \(I \otimes P_{v_2}\) and \(I \otimes P_{v_2^\perp}\) where $$\eqb{ P_{v_2} & \eql & \otr{v_2}{v_2} & \mbx{Pass} \\ P_{v_2^\perp} & \eql & \otr{v_2^\perp}{v_2^\perp} & \mbx{Absorb} \\ }$$

  • Thus, in combination, the four possibilities are:
    1. Both pass (P-P): $$ \parenl{ I \otimes P_{v_2}} \parenl{ P_{v_1} \otimes I } \eql P_{v_1} \otimes P_{v_2} $$
    2. Both are absorbed (A-A): $$ \parenl{ I \otimes P_{v_2^\perp}} \parenl{ P_{v_1^\perp} \otimes I } \eql P_{v_1^\perp} \otimes P_{v_2^\perp} $$
    3. Alice's photon is absorbed, Bob's passes (A-P): $$ \parenl{ I \otimes P_{v_2}} \parenl{ P_{v_1^\perp} \otimes I } \eql P_{v_1^\perp} \otimes P_{v_2} $$
    4. Alice's passes, Bob's is absorbed (P-A): $$ \parenl{ I \otimes P_{v_2^\perp}} \parenl{ P_{v_1} \otimes I } \eql P_{v_1} \otimes P_{v_2^\perp} $$

  • For any 2-qubit state \(\ksi\), the probabilities of these four events are:
    1. Both pass: \(\swichh{\psi}{P_{v_1} \otimes P_{v_2}}{\psi}\)
    2. Both absorb: \(\swichh{\psi}{P_{v_1^\perp} \otimes P_{v_2^\perp}}{\psi}\)
    3. Pass-absorb: \(\swichh{\psi}{P_{v_1} \otimes P_{v_2^\perp}}{\psi}\)
    4. Absorb-pass: \(\swichh{\psi}{P_{v_1^\perp} \otimes P_{v_2}}{\psi}\)

  • Finally, we can now focus on the probability that both see the same result: $$\eqb{ \mbox{Pr[Same result]} & \eql & \mbox{Pr[both pass]} \; + \; \mbox{Pr[both are absorbed]} \\ & \eql & \swichh{\psi}{P_{v_1} \otimes P_{v_2}}{\psi} \; + \; \swichh{\psi}{P_{v_1^\perp} \otimes P_{v_2^\perp}}{\psi} }$$

  • What remains: doing the calculation.

  • The calculation shows that $$ \mbox{Pr[Same result]} \eql \cos^2 \parenl{\theta_1 - \theta_2} $$ Or the other way around: \(\cos^2 (\theta_2 - \theta_1)\)
 

In-Class Exercise 2: One of the solved problems shows how to calculate the result for \(\swich{\psi}{P_{v_1} \otimes P_{v_2}}{\psi}\). Show how the same approach can be used to calculate \(\swich{\psi}{P_{v_1^\perp} \otimes P_{v_2^\perp}}{\psi}\), writing out all 16 terms (just once) for clarity. Then, complete the steps in showing \(\mbox{Pr[Same result]} = \cos^2 (\theta_2 - \theta_1)\).
 

The result comports with intuition:

  • When \(\theta_1=0\), Alice's measurement is the standard basis \(\kt{0},\kt{1}\).

  • In this case, the 2-qubit outcome will be one of \(\kt{00},\kt{11}\).

  • Because of this, Bob's measurement at angle \(\theta_2\) will apply to his qubit which will be in one of \(\kt{0},\kt{1}\).

  • Recall that we derived the pass-through probability as \(\cos^2\theta_2\).

  • This is exactly what's predicted: when \(\theta_1=0\), $$ \cos^2 (\theta_1 - \theta_2) \eql \cos^2 \theta_2 $$

  • When the angles are close, \(\theta_1 \approx \theta_2\) we should get high probability of agreement, which is indeed the case: $$ \cos^2 (\theta_1 - \theta_2) \approx \cos^2 0 \eql 1 $$
 

In-Class Exercise 3: What is the probability of agreement when:

  1. \(\theta_1=-\frac{\pi}{3}, \theta_2=0\)?
  2. \(\theta_1=\frac{\pi}{3}, \theta_2=-\frac{\pi}{3}\)?
 


9.4   Hidden variables, part II: what quantum mechanics predicts

 

With that background, we now move to the interesting part of the experiment:

  • Alice and Bob each choose their angle randomly from amongst \(-60^\circ, 0, 60^\circ\).

  • There are 9 possible cases: $$ \begin{array}{|c|c|}\hline \mbox{Alice: } \theta_1 & \mbox{ Bob: } \theta_2 \\\hline 60^\circ & 60^\circ \\ 0^\circ & 0^\circ \\ -60^\circ & -60^\circ \\ -60^\circ & 60^\circ \\ -60^\circ & 0^\circ \\ 0^\circ & 60^\circ \\ 0^\circ & -60^\circ \\ 60^\circ & 0^\circ \\ 60^\circ & -60^\circ \\\hline \end{array} $$

  • The probability of agreement in each of the 9 cases is given by: \begin{array}{|c|c|}\hline \mbox{Alice: } \theta_1 & \mbox{ Bob: } \theta_2 & \theta_1 - \theta_2 & \cos(\theta_1 - \theta_2) & \cos^2(\theta_1 - \theta_2) \\\hline 60^\circ & 60^\circ & 0 & 1 & 1\\ 0^\circ & 0^\circ & 0 & 1 & 1\\ -60^\circ & -60^\circ & 0 & 1 & 1\\ -60^\circ & 60^\circ & -120^\circ & \frac{1}{2} & \frac{1}{4}\\ -60^\circ & 0^\circ & -60^\circ & \frac{1}{2} & \frac{1}{4}\\ 0^\circ & 60^\circ & -60^\circ & \frac{1}{2} & \frac{1}{4}\\ 0^\circ & -60^\circ & 60^\circ & \frac{1}{2} & \frac{1}{4}\\ 60^\circ & 0^\circ & 60^\circ & \frac{1}{2} & \frac{1}{4}\\ 60^\circ & -60^\circ & 120^\circ & \frac{1}{2} & \frac{1}{4}\\\hline \end{array}

  • We can summarize this as: $$ \mbox{Pr[pass/absorb outcomes agree]} \eql \left\{ \begin{array}{cc} 1 & \;\;\;\;\; \theta_1 = \theta_2 \\ \frac{1}{4} & \;\;\;\;\; \theta_1 \neq \theta_2 \\ \end{array} \right. $$

  • Since Alice and Bob pick their angles randomly, any of the 9 configurations above is equally likely to occur.

  • Thus, the probability of agreement is: $$\eqb{ \mbox{Pr[agree]} & \eql & \mbox{Pr[agree|\(\theta_1=\theta_2\)]} \, \mbox{Pr[\(\theta_1=\theta_2\)]} \; + \; \mbox{Pr[agree|\(\theta_1\neq\theta_2\)]} \, \mbox{Pr[\(\theta_1\neq\theta_2\)]} \\ & \eql & 1 \cdot \smf{3}{9} \; + \; \smf{1}{4} \cdot \smf{6}{9} \\ & \eql & \smf{1}{2} }$$
 


9.5   Hidden variables, part II: What hidden variables predict

 

We now suppose each photon in an entangled pair carries in it some local hidden state, inaccessible to quantum theory:

  • If this is true, then the following must be true:
    • Since the photons are identical, it must be the same variable in each but with possibly different values.
    • The action of pass/absorb must depend on this variable, otherwise we're just left with quantum theory.
    • The action is local, by assumption.

  • Let's look a little more closely at the locality assumption:
    • If locality is true, then the value of the carried variable at Alice can have no influence on what occurs at Bob's end.
      (And vice-versa.)
    • Also, if the value happens to be the same for each photon, we should observe the same outcome (whether pass or absorb)

  • Next, there is another assumption that comes with buying into local-hidden-variables:
    • The entangled pair is the Bell state \(\isqt{1} \parenl{ \kt{00} + \kt{11} }\)
    • If we simply switched the photons sent to Alice and Bob, we should see the same statistical results.
    • That is, send the "right" photon to Alice, the "left" to Bob.
    • The photons are considered identical, at least in reacting to a filter.

  • The consequence of "identical reaction" is:
    • The hidden states that travel with the two photons are identical.
    • Thus, both should react the same way (pass or absorb) when encountering the same polarizer angle.
 

Now let's analyze what happens with hidden variables:

  • First, consider the possibilities when a hidden variable "encounters" a filter set at one of the three angles.

  • For any given angle, a hidden variable produces either "pass" or "absorb".

  • Thus, for the three angle choices, a particular hidden variable must have three outcomes from one of the rows in: $$ \begin{array}{|c|c|c|}\hline -60^\circ & 0 & 60^\circ \\\hline P & P & P \\ {\bf P} & {\bf P} & {\bf A} \\ P & A & P \\ P & A & A \\ A & P & P \\ A & P & A \\ A & A & P \\ A & A & A \\\hline \end{array} $$ where P = pass, and A = absorb.
    • For example, consider the 2nd row: P P A.
    • This is saying that the hidden variable happens to have the effect:
      • Pass when \(\theta=-60^\circ\)
      • Pass when \(\theta=0^\circ\)
      • Absorb when \(\theta=60^\circ\)

  • To summarize: whatever the particular state of the hidden variable (or variables), the outcomes will be one of the rows in the above table.

  • Then, we could call a hidden state a PPA-state if it has the effect described in the second row (PPA).

  • So, there are 8 types of pairs: PPP, PPA, PAP, PAA, APP, APA, AAP, AAA.

  • In any particular experiment, one of these types will arrive at Alice and Bob.

  • Now let's return to the 9 possible angle configurations and focus only on the PPA state.

  • We'll ask the question: what happens when two identical PPA photons encounter randomly chosen angles from \(-60^\circ, 0, 60^\circ\)?

  • First, let's note what identical-PPA means: $$ \begin{array}{|l|c|c|c|}\hline \; & -60^\circ & 0 & 60^\circ \\\hline \mbox{Alice's photon} & P & P & A \\ \mbox{Bob's photon} & P & P & A \\\hline \end{array} $$ For example, suppose Alice has set her angle \(\theta_1=0\) and Bob has set \(\theta_2=60^\circ\):
    • The left (Alice's) PPA photon will pass.
    • The right (Bob's) PPA photon will be absorbed.

  • We can describe every one of the possible outcomes in a table that shows all the 9 angle settings: $$ \begin{array}{|c|c|}\hline \mbox{Alice: } \theta_1 & \mbox{ Bob: } \theta_2 & \mbox{Alice's photon} & \mbox{Bob's photon} & \mbox{Agreement} \\\hline 60^\circ & 60^\circ & A & A & agree \\ 0^\circ & 0^\circ & P & P & agree \\ -60^\circ & -60^\circ & P & P & agree \\ -60^\circ & 60^\circ & P & A & \\ -60^\circ & 0^\circ & P & P & agree \\ 0^\circ & 60^\circ & P & A & \\ 0^\circ & -60^\circ & P & P & agree \\ 60^\circ & 0^\circ & A & P \\ 60^\circ & -60^\circ & A & P \\\hline \end{array} $$ So, a PPA pair of photons will agree 5/9-ths of the time.

  • We can work out such tables for all 8 types of photons.
 

In-Class Exercise 4: Write out the table for APA photons (when both are APA) and for AAA photons (when both are AAA).
 

Let's continue the analysis:

  • For PPP and AAA pairs, the probability of agreement is 1.

  • For all others, the probability of agreement is \(\frac{5}{9}\).

  • Because these are hidden variables, we cannot make any assumptions about which states occur more frequently and how they occur.

  • But we can state confidently that: the probability of agreement is at least \(\frac{5}{9}\).

  • So, now we have two theories with differing predictions of photon-action-agreement (pass or absorb):
    • Quantum theory: $$ \mbox{Pr[agree]} \eql \smf{1}{2} $$
    • Hidden-variable theory: $$ \mbox{Pr[agree]} \geq \smf{5}{9} $$

  • The latter is an example of a Bell Inequality.
    (There are many such experimental designs, each with their own Bell Inequality.)

  • So, who's right?

  • The only way to decide: multiple carefully conducted experiments, repeated and analyzed.

  • In all cases tested so far, quantum theory has always been correct, with very high statistical accuracy.

  • That is, there is no known case where quantum theory has been wrong.
So, what then is quantum mechanics and how does it relate to what we've seen in quantum computing?
 


9.6   Towards the continuum, part 1: two types of infinity

 

Why do we need this extension to the continuum?

  • For discrete physical phenomena like Stern-Gerlach, the theory developed so far with linear algebra suffices.

  • But most physical quantities of interest are modeled by real numbers, such as position and momentum.
 

First, some notational clarification:

  • We have been using a subscript to both identify vector components and index a collection of vectors, as in $$ \kt{w} \eql \sum_i c_i \kt{v_i} $$
    • Thus, for example if $$ \kt{w} \eql c_1 \kt{v_1} + c_2 \kt{v_2} + c_3 \kt{v_3} $$ then in the v-basis $$ \kt{w} \eql \mat{c_1 \\ c_2 \\ c_3} $$
    • Thus \(c_i\) (as in \(c_1, c_2, c_3\)) refer to coefficients in a linear combination
    • Or, equivalently, vector components when expressing \(\kt{w}\) in the v-basis.
    • And with \(\kt{v_i}\), the subscript \(i\) refers to which vector in the collection.

  • This dual-use of subscripts did not matter in the discrete case because \(n\)-component vectors need an \(n\)-vector basis.

  • But in the continuous case, it will be different.

  • Let's rewrite the discrete version to separate out components from collections:
    • We'll write $$ \kt{w} \eql c(1) \kt{v_1} + c(2) \kt{v_2} + c(3) \kt{v_3} $$
    • Here, we use parentheses for components, and subscripts for indexing something from a collection: $$\eqb{ c(i) & \eql & \mbox{\(i\)-th component of } \kt{w} & \eql & \mbox{\(i\)-th coefficient in linear-comb }\\ v_i & \eql & \mbox{\(i\)-th vector in a basis } \kt{v_1},\ldots,\kt{v_n} & & \\ }$$
    • This way, we can even refer to the components of each \(\kt{v_i}\) as in: $$ w(1) \eql c(1) v_1(1) + c(2) v_2(1) + c(3) v_2(1) $$ where $$ v_i(j) \eql \mbox{\(j\)-th component of vector} \kt{v_i} \\ $$

  • With this slightly changed notation, let's point out:
    • We've written $$ \kt{w} \eql \sum_i c(i) \kt{v_i} $$ where $$ c(i) \eql \inr{v_i}{w} $$
    • Recall: $$ \kt{w} \eql \sum_i \inr{v_i}{w} \, \kt{v_i} $$

  • Let's recall our standard inner-product with our new notation:
    • Let $$\eqb{ \kt{w} & \eql & \sum_i c(i) \kt{v_i} \\ \kt{u} & \eql & \sum_i b(i) \kt{v_i} \\ }$$ be two vectors expressed in the v-basis.
    • Then, $$ \inr{w}{u} \eql \inrh{ \sum_i c(i) \kt{v_i} }{ \sum_i b(i) \kt{v_i} } \eql \sum_i c^*(i) b(i) $$ as expected (See Module-2 solved problems.)

  • Next, for an orthonormal basis \(\kt{v_1},\ldots,\kt{v_n}\) $$ \inr{v_i}{v_j} \eql \left\{ \begin{array}{cc} 1, & \;\;\;\; i=j \\ 0, & \;\;\;\; i\neq j \\ \end{array} \right. $$ We'll invent functions \(\delta_i(j)\) where $$ \delta_i(j) \eql \left\{ \begin{array}{cc} 1, & \;\;\;\; i=j \\ 0, & \;\;\;\; i\neq j \\ \end{array} \right. $$ and write $$ \inr{v_i}{v_j} \eql \delta_i(j) $$

  • Notice that the \(\delta_i(j)\) are also vectors:
    • For example, \(\delta_3()\) is $$ (0,0,1,0, \ldots, 0) $$
    • That is, only \(\delta_3(3)=1\). The rest are \(0\)s.

  • If an operator \(A\) has an eigenbasis \(\kt{\phi_i}\) with eigenvalues \(\lambda(i)\), then we know two things:
    1. The action of \(A\) on its own eigenvectors: $$ A \kt{\phi_i} \eql \lambda(i) \kt{\phi_i} $$
    2. Any vector can be expressed in this eigenbasis: $$ \kt{w} \eql \sum_i \alpha(i) \kt{\phi_i} $$
 

With this (slightly changed) notation, let's first look at countably infinite vectors:

  • Here, a countably infinite vector has components labeled by the integers: $$ \kt{w} \eql \mat{c(1) \\ c(2) \\ \vdots \\ c(i) \\ \vdots } $$

  • And a basis will have a countably infinite collection of vectors, where expressing in terms of such a basis looks like: $$ \kt{w} \eql \sum_{i=1}^\infty c(i) \kt{v_i} $$

  • An operator is expressed as an infinite dimensional matrix (both rows and columns): $$ A\kt{w} \eql \mat{ a_{11} & a_{12} & \ldots \\ a_{21} & a_{22} & \ldots \\ \vdots & \ddots & \ldots } \mat{c(1) \\ c(2) \\ \vdots } $$

  • One problem we need to worry about: do sums converge?

  • The theory is developed in a way to restrict attention to vectors (and operators) where sums are convergent, as in $$ \kt{w} \eql \sum_{i=1}^\infty c(i) \kt{v_i} \; \lt \; \infty $$

  • With that restriction, just about everything we've seen in finite discrete linear algebra applies.

  • For example, standard basis vectors are $$\eqb{ \kt{v_1} & \eql & (1, 0, 0, \ldots) \\ \kt{v_2} & \eql & (0, 1, 0, \ldots) \\ \kt{v_3} & \eql & (0, 0, 1, \ldots) \\ \vdots & & \\ }$$

  • However, this type of infinity has found limited use.

  • Variables of interest in modeling physical phenomena tend to be real valued and therefore continuous.
 


9.7   Towards the continuum, part 2: what do we need from a continuous linear algebra?

 

Let's examine a few core concepts from discrete linear algebra and ask what they would look like in the continuous domain.

We'll state these in the form of puzzles to resolve:

  • Puzzle #1: what do linear combinations look like?
    • In the discrete case, we wrote $$ \kt{w} \eql \sum_i c(i) \kt{v_i} $$
    • In the continuous case, we'd expect a sum to be replaced by an integral, along the lines of $$ \kt{w} \eql \int_x c(x) \kt{v_x} dx $$
    • However, what does it mean to integrate the vector \(\kt{v_x}\)?
    • And what is a continuous vector \(\kt{v_x}\) anyway?
    • What do its components look like?

  • Puzzle #2: what do inner products look like?
    • If we've solved the first puzzle and it's true that $$\eqb{ \kt{w} & \eql & \int_x c(x) \kt{v_x} dx \\ \kt{u} & \eql & \int_x b(x) \kt{v_x} dx \\ }$$ then is $$ \inr{w}{u} \eql \int_x c^*(x) b(x) dx $$ analogous to the discrete case where $$ \inr{w}{u} \eql \sum_i c^*(i) b(i)? $$

  • Puzzle #3: What does orthonormality look like?
    • We've described orthonormality as $$ \inr{v_i}{v_j} \eql \left\{ \begin{array}{cc} 1, & \;\;\;\; i=j \\ 0, & \;\;\;\; i\neq j \\ \end{array} \right. $$ in the discrete case.
    • Is the continuous equivalent to the following, $$ \inr{v_x}{v_y} \eql \left\{ \begin{array}{cc} 1, & \;\;\;\; x=y \\ 0, & \;\;\;\; x\neq y \\ \end{array} \right. $$ where \(x,y\) are real numbers?

  • Puzzle #4: How do probabilities emerge from coefficients, and how does unit length play a role?
    • For example, is $$ \inr{w}{w} \eql \int_x c^*(x) c(x) dx \eql 1? $$
    • How does one calculate probabilities?

  • Puzzle #5: What do operators and their eigen-stuff look like?
    • Will we have the continuous equivalent $$ A \kt{\phi_x} \eql \lambda(x) \kt{\phi_x}? $$
    • And will any vector be expressible in the continuous basis \(\kt{\phi_x}\)?

  • Puzzle #6: How do outer-products and projectors work in the continuous case?

  • Puzzle #7: How do tensor products work in the continuous case?
 


9.8   Towards the continuum, part 3: some aspects of continuous linear algebra

 

We'll now examine how some of the above puzzles are resolved.

Before that, let's step back and recall the difference between a vector and its numerical realization:

  • With real-vectors, a vector like \(\kt{w}\) is an "arrow":
    • One can add arrows (via parallelogram) and scale them (stretch).
    • However, to be useful and to compute things like inner-products, we need to "numerify" vectors.

  • The first step in numerification is to choose a basis, \(\kt{v_1},\ldots,\kt{v_n}\), preferably orthonormal.
        \(\rhd\) This part is still abstract because the \(\kt{v_i}\)'s are arrows.

  • Then, express a vector in terms of this basis: $$ \kt{w} \eql \sum_i c(i) \kt{v_i} $$

  • Then, the numbers \(c(i)\) are the numerical realization of the abstract vector \(\kt{w}\): $$ \mbxx{(abstract)} \;\;\;\; \kt{w} \eql \mat{c(1) \\ c(2) \\ \vdots \\ c(n) } \;\;\;\; \mbxx{(concrete - with numbers)} $$

  • One can now define and calculate the inner product, apply operators, and so on.

  • Any basis vector when expressed in the same basis looks like a standard-basis vector, for example $$ \kt{v_2} \eql 0\cdot \kt{v_1} + 1\cdot\kt{v_2} + 0\cdot\kt{v_3} + \ldots \eql \mat{0\\ 1\\ 0\\ \vdots} $$

  • However, "numerification" occurs in different ways for different mathematical objects.

  • For example, consider \(\mathbb{M}_{m\times n}\), the set of all matrices with \(m\) rows and \(n\) columns:
    • Let \(D_{ij}\) be a matrix with all zeroes except for a 1 in row \(i\), column \(j\).
    • Then any matrix \(A \in \mathbb{M}_{m\times n}\) can be expressed in terms of this basis: $$ A \eql \sum_{ij} a_{ij} D_{ij} $$
    • For example: $$ \mat{2 & 3\\ 4 & 0} \eql 2\mat{1 & 0\\0 & 0} + 3\mat{0 & 1\\0 & 0} + 4\mat{0 & 0\\1 & 0} $$
    • In this way, any matrix \(A\) can be treated as an abstract vector \(\kt{A}\).

  • A slightly more involved but very useful example comes from treating the set of "well behaved" real functions on the interval \([0,T]\) as a vector space:
    • Treat each function as a vector.
    • Define as a basis, the countably infinite collection of functions $$ \setl{ \sin\parenl{ \smm{\frac{2\pi n x}{T}}}, \cos\parenl{\smm{\frac{2\pi n x}{T}}} } $$ (The constants come from fitting the trig functions into the interval \([0,T]\))
    • Here, the basis functions are mutually orthogonal when the inner product of two functions is defined as the integral $$ \inr{f}{g} \eql c \int_{x\in T} f(x)g(x) dx $$ where \(c\) is a constant to enforce the unit-length part of orthonormality.
    • Then, any function \(f\) in this space can be written as the countably infinite sum: $$ f(x) \eql \sum_{n=0}^\infty a_n \sin\parenl{ \smm{\frac{2\pi n x}{T}} } + b_n \cos\parenl{ \smm{\frac{2\pi n x}{T}} } $$
    • Note: the vectors are continuous but the basis is countably infinite.

    This is the famous Fourier series expansion of a function, the first of the three Fouriers.

 

Now let's return to the continuous case:

  • Analogous to the discrete case, we "numerify" a vector with a continuous collection of coefficents \(c(x)\):
    • Since \(x\) is real-valued, \(c(x)\) is just our familiar function.
    • Since coefficients are complex, the output of the function is a complex number.
    • Let's contrast with the discrete case:

  • In this case:
    • We'll have a set of basis vectors indexed by real numbers: \(\kt{v_x}\).
    • And coefficients \(c(x)\) such that a vector is expressed in this basis using an integral. $$ \kt{w} \eql \int_x c(x) \kt{v_x} dx $$
    • And an inner product for two "numerified" vectors is simply $$ \inr{w}{u} \eql \int_x c^*(x) b(x) dx $$ where \(\kt{w}\) is numerified by the function \(c(x)\) and \(\kt{u}\) by \(b(x)\).
    • When doing so, one scales by appropriate constants to maintain unit-lengths.
 

Next, let's examine the connection between the coefficient function \(c(x)\) and inner products:

  • In the discrete case: $$\eqb{ \inr{v_i}{w} & \eql & c(i) \\ & \eql & \sum_j \delta_i(j) c(j) }$$ where $$ \delta_i(j) \eql \left\{ \begin{array}{cc} 1, & \;\;\;\; i=j \\ 0, & \;\;\;\; i\neq j \\ \end{array} \right. $$ are functions we defined to "pick off" the \(i\)-th coefficient.

  • For example:
    • Suppose we want to pick off \(c(2)\).
    • Then $$\eqb{ \delta_2(1) & \eql & 0 \\ \delta_2(2) & \eql & 1 \\ \delta_2(3) & \eql & 0 \\ }$$ so that $$\eqb{ \delta_2(1) \cdot c(1) + \delta_2(2) \cdot c(2) + \delta_2(3) \cdot c(3) & \eql & 0\cdot c(1) + 1 \cdot c(2) + 0\cdot c(3) \\ & \eql & c(2) }$$

  • Therefore, what's needed in the continuous case is an equivalent \(\delta\) function with the property $$ c(x) \eql \int_y \delta_x(y) c(y) dy $$ where integrating the product \(\delta_x(y) c(y)\) picks off \(c(x)\).

  • For the moment, let's assume such a function exists:
    • We don't really need to know what this function is.
    • Just how it acts inside integrals like the one above.
    • Here, it "picks off" the coefficient \(c(x)\) from the continuum \(c(y)\) via the integral.

  • This is called the Dirac Delta function, which turns out not to be a "function" in the usual sense:
    • There are a variety of such "functions".
    • Each is represented as a collection dependent on some parameter that controls their "size" in the limit.
    • For example: $$ \delta_0^{(n)}(x) \eql \smf{n}{\sqrt{\pi}} e^{-n^2x^2} $$ The integral of each is 1 but as \(n\to\infty\), the shape approaches that of a "spike", as in

    • We have written a generic such function as \(\delta_x(y)\) where \(y\) is a variable and spike is centered at \(x\).
    • It is often written as \(\delta(y-x)\).

  • One implication of using this function:
    • The inner products of basis vectors in the discrete case is: $$ \inr{v_i}{v_j} \eql \delta_i(j) $$
    • For discrete vectors, \(\delta_i(j)\) is either 0 or 1.
    • In the continuous case: $$ \inr{v_x}{v_y} \eql \delta_x(y) $$

  • To see why this integral-use approach is needed, suppose we were to define \(\delta_x(y) = 1\) only when \(x=y\) and \(\delta_x(y) = 0\) otherwise:
    • This is a function that has a spike of height \(1\) at \(x\).
    • And it's \(0\) everywhere else.
    • Unfortunately, this leads to a technical difficulty with the "picking off" part: $$\eqb{ c(x) & \eql & \int_y \delta_x(y) c(y) dy\\ & \eql & \int_{y\neq x} \delta_x(y) c(y) dy \; + \; \int_{y=x} \delta_x(y) c(y) dy \\ & \eql & \int_{y\neq x} 0\cdot c(y) dy \; + \; \int_{y=x} 1\cdot c(y) dy \\ & \eql & 0 }$$

  • The technical difficulty can be resolved by introducing a "generalized function" with certain properties, which we won't go into here, to make the integral work, i.e., so that $$ c(x) \eql \int_y \delta_x(y) c(y) dy $$

  • From our point of view, all we need to know is how to apply \(\delta_x(y)\) inside an integral.
 

Before tackling operators, let's simplify notation:

  • We've used \(c(x)\) to numerify a continuous vector \(\kt{w}\).

  • Instead of using two symbols, we'll simply write $$ \kt{w} \eql \kt{c(x)} $$ and dispense with the former.

  • Thus, if \(\psi(x)\) is a function that represents coefficients, then the corresponding vector is \(\kt{\psi(x)}\).
 

Let's now review what we have so far:

  • Puzzle #1: what do linear combinations look like? We do in fact have $$ \kt{w} \eql \int_x c(x) \kt{v_x} dx $$

  • Puzzle #2: what do inner products look like? As anticipated, we have $$ \inr{w}{u} \eql \int_x c^*(x) b(x) dx $$ with the appropriate scaling (normalization) to maintain unit-length.

  • Puzzle #3: What does orthonormality look like? We now have this all-purpose Delta "function" that let's us define $$ \inr{v_x}{v_y} \eql \delta_x(y) $$

  • Puzzle #4: How do probabilities emerge from coefficients?
    • In the discrete case, the squared-magnitude of a coefficient gave us the probability associated with the corresponding eigenvector.
    • For example, if we expressed a vector in terms of a measurement eigenbasis $$ \ksi \eql \sum_i \alpha(i) \kt{\phi_i} $$ then the probability of seeing \(\kt{\phi_i}\) is \(\magsq{\alpha(i)} = \alpha^*(i)\alpha(i)\).
    • If in the continuous case, the coefficient function when expressing \(\ksi\) in terms of an eigenbasis is $$ \ksi \eql \int_x \alpha(x) \kt{\phi(x)} dx $$ then the probabilities arise through $$ \mbox{Pr[observe outcome in \([\phi(x), \phi(x+\epsilon)]\)]} \eql \int_x^{x+\epsilon} \alpha^*(y) \alpha(y) dy $$
    • That is, via integrating the probability density function implied by the eigenbasis.
    • Appropriate normalization is introduced so that $$ \int_y \alpha^*(y) \alpha(y) dy \eql 1 $$
 

Before examining operators, let's ask: what continuous basis should we use for calculations?

  • In the discrete case, it did not matter much:
    • For example, in the Stern-Gerlach experiment, a z-aligned apparatus uses the standard-basis.
    • This makes the x-aligned apparatus the H-basis.
    • But we could also use the S-basis for x-aligned, and H-basis for vertical.

  • The other aspect of linear algebra that did not matter was: eigenvalues:
    • We did not use this for quantum computing because we focused only on the states that were measurement outcomes.
    • But an actual experimental measurement also produces an eigenvalue.
    • In fact, the eigenvalue is the actual physically observable quantity.

  • In the continuous case, actual physical observables do matter both in theory and in practice.

  • The most intuitively fundamental physical observable is location: where in 3D space, an object is.

  • In one dimension (to simplify), this is an x-value (along the x-axis).

  • Thus, if a real number like \(x\) is an observed eigenvalue, what are the corresponding eigenvectors and operator?
    • Suppose we could find an operator \(A_{pos}\) (for position) and eigenvectors \(\kt{\phi_x}\) so that $$ A_{pos} \kt{\phi_x} \eql x \kt{\phi_x} $$
    • Then, the observed position \(x\) is the result (eigenvalue) of a position measurement, with resulting state \(\kt{\phi_x}\).

  • This should immediately raise a question:
    • Don't measurements result in probabilistic outcomes?
    • If so, does that mean the position of a quantum object is necessarily probabilistic, according to the theory?

  • This set of eigenvectors \(\kt{\phi_x}\) is the default starting basis called the position basis.

  • What do the eigenvectors \(\kt{\phi_x}\) look like? Do they look like the all-zeroes-but-one-1 eigenvectors we've seen in the discrete case?

  • The reasoning is a bit tricky:
    • We've already seen that we can't have spike=1 functions.
    • Consider two x-values \(x\) and \(x^\prime\).
    • Then, in the integral $$ \int_y \delta_{x^\prime}(y) \delta_x(y) dy $$ the first delta-function picks off the value \(x^\prime\) in the second, so $$ \int_y \delta_{x^\prime}(y) \delta_x(y) dy \eql \delta_x(x^\prime) $$
    • But, by our definition of inner-products for orthonormal vectors, $$ \inr{\phi_x}{\phi_{x^\prime}} \eql \delta_x(x^\prime) $$ Thus, the integral above (on the left) is in fact the inner product \(\inr{\phi_x}{\phi_{x^\prime}}\).
    • Which means the two functions in the integral are in fact the functions that define the eigenvectors: $$\eqb{ \phi_x(y) & \eql & \delta_x(y) & \mbx{eigenvector for \(x\)}\\ \phi_{x^\prime}(y) & \eql & \delta_{x^\prime}(y) & \mbx{eigenvector for \(x^\prime\)}\\ }$$
    • Thus, when the position eigenbasis is represented in its own basis, the corresponding coefficient functions are delta-functions.

  • If this all seems rather complicated, that's because it is!

  • None of this was fully worked out until Paul Dirac put the pieces together.
    • And, as it turns out, Dirac's version was not mathematically rigorous.
    • The formal rigorous underpinning would take years to address.

  • Finally, let's simplify notation once more:
    • We've used the continuous position basis \(\kt{\phi_x}\) indexed by position \(x\).
    • Again, instead of two symbols, one directly writes the \(\kt{x}\) as a basis vector.
    • Then, $$ A_{pos} \kt{x} \eql x \kt{x} $$ which at first is a bit confusing and takes some getting used to.
    • With this simplification, we can write $$ \inr{x}{\psi} \eql \int_y \delta_x^*(y) \psi(y) dy \eql \psi(x) $$ where the vector \(\psi(y)\) is the "numerification" of \(\kt{\psi}\) in the position basis.

  • This has a nice interpretation:
    • Recall that \(\inr{v}{w}\) is the coefficient of \(\kt{w}\) along \(\kt{v}\).
    • Then, in \( \inr{x}{\psi}\) we get the coefficient of \(\kt{\psi}\) along the x-axis, which is simply \(\psi(x)\).

  • And where did this \(\psi(x)\) come from?
    • We will see that \(\psi(x)\) is determined by Schrodinger's equation that, in some sense, describes the "physics" of the object of interest.
    • Because we have unit-lengths, we will normalize to enforce $$ \int_x \psi^*(x) \psi(x) dx \eql 1 $$
    • Which makes \(\psi^*(x) \psi(x) = \magsq{\psi(x)}\) a probability density function, from which we can calculate $$ \mbox{Pr[observe position in \([a, b]\)]} \eql \int_a^b \magsq{\psi(x)} dx $$

  • Let's summarize:
    • The preferred starting point is the position basis, whose vectors are denoted by \(\kt{x}\), one for each value of possible locations \(x\).
    • The actual observed eigenvalues are the locations \(x\).
    • This means a measurement Hermitian \(A_{pos}\) must have \(\kt{x}\) as eigenvectors and \(x\) as eigenvalues: $$ A_{pos} \kt{x} \eql x \kt{x} $$
    • The eigenvectors \(\kt{x}\) when numerified turn out to be the delta functions \(\delta_x(y)\).
 

Next, let's look at operators in general:

  • We need to address several questions at the very least:
    1. What is an operator in the continuous case and how does one "numerify" it? That is, how does one calculate $$ A \kt{\psi(x)} \eql ? $$ for an operator \(A\)?
    2. How does one determine the coefficients in expressing a vector in this eigenbasis $$ \kt{\psi(x)} \eql \int_{y} \alpha(y) \kt{\phi_y(x)} dy? $$
    3. What kinds of operators are useful?

  • Let's start with what an operator does:
    • When numerified, a vector \(\kt{\psi(x)}\) is a function \(\psi(x)\).
    • Thus, when numerified, an operator is something that acts on a function to produce a function.
    • What kinds of operators do this?
    • There are many, but certainly differentiation is one of them: $$ \frac{d}{dx} \psi(x) \eql g(x) \;\;\;\;\; \mbox{\(g(x)\) is some function} $$ where \(\frac{d}{dx}\) is an operator.

  • Just like the H-basis was the second important qubit basis, in the continuous vector world, the second important basis is the momentum basis.
    • A particle's position describes current location.
    • Momentum describes its current motion.

  • Again, momentum is a physically observable real number: let's call this \(p\):
    • Which means there is some Hermitian \(A_{mo}\) operator such that for any momentum eigenstate \(\kt{p}\): $$ A_{mo} \kt{p} \eql p \kt{p} $$
    • The question of course is: what are these momentum eigenstates when written in the position basis, and what is the operator?
    • Relatedly: can position be described in the momentum basis?

  • Without getting into the technical details, we'll state that each momentum eigenstate numerified (in the position basis) works out to the function $$ \kt{p} \eql \smf{1}{\sqrt{2\pi}} e^{\frac{ipx}{\hbar}} $$

  • We can now ask the question: suppose we have some generic state \(\kt{\psi(x)}\) in the x-basis, what does this look like in the momentum basis?
    • Let \(\tilde{\psi}(p)\) denote the coefficient function (of \(p\)) of the same state in the momentum basis.
    • Then one can show that $$\eqb{ \tilde{\psi}(p) & \eql & \smf{1}{\sqrt{2\pi}} \int e^{\frac{-ipx}{\hbar}} \, \psi(x) \, dx\\ \psi(x) & \eql & \smf{1}{\sqrt{2\pi}} \int e^{\frac{ipx}{\hbar}} \, \tilde{\psi}(p) \, dp\\ }$$
    • This allows going to-and-from one basis to another.
    • Notice: the calculation involves doing something to one function to produce the other.
    • This pair of actions is together called the Fourier transform, the second of the three Fouriers.

  • Lastly, we should ask: where did the exponentials come from?
    • The momentum operator turns out to be the differential operator \(-i\hbar \frac{d}{dx}\).
    • When applied to the eigenvector-eigenvalue equation, that equation becomes a differential equation whose solution is exponential.
 

Other puzzles we haven't resolved but won't have time for:

  • Outer products and projectors.
  • Tensor products.
 

The fact that momentum is involved suggests that something is "moving" or changing with time:

  • If the state is changing with time, we need an additional variable for time.

  • Let $$\eqb{ \kt{\psi(x,0)} & \eql & \mbox{state at time 0} \\ \kt{\psi(x,t)} & \eql & \mbox{state at time t} \\ }$$

  • Now, what used to be the coefficient function \(\psi(x)\) has an additional variable \(t\) to denote time: \(\psi(x,t)\)

  • This raises the question: what determines how the state changes with \(t\)?
 


9.9   Towards the continuum, part 4: postulates of quantum mechanics

 

About postulates:

  • They're called postulates because there is no underlying theory from which these postulates can be derived.

  • In some sense, they are assumptions from which the rest of the theory is developed.

  • Interestingly, there is a bit of divergence amongst physicists in how best to state the postulates.

  • Now, one can describe the postulates with or without relativity.
        \(\rhd\) We will describe the (simpler) non-relativistic postulates.

  • Another issue is whether the system under consideration is isolated from external influences.
        \(\rhd\) We'll assume isolation.

  • We'll state them for the one-dimensional case and then comment.
 

The (non-relativistic) postulates of quantum mechanics:

  1. State description postulate. The state of an isolated quantum-mechanical system is described by a continuous complex vector \(\kt{\psi(x,t)}\).

  2. Measurement postulate. Every physical measurement is described by a Hermitian operator \(M\) such that:
    • The eigenvectors of \(M\) are the outcome states of measurement.
    • The (real) eigenvalues of \(M\) are the physically observed quantities.
    • For any state \(\kt{\psi(x,t)}\) and outcome eigenvector \(\kt{\phi(x)}\), the probability of observing that outcome is proportional to \(\magsq{\inr{\phi(x)}{\psi(x,t)}}\).

  3. State change postulate. The dependence on time is prescribed by the Schrodinger equation $$ i\hbar \frac{d}{dt} \kt{\psi(x,t)} \eql H(t) \kt{\psi(x,t)} $$ where \(H(t)\) is an operator, potentially itself time-varying, called the quantum Hamiltonian operator constructed in a particular way:
    • Classical quantities like position \(x\) and momentum \(p\) are replaced by corresponding quantum operators like \(A_{pos}\) and \(A_{mo}\) to form \(H\).
 

Now let's point out a few things:

  • We've only stated the postulates in the simplest possible way, for one dimension, without relativity, and for a confined system.

  • The state description should be somewhat familiar to us, except for the dependence on time.

  • The Schrodinger equation is entirely new, about which we'll say more below.

  • The coefficient function \(\kt{\phi(x,t)}\) is often called the wavefunction.

  • Some books include tensoring as a postulate, others state that it's a consequence of the statistical independence of subsystems (and is not needed as a postulate).
 

Schrodinger's equation and unitary evolution:

  • First, observe: it's a differential equation.
    • One might wonder: is it derived from something more elementary, like many other differential equations in science and engineering.
    • The answer: no, there is no underlying principle that leads to this equation.

  • Second, a state that's changing according to the equation is changing unitarily:
    • When \(H(t) = H\) (not varying with time), one can show that $$ \kt{\phi(x,t)} \eql U(t) \kt{\phi(x,0)} $$ where $$ U(t) \eql e^{-iHt/\hbar} $$ is a unitary operator.
    • Thus, the state is changing as if a unitary operator is being applied.
    • A similar but more technical argument shows that the state changes unitarily even if \(H(t)\) varies with time.
 

Let's focus on an important takeaway:

  • The two key phases of a quantum system carry over into the continuous domain:
    • The system evolves unitarily until a measurement is made.
    • A measurement drastically and probabilistically alters the state into one of the measurement's eigenstates.

  • But what is the nature of this drastic change and how does it happen?

  • For example:
    • When a polarized photon goes through a filter, what exactly causes the drastic change in polarization?
    • When a Stern-Gerlach atom travels through the apparatus in superposition, and one sees two separated landing spots, where does the measurement occur?

  • Another example:
    • When an electron goes through the double slit, it does so in superposition, and yet lands in one single spot.
    • The wavefunction shows some spread and yet it collapses to one single location.

  • And why is measurement probabilistic?

  • These two issues, collapse and non-determinism, together constitute the biggest unresolved mystery in quantum mechanics: the measurement problem.
 


9.10   Schrodinger's cat

 

Let's return to the famous double-slit experiment:

  • Recall the interference pattern:

  • The pattern is in fact the probability distribution (think: continuous histogram) of where a single particle (photon or electron) may land:

  • What's important to note:
    • The entirety of the photon or electron lands in one location.
    • There is no splitting of such a particle.

  • Let's measure landing location from the bottom of the screen:

  • Let's rephrase this as:
    • Some wavefunction \(\kt{\phi(x)}\) describes the probability of finding the particle at location \(x\):

    • A measurement (landing on the screen) results in a particular position eigenvalue outcome \(x\).

  • One could say that the wavefunction collapses upon measurement.

  • Another way to describe this:
    • The quantum object (photon or electron) behaves like a (probabilistic) wave before measurement.
    • Then, it abruptly behaves like a particle.
    This is sometimes called wave-particle duality.
 

Next, let's examine why this is a problem in quantum mechanics:

  • Recall what QM says about the behavior of a quantum object:
        \(\rhd\) There are only two kinds: unitary evolution and measurement.

  • Let's think about the unitary part for a moment in the context of two quantum objects that interact.

  • Suppose one has two particles each in state \(\kt{0}\).

  • Then, there are only two possibilities for unitary interaction:
    • They tensor: \(\kt{0} \otimes \kt{0} = \kt{00}\).
    • Or they get entangled, such as: \(\isqt{1}(\kt{00} + \kt{11})\).
      (There are many ways to entangle.)

  • Let's write this formally to see the unitary interaction algebraically and visually:
    • Tensoring with \(U = I \otimes I\)

      $$ \parenl{I \otimes I} \parenl{\kt{0} \otimes \kt{0}} \eql \parenl{\kt{0} \otimes \kt{0}} $$

    • Entangling example:

      $$ \cnot \, \parenl{H \otimes I} \parenl{\kt{0} \otimes \kt{0}} \eql \isqts{1}\parenl{\kt{00} + \kt{11}} $$

  • Algebraically we see that unitary interaction involves combining wavefunctions (tensoring or entangling).

    Note: we can consider tensoring as a special case of the more general entangling.

  • Now consider this reasoning:
    • All physical objects at the micro-level (atoms, electrons, photons) are thought to be quantum objects:
          \(\rhd\) There's no known non-quantum object at that granularity.
    • So all must be subject to quantum theory.
    • Then, when a photon lands on a screen, its interaction with the quantum objects on the screen must be determined by quantum theory.
    • That is, either tensoring or entangling.

  • The troubling question: then, when, where and how is measurement occuring?

  • Another way to ask the question:
    • When the particle's wavefunction "encounters" the wavefunctions of the screen's constituents, what is the nature of that interaction?
    • QM says it should tensor or entangle.
    • Then, how did measurement occur? When and where?
    • We do, after all, see a single whole-particle at a position.
 

The Copenhagen diktat:

  • To resolve the conundrum, physicist Niels Bohr and others said the following:
    • There is a sharp unknowable boundary between unitary evolution and measurement.
      (Equivalently, between wave and particle behavior.)
    • When measurement occurs, the unitarily-evolving wavefunction instantaneously "collapses" to a single point.
    • Knowing the boundary is not needed for calculations and predictions.
    • It is futile to delve into the measurement conundrum, so it's best to ignore it.

  • In some sense, this was a "don't ask, don't tell" rule.

  • For years, nobody questioned this rule.

  • This has come to be called the Copenhagen interpretation.
 

But Schrodinger and Einstein did not like this state of affairs:

  • Schrodinger created a thought experiment, now famously called Schrodinger's Cat.

  • In this thought experiment:

    Modified from Wikimedia commons

    • A cat is placed in a box and the lid is shut.
    • Inside, there is a quantum device that may probabilistically eject a quantum particle.
    • If ejected, the particle causes a geiger counter to click.
    • That in turns releases poison gas.

  • One can model this simply as
    • Let the randomness of poison be described as $$ \kt{p} \eql \isqts{1} \parenl{ \kt{0} + \kt{1} } $$ where the \(\kt{1}\) state indicates release of particle.
    • Let the cat's two possible states be modeled as $$\eqb{ \kt{0} & \eql & \mbox{alive} \\ \kt{1} & \eql & \mbox{dead} \\ }$$
    • Then, the quantum entanglement can be written as: $$ \isqts{1} \parenl{ \kt{00} + \kt{11} } $$ where $$\eqb{ \kt{00} & \eql & \mbox{particle not released, cat alive} \\ \kt{11} & \eql & \mbox{particle released, cat dead} \\ }$$

  • Schrodinger asked: is the cat in a superposition of two states until observed (measured)?
    • He did this to highlight the difficulty of reasoning about measurement.
Ideas about resolving the measurement conundrum are typically called interpretations.

Let's look at a few of these next.
 


9.11   Interpretations

 

We have seen the Copenhagen interpretation.

We'll now briefly describe two historically important ones: Pilot-Wave, and Many-Worlds.

And mention a few others.
 

The De-Broglie-Bohm pilot wave theory:

  • Often credited to David Bohm, a former student of Oppenheimer in 1951.

  • It turns out that a version of the theory had been proposed earlier by Louis de Broglie.

  • They differ slightly in the details.

  • Both theories involve a global hidden variable called a pilot wave.

  • The main ideas:
    • Every quantum object has an associated pilot wave that guides its behavior.
    • When particles interact, they have a joint pilot wave in so-called configuration space.
    • This wave extends throughout space and evolves through Schrodinger's equation.
    • The quantum object is a particle whose motion is guided by the pilot wave.
    Thus, every particle is both a wave and a particle, with the wave directing the particle's motion.

  • Bohm and De-Broglie showed that their versions produce the same quantitative results as QM-with-Copenhagen.
        \(\rhd\) On this, there's agreement amongst physicists.

  • In pilot-wave theory, randomness is explained via a dependence on initial particle properties.

  • What pilot-wave says about the double-slit:
    • A particle goes through only one slit.
    • The pilot wave, which does go through both slits, determines which slit.

  • What Bohm says about Schrodinger's cat:
    • The cat will be in only one of the two states.
          \(\rhd\) There is no entangled state.
    • The pilot wave of all involved particles determines which state.

  • But Bohm's theory has also been criticized:
    • There's no evidence of this mysterious pilot wave.
    • To make the theory work, one has to believe that the pilot wave for a single particle extends across the whole universe.
    • There is no explanation of how the initial particle properties arise.
    • The theory uses location as the primary variable, with all others secondary.

  • This is an area of active research (and debate):
    • When particles interact, it's only the waves that interact, according to the theory.
 

Many Worlds:

  • Developed by Hugh Everett, a student of QM pioneer John Wheeler.

  • The main, and rather radical, idea:

    • Whenever measurement occurs for any particle, the universe splits into multiple copies, one for each possible outcome.

  • Thus, there is one universe in which Schrodinger's cat is alive, and one in which its dead.

  • The copies are identical in every other respect except for the outcomes.

  • What the theory hopes to do:
    • Provide the simplest, most elegant explanation of measurement.
    • Remove the non-determinism altogether.
 

In-Class Exercise 5: What are your thoughts about the many-worlds idea?
 

There are now a host of theories, in various stages of debate, acceptance, and experimentation, such as:

  • Super-determinism:
    • We used an independence assumption (between identical photons) in showing hidden-variable theory was wrong.
    • However, one can argue that photons carry hidden variables that are correlated with the measuring device's hidden variables.
    • These can be adjusted to produce the same predictions as quantum theory.
    • These variables get "set" at creation: the big bang.
    • Like many-worlds, it's un-testable.

  • Penrose's gravitation model:
    • Gravitational effects play a role in determining a measurement outcome, addressing why collapse occurs and where.

  • The GRW theory:
    • Named after authors Ghirardi, Rimini and Weber.
    • It tweaks Schrodinger's equation to add a variable that determines collapse.

  • In the above two, the wave function collapse is non-instantaneous, and based on physical principles.
 


9.12   Delayed choice and quantum eraser

 

We'll end this QM teaser by describing two intruiging experiments.

Recall, the Mach-Zehnder set up that showed interference:

  • Consider a single photon released from the source.

  • In the first scenario, on the left, a photon has a 0.5 probability to take either path.

  • In the second scenario, there is interference, and no photon ever arrives at D2.

  • Let's first examine the second case:

    • Each splitter is a unitary operation on the photon's wavefunction.
    • The resulting wavefunction is a superposition of the four possible paths.
    • The coefficients of the paths to D2 add up to \(0\) because of the careful set up with distances.

  • The first scenario is just as interesting:

    • The two detectors together act as a measurement device.
    • Only one of them goes off, probabilistically.

  • A skeptic might ask: doesn't the photon go one way or the other at the first beam splitter in Scenario 1?
 

The delayed choice experiment:

  • In this experiment, the second splitter can be inserted after the photon has left the first, and before it reaches the detector:

    • For this set up, the paths have to be long enough to allow for electronic insertion of the second splitter

  • What is observed?
    • Photons for which BS2 is not inserted behave like Scenario 1: randomly landing on either D1 or D2.
          \(\rhd\) 50% land on D2.
    • Photons for which BS2 is inserted all go to D1.
          \(\rhd\) Interference is observed.

  • What to make of this?
    • The results are consistent with QM.
    • But it forces us to acknowledge that the wavefunction is a strange entity that spreads out (both BS1 paths).
    • And that a unitary applied later (BS2) has immediate effect.
 

Finally, let's look at an even stranger result, the quantum eraser experiment:

  • First, consider the following set up:

    • A Polarizing Beam Splitter (PBS) will send horizontally and vertically polarized photons in different directions.
    • A source of entangled photons produces polarization-entangled photons with state $$ \ksi \eql \isqts{1} \parenl{ \kt{V}\kt{V} + \kt{H}\kt{H} } $$
    • Thus, if the left photon is detected at D3, then we know the right photon takes path-1 in the figure.
          \(\rhd\) Thus, measuring the left photon provides which-path information.

  • What QM predicts: when which-path is knowable, no interference occurs.

  • Thus, in the above case, even if we don't actually use the left-side measurements, there will be no interference.
        \(\rhd\) 50% at D1, 50% at D2

  • In the next set up, the left photon's polarization is scrambled randomly (perhaps by measuring in the \(45^\circ\) basis):

    What this means:

    • A \(45^\circ\)-polarized photon will have equal chance of going up or straight on the left side.
    • Thus, the left-side measurement reveals nothing about the path taken on the right.

  • What is observed in this case?
        \(\rhd\) Interference!

  • But this interference is a bit subtle:
    • Overall, there are still 50% at D1, 50% at D2.
    • But if one separates out the photons on the right, whose left-partners went to D3, then those photons will show an interference pattern (100% of them at D1).
    • Similarly, the others will show the opposite interference pattern (100% of them at D2).

  • Now for the next set up:

    Here:

    • A random generator turns on-and-off the polarization scrambler.
    • A significant distance is introduced between the two parts so that left photons reach the scrambler long after the right photons have reached either D1 or D2.

  • Note: the decision about whether to scramble or not occurs after the right side measurement.

  • What is observed?
    • When left photons are scrambled, the right side equivalents show interference.
    • For the unscrambled left photons, the equivalent right side ones do not exhibit interference.

  • The term retrocausality has been used to label this type of phenomenon.
With that, we will conclude our quick peek at quantum mechanics.
 


9.13   The three Fouriers

 

We've seen two of the three so far:

  • The first applies to the vector space of (well-behaved) functions:
    • The (orthonormal) basis consists of sines and cosines.
    • Any vector \(f(x)\) in this space can be expressed as: $$ f(x) \eql \sum_{n=0}^\infty a_n \sin\parenl{ \frac{2\pi n x}{T} } + a_n \cos\parenl{ \frac{2\pi n x}{T} } $$
    • It turns out, it is often more convenient to use the complex exponentials \(e^{i\frac{2\pi nx}{T}}\) as the orthonormal basis and write $$ f(x) \eql \sum_{n=-\infty}^\infty c_n e^{i\frac{2\pi nx}{T}} $$

  • The second Fourier also applies to the vector space of (well-behaved) functions but is a way to transform coordinates in one basis to another: $$\eqb{ \kt{\tilde{\psi}(p)} & \eql & \smf{1}{\sqrt{2\pi}} \int e^{\frac{-ipx}{\hbar}} \, \psi(x) \, dx\\ \kt{\psi(x)} & \eql & \smf{1}{\sqrt{2\pi}} \int e^{\frac{ipx}{\hbar}} \, \tilde{\psi}(p) \, dp\\ }$$ This, we saw, was a way to convert back and forth from coordinates in the position basis to the momentum basis.

  • Then, one might ask: is there a Fourier for plain, finite-dimensional vectors?
        \(\rhd\) There indeed is.
 

The Discrete Fourier Transform:

  • Let's start with the goal:
    • The vector space of interest: N-dimensional complex vectors of the form $$ \kt{w} \eql \mat{w(1) \\ w(2) \\ \vdots \\ w(N)} $$
    • A special basis of orthonormal vectors $$ \kt{v_1}, \ldots, \kt{v_N} $$ where any \(\kt{w}\) can be expressed in this basis as: $$ \kt{w} \eql \sum_k c(k) \kt{v_k} $$

  • Note: we're using \(N\) instead of \(n\) because later when we look at the quantum version, we'll use \(n\) qubits to represent \(N=2^n\) basis vectors.

  • Now define the special number \(\omega = e^{i \frac{2\pi}{N}}\):
    • This will turn out to be one of the n-th roots of \(1\), as we'll later show.
    • Note that $$ \omega^* \eql e^{- i\frac{2\pi}{N}} $$

  • The special basis is $$ \kt{v_0} \eql \mat{ 1 \\ 1 \\ 1 \\ \vdots \\ 1} \;\;\;\;\;\; \kt{v_1} \eql \mat{ 1 \\ \omega \\ \omega^2 \\ \vdots \\ \omega^{N-1}} \;\;\;\;\;\; \kt{v_2} \eql \mat{ 1 \\ \omega^2 \\ \omega^4 \\ \vdots \\ \omega^{2(N-1)}} \;\;\;\;\;\; \ldots \;\;\;\;\;\; \kt{v_{N-1}} \eql \mat{ 1 \\ \omega^{N-1} \\ \omega^{2(N-1)} \\ \vdots \\ \omega^{(N-1)(N-1)}} $$

  • Here, the indexing starts at \(0\) to match the powers of \(\omega\): $$ \kt{v_k} \eql \mat{ (\omega^k)^0 \\ (\omega^k)^1 \\ (\omega^k)^2 \\ \vdots \\ (\omega^k)^{N-1} } $$

  • One can show that this is an orthogonal (but not unit-length) basis.

  • Then, any \(\kt{w}\) can be expressed as $$ \kt{w} \eql \sum_{k=0}^{N-1} c(k) \kt{v_k} $$ where the coefficients work out to $$\eqb{ c(k) & \eql & \sum_{j=0}^{N-1} (\omega^*)^{jk} \\ & \eql & \sum_{j=0}^{N-1} e^{- i \frac{2\pi jk}{N}} }$$

  • Now, let's place the special basis vectors \(\kt{v_k}\) as columns of a matrix \(D\): $$ D \eql \mat{ & & & \\ \vdots & \vdots & \ldots & \vdots \\ \kt{v_0} & \kt{v_1} & \ldots & \kt{v_{N-1}}\\ \vdots & \vdots & \ldots & \vdots \\ & & & } \eql \mat{ 1 & 1 & 1 & \vdots & 1 \\ 1 & \omega & \omega^2 & \vdots & \omega^{N-1} \\ 1 & \omega^2 & \omega^4 & \vdots & \omega^{2(N-1)} \\ \vdots & \vdots & \vdots & \vdots & \vdots \\ 1 & \omega^{N-1} & \omega^{2(N-1)} & \ldots & \omega^{(N-1)(N-1)} } $$

  • Then, we can rewrite the expression of \(\kt{w}\) as $$ \mat{w(0) \\ w(1) \\ \ldots \\ w(N-1)} \eql \mat{ 1 & 1 & 1 & \vdots & 1 \\ 1 & \omega & \omega^2 & \vdots & \omega^{N-1} \\ 1 & \omega^2 & \omega^4 & \vdots & \omega^{2(N-1)} \\ \vdots & \vdots & \vdots & \vdots & \vdots \\ 1 & \omega^{N-1} & \omega^{2(N-1)} & \ldots & \omega^{(N-1)(N-1)} } \mat{c(0) \\ c(1) \\ \ldots \\ c(N-1)} $$

  • From this we see that the matrix \(D\) converts from one set of coordinates to another:
    • The \(w(k)\) coordinates are standard coordinates.
    • The \(c(k)\) coordinates are Fourier coordinates.
    The conversion here is from Fourier to standard.

  • One can show that \(D^{-1} = \frac{1}{n} D^\dagger\), and so \(D^{-1}\) will convert from standard to Fourier.

  • This pair of transforms are called
    • \(D^{-1}\): the Discrete Fourier Transform (DFT)
    • \(D\): the Inverse Discrete Fourier Transform (i-DFT)

  • The quantum version is called the QFT and plays a key role in Shor's algorithm.
 

About the Fouriers:

  • The three Fouriers have proven extraordinarily useful across a wide variety of domains.

  • Theoretically, they help construct and understand solutions to differential equations, including, as we see, Schrodinger's equation.

  • Practically, they are invaluable as a solution tool.

  • Perhaps none more so than the DFT, which is a workhorse of engineering.

  • The DFT, which is an \(O(N^2)\) calculation (matrix-vector multiplication), can be cleverly implemented in time \(O(N \log N)\):
    • This is called the Fast Fourier Transform (FFT).
    • It is one of the most useful algorithms of all time.

  • The three Fouriers, with theory and applications, have spanned over two centuries.

  • They are now joined by the fourth Fourier: the Quantum Fourier Transform (QFT), whom we shall meet in Shor's algorithm.
 




© 2022, Rahul Simha