\( \newcommand{\blah}{blah-blah-blah} \newcommand{\eqb}[1]{\begin{eqnarray*}#1\end{eqnarray*}} \newcommand{\eqbn}[1]{\begin{eqnarray}#1\end{eqnarray}} \newcommand{\bb}[1]{\mathbf{#1}} \newcommand{\mat}[1]{\begin{bmatrix}#1\end{bmatrix}} \newcommand{\nchoose}[2]{\left(\begin{array}{c} #1 \\ #2 \end{array}\right)} \newcommand{\defn}{\stackrel{\vartriangle}{=}} \newcommand{\rvectwo}[2]{\left(\begin{array}{c} #1 \\ #2 \end{array}\right)} \newcommand{\rvecthree}[3]{\left(\begin{array}{r} #1 \\ #2\\ #3\end{array}\right)} \newcommand{\rvecdots}[3]{\left(\begin{array}{r} #1 \\ #2\\ \vdots\\ #3\end{array}\right)} \newcommand{\vectwo}[2]{\left[\begin{array}{r} #1\\#2\end{array}\right]} \newcommand{\vecthree}[3]{\left[\begin{array}{r} #1 \\ #2\\ #3\end{array}\right]} \newcommand{\vecfour}[4]{\left[\begin{array}{r} #1 \\ #2\\ #3\\ #4\end{array}\right]} \newcommand{\vecdots}[3]{\left[\begin{array}{r} #1 \\ #2\\ \vdots\\ #3\end{array}\right]} \newcommand{\eql}{\;\; = \;\;} \definecolor{dkblue}{RGB}{0,0,120} \definecolor{dkred}{RGB}{120,0,0} \definecolor{dkgreen}{RGB}{0,120,0} \newcommand{\bigsp}{\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;} \newcommand{\plss}{\;\;+\;\;} \newcommand{\miss}{\;\;-\;\;} \newcommand{\implies}{\Rightarrow\;\;\;\;\;\;\;\;\;\;\;\;} \newcommand{\prob}[1]{\mbox{Pr}\left[ #1 \right]} \newcommand{\exval}[1]{\mbox{E}\left[ #1 \right]} \newcommand{\variance}[1]{\mbox{Var}\left[ #1 \right]} \newcommand{\kt}[1]{\left\vert #1 \right\rangle} \newcommand{\br}[1]{\left\langle #1 \right\vert} \newcommand{\bkt}[2]{\left\langle #1 \middle\vert #2 \right\rangle} \newcommand{\pss}{\large\psi} \)


Math Review

 


Note: some parts of the review (such as calculus) aren't needed for quantum computing, but are presented as prelude to following one's curiosity about quantum mechanics.
 


M.1   Sets and numbers

 

About sets:

 

Sets of numbers:

  • We're not going to go deep into numbers or number theory; we'll only outline a few important sets by way of background.

  • Natural numbers: \(\mathbb{N} = \{1, 2, 3, \ldots\}\)
    • A set by itself has limited value until we 'do' something with its elements.
    • Thus we define the common and intuitive operators \(+, -, \times, \div\).
    • In doing so, we run into our first difficulty: clearly \(5 - 2\) is in the set \(\mathbb{N}\) but what about \(2 - 5\)?

  • The difficulty is resolved by expanding our notion of number.

  • Integers: \(\mathbb{Z} = \{\ldots, -3, -2, -1, 0, 1, 2, 3, \ldots\}\), the set of integers.
    • Now all additions, subtractions and multiplications give us members of the same set.
    • But division requires an expanded set of numbers.

  • Rational numbers: \(\mathbb{Q} = \{\frac{p}{q}: \mbox{p and q are integers}\}\)
    • Now all four basic operators applied to any two members of \(\mathbb{Q}\) results in a member of \(\mathbb{Q}\).
    • Yet, there are other operations we seek to perform, and soon discover that, for example, \(\sqrt{2}\) is not rational.

  • Real numbers:
    • Real numbers are actually tricky to define rigorously.
    • One can't simply add all square roots to the rationals: what about the number that describes the ratio of a circle's circumference to its diameter?
    • We'll stick to an intuitive notion: the reals \(\mathbb{R}\) are all the numbers we get from common operations such as: arithmetic, powers, roots, limits.
    • Once again, we run into a difficulty:
      • We know that the equation \(x^2 = 2\) has a real number solution.
      • But \(x^2 = -2\) does not.

  • Which finally leads to complex numbers, commonly denoted by \(\mathbb{C}\):
    • By introducing a symbol \(i\) to represent \(\sqrt{-1}\), \((\sqrt{2}i)^2 = -2\).
    • One representation of a complex number: \(a + bi\).
    • Another (the polar representation): \(r e^{i\theta}\).

  • And to a natural question: do we need anything beyond complex numbers?
    (We'll address this below.)
 

Let's point out a few highlights regarding sets of numbers:

  • It is possible to reason about the cardinality (size) of infinite sets.

  • From which:
    • \(\mathbb{N}, \mathbb{Z}, \mathbb{Q}\) all have the same cardinality, commonly called countable.
    • The cardinality of the reals (and complex numbers) is larger, called uncountable.
    • There are other oddities when reasoning about the different sizes of infinity, for example: the number of points on the plane has the same cardinality as the number of points on the line.

  • Types of real numbers:
    • Consider a polynomial like: \(p(x) = -3 + x^2 + 3x^5\).
    • It has long been of interest to solve equations like \(p(x) = 0\) or to reason whether solutions can in fact be found.
    • A algebraic real number \(\alpha\) is a number for which at least one polynomial \(p(x)\) exists such that \(p(\alpha) = 0\).
    • Thus, every algebraic number is the solution to some polynomial equation, of which there are only a countable number.
    • Therefore, the other reals (which are uncountable) can't be solutions to a polynomial equation: we call these transendental numbers.
    • Examples are: \(\pi, e\).

  • Polynomial equations may not have a real solution. If, however, we include complex numbers as possible solutions to polynomial equations, then every such equation has at least one solution.
    • This is known as the Fundamental Theorem of Algebra.
    • The theorem says more: a degree-n polynomial has exactly n complex roots (allowing for multiplicity).
 

Let's now turn to a few questions:

  • Since the reals include the integers, why bother with integers?
    • Recall the difficulty: division.
    • We can in fact modify the division operator to give us two operators that result in integers:
      • The integer div operator, as in \(7 \mbox{ div } 3 = 2\)
      • The integer remainder operator, as in \(7 \mbox{ mod } 3 = 1\)
    • Of these, the remainder operator turns out to be the interesting one, leading to modern applications like cryptography.

  • Since complex numbers include the reals, why bother with reals?
    • Reals are a bit easier to understand.
    • And our real world has measurable quantities (length, temperature, time) for which real numbers are well-suited.
    • Thus, even if we work with complex numbers, we'll want to know how to extract a real part that's relevant to an application.

  • Do we need anything beyond complex numbers?
    • Surprisingly, the answer for us is: yes.
    • We'll see that when we need to explain quantum phenomena, what's needed is a bigger structure: a complex vector.

  • One way to think of a complex number \(a + bi\) is that it is a pair of real numbers \((a,b)\).

  • Then, a 2-component complex vector \((a+bi, c+di)\) can be considered a quadruple of real numbers. This is what will be needed, as we'll see, to represent a qubit.
 


M.2   Functions

 

What is a function?

  • Intuitively, a function is something that "computes" a result (another number) given an input.

  • For example, \(f(x) = x^2\) takes a number \(x\) and produces its square.

  • Alternatively, we can think of a function as an association: the square function associates every number with its square, e.g., $$ \begin{array}{lll} 2 \;\;\;\;\; & \;\;\;\;\; \mbox{is associated with} \;\;\;\;\; & \;\;\;\;\; 4\\ 3 \;\;\;\;\; & \;\;\;\;\; \mbox{is associated with} \;\;\;\;\; & \;\;\;\;\; 9\\ 3.126 \;\;\;\;\; & \;\;\;\;\; \mbox{is associated with} \;\;\;\;\; & \;\;\;\;\; 9.771876\\ \end{array} $$ Here, there is no specified computation.
 

The "association" view becomes useful when we consider coordinates:

  • Note: Greek geometry developed entirely without coordinates, and for that matter, even numbers.
    (Their later geometry did feature numeric quantities: areas and volumes.)

  • Many sophisticated results were derived without algebra: geometrical constructions, proofs of properties, conics, areas and volumes.
         \(\rhd\) Example: trisecting a line segment (not trivial)

  • Meanwhile, algebra, which arrived from the east was a powerful way of representing rules and working with expressions:
    • For example, a general answer can be given algebraically to the polynomial equation \(ax^2 +bx + c = 0\).

    Enter Descartes ("I think, therefore I need coordinates"), with an astonishingly simple idea: associate two numbers with every point in the plane
         \(\rhd\) distances from two fixed lines (called axes)

    • Which allowed algebraic representations of geometric ideas
           \(\rhd\) lines, curves, surfaces, area, volumes
    • Coordinates also led to an appealing visualization of functions:

    • Descartes introduced the modern convention for notating basic algebra, such as variables (\(x,y,z\)), exponentiation (\(x^y\))
 

More about functions:

  • We use the term locus to describe a set of points satisfying a certain property.

  • Example: consider all points \((x,y)\) such that \(2x - 3y - 6 = 0\)

    • We write this as \(\{(x,y): 2x-3y-6=0\}\).

    • Or as a function: \(f(x) = (2x - 6) / 3 \)

    • This turns out, geometrically, to be a line.

  • Functions became much more useful and powerful when associated with coordinate geometry:
    • One could now algebraically describe the curves that had already been well-studied: $$\eqb{ \frac{x^2}{r^2} + \frac{y^2}{r^2} = 1 & \;\;\;\;\;\; & \mbox{circle} \\ \frac{x^2}{a^2} + \frac{y^2}{b^2} = 1 & \;\;\;\;\;\; & \mbox{ellipse} \\ \frac{x^2}{a^2} - \frac{y^2}{b^2} = 1 & \;\;\;\;\;\; & \mbox{hyperbola} \\ y^2 - 4ax = 0 & \;\;\;\;\;\; & \mbox{parabola} \\ }$$ And, moreover, exploit algebraic results to rigorously derive new properties in full generality.
    • There soon followed even more algebraic descriptions, such as the parametric representation: \(x(t) = r\cos(t), y(t)=r\sin(t)\) (circle).

  • The general idea of and notation for single-variable functions is easily extended to:
    • Multiple variables, e.g. \(f(x,y,z)\)
    • Function composition, e.g., \(f(g(x))\)
    • An inverse of a function, \(f^{-1}(x)\)
 

Special classes of functions:

  • With the development of the mathematics of functions, some particular functions grew in significance.

  • For example, polynomials occur in many contexts, and are easy to understand, easy to calculate by hand.

  • An important question then arose: are polynomials good enough? That is, can any function be well-approximated by some polynomial?
    • The Stone-Weierstrass theorem confirmed the existence of a polynomial \(p(x)\) such that \(|p(x) - f(x)| \lt \epsilon\) when given \(f(x)\) and \(\epsilon\) in the interval of interest.
    • Bernstein showed that a linear combination of (the later to be named) Bernstein polynomials could arbitrarily closely approximate any continuous function.

  • More significantly, the trigonometric "polynomials", the functions \(\sin(2\pi\omega t)\) and \(\cos(2\pi\omega t)\) have come to dominate applied mathematics.
    • Fourier's result: $$ f(t) \;\; = \;\; a_0 + \sum_{k=1}^\infty a_k \sin(2\pi kt/T) + \sum_{k=1}^\infty b_k \cos(2\pi kt/T) $$ for any continuous function \(f(t)\) over an interval \([0,T]\).
    • Thus, a linear combination of the basic trig's can be used to approximate a function as closely as desired.

  • A discrete, vectorized version, the Discrete Fourier Transform (DFT), is one of the most useful mathematical tools in science and engineering.

  • And, as we will see, the quantum DFT is a foundational building block of quantum computing.
    • It should be surprising that sine's and cosine's have something to do with factoring integers.
 


M.3   Calculus

 

The historical background:

  • Scientists before Newton already knew from data that the planets traveled in ellipses and that a ball thrown at an angle traverses a parabola.

  • For example: Galileo performed detailed experiments to understand the paths traversed by moving objects, and Kepler fitted ellipses to planetary observations.

  • However, none were able to explain why motion resulted in these particular (well-known at the time) curves.
 

Calculus begins with limits:

  • Consider the sequence of numbers \(f(n) = \frac{10n}{2n+1}\), one number for each natural number \(n \geq 1\).

  • The first few happen to be (to 2 decimal places): $$\eqb{ f(1) & \eql & \frac{10}{3} & \eql & 3.33 \\ f(2) & \eql & \frac{20}{5} & \eql & 4.00 \\ f(3) & \eql & \frac{30}{7} & \eql & 4.29 \\ f(4) & \eql & \frac{40}{9} & \eql & 4.44 \\ f(5) & \eql & \frac{50}{11} & \eql & 4.55 \\ \vdots & & \vdots & & \vdots \\ f(100) & \eql & \frac{10000}{201} & \eql & 4.98 \\ }$$ If we rewrite the n-th term as $$ f(n) \eql \frac{10}{2 + \frac{1}{n}} $$ then it's easier to see that as \(n\) grows towards \(\infty\), the \(n\)-th term approaches \(5\).

  • We write this formally as $$ \lim_{n\to\infty} f(n) \eql \lim_{n\to\infty} \frac{10}{2 + \frac{1}{n}} \eql 5 $$

  • It is not necessary to have an integer sequence:
    • We could define $$ f(\Delta) \eql \frac{10}{2 + \Delta} $$ where \(\Delta\) is a real number.
    • Then, $$ \lim_{\Delta\to 0} f(\Delta) \eql \lim_{\Delta\to 0} \frac{10}{2 + \Delta} \eql 5 $$

  • Because sequences can be strange, a more rigorous definition has now come to be accepted: the limit $$ \lim_{x\to a} f(x) \eql L $$ if for any \(\epsilon \gt 0\) there exists a \(\delta \gt 0\) such that \(|f(x) - L| \lt \epsilon\) whenever \(|x-a| \lt \delta\).

  • This definition and a deeper analysis of limits has led to a theory of limits:
    • Many frequently occuring limits have been worked out.
    • The theory has been developed to work with combinations, ratios, and compositions of functions.
      (Remember L'Hopital's Rule?)

  • A special case of limits, the sequence of partial sums turns out to be especially useful:
    • For example, suppose we define $$ S(n) \eql \sum_{i=0}^n \frac{1}{2^i} $$
    • Then, \(S(0), S(1), S(2), S(3), \ldots\) is a sequence of numbers, each representing a partial sum of the infinite series \(\sum_{i=0}^\infty \frac{1}{2^i}\).
    • This has a well-defined limit so we can in fact directly write $$ \sum_{i=0}^\infty \frac{1}{2^i} \eql 2 $$ We call this an infinite series.
    • Infinite series of various kinds have proven extremely useful.
    • Here's one to remember: $$ e^x \eql 1 + x + \frac{x^2}{2!} + \frac{x^3}{3!} + \ldots $$
 

Derivatives:

  • Newton (and Leibniz) began with a simple question:
    • A object moving in a straight line with constant velocity \(10\) has velocity \(v(t) = 10\) at all times \(t\) during its travel.
    • What is \(v(t)\) when the velocity is not constant, as in ball dropping to the ground?
    • If the ball dropped from height \(h\) lands at time \(T\) then clearly \(v(t) \neq \frac{h}{T}\) because it's changing.

  • The first insight came by pondering the precise relationship between instantaneous velocity \(v(t)\) and the distance \(d(t)\) traveled by time \(t\): $$ \lim_{\Delta t\to 0} \frac{d(t+\Delta t) - d(t)}{\Delta t} $$

  • The most critical realization is that the result of taking the limit is a function: you get one value on the right for every \(t\). Thus, one can give this function a name, \(v(t)\): $$ v(t) \; \defn \; \lim_{\Delta t\to 0} \frac{d(t+\Delta t) - d(t)}{\Delta t} $$

  • Even more importantly, many such "rate" limits can be worked out analytically, for example, for \(f(x) = x^2\): $$ \lim_{\Delta x\to 0} \frac{ (x+\Delta x)^2 - x^2}{\Delta x} \eql \lim_{\Delta x\to 0} \frac{2x\Delta x + (\Delta x)^2}{\Delta x} \eql 2x $$

  • This function came to be called the derivative, with various alternative notations: $$ f^\prime(x) \eql \frac{df}{dx}(x) \; \defn \; \lim_{\Delta x\to 0} \frac{f(x+\Delta x) - f(x)}{\Delta x} $$ Here, we want to know: what function \(f\) satisfies the above equation?

  • With this definition and the ability to work out limits, a great number of derivative functions can be calculated, and one can take derivatives of derivatives such as \(f^{\prime\prime}(x) = \frac{df^\prime}{dx}(x)\)

  • For Newton (and physics), this mattered because the physical world is rife with examples where a law is more easily observed in derivative form:
    • For example, to a good approximation, all motion has constant acceleration (\(d^{\prime\prime}(t) = \mbox{ constant})\).

  • With the additional critical insight that any 2D or 3D motion can be cleanly decomposed (projected) along orthogonal axes (linear algebra!), it now becomes easier to put together a theory of motion along curves.

  • There are of course many more steps from here to finally showing how gravitation results in elliptical planetary motion, but not too many (the notion of force, for example):

  • Newton's triumph of course was a complete quantification of the motion of two bodies relative to each other under gravitation, resulting in conic-section trajectories.

  • Soon, other laws followed, conveniently stated as a differential equation (an equation involving derivatives) such as $$ m \; f^{\prime\prime}(t) \eql -k \; f(t) $$ (Simple harmonic motion.)

  • Why are differential equations important?
    • Laws and insights are convenient to formulate using derivatives.
    • For example, in the above differential equation, the left side expresses force in terms of acceleration \(f^{\prime\prime}(t)\) and the right side expresses force in terms of distance \(f(t)\).
    • It's the fact that forces are equal that is simple to express: this is what becomes the differential equation.
    • A single differential equation can apply to many different situations, and solution techniques for one can apply to others.

  • Once a differential equation is written then, of course, we want to know: what function solves the equation?
 

Integration:

  • The key takeaway from derivatives and limits is that taking limits simplifies expressions.

  • Now, Newton also asked about the reverse direction in reasoning: if \(v(t)\) is a given velocity function, can one derive the distance function \(d(t)\) from it?

  • If one divides the interval \([0, T]\) into \(n\) small intervals of size \(\delta\) then:
    • The velocity in the \(k\)-th interval approximately \(v(k\delta)\).
    • The distance moved during this time is approximately \(v(k\delta)\delta\).
    • Thus the total distance moved is approximately $$ \sum_k v(k\delta)\delta $$
    • Then, letting \(\delta\to 0\) (and therefore \(k\to\infty\)), one writes the limit as $$ \int_0^T v(t) dt $$

  • Since we know that \(d^\prime(T) = v(T)\), differentiation and integration are related (through the Fundamental Theorem of Calculus).
 

Multiple variables:

  • Because many phenomena involve multiple variables, a natural extension to single-variable functions, is a multi-variable function like \(f(x_1,x_3,\ldots,x_n)\).

  • In this case, the "rate of change" for each variable, holding the others still, is called a partial derivative: $$ \frac{\partial f}{\partial x_i} \; \defn \; \lim_{\Delta x_i\to 0} \frac{f(x_1,x_2,\ldots,x_i+\Delta x_i,\ldots, x_n) - f(x_1,x_2,\ldots,x_i,\ldots, x_n)}{\Delta x_i} $$

  • For example, if \(f(x,y) = 3x^2y + xy^3\) then $$ \frac{\partial f}{\partial x} \eql 6xy + y^3 $$

  • Then, laws can be written in terms of partial derivatives and one must then scramble to find solutions.

  • For example, here's one commonly used partial differential equation: $$ \frac{\partial^2 f(x,t)}{\partial x^2} \eql \frac{1}{v^2} \frac{\partial^2 f(x,t)}{\partial t^2} $$
 

Summary:

  • Calculus has proven extraordinarily useful as a mathematical tool to quantify all kinds of physical phenomena.

  • When laws are expressed using calculus, the expressions have explanatory power.

  • Solving a differential or integral equation, however, is another matter:
    • The vast majority of practical applications are hard to solve exactly with a clean looking function.
    • Either approximations are used, or a computational approach produces numerical values.
    • Many approximations, such as a Fourier series, have their own explanatory (and computational) power.

  • The applications of calculus go well beyond physical phenomena:
    • One powerful example is: probability.
    • Reasoning about probability with continuous variables involves calculus, as it turns out.

  • We aren't going to need calculus at all for quantum computing, but it's certainly needed for quantum mechanics.
 


M.4   Probability

 

For quantum computing we need only the most elementary notions in discrete probability.

For quantum mechanics, however, one needs more: an understanding of continuous probability (density functions and the like).
 

Let's review some discrete probability:

  • Suppose we have a jar that has 5 cookies each labeled "1" and 10 cookies labeled "4".
    • We now ask the question: if we reach in and remove a cookie randomly, what are the chances we'll pick a type "1" cookie?
    • We know from elementary reasoning that it's \(\frac{5}{15}\)
    • And we write this as: $$ \mbox{Probability we get a "1" cookie} \eql \frac{5}{15} $$
    • And $$ \mbox{Probability we get a "4" cookie} \eql \frac{10}{15} $$

  • Let's generalize this a bit:
    • Suppose our setup has \(n_1\) cookies labeled "1", \(n_2\) cookies labeled "4", ..., \(n_i\) cookies labeled "\(i^2\)", ... \(n_k\) cookies labeled "\(k^2\)".
      (Made by a mathematician baker, obviously. And square-shaped instead of round.)
    • We pick one cookie randomly.
    • Then $$ \mbox{Probability we get a cookie labeled "\(i^2\)"} \eql \frac{n_i}{n_1 + n_2 + \ldots + n_k} $$

  • Let's use the symbol $$ p_i \; \defn \; \frac{n_i}{n_1 + n_2 + \ldots + n_k} $$ This is the probability of obtaining the i-th category (that happens to have the peculiar value "\(i^2\)").

  • This example shows what we mean by discrete probability:
    • Typically, there are a fixed number of categories, like the k categories above.
    • There's a probability associated with each category.
    • The probabilities sum to 1: $$ p_1 + p_2 + \ldots + p_k \eql 1 $$

  • Now suppose we replace the randomly selected cookie after each random drawing, then we could repeat the experiment many times:
    • Each time, we'll get a value: if we pick the i-th category, the value is \(i^2\) in our strange example.
    • For example, in our first 2-cookie-types example, we'd get either "1" or "4".
    • We could ask: what is the average value over a large number of repetitions?
    • For the 2-cookie-types example, we'd expect the average to be closer to 4 (since there are more "4" cookies).

  • Suppose, in our 2-cookie case, we performed 10 repetitions (called trials, in the jargon) and obtained
           First trial:     1
           2nd trial:       4
           3rd:             4
           4th:             1
           5th:             1
           6th:             4
           7th:             4
           8th:             1
           9th:             4
           10th:            4
           
    The average with this particular data turns out to be $$ \frac{1 + 4 + 4 + 1 + 1 + 4 + 4 + 1 + 4 + 4}{10} \eql \frac{28}{10} \eql 2.8 $$

  • Of course, if we ran another 10 trials, we might get something different.

  • Next, let's perform a calculation with the probabilities: $$ \mbox{value}_1 p_1 + \mbox{value}_2 p_2 \eql 1 p_1 + 4 p_2 \eql 1 \frac{5}{15} + 4 \frac{10}{15} \eql 3 $$

  • For the k-cookie case, this calculated value becomes $$ \sum_{i=1}^k i^2 p_i $$ This is a sort-of weighting of values by their probabilities, and not the result of number-crunching with data.

  • This calculation with probabilities is called the expected value.
    • What does this have to do with the average from experimental data?
    • If you have a lot of data, then the average from the data should be pretty close to the expected value obtained by calculation with the probabilities.
    • This is sometimes called the Law of Large Numbers.

  • Statistians would use the term sample mean for the average obtained from the data, and population mean for the calculated expected value.
 

Let's define the term random variable:

  • Textbook probability problems sometimes feature non-numeric entities like a deck of cards (where the average makes no sense).

  • We will restrict our attention to only experiments with numeric outcomes, even if they are numbers on cookies.
    • Of course, in the real world, we have measurements like temperature (a real number) or the number of students in room (an integer).

  • We use the term random variable, typically with a symbol like \(X\), to denote the outcome of an experiment:
    • Thus, in our 2-cookie experiment: $$ X \;\defn \; \mbox{value on cookie} $$ and thus $$ X \;\in \; \{1, 4\} $$
    • We can list the probability for each outcome: $$\eqb{ \prob{X = 1} & \eql \frac{5}{15} \\ \prob{X = 4} & \eql \frac{10}{15} \\ }$$
    • And the expected value of this random variable then is written as: $$ \exval{X} \eql 1 \prob{X=1} + 4 \prob{X=4} $$
    • In general, for the k-cookie version this would become $$ \prob{X = i} \eql \frac{n_i}{n_1 + n_2 + \ldots + n_k} $$ This is called a discrete probability distribution.
    • With expected value: $$ \exval{X} \eql \sum_{j} j \: \prob{X=j} $$

  • There some "well known" distributions that have a particular form and that tend to get used often. These have names like: Binomial, Geometric, and so on.
 

Even though not needed for quantum computing, let's say a few things about continuous random variables:

  • In this case, a continuous random variable will take on real-number values: \(X\in \mathbb{R}\).

  • For example, if the outcome is a body-temperature measurement, we could have \(X\in [90, 108]\), with \(X=98.6\) being an example value.

  • Distributions, however, are specified differently, using the tools of calculus:
    • For a continuous random variable, the probability for a particular value is typically \(0\): \(\prob{X=a}=0\).
    • Instead, one obtains the probability that a value obtained from an experiment lies in a range: $$ \prob{X \in [a,b]} \eql \int_a^b f_X(x) \; dx $$ where the function \(f_X\) is the probability density function of the random variable \(X\).

  • In a similar manner, the expected value is calculated as: $$ \exval{X} \eql \int_{-\infty}^{\infty} x \;f_X(x) \; dx $$

  • Why densities are used, and how we work with them, are not trivial to understand.

  • Here, for example, is the most well-known density: $$ f(x) \eql \frac{1}{\sqrt{2\pi} } e^{- \frac{x^2}{2}} $$
    • This is the standard normal distribution.
    • It turns out that \(\exval{X}=0\) for this density.
 

Although not needed for this course, you are encouraged to review basic probability: along with linear algebra, it is the one of the most applicable areas of math. Here is a one-module review that you can use.

With this topic, we have reviewed material you are likely to have already seen before.

We'll now turn to some ideas you may or may not have seen.
 


M.5   Functions and operators

 

Let's return to derivatives for a moment and think of the action of calculating a derivative as:

The box represents something that takes as input a function and produces as output a function
     \(\rhd\) In this case, the derivative of the input function.

The idea of an operator:

  • Whereas a function's input and output are numbers, we'll use the term operator for action of taking a function and returning a function.

  • Our example showed the derivative operator.

  • In notation, we'll write the derivative operator as $$ {\bf D} \; f(x) \eql f^\prime(x) $$ In other words, the operator \({\bf D}\) applied to the function \(f(x)\) produces its derivative \(f^\prime(x)\) as output.

  • Clearly, one can imagine and design all kinds of operators.

  • For example, here's a simple one: \({\bf S}_\alpha\) is the scalar operator that multiples a function by the number \(\alpha\): $$ {\bf S}_\alpha \; f(x) \eql \alpha f(x) $$

  • This idea of the input/output being something other than a number is one we've seen before with vectors: $$ {\bf A} \; {\bf x} \eql {\bf y} $$ Here, one can think of the effect of left-multiplying by matrix \({\bf A}\) as:

  • The notation resembles our just-introduced operator notation as well.

  • Accordingly, we'll broaden our definition of operator to apply both to functions and vectors, with the same notation: $$\eqb{ {\bf D} \; f(x) & \eql & f^\prime(x) & \;\;\;\;\;\;\;\mbox{Functions}\\ {\bf A} \; {\bf x} & \eql & {\bf y} & \;\;\;\;\;\;\;\mbox{Vectors} }$$

  • Why is this useful?
    • We will later see that a function can be thought of as a "continuous" version of a vector.
    • Thus, when transforming one function to another, we'll extend the "multiply by a matrix on the left" to "apply an operator on the left".
    • What is interesting is that a general theory of operators applies both to vectors and functions.
    • Quantum mechanics makes significant use such operators.

  • For example, the multiply-by-scalar operator can be written to apply to either functions or vectors: $$\eqb{ {\bf S}_\alpha \; \kt{f(x)} & \eql & \alpha \kt{f(x)} & \;\;\;\;\;\;\;\mbox{Functions}\\ {\bf S}_\alpha \; \kt{\bf x} & \eql & \alpha \kt{\bf x} & \;\;\;\;\;\;\;\mbox{Vectors} }$$

  • In the quantum literature, it's common to not distinguish and use a single symbol like \(\pss\) $$ {\bf S}_\alpha \; \kt{\pss} \eql \alpha \kt{\pss} \;\;\;\;\;\;\;\mbox{Function or vector} $$

  • We'll later see that \(\pss\) is a unifying way to refer to the state of a system whether discrete (vector) or continuous (function).

  • And an operator's effect will be analyzed using its eigenvectors and eigenvalues, whether discrete (matrix) or continuous (operator).
         \(\rhd\) This will turn out to be the key to understanding the quantum world.
 

A point about notation:

  • We've introduced notation for function operators like \({\bf D} \; f(x) \eql f^\prime(x)\) to see that it's similar to vector operators \({\bf A} \; {\bf x} \eql {\bf y}\).

  • But just as often, one peels back \(\frac{d}{dx}\) from \(\frac{df}{dx}\) and treat \({\bf \frac{d}{dx}}\) as an operator: $$ {\bf \frac{d}{dx}} \; f(x) \eql f^\prime(x) $$

  • As an example, one could define the operator \( \left( \frac{\partial^2}{\partial x^2} - \frac{1}{v^2} \frac{\partial^2}{\partial t^2} \right) \) and apply it to a function \(f(x,t)\): $$ \left( \frac{\partial^2}{\partial x^2} - \frac{1}{v^2} \frac{\partial^2}{\partial t^2} \right) \; f(x,t) \eql \frac{\partial^2 f(x,t)}{\partial x^2} - \frac{1}{v^2} \frac{\partial^2 f(x,t)}{\partial t^2} $$
 

Operators with "nice" properties:

  • There are four properties that turn out to be critical when considering operators: linear, Hermitian, unitary, bounded.

  • A general operator \({\bf A}\), whether a function operator or vector operator, is linear if:
    • For vector operator \({\bf A}\) and vectors \({\bf x}\) and \({\bf y}\) it's true that $$ {\bf A}\; (\alpha {\bf x} + \beta {\bf y}) \eql \alpha {\bf Ax} + \beta {\bf Ay} $$ for all scalars \(\alpha, \beta\).
    • For function operator \({\bf A}\) and functions \(f\) and \(g\) it's true that $$ {\bf A}\; (\alpha f + \beta g) \eql \alpha {\bf A}f + \beta {\bf A}g $$ for all scalars \(\alpha, \beta\).

  • As an example, consider the function operator \({\bf \frac{d^2}{dx^2}}\) (the second derivative):
    • This is a linear operator because $$ { \frac{d^2}{dx^2}} \; (\alpha f + \beta g) \eql \alpha \frac{d^2}{dx^2}f + \beta \frac{d^2}{dx^2}g $$
    • For example, if \(f(x) = 3x^2, \; g(x)=e^{10x}\), $$\eqb{ {\frac{d^2}{dx^2}} \; (\alpha f + \beta g) & \eql & { \frac{d^2}{dx^2}} \; (\alpha 3x^2 + \beta e^{10x} ) \\ & \eql & \alpha \frac{d^2}{dx^2} 3x^2 \;+ \; \beta \frac{d^2}{dx^2}e^{10x}\\ & \eql & \alpha \frac{d^2}{dx^2}f + \beta \frac{d^2}{dx^2}g }$$

  • Generally, anything linear is easier to work with than anything nonlinear, but the biggest significance is in interpreting the linearity above:
    • Suppose we have some linear differential equation like $$ \frac{\partial^2 f(x,t)}{\partial x^2} \eql \frac{1}{v^2} \frac{\partial^2 f(x,t)}{\partial t^2} $$
    • We notice that, in linear operator form, it can be written as: $$ \left( \frac{\partial^2}{\partial x^2} - \frac{1}{v^2} \frac{\partial^2}{\partial t^2} \right) \; f(x,t) \eql 0 $$
    • Since the operator is linear, if functions \(g(x,t)\) and \(h(x,t)\) happen to be solutions to the equation, so is any \(\alpha g(x,t) + \beta h(x,t) \).
    • In this particular case, you can work through and show that \(\sin(x - vt)\) and \(\cos(x - vt)\) are solutions, which means \(A \sin(x - vt) + B \cos(x - vt)\) is a solution.
    • In an actual application of the differential equation there will be other constraints (called boundary conditions) that need to be satisfied.
    • For example: the value at \(t=0\) is \(0\).
    • Then one applies these conditions to solve for the constants \(A\) and \(B\), which then provide the actual solution for the problem at hand.
    • Not only is this useful for a particular problem, it turns out to be theoretically critical since many key features of the quantum world (quantization) arise through this.

  • The next two properties depend on understanding complex vectors and matricies, which we will cover in the course. For now, we'll just list these properties.

  • The second property is whether an operator is Hermitian:
    • A complex matrix (and therefore operator) \({\bf A}\) is Hermitian if \({\bf A}^\dagger = {\bf A}\), where \({\bf A}^\dagger\) is the conjugated transpose.
    • An equivalent definition for function operators needs more background, which we will defer to later.

  • The third property is unitarity, whether the operator is unitary or not.
    • A complex matrix (and therefore operator) \({\bf A}\) is unitary if \({\bf A}^\dagger = {\bf A}^{-1}\), where \({\bf A}^\dagger\) is the conjugated transpose.
    • Unitary operators are how we build quantum computations.

  • The fourth property is boundedness, which we will also defer, except to note that all matrix operators are bounded.

  • Of these, the first three matter in quantum computing but the fourth also matters in quantum mechanics.

  • One of the key results to come from the first two properties is that eigenvectors are orthogonal, and eigenvalues are real (no imaginary part).
         \(\rhd\) These eigenvectors (and eigenvalues) are the essence of any quantum system.
    We will learn more about this topic later.
 


M.6   Waves and wavefunctions

 

First, we point out: nothing in this section is needed for the course; it's here to stimulate curiosity.

Let's start this section with the obvious questions: what do we mean by a "wave", and what does it have to do with quantum computing?

We'll first address the second question:

  • Waves are not really needed for most of quantum computing.

  • One particular formulation of quantum computing, adiabatic quantum computing, along with its particular hardware, does need a bit of "wave" background to understand.

  • But most importantly, for the curious, we'll present a bit of background on quantum mechanics: this will feature Schrodinger's equation, which is about a type of "wave" (ominous quotes intended).

  • And, here, an understanding of waves will reveal one of the most startling and counter-intuitive aspects of the quantum world:
    • In the macroworld of big objects, a mathematical wave (a differential equation) describes a real physical wave, like a sound wave or a pond ripple.
    • Thus, there's some physical thing that's doing the "waving".
    • However, in the quantum world, there is no physical thing that "waves". Instead, most peculiarly, what does the "waving" are probabilities that make the physical phenomena random.
    • This will take some background to explain - we'll get there in time.

  • Also, waves are generally useful to learn about because you might one day decide to dig deeper into quantum technologies or optics.
 

Let's start with some physical phenomena:

  • The most commonly visible wave is probably some kind of water wave, like ripples from a pebble dropped in a pond:

    • To simplify, we'll look at a cross section.
    • Notice: a particular water molecule doesn't actually move away from the center, just up and down.
    • This is called a transverse wave, where the wave movement is perpendicular to the matter that moves.

  • Consider a particular water molecule on the surface, some distance from the center:
    • This molecule moves up and down (strictly vertically).
    • Think of this as away and towards its mean position.
    • At some particular time, it will be at its highest position
           \(\rhd\) maximum positive displacement, a crest.
    • At some later time, it will be at the lowest position vertically
           \(\rhd\) maximum negative displacement, a trough.

  • Suppose \(x\) describes position (distance) away from the center, and \(t\) is time.
    • Let \(f(x,t)\) = the displacement at position \(x\) at time \(t\).
    • For a fixed time \(t\), if we observe displacement along \(x\), it will vary up and down.
    • For a fixed position \(x\), if we observe the change in time, again, there will be an up and down repetition.
           \(\rhd\) Both are similar!

  • Another kind of wave comes from sound:

    • It's harder to "see" the wave here.
    • Note: matter moves in the same direction as the wave
           \(\rhd\) This type is called a longitudinal wave.
    • Again, we observe similar shapes for changing \(t\) with fixed \(x\) and for changing \(x\) for fixed \(t\).

  • A more artificial, lab-suited wave:

    • One end (the right) of a rope is attached to a wall.
    • The other end is moved up and down.
    • This will cause the rope to exhibit a wave like shape.
 

The wave equation:

  • We'll now, to simplify, make a uniformity assumption: that the disturbance pattern is the same everywhere and continues indefinitely.

  • With uniformity, let's examine how \(f(x,t)\) looks at two different \(x\) values:

    • These are examples of how \(f(x,t)\) changes with \(t\) for fixed \(x\).
    • Consider the value of \(f(x_2,t_2)\).
    • Since the wave moved rightwards, this value must have been at a previous \(x_1\) at an earlier time \(t_1\).
    • If the wave velocity is \(v\) then \(x_2 = x_1 + v(t_2 - t_1)\).
    • Thus, \(f(x_2,t_2) = f(x_2 - v(t_2-t_1), t_1)\).
    • Let's set \(t_1 = 0\) to make an observation: \(f(x_2,t_2) = f(x_2 - v t_2, 0)\).
    • Now for the key observation: the value of f at these two x-values will be the same no matter what \(t_2\) is.
    • Now define a new function \(h(x)\) to reflect this equality: \(f(x,t) = h(x-vt)\).
    • Think of \(h\) as the 'shape' function.
    • With this, a few steps of applying partial derivatives shows that $$ \frac{\partial^2 f(x,t)}{\partial x^2} \eql \frac{1}{v^2} \frac{\partial^2 f(x,t)}{\partial t^2} $$
    This is called the (one dimensional) wave equation.

  • One can derive the wave equation from physical principles (for example, rope tension, forces, and acceleration), with some assumptions.

  • The wave equation is quite general and broadly used in many applications.
 

Solutions to the wave equation:

  • Solutions to the wave equation are inevitably some form of sinusoid, or more frequently, linear combinations of them.

  • For example: $$ f(x,t) \eql A \sin(\omega(x-vt)) + B \cos(\omega(x-vt)) $$

  • One important variation is: $$ f(x,t) \eql A \sin(kx - \omega t) + B \cos(kx - \omega t) $$ where the parameters \(k\) and \(\omega\) turn out to have physical meaning:
    • The argument to a sinusoid is always an angle, as in \(\sin(\theta)\).
    • This 'angle' argument to a sinusoid is often called the phase.
    • Here, \(k\) is a parameter that converts \(x\) into phase (\(k\) is a sort of phase-to-distance ratio).
    • And \(\omega\) is a time-to-angle converter called frequency.
           \(\rhd\) A higher frequency results in the phase changing more quickly.
           \(\rhd\) Which means more cycles per unit time.

  • Perhaps the most important variation is the complex version: $$ f(x,t) \eql C e^{i(kx - \omega t)} $$

  • In case you were wondering about the negative sign in \((kx - \omega t)\), this comes about with a rightward moving wave (increasing x). When a move moves leftward we get \((kx + \omega t)\), so the more general solution is sometimes written as $$ f(x,t) \eql C e^{i(kx \pm \omega t)} $$
 

It will be useful to know what happens when waves 'encounter' other things, especially, other waves:

  • When two rigid physical objects (like billiard balls) move and collide (interact), the two objects most often change their velocity and direction of motion.

  • Let's investigate what happens when two waves interact, and when they interact with a certain non-moving rigid object.

  • Let's start with two pond ripples interacting:

    • Unlike rigid objects, waves don't "bounce" off of each other.
    • Instead, each wave continues and at any particular position, the disturbances caused by each add.

  • Let's look at one dimensional addition of two waves:
    • First, something that resembles the physical situation:

    • Let's draw these one below another and add:

      In this case, crests (and troughs) from each wave occur at the same positions.
           \(\rhd\) The crest heights and trough depths double.

    • If they're 'out of phase' one can get something like:

    • If the phase is adjusted so that crests line up with troughs, you can get perfect cancelation!

  • This phenomenon of wave addition is sometimes called interference.

  • The resulting pattern (a sort of wave) is called an interference pattern.

  • Here's a different one that results from two sources on a wall (think: two swimmers practicing leg kicks by the side).

  • And one more when one wave "interacts" with a barrier that has a tiny aperture:

    We will revisit this scenario in the famous double slit experiment.

 

Summary: why did we spend time on waves?

  • There is an equation that's similar to the wave equation, and whose solutions are also complex exponentials: $$ i\hbar \frac{\partial f(x,t)}{\partial t} \eql - \frac{\hbar^2}{2m} \frac{\partial^2 f(x,t)}{\partial x^2} + V(x,t) \; f(x,t) $$

  • This is the famous Schrodinger equation, the foundational equation of quantum mechanics.
    • Even though the function \(f(x,t)\) is not really a wave in the sense of what we've seen, it's called a wavefunction.
    • The symbol \(\Psi\) is what's commonly used and the equation is often written as: $$\eqb{ i\hbar \frac{\partial \Psi(x,t)}{\partial t} & \eql & \left( - \frac{\hbar^2}{2m} \frac{\partial^2}{\partial x^2} + V \right) \Psi(x,t) \\ & \eql & {\bf H} \; \Psi(x,t) }$$ where the term in parenthesis is a linear operator \({\bf H}\) called the Hamiltonian

  • What is truly strange is that \(\Psi(x,t)\) indirectly represents probabilities and these will vary in a wave-like manner.

  • The strangest part of this strangeness is that such \(\Psi(x,t)\) quantities can be added much like we added waves to cause interference.

  • Here's the surprising implication for quantum computing: it is this interference, as we will see, that actually forms the basis of many quantum computing algorithms.

  • Note: it might be a relief to know that, in quantum computing, everything is discrete, and we will not need Schrodinger's equation at all.
         \(\rhd\) Instead, we'll be adding the scalars in linear combinations of vectors.


Back to main review page


© 2022, Rahul Simha