You are on page 1of 60

Table of Contents

What is Quantum Mechanics all about? 4

The Success and Failure of Classical Physics 7

The First Hints of Quantization 12

Quantum Weirdness in all its Glory: the Double Slit Experiment 20

The Schrodinger Equation 27

The Collapse of the Wave Function 32

The Electron: Is it a Wave or a Particle? 34

How Quantum Mechanics Explains the Stability of Matter 38

Quantum Tunneling and our Sun 42

Quantum Superposition 46

Schrodinger’s Cat: The Cat that is Both Alive and Dead 49

Interpretations of Quantum Mechanics 52

Heisenberg’s Uncertainty Principle and the Limit of Our Knowability 55

The Relationship Between Classical and Quantum Mechanics 57

2
Meet the writer:
My name is Leonid Sarieddine; I majored in physics and pure
mathematics at the American University of Beirut and am currently doing my
master’s degree there as well. Throughout my studies, I have been constantly
fascinated by quantum mechanics since it is such a weird, yet fundamental,
subject. This book gave me the opportunity to try and explain it in a manner
which, hopefully, anybody can understand; or at least understand the basics
of it.

3
What is Quantum Mechanics all about?

At the turn of the twentieth century, scientists began to realize that the
subatomic1 world behaved radically different from our everyday experience. There
are many things in our day-to-day life which we take for granted, one of the most
obvious ones being that any object which we observe always has a definite position
and can trace a certain path when it is in motion. For example, when you throw a
pen up in the air, you can see it gradually going up and then back down to the
ground. You’ll be perhaps surprised to know that no such thing can be said of
subatomic particles. It was gradually concluded that subatomic particles don’t
follow continuous and smooth paths throughout space like our everyday
macroscopic objects do; rather, subatomic particles such as an electron can
seemingly “go through” two different holes at the same time, and yet occupy a
single specific position at a later time, almost as if it had briefly split into two
entities and then recombined!

The story gets even weirder from there. An electron can be in multiple states
of being at the same time; it can be in a state possessing energy 𝐸1 while
simultaneously also being in a state of being possessing energy 𝐸2. Make no
mistake about what i’m trying to tell you: I'm not saying that the electron had
energy 𝐸1 and then, at a later time, acquired energy 𝐸2; I'm saying it’s in both states
of being at the same time! If you are confused by this then you’re in good company;
such behavior is simply incomprehensible to us but we will see how physicists came
to these conclusions despite these unthinkable consequences.

Because of its absurdness, many people even joke about learning quantum
mechanics. I’ve seen Richard Feynman’s quote too many times: “I think I can safely
say that nobody understands quantum mechanics”. Don’t worry, you didn’t just
waste your money. Everyone who learns quantum mechanics is learning the

1
smaller than or occurring within an atom.

4
behavoir of nature, or more specifically, what we know about the behavoir of
nature. It’ll be clear to you while reading this book that the theory intuitively doesn’t
make sense, and that was Feynman’s idea. But in my opinion, that’s what make the
theory so exotic; Nature is behaving in a way that we can’t comprehend. This
sentence is both exciting and dangerous. If quantum mechanics, which we know is
true by the experiments done, opposes our intuition, what else might be
analogous? If a definitness, which should be an absolute fact, turned to be
debatable, imagine what might be hidden…

It’s been observed countless times that the electron’s behavior depends on
whether you’re looking at it or not2. If you are not looking at it, then it tends to
behave like a wave – an entity which is smeared out in space, capable of interfering
with itself just like a water wave can, for example; but then when you suddenly look
at it, it behaves like a particle occupying a single position in space. Take a moment
to reflect about how weird this is: You – and indeed everything around you – are
made of electrons and yet these tiny entities operate in ways that are completely
incomprehensible to us. We’ll discuss precisely how this was all observed in the
famous double slit experiment. Reality, as it turns out, can be stranger than fiction!
Indeed, even a science fiction writer would be hard-pressed to come up with such a
bizarre world that we find ourselves in.

This radical departure from everyday life came as a shock to the physics
community, but physics is an empirical3 science, and as experimental evidence
piled up in favor of the quantum picture of reality, we had no choice but to
abandon classical physics in favor of this new strange view of nature. You may be
asking yourself that if we can’t even attach the notion of a path or position to the
electron, then what the hell is it anyway? It can’t be a wave since sometimes it can
have a well-defined position – when we measure the electron we do indeed find
that it occupies a particular point in space – but on the other hand it can’t be a
particle since a particle doesn’t exhibit the wave phenomena, and other strange
behavior that we mentioned before, associated with the electron. The fact of the

2
By “looking at it”, I really mean measuring it.
3
Empirical is defined as : based on, concerned with, or verifiable by observation or
experience rather than theory or pure logic.

5
matter is, a subatomic particle like an electron has a dual nature to it: it behaves
sometimes as a particle, other times as a wave, but in actuality, it is neither and we
are naive in assuming that we can describe the electron as either a wave or a
particle. Keep in mind, we are the ones who defined what is a wave and what is
matter, is it really that surprising that the subatomic world didn’t directly conform
to our definitions? As we will see later, a complete description of the electron can
only be given by understanding the Schrodinger equation and the wave function
that describes it.

You may dismiss this rather bizarre behavior of subatomic particles as being
completely irrelevant to our everyday life. Why should you care about these weird
quantum effects? Well, for one thing, it turns out we would all be dead without
those weird quantum phenomena. Indeed, as we will explore later in this book, the
reason the sun shines is really because subatomic particles are weird. If electrons
behaved like classical particles do – by that I mean just like, say, billiard balls which
do not exhibit the aforementioned strange behavior – then our sun would have
been dead a long time ago, and even worse, all matter would have collapsed into
itself. There would be no planets, no stars, and no life as we know it. It is perhaps
quite ironic that we owe our very existence to the fact that our world is “strange” at
the subatomic level. Of course, the word “strange” is relative. It is strange to us
because we don’t experience such bizarre behavior in our everyday life but it is
quite remarkable that a world such as ours could emerge from an underlying
subatomic world that behaves completely differently from ours.

6
The Success and Failure of Classical
Physics

The story of physics could be said to begin with Sir Isaac Newton in the 17th
century. Newton wrote down an equation which could predict the motion of any
object in the universe with extreme precision. His celebrated law of motion(𝑓 = 𝑚𝑎)
is still used today in many areas of physics and engineering. At first sight, one might
dismiss the law as being not that significant, however if we write the law in all its
generality one begins to grasp it importance:

2
𝑑𝑥
𝑓=𝑚 2
𝑑𝑡

2
𝑑𝑥
The term 2 is the acceleration of the object. This is because velocity is the
𝑑𝑡

rate of change of position, and acceleration is the rate of change of velocity; hence
when you put these two together, you are led to the fact that acceleration is the
rate of change of the rate of change of position, which is what this second derivative
term is telling us. A derivative means rate of change, a second derivative means a
rate of change of a rate of change which is what you see in the equation.

The above equation directly relates the position of the object 𝑥 at a particular
time, and the external force 𝑓 applied on that object at that instant . So the
importance of that equation is not that it relates acceleration to external forces but
rather that it relates the position – through its rate of change – to the external force
applied allowing us to predict the position of any object as a function of time.
Predicting the position as a function of time exactly means knowing the trajectory
of that object. This is because knowing the position of the object at every instant of
time is knowing the path that it will follow. In other words, if we know the forces

7
applied on our object, then we can predict its motion and trajectory for all future
time if we solve this equation. To give you an idea of how one might apply this
equation, the following questions could all be answered by solving this equation4:
“When will a lunar eclipse happen?”, “what position in space will the Earth be after a
certain time?”, “what is the trajectory of this particular galaxy? Will it collide with
another galaxy?”, “how much thrust power should a rocket have so that it escapes
Earth’s orbit?”, “will this asteroid collide with the Earth?”, “Will this beam of steel
break if I applied a certain amount of force on it?”, “is this building stable or will it
collapse shortly?”. And the list goes on. I hope you are at least somewhat impressed
by now.

The leading thinkers of Newton’s time were astonished by this law since they
were mesmerized by its predictive power: Suddenly, the universe seemed to work
like a mechanical clock, everything was perfectly predictable, from the motion of an
apple falling from a tree to the motion of planets around distant stars. In other
words, for any given isolated object, if we know the initial conditions, we not only
can predict the future, but also learn about its past. Let’s make this even more
interesting, what I’m saying is even more powerful than you might think. This is
because5 every atom at the moment has specific initial conditions, and using what
I’ve just explained, shouldn’t the future be determined based on these initial
conditions? This is of course assuming humans with their consciousness are
nothing but an extremely complex set of atoms. We just described scientific
determinism which is a bit philosophical. But since it’s a bit far from our purposes,
I’ll leave you to debate this with yourself.

Now returning to the great Newton, what’s even more remarkable about his
discovery is that he also created the mathematics needed to properly describe the
dynamics of objects. Most physicists throughout history used already existing
mathematics to come up with new laws of nature, but Newton had to do double
work: he created the mathematics (calculus), and then discovered the laws of
4
By “solving an equation” I don’t mean necessarily analytically – using pen and paper – I also
mean numerically using, for instance, computer simulations. Essentially this means plugging the
equation in a computer and seeing the result that comes out.
5
I’m neglecting the fact that we don’t have enough computational power for this.

8
nature with the help of those mathematical tools he had discovered. Although, one
could argue that calculus would have been discovered without Newton – after all,
Leibnitz, who was another great mathematician at the time, independently
discovered calculus as well; and both Newton and Leibnitz continued the work of
Descartes and Fermat who have already laid out some of the basic ideas underlying
calculus – but the way Newton used those mathematical tools to formulate laws of
nature was truly profound. In a nutshell, Newton was the first individual to tell us
how to do real physics in a quantitative manner and essentially reduced the
physical problem of predicting the motion of objects to the purely mathematical
problem of solving a differential equation6 which is what Newton’s law actually is.
Nowadays, pretty much all physical theories are formulated in terms of these
“differential equations”. Some are more sophisticated than others, but all have the
same underlying idea that Newton used. Electromagnetism, Quantum Mechanics,
and General Relativity all are governed by differential equations. So essentially,
Newton was the first to observe nature to be modeled by the mathematical tools of
calculus and founded the first field of theoretical physics which we nowadays call
“classical mechanics”.

Newton’s laws of mechanics reigned supreme for more than 200 years until
the turn of the twentieth century when two revolutions in physics started to
happen: relativity and quantum mechanics. Einstein’s special theory of relativity
discovered that Newton’s laws of mechanics are only approximately true and
continue to hold only for objects moving at speeds low compared to the speed of
light (by low I mean something like 10% of the speed of light or lower). It was
perhaps not that much of a surprise that it took more than 200 years to discover
this flaw since in our everyday life nothing even comes close to the speed of light.
The speed of light is about 300,000 km/s or 1,080,000,000 km/h (yes, more than a
billion km per hour!). So you can imagine how the speed of your car, or any
everyday object for that matter, would pale in comparison to the speed of light.
Even our fastest rockets could travel no more than something like 0.00003 the
speed of light. To put things even more in perspective; a light beam could travel
around the whole earth 7 times in one second! And it is only in that extreme regime
(when objects start to approach the speed of light) does relativity start to predict

6
You can think of a “differential equation” as any equation which contains derivatives in it.

9
deviations from Newton's law of motion. But classical physics had an even bigger
problem than what Einstein’s relativity suggested. There were indications that
Newton’s laws are completely and utterly false when applied to the subatomic
world which is such a plot twist for physicists at the time. Again, perhaps you’ll not
be surprised to know that it took also more than 200 years to realize this since even
in the mid-1800s, there was still dispute about whether atoms actually existed and
it wasn’t until 1897 that subatomic particles – in particular, the electron – were
discovered by Thomson.

To illustrate one of the problems facing classical physics, consider the


following: the hydrogen atom consists of one proton (also called the nucleus) and
one electron. We know that the electron (which is negatively charged) and the
proton (which is positively charged) attract each other via the electromagnetic
force. Hence we can plug in the electromagnetic force in Newton’s law of motion to
solve for the motion of the electron around the nucleus – the proton is much
heavier than the electron and thats why I say that the electron orbits around the
proton. This situation is entirely analogous to how the earth orbits around the
much heavier sun due to their mutual gravitational attraction. When we solve the
equation and take into account what we know about classical radiation, we find
−11
that the electron spirals into the nucleus within a time frame of 10 seconds (i.e.
instantly). But if that were the case, then all atoms in existence would collapse since
all atoms are made of electrons and protons; electrons would just spiral into the
nucleus making the atoms collapse. Then there would be no life or planets or much
of anything in the universe! Clearly, there’s something wrong with classical physics.
It’s not that classical physics is only approximately true, but rather it’s completely
false even when analyzing this simple system of two subatomic particles.

10
The above figure illustrates the prediction of classical physics: According to
Newton’s law of motion, an electron should spiral toward the nucleus thus
collapsing the atom.

11
The first hints of Quantization

In the late 1800s, the physicist Max Planck was working on a very practical
problem of how to maximize the efficiency of light bulbs. Basically, he wanted to
know what’s the maximum amount of light he could get for a given amount of
energy. By that time, physicists already knew that light is actually an
electromagnetic wave which is just a wave consisting of oscillating electric and
magnetic fields. If that is a bit abstract to you, you could think of an electromagnetic
wave as a wave of energy hurtling through space. Electromagnetic waves of
different wavelengths7 correspond to different forms of light ranging from visible
light, to infrared, ultraviolet, x-rays, gamma rays, etc.

As is illustrated in the figure, the visible light (i.e. everything that we can see
with our eyes) are just electromagnetic waves which have wavelengths between

7
You can describe a wave using frequency or wavelength; frequency and wavelength are
inversely proportional to each other.

12
400 and 700 nanometers. For example, if you are sitting in your room right now,
what you see is visible light coming from the sun or light bulbs, as well as visible
light reflected from the surface of various objects such as your table, couch,
computer and other things that you can directly see. What you don’t see, however,
are the countless forms of light that are in your room which are not visible light, like
for example radio waves or microwaves. In fact, your cellphone is constantly
transmitting and receiving microwaves without you directly seeing those waves;
that’s how you communicate with your friends and family – through invisible
electromagnetic waves!

Now, maximizing the efficiency of light bulbs requires understanding the


intensity of light and radiation. This, however, turned out to be a difficult problem
in physics and several brilliant physicists attacked the problem beforehand but
failed. The problem grew up to be embarrassingly disastrous for the physics
community as it was dubbed “the ultraviolet catastrophe”. To illustrate the failure,
here is a graph8 comparing the experimental results to those obtaining by
theoretical prediction:

8
Technically, this refers to what is called Blackbody radiation, an object which is sort of an
idealized emitter and absorber of radiation, but again, this technicality is not important for the point
at hand.

13
This is the graph of the intensity of light (at a specific temperature of 5000k)
as a function of wavelength. Notice how different the experimental and theoretical
curves are. This was particularly distressing since the theoretical prediction was
obtained using a combination of two well-established theories of physics:
electromagnetism (theory of light) and thermodynamics (theory of heat); so both of
them or one of them must at least be partially wrong.

Remember that for physicists, just having the experimental data isn’t enough
and experiments must be based on theoretical predictions to have a sensible
theory capable of providing systematic calculations when the need arises and
capable of predicting genuinely new phenomena. For example, when Maxwell
discovered the laws of electricity and magnetism, he predicted the existence of
electromagnetic waves traveling with the speed of light; and then he postulated
that light is in fact an electromagnetic wave. This is an instance of how a theory can
produce predictions which we didn’t even think of testing before we had a theory.

Now it became Planck’s turn to attack the problem. After several failed
attempts, Planck used a desperate trick and assumed that light waves carry energy
in discrete packets proportional to a given frequency. So to be more concrete, he
assumed that a wave of frequency f must also have an elementary unit of energy:
𝐸 = ℎ𝑓

Here h is a constant to be determined experimentally (proportionality


constant). Accordingly, the possible energies that the wave could take are only
integer multiples of this elementary unit. This means that for a given frequency,
energy is not continuous but discrete, and can’t be subdivided into arbitrarily many
smaller pieces. I repeat this for emphasis: For a light wave of a given frequency,
only integer multiples of the elementary unit of energy are allowed, all other energy
levels are forbidden. A light wave of frequency 𝑓 can never possess an energy of
magnitude, say, 3.5ħ𝑓 since 3.5 is not an integer. The most common analogy of
quantiziation given is a staircase, which can only be made up of an integer amount
of steps. This is not what we observe in everyday life. In our world, energy can be
transferred continuously between objects; when you heat up your meal on a stove,
there’s no reason why you can’t supply a given amount of heat-energy. The fact of

14
the matter is, energy is quantized at the subatomic level, but we don’t see it since
the elementary unit of energy is so small that we can’t tell the difference whether
energy is supplied in a continuous or discrete manner at our macroscopic scale. We
will describe this in detail below.

Planck hoped that the Planck constant would eventually turn out to be 0
indicating that light doesn’t emit energy in discrete packets – but as we now know,
this didn’t turn out to be the case and Planck’s “desperate trick” revolutionized all of
physics. Instead, Planck found out that his new trick was able to fit the data
perfectly9 and sparked the birth of quantum physics. Basically, Planck found out
that if he assumed energy comes only in discrete/quantized packets, then he could
derive a formula which would perfectly agree with the experimental curve in the
graph above.The Planck constant was found experimentally to have the value:

−34
ℎ = 6. 626 * 10

This constant is a fundamental constant of nature and sets the scale of the
quantum world. Observe how small this constant is; the fact that quantum effects
occur only at very small scales – as opposed to our macroscopic world scale – can
be traced to the smallness of the planck’s constant. For instance, as we will later
learn in the uncertainty principle, planck’s constant sets the scale with which you
can measure a particle’s position. So essentially, it sets the scale at which even the
notion of length can be defined leading to a sort of pixelated view of our physical
world.

The smallness of Planck’s constant also explains why we don’t observe


quantization on every day scales and why we have missed our quantized nature of
14
reality for so long. Let’s take a visible light wave; say it has a frequency of 5 * 10
Hz. If we multiply this frequency by Planck’s constant, we observe that the possible

9
Essentially what Planck did was reverse engineered the problem: He looked at the
experimental curve and data first and tried to fit that data with a theoretical curve.This was how he
was led to quantization essentially.

15
energies that such a wave of this frequency can have are integer multiples of
−19
3. 313 * 10 . Hence it can take on the following energies:

−19
3. 313 * 10
−19
2 * 3. 313 * 10
−19
3 * 3. 313 * 10
.
.
.
−19
100 * 3. 313 * 10
.
.
.

The dots are meant to indicate etc.

Notice that if you take any two consecutive energies (say the nth energy and
n+1-th energy) and then you subtract them then you get the elementary unit of
energy:

−19
𝐸𝑛+1 − 𝐸𝑛 = 3. 313 * 10

But this value is so small that it can hardly be detected at all! It appears as if the
energy is continuous because the spacing between two consecutive possible
energies is so small that it is almost unobservable. Of course, sophisticated physics
equipment can detect them – and we have – but that explains why we missed the
discretization for so long, it’s that the “elementary unit of energy” is so small that it
is easy to miss.

After Planck, Einstein came up with an even more radical view: Light waves
don’t just possess a discrete set of energies, but they are discrete particle-like
packets of energy. These discrete energy packets are what we now call “photons”.
This was extremely controversial, and Planck initially refused this point of view

16
since it was well established at that point that light is a wave. Einstein came to this
conclusion by explaining the so-called photoelectric effect.

In the photoelectric effect, light waves are sent to hit a metal surface (which
contains electrons). And according to classical physics, if the light beam is intense
enough, then electrons should be ejected out of the metal surface:

So we know that light can eject electrons out of metals. But, how much light
is needed to do that? This should be intuitive: If you initially hit a surface with a
beam of light and keep increasing the power of the beam, eventually the beam will
be strong enough to knock the electrons out of the metal surface. But surprisingly,
that’s not what was observed; it turned out that knocking the electrons out of the
metal surface didn’t depend on the intensity at all (you can think about intensity as
the amount of light the metal received), but rather depended on the frequency of
light waves. It was observed that there is a frequency threshold, above which the
electrons were knocked out, and below which they were not knocked out
regardless of the intensity of the beam. Concretely, this means that if you shine a
beam of light of a frequency below the threshold frequency and keep increasing

17
the intensity of that light, (meaning you shine more and more amounts of this light
of this specific frequency) you still wouldn’t be able to knock out even a single
electron. But shining even a small amount of light consisting of a frequency above
the threshold frequency would result in electron emission. I like to imagine this as
normal collisions, photons were proved by special relativity to have momentum
which can be thought of as colliding power. The photoelectric effect is essentially
photons entering the metal atom, colliding with an electron strongly enough, and
𝐸
ejecting it out of the atom. The momentum of a photon is given to be 𝑐
where c is
ℎ𝑓
the speed of light and energy 𝐸 = ℎ𝑓 as mentioned above. So, momentum is 𝑐

which means that it is proportional to frequency. Now, the photoelectric effect


𝑥
shows us that the collision of a 10 (any number) photon with a below-threshhold
frequency/momentum can’t push the electron enough out of the atom. It might,
increase its energy level, but ionizing the atom requires a threshold momentum. On
the other hand, only one photon at threshold-momentum will be able to ionize the
atom. What will happen if the photon had much more momentum? The collision
will cause the electron to acquire velocity and we might get a current if the medium
permitted that. All of what I just mentioned can be summorized by Einstein’s
equation:

𝐾. 𝐸 = ℎ𝑓 − ℎ𝑓0

Where 𝑓0 is the threshold frequency. The equation tells us that the photon
energy absorbed by the atom ℎ𝑓 minus the threshold energy (called workfunction)
will be converted to kinetic energy of the electron ejected.

Thus, Einstein argued, light should come in discrete particle-like energy


packets whose energy is proportional to frequency, as Planck previously
hypothesized, and even a small amount (i.e. small intensity) of those light packets
ought to knock electrons as long as their frequency (and hence, their energy) is
above a certain threshold. This is vastly different from the classical world where a
particle (perhaps through collisions) can absorb any amount of energy and there is
no lower limit for the energy that it can absorb; in the subatomic world this is not
the case: the electron will not absorb any energy below a certain threshold.

18
This also settled an age-old debate dating back to Newton who believed that
light was a particle and not a wave as believed by his rivals. Afterward, a series of
experiments definitively proved the wave nature of light and it was regarded as
such since then; Maxwell’s electromagnetic theory supported the fact that light was
a wave (an electromagnetic wave) but Einstein’s discovery seemed to indicate that
light is a particle which means that light – much like the electron, as we will see later
on – has a dual nature to it, it sometimes behaves as a wave and other times as a
particle. For this discovery, Einstein won the 1921 Nobel prize10.

10
Contrary to popular belief, Einstein never won a nobel prize for his extraordinary work on
special relativity or even the more extraordinary theory of General relativity

19
Quantum weirdness in all its glory: the
Double slit experiment

After the discovery of the dual wave-particle nature of light, a physicist by the
name of Louis de Broglie proposed what seemed like a ridiculous notion at first: He
claimed that all particles – and not just light – have a dual nature to them. Even
seemingly concrete particles like the electron are in fact “matter waves” and exhibit
both wave and particle phenomena just like light does. An experiment was set out
to test this hypothesis; the results of the experiment confirmed de Broglie’s
hypothesis and its implications were highly counterintuitive so we shall explain the
experiment in detail.

The setup of the experiment is as follows: We have an electron gun which


shoots electrons, and we have a wall with two slits in them. Behind the wall with
two slits, there’s a screen that serves as our detector (i.e. it detects where the
electrons will finally land), we call it “the detector screen”. So basically the point of
the experiment is to shoot electrons from the gun, letting them pass through the
slits of the wall so that they eventually hit the detector screen and see where they
land. If we believe that electrons obey the laws of classical physics we should
predict a result similar to the one given below:

20
That is, if we just start shooting with our electron gun in random directions –
we shoot them one at a time so that the electrons don't hit each other – we expect
about half of them to pass through the right slit and the other half through the left
slit which would obviously create two strips of final electron positions on the
detector screen as shown in the figure. However, this is not what actually happens if
we do the experiment! The above figure is simply an illustration of what we
expected to happen but not what actually happened.

Unlike what we expected, we see more than two strips on the detector
screen even if we fired the electrons one by one. What actually happens is seen in
the figure below:

21
Notice that we observe that electrons don’t just land on two strips of the
detector screen as we expected them to, but rather we see many strips on the
detector screen indicating that electrons seemingly didn’t just follow straight-line
paths like they were supposed to, but for some reason, their path was somehow
“curved” so that some of them landed on many different parts of the detector
screen. You may think that the reason behind this is because electrons bounced off
each other causing their path to differ from their expected straight-line path but
again I remind you that we fired the electrons one at a time to make sure this didn’t
happen.

Upon further inspection, physicists realized that what we see on the screen is
actually an “interference pattern” similar to a pattern exhibited by waves when they
interfere with each other, like water waves for example. You see, when two waves
come into contact, they create a new wave which is a superposition of the two
constituent waves. The two waves can constructively interfere or destructively
interfere depending on how they meet (or more precisely, depending on their
wavelength and phase):

22
In the figure above, the two waves are green and blue and red is the
resultant wave. In our case with water waves, we would get the below result:

We send a water wave toward the two slits, the wave then splits into two
waves which start interfering with each other creating the interference pattern
observed in the detector screen. Observe carefully how the interference of the two
water waves creates many dark strips on the detector screen. But notice that this is
exactly analogous to the interference pattern observed when you shoot the

23
electron gun into the two slits. Hence we are led to imagine the following “picture”
of wave-like electrons:

It’s as if the electron becomes a smeared-out wave when traveling toward the
two slits and then – just like with water waves – splits into two waves that start
interfering with each other to create the interference pattern shown on the screen.
It must be noted that what we detect on the screen is a “particle”, i.e. an entity with
a well-defined position. Hence, even though the electron starts out as a wave, it
eventually localizes to a single particle when it hits the screen. It’s a peculiar feature
of subatomic particles that whenever they are “measured” – in our case, the
detector screen represents the process of measurement – they suddenly become
localized particles, but before measurement, they behave as waves! In a nutshell,
the electron is a wave when we’re not looking at it, but becomes a particle when we
look at it.

This is what we mean by the dual nature of the electron: It behaves as a wave when
it's going through the slits and then becomes a particle (i.e. a localized entity) when
it hits the screen. You may ask what happens if we try to “see” what the electron is
doing before it goes through the slits; maybe we can catch it in the act of trying to
behave as a wave directly! But as mentioned before, whenever you try and “look at

24
it” (measure it) then it suddenly becomes a particle and all you would see is a
particle going through one of the slits; hence performing such an experiment –
where you try and measure the electron before it gets to the slits – just reproduces
the result expected from classical physics, where the electron behaves like a bullet:

Because you measured its position before hitting the screen, the electron
was forced to become a particle before going through the slits, and thus, there
would be no interference or wave phenomena, just straight-line trajectories like
those predicted by classical physics.

Notice the clever way physicists managed to conclude the wave-like behavior
of electrons: They only ever measured the final position of the electrons on the
detector screen – where the electron was a particle since it was localized at a
specific position – but by observing the shape of the pattern on the screen, they
were able to conclude that such behavior can only come from waves and not
particles. So, even though we can’t directly observe the electron acting as a wave,
we can conclude that it’s a wave-based on the data shown on the detector screen.
Again, remember that the electrons were shot one by one so there was no room for
them to “hit each other” or anything of that sort, yet interference effects were clear
as day and the wave conclusion was inevitable.

25
Notice how this experiment makes the notion of a path of a particle
ill-defined. There is no trajectory or path that the electron follows. The very concept
of the trajectory of a particle through space makes no sense for a subatomic
particle. If you ask a physicist “what is the path that the electron takes through
space”, his answer would probably be “your question doesn't make sense. There is
no such thing as a path when it comes to the electron”.

As should be apparent at this point, the process of measurement has a


fundamental effect on the behavior of the electron. This is drastically different from
classical behavior: In classical mechanics, if you have an object moving through
space, it wouldn’t behave as a wave or particle depending on whether you’re
observing it or not. It’s also worth mentioning that if you perform this experiment
of shooting electrons you will not be able to predict with 100% certainty where
exactly the electron will land on the detector screen; at best you can ask for the
probability of landing in a particular position on the screen.

At this point, you may think that because of how small subatomic particles
are, there’s a possibility that we’re just not measuring stuff accurately (after all,
−13
subatomic particles operate on scales of the order of 10 meters or less), but let
me assure you that our equipment today is extremely precise and these
phenomena have been observed countless number of times and even parts of our
modern day technology like computers and phones are in fact based on these
weird quantum phenomena to some extent. Even a million years from now, when
we have developed far more advanced technology, this strange behavior of
elementary particles will still be ever present and there’s nothing we can do about
it; it’s simply how nature chooses to operate.

26
The Schrodinger equation

Before 1925, there was no formal procedure to predict the properties of


subatomic particles like Newton's law of motion did for classical physics. A precise
mathematical formulation should ideally consist of an equation that, given the state
of the system at present time, would predict the state of the system for all future
times. Newton’s law of motion did exactly that: If you give it the initial position and
initial velocity of a particle, it can predict its state of motion (i.e. its position, velocity,
and any other quantity of interest) for all future time. Even though some progress
was made in the early 1900s, such a predictive law was still missing for quantum
mechanics. The concept of a trajectory didn’t even make sense for subatomic
particles so it wasn’t obvious that such a law existed at all, and even if it did, what
would it even look like?

In the summer of 1925, a young physicist by the name of Werner Heisenberg


made a remarkable breakthrough and formulated what is now known as matrix
mechanics. Erwin Schrodinger soon followed by publishing his now-famous
Schrodinger equation which is precisely what physicists were looking for in the
same spirit as Newton’s law of motion. This equation, however, was very different
from Newton’s law of motion. Instead of describing how the trajectory of a particle
evolved in time, it instead described the evolution of a probability function
associated with subatomic particles. This was the best we could come up with and it
agreed with the experiments: Even though we can’t say what’s the precise position
the electron will be after some time, this new law could predict the probability that
it will be in a particular place/position at a particular time. Classical physics was
completely deterministic, if someone falls from the 10th floor, classical physics
allows you to calculate with certainty how much time it takes for that person to hit
the ground; quantum mechanics on the other hand could only tell you the
probability that, say, a particle will hit the ground after a given amount of time.
So the best we could hope for in this new quantum view of reality is to
probabilistically predict the properties of the electron/subatomic particles. The

27
governing equation in quantum mechanics is the so called Schrodinger equation (in
one dimension)11:

∂Ψ(𝑥,𝑡)
𝑖ħ ∂𝑡
= 𝐸Ψ(𝑥, 𝑡)

Where Ψ(𝑥, 𝑡) is called “the wave function” and is the quantity of interest in
∂Ψ(𝑥,𝑡)
quantum mechanics. ∂𝑡
refers to the time rate of change of Ψ(𝑥, 𝑡) and E is the
energy of the particle. The symbol 𝑖 is the imaginary number but there’s nothing
“imaginary” about the equation, the imaginary number is there just because it is
convenient mathematically; the equation could be written without reference to
imaginary numbers but would look a bit more complicated. The ħ that appears in
the equation is called planck’s constant which was introduced earlier.

2
The physical interpretation of the wave function is as follows: |Ψ(𝑥, 𝑡)| is the
probability12 that you will find the electron at position 𝑥 at a time 𝑡.

To give a quick and easy example, a typical wave function might look
something like this:

11
This is actually a special case of the Schrodinger equation; the full equation is fairly more
complicated but this form is sufficient for our purposes – it highlights the physical content of the
equation.
12
Technically, it’s a probability density, i.e. probability per unit length.

28
This figure is a plot of Ψ as a function of position (at some particular fixed
instant of time). Hence the y-axis refers to values of Ψ and the x-axis refers to
values of position; for every value of position on the x-axis, there is a corresponding
value of Ψ which tells us how likely you are to find the electron at this position. The
figure indicates that the electron is very likely to be near the origin since the
2
function Ψ – and hence its square |Ψ| – takes relatively high values at the origin
and near it, and then it plummets to near zero away from the origin hence
indicating that you are unlikely to find the electron away from the origin if you do
the measurement.

Now that we know what Ψ means physically we can state in simple words
what the Schrodinger equation is telling us: The ratio of the rate of change of the
probability13 function to the probability function itself is directly proportional to the
energy of that particle

This equation, named after its discoverer Erwin Schrodinger, is the


culmination of many years of hard work to get a precise mathematical formulation
of quantum mechanics. Note that doing just experiments is not enough: to have a
truly successful theory you must have a way of predicting what should happen in
your experiment, then if the experiment indeed reproduces your predicted results
you can call your theory a success. The Schrodinger equation provides just that; it
provides us with a way to predict the properties of our system which we can test
with our experiments.

You may be wondering: What exactly does it mean to predict probabilities


rather than actual values? It means the following: Suppose we have some quantum
particle trapped in a box that is made of two parts: right and left. And suppose, for
simplicity, our equation predicts that the particle should lie to the right with 70%
certainty and to the left with 30% certainty (this isn’t a realistic scenario but I only
intend to illustrate the point). Then to experimentally test this prediction, you take
this particle, put it in the box, and measure its position to see whether it lies to the
left or to the right. Then you redo the experiment many many times and see

13
Again, I am being a bit imprecise because the square of the wave function – and not the
wave function itself – represents probability.

29
whether 70% of the time it lies to the right and 30% of the time it lies to the left. It’s
intuitively obvious that increasing the number of trials makes the experiment more
authentic. For example, doing just 5 trials hardly seems like a fair experiment. You
might end up with, say, 3 of them to the left and 2 to the right; but that’s only due
to the fact that the number of trials was too low. If you go up to, say, 100 trials, you
will see it (or at least, you should) approach the predicted result. If you don’t, then
there’s a problem with the theoretical model as opposed to the experiment. In the
case of quantum mechanics, the experiments match the results predicted by the
Schrodinger equation incredibly well.

Before Schrodinger, the famous physicist Neils Bohr came up with a model of
the atom which reproduced some of the experimental results for quantum
mechanics. Bohr’s model predicted, for instance, the energy levels of the hydrogen
atom, but it was mostly an ad hoc approach. For example, Bohr’s model starts with
the assumption that angular momentum is quantized, but the Schrodinger
equation actually predicts the quantization of angular momentum (quantization
comes from imposing the single-valuedness of the wave function). So whereas in
Bohr’s case it was an assumption, in Schrodinger’s case it was a prediction. A more
successful approach was via the so-called Bohr-Sommerfeld model which quantizes
a quantity called “the action” but again, it makes certain assumptions and fails to
reproduce many of the other results. Finally, the first giant leap came through
Werner Heisenberg in his monumental 1925 paper. The Nobel prize winner Steven
Weinberg said this of Heisenberg’s breakthrough:

‘If the reader is mystified at what Heisenberg was doing, he or she is not
alone. I have tried several times to read the paper that Heisenberg wrote on
returning from Heligoland, and, although I think I understand quantum
mechanics, I have never understood Heisenberg’s motivations for the
mathematical steps in his paper. Theoretical physicists in their most
successful work tend to play one of two roles: they are either sages or
magicians ... It is usually not difficult to understand the papers of
sage-physicists, but the papers of magician-physicists are often
incomprehensible. In this sense, Heisenberg’s 1925 paper was pure magic.’’

30
Despite the fact that Heisenberg’s paper contains several “leaps of logic” of
which Weinberg speaks of, many of his steps can be at least somewhat motivated.
Perhaps the biggest leap of all is the fact that Heisenberg, instead of thinking of the
position of a particle as a number (like in classical physics), he regarded it as a
matrix; but that was the result of him trying to reproduce the so called
Ritz-combination law of frequencies (which was experimentally observed) and he
was inevitably led to this algebra of matrices despite the fact that Heisenberg
himself didn’t even know matrix algebra!14 In fact, it was the mathematician Max
Born who pointed out to Heisenberg that what he was doing was in fact matrix
algebra and he set out – along with his assistant Pascal Jordan – to put Heisenberg’s
work on a more rigorous footing. This was the birth of matrix mechanics, the first
precise mathematical formulation of quantum mechanics. Shortly afterward, Erwin
Schrodinger published his famous equation and the two approaches – The
Schrodinger approach and matrix mechanics – were shown to be equivalent in the
sense that they reproduce the same results; essentially this just means that both of
them are a different way of stating the same thing. But physicists preferred
Schrodinger’s approach since it was easier to use, it was also mathematically
cleaner since it just involved solving a single (albeit hard) equation meanwhile
Heisenberg’s matrix mechanics involved messier methods like working with infinite
arrays of numbers (infinite matrices) and the like.

14
If you happen to know Linear Algebra and are wondering how come a great physicist like
Heisenberg didn’t know matrices, then you must know that at the time, matrices were not common
knowledge among physicists. In fact, only after quantum mechanics did matrices become common
in physics.

31
The Collapse of the Wave Function

As we already established, the goal of quantum theory is to solve for the


wave function Ψ(𝑥, 𝑡). Once we know Ψ(𝑥, 𝑡), we can calculate the probability of the
particle being in any region of space.

But what happens when we actually measure the position of the particle? As
hinted in the introduction and as we clearly observed in the double slit experiment,
when we measure the position of the particle, it becomes localized to a specific
position. Notice how weird this is: before doing the measurement, the electron’s
2
position – mathematically encoded in the probability function |Ψ(𝑥, 𝑡)| – is smeared
out. Indeed, even the notion of “electron’s position” is ill-defined before we actually
measure it. The electron is not located in any particular place, it is smeared out
throughout all of space weighted by different probabilities in accordance with the
wave function Ψ. So, what we mean when we say that an electron is a “probability
wave” is that the electron is described by a probability function Ψ; before we
measure the electron’s position, we don’t know where it is and even asking the
question of “where it actually is” doesn’t make sense; all we know is that there’s a
probability that the electron might be in a particular region of space depending on
the value of Ψ in that place. You may object and claim that maybe even before we
measure the electron’s position, it must be located “somewhere” we don’t know
about and all our talk of probabilities is just a result of our ignorance of not
knowing the electron’s hiding place. But again: I refer you to the double-slit
experiment which definitively rules out this possibility; if the electron did indeed
have a well defined position before the measurement, then it wouldn’t exhibit
those interference effects observed in the double-slit experiment.

This naturally leads us to the concept of wave function collapse: Before


measurement, the wave function (and its square) is smeared out through space –
indicating the probability that the electron might be located at each point in space –
but suddenly when we measure the position of the electron we know its precise

32
position and hence the wave function “collapses” into a spike at a single point. This
is illustrated in the figure below:

This discontinuous behavior of the collapse of the wave function – from


being smeared out smoothly through space to suddenly collapsing into a spike – is
called the “measurement problem”. In particular, it’s not exactly clear why the wave
function must collapse at all upon measurement; this whole process of
measurement is shrouded in mystery because, in classical physics, we weren’t used
to nature changing its behavior depending on whether we measure it or not, but
that’s exactly how it is in quantum physics. This is one of the most famous and as
yet unsolved problems arising out of quantum physics.

33
The Electron: Is it a wave or a Particle?

The remarkable thing is that Schrodinger’s equation, at least in some sense,


solves the central mystery of quantum mechanics: Is the electron a wave or a
particle? The answer is that it’s neither and it doesn’t really matter anyway! We, as
humans, are insisting that the electron should be either a wave or a particle based
on our experience of the macroscopic world we see. In our everyday life, we don’t
see objects behaving sometimes as waves and other times as particles, hence, we
divide our objects into two criteria: waves and particles. But this is merely an
artifact of our experience as humans living in a non-quantum mechanical world.
Who says that the subatomic quantum mechanical world can even be described in
terms of our primitive wave-or-particle terminology that we describe every-day-life?
In all likelihood, the English language is simply not adequate enough to tell you
what a subatomic particle like an electron is. Fortunately for us, nature seems to
speak the language of mathematics, which is the only language suitable for
describing the quantum world, at least to the best of our knowledge it is.

The brilliant physicist Eugene Wigner famously remarked on the


“unreasonable effectiveness of mathematics in the natural sciences”; by this, he
was referring to the fact that there is no a priori reason why mathematics must
yield such a precise description of nature. It often happens that mathematicians
come up with a mathematical theory having no applications whatsoever in mind,
and these – once dubbed as being completely useless – mathematical theories turn
out to be the precise things needed to describe nature. Famous examples include
differential geometry (mathematics of curved spaces) which turned out to be able
to describe the curvature of space-time that we live in and group theory which is
essential to describe symmetries that is central to modern theoretical physics.

Schrodinger equation is remarkably successful in describing the electron


and there is no ambiguity in it, nor is there a need to say whether the wave function
is describing a “wave” or a “particle”. If somebody asks you what’s an electron, you

34
might as well just say it’s an entity described by the Schrodinger equation, full stop.
We tend to describe the electron as being a “wave” or a “particle” just because it
paints nice intuitive pictures to which we can relate, and sometimes these intuitive
pictures are very useful – indeed, they remain ubiquitous in physics textbooks – so
long as we understand that these are incomplete descriptions, whereas the
complete description can only be found in the Schrodinger equation. You may find
this unsatisfactory but remember that nature is under no obligation to cater to your
intuition. In the physics community, we all go through the steps of initially being
shocked by quantum mechanics – proceeding to a stage of frustration with its
counterintuitive consequences, until we’ve learned enough of quantum mechanics
to be able to adjust our intuition and live with its consequences.

You might find this unsatisfactory and say that you can’t identify the entity
(electron) with the object describing it (wavefunction), but I would argue otherwise.
This is because, it’s been well established by experiments that all electrons in the
universe have exactly the same charge and mass, and the only thing distinguishing
two electrons is their wave function; you might ask what if two electrons have the
same wave function? It’s a fundamental law of nature – called the Pauli exclusion
principle – that says this simply cannot happen; no two electrons can occupy the
same state – or in other words, have the exact same wave function – hence the
wave function completely characterizes an electron which allows us to identify the
electron with its wavefunction. There is a subtle point here that may have gone
unnoticed: This complete characterization is unlike anything we experience in our
everyday life; for example, you might be tempted to say that these two golf balls
are the same:

35
They certainly look the same but in actuality, they aren’t because if you zoom well
enough you are bound to eventually find differences. Meanwhile, electrons are
“actually” the same apart from their wavefunction. The only properties that the
electron possesses are charge, mass, and wavefunction (which includes an intrinsic
property called the spin). You can’t say one electron is “more curved” than the
other, or that one electron “is more round” than the other; there are no properties
that could be attributed to one unique electron. This is why you might object to
saying that the entity cannot be identified with the object describing it – it’s because
we don’t experience such objects in our daily lives; anything that we interact with
cannot be identified with a property (or indeed, any number of properties) that we
try to describe it with.

It is worthwhile to note that even though the “wave function” contains the
term “wave” in it, it doesn’t mean that it’s a function that describes a wave.
Sometimes, depending on the problem at hand, the wave function does indeed turn
out to be a function which has wave properties, and by that I mean it’s a function
which has a well defined frequency and wave-length, like for example 𝑠𝑖𝑛(𝑥); but
other times it’s a function which isn’t a description of any sort of wave. An example
2
−𝑥
of this would be 𝑒 where 𝑒 is euler’s number otherwise known as the exponential;
then it’s clear that we can’t assign a wavelength to such a function because, for one
thing, it isn’t a periodic function and remember that wavelength is the distance

36
after which the wave repeats itself. So keep in mind that “wave function” is not
necessarily a wave; it’s just a function that solves the Schrodinger equation. So,
subatomic particles are not waves or particles, they are described by wavefunctions
and can behave like waves if the wave function is wave-like and can behave as
particles if the wave function is like a spike (as we have seen in the previous figure)
or they can behave like neither if the wave function is neither wave-like nor like a
spike.

37
How Quantum Mechanics Explains
the Stability of Matter

Consider the hydrogen atom; this atom is composed of an electron and a


proton (or nucleus). You may think of an atom as mostly empty space occupied by
an electron “orbiting the nucleus”. Of course, now we know that this “orbit” picture
is quite faulty because of quantum mechanics, but you may think of an electron as
a sort of cloud surrounding the nucleus where the cloud is meant to indicate that
the electron can be “anywhere” though it is more likely to be in some regions than
others depending on its wave function.

The picture above displays the quantum picture of the hydrogen atom –
regions of space with a high density of blue dots represent a high likelihood of the
electron being there. So this figure indicates that the electron is likely to be in

38
regions near the nucleus and very unlikely to be in regions that are far away from
the nucleus.

The Schrodinger equation predicted that the electron can’t possess any
energy but only a specific set of discrete energy levels. Before elaborating on that,
let me clarify a potential question that may be on your mind: If the electron is a sort
of fuzzy probability cloud, can we speak about its energy? Yes, we can, because
remember when we are talking about its energy, we really mean what is the energy
of the electron when we look at it. This is not unlike asking the question “where is
the electron likely to be located?”. Before we measure its location, the electron is a
fuzzy probability cloud, but when we measure its position, the wave function
collapses into a single point making the question of finding its position well defined.
Similarly here, even though the electron is a probability cloud, we can ask what
possible energies could the electron possess after making our measurement.

The figure below illustrates the first few possible energy levels the electron
can possess:

Note that 𝑒𝑉 is a unit of energy

39
−19
1 𝑒𝑉 = 1. 6 * 10 𝐽

where 𝐽 stands for Joules, the standard unit of energy.

The lowest possible energy level is 𝐸1 =− 13. 6 𝑒𝑉 and is usually called the
ground state energy. The second lowest, usually called the first excited energy
state, is 𝐸2 =− 3. 4 𝑒𝑉 so this means that anything between − 13. 6 𝑒𝑉 and − 3. 4 𝑒𝑉
is forbidden, the electron can’t possess any energy strictly between
− 13. 6 𝑒𝑉 𝑎𝑛𝑑 − 3. 4 𝑒𝑉. So for instance, the electron can’t possess an energy of
− 5 𝑒𝑉, since it’s inside the forbidden region. Similarly the electron can’t possess
− 2. 5 𝑒𝑉 since − 2. 5 is between the second and third lowest energy levels, − 3. 4
and − 1. 52 respectively.

So suppose you have the electron in the ground state possessing energy
𝐸1 =− 13. 6 𝑒𝑉, and you shine a beam of light on the hydrogen atom to try and
excite the electron (i.e. to try and increase its energy); if the beam of light –
consisting of photons – possess an energy of, say, 5 𝑒𝑉, then according to the
classical picture the electron would simply absorb that beam of photons and
increase its energy going from an energy of 𝐸1 =− 13. 6 𝑒𝑉 to 𝐸𝑛𝑒𝑤 =− 8. 6 𝑒𝑉. But
quantum mechanics, on the other hand, forbids this; since − 8. 6 𝑒𝑉 is in the
forbidden energy range for the electron so the electron simply wouldn’t absorb the
incoming photons and would simply ignore them. If the beam of photons possess
an energy of 10. 2 𝑒𝑉 then the electron would indeed absorb the photons since
− 13. 6 + 10. 2 = − 3. 4, and − 3. 4 𝑒𝑉 is an allowed energy level for the electron
since it’s the first excited state.

This important observation explains the stability of matter. Remember that


we were faced with the following dilemma: If the electron is seen as a classical
particle orbiting the nucleus, then the electron will keep losing energy – as it emits
light – and spiral into the nucleus thereby collapsing the atom. However, quantum
mechanics explains why this cannot happen: It’s because the electron can’t possess
energy lower than the ground state 𝐸1 =− 13. 6 𝑒𝑉. So the electron cannot emit any

40
light or photons or lose any energy if it is already in the ground energy state. If the
electron is, say, in the 2nd excited state 𝐸3 =− 1. 52 then it can emit a photon of
energy 1. 88 𝑒𝑉 making the electron transition to the first excited state 𝐸2 =− 3. 4 𝑒𝑉
; because − 1. 52 − 1. 88 =− 3. 4 and − 3. 4 is an allowed energy for the electron,
this photon emission is allowed. Similarly, it can release a photon of energy
12. 08 𝑒𝑉 making a transition to the ground state with energy 𝐸1 =− 13. 6 𝑒𝑉. But if
the electron is already in the ground state, then no lower energy level is accessible
and hence it can no longer emit any photons.

It must be noted that these energy levels of the hydrogen atom were known
well before the Schrodinger equation predicted them; for one thing it was known
experimentally and even Bohr’s model of the atom was able to predict them but as
mentioned before, Bohr’s model was an ad hoc approach; essentially Bohr’s model
made assumptions specifically to reproduce the data regarding the energy levels of
the hydrogen atom, and failed when it came to more complicated systems other
than the hydrogen atom. Schrodinger equation, on the other hand, was much more
general and applicable to any system whatsoever.

Apart from the quantization of energy, other measurable quantities such as


angular momentum were also found to be quantized. Schrodinger’s equation also
predicted the quantization of angular momentum. Let me take a minute to
illustrate how remarkable this is: The Schrodinger equation, when applied to the
hydrogen atom problem, predicted the possible energy levels of the atom as well as
the possible values of the angular momentum, and also predicted the wave
function – or probability function – of the electron inside the atom, and they all
agreed perfectly with experiments! It was a true triumph and a testament to the
power of theoretical physics. You start with a pen and paper and then use them to
uncover the deep laws nature chooses to obey even at subatomic levels!

Even more shocking perhaps, is that entirely new measurable quantities like
“spin” emerged. Spin is a tricky concept since, unlike what the name suggests, it’s
not exactly the rotation of a particle around its axis but rather its sort of an intrinsic
property of subatomic particles like mass or charge.

41
Quantum Tunneling and our Sun

Every human, animal, and plant on Earth depends on the sun for survival; the
surprising thing is, as it turns out, the sun wouldn’t shine without weird quantum
phenomena. As we will shortly explain, what powers our sun is nuclear fusion and
the only reason nuclear fusion occurs at all is because of a phenomena called
“quantum tunneling”.

Before getting into the quantum stuff, let’s talk a bit about nuclear
fusion. Our sun is a sphere of hot gas mostly made of hydrogen and helium atoms.
At the core of the sun, hydrogen atoms – which contain a single proton – are
converted into helium atoms – which contain 2 protons and 2 neutrons. It takes
four hydrogen atoms to fuse together to make up a helium atom. However the net
mass of four hydrogen atoms is greater than a helium atom and hence this extra
mass is converted into energy15 and released in the form of light.

15
The energy to mass converging is done using Einstein’s E = mc2.

42
So the process of fusing the four protons together to create a helium atom
also releases energy in the form of light, which makes our sunshine. Indeed, this is
how all stars in the universe shine, through this process of nuclear fusion. Fusing
two protons together requires enormous amounts of energy, since the two
protons, which have the same positive charge, tend to repel each other, and hence
you need energy – lots of it – to overcome their repulsion and fuse together.

At first sight, the core of our sun seems like the perfect place to fuse protons
together; after all, the core of the sun is extremely hot, around 20 million degrees
celsius; this would imply that under such extreme hot conditions protons acquire
enormous amounts of kinetic energies which could hypothetically allow them to
overcome the electric repulsion between them. This does not turn out to be the
case; even the core of the sun is not hot enough for the protons to overcome their
electric repulsion. You’d think that given the amount of hydrogen atoms in the sun’s
core, the density of the core, and the temperature of the core, surely it’s enough to
fuse at least a small fraction of the hydrogen atoms together? The sun’s core has

43
55
more than 10 protons; so even if 0.00001% of them fuse together, this would still
imply an enormous amount of nuclear reactions which could power the sun.
Unfortunately, the answer is no. If you do the calculations, according to classical
physics, no amount of nuclear reactions whatsoever should occur in the core of our
sun or pretty much any other star for that matter. There are stars which are less
massive than our sun and have far less temperatures in their core, yet still manage
to somehow successfully undergo nuclear fusion. So what the hell is going on?

The answer, perhaps surprisingly, is quantum mechanics. Quantum


mechanics allows subatomic particles to “tunnel through” to otherwise classically
forbidden regions. By this I mean the following: Using pure classical mechanics,
even at temperatures as high as 20 million degrees, protons still don’t have enough
kinetic energy to overcome the electric repulsion between them, hence fusing
together is simply not possible classically; It’s “forbidden”. According to quantum
mechanics, on the other hand, overcoming the electric repulsion between them is
highly unlikely but still possible. Remember that quantum mechanics is inherently a
probabilistic theory. In this case, it tells us that overcoming the repulsion is very
very very unlikely but there’s still a tiny chance that it might happen; and if it does
overcome the repulsion barrier we say that the particle “tunneled through the
barrier”, hence the term “quantum tunneling”.

55
In the sun’s core there are more than 10 protons running around each
other each millisecond; at any given moment two protons are highly unlikely to fuse
together but there’s simply too many of them and hence nuclear reactions do occur
and they occur in vast amounts! To illustrate it by an analogy consider the following:
Suppose you buy a lottery ticket, the chances that you will win is probably
something like 1 in a million; but if you buy a lottery ticket billion times over,
chances are that you’ll win the lottery more than a hundred or so times. What
happens in the sun is similar but on a much bigger scale even: The probability of
−30 55
fusion is something of the order of 10 . But there are more than 10 protons
running around each other at each moment hence, as is apparently obvious, the
number of nuclear reactions that occur at each moment is still huge, and this is
what accounts for the sun shining. In some sense, the sun has to win the lottery for

44
a nuclear reaction to occur, but the sun has plenty of lottery tickets in supply (lots
of protons) and hence is constantly winning!

45
Quantum superposition

Before telling you what superposition is, I must first define what a quantum
state is. A quantum state can be thought of as a particular solution to the
Schrodinger equation, i.e., some Ψ(𝑥, 𝑡) which is a solution to the Schrodinger
equation. But in general, Schrodinger’s equation may have more than one solution,
there might be many: Ψ1 , Ψ2 , Ψ3 etc.

One particular important property of the Schrodinger equation is that it is


what we call a linear (partial) differential equation. This means if Ψ1 is a solution to
the equation, and Ψ2 is another solution then it follows that Ψ1 + Ψ2 is another

solution as well16. If you know some basic properties of the derivatives you can
verify this for yourselves by noting that if:

∂Ψ1
𝑖ħ ∂𝑡
= 𝐸Ψ1

∂Ψ2
𝑖ħ ∂𝑡
= 𝐸Ψ2

Then it follows that by adding the two equations together:

∂(Ψ1+Ψ2)
𝑖ħ ∂𝑡
= 𝐸(Ψ1 + Ψ2)

16
For those who realized that this is not a state of unit norm, yes, I am bypassing this little
technicality because we can always multiply them by constants to make it so; mentioning the
normalization just distracts from the main point and doesn’t add much to the discussion at hand.

46
Hence we clearly see that if each of Ψ1and Ψ2 satisfies the equation
individually then so does their combination. In fact it’s not hard to show that any
combination of the form 𝑎Ψ1 + 𝑏Ψ2 where a and b are ordinary numbers also
satisfies the equation. This can also be shown just by adding the two equations
together. We call such a combination 𝑎Ψ1 + 𝑏Ψ2 a superposition of two states.
Hence, when we say “superposition”, we mean such a combination of states for
some numbers a and b. Physically it means that the electron might be in any one of
those states before measurement. For example the state Ψ1can be the state where
the electron has some energy, say, 𝐸1, and Ψ2 could be the state where the electron
has energy 𝐸2. Then being in a superposition of those two states just means that
the electron is in a state, which upon measurement, will yield energy either 𝐸1 or 𝐸2.
What about a and b? They actually represent the probability of finding the electron
at Ψ1 and Ψ2 respectively and accordingly, the probability of measuring 𝐸1 and 𝐸2.

Obviously a particle being in a superposition of two states has no classical


analogue; in classical physics, a particle can only be in one state at a time. Yes, the
particle’s energy may change over time but at any one particular instant in time, it
has a fixed position, velocity and energy. Meanwhile in quantum mechanics, we’re
saying that at a fixed instant of time a particle is in a superposition of states and
only after measuring a specific property of the quantum particle – like for example,
its energy – does the particle “collapse” into a state with a definite value of the
aforementioned property (in our example, a definite value of energy). So before the
measurement takes place, this particle is sort of like a mysterious cloud of possible
energies — we can predict using the Schrodinger equation what’s the probability of
obtaining those possible energies when we measure its energy but that’s the best
that we can do. Again it’s important to emphasize how absurd some of the
implications of this are: It’s not that the particle “in actuality” has a specific energy
and we’re just not aware of it, but rather the particle “actually is” in a superposition
of various energy states. This is entirely analogous to what we have discussed
before. We said that before measuring the position of the particle, all we know is
that it is smeared out across all space weighted by probabilities; one could say that
in this case, the particle is in a superposition of “position states” i.e. there are many

47
(in fact, infinitely many) possible positions the particle may be, the wave function
allows us to calculate the probability of being in a particular position or region of
space, and only after measuring its position does the particle collapse into a specific
position. Similarly, a particle can be in a superposition of energy states, and only
after measuring its energy does the particle collapse into one energy state.

48
Schrodinger’s cat: The cat that is
both alive and dead

It’s no secret that even the founders of quantum mechanics were baffled
themselves by its implications. The macroscopic world seemed so different than the
bizarre subatomic world of particles so Schrodinger devised the following thought
experiment to illustrate the absurdity of quantum mechanics when it is applied to
our every-day large scale world:

A cat is placed in a box containing a little amount of radioactive substance. If


one of the radioactive substance atoms decays, it will set out a chain reaction which
causes poison to be released which would definitely kill the cat17. Whether one of
the atom decays or not is governed by the laws of quantum mechanics (i.e. it is
probabilistic), hence essentially the atom is in a superposition of two states,
“decaying” and “not decaying” much like before we had a particle in a superposition
of two energy states. Now comes the question: Is the cat alive or dead? It seems like
an absurd question to ask but the laws of quantum mechanics force us to do so.
The atom is in a superposition of two states, which also seemingly forces the cat to
be in a superposition of two states: If the atom decays then the cat dies, if the atom
doesn’t decay then the cat lives, hence since the atom – being governed by the
probabilistic laws of quantum mechanics – seems to be in a superposition of
“decays” and “not decays”, it seems that the cat is in a superposition of “alive” and
“dead”!

17
The details of how that happens doesn’t matter but if you must know: Schrodinger
imagined that the decay of the material triggers a geiger counter – a device which can measure
radioactivity – which releases a hammer that shatters a glass of cyanide which is extremely
poisonous, thereby killing the cat.

49
Schrodinger regarded this as complete nonsense (that was the whole point
of his thought experiment! To illustrate absurdity). Surely, the cat can’t be both alive
and dead at the same time and there is good reason to believe why this is so. After
all, this thought experiment can be replicated but with a human being replaced by a
cat, and surely a human cannot be in a superposition of “alive” and “dead”. Because
if the human lived, he or she will not remember being “partially alive and partially
dead”, they will only recall their regular life like a normal person does. The majority
of the physics community would say that the cat – being a macroscopic object – is
either alive or dead even before you measure its state but there is no complete
consensus on the subject matter since Schrodinger’s cat experiment is really the
same as the measurement problem which we mentioned before. The questions still
remain of how and why exactly does the wave function collapse? Does it require
conscious observers18 to collapse the wave function? One of the leading quantum
physicists Eugene Wigner believed so at the beginning but later abandoned this
view. There is no shortage of opinions on this delicate issue; however, even after
nearly 100 years of intense debate and research, there is no resolution to this issue.
The question of the wave function collapse and the measurement problem remains
one of the hardest (perhaps the hardest) questions in all of physics.

18
Then you go down the (perhaps even bigger) rabbit hole of trying to define what exactly is
consciousness.

50
Our macroscopic world is concrete and unambiguous – objects have definite
positions and cannot occupy two states at once, whereas the quantum world is
uncertain and probabilistic where objects can be in many states at once; this
experiment tries to entangle these two radically different worlds together by tying
the fate of a macroscopic object (the cat – which is part of our certain macroscopic
world) to that of a microscopic object (the atom – governed by the probabilistic
world of quantum mechanics).

51
Interpretations of Quantum
Mechanics19

Before we start this section, it’s important to summarize what we previously


explained. Quantum mechanics is mainly described by Schrodinger’s equation
where the solution is a wave function that gives us a probability distribution of the
intended physical variable. In other words, there’s no definite answer to where the
particle is or how much velocity it has. The model says that the whole system is in a
superposition of many states weighted by their corresponding probabilities.

All this is very interesting, but what do we make of it? If we did an experiment
100 times (under the same conditions), we’ll have different answers corresponding
to the wave function’s probabilistic distribution. Why is this happening? How can
the same initial conditions yield different results? Is probability a part of how nature
operates? Quantum mechanics is consistent till this point with all experiments
made. So, the question is not if quantum mechanics works, but how quantum
mechanics works. As Richard Feynman says: “I think I can safely say that no one
understands quantum mechanics”. Physicists, to this day, don’t have a general
consensus on why the quantum phenomenon exists.

It’s important to differentiate between two ways we can understand a


scientific theory. The first way is from an ontic point of view, stating that the
theory is indeed 100% true (it’s exactly how nature behaves). The second way is
from an epistemic point of view where we take into consideration the boundaries
of the human brain, assuming that the phenomenon is deeper than what we
perceive and a higher order of knowledge may exist that we’re not aware of. For

19
This section is directly taken off the first volume of the “Elegant Description of Reality”
Book. If you read that book before, I think even repeating it after reaching this far will be really
beneficial and give you more understanding quantum interpretations.

52
example, the indeterministic and indefinite nature of particles in quantum physics
can be either seen as a limitation of the human mind in analyzing the real picture,
or quality of nature implying that nature works in probabilities.

The first and most popular interpretation advocated by Heisenberg and Bohr
is the Copenhagen interpretation. They explain that after an hour, the cat is both
dead and alive, a superposition between the two states, and it will stay that way
until an observation (measurement) is made where reality collapses into one state.
So, when the chamber is closed, the cat is simultaneously dead and alive (this was
the absurd part according to Schrodinger). It’s important to note that a
measurement is an interference in the state of the studied case. For example, a
photon used in the measurement is what collapses the wave function from a
superposition of states to a single state. This collapsing is the movement from
quantum to classical physics. This interpretation, however, is more mathematical
than physical. And here, the Copenhagen interpretation splits between a weak and
strong Copenhagen interpretation.

Heisenberg, who’s behind the weak interpretation, didn’t believe that nature
behaved in the way the equations implied, however, we know that they work. So 35
he preferred the stance to “shut up and calculate” meaning that we have an
incredible framework that agrees with every experiment. Diving more deeply won’t
take us anywhere so it’s an epistemic look at the theory. Bohr on the other hand,
who’s behind the strong interpretation, advocated the full ontic view on the
formalism of quantum mechanics. He doesn’t only say that the wave function
describes the behavior of an electron, but also, it has a real physical meaning and it
is the form of existence of an electron. So, Heisenberg says that this formalism is
only mathematical and it’s not real, while Bohr says that the formalism is the
description of how reality exists itself. Of course, this is before the collapse where
we think classically of the electron.

Now, let’s dwell more on the ontic vs epistemic point of view. We know that
if we throw a dice, it will have a 1 6 probability of hitting each number and we don’t
know for sure which one we’ll get after the throw until we look. Describing this in
quantum terminology, the system is made out of six possible states each having a

53
probability of 1 6. According to the Copenhagen interpretation, when the dice fall,
it’s in a superposition state of the 6 possibilities and only when we look, we’re
collapsing the wave function to one state. But even in this case, theoretically, we
can study the initial velocity, position, and rotation of the throw along with the air
resistance, so that we don’t need to use probability as we can predict exactly what
we’ll get. Einstein didn't accept that nature is throwing dice, and preferred the
explanation that we don’t understand how nature picks each event. And here you
can see how the ontic vs epistemic views can determine our interpretation of the
Copenhagen interpretation.

The next interpretation is the many-worlds interpretation which will seem


like a science fiction movie idea at first. It tells us that when we open the chamber
of the cat and we find what happened, two dimensions (worlds) are created where
one is entangled (linked) with the dead cat and the other one is entangled with the
living cat. The two new realities have no interaction in any way. From here we can
36 go to another interpretation called the quantum information theory. Unlike the
Copenhagen theory where many states exist and an observer or measuring
instrument collapses the system into one state, information theory says that
quantum mechanics describes what the observer sees rather than the real world
itself. So, what we acquire in our observation or measurement is what we
encounter and it might have some errors in an abstract sense.

The Copenhagen interpretation can also be seen as reversible according to


Heisenberg. This means that after the collapse of the wave function, it remains in
the same state and the collapse is only in the observer’s measurement. Bohr on the
other side believes it’s an irreversible process and the measurement permanently
alters the state of the system. Many worlds theory is reversible, the constant
splitting of reality won’t alter the wave function and continuous splitting will keep
occurring infinitely.

54
Heisenberg’s uncertainty principle and
the limit of our knowability

We have talked about how we can measure the position of an electron,


similarly we can measure the velocity and momentum of a particle. The momentum
𝑝 of a particle is just the product of its mass and velocity 𝑝 = 𝑚𝑣. Heisenberg’s
uncertainty principle says that you cannot measure the position and momentum of
the particle at the same time with perfect precision.

ħ
∆𝑥∆𝑝 ≥ 2

Where ∆𝑥 refers to the uncertainty in position and ∆𝑝 refers to the


uncertainty in momentum. It’s important to realize that the uncertainty principle
has nothing to do with the precision of our equipment but rather with the intrinsic
properties of subatomic particles, it is simply not possible to measure both position
and momentum with extreme precision no matter how sophisticated your
equipment might be.

Suppose you wanted to measure the position of a particle with near perfect
precision, that will mean that you would want ∆𝑥 to be extremely small, basically
very near zero, but this would imply that ∆𝑝 – the uncertainty in momentum – has
ħ
to be very large because the inequality ∆𝑥∆𝑝 ≥ 2
must always be satisfied. This
means that the more precisely you try to measure the position of the particle, the
less precise you can measure its momentum. And the opposite is also true; the
more precisely you try to measure a particle’s momentum the less precise you can
measure its position. This also implies that you cannot have a 100% certain
measurement – again, regardless of how sophisticated your equipment might be –
because that would immediately violate the inequality. A 100% certain

55
measurement in position would mean that ∆𝑥 = 0 which would imply that
ħ
∆𝑥∆𝑝 = 0 and that would in turn imply that 0 ≥ 2
, obviously a contradiction.

The uncertainty principle essentially reinforces the idea that the subatomic
world is “fuzzy” and seems to point toward a pixelated view of the world. It
essentially sets the scale at which even the notion of length can be well defined.
This is because, as we have emphasized, the uncertainty principle has nothing to do
with equipment, but is a statement about the intrinsic knowability of position and
momentum; it tells you the limit of what is possible to know about a particle’s
position and momentum regardless of how advanced your technology might be.

56
The Relationship between classical
and quantum mechanics20

At this point an obvious question comes to mind: If the world is inherently


quantum mechanical, how does the classical everyday life emerge from this
quantum weirdness? By that I mean the following: Consider a golf ball. This ball is
24
made of more than 10 atoms which are in turn made of subatomic particles which
in turn obey the laws of weird quantum mechanics. But the ball does not, the ball
perfectly obeys the laws of classical physics i.e. Newton’s laws so how come that the
constituent particles obey one set of laws but the entity that the particles constitute
(i.e. the golf ball) obey another set of laws? We will now see how quantum
mechanics laws of nature – in particular, the Schrodinger equation –contains the
classical laws.

To answer this question we must first introduce the notion of an expectation


value. Suppose we have some quantity x which can take n different values, each
with a different probability. We denote the possible values of 𝑥 by 𝑥1,𝑥2,𝑥3,...,𝑥𝑛and
their corresponding probability by 𝑝1,𝑝2,𝑝,...,𝑝𝑛. The expectation value of x, written
≺ 𝑥 ≻, is just the sum over all possible values of 𝑥 weighted by their probability.

𝑛
≺𝑥 ≻ = ∑ 𝑥𝑖𝑝𝑖 = 𝑥1𝑝1 + 𝑥2𝑝2 + .... + 𝑥𝑛𝑝𝑛
𝑖=1

So for a quick and easy example, suppose x can take the values of 1,2,3, with
probabilities 0.5, 0.25, 0.25 respectively. Then:

≺𝑥 ≻ = 1 * 0. 5 + 2 * 0. 25 + 3 * 0. 25 = 1. 75

20
This topic is slightly more advanced than the rest as you’ll see. Don’t get demotivated by
the mathematics inside it, it will really seem simple at the end.

57
So basically, the expectation value of a certain quantity 𝑥 represents a
weighted average of the possible values of that quantity. By “weighted” I mean each
value is weighted by its probability.

For our purposes, we are interested in the expectation value of the position
of a particle. Since the particle can take infinitely many possible positions (it can
take any real number) then the sum is replaced by an integral, hence

+∞
2
≺𝑥 ≻ = ∫ 𝑥|Ψ(𝑥, 𝑡)| 𝑑𝑥
−∞

2
Remember that |Ψ(𝑥, 𝑡)| has the interpretation of probability, hence when
we multiply it by x and integrate (sum) over all values of x it’s exactly analogous to
the previous formula when x takes a finite set of discrete values.

Why am I telling you this? Because as it turns out, the expectation value of
position obeys Newton’s classical law of motion!

2
𝑑 ≺𝑥≻
2 = ≺𝑓 ≻
𝑑𝑡

Hence the acceleration of the weighted average of the position is equal to


the weighted average of the force. Notice that ≺ 𝑥 ≻ depends on Ψ(𝑥, 𝑡) which in
turn is governed by the Schrodinger equation. Hence the above formula basically
says that the Schrodinger equation implies that Newton’s laws hold in an average
sense. This consequence of quantum mechanics is called Ehrenfest’s theorem.

To be more concrete how this comes about; start out with the definition of
the expectation value of position:

+∞
2
≺𝑥 ≻ = ∫ 𝑥|Ψ(𝑥, 𝑡)| 𝑑𝑥
−∞

58
Then compute its rate of change. But notice computing the rate of change of
≺ 𝑥 ≻ also involves computing the rate of change of Ψ(𝑥, 𝑡) which is governed by
the Schrodinger equation. Hence using Schrodinger equation you can compute the
rate of change of Ψ(𝑥, 𝑡), which would lead to the rate of change of ≺ 𝑥 ≻. When
you compute that rate of change twice (i.e. find the second derivative of ≺ 𝑥 ≻
using the Schrodinger equation) you find that Newton’s law of motion is indeed
satisfied in an average sense:

2
𝑑 ≺𝑥≻
2 = ≺𝑓 ≻
𝑑𝑡

All this is well and good but now we can ask an even deeper question, which
lies at the heart of classical-quantum correspondence: is the converse statement
true? That is, if we start from Ehrenfest’s theorem – which tells us that the
expectation value follows the classical laws of motion – not knowing anything about
quantum mechanics, can we derive the Schrodinger equation? The answer is
remarkably yes21! So if we start from the assumption that the weighted average of
the position of a particle satisfies the classical law of motion and ask what’s the
corresponding equation governing the probability function, then the answer is
Schrodinger’s equation. In a nutshell, Imposing that Newton’s laws hold in an
average sense directly leads to the Schrodinger equation.

As you can see, Schrodinger’s equation might look different from classical
mechanics, but as we observed, it encodes classical physics in it, as it should;
because as we noted before, the classical world must emerge out of the quantum
world. Perhaps you’ll also be interested to know that this is just the tip of the
iceberg; there are deeper connections between the theories which are rooted in the
symmetries of space-time. But that is a complicated topic which I am not going to
delve into, but I thought it was worth mentioning. With that being said, I would like
to conclude this book with the following quote:

21
Proving Ehrenfest’s theorem from Schrodinger’s equation is easy if you know calculus; but
the proof of the converse statement is not obvious at all since it requires the use of a sophisticated
mathematical theorem called Stone’s theorem on one-parameter unitary groups. Once you know
the theorem, you are directly led to the Schrodinger equation.

59
“Not only is the universe stranger than we think, it is stranger than we can
think”

– Werner Heisenberg

60

You might also like