Waves and Electromagnetism Wavefunctions Lecture Notes: Alessandro de Angelis University of Udine, December 2012
Waves and Electromagnetism Wavefunctions Lecture Notes: Alessandro de Angelis University of Udine, December 2012
Waves and Electromagnetism Wavefunctions Lecture Notes: Alessandro de Angelis University of Udine, December 2012
Wavefunctions
Lecture notes
Alessandro De Angelis
1
2.8 Composition of waves . . . . . . . . . . . . . . . . . . . . . . . . 38
2.8.1 Boundary conditions and steady waves . . . . . . . . . . . 38
2.8.2 Beats . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
2.8.3 Group velocity and phase velocity . . . . . . . . . . . . . 42
2.9 Waves in three dimensions . . . . . . . . . . . . . . . . . . . . . . 44
2.9.1 Spherical waves . . . . . . . . . . . . . . . . . . . . . . . . 44
5 Geometrical optics 64
5.1 Light propagation through different materials; transmission of
electromagnetic waves . . . . . . . . . . . . . . . . . . . . . . . . 64
5.2 Huygens’ principle . . . . . . . . . . . . . . . . . . . . . . . . . . 64
5.3 The laws of reflection and refraction . . . . . . . . . . . . . . . . 65
5.3.1 The Fermat principle . . . . . . . . . . . . . . . . . . . . . 68
2
7.4.2 Time dilation . . . . . . . . . . . . . . . . . . . . . . . . . 86
7.4.3 Length contraction . . . . . . . . . . . . . . . . . . . . . . 87
7.5 Invariance of the interval . . . . . . . . . . . . . . . . . . . . . . . 87
3
11 The Schrödinger equation 101
11.1 Formulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
11.2 Interpretation of the wavefunction . . . . . . . . . . . . . . . . . 101
11.3 The operation of measurement; collapse of the wavefunction . . . 101
11.4 Reality, ortodoxy, agnosticism . . . . . . . . . . . . . . . . . . . . 101
11.5 Expectation values . . . . . . . . . . . . . . . . . . . . . . . . . . 101
11.6 Momentum, and operators . . . . . . . . . . . . . . . . . . . . . . 101
11.6.1 Angular momentum . . . . . . . . . . . . . . . . . . . . . 101
11.7 The Hamiltonian operator . . . . . . . . . . . . . . . . . . . . . . 102
11.7.1 * A quantum view of Nöther’s theorem . . . . . . . . . . 102
4
Chapter 1
5
Figure 1.1: The temperature field; isotherms.
a measure of how much heat is flowing, and can be defined as the amount
of thermal energy that passes, per unit time and per unit area, through an
infinitesimal surface element, perpendicular to the direction of flow. The vector
points in the direction of flow. In symbols: if ∆J is the thermal energy that
passes per unit time through the surface element ∆a, then
where ~ef is the versor of the heat flow. Examples of the heat flow vector are
also shown in Figure 8.1.
in <3 , where each of the partial derivatives is evaluated at the point (x, y, z).
One can think of the symbol ∇ as being “applied” to a real-valued function T
to produce a 3-dimensional function ∇T .
Is ∇T (x, y, z) a vector? Of course it is not generally true that any three
numbers form a vector. It is true only if, when we rotate the coordinate system,
6
the components of the vector transform among themselves in the correct way
for a vector. So it is necessary to analyze how these derivatives are changed by
a rotation of the coordinate system. We shall show that ∇T (x, y, z) is indeed
a vector: the derivatives do transform in the correct way when the coordinate
system is rotated. We can observe that, by the total differential theorem,
∂T ∂T ∂T
dT = dx + dy + dz (1.2)
∂x ∂y ∂z
for any d~r = (dx, dy, dz). And since df is a scalar and d~r is a vector, ∇T (we
can now call it ∇T~ ) must be a vector (of course a demonstration based on the
transformations of coordinates is also possible, but ennoying).
Is ∇~ a vector? Strictly speaking, no, since ∂ , ∂ and ∂ are not numbers.
∂x ∂y ∂z
But it helps to think of it as a vector, as we shall see. The process of “applying”
∂ ∂ ∂
∂x , ∂y , ∂z to a real-valued function T (x, y, z) can be thought of as multiplying
the quantities:
∂ ∂T ∂ ∂T ∂ ∂T
(T ) = , (T ) = , (T ) =
∂x ∂x ∂y ∂y ∂z ∂z
~ ) · (d~r) .
dT = (∇f (1.3)
7
Figure 1.2: Divergence.
8
Figure 1.3: Curl.
~ · (∇
2. ∇ ~ × F~ )
~ · (∇T
3. ∇ ~ )
~ ∇
4. ∇( ~ · F~ )
~ × (∇
5. ∇ ~ × F~ )
You can check that these are all the legal combinations.
1. and 2. ∇ ~ × (∇T
~ ) and ∇ ~ · (∇
~ × F~ ). The first two are identically zero. Let
us check it for the first one: one has, by the Schwartz’s lemma:
We when thus demonstrated that the curl of a gradient is zero, which is easy
to remember because of the way the vectors work. It can be demonstrated that
9
if the curl of a field F~ is zero, then this field is always the gradient of a scalar
field: there is some scalar field φ such that F~ = ∇φ.~
Also ∇~ · (∇~ × F~ ) = 0, and there is a similar theorem stating that if the
divergence of F~ is zero, F~ is the curl of some vector field A.
~
3. ∇~ · (∇T
~ ). Let us examine the third expression now. For a real-valued
function f (x, y, z), the Laplacian1 of T , denoted by ∇2 T , is defined as
2 2 2
∇2 T (x, y, z) = ∇ ~ )= ∂ T +∂ T +∂ T .
~ · (∇T (1.6)
∂x2 ∂y 2 ∂z 2
Sometimes the notation ∆T is used instead.
The Laplacian can be thought as a scalar operator
2
∂2 ∂2
2 ∂
∇ = + 2+ 2
∂x2 ∂y ∂z
~ ∇
4. ∇( ~ · F~ ). It is a possible vector field, which may occasionally come up (for
example, see next point).
~ × (∇
5. ∇ ~ × F~ ). Let us compare this expression with the vector identity
~ × (B
A ~ × C)
~ = B(
~ A~ · C)
~ − (A
~ · B)
~ C~.
~ and B
In order to use this formula, we should replace A ~ by the operator ∇
~ and
~ ~
put C = F . If we do that, we get
~ × (∇
∇ ~ × F~ ) = ∇(
~ ∇~ · F~ ) − ∇2 F~ . (1.7)
work was pivotal to the development of mathematical astronomy and statistics. He was one
of the first scientists to postulate the existence of black holes and the notion of gravitational
collapse.
10
proportional to d, the distance between the plates (for a given temperature
difference, the thinner the slab the greater the heat flow). Letting J be the
thermal energy that passes per unit time through the slab, we write
J = k(T2 − T1 ) A/d .
11
Figure 1.4: Spherical coordinates.
12
Figure 1.6: Line integral and gradient theorem.
gradient. The relation contains a very simple idea: since the gradient repre-
sents the rate of change of a field quantity, if we integrate that rate of change,
we should get the total change (like in the fundamental theorem of calculus).
Let us first introduce the concept of line integral for a vector field. The line
integral from a point a to a point b of the curve is nothing but the integral of
the dot product of the value of the function times the line element (i.e., at each
point the value of the function is weighted by the cosine of the angle formed by
the vector itself and by the tangent to the line):
Z b
~
E(x, y, z) · d~l
a
(see Figure 1.5). For example, the work done by the force F~ from a to b along
a given path is the line integral of F~ along that path.
Suppose we have a scalar function of three variables T (x, y, z). Starting at
point a, we move by a small distance d~l1 (Figure 1.6). The function T will
change by an amount
dT = ∇T~ · d~l1 .
13
A geometrical interpretation will make use of an example. Suppose you
wanted to determine the height of the Eiffel Tower. You could climb the stairs,
using a ruler to measure the rise at each step, and adding them all up, or you
could place altimeters at the top and the bottom, and subtract the two readings;
you should get the same answer either way.
Line integrals ordinarily depend on the path taken from a to b. But the right
side of in the gradient theorem makes no reference to the path - only to the end
points. Evidently, gradients when the special property that their line integrals
are path independent:
Rb
Corollary 1: a (∇T ~ ) · d~l is independent of the path taken from a to b.
Corollary 2: (∇T )·d~l = 0, since the beginning and end points are identical,
~
H
in this case tradition dictates that “outward” is positive, but for open surfaces
it is, again, arbitrary.
We will identify sometimes a flat surface with a vector perpendicular to the
surface itself, and with intensity equal to the area of the surface. The expression
of the surface integral becomes then
Z
~h · d~a .
S
If ~h describes the flow of a fluid (mass per unit area per unit time), then
~h · d~a represents the total mass per unit time passing through the surface -
R
S
hence the alternative name, flux Φ. In the case of heat flow, we may think: ~h is
the “current density” of heat flow and the surface integral of it is the total heat
current directed out of the surface: that is, the thermal energy per unit time
(joule per second).
14
Figure 1.7: The closed surface S defines the volume V. The unit vector ~n is the
outward normal to the surface element da, and ~h is the heat-flow vector at the
surface element.
We can generalize this idea to the case to any vector field; for instance, it
might be the electric field. We can certainly still integrate the normal component
of the electric field over an area if we wish. Although it does not appear to be
the flow of anything, we still call it the “flux”. We say flux of E~ through the
surface S the quantity Z
Φ= E~ · d~a .
S
We thus generalize the word “flux” to mean the “surface integral of the normal
component” of a vector.
They also write d3 V instead of dV. We will use the simpler notation, and assume that you
remember that an area has two dimensions and a volume has three.
15
Figure 1.8: A volume V contained inside the surface S is divided into two
pieces by a “cut” at the surface Sab . We now when the volume V1 enclosed in
the surface S1 = Sa + Sab and the volume V2 enclosed in the surface
S2 = Sb + Sab .
Since ~n1 = −~n2 , the sum of the fluxes through S1 and S2 is just the sum of two
integrals which, taken together, give the flux through the original surface S.
16
~ out of a small cube.
Figure 1.9: Computation of the flux of C
We can similarly subdivide again the volume - say by cutting V1 into two
pieces. You see that the same arguments apply. So for any way of dividing
the original volume, it is generally true that the flux through the outer surface,
which is the original integral, is equal to a sum of the fluxes out of all the little
interior pieces. Then we can divide it by a large number of small cubes.
(since we are considering a small cube, we approximate the integral by the value
of Cx at the center of the face - which we call the point (1) - multiplied by the
17
area of the face). Similarly, the flux out of face 2 is approximately
Φ2 = Cx (2) ∆y∆z .
Cx (1) and Cx (2) are, in general, slightly different. If ∆x is small enough, we
can write
∂Cx
Cx (2) = Cx (1) + ∆x
∂x
(there are, of course, higher order terms, but we are considering the limit for
∆x → 0). So the flux through faces 1 and 2 is
∂Cx
Φ1 + Φ 2 = ∆x∆y∆z .
∂x
The derivative should really be evaluated at the center of face 1; that is, at
(x, y + ∆y/2, z + ∆z/2). But in the limit of an infinitesimal cube, we make a
negligible error if we evaluate it at the corner (x, y, z).
Applying the same reasoning to each of the other pairs of faces, we find
∂Cy
Φ3 + Φ 4 = ∆x∆y∆z ,
∂y
and
∂Cz
Φ5 + Φ 6 = ∆x∆y∆z .
∂z
The total flux through all the faces is the sum of these terms. We find that
Z
~ · d~a = ∂Cx + ∂Cy + ∂Cz ∆x∆y∆z
C
cube ∂x ∂y ∂z
and so we can say that for an infinitesimal cube
dΦ = (∇ ~ · C)dV
~ .
We have shown that the outward flux from the surface of an infinitesimal
cube is equal to the divergence of the vector multiplied by the volume of the
cube. We now see the meaning of the divergence of a vector. The divergence of
~ per unit volume,
a vector at the point P is the flux - the outgoing “flow” of C
in the neighborhood of a point P .
We have connected the divergence to the flux out of an infinitesimal volume.
For any finite volume we can use the fact we proved above - that the total
flux from a volume is the sum of the fluxes out of each part. We can, that is,
integrate the divergence over the entire volume. This gives us the theorem that
the integral of the normal component of any vector over any closed surface can
also be written as the integral of the divergence of the vector over the volume
enclosed by the surface. This theorem is named after Gauss3 .
I Z
~
C · d~a = ~ · C)
(∇ ~ dV
S V
3 Carl Friedrich Gauss (Brunswick, 1777 - Göttingen, 1855) was a German mathematician,
generally regarded as one of the greatest mathematicians of all time for his contributions
to number theory, geometry, probability theory, geodesy, planetary astronomy, the theory of
functions, and potential theory (including electromagnetism). Gauss was the only child of
poor parents. He was rare among mathematicians in that he was a calculating prodigy.
18
~ around the curve Γ. Right: The
Figure 1.10: Left: the circulation of C
circulation around the whole loop is the sum of the circulations around the
two loops Γ1 and Γ2 .
19
Figure 1.11: Some surface bounded by the loop Γ is chosen. The surface is
divided into a number of small areas, each approximately a square. The
circulation around Γ is the sum of the circulations around the little loops.
The integral along Γab will have, for the curve Γ2 , the opposite sign compared
to Γ1 because the direction of travel is opposite - we must take both our line
integrals with the same “sense” of rotation. Following the same kind of argument
we used before, you can see that the sum of the two circulations will give just
the line integral around the original curve Γ: the parts due to Γab cancel. The
circulation around the one part plus the circulation around the second part
equals the circulation about the outer line.
We can continue the process of cutting the original loop into any number
of smaller loops. When we add the circulations of the smaller loops, there is
always a cancellation of the parts on their adjacent portions, so that the sum is
equivalent to the circulation around the original single loop.
Now let us suppose that the original loop is the boundary of some surface.
There are, of course, an infinite number of surfaces which all have the original
loop as the boundary. Our results will not, however, depend on which surface
we choose. First, we break our original loop into a number of small loops that
all lie on the surface we have chosen. No matter what the shape of the surface,
if we choose our small loops small enough, we can assume that each of the small
loops will enclose an area which is essentially flat. Also, we can choose our small
loops so that each is very nearly a square (Figure 1.11). Now we can calculate
the circulation around the big loop by finding the circulations around all of the
little squares and then taking their sum.
20
Figure 1.12: Computing the circulation of C around a small square.
Since
∂Cx
Cx (3) ' Cx (1) + ∆y
∂y
and
∂Cy
Cy (4) ' Cy (2) + ∆x
∂x
one can write, at first order,
I
~ ~ ∂Cy ∂Cx
C · dl = − ∆x∆y .
∂x ∂y
21
Neglecting second order terms, the derivative can be evaluated at (x, y). The
above expression can be written in vector form:
I
~ · d~l = ∇
C ~ ×C~ · ∆~a .
The circulation around any loop Γ can now be easily related to the curl of the
vector field. We fill in the loop with any convenient surface S, as in Figure 1.11,
and add the circulations around a set of infinitesimal squares in this surface;
the sum can be written as an integral.
Our result is a very useful theorem called Stokes’ theorem4
I Z
~ · d~l = (∇
C ~ × C)
~ · d~a (1.12)
Γ S
Ireland, he spent all of his career at University of Cambridge, where he served as the Lucasian
Professor of Mathematics from 1849 until his death. Stokes made seminal contributions to
fluid dynamics (including the Navier-Stokes equations), optics, and mathematical physics.
22
Let’s show something else. Suppose we have any scalar field φ. If we take its
~ the integral of this vector around any closed loop must be zero:
gradient, ∇φ,
I
~ · d~l = 0 .
∇φ
loop
over any surface; but this means that the integrand is zero:
~ × (∇φ)
∇ ~ = 0.
and thus
~ = ∇ϕ
C ~ .
Note that I can choose arbitrarily the point A; this reflects on the fact that,
adding an arbitrary constant to ϕ, the above relation is still valid.
(i.e., its value must be infinite at the origin, see Figure 1.13) for a sketch.
5 Paul Dirac (1902 -1984) was an English theoretical physicist who made fundamental con-
tributions to the early development of both quantum mechanics and quantum electrodynamics.
Among other discoveries, he formulated the Dirac equation, which describes the behaviour of
fermions, and predicted the existence of antimatter. Dirac shared the Nobel Prize in Physics
for 1933 with Erwin Schrödinger, for his contributions to Quantum Mechanics.
23
Figure 1.13: A sketch of the function δ(x).
δ 3 (~r) = δ(x)δ(y)δ(z) .
24
Part I
Waves and
electromagnetism
25
Chapter 2
Waves
natural philosopher, alchemist and theologian, who has been considered by many to be the
greatest and most influential scientist who ever lived. His monograph Philosophiae Naturalis
Principia Mathematica, published in 1687, laid the foundations for most of classical mechanics.
Newton built the first practical reflecting telescope and developed a theory of colour based on
the observation that a prism decomposes white light into the many colours that form the visible
spectrum. He also formulated an empirical law of cooling and studied the speed of sound. In
mathematics, Newton shares the credit with Leibniz for the development of differential and
integral calculus. Newton was also deeply involved in occult studies and interpretations of
religion.
2 Christiaan Huygens (1629 - 1695) was a prominent Dutch mathematician, astronomer,
physicist and horologist. His work included early telescopic studies elucidating the nature of
the rings of Saturn and the discovery of its moon Titan, the invention of the pendulum clock
and other investigations in timekeeping, and studies of both optics and the centrifugal force.
Huygens achieved note for his argument that light consists of waves.
26
Figure 2.1: A perturbation moving right with speed c.
∂ξ dξ ∂u± dξ
= =
∂x du± ∂x du±
∂ξ dξ ∂u ± dξ
= ±v
=
∂t du± ∂t du±
27
Figure 2.2: Perturbation along a string.
2
d2 ξ
∂ ξ d dξ ∂u±
= =
∂x2 du2±
du± du± ∂x
2
d2 ξ
∂ ξ d dξ ∂u±
(±v) = v 2 2
2 =
∂t du± du± ∂t du±
∂2ξ 1 ∂2ξ
⇒ 2
= 2 2 (2.1)
∂x v ∂t
This is called the d’Alembert’s3 equation, or wave equation, in 1 dimension. A
plane wave is defined as a function satisfying the equation (2.1).
Since the wave equation is linear, the sum of two or more waves is still a
wave. The general solution ξ(x, t), however, is not in general a function that
moves left or right, but a linear combination of tho functions ξ+ and ξ− moving
right and left respectively:
ξ(x, t) = ξ+ (x − vt) + ξ− (x + vt) . (2.2)
28
Because of the little bending, the two angles α and α0 are small, so we can write:
and then:
Fx = dm ax ' 0
(2.5)
Fξ = dm aξ ' T (tan α0 − tan α).
We see that the horizontal net force is very small compared to the vertical
one. We can associate tan α to the slope of the string, i.e., to the derivative
with respect to x: ∂ξ/∂x, so we have:
∂ ∂ξ ∂2ξ
Fξ = T dx = T 2 dx . (2.6)
∂x ∂x ∂x
Let µ = M/L be the linear density of the string. We have that the mass
of element AB is µdx; moreover the vertical acceleration is ∂ 2 ξ/∂t2 . By the
second law of motion we can write:
∂2ξ ∂2ξ
µdx = T 2 dx. (2.7)
∂t2 ∂x
Thus,
∂2ξ T ∂2ξ
2
= . (2.8)
∂t µ ∂x2
As we can see this is the d’Alembert wave equation so weqnow know that
the perturbabion moves through the string with velocity v = Tµ .
29
2.3 Sinusoidal waves and the Fourier theorem
A periodical phenomenon is something that repeats itself equally on regular
periods of time. More precisely in mathematics a function f (t) is said to be
periodical of period T if:
Well known examples of periodical functions are the trigonometric functions sin
and cos.
Sinusoidal (or, which is equivalent, cosinusoidal) waves are very important
because of the Fourier theorem.
The Fourier theorem states that every periodical function f (t) of period
T = 2π/ω which is finite, continuos and differentiable can be expressed by the
Fourier series:
∞
X ∞
X
f (t) = a0 + (an cos nωt + bn sin nωt = a0 + cn cos(nωt + ϕn )) (2.12)
n=1 n=1
30
where:
p
cn = a2n + b2n (2.13)
an
tan ϕn = (2.14)
bn
Z T
1
a0 = f (t)dt (2.15)
T 0
T
2 T
Z Z
2
an = f (t) cos(nωt)dtbn = f (t) sin(nωt)dt (2.16)
T 0 T 0
Thus a generic wave can be written as a linear combination of sine waves.
Let us analyze the proprieties of (co)sinusoidal waves:
where ξ˜0 = ξ0 eiδ (the amplitude absorbs the phase). As we stated before, the
real part of the complex wave is the physical wave in this context. Since the
wave equation is linear, we can carry on our calculations using exponentials (the
advantage of the complex notation is that exponentials are much easier to ma-
nipulate than sines and cosines), and then go back to the cosine representation
when we want.
31
• k = 2π/λ is defined as the wave number.
• ω = kv = 2π/T is the angular frequency.
• δ is the phase, and it depends on the choice of the starting time.
2 2
In particular, if ξ0y = ξ0x the wave is said to be circularly polarized.
32
Figure 2.5: A stationary source of waves and a moving source.
33
server. Suppose that the source and the observer move with constant relative
velocity, and that the direction of the velocity lies on the straight line joining
the two objects. We want to analize how the Doppler effect manifests itself.
We make also the semplification that one of the two objects is stationary with
respect to the medium in which the wave propagates (for example, in the case
of sound, the air of the atmosphere). We can immediately note that the phe-
nomenon depends on who is moving and who is stationary with respect to the
medium in which the wave propagates.
Source moving, observer at rest. Let us consider first the case in which
the source moves with velocity vs < v in the direction of the observer. The
length of the wave received λ0 changes. We have that
vs
λ0 = λ − .
f
Thus:
v v vs
= − ;
f0 f f
1
=⇒ f 0 = f · ;
1 − vvs
These conditions hold only when the velocity of propagation v is larger than vs .
In particular if vs v, we can make the approximation
vs
f0 ' f 1 + .
v
Source at rest, observer moving. Let us consider the case in which the
source lies on a fixed point and the observer moves with velocity vo in the
direction of the source. The velocity of the wave relative to the observer is
v + vo , so we get
v + vo v + vo vo
f0 = = =f 1+ .
λ v/f v
We can notice that only in the first approximation the variation of frequency
is the same in both cases. The case in which both the source and the observer
move is more complicated, as the case in which the direction of the movement is
not directed along the line joining them; however, there is nothing conceptually
new.
34
Figure 2.6: Experimental plot of the relative velocity (in km/s) of known
astrophysical objects as a function of the distance from the Earth (in Mpc; 1
pc ' 3.3 ly). The line is a fit to Hubble’s Law.
35
Figure 2.7: Redshift of emission spectrum of stars and galaxies at different
distances. Using this shift we can calculate the relative velocity between these
objects and the Earth.
we can compute the relative velocity of these objects with respect to us, and
then the distance, thanks to the so-called Hubble’s law.
In 1929 Edwin Hubble4 , studying the emission of galaxies, observed from
their Doppler redshift that objects in Universe move away with us with velocity
v = H0 d , (2.23)
where d is the distance between the objects, and H0 is a parameter called the
Hubble constant (whose value is known today to be about 24km/s/M ly). The
above relation is called Hubble’s law.
To give an isea of what H − 0 means, the speed of revolution of the Earth
around the Sun is about 30 km/s; Andromeda, the large galaxy closest to the
Milky Way, has a distance d of about 2.5 Mly. However, the Hubble’s law is
just statistical and working for large distances, where gravitational attraction
becomes negligible: Andromeda is indeed approaching us.
Dimensionally we note that H0 is a frequency: H0 ' (14 × 109 years)−1 .
A simple interpretation of this law is that, if the Universe has always been
expanding at a constant rate, about 14 · 109 years ago its volume was zero. This
result is consistent with present estimates of the age of the Universe within the
so-called big bang theory.
The redshift
λ0
z= −1
λ
is also used as a metric of distance of objects.
4 Edwin Hubble (1889 - 1953) was an American astronomer who played a crucial role in
establishing the field of extragalactic astronomy and is generally regarded as one of the most
important observational cosmologists of the 20th century.
36
Figure 2.8: The Cherenkov cone.
To find the value of cos θ let us consider two positions of the source S1 and
S2 , and the corresponding points P and Q on the wave front: the wave emitted
in S1 has, in P , the same phase of the one emitted in S2 in Q. For the same
reason also the points S10 and S2 have the same phase. The time that the source
spends to go from S1 to S2 is equal to the time that the wave spends to go from
S1 to S10 . If we call a the distance S1 S2 we have S1 S10 = a cos θ. Thus we get:
a a cos θ
=
vs v
v
=⇒ cos θ = .
vs
As an example, when cosmic rays interact with the atmosphere they generate
showers of particles. The charged particles radiate light, and some of them
are faster than light in the atmosphere, thus generating cones of collimated
Cherenkov light. This light is detected by special-purpose telescopes.
5 Pavel Alekseyevich Cherenkov (1904-1990) was a Soviet physicist who shared the Nobel
Prize in physics in 1958 with Ilya Frank and Igor Tamm for the discovery of Cherenkov
radiation, made in 1934. The discovery was made during Cherenkov’s thesis, directed by the
academician Nikolai Vavilov; when the Nobel prize was assigned, however, Vavilov was dead
since 15 years.
37
2.8 Composition of waves
Since the wave equation is linear, when we want to sum two waves we can
just sum the functions representing them. We should remember that energy is
proportional to the square of the amplitude: this can create effects of positive
and negative interference that we shall discuss later.
38
istics but opposite velocity. If we sum the two waves we obtain a wave with
constant speed zero, also called stationary wave.
Let us analyze mathematically what happens. Since v = ω/k, and ω remains
the same, the change of velocity from +v to −v corresponds to a change of wave
number from +k to −k. So the two wave equations are:
ξ+ (x, t) = ξ0 sin (kx − ωt) ;
ξ− (x, t) = ξ0 sin (−kx − ωt) .
Using the appropriate trigonometric formulae, the sum of the waves is:
ξ(x, t) = ξ+ (x, t) + ξ− (x, t) =
= ξ0 sin (kx − ωt) + ξ0 sin (−kx − ωt) =
= 2ξ0 sin (−ωt) cos (kx) =
= −2ξ0 sin (ωt) cos (kx) .
The oscillation is maximum in the points x such that cos(kx) = 1, while is
minimum in the points that satisfy the condition cos(kx) = 0. These points are
called respectively antinodes and nodes.
Observe that only waves that have nodes on the extremities of the rope will
survive: otherwise the amplitude of the sum of the waves in an extreme is not
null, and so also the energy (dispersed) is not null, but this contradicts the
conservation of energy, because we are supposing the reflected wave to have the
same amplitude ξ0 .
Imagine you are perturbing a violin string: the extremities P and Q are
fixed and you generate different waves displacing the string in different points.
The wave you have generated can be expressed as the sum of sinusoidal waves,
from the Fourier’s theorem. All these waves have a same property: the points
P and Q are nodes for them. After a transient, only the waves for which the
points P and Q are nodes survive, otherwise the wave, when reflected, generates
a destructive interference. So only discrete values of λ are permitted, those for
which the boundaries are nodal points. These are called the harmonics:
• the fundamental frequency with frequency f1 is the wave with maximum
wavelenght: it is such that λ2 = L where λ is the wavelenght and L is the
string lenght. Observe that λ = 2L;
2λ
• the second harmonic f2 is a wave with wave lenght such that 2 = L, that
is λ = L;
3λ
• the third harmonic f3 is such that 2 = L, that is λ = 32 L;
4λ
• the fourth harmonic f4 has a wave lenght such that 2 = L, i.e. λ = 21 L;
• and so on . . .
The wave generated perturbing the violin string is, after a transient, the sum
of these waves, as we can see in Figure 2.12. The fundamental frequency deter-
mines the pitch of the note, and together with the higher harmonics determine
39
Figure 2.11: The first four harmonics.
40
Figure 2.12: A string is plucked in a certain point. This creates a wave that is
sum of three waves: the foundamental frequency, the second and the third
harmonics. Their sum produce a determined sound.
the timbre of the sound. We can obtain sounds with same pitch, but different
timbre, just plucking the string in different places.
2.8.2 Beats
We will now analyze another interesting example of wave composition. Let us
consider two sinusoidal waves that propagate in the same direction, with the
same amplitude ξ0 and velocity v, but slightly different frequencies ωi .
The waves are defined by the following equations:
41
Figure 2.13: Graphic representation of beats.
we obtain
where Ξ(k) is a function that takes a large value in a region of area ∆k̄ around a
certain point k̄, and goes to zero elsewhere. For example f could be a Gaussian
function with very low variance so that the funcion has a great peak near k̄. An
example of wave packet is represented in Figure 2.14.
We want now to analize the speed of propagation of a wave packet. To do
this we observe that beats are a simple wave packet, made by two waves. So we
begin studying this case, that is simpler than the general one. We know that
42
Figure 2.14: A wave packet.
Figure 2.15: In the example of beats the envelope wave (green) is given by the
cosinus with longer wavelenght.
the speed of a wave whose equation is f (x, t) = ξ0 cos(kx − ωt) is given by:
ω
v= . (2.27)
k
In the previous section we obtained that beats have an equation that is the
product of two cosinusoidal functions. The speed of the envelope is given by the
speed of the factor with longer wavelenght, i.e., lower wavenumber. If we return
to (2.26), we have to compare the numbers hki and ∆k 2 , to understand which
term has the lower wavenumber. It is easy to prove that ∆k 2 ≤ hki, so we have
∆k ∆ω
to consider the factor cos 2 x − 2 t . Using the (2.27) we find the velocity
∆ω
venvelope = .
∆k
If we want to extend the result to general wave packets, we have to take the
limit for ∆k that goes to zero. So the velocity of the envelope, also called group
velocity, is given by:
d(ω(k))
vg = . (2.28)
dk
The phase velocity is the speed of propagation of a phase, for example of the
point P represented in Figure 2.15. This velocity, in the case of beats, is equal
to the propagation velocity of the factor with shorter wavelenght and higher
43
wavenumber, i.e., we have to look at the term cos(hkix − hωit). Remembering
that v = ω1 /k1 = ω2 /k2 the phase velocity of beats is:
hωi ω1 + ω2 vk1 + vk2 ω1 ω2
v= = = = = = v.
hki k1 + k2 k1 + k2 k1 k2
Note that the phase velocity is equal to the speed of the two waves that generate
beats. In general, in a dispersive medium (where v is a function of ω) the phase
and group velocities can be different.
44
where S(r) = 4πr2 is the surface of the wavefront.
Therefore the equation of a spherical wave in a nondispersive medium is
ξ0
ξ(r, t) = sin(kr − ωt).
r
45
Chapter 3
together with the equation describing the motion of a particle of electrical charge
q in an electromagnetic field
F~ = q(E~ + ~v × B)
~ (3.5)
(Lorentz2 force), provide a complete description of electromagnetic field and of
its dynamical effects.
We want to write Maxwell’s equations in a local form, and to transform the
integro-differential equations above into purely differential equations.
1 James Clerk Maxwell (1831 - 1879) was a Scottish physicist. His most prominent achieve-
46
3.1 Charge density and current density
We first write in terms of local variables the electric charge and the electric
current.
The charge density ρ(t, ~r) is defined as the charge per unit volume in a point
~r at a time t: Z
q= ρ(t, ~r) dV . (3.6)
V
The current density ~(t, ~r) is defined as the intensity of electrical current per
unit surface: Z
~.
I = ~(t, ~r) · dS (3.7)
S
and thus
~ · E~ = ρ .
∇ (3.8)
0
In the same way we get from Equation (3.3), by applying Gauss’ theorem,
I
B~ · dS
~ = 0 =⇒ ∇
~ ·B
~ = 0. (3.9)
Equations
I Z
d
E~ · d~l = − B~ · dS
~
C(S) dt S
I Z
~ · d~l = d
B µ0 I + 0 µ0 E~ · dS
~
C(S) dt S
47
and thus
~
∂B
~ × E~
∇ = −
∂t
~ ×B
~ ∂ E~
∇ = 0 µ0 + µ0~ .
∂t
These are called respectively the law of Faraday3 -Lenz4 law and the Ampère5 -
Maxwell law.
Thus Maxwell’s equations can be written in differential form as:
~ · E~ = ρ
∇
0
~
~ × E~ = − ∂ B
∇
∂t
∇~ ·B ~ =0
~
~ ×B
∇ ~ = 0 µ0 ∂ E + µ0~
∂t
These equations allow calculating the electric and magnetic fields from the
charge distribution ρ(t, ~r) and the current density ~(t, ~r).
~ ·~ + ∂ρ = 0 .
∇ (3.10)
∂t
3 Michael Faraday (1791 - 1867) was an English scientist who contributed to the fields of
48
The equation above is a continuity equation for charge. If there is a net
electric current is flowing out of a region, then the charge in that region must
be decreasing by the same amount. Charge is conserved.
~ =∇
B ~ ×A
~.
E~ = −∇φ
~
φ0 = φ + C .
The new potential φ0 gives the same electric fields, since the gradient of a con-
stant is zero; φ0 and φ represent the same physics.
Similarly, we can have different vector potentials A ~ which give the same
~ ~
magnetic fields. Again, because B is obtained from A by differentiation, adding
a constant to A ~ does not change anything physical. But we can add to A ~ any
field which is the gradient of some scalar field, without changing the physics:
~ =∇
B ~ ×A
~ =⇒ ∇
~ × (A
~ + ∇ψ)
~ ~ ×A
=∇ ~+∇
~ × ∇ψ
~ =∇
~ ×A
~=B
~.
~ by arbitrarily
It is usually convenient to take some of the freedom out of A
placing some other condition on it (in much the same way that we found it
convenient - often - to choose to make the potential zero at large distances).
We can, for instance, restrict A ~ by choosing arbitrarily what the divergence
~
of A must be. We can always do that without affecting B. ~ This is because
~ ~ 0 ~
although A and A have the same curi, and give the same B, they do not need
to have the same divergence. By a suitable choice of ψ we can make ∇~ ×A ~ 0 any
well-behaved function we wish.
What should we choose for ∇ ~ · A?
~ The choice should be made to get the
greatest mathematical convenience and will depend on the problem we are doing.
For magnetostatics, we make the simple choice
~ ·A
∇ ~=0 (3.11)
49
(later, when we take up electrodynamics, we shall make a different choice). Our
complete definition of A is then, for the moment,
~ ×A
∇ ~=B
~ and ∇
˜ · Ã = 0 .
and the Biot-Savart6 law, stating that the elementary magnetic field generated
by a current I over an element of conductor d~l at a radius ~r is
µ d~l × ~u
~ = 0 r
dB I . (3.13)
4π r2
from the Maxwell equations? It turns out that this is indeed possible.
We do the demonstration for the Coulomb’s law. The differential form of the
Maxwell’s law is equivalent to its integral form (3.1). Due to symmetry reasons,
the field must be directed radially. Then if we choose as a surface a sphere of
radius r, equation (3.1) becomes
~ 4πr2 = q
|E| ,
0
which demonstrates the statement.
6 Jean-Baptiste Biot (1774 - 1862) was a French physicist, astronomer, and mathemati-
cian who established the reality of meteorites, made an early balloon flight, and studied the
polarization of light. Félix Savart (1791 - 1841), professor at Collège de France, was the co-
originator of the Biot-Savart Law, along with Biot. Together, they worked on the theory of
magnetism and electrical currents. Their law was developed about 1820.
50
Chapter 4
~ ·B
∇ ~ =0
• Ampère-Maxwell’s law:
~
~ = ε0 µ0 ∂ E + µ0~ .
~ ×B
∇
∂t
Now we consider a region without charges and we are going to analyze the
situation. We can write the Maxwell’s equations in vacuo as:
~ · E~ = 0
∇ (4.1)
~
~ × E~ = − ∂ B
∇ (4.2)
∂t
51
~ ·B
∇ ~ =0 (4.3)
~ ×B
∇ ~ = ε0 µ0 ∂E (4.4)
∂t
A trivial solution would be that the electromagnetic field is everywhere zero.
Let us examine the characteristics of nontrivial solutions.
We suppose that there would be field generators outside the no-charges re-
gion. Consequently the region inherits information about the charges in the
Universe. Equation 4.2 can be written as:
~ × E~ = − ∂ ∇
~ × ∇
∇ ~ ×B ~
∂t
∂ 2 E~
= −ε0 µ0 2 thanks to equation 4.4
∂t
On the other hand
∂ 2 E~
~ × ∇
~ × E~ = ∇
~
~ · E~ −∇2 E~ = −∇2 E~
−µ0 ε0 = ∇ ∇
∂t2 | {z }
=0, eq. 4.1
∂2B~ ∂ ~
−µ0 ε0 2
= ε0 µ0 ∇ × E~ = ∇
~ × ∇~ ×B
~ =∇~ ∇ ~ ·B
~ −∇2 B
~ = −∇2 B
~
∂t ∂t | {z } | {z }
eq.4.2 =0,eq. 4.3
52
we can write the velocity of this wave:
1
1 4πε0 9 · 109 m2
v2 = = µ0 ' ' 9 · 1016 .
ε0 µ0 4π 10−7 s2
Then: r
1
v= ' 3 · 108 m/s (4.7)
ε0 µ0
that is consistent with the value of the speed of light. It becomes then natural
to assume that light is an electromagnetic wave.
Hertz1 presented the decisive experimental confirmation that light is a Max-
well’s wave in 1888, by generating light through an oscillating electromagnetic
field. He wrote: “The connection between light and electricity is now estab-
lished ... In every flame, in every luminous particle, we see an electrical process
... Thus, the domain of electricity extends over the whole of nature. It even
affects ourselves intimately: we perceive that we possess ... an electrical organ -
the eye.” By 1900, then, three great branches of physics, electricity, magnetism,
and optics, had merged into a single unified theory (and it was soon apparent
that visible light represents only a tiny window in the vast spectrum of electro-
magnetic radiation, from radio though microwaves, infrared and ultraviolet, to
X-rays and gamma rays.)
The Ampère-Maxwell equation (4.4) becomes, if we do this assumption,
~ ~
~ = ε0 µ0 ∂ E = 1 ∂ E .
~ ×B
∇
∂t 2
c ∂t
Thus the speed of light enters in the fondamental laws of nature.
This fact has an important consequence: in classical mechanics the Maxwell’s
equations are not the same for different reference frames in relative motion with
constant velocity. Indeed, in classical physics, all the laws depend on acceler-
ation, which is invariant between the inertial reference frames (the coordinates
are transformed from one reference frame to another by means of the Galilei2
transformations); velocities cannot enter in a fundamental law valid in all iner-
tial frames, since they are not an invariant quantity.
Since the Maxwell’s equations violate the Galilean relativity, if we assume
that they are correct, there could be two possibilities (not mutually exclusive):
• Relativity is not a legitimate assumption: there is a preferred reference
frame, in which electromagnetism can be described (this has been called
1 Heinrich Hertz (1857 - 1894) was a German physicist who clarified and expanded James
philosopher who played a major role in the scientific revolution. His achievements include
improvements to the telescope and consequent astronomical observations and support for
Copernicanism. His contributions to observational astronomy include the discovery of the
phases of Venus, of the four largest satellites of Jupiter (named the Galilean moons in his
honour), and the observation and analysis of sunspots. Galilei also worked in military science
and technology.
53
the aether). For example, this would be the case if electromagnetic fields
would be the perturbation of a static medium.
• The relativistic transformations of Galilei do not work.
We shall see that the answer to this question brought to us one of the most
fascinating theories of the XX century - Einstein’s3 special relativity.
∂ E~ ∂ E~ ~
∂B ~
∂B
= = = =0
∂x ∂y ∂x ∂y
(the waves must be plane waves). Thus the Maxwell’s equations become:
→
−
∂ Ez
=0 (4.8)
∂z
∂ E~ ~
∂B
ûz × =− . (4.9)
∂z ∂t
Indeed
→
− −
→! →
− −
→! −
→ −
→!
~ × E~ = ûx ∂ Ez ∂ Ey ∂ Ez ∂ Ex ∂ Ey ∂ Ex
∇ − − ûy − + ûz − =
∂y ∂z ∂x ∂z ∂x ∂y
∂ E~ ~
∂B
= ûz × =− .
∂z ∂t
Analogously for the magnetic field:
−
→
∂ Bz
=0 (4.10)
∂z
3 Albert Einstein (1879 - 1955) was a German-born physicist who deeply changed the rep-
resentation of the Universe by the human species. While best known for his mass-energy
equivalence formula E = mc2 (published in 1905), he received the 1921 Nobel Prize in Physics
for his discovery of the law of the photoelectric effect (also in 1905), which was pivotal in es-
tablishing quantum theory within physics. Near the beginning of his career, Einstein thought
that Newtonian mechanics could not reconcile the laws dynamics with the laws of the electro-
magnetic field. This led to the development of his special theory of relativity (again in 1905).
He realized, however, that the principle of relativity could also be extended to gravitational
fields, and with his subsequent theory of gravitation in 1916, he published a paper on the gen-
eral theory of relativity. He was visiting the United States when Adolf Hitler came to power
in 1933, and did not go back to Germany, where he was a professor in Berlin. He settled in
the U.S., becoming a citizen in 1940. On the eve of World War II, he helped the set up of
the Manhattan Project, and ultimately the construction of the atomic bomb. Later, however,
he highlighted the danger of nuclear weapons. Einstein was affiliated with the Institute for
Advanced Study in Princeton until his death.
54
~
∂B 1 ∂ E~
ûz × = 2 . (4.11)
∂z c ∂t
We observe that
~
! →
− →
−
∂B 1 ∂ Ez ∂ Ez
0 = ûz · ûz × = 2 =⇒ =0
∂z c ∂t ∂t
! −
→ −
→
∂ E~ ∂ Bz ∂ Bz
0 = ûz · ûz × =− =⇒ =0
∂z ∂t ∂t
and thus we can conclude that Ez and Bz are constants, and in particular they
do not depend on the variables x, y, z, t. If we postulate that the Universe has
finite energy, we conclude that Ez = Bz = 0. Recall that the energy density
associated with the presence of electric and magnetic field are respectively
ε0 E 2 B2
uE~ = uB~ = . (4.12)
2 2µ0
Unless the Universe has infinite energy, Ez = Bz = 0: the electromagnetic wave
must be transverse.
We rewrite all the equations obtained to understand the relations between
the electric field and the magnetic field in an electromagnetic wave.
Ez = 0 (4.13)
Bz = 0 (4.14)
∂Ex ∂By
=− (4.15)
∂z ∂t
∂Ey ∂Bx
= (4.16)
∂z ∂t
∂Ex ∂By
= c2 (4.17)
∂t ∂z
∂Ey ∂Bx
= −c2 (4.18)
∂t ∂z
If we set υ = z − ct we have that ∂υ ∂υ
∂z = 1 and ∂t = −c. We can obtain from
equation 4.15
∂By ∂Ex ∂Ex ∂υ ∂Ex
=− =− =−
∂t ∂z ∂υ ∂z ∂υ
Z Z Z
∂By ∂Ex 1 ∂Ex
⇒ By = dt = − dt = dυ
∂t ∂υ c ∂υ
Ex
⇒ By = + constant
c
The constant must be equal to zero, not to have infinite energy; in the end we
have then
Ex
By = . (4.19)
c
55
Analogously from Equation (4.16)
Ey
Bx = − . (4.20)
c
In conclusion we have obtained that
~
|E|
~ =
|B|
c
and
~ · E~ = Ey Ex
B − Ex + Ey + 0 = 0.
c c
In other words, the electric wave is perpendicular and proportional to the mag-
netic wave.
We also observe from (4.19) and (4.20) that
E~ × B
~ = EB u~z : (4.21)
u = 2uE~ = ε0 E 2 .
56
4.2.2 The Poynting vector
Let u be the energy density per volume unit, the energy carried by a wave,
perpendicularly to its direction, per time and surface unit is
dU dU dx
I= = = uv.
dSdt dSdx dt
We call I the intensity and it is measured in W/m2 .
So far we assumed that the origin of the energy flux is coincident to the
wave source, ignoring possible dissipative effects. The energy carried by an
electromagnetic wave in vacuum per time and surface unit is
1
S = uc = ε0 E 2 c = ε0 EBc2 = EB
µ0
1 ~~
S~ = ~
E ×B
µ0
has the same direction of the wave and its magnitude is equal to the power per
surface unit. This vector is called the Poynting’s vector 4 .
A sine wave of the form E = E0 cos(kz − ωt), B = B0 cos(kz − ωt) has an
average power of EB/2µ0 .
The electromagnetic wave also carries momentum. Based on considerations
supported by Einstein’s relativity theory (see later), the momentum carried per
~
surface unit is related to energy as S/c.
57
Figure 4.1: Sketch of the apparatus used to study the photoelectric effect.
58
photoelectric current I that is measured with meter A. We adjust the poten-
tial difference V by moving the sliding contact so that the collector is slightly
negative with respect to the target. This potential difference acts to slow down
the ejected electrons. We then vary V until it reaches a certain value, called the
stopping potential Vstop , at which point the reading of meter A has just dropped
to zero. When V = Vstop , the most energetic ejected electrons are turned back
just before reaching the collector.
Then Kmax , the kinetic energy of
these most energetic electrons, is
Kmax
Kmax = eVstop
59
additional energy (hν − Φ) that the electron acquires from the photon appears
as kinetic energy K of the electron. In the most favourable circumstance, the
electron can escape through the surface without losing any of this kinetic energy
in the process; it then appears outside the target with the maximum possible
kinetic energy Kmax .
Therefore we can conclude that the a light wave of frequency ν is made of
quanta of energy
h
hν = ω = ~ω
2π
where h is the Planck’s constant, which has value
h ' 6.63 × 10−34 J s = 4.14 × 10−15 eV s .
in the standard SI the unit of electrical current, the ampere A, is used instead of the coulomb;
the two definitions are however conceptually equivalent.
60
Note the unit of mass, in which the relation E = mc2 establishing equivalence
of mass and energy (we shall discuss this relationship later) is used implicitly:
what one is doing here is to use 1 eV ' 1.60 × 10−19 J as the new fundamental
unit of energy.
With these new units, the mass of a proton is about 0.938 GeV/c2 , and the
mass of the electron is about 0.511 MeV/c2 . The fundamental energy level of
Hydrogen is about −13.6 eV.
In addition to this, nature is providing us with two constants which are
particularly appropriate for the world of fundamental physics: the speed of light
c ' 3.00×108 m/s = 3.00×1023 fm/s, and the Planck’s constant ~ ' 1.05×10−34
J s ' 6.58 × 10−16 eV s. It seems then natural to express speeds in terms of
c, and angular moments in terms of ~. When we do this we switch to the
so-called Natural Units (NU). The minimal set of natural units (not including
electromagnetism) could then be
Speed 1c 3.00 × 108 m/s
Angular momentum 1~ 1.05 × 1034 J s
Energy 1 eV 1.60 × 10−19 J
In such a system, ~ = c = 1.
After these conventions, just one unit can be used to describe the mechanical
Universe: we choose energy, and thus we can express al mechanical quantities
in terms of eV and of its multiples. It is immediate to express momenta and
masses directly in terms of energy; related to lengths and energies, we can use
the fact that
(the first relation can also be written as ~c ' 0.197 GeV fm. By choosing natural
units, all factors of ~ and c may be omitted from equations, which leads to
considerable simplifications. For example, the relativistic energy relation
E 2 = p 2 c2 + m 2 c4
becomes
E 2 = p2 + m2 .
To express 1 m and 1 s in NU, we can just write
1m 12 −1
1m = ~c ' 5.10 × 10 MeV
1s 21 −1
1s = ~ ' 1.52 × 10 .MeV
Both length and time are thus, in natural units, expressed as inverse of energy.
The first relation can also be written as 1fm ' 5.10GeV−1 : note that when you
when a quantity expressed in MeV−1 , in order to express it in GeV−1 , you must
multiply (and not divide) by a factor of 1000.
61
mks NU
Quantity p q r n
Action (~) 1 2 -1 0
Velocity (c) 0 1 -1 0
Mass 1 0 0 0
Length 0 1 0 -1
Time 0 0 1 -1
Momentum 1 1 -1 1
Energy 1 2 -2 1
The NU and SI dimensions are listed for some important quantities in Table
4.1.
Finally, let us discuss how to treat electromagnetism. To do so, we must
introduce a new unit, charge for example. We can redefine the unit charge by
observing that
e2
4π0
has the dimension of [J m], and thus is a pure number in NU. By dividing by
~c one has:
e2 1
' .
4π0 ~c 137
Imposing that the electric permeability of vacuum 0 = 1 (thus automatically
µ0 = 1 for the magnetic permeability of vacuum, since from Maxwell’s equations
62
0 µ0 = 1/c2 ) we obtain the new definition of charge, and with such a definition:
e2 1
α= ' .
4π 137
This is called the Lorentz-Heaviside convention. Elementary charge in NU be-
comes then a pure number:
e ' 0.303 .
63
Chapter 5
Geometrical optics
In cases where the wavelength is small compared to other length scales in a phys-
ical system, light waves can be modeled by light rays, moving on straight-line
trajectories representing the direction of the propagation. This is the domain
of the so-called geometrical optics.
mental understanding of electrical circuits, spectroscopy, and the emission of black-body (he
coined the term “black body”) radiation by heated objects.
64
Figure 5.1: Huygens’ principle.
The Huygens’ principle states that all points on a given wave front can be
taken as point sources for the production of spherical secondary waves, called
wavelets, which propagate in the forward direction with the speed characteristic
of waves in that medium.
The corrections introduced by Kirchhoff were that the amplitude varies ac-
cording to a decreasing function of the angle θ with respect to the direction of
propagation (f (θ) ' (1 + cos θ)/2), and that the phase of the emitter is antic-
ipated by π/2 with respect to the phase of the wavefront. In many problems,
however, these two points can be neglected.
His name has been attached to the laws of reflection and of refraction of light.
65
To demonstrate it, we can look at the Figure 5.3: when the wave front
touches the surface of reflection, every point behaves like a source of a new
wave, that propagates form the point from which the principal wave came. The
result of these propagations is a wave that appears like the original wave, just
reflected by an angle equal to the angle θ1 .
A second application of the rule that Huygens postulated is the demonstra-
tion of the phenomenon of refraction.
Theorem 2. If a wave passes from a medium with rifraction index n1 to one
with index n2 , then the direction of the light obeys the following equation
where θ1 and θ2 are the angles formed by the direction of the wave with the
normal of the surface, before and after the change of medium of propagation
respectively (vi = c/ni is the speed of light in the two media).
Let us look at Figure 5.3. If we call T the time at which the wavefront
begins to cross the second medium, and T + ∆t the time at which a second ray
hits the second medium, then in a time of ∆t the wave formed by the wavelets
has moved inside the second medium of a distance of v2 ∆t. We obtain that
AC sin θ1 = v1 ∆t, and AC sin θ2 = v2 ∆t. From that we obtain
sin θ1 v1 n2
= = . (5.2)
sin θ2 v2 n1
Note that if n1 > n2 , an angle of incidency θc exists, called the critical angle,
for which the wave does not propagate in the second medium. This is because
the equation (5.2) becomes:
n1
θ2 = arcsin sin θ1 .
n2
This principle is used in optical fibres: the waves in a tiny tube cannot exit from
the fibres because of the elevate refraction index.
We also observe that in the transmission between two media, frequency of
waves does not change, consistent with Huygens’ principle: in fact the wavelets
have all the same frequency as the frequency of the principal wave.
66
Figure 5.2: The phenomenon of reflection.
67
Figure 5.3: The phenomenon of refraction.
ematician who is given credit for early developments that led to infinitesimal calculus. He
made notable contributions to analytic geometry, probability, and optics, and in particular to
number theory.
68
Figure 5.4: The Fermat’s Principle.
our path in which the medium changes. The total time is given by
√ p
a 2 + x2 b2 + (c − x)2
t= + ;
v1 v2
dt x c−x
= √ − p =
dx 2
v1 a + x 2 v2 b + (c − x)2
2
sin θ1 sin θ2
= − .
v1 v2
Imposing the derivative of the function to equal 0, we get
sin θ1 v1 n2
= = .
sin θ2 v2 n1
that is the law of refraction.
69
Chapter 6
Light waves interfere with each other much like mechanical waves do: inter-
ference associated with light waves arises when the electromagnetic fields that
constitute the individual waves combine.
Since light frequency is very high compared to the sensitivity of the human
eye (and of most instruments: the typical frequency of 1015 Hz is very difficult
to reach), for interference between two sources of light to be observed, there are
two conditions which must be met: the sources must be coherent (they must
maintain a constant phase with respect to each other), and the waves must have
identical wavelengths.
notable scientific contributions to the fields of vision, light, solid mechanics, energy, physiology,
language, musical harmony, and in the decipherment of Egyptian hieroglyphs.
70
Figure 6.1: Arrangement of Young’s interference experiment.
the screen. Dark regions - called dark bands, dark fringes, or (loosely speaking)
minima - result from fully destructive interference and are visible between ad-
jacent pairs of bright fringes (maxima and minima more properly refer to the
center of a band.) The pattern of bright and dark fringes on the screen is called
an interference pattern. Figure (6.2) is a photograph of part of the interference
pattern that would be seen by an observer standing to the left of screen C at a
distance L d, where d is the distance between S1 and S2 .
Where are the dark fringes and the bright fringes located? Let us examine
the difference in optical path between the two rays in Figure (6.3).
When one wave travels an integer number of wavelengths farther than the
other, the waves arrive in phase, and a bright fringe occurs. This condition is
verified for
δ = r2 − r1 ' d sin θ = mλ
(this assumes the paths are parallel; they are not exactly parallel, but the above
is a very good approximation since L d). The absolute value of the integer
m is called the order of the maximum.
When destructive interference occurs, a dark fringe is observed. This needs
a path difference of an odd half wavelength
1
δ = r2 − r1 ' d sin θ = m + λ.
2
71
Figure 6.2: Interference pattern.
72
Figure 6.4: N-slits interference.
(with m = 0, ±1, ±2, ...). Young’s setup is thus effective to measure the wave-
length of light, which is amplified by a factor L/d.
What happens in between the maxima and the minima? Let us solve the
general problem of the interference of N equally spaced sources in phase, each
of amplitude A (Figure (6.4). Let A(θ) be the common amplitude of each wave
(which can be a function of θ due to the attenuation).
The total field Etot as a function of θ is
N
X N
X N
X
Etot = En = A(θ)ei(krn −ωt) = A(θ)ei(kr1 −ωt) eik(n−1)d sin θ . (6.3)
n=1 n=1 n=1
With ζ = eikd sin θ , the sum is a sum of the geometrical series 1 + ζ + ζ 2 + ... +
73
ζ N −1 = (ζ N − 1)/(ζ − 1). Thus
eikN d sin θ − 1
Etot = A(θ)ei(kr1 −ωt)
eikd sin θ − 1
eik(N/2)d sin θ eik(N/2)d sin θ − e−ik(N/2)d sin θ
= A(θ)ei(kr1 −ωt) ik(1/2)d sin θ ik(1/2)d sin θ
e e − e−ik(1/2)d sin θ
sin (N/2)kd sin θ
= A(θ)ei(kr1 −ωt) eik((N −1)/2)d sin θ .
sin (1/2)kd sin θ
with
2πd sin θ
α ≡ kd sin θ = .
λ
The amplitude at θ = 0 is
and thus
Atot (θ) A(θ) sin (N α/2)
= .
Atot (0) A(0) N sin (α/2)
When we go to intensities,
2
Itot (θ) A(θ) sin (N α/2)
= .
Itot (0) A(0) N sin (α/2)
Even for large angles, the effect of A(θ) is to simply act as an envelope
function of the oscillating sine functions. We can always bring A(θ) back in if
we want to, but the more interesting behavior of Atot (θ)) is the oscillatory part.
We are generally concerned with the locations of the maxima and minima of
the oscillations and not with the actual value of the amplitude. We can pose,
for the moment, Atot (θ) ' Atot (0).
What does the Itot (θ)/Itot (0) ratio look like as a function of θ? The plot for
N = 4 is shown in Figure (6.5). If we are actually talking about small angles,
then we have α = kd sin θ ' kdz/D.
74
Figure 6.5: 4-slits interference.
6.2 Diffraction
Any wave passing through an opening experiences diffraction. Diffraction means
that the wave spreads out on the other side of the opening rather than the
opening casting a sharp shadow.
Diffraction is most noticeable when the opening is about the same size as
the wavelength of the wave. If light passes through a narrow slit, it produces a
characteristic pattern of light and dark areas called a diffraction pattern, which
can be explained by the interference of light traveling different optical paths.
Light passing a sharp edge also exhibits a diffraction pattern. Huygens’
Principle describes this spreading out, and a Huygens’ construction can be used
to quantify the diffraction phenomenon. For example, Figure (6.6) shows co-
herent light incident on an opening, which has dimensions comparable to the
wavelength of the light. Rather than casting a sharp shadow, light spreads out
on the other side of the opening. We can describe this spreading out by using a
Huygens’ construction and assuming that spherical wavelets are emitted at sev-
eral points inside the opening. The resulting light waves on the right side of the
opening undergo interference and produce a characteristic diffraction pattern.
Light waves can also go around the edges of barriers. In this case, the light
far from the edge of the barrier continues to travel like the light waves shown in
Figure (6.6). The light near the edge of the barrier seems to bend around the
barrier and is described by the sources of wavelets near the edge.
75
Figure 6.6: Diffraction through a single slit.
76
limit of the N −slit result.
We shall perform this calculation by an integral over all the phases from
the possible paths from different parts of the wide slit. Let the slit run from
y = −a/2 to y = a/2, and let B(θ)dy be the amplitude that would be present
at a location θ on the screen if only an infinitesimal slit of width dy was open.
So B(θ) is the amplitude (on the screen) per unit length (in the slit): B(θ)dy is
the analog of the A(θ) in the case of interference. If we measure the pathlengths
relative to the midpoint of the slit, then the path that starts at position y is
shorter by y sin θ. It therefore has a relative phase of e−ky sin θ .
Integrating over all the paths that emerge from the different values of y
(through slits of width dy) gives the total wave at position θ on the screen as
(up to an overall phase from the y = 0 point, and ignoring the temporal part of
the phase)
Z +a/2
Etot (θ) = dy B(θ) e−iky sin θ .
−a/2
This is the continuous version of√the discrete sum in the case of the multi-slit
interference. B(θ) falls off like 1 = r; however, we shall assume that θ is small,
which means that we can assume cosθ ' 1.
Z +a/2
B(0) −ik(a/2) sin θ
Etot (θ) ' B(0) dy e−iky sin θ = e − eik(a/2) sin θ
−a/2 −ik sin θ
−2i sin ka sin θ
2
= B(0)
−ik sin θ
sin 1 ka sin θ
= B(0)a 1 2 .
2 ka sin θ
There is no phase here, so this itself is the amplitude Atot (θ). Taking the
usual limit at θ = 0, we obtain Atot (0) = B(0)a. Therefore, Atot (θ)/Atot (0) =
sin(β/2)/(β/2), where β = ka sin θ. Since the intensity is proportional to the
square of the amplitude, we arrive at the equation
2
Itot (θ) sin(β/2)
=
Itot (0) β/2
with β = ka sin θ = 2πa sin θ/λ.
From Figure (6.8), we see that most of the intensity of the diffraction pattern
is contained within the main bump where β < 2π. Numerically, you can shown
that the fraction of the total area that lies under the main bump is about 90%.
So it makes sense to say that the angular width of the pattern is given by
λ
sin θ ' .
a
77
Figure 6.8: The function sin2 (β/2)/(β/2)2 .
The intensity of the pattern on the screen is the result of the combined effects
of interference and diffraction.
78
• To an observer standing at the center of the platform when the subway car goes by, the car
has a length L that can be calculated by measuring the time that passes between the time
when the front end of the car reaches the center of the platform and the time when the rear
end of the car reaches the center of the platform: L=Vt
• The interval t is a proper time interval for the observer standing on the platform, because the
two events whose times are being measured both happen at the center of the platform.
Therefore the time intervals are related by
L'
t'= t L =
• Let’s look at two inertial reference systems K and K' with coordinate axes XYZ and
X’ Y’ Z’, where the system K' moves relative to K along the X(X') axis.
• Suppose signals start out from some point A on the X' axis in two opposite
directions. Since the velocity of propagation of a signal in the K' system, as in all
inertial systems, is equal (for both directions) to c, signals will reach points B and C,
equidistant from A, at the same time (in the K' system)
• The same two events (arrival of the signal at B and C) can by no means be
simultaneous for an observer in the K system. In fact, the velocity of a signal
relative to the K system has, according to the principle of relativity, the same value
c, and since the point B moves (relative to the K system) toward the source of its
signal, while the point C moves in the direction away from the signal (sent from A
to C), in the K system the signal will reach point B earlier than point C.
• We shall frequently use the concept of event. An event is described by the place
where it occurred and the time when it occurred. Thus an event occurring in a
certain material particle is defined by the three coordinates of that particle
• We designate the time in the systems K and K' by t and t'
Intervals
Transformation of intervals
As already shown, if ds = 0 in one inertial system, then ds' = 0 in any other system.
On the other hand, ds and ds' are infinitesimals of the same order. From these two
conditions it follows that ds2 and ds'2 must be proportional to each other:
Timelike events
Spacelike events
• In the same way, all events in the region bOd are in the
Future and past
absolute past relative to O; i.e. events in this region
occur before the event O in all systems of reference. (2)
• Next consider regions dOa and eOb. The interval
between any event in this region and the event O is
spacelike. These events occur at different points in
space in every reference system. Therefore these
regions can be said to be absolutely remote relative to
O.
• Note that if we consider all three space coordinates
instead of just one, then instead of the two intersecting
lines of Fig. 2 we would have a “cone" x2 + y2 + z2 –
c2t2 = 0 in the 4-dimensional coordinate system x, y, z.
t, the axis of the cone coinciding with the t axis. (This
cone is called the light cone.) The regions of absolute
future and absolute past are then represented by the two
interior portions of this cone.
• Two events can be related causally to each other only if
the interval between them is timelike; this follows
immediately from the fact that no interaction can
propagate with a velocity greater than the velocity of
light. As we have just seen, it is precisely for these
events that the concepts "earlier" and "later" have an
absolute significance, which is a necessary condition for
the concepts of cause and effect to have meaning.
Transformation of velocities
Exercises at home
• Now, the signal transmitted by one observer, and received by another, is a light
wave. It makes a difference compared to the situation in Doppler effect with sound
waves: there is no medium, and the velocity of the wave (i.e., of light) is the same
for both observers. And the the “transmitter” and “receiver” move relative to each
other with such a speed that relativistic effects have to be taken into account.
• We will consider the following situation: the “transmitter” is in the frame O that
moves away with speed -u (meaning: to the left) from the frame O’ in which the
“receiver” located. At a moment the “transmitter” starts to broadcast a light wave.
--(/(-.
•"'*
)
&
• $
"'0
"'1+0
,"'
•
$’
•
)
t0$
+
,&
$
1#t0
t'(c + u)
'=
t 0
c t'(c + u) t 0 1
'=
= '=
' t 0 t' 1+ u /c
t'
=
t 0
1 u 2 /c 2
'=
1+ u /c
(1 u /c)(1+ u /c)
=
(1+ u /c)(1+ u /c)
1 u /c
=
1+ u /c
-.
! "
4-vectors
4-tensors
Covariant derivatives
μ
μ ; μ
x x μ
μ 1 2
μ = 2 2 2 =
c t
4-dimensional velocity
E = h
( (
E=mc2
Relativistic momentum
Relativistic momentum
The photon
Relativistic force
E i = 0 A i + i A 0 F 0i = E i
1ˆ 2ˆ 3ˆ
B i = 1 2 3
1
A A2 A3
μ F μ = μ ( μ A A μ ) = (μ μ ) A ( μ A μ )
Ma μ A μ = + A = 0
t
μ F μ = j
• Non ce ne sarebbe bisogno, ma…
dp 0 dp 0 m
qeEu = ds dt = dt 1 v 2 = Fv
d
Elastic collision: target at rest
Compton scattering
• If the incident particle 1 has 0 mass (photon)
The fall of Classical Physics
1
Classical physics at the end of XIX Century
Scientists are convinced that the particle and wave model can describe the
evolution of the Universe, when folded with
Newton’s laws (dynamics)
Description of forces
Maxwell’s equations
Law of gravity.
…
We live in a 3-d world, and motion happens in an absolute time. Time and
space (distances) intervals are absolute.
The physics entities can be described either in the particle or in the wave
model.
Natura non facit saltus (the variables involved in the description are
continuous).
Something is wrong
Relativity, continuity, wave/particle (I)
2
Something is wrong
Relativity, continuity, wave/particle (IIa)
In the beginning of the XX
century, it was known that
atoms were made of a
heavy nucleus, with
positive charge, and by
light negative electrons
Electrostatics like gravity:
planetary model
All orbits allowed
But: electrons, being
accelerated, should radiate
and eventually fall into the
nucleus
Something is wrong
Relativity, continuity, wave/particle (IIb)
If atoms emit energy in the form of photons due to
level transitions, and if color is a measure of energy,
they should emit at all wavelengths – but they don’t
3
Something is wrong
Relativity, continuity, wave/particle (III)
4
I
Light behaves like a particle,
sometimes
5
The Compton Effect
11
6
Compton Effect, Explanation
The results could be explained, again, by treating the
photons as point-like particles having
energy hƒ
momentum hƒ / c
Assume the energy and momentum of the isolated
system of the colliding photon-electron are conserved
Adopted a particle model for a well-known wave
The unshifted wavelength, o, is caused by x-rays
scattered from the electrons that are tightly bound to
the target atoms
The shifted peak, ', is caused by x-rays scattered
from free electrons in the target
13
Blackbody radiation
Every object at T > 0 radiates electromagnetically, and
absorbes radiation as well
Stefan-Boltzmann law:
Blackbody: the
perfect absorber/emitter
“Black” body
7
Experimental findings
& classical calculation
Wien’s law: the emission
peaks at
Ultraviolet catastrophe
15
Experimental findings
& classical calculation
Classical calculation (Raileigh-
Jeans): the blackbody is a set of
oscillators which can absorb any
frequency, and in level transition
emit/absorb quanta of energy:
No maximum; a ultraviolet
catastrophe should absorb all
energy Experiment
16
8
Planck’s hypothesis
Only the oscillation modes for which
E = hf
are allowed…
17
Interpretation
Elementary oscillators can have only E n
quantized energies, which satisfy 4hf 4
E=nhf (h is an universal constant, n is 3
3hf
an integer –quantum- number)
2hf 2
Transitions are accompanied by the
hf 1
emission of quanta of energy (photons)
18
9
Which lamp emits e.m. radiation ?
1) A
2) B
3) A & B
4) None
19
has been
recorded
millions of
times…
20
10
Bremsstrahlung
"Bremsstrahlung" means in German
"braking radiation“; it is the radiation
emitted when electrons are decelerated or
"braked" when they are fired at a metal
target. Accelerated charges give off
electromagnetic radiation, and when the
energy of the bombarding electrons is
high enough, that radiation is in the x-ray
region of the electromagnetic spectrum. It
is characterized by a continuous
distribution of radiation which becomes
more intense and shifts toward higher
frequencies when the energy of the
bombarding electrons is increased.
21
Summary
The wave model cannot explain the behavior of light
in certain conditions
Photoelectric effect
Compton effect
Blackbody radiation
Gamma conversion/Bremsstrahlung
Light behaves like a particle, and has to be
considered in some conditions as made by single
particles (photons) each with energy
11
II
Particles behave like waves,
sometimes
23
24
12
Diffraction of electrons
Davisson & Germer 1925:
Electrons display diffraction patterns !!!
25
de Broglie’s wavelength
What is the wavelength associated to a particle?
de Broglie’s wavelength:
13
Atomic spectra
Why atoms emit according to a discrete energy spectrum?
Balmer
Something must
be there... 27
2r=n n=1,2,3,…
14
Hydrogen (Z=1)
v
m
r
F
NB:
• In SI, ke = (1/40) ~ 9 x 109 SI units
• Total energy < 0 (bound state)
• <Ek> = -<Ep/2> (true in general for bound states, virial theorem)
29
Energy levels
The radius can only assume
values
30
15
Transitions
An electron, passing from an orbit of energy Ei
to an orbit with Ef < Ei, emits energy [a photon
such that f = (Ei-Ef)/h]
31
16
Limitations
Semiclassical models wave-particle duality can explain
phenomena, but the thing is still insatisfactory,
When do particles behave as particles, when do they behave
as waves?
Why is the atom stable, contrary to Maxwell’s equations?
33
Wavefunction
34
17
Wavefunction - II
We want a new kind of “waves” which can account for particles, old waves,
and obey to F=ma.
And they should reproduce the characteristics of “real” particles: a
but of probability
Waves such that F=ma? We’ll see that they cannot be a function in R,
but that C is the minimum space needed for the model.
35
SUMMARY
Close to the beginning of the XX century, people thought that
physics was understood. Two models (waves, particles). But:
Quantization at atomic level became experimentally evident
Particle-like behavior of radiation: radiation can be considered in some
conditions as a set of particles (photons) each with energy
18
L’equazione di Schroedinger
37
38
19
L’equazione di S.
39
20
L1 – The Schroedinger
equation
• A particle of mass m on the x-axis is subject to
a force F(x, t)
• The program of classical mechanics is to
determine the position of the particle at any
given time: x(t). Once we know that, we can
figure out the velocity v = dx/dt, the momentum
p = mv, the kinetic energy T = (1/2)mv2, or any
other dynamical variable of interest.
• How to determine x(t) ? Newton's second law:
F = ma.
– For conservative systems - the only kind that occur
at microscopic level - the force can be expressed as
the derivative of a potential energy function, F = -dV/
dx, and Newton's law reads m d2x/dt2 = -dV/ dx
– This, together with appropriate initial conditions
(typically the position and velocity at t = 0),
determines x(t).
1
• A particle of mass m, moving along the
x axis, is subject to a force
F(x, t) = -dV/ dx
• Quantum mechanics approaches this
same problem quite differently. In this
case what we're looking for is the wave
function, (x, t), of the particle, and we
get it by solving the Schroedinger
equation:
~ 10-34 Js
• In 3 dimensions,
2
The statistical interpretation
• What is this "wave function", and what can it tell you?
After all, a particle, by its nature, is localized at a point,
whereas the wave function is spread out in space (it's
a function of x, for any given time t). How can such an
object be said to describe the state of a particle?
• Born's statistical interpretation:
9
Will a normalized function stay as such?
10
Expectation values
12
More on operators
Compound operators
• Kinetic energy is
13
Angular momentum
14
Exercise
• A particle is represented at t=0 by the wavefunction
(x, 0) = A(a2-x2) |x| < a (a>0).
=0 elsewhere
a Determine the normalization constant A
b, c What is the expectation value for x and for p at t=0?
15
Exercise (cont.)
• A particle is represented at t=0 by the wavefunction
(x, 0) = A(a2-x2) |x| < a (a>0).
=0 elsewhere
d, e Compute <x2>, <p2>
f, g Compute the uncertainty on x, p
h Verify the uncertainty principle in this case
16
L2 – The time-independent
Schroedinger equation
17
18
, soluzione della prima equazione (eq.agli autovalori detta
equazione di S. stazionaria), e’ detta autofunzione
19
3 comments on the stationary solutions: 1
20
3 comments on the stationary solutions: 2
2. They are states of definite energy. In
mechanics, the total energy is called the
Hamiltonian:
H(x, p) = p2/2m + V(x).
The corresponding Hamiltonian operator,
obtained by the substitution p -> p operator
23
The infinite square well
24
Infinite square well, 2
25
Infinite square well, 3
But…
26
Infinite square well, 4
27
11/6/12
10
11
1
11/6/12
x(m)
Imagine that you're holding one end of a long rope, and you generate a
wave by shaking it up and down rhythmically.
Where is that wave? Nowhere, precisely - spread out over 50 m or so.
What is its wavelength? It looks like ~6 m
x(m)
By contrast, if you gave the rope a sudden jerk you'd get a relatively narrow
bump traveling down the line. This time the first question (Where precisely is
the wave?) is a sensible one, and the second (What is its wavelength?)
seems difficult - it isn't even vaguely periodic, so how can you assign a
wavelength to it?
12
=> Uncertainty is a characteristic of the wave representation
x(m)
x(m)
The more precise a wave's x is, the less precise is , and vice versa. A
theorem in Fourier analysis makes this rigorous…
This applies to any wave, and in particular to the QM wave function. is
related to p by the de Broglie formula
13
2
11/6/12
14
15
3
11/6/12
16
17
4
11/6/12
The Schrödinger equation outside the finite well in regions I and III is
or using
18
and the wave function must be smooth where the regions meet.
19
5
11/6/12
Penetration Depth
The penetration depth is the distance outside the potential well where
the probability significantly decreases. It is given by
mx m
Et (V0 E) = (V0 E)
p 2m(V0 E) 2m(V0 E) 2
20
21
6
11/6/12
The wave function will consist of an incident wave, a reflected wave, and a
transmitted wave.
The potentials and the Schrödinger wave equation for the three regions are
as follows:
As the wave moves from left to right, we can simplify the wave functions to:
22
23
7
11/6/12
Tunneling
Now we consider the situation where classically the particle does not have enough
energy to surmount the potential barrier, E < V0.
The quantum mechanical result is one of the most remarkable features of modern
physics, and there is ample experimental proof of its existence. There is a small,
but finite, probability that the particle can penetrate the barrier and even emerge
on the other side.
The wave function in region II becomes
24
Uncertainty Explanation
Consider when L >> 1 then the transmission probability becomes:
25
8
QM in 3 dimensions
1
V(r): separation of variables
2
ˆ = 2 2 ,
,
ˆ
ˆ
Y = ll 1Y 2 = ll 1
2
ˆ
2 . 3
The angular equation
4
The angular equation -
= i ,
ˆ
ˆ = m
ˆ = m
ˆ
. 5
The angular equation -
6
7
ˆ = l(l + 1)
8
Spherical harmonics
= (-1)m for m>=0 and =1 for m<0. The Y are orthogonal, so
9
The radial equation
10
The H atom
u(r) = rR(r)
11
Asymptotic behavior
u(r) = rR(r)
12
The radial solution
u(r) = rR(r)
n>l
14
Ground state
15
n>1
16
n > 1 (continued)
17
Example
18
Graphs
19
Angular momentum
20
21
22