Class Notes For Modern Physics Part 2
Class Notes For Modern Physics Part 2
J. Gunion
U.C. Davis
J. Gunion
The particle nature of matter
Here, I will try to say a few additional words about the Thomson and
Rutherford experiments. Please read the material in the book on the
Millikan experiment.
Thomson
F eE Ve l
ay = = = and t= . (1)
me me me d vx
This gives
V le vy Vl e
vy = , tan θ = = . (2)
mevxd vx vx2 d me
E V
qE = qvxB , ⇒ vx = = . (3)
B dB
J. Gunion 9D, Spring Quarter 5
~ that gives exact balance, we get
Thus, using the B
Rutherford
Based on his own experiments and those of others, in which it was clear that
an atom was not a simple object, but had balancing negative and positive
charged particles in it, with the negative one having a much smaller mass
than the positive one, Thomson proposed the “plum-pudding” picture of
the atom.
Of course, this picture failed to explain the rich line spectra that people
were finding for excited atoms, in particular even the simplest Hydrogen
atom.
(2e)(Ze)
F =k (5)
r2
(where α has charge 2e in magnitude, the nucleus has charge Ze in
magnitude, and k is Coulomb’s constant). He predicted the following
result for scattering:
k2Z 2e4N nA
∆n = 4 , (6)
4R2( 12 mαvα
2 )2 sin (φ/2)
where R and φ appear in the figure, N is the number of nuclei per unit
area of the foil (and is thus proportional to the foil thickness), n is the
total number of α particles incident on the target per unit time, ∆n is
the number of α particles entering the detector per unit time at an angle
φ, and A is the area of the detector. The velocity vα is determined from
the accelerating potential difference between the α emitter and the gold
foil (or other) target: Kα = 12 mαvα 2
= (2e)V (non-relativistic ok here,
and use charge of α = 2e).
1 2 (Ze)(2e)
mαvα =k (7)
2 dmin
to solve for dmin. When Rutherford’s prediction of eq. (6) starts to fail,
the corresponding dmin presumably is the point at which the α is actually
impinging into the nucleus itself rather than just Coulomb scattering from
it.
Example
Estimate the radius of the Aluminum nucleus.
In 1919, Rutherford was able to show a breakdown in eq. (6) for 7.7 MeV
α particles scattered at large angles from aluminum (Z = 13). Estimate
the radius of the Al nucleus.
Answer: assume all the α kinetic energy, Kα, goes into potential energy
2Ze2
dmin = k
Kα
2(13)(1.6 × 10−10 C)2
= (8.99 × 109 N · m2/C 2)
(7.7 × 106 eV )(1.60 × 10−19 J/eV )
= 4.9 × 10−15 m . (8)
As mentioned earlier, many spectral lines had been seen, coming from the
sun, coming from excited atoms, and so forth. There was no explanation
yet. ***Do spectral line demo.***
Bohr
Bohr followed the lead of Planck and Einstein by assuming that if light
was quantized then why shouldn’t atomic electronic orbits be quantized
in some way. Then, spectral lines could arise when an electron jumped
from one such electronic orbit to another orbit (of lower energy) by
emitting a photon of definite frequency given by ∆E = hf .
Let us see why this works. (For atoms, we can use non-relativistic
procedures with adequate accuracy.)
1. The electric potential energy of the e− is U = −ke2/r, where
k = 1/(4π0).
2. The total energy of the atom is the sum of the potential and kinetic
energies,
1 2 e2
E = K + U = me v − k . (12)
2 r
3. Meanwhile, Newton’s force law says force=centripetal acceleration, or
ke2 v2 1 2 ke2
= me , which ⇒ K= me v = . (13)
r2 r 2 2r
ke2
E=− . (14)
2r
5. Next, we solve for v in terms of n and r using the equations above,
1 2 ke2
mevr = nh̄ and 2 mev = 2r , to obtain
n2h̄2
rn = , n = 1, 2, 3, . . . (15)
me ke2
h̄2
a0 = = 0.0529 nm . (16)
me ke2
6. Finally, substitute the form of rn into the equation for E just above,
i.e. eq. (14), to obtain
2
ke 1 13.6
En = − =− eV . (17)
2a0 n2 n2
One finds that ke2/(2a0h) = R, the Rydberg constant and one gets a
theoretical post-diction of the Balmer formula.
An Example
Suppose the stellar atmosphere has a temperature of order T =
79, 000 K. (a) Is it reasonable to expect that a lot of the Hydrogen atoms
3
kB T = (1.5)(8.62 × 10−5 eV /K)(79000) = 10.2 eV . (19)
2
hf = E2 − E1 ⇒
hc (4.136 × 10−15 eV · s)(3 × 108 m/s)
λ = =
E2 − E1 10.2 eV
= 1.22 × 10−7 m = 122 nm , (21)
2
ke 4
En = . (24)
2a0 n2
1 f Ei − Ef
= =
λ c hc !
2
ke 4 4
= −
2a0hc n2i n2f
ke2
1 1
= − (25)
2a0hc (ni /2)2 (nf /2)2
vn=1 1 h̄
= from mvr = nh̄
c c mr1
1 ke2 h̄2
= from r1 = me ke2
c h̄
ke2 1
= ≡α= . (26)
h̄c 137
Note how small v1/c is. The non-relativistic approximation employed by
Bohr was ok.
The quantitiy α is sometimes called the fine structure constant. It is a
very useful characterization of the strength of the E&M force. Other
forces, such as the strong and weak forces that we will learn more about
late in the quarter, have different strengths.
Such dimensionless ratios constructed using known physical constants
(here, h̄, c, k and e) are typically of deep theoretical significance. To
construct α, we needed the new fundamental constant h̄.
For small V , these collisions were elastic and the e−’s retained most of
When V is increased further, more and more electrons reach the collector
until once again the current suddenly dips. What is happening is that V
is large enough for the accelerated e−’s to have two inelastic collisions
with two subsequent Hg atoms.
How could they check this? Well, if they really had excited the Hg atomic
electron to a higher level, it should emit a photon of the corresponding
frequency when this electron fell back down to its original lower level.
hc hc 1240 eV · nm
e∆V = ∆E = hf = , ⇒ λ= = = 253 nm .
λ ∆E 4.9 eV
(28)
This is the precise wavelength they observed. In 1925, they were awarded
the Nobel prize for this confirmation of Bohr’s theory.
The next big question was why should angular momentum be quantized
in the manner proposed by Bohr?
What turns out to be the fundamental idea was that developed by de
Broglie in 1925.
He speculated that if light, a wave phenomenon originally, also had a
particle-like nature, then why not the reverse?
He also was looking for a way to explain the integers and quantization
that emerged in Bohr’s atomic theory, which concerned electrons circling
a nucleus. The only way that integers had cropped up in the past was
h E
λ= and f = . (29)
p h
(We will return to the problem with this that arises if you compute wave
velocity as f λ = E/p using the relativistic formulae, p = γ(u)m0u and
E = γ(u)m0c2, which would give E/p = c2/u > c.)
De Broglie noticed that if we employ the photon-like formula p = mve =
h/λ and plug this into mevr = nh̄, we find
h h 2πr
r=n ⇒ = n. (30)
λ 2π λ
In other words, the circumference of the electron orbit must contain an
integer number of electron wavelengths, which, in turn, implies that the
Since I want to make sure we are all on the same page, I will give a very
brief review of wave equations and the E&M wave equation in particular.
The latter material was in sections 32.2 and 32.3 of University Physics
Part 2 by Young and Freedman, which is I believe the text employed for
your earlier courses. Had you covered chapters 35 and 36 of this text,
which I understand you probably did not, you would have seen a very
detailed discussion for light of the interference phenomena and so forth
that we have already talked about. But, you did, I believe, cover these
phenomena for mechanical waves, which is where we begin.
First, recall the differential equation that you studied and understood for
mechanical waves on a string or in water.
∂ 2y(x, t) 1 ∂ 2y(x, t)
= , (31)
∂x2 v2 ∂t2
where y(x, t) is the displacement of the string, or ... at location x and
time t, and v is the velocity with which the wave moves along the string.
∂ 2f ∂ 2f ∂ 2f 1 ∂ 2f
= f 00(θ) and = v 2f 00(θ) ⇒ = . (32)
∂x2 ∂t2 ∂x2 v2 ∂t2
2π
f = sin (x − vt) (33)
λ
2π 1 v
vT = 2π , or f = = . (34)
λ T λ
J. Gunion 9D, Spring Quarter 36
Often, it is more convenient to write
2π
(x − vt) ≡ kx − ωt , (35)
λ
are both also solutions to the wave equation and the earlier sin form can
be written as (using eib = cos b + i sin b)
1 h i
sin(kx − ωt) = ei(kx−ωt) − e−i(kx−ωt) . (37)
2i
+∞
1
Z
f (x − vt) = √ dkfe(k)eik(x−vt) , (38)
2π −∞
where k can run over negative as well as positive values and ω(k) = vk
is implicitly required by this form. The functional form fe(k) defines√the
Fourier decomposition of the wave solution f (x − vt). The inverse 2π
is just a conventional choice.
The above might be complex. So, for a real observable like string
displacement, we would take the real part of f (x − vt). It will still be
a solution of the wave equation since it will still be a function only of
x − vt.
1
fe = 0, k < k0 − ∆k (39)
2
1 1
fe = 1, k0 − ∆k ≤ k ≤ k0 + ∆k (40)
2 2
1
fe = 0, k > k0 + ∆k . (41)
2
This fe(k) is plotted below. (The book, Example 5.7, uses the notation
fe(k) = a(k).)
What you see in the figure for Re[f (x)] is the rapid oscillation of
the cos(k0x) = Re[eik0x] factor within an envelope described by the
sin(∆k x/2)/(∆k x/2) factor. The latter has its first nodes at ∆k x/2 =
±π, i.e. x = ±2π/∆k. (There is no node at x = 0 because
1
∆x∆k = . (44)
2
1
∆x∆k ≥ . (45)
2
2π 2πp p
k= = = , (46)
λ h h̄
eq. (45) can be rewritten as
h̄
∆x∆p ≥ . (47)
2
This is the famous Heisenberg uncertainty principle that was first proposed
for matter, and only later was it realized that it was already present in
the description of light waves as photon packets with p = h/λ.
We will return to a thorough discussion of the implications of this kind of
uncertainty principle. However, you should at this point take note of the
Z +∞
1 ikx
f (x ) = √ fe(k)e dk
2π Z−∞
k0 +∆k/2
1 ikx
= √ e dk
2π k0−∆k/2
1 1 h i(k0+∆k/2)x i(k0 −∆k/2)x
i
= √ e −e
2π ix
1 eik0x
1
= √ 2 sin ∆k x
2π x 2
∆k sin(∆k · x/2) ik0x
= √ e . (48)
2π (∆ k · x/ 2)
D (D/2)sin θ
2
θ
Path difference between top of slit and middle of slit = (D/2)sinθ = λ/2
for full cancellation requires θ=λ/D
py py λ θh λh h
θ∼ ∼ ⇒ py ∼ ∼ ∼ (49)
px h λ Dλ D
from which we find (can’t expect to get 2π type factors right here)
h
∆py ∆y ∼ py D ∼ D ∼ h. (50)
D
Let me now give a brief review of the E&M wave equation. Had I known
that this was only given very brief attention in your previous course, I
would have surely begun this quarter’s lectures with the following review.
One starts with the two Maxwell equations:
d
I Z
~ · d~l = −
E ~ · n̂ dA
B (51)
C dt S
d
I Z
~ · d~l = µ00
B ~ · n̂ dA
E (52)
C dt S
∂Ey ∂Bz
= − (53)
∂x ∂t
∂Bz ∂Ey
= −µ00 (54)
∂x ∂t
I give a brief derivation of the first equation. We apply eq. (51) to the
case of E~ = ŷEy and B ~ = ẑBz . The equation says that a time varying
Bz (using n̂ = ẑ and very small size dx by dy loop in the x, y plane) can
generate an E ~ field that circulates around the small loop. Applying this
to the E~ = ŷEy case, we find that Ey must vary with x. In short, the
time varying Bz field is generating a spatial variation of Ey as a function
of x. Of course, Ey will end up with time dependence that matches that
of the time derivative of Bz . A figure showing how this application works
is below.
E_y(x) dy
n E_y(x+dx)
B_z(x)
dx
z
x
Figure 10: Set up for deriving eq. (53).
∂Fy ∂Fx
~ ×F
(∇ ~ )z = − (56)
∂x ∂y
∂Fx ∂Fz
~ ×F
(∇ ~ )y = − ; (57)
∂z ∂x
these will be needed in the differential forms of eqs. (51) and (52),
respectively.
Using this theorem in eqs. (58) and (59) and the fact that the surface
S, and its normal n̂, can be thought of as being arbitrary, the integrands
~
∂B
~ ×E
∇ ~ = − (58)
∂t
~
∂E
~ ×B
∇ ~ = µ00 , (59)
∂t
∂Ey ∂Bz
~ × E)
(∇ ~ z= , ~ × B)
(∇ ~ y=− . (60)
∂x ∂x
Substituting the above into eqs. (58) and (59), respectively, we get
eqs. (53) and (54), repeated below.
∂Ey ∂Bz
= − (61)
∂x ∂t
∂Bz ∂Ey
− = µ00 (62)
∂x ∂t
J. Gunion 9D, Spring Quarter 52
Let us now consider the equation that can be derived from eqs. (53) and
∂ ∂ 2 Ey ∂ ∂Bz ∂Bz
(54). Take ∂x eq. (53) ⇒ ∂x2 = − ∂t ∂x
and substitute for ∂x
using
∂Ey
eq. (54), which states ∂B
∂x
z
= −µ0 0 ∂t to obtain:
∂ 2 Ey ∂ 2 Ey
= µ00 . (63)
∂x2 ∂t2
∂ 2 Bz ∂ 2 Bz
= µ00 . (64)
∂x2 ∂t2
1
Bz = A sin(kx − ωt) , (66)
c
by employing either eq. (53) or eq. (54). For example, eq. (54) states
1 ∂Ey
that ∂B
∂x
z
= − c2 ∂t
. Substituting in the above forms we get
∂Bz 1
= A k cos(kx − ωt)
∂x c
1 ∂Ey 1
− = − 2 A(−ω) cos(kx − ωt)
c2 ∂t c
1 ω
= A cos(kx − ωt)
c c
1
= A k cos(kx − ωt) using ω/k = c . (67)
c
The fact that Ey and Bz are exactly “in phase” all the time, is one of
the remarkable features of E&M radiation. But, it had to be true in
order for one to “feed” the other.
1 ~ ·B
1B ~
uE = ~ ·E
0 E ~, uB = , (68)
2 2 µ0
~ = |E|/c.
respectively. As we have seen above, for the travelling wave, |B| ~
So, u = uE + uB can be written in a variety of forms:
1 ~ 2
1 |B|
u = ~ 2+
0|E|
2 2 µ0
~ 2
= 0|E|
~ 2
|B|
=
µ0
0
r
= ~ B|
|E|| ~ . (69)
µ0
Suppose the maximum |E| ~ value for a traveling sinusoidal wave, moving in
the x direction is Emax = |E|~ = 100 N/C and occurs at t = 0, x = 0.
Give a value for the amount of energy impacting a screen perpendicular
to the x axis per unit area per unit time at t = 0, x = (2/3) × λ.
Answer: Since the field is maximum at t =0, x = 0, it is convenient
to use the form Ey = Emax cos 2λπ (x − ct) . Substituting t = 0, x =
(2/3)λ gives Ey = Emax cos(4π/3) = − 21 Emax. From our earlier
equations, we have
S = cu = c0Ey2
1
= (3 × 108 m/s)(8.85 × 10−12 C 2/N · m2)(− 100 N/C)2
2
J. Gunion 9D, Spring Quarter 56
= 6.6375 J/(m2 · s) . (70)
Of course, as time passes, at this same location, the S value will oscillate
up and down and so the average energy per unit area per unit time will
be Saverage = 21 Smax = 12 c0Emax2
. This is what we usually call the
intensity of the E&M wave, but we see that a more accurate name
would be average intensity.
We do not know how many photons this corresponds to (on average)
until λ is specified. Also note that we could compute (at any instant)
Bz using Bz = Ey /c.
We have stated that the relation between the energy and the momentum
carried by an E&M wave is p = E/c, where in the continuous wave
vision E is the same as S. p will then be so much momentum per unit
area per unit time.
To derive this relation between the energy and momentum carried by
an E&M wave, is a bit of an exercise. I give it below in case you are
interested.
dpx Ey
= Qvy Bz = Qvy . (72)
dt c
dU
∆U = QEy ∆y , ⇒ = QEy vy . (73)
dt
Substituting the result of solving this equation for vy into the previous
S 6.6375J/(m2 · s)
p= = = (2.212 × 10−8 kg · m/s)/(m2 · s) . (75)
c 3 × 108 m/s
General Lesson
We have seen that a single slit will have a minimum at sin θ = λ/D
coming from the complete cancellation of various Huygen’s wavelet
amplitudes from different parts of the slit. At θ = 0, all the wavelets
arrive in phase at the central point of the screen and simply add up to
give you a maximum E field, call this maximum E0. E0 will have some
wave-like form, of course, and so will oscillate up and down as time passes
according to some form like
2π
E0 = Emax sin (xscreen − ct) . (76)
λ
One can also add up the wavelets for any other θ. I will not go through
the derivation, but the result is
2
sin[πD(sin θ)/λ] sin[πD(sin θ)/λ]
E = E0 , ⇒ I = I0 (77)
πD(sin θ)/λ πD(sin θ)/λ
Next consider the case of two such slits. If we cover up either one
of the slits, we will get a uniform E0 on the detecting screen and the
corresponding I0 ∝ |E0|2, just as discussed above. But if we now open
up both slits, we will get our two-slit interference pattern. Let us call the
distance from the upper slit, slit #1, to the screen L. Then the distance
from the lower slit #2 to the screen is L + S sin θ.
L + S sin θ
S θ
S sin θ
First, let us note that the two waves cancel when S sin θ = (n + 12 )λ,
Of course, for general θ, L = xscreen/ cos θ, but this is not needed for
the above discussion.
Now, let us return to the question posed at the end of the last class, for
which we focus on the case of θ = 0. Then,
2π
E 1+2 = 2Emax sin (xscreen − ct) , ⇒ I 1+2 = 4c0|E0|2
λ
(78)
implying that I 1+2 = 4I0, where I0 is that from just one slit! Here, I0
denotes the instantaneous intensity. For the average intensity, the same
applies:
1
hI 1+2i = 4hI0i = 4 I0max . (79)
2
In contrast, as I have said before, the matter waves that we shall come
to do not have any such direct physical interpretation.
~ and B
Matter waves are sort of like putting the E ~ together in the form
~ =E
F ~ + icB
~. (80)
~ , we get
If we take the absolute square of this F
~ | 2 = (E
|F ~ + icB)
~ · (E
~ − icB)
~ = |E|
~ 2 + c2|B|
~ 2 (81)
Ψ = Ψ1 + iΨ2 . (82)
First, we note again that the wave equation for X being either Ey or Bz
takes the form
∂ 2X 1 ∂ 2X
2
= 2 2
. (84)
∂x c ∂t
Suppose we write the energy momentum relationship for light in the form
2 E2
p = c2 , multiply this times X and then make the replacements
∂ h̄ ∂
E → ih̄ and p→ . (85)
∂t i ∂x
J. Gunion 9D, Spring Quarter 65
Then,
E2 ∂ 2X 1 ∂ 2X
p2X = 2
X ⇒ −h̄2 = , (86)
c ∂x2 c2 ∂t2
which contains our wave equation. A hand-waving motivation for these
identifications is to note that for a wave solution of the type eik(x−ct)
(our general form for the case of v = c) that we were discussing earlier,
it is certainly the case that
h̄ ∂
eik(x−ct) = h̄keik(x−ct) = peik(x−ct) , (87)
i ∂x
where we used
h 2π h
h̄k = = =p (88)
2π λ λ
for a photon within an E&M wave. Similarly,
∂
ih̄ eik(x−ct) = h̄kceik(x−ct) = Eeik(x−ct) , (89)
∂t
h̄kc = pc = E (90)
More de Broglie
1q 2 2 2 4
1p 2 2
p = (eV + mec ) − mec = e V + 2eV mec2
c
√ s c
2eV mec2 eV
= 2
+ 1. (91)
c 2mec
h hc 1
λ = =√ q
p 2eV me c2 eV
+1
2m e c 2
h 1 1
= √ p q . (92)
2me × 1eV V (volts) eV
+1
2me c2
h 6.63 × 10−34 J · s
√ = p
2me · 1eV 2(9.11 × 10−31 kg)(1.6 × 10−19 J )
= 1.227 nm . (93)
1.227 nm
λ= √ = 1.67 × 10−10 m . (94)
54
What did they see? From X-ray measurements DG knew that their
atomic spacing was d = 2.15 × 10−10 m. As illustrated in the next
figure, they found constructive interference for φ = 50◦ corresponding to
Because of equal entry and exit angles for the e− (or any other particle,
e.g. neutron), the waves from any two atoms on any one horizontal crystal
2D cos θ = mλ (96)
More Examples
6.63 × 10−34 J · s
λairplane = = 1.23 × 10−40 m
(18000 kg)(300 m/s)
6.63 × 10−34 J · s
λelectron = = 2.43 × 10−6 m .(97)
(9.11 × 10−31 kg)(300 m/s)
The latter is something you can hope to measure using the kind of
techniques just described. The former is not something you could ever
measure — we do not need to worry about wavelengths and wave patterns
in our everyday world!
(II) Consider a two-slit experiment using electrons. The slits are assumed
to be very narrow compared to the wave-length of the electrons. Beyond
the slits is a bank of e− detectors. At the center detector, directly in the
path the beam would follow if unobstructed, 100 electrons per second
are detected. Suppose that as the detector angle varies, the number per
Answers:
h h
(a) At the minimum, we require S sin θX = 12 λ. We need, λ = p
= mv
and we get v from
1 −19 1
K= 2
me v , ⇒ (1 eV )×(1.6×10 J/eV ) = (9.11×10−31 kg)v 2
2 2
(98)
which gives v = 5.93 × 105 m/s. From this we get
h 6.63 × 10−34 J · s
λ= = = 1.23 × 10−9 m. (100)
p 5.40 × 10−25 kg · m/s
(b) In the discussion that follows, we do not write the wave form explicitly.
But, there is always a wave form present. In the present case of e−
waves the wave form would be something like
initially, i.e. before passing through a slit, and would afterwards be similar
in form with kx − ωt replaced by kr − ωt, where r is the distance from
a slit. The important point will be that whatever the wave form, the
two slits will have equivalent wave forms, that are in phase at a central
Since the two slits are very narrow and the waves from the two slits add
equally at this point of constructive interference, the amplitude of either
individual wave must be half the total:
With one slit closed, the electron flux at the central detector would be
1/4 the two slit flux, or 25/s. Note the importance of assuming that the
slits are really narrow. In this case, this same flux would apply for all the
electron detectors, regardless of angle, when only one slit is open.
(c) If one slit were open and its width (already very narrow) were reduced
to 0.36 of its original size, all detectors would register an electron flux
that is 0.36 × 25/s, or 9/s. In equation form, this means that
With both slits open (but with slit #1 at only 36% of its original size),
we have two waves of different amplitudes, one proportional to 5 (slit
#2 of original width) and one proportional to 3 (slit #1). At points of
constructive interference, such as the central detector, where the waves
1
(64/s + 4/s) = 34/s , (108)
2
i.e. the sum of the 9/s and the 25/s expected from each slit alone.
To repeat, to find the probability (or flux) at a given location, we do
not add the probabilities (or fluxes) from each slit at that location; these
ch 1.24 × 103 eV · nm
λ= ∼ = 1.24 × 10−7 nm . (110)
E 10 × 109 eV
Let us review once more the HUP. We have found by example that for any
wave pattern it is always true that
1
∆k∆x ≥ , (111)
2
where I have stated that the minimum arises for Gaussian wave packets.
We then input either Planck (photons) or de Broglie (matter waves) via
the relation
h 2π 1
p = = h̄ = h̄k , ⇒ ∆p∆x ≥ h̄ . (112)
λ λ 2
Another uncertainty relation involves the uncertainty in energy of a wave
packet, ∆E, and the time, ∆t, taken to measure that energy. Using
Gaussian or other wave forms that are functions of kx − ωt and that are
of finite extent in ∆t, we can derive the wave result that
1
∆ω∆t ≥ , (113)
2
J. Gunion 9D, Spring Quarter 85
where once again the minimum is for Gaussian forms.
We now input the relation
1
E = hf = h̄(2πf ) = h̄ω , ⇒ ∆E∆t ≥ h̄ . (114)
2
This result states that the precision with which we can know the energy
of some system is limited by the time available for measuring the energy.
∆px∆x
Here we consider an idealized (thought) experiment in which we try
to measure the position of a particle using photons. A more careful
treatment is given in the book. Here, I just give the idea of the
argument.
h
• The photon carries momentum given by p = λ .
• The matter particle tends to pick up some portion of this momentum
(depending upon angle of incidence which is determined by size of lens
h
(∆p)particle being probed ∼ . (115)
λ
• Also, the position of the particle can not be determined to any greater
precision than the wavelength λ of the light:
• Multiplying, we get
∆E∆t
A similar argument is possible for the ∆E∆t relation.
• Consider a wave of frequency f incident on a particle at rest.
• Suppose that the minimum uncertainty in the number of waves we can
count is 1 wave.
# we count
Since f = , we get
time interval
1
∆f = , (118)
∆t
hc
∆E ∼ ; (121)
λ
J. Gunion 9D, Spring Quarter 89
i.e. if you want to not change the energy of the particle the photon is
probing, you must keep λ large. But, then this means it takes longer
for the wave front arrival to be clearly defined. The result is
hc λ
∆E∆t = ∼ h. (122)
λ c
e− in a Hydrogen atom
Is there any relation between the energy levels of the Hydrogen atom
and the uncertainty principle? Let’s see.
We suppose that the electron is confined in a one-dimensional sense
to a region of order ∆x. Then, let us employ the HUP in the form
∆px ∼ h̄/∆x. (I have chosen the numerical factor to give me the
prettiest results.) The associated kinetic energy is
(∆px)2 h̄2
K ≥ (K)∆px ≡ > . (123)
2me 2me (∆x)2
h̄2 ke2
(K)∆px ∼ ∼ − . (124)
2me(∆x)2 ∆x
ke2
U =− , (126)
∆x
J. Gunion 9D, Spring Quarter 91
which can be combined with our minimum K for the particle to compute
the total energy:
h̄2 ke2
E =K+U = − . (127)
2me (∆x)2 ∆x
∂E h̄2 ke2
= 0 = −2 + (128)
∂∆x 2me (∆x)3 (∆x)2
which is solved by
h̄2
∆x = = a0 , (129)
me ke2
which, after substitution into the above form for E gives
k2e4me
E=− 2 (130)
2h̄
J. Gunion 9D, Spring Quarter 92
which is precisely the E0 energy level of the first Bohr orbit.
The Unstable Z boson
The Z boson is an unstable (i.e. a particle that decays) with mass
mZ ∼ 91 × 109 eV . The average lifetime of the Z is
35
N=2 DELPHI
30 N=3
N=4
25
20
15
10
5
88 89 90 91 92 93 94 95
Energy, GeV
Figure 17: e+e− → hadrons as a function of Me+e− . The Z peak is
centered about mZ = 91.13 × 109 eV = 91.13 GeV and has a width of
roughly 2 to 3 × 109 eV .
1 6.582 × 10−16 eV · s
∆E ≡ ∆mZ ∼ h̄ = ∼ 2.3×109 eV = 2.3 GeV ,
τZ 2.9 × 10−25 s
(132)
and this is what is explicitly seen in the plot. The plot also shows how
the peak would get narrower (broader), relative to its height, if certain
decay modes are eliminated (added).
Things we know
Assume equal slit widths.
1. It is only when we have both slits open that the interference pattern
develops.
2. Even if we send only one e− at a time, if both slits are open the e−
hits at the detector bank will accumulate where the interference wave
1. Suppose you have both slits open, but you try to measure unambiguously
which slit a given e− passes through.
⇒ you disturb the e−.
2. For example, place some detecting particles on the right side of the
slit. Use the recoil of one of these particles to determine which slit the
e− goes through.
3. To decide which slit, need to measure the detecting particle’s position
with ∆y D (D in the figure is the separation between slits, not the
size of an individual slit).
4. During the collision, the detecting particle suffers a change in momentum
∆py , equal and opposite to the change in momentum experienced by
the e− passing through the slit.
py h
tan θ ∼ θ = = . (133)
px 2pxD
∆py h h
θ= or ∆py (134)
px 2pxD 2D
h h
(∆py ∆y)detecting particle ·D = . (137)
2D 2
This is in clear violation of the uncertainty principle.
If ∆y is small enough to determine which slit the electron goes through,
∆py will be so large that the e−’s will be deflected all over the place
and the interference pattern will be destroyed.
I will be, and have been for that matter, kind of combining the material
appearing in Chapters 5 and 6. You should read this material as a unit. For
example, I have set up the ingredients for writing down the Schroedinger
equation that first appears in Sec. 6.3, where it is introduced more or less
by fiat. I will now go a little bit beyond just simply writing it down and tell
you one way that you can kind of understand it.
h̄ ∂ ∂
px = , E = ih̄ . (138)
i ∂x ∂t
2 2
2 p 2 p
E = m0 c + , ⇒ EΨ = m0 c + Ψ
2m 2m
∂Ψ h̄2 ∂ 2Ψ
⇒ ih̄ = m0 c2 Ψ − . (139)
∂t 2m0 ∂x2
Ψ = Aei(kx−ωt) , (140)
i(kx−ω (k)t 2
2 2
Aei(kx−ω(k)t) ,
ih̄(−iω(k))Ae = m0c − h̄ (ik) (141)
implying
m0 c2 h̄k2
ω(k) = + . (142)
h̄ 2m
Note how this is consistent with the de Broglie / Planck relations
2π 2πp p E E
k= = = , and ω = 2πf = 2π = (143)
λ h h̄ h h̄
or equivalently
h 2π
p= = h̄ = h̄k , E = hf = h̄2πf = h̄ω (144)
λ λ
(h̄k)2
E = h̄ω = m0c2 + . (145)
2m0
∆x ω(k) m0 c2 h̄k m0 c2 p
= = + = + (147)
∆t k h̄k 2m0 p 2m0
has any meaning. The answer is no. We will have to deal with wave
This is because
|Ψ|2 = A2 (148)
is completely independent of x and so the particle has a uniform
probability of being anywhere along the x axis! Obviously, it is nonsense
to discuss the velocity of a uniform probability distribution.
We must now face the subtle issue of how to construct a wave form
that can describe an actual physical particle and how it is we determine
the velocity of the particle. This will bring us to consider the difference
between group and phase velocity.
We considered in Fig. 8 and surrounding material how to create a photon-
like object by adding together E&M type wave patterns. There, the
group and phase velocities were both equal to c and we did not distinguish
or even discuss. For massive particles one must be careful.
The book has a discussion using two sin waves. However, I prefer to
use the plane wave form we have just been discussing, which is an actual
solution of the SE (unlike the sin or cos forms alone).
To define a “particle” we clearly cannot use a single plane-wave solution
for which the particle has no preferred location.
We must superimpose a least two plane-wave solutions.
Let us begin with exactly two and see what happens.
k1 = k0 + dk , k2 = k0 − dk , ω1 = ω0 + dω , ω2 = ω0 − dω .
(150)
• Then,
h i
Ψ = A ei[(k0+dx)x−(ω0+dω)t] + ei[(k0−dx)x−(ω0−dω)t]
h i
i(k0 −ω0 t) i[(dk)x−(dω )t] −i[(dk)x−(dω )t]
= Ae e +e
= Aei(k0x−ω0t)2 cos [(dk)x − (dω)t] . (151)
ω0 h̄ω0 E0
vp ≡ = = , (152)
k0 h̄k p0
dω
vg ≡ , (153)
dk
which is called the group velocity.
• It is this latter term that defines the envelope of the wave and tells us
where the probability is. Indeed,
From this we see that for a massless photon, vp = c, but that for a
massive particle vp > c. However, this is not a problem since vp does
not describe where the probability is!
E(p = h̄k) 1
q
ω(k) = 2πf = = h̄2k2c2 + m20c4 ,
⇒ h̄ h̄
dω(k) 1 h̄2kc2
vg = = q
dk h̄ h̄2k2c2 + m2c4
0
c c2
= q = , (158)
1+ m 0c
2 vp
h̄k
E γ(u)m0c2 c2 c2
vp = = = ⇒ vg = = u! (159)
p γ(u)m0u u vp
should be the probability that the particle will be found in the infinitesimal
interval dx about the point x. Once again, I stress that Ψ itself is
not something you can observe and even |Ψ|2 is only observable in a
probabilistic sense.
Because of its relation to probabilities, we will typically insist on the
following.
1. Ψ should be single valued and a continuous function of x and t. In this
way, no ambiguities will arise concerning the predictions of the theory.
2. In fact, we will typically require that not only should Ψ be continuous,
but it should also be smooth in the sense that its first derivatives are
finite.
Similarly, Z b
Pab = |Ψ(x, t)|2dx (162)
a
should be the probability of finding the particle in the interval a ≤ x ≤
b.
In this formulation of QM, the fundamental problem of QM is the
following: Given the wavefunction at some initial instant, say t = 0, what
is the wavefunction at any subsequent time t.
To answer this question requires a dynamical equation for Ψ(x, t). We
1. We start with Z ∞
ikx
Ψ(x, 0) = Ψ(k)e
e dk (163)
−∞
√
with Ψ(k) = (Cα/ π) exp[−α2k2].
e
Since Ψ(k)
e is centered about k = 0, we will be constructing a wave
packet that has a central momentum and velocity of 0 (motionless
particle), although there will be a spread of velocities and momenta
about this central value.
2. We can perform the k integral if we carefully examine it and use the
Cα ∞ x2
Z 2
−(αk− 2ix ) −
Ψ(x, 0) = √ dke α 4α2
π −∞
C −x2/4α2 ∞ −z2
Z
= √ e e dz defining z = αk − ix/(2α)
π −∞
2 2 2
= Ce−x /4α = Ce−(x/2α) . (164)
R∞ 2 √
Above, we used −∞
e−z dz = π — see any integral table.
A Gaussian form has begot another Gaussian form.
3. We now state that the appropriate way to define the width of a
Gaussian form is to take the probability and write it in the form
k2 1 x2
2α2k2 = 2
, and 2 2
x 2
= (166)
2(∆k) 4α 2(∆x)2
1. For t 6= 0, use
Z ∞
i(kx−ω (k)t)
Ψ(x, t) = Ψ(k)e
e
−∞
Z ∞ h̄k 2
i kx− 2me t
= Ψ(k)e
e dk
−∞Z
∞
C −α2 k2 +ikx−ih̄tk2 /(2me )
= √ e dk (169)
π −∞
I will not go through the details of the variable shift and integration.
The fact that the k2 coefficient is complex is not any particular
problem. You just R∞have to trust your√square completion process and
use the standard −∞ exp[−y 2]dy = π at the appropriate point.
3. The important component of interest is the residual coming from the
last term above which says that
x2
Ψ(x, t) ∝ exp − . (171)
ih̄t
4 α2 + 2m e
" # " #
2 2
x x
2
|Ψ(x, t)| ∝ exp − 14
ih̄t
× exp − 14
2
α + 2m e α2 − 2im
h̄t
e
1 2 1 1
= exp − 4 x 2
+ 2
α + ih̄t/(2m
e) α − ih̄t/(2me)
2α2
= exp − 14 x2
h i2
α4 + 2h̄t
me
2
1 x
= exp −
i2
2 α2 + h̄t
h
2m e α
2
1 x
≡ exp −
2 [∆x(t)]2
2
h̄t
⇒ [∆x(t)]2 = [∆x(0)]2 + , (172)
2me∆x(0)
where we used α = ∆x(0). In getting from the 2nd to the 3rd line
Non-stationary case
If we want a particle whose central location is moving, we simply modify
our input form of Ψ(k)
e so that it is a Gaussian (or other choice) centered
about k = k0. The packet would then move with a group velocity given,
in the NR case, by
2
dω(k) d h̄k h̄k0
vg = = = . (174)
dk k = k0 dk 2m0 k = k0 m0
The wave packet would, just as in the k0 = 0 case discussed above, spread
out as time passed due to the presence of a distribution of momenta and
velocities for individual subcomponents of the wave packet.
Definition of ∆x?
So, let us now use the probability interpretation of Ψ(x, 0) to actually
2 2
2
2 2
P (x) = |Ψ(x, 0)|2 = Ce−x /4α = C 2e−x /2α . (175)
for the particular form of P (x) above due to the fact that P (x) is even
in x → −x, whereas as x changes sign under x → −x.
We now come to the “standard” definition of ∆x:
2 2
(∆x)2 = h(x − hxi)2i = hx2i − 2hxhxii + hxi = hx2i − hxi . (177)
dN
P (t) ≡ = N0e−t/τ , (179)
dt
J. Gunion 9D, Spring Quarter 119
where by convention τ is referred to as the particle lifetime.
Of course, we must remember that P (t) is notpthe amplitude, but rather
the probability. The amplitude will be f (t) ∝ P (t) ∝ e−t/2τ .
If we are talking about solutions to the SE, it must be that this time
structure is a superposition of plane-wave type solutions (these are all we
have — they form a complete set). In the case of a time structure, this
means that there must exist a fe(ω) such that
Z ∞
f (t) = dω fe(ω)e−iωt . (180)
−∞
∞
1
Z
fe(ω) = dteiωtf (t) (181)
2π −∞
as we easily show.
Proof
∞ Z ∞
1
Z
0
f (t) =? dω dt0eiωt f (t0) e−iωt
2π −∞ −∞
Z ∞ Z W
1 0 0 iω (t0 −t)
= dt f (t ) lim dωe
2π −∞ W →∞ −W
Z ∞ iW (t0 −t) −iW (t0 −t)
1 e − e
= dt0f (t0) lim
2π Z−∞ W →∞ i(t0 − t)
∞
1 0 0 2 sin[W (t0 − t)]
= dt f (t ) lim
2π −∞ W →∞ t0 − t
= f (t) . (182)
The last step takes a bit of work. Note that if t0 6= t then the
sin[W (t0 − t)] is very rapidly oscillating as a function of t0 (for
R very large
W ) and in the large W limit nothing will survive in the dt0 integral
for any tiny dt0 interval where t0 6= t. Assuming that f (t) is a smoothly
∞
1 2 sin[W (t0 − t)]
Z
dt0f (t0) lim 0−t
2π −∞ W →∞
Z ∞ t
0
1 0 2 sin[W (t − t)]
= f (t) lim dt
2π W →∞ −∞ t0 − t
= f (t) , (183)
where for the last step we have simply looked up the integral in a table
and found that its value is 2π independent of t0 − t and W .
QED
So, now let us return to the computation of interest:
∞
1
Z
fe(ω) = dteiωtf (t)
2π Z−∞
∞
1
∝ dteiωte−t/2τ
2π Z0
∞
1
= dtet(iω−1/2τ )
2π 0
1
P (ω) = |fe(ω)|2 ∝ 1 . (185)
ω2 + 4τ 2
p2
E= +U. (187)
2m0
which is to say the time and space dependence can be separated. (In
the U = 0 case, ψ(x) = eikx and φ(t) = e−iωt.) Substituting this form
into eq. (188) and dividing by ψφ, we obtain (I will drop the subscript 0
on m0, and simply write m in what follows.)
The lhs depends only on x and the rhs depends only on t. This is only
possible if both sides are equal to the same constant, which we call E
dφ(t) −i E
ih̄ = Eφ(t) : ⇒ φ(t) = e h̄ t = e−iωt . (191)
dt
h̄2 d2ψ(x)
− + U (x)ψ(x) = Eψ(x) , (192)
2m dx2
d2ψ 2m
= 2 [U (x) − E]ψ(x) . (193)
dx2 h̄
2 2
P (x, t) = |Ψ(x, t)| = |ψ(x)| (194)
Particle in a box
0, 0≤x≤L
U (x) = . (195)
∞ , x < 0 or x > L
d2ψ 2mE
=− 2 ψ(x) = −k2ψ(x) . (196)
dx2 h̄
5. At x = 0, ψ(0) = 0 ⇒ B = 0.
At x = L, with B = 0, and assuming A 6= 0, ψ(L) = 0 ⇒ kL = nπ.
This latter condition can be written in what should now be a not-
unexpected manner. Namely:
nπ 2π nπ n
kL = nπ ⇒ k= ⇒ = → λ = L,
L λ L 2
(198)
which is to say that we must fit a half-integer number of wave-lengths
in between the box edges.
RL 2
p
To normalize, we require 0
|ψn(x)| dx = 1 which gives A = 2/L.
Example 1
Could we ever expect to see any of this stuff for a macroscopic object?
Answer: no!
For example, consider a 1.00 mg object confined to move between two
rigid walls separated by 1.00 cm. (a) Calculate the minimum speed of
the object. (b) If the speed of the object is (an observable amount of)
3.00 cm/s, find the corresponding, value of n.
Solution:
(a) Treating this as a particle in a box, the energy of the particle can
The minimum energy results from taking n = 1. For the above m and
L, we calculate
q
v1 = 2(5.49 × 10−58 J )/(1.00 × 10−6 kg) = 3.31 × 10−26 m/s ,
(204)
an immeasurably small speed. The object would take 3 × 1023 s, or about
1 million times the present age of the universe, to move the 1.00 cm
between the walls.
1 1
E=K= 2
mv = (1.00×10−6 kg)(3.00×10−2 m/s)2 = 4.50×10−10 J .
2 2
(205)
This too, to be allowed, would have to be one of the En values. To
determine which one, we solve for n, and obtain
√
8mL2E q
n= = (8.00 × 10−10 kg · m2)(4.50 × 10−10 J ) = 9.05×1023 .
h
(206)
This is an enormous number. Indeed, the value of n is so large that we
would never be able to distinguish the quantized nature of the energy
levels. The difference between the energies for n = 9.05 × 1023 and
n0 = 9.05 × 1023 + 1 is only about 10−33 J , much too small to be
detected experimentally.
This is an example that illustrates Bohr’s correspondence principle, which
asserts that quantum predictions must agree with classical results for
large masses and lengths.
Example 2
(b) In this state, what is the probability that the e− would be found
within 10 cm of the left-hand wall?
2
We have P (x) = |ψ(x)| . For the general case,
r !2
2 2 nπx
|ψn(x)| = sin . (208)
L L
So,
L/3
L 2 nπx
Z
2
P 0≤x≤ = sin dx
3 L L
Z0 L/3
2 1 2nπx
= 1 − cos dx
0 L 2 L
which yields n = 4.89 × 108. Then, using the general result above, we
We have also seen that the wave functions and associated probabilities
can be used to compute expectation values, such as that we considered
when computing (∆x)2 ≡ h(x − hxi)2i. In general, any such average is
an average of some dynamical property that can be measured, such as
position, momentum, energy, . . . . We always define
Z
hQi = Ψ∗(x, t)QΨ(x,
b t)dx. (214)
all x
The symbol Q b stands for the operator associated with the observable Q;
for each observable there is a unique operator. For position it is simply
x. But, in many cases Q b is a differential operator. Thus, its location
above is absolutely critical.
h̄ ∂ b = ih̄ ∂ .
pb = , E (215)
i ∂x ∂t
J. Gunion 9D, Spring Quarter 141
We will not go through it here, but it can be shown that (dropping the
time dependent part of the wave function)
h̄ ∂
Z Z
dxψ ∗(x) ψ(x) = dkψ(k)h̄k
e ψ(k)
e . (216)
i ∂x
dhxi d
Z
hpi = m =m Ψ∗(x, t)xΨ(x, t)dx (217)
dt dt
dΨ dΨ∗ ∂ 2Ψ ∂ 2 Ψ∗
then use the SE to evaluate and
dt dt
in terms of ∂x2
and ∂x2
, and
then do some partial integrations.
p2
Let us use pb to compute hpi and h 2m i
for the n = 1 ground state of the
infinite well.
h̄ ∂
Z
∗
hpi = ψ (x) ψ(x)dx
i ∂x
J. Gunion 9D, Spring Quarter 142
r ! r !
L
2 πx h̄ ∂ 2 πx
Z
= sin sin dx
0 L L i ∂x L L
r ! r !
L
2 πx h̄ π 2 πx
Z
= sin cos dx
0 L i LL L L
Z L
h̄ 2π πx
πx
= 2
sin cos dx = 0 (218)
iL 0 L L
which is the answer we expected since the particle is not going anywhere.
Classically, it is just bouncing back and forth. Quantum mechanically,
the particle is a stationary wave form in a fixed box.
Now let us compute
∂2
Z
2 ∗ 2
hp i = ψ (x)(−h̄ ) 2 ψ(x)
∂x
! !
Z L r 2
r
2 πx −π 2 πx
= sin (−h̄2) 2 sin dx
0 L L L L L
!2
2 2Z L
r
π h̄ 2 πx
= 2
sin
L 0 L L
where the 1 just comes from the fact that the remaining integral is just
the total probability integral.
First, let us note that this result implies that
p2 h̄2π 2
hEi = h i= = E1 (220)
2m 2mL2
as we expect.
Next, we note that since hpi = 0, we have
πh̄
q
2 2
∆p = hp i − hpi = . (221)
L
πh̄
∆x∆p = 0.181L = 0.568h̄ , (223)
L
which is pretty close to the smallest (Gaussian wave function) result. Of
course, as n increases, ∆x∆p increases, with
L
lim ∆x = √ . (224)
n→∞ 12
This latter limit is the same as the classical limit, in which we would
compute hx2i using the uniform probability P = L1
L
1 1
Z
2
hx i = x2 dx = L2 ,
Z0 L L 3
1 L
hxi = x dx =
0 L r 2
1 1 L
⇒ ∆x = L − =√ . (225)
3 4 12
" r #
h̄ ∂ h̄ nπ 2 nπx
pbpψ
b n(x) = cos
i ∂xi L L L
2 r
nπ 2 nπx
= (−h̄2) sin
L l L
b 2i =
b then hQ
Note the general result that if Ψ is an eigenstate of Q,
2
2
q = hQi implying that
b
2 2 2
(∆Q) = hQ i − hQi = 0 .
b b (229)
Similar manipulations apply for the energy operator. For stationary states,
energy is always an eigenvalue. Our time dependence is Ψ(x, t) =
ψ(x)φ(t) = ψ(x)e−iωt for which
∂
EΨ(x, t) = ih̄ Ψ(x, t) = h̄ωΨ(x, t) ,
b (230)
∂t
in agreement with the standard energy-frequency relationship. In
contrast, pb 2 does not always give a definite eigenvalue when operating
on a stationary state. It just happens to for this infinite well case, where
U = 0 is a constant inside the well.
If a wave function, Ψ is such that both the momentum and the position
are well defined, in operator language this means that Ψ should be an
eigenstate of both the momentum and the position operators:
h̄ ∂Ψ
pΨ
b = = pΨ , x
bΨ = xΨ . (231)
i ∂x
In this case, it should not matter in which order we operate the momentum
and position operators on the wave function. That is, we would for
consistency have
h̄ ∂ h̄ ∂
pb
bxΨ = x
bpΨ
b , ⇒ [xΨ] = x Ψ ,. (232)
i ∂x i ∂x
∂
[E, t] = ih̄ , t = ih̄
b b (235)
∂t
implying that energy and time cannot be both precisely known, since
a wave function Ψ(x, t) cannot simultaneously have precise E
b and tb
eigenvalues.
d2ψ 2m
= 2 (U − E)ψ (237)
dx2 h̄
The form e−αx is also a solution of the equation, but becomes infinite
if x → −∞. This choice would not allow us to normalize the wave
function — the net probability integral would diverge. Thus, we must
choose only the e+αx form for x < 0.
2. In the region x > L, we have exactly the same SE, but now we must
only have the e−αx solution which decays exponentially for x → +∞.
which we rewrite as
α2
2α
tan kL 1 − = . (241)
k2 k
J. Gunion 9D, Spring Quarter 152
For specified U and L, this equation can only be solved for very special
values of E (which is contained in both α and k).
Given a value of E that solves eq. (241), let us examine the behavior of
ψ. The crucial point is that ψ “leaks” into the x < 0 and x > L regions
where classically the particle cannot go. The distance, δ, from one of the
side boundaries of the well at which ψ declines to 1/e of its value at the
boundary is called the penetration depth. That is, δ is defined by
Example
An electron is in a potential well with L = 0.200 nm and U = 100 eV .
Find the possible values of E for which the electron is bound to the well.
We are looking for values of E < U that solve eq. (241). The left and
right hand sides of eq. (241) are plotted below.
20 40 60 80 100
-2.5
-5
-7.5
-10
Figure 25: Plot of left (red) and right (blue) sides of eq. (241). The
horizontal axis is E in eV .
h̄ (197.3 eV · nm/c)
δ = p =p
2m(U − E) 2(511 × 103 eV /c2)(100 eV − 93.833 eV )
= 0.07859 nm , (244)
d2ψ 2m
= 2 (U − E)ψ , (245)
dx2 h̄
with E > U , the solutions are now of the form eilx or e−ilx (it will
prove more convenient to use these exponential forms than the equivalent
A sin lx and B cos lx forms), with l2 = 2h̄m
2 (E − U ). Similar results apply
to x > L. In both regions we have oscillatory behavior and there is no
damping for |x| → ∞. We will have more to say about related situations
in the next chapter.
d2U
1
U (x) = U (a) + k(x − a)2 , where k= (246)
2 dx2 x=a
Let us try
2
ψ(x) = C0e−αx . (248)
For this form
d2ψ d h −αx2
i 2
= C0(−2αx)e = C0[−2α + 4α2x2]e−αx (249)
dx2 dx
which has the same structure as the other side of the SE equation,
2m 1 2
2 mω 2x2 − E C0e−αx , (250)
h̄ 2
2m 1 mω
4α = 2 ( mω 2) ,
2
or α=
h̄ 2 2h̄
2mE mω 1
2 = 2α = , or E = h̄ω . (251)
h̄ h̄ 2
In short, we have
1 − mωx
2
E0 = h̄ω , ψ0(x) = C0e 2h̄ , (252)
2
The most important point about the ground state is that E0 6= 0. That
is, there is a minimum energy for the QM oscillator state. This energy
is sometimes called the zero-point energy. It has important implications.
It means for example that a crystal lattice of ions, each of which can be
2 h̄2 2 2 h̄
|xmax| ∼ <
∼A , ⇒ A >
∼ mω , (254)
m2 ω 2 A2
1
En = n+ h̄ω . (256)
2
The important aspect of this prediction is the equal spacing of the energy
levels. For this system, if the state is in a n 6= 0 state, when it decays to
its next lowest state, the frequency of the radiation will always be given
by hf = h̄ω, no matter what the starting n value.
J. Gunion Fig.
9D,7−1b, p.232
Spring Quarter 166
Imagine a particle incident on the barrier from the left. Classically, if
E > U , then the particle will temporarily have reduced velocity as it
passes the barrier (there is a force to slow it down at the left-hand edge
of the barrier), but it will regain its initial velocity once it has passed the
barrier (there is a force to speed it up at the far right-hand edge of the
barrier).
If E < U , the particle cannot penetrate the barrier classically and the
particle will simply bounce off the barrier and reverse direction.
In QM, all regions are accessible to the particle even if E < U . The QM
penetration of the barrier is called tunneling.
To explore this quantitatively, we must set up the situation and impose
appropriate boundary condition requirements at the edges of the barrier.
We will imagine:
1. A wave is coming in from the left.
2. It hits the barrier.
3. Some of the wave is reflected, but some penetrates into the barrier.
4. The penetrating wave can make it over to the right-hand side of the
barrier with some reduced amplitude.
5. There is thus some wave moving to the right in the region past the
The above set-up is realized using the plane wave solutions of the SE as
To the right side of the barrier, the physical setup envisioned means that
we should allow only the wave moving to the right:
(Ψ∗Ψ)transmitted F ∗F |F |2
T = = = . (260)
(Ψ∗Ψ) incident A∗ A |A|2
When the wave is in the barrier region itself, the SE takes a different
form.
d2ψ 2m(U − E) 2
2
= 2 ψ(x) ≡ α ψ(x) , (262)
dx h̄
where we are considering the situation with U > E so that the α2 on
the rhs above is positive (as opposed to the negative −2h̄mE2 that applies
outside the barrier region). The solutions of this equation are exponential
decay or increase. We should allow for both by writing
∂Ψ
We employ continuity of Ψ and of ∂x
at the right and left sides of the
Example
Two wires are separated by an insulating layer. Modeling the latter as a
◦
For L = 50 A (i.e. 5 nm), we get T = 0.963 × 10−38, a very small
number.
◦
For L = 10 A, we get T = 0.657 × 10−7. Changing the barrier thickness
by just a factor of 5 has a huge effect.
Now suppose that we have the above situation and that a 1 mA current
of electrons is incident on the insulating barrier layer. How much of
Answer: Each of the e−’s in the current has the probability given by
T = 0.657 × 10−7 to pass through the insulating layer. The cumulative
effect will be a transmitted current of
This is just the right amount of kinetic energy to get the particle over
the U − E barrier; if δ = ∆x were any bigger then ⇒ not enough K to
get over barrier.
2ik 2
C= = α . (272)
ik − α 1− ik
α
1 +1
C 0 + D0 = (−αC 0 + αD 0) , ⇒ D0 = C 0 ik
α
ik ik
−1
Armed with this result, we can easily compute (see 3rd b.c. equation)
4α 2 2
0 0 2 −2αL 4kδ
2
|F | = |C + D | = k
e = e−2L/δ . (274)
α 2 2 1 + (kδ)2
1+ k2
Of course, the thing that is a bit tricky for you is all the use of absolute
squares of complex numbers. I go through this example so that you can
see that this complex stuff is essential for actual computations in QM.
In any case, you can now see that my earlier approximation of keeping
only the exponential factor is only a rough approximation. However, for
large L/δ, and (kδ)2 = E/(U − E) ∼ O(1), dropping the multiplicative
factor is a small perturbation on the very small exponential, thereby
justifying the use of
√
T (E) ∼ e−2L/δ = e−2L 2m(U −E )/h̄
. (275)
However, the real view is that the e− wave function actually penetrates
for a distance of order δ into the region outside the metal.
h̄
The e− has K = E inside the metal, so δ = √ .
2m(U −E )
The STM uses a probe to look at the exponential decay in this region
outside the metal.
exp[−x/δ ]
probe
e2V
i= e−2L/δ , (277)
4π 2Lδh̄
and if L changes by even a small amount (i.e. if there is variation in the
surface defined by the e− wave functions), we will see a change in i.
◦
In the above example, if L → L + 0.01A then i changes by a fractional
◦
−0.01×2/1 A
amount of e ' 0.98. Such a 2% change in i is measurable
using appropriate amplification techniques. Thus, one can have sensitivity
◦
to surface details at the 0.01A level.
In most physical cases, U is not actually a constant, but rather has some
non-trivial x dependence. We will do the example of α decay shortly.
The generalization of our approximate formula for T (E) that applies in
this case is
" √ #
2 2m
Z p
T (E) ∼ exp − U (x) − E dx . (278)
h̄ barrier
This reduces to our previous expression if the barrier region has length L
and U (x) = U is a constant in that region.
α decay simplified
Many nuclei heavier than lead naturally emit an α particle, but emission
rates vary by factors of 1013, whereas the energies of the α’s range only
from 4 to 8 M eV . Why?
2kZe2
Vcoulomb(R) =
R
2(90)(1.6 × 10−19 C)2(9 × 109 N · m2/C 2)
=
7 × 10−15 m
10−6 M eV
×
1.6 × 10−19 J
= 37 M eV . (279)
37 M eV
average potential height = ∼ 18 M eV ,
2
R1 − R (62 − 7) f m
average barrier width = = ∼ 28 f m .
2 2
(281)
Then,
1
T (E) = h i
U2
1 + 14 E (U −E ) sinh2 αL
4.2 M eV 4.2 M eV −14
∼ 16 1− exp −2(2.8 × 10 m)/δ ,
18 M eV 18 M eV
(282)
where
h̄ 6.58 × 10−22 M eV · s
δ = p =p
2mα(V − E) 2(3727 M eV /c2)(18 − 4.2) M eV
This gives
where we should note that the output v implies that the NR approximation
is ok.
T (E)τ 10−21 s
= 1, ⇒ τ = ∼ 1020 s . (289)
tcrossing 10−41
The actual 238U lifetime is τ = 4.5 × 109 yr ∼ 1017 s. Thus, our first
estimate, while very crude, is in the right ballpark.
" √ #
2 2m
Z p
T (E) ∼ exp − U (x) − E dx . (290)
h̄ barrier
r s
E0 ZR
T (E) = exp −4πZ +8 , (291)
E r0
h̄2
where r0 = is a kind of a Z = 1 Bohr radius for the α particle:
mα ke2
r0 ∼ a0/7295 ∼ 7.25 f m since mα ∼ 7295me, and the associated
Z = 1 kinetic energy is
ke2 ke2 a0
E0 = = = 13.6 eV × 7295 = 0.0993 M eV . (292)
2r0 2a0 r0
ln 2
τhalf life ≡ τ1/2 = (293)
fcrossing T (E)
Example
As an example of using the precise formula, consider T h which ejects
an α particle with energy E = 4.05 M eV . The nuclear radius is
R = 9.00 f m.
The daughter atomic number is Z = 88. We compute
" r r #
0.0993 9.00
T (E) = exp −4π(88) + 8 88
4.05 7.25
= exp[−89.542] = 1.3 × 10−39 . (294)
and
0.693 0.693
τ1/2 = = = 5.4 × 1017 s = 1.7 × 1010 yr .
fcrossing T (E) 1.3 × 10−18
(296)
which compares favorably with the actual value of 1.4 × 1010 yr.