Download The ingredients of an atom

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Tight binding wikipedia , lookup

Transcript
Chapter 1
The ingredients of an atom
1.1
Introduction
Atomic Physics is the product of our attempts to understand the visible world. What we see is determined
by how light interacts with matter. What we understand about the nature of matter has been learned
largely from the study of atomic spectra. Optics has therefore played a key role in the development of
atomic physics and quantum theory which was developed in response to certain puzzles presented by
atomic spectra. Atomic physics remains one of the most important testing grounds for quantum theory
and its derivative: QED, quantum electrodynamics. It is therefore still an area of active research, both
for its contribution to fundamental physics and to technology. Many branches of science rely heavily on
atomic physics. The following list gives a few examples:
Physics Astrophysics, plasma physics, atmospheric physics, solid state physics, chemical physics and
radiation physics.
Other science Chemistry (analysis, reaction rates), biology (molecular structure, physiology), materials
science, energy research, fusion studies.
Applications Lasers, X-ray technology, NMR, pollution detection, medical applications of devices
(lasers, NMR etc.).
In this chapter we will give a brief overview of the main physical ingredients of an atom. These
ingredients will be revisited in more detail in subsequent chapters, but at this stage we just want to
estimate the relative size of the main effects, making use of semi-classical ideas based on the Bohr model.
1.2
Atomic gross structure
The most important quantity in an atom is the binding energy of the electron due to the Coulomb
attraction to the nucleus. This determines the gross structure of the emission spectra obtained from
the atom. As a first guess, we can just use the quantized energies derived from the Bohr model. This
tells us that the Coulomb energies are of order 1 – 10 eV, or 104 − 105 cm−1 in wave number units. (See
section 1.9 for a discussion of atomic energy units.)
1.2.1
The Bohr model
In 1911 Rutherford discovered the nucleus. This then led to the idea of atoms consisting of electrons in
classical orbits in which the centrifugal forces are balanced by the Coulomb attraction to the positive
nucleus, as shown in figure 1.1. The problem with this idea is that the electron in the orbit is constantly
accelerating. Accelerating charges emit radiation called bremsstrahlung, and so the electrons should
be radiating all the time. This would reduce the energy of the electron, and so it would gradually spiral
into the nucleus, like an old satellite crashing to the earth.
In 1913 Bohr produced his model for the atom. The key new elements of the model are:
• The angular momentum L of the electron is quantized in units of h̄ (h̄ = h/2π):
L = nh̄ ,
where n is an integer.
1
(1.1)
2
CHAPTER 1. THE INGREDIENTS OF AN ATOM
-e
v
F
r
+Ze
Figure 1.1: The Bohr model of the atom considers the electrons to be in orbit around the nucleus. The
central force is provided by the Coulomb attraction. The angular momentum of the electron is quantized
in integer units of h̄.
• The atomic orbits are stable, and light is only emitted or absorbed when the electron jumps from
one orbit to another.
When Bohr made these hypotheses in 1913, there was no justification for them other than they were
spectacularly successful in predicting the energy spectrum of hydrogen. With the hindsight of quantum
mechanics, we now know why they work. The first assumption is equivalent to stating that the orbit
must correspond to a fixed number of de Broglie wavelengths. For a circular orbit, this can be written:
2πr = integer × λdeB = n ×
h
h
=n×
,
p
mv
(1.2)
which can be rearranged to give
h
.
(1.3)
2π
The second assumption is a consequence of the fact that the Schrödinger equation leads to time-independent
solutions (eigenstates).
The derivation of the quantized energy levels proceeds as follows. Consider an electron orbiting a
nucleus of mass mN and charge +Ze. The centrifugal force is balanced by the Coulomb force:
L ≡ mvr = n ×
mv 2
Ze2
=
r
4π²0 r2
F =
(1.4)
As with all two-body orbit systems, the mass m that enters here is the reduced mass:
1
1
1
=
+
.
m
me
mN
The energy is given by:
(1.5)
1
En
=
=
=
kinetic energy + potential energy
Ze2
2
1
mv
−
2
4π²0 r
mZ 2 e4
− 2 2 2,
8²0 h n
(1.6)
where we made use of eqns 1.3 and 1.4 in the last line. This can be written in the form:
R0
n2
(1.7)
¶
m 2
Z R∞ ,
me
(1.8)
m e e4
.
8²20 h2
(1.9)
En = −
where R0 is given by:
µ
R0 =
and R∞ is the Rydberg constant:
R∞ =
1 In atoms the electron moves in free space, where the relative dielectric constant ² is equal to unity. However, in solid
r
state physics we frequently encounter hydrogenic systems inside crystals where ²r is not equal to 1. In this case, we must
replace ²0 by ²r ²0 throughout.
1.2. ATOMIC GROSS STRUCTURE
Quantity
3
Symbol
Formula
Numerical Value
R∞
me e4 /8²20 h2
2.18 × 10−18 J
13.6 eV
109,737 cm−1
Bohr radius
a0
²0 h2 /πe2 me
5.29 × 10−11 m
Fine structure constant
α
e2 /2²0 hc
1
137.04
Rydberg constant
Table 1.1: Fundamental constants that arise from the Bohr model of the atom.
The Rydberg constant is a fundamental constant and has a value of 2.18 × 10−18 J, which is equivalent
to 13.6 eV or 109,737 cm−1 .
R0 is the effective Rydberg constant for system in question. In the hydrogen atom we have an electron
orbiting around a proton of mass mp . The reduced mass is therefore given by
m = me ×
mp
= 0.999 me
me + mp
and the effective Rydberg for hydrogen is:
RH = 0.999R∞ .
(1.10)
Atomic spectroscopy is very precise, and 0.1% factors such as this are easily measurable. Furthermore,
in other systems such as positronium (an electron orbiting around a positron), the reduced mass effect is
much larger, because m = me /2.
By following through the mathematics, we also find that the orbital radius and velocity are quantized.
The relevant results are:
n2 me
rn =
a0 ,
(1.11)
Z m
and
Z
vn = α c .
(1.12)
n
The two fundamental constants that appear here are the Bohr radius a0 :
a0 =
h2 ²0
,
πme e2
(1.13)
α=
e2
.
2²0 hc
(1.14)
and the fine structure constant α:
The fundamental constants arising from the Bohr model are related to each other according to:
a0 =
and
R∞ =
h̄ 1
,
me c α
(1.15)
h̄2 1
.
2me a20
(1.16)
The definitions and values of these quantities are given in Table 1.1.
1.2.2
The need for quantum mechanics
A simple back-of-the-envelope calculation can easily show us that the Bohr model is not fully consistent
with quantum mechanics. In the Bohr model, the linear momentum of the electron is given by:
¶
µ
nh̄
αZ
mc =
.
(1.17)
p = mv =
n
rn
4
CHAPTER 1. THE INGREDIENTS OF AN ATOM
E2
E2
hn
E1
hn
E1
absorption
emission
Figure 1.2: Absorption and emission transitions.
However, we know from the Heisenberg uncertainty principle that the precise value of the momentum
must be uncertain. If we say that the uncertainty in the position of the electron is about equal to the
radius of the orbit rn , we find:
h̄
h̄
∆p ∼
≈
.
(1.18)
∆x
rn
On comparing Eqs. 1.17 and 1.18 we see that
|p| ≈ n∆p .
(1.19)
This shows us that the magnitude of p is undefined except when n is large. This is hardly surprising.
The Bohr model is a mixture of classical and quantum models, and we can only expect the arguments to
be fully self-consistent when we approach the classical limit at large n. For small values of n, the Bohr
model fails when we take the full quantum nature of the electron into account.
The integer n that arises from the Bohr model is known as the principal quantum number. In
more accurate calculations of the energy based on the Schrödinger equation, we will find that the gross
energy of many-electron atoms depends on other quantum numbers in addition to n. The most important
of these is the orbital quantum number l, which can take integer values up to n − 1.
1.3
Absorption and emission spectra
A photon is absorbed or emitted when an electron jumps between two quantum states, as shown in
figure 1.2. These are called radiative transitions. The energy of the photon is equal to the difference
in energy of the two levels namely (E2 − E1 ):
hν = E2 − E1 .
where E1 and E2 are the energies of the lower and upper levels of the atom respectively.
In the case of the hydrogen atom we have from eqn 1.7:
µ
¶
1
1
hν = RH
−
,
n21
n22
(1.20)
(1.21)
where n1 and n2 are the principal quantum numbers of the two states involved. In absorption we start
from the ground state, so we put n1 = 1. In emission, we can have any combination where n1 < n2 .
Some of the series of spectral lines have been given special names. The emission lines with n1 = 1 are
called the Lyman series, those with n1 = 2 are called the Balmer series, etc.
1.4
Angular momentum and magnetic moments
The quantization of the angular momentum was a fundamental postulate of the Bohr model (c.f. Eq. 1.1).
The full quantum mechanical treatment of the atom gives the following form for the magnitude of the
angular momentum L:
p
(1.22)
L = l(l + 1)h̄ ,
where l is the orbital quantum number. l can take integer values up to (n − 1). The component of the
angular momentum along a particular axis (usually taken as the z axis) is quantized in units of h̄ and its
value is given by the magnetic quantum number ml :
Lz = ml h̄ .
(1.23)
1.4. ANGULAR MOMENTUM AND MAGNETIC MOMENTS
5
z
| L | = l (l + 1) h
L z = mlh
x,y
Figure 1.3: Vector model
p of the angular momentum in an atom. The angular momentum is represented
by a vector of length l(l + 1)h̄ precessing around the z-axis so that the z-component is equal to ml h̄.
m
I
+Ze
r
v
-e
Figure 1.4: The orbital motion of the electron around the nucleus is equivalent to a current loop, which
generates a magnetic dipole moment.
ml can take integer values from −l to +l. We will see where these relationships come from in subsequent
chapters. The quantization conditions given in equations 1.22 and 1.23 give rise to the vector model of
angular momentum illustrated in figure 1.3.
The orbital motion of the electron causes it to have a magnetic moment. An electron in a circular
Bohr orbit is equivalent to a current loop, as illustrated in figure 1.4. We know from electromagnetism
that current loops behave like magnets. The electron in the Bohr orbit is thus equivalent to a little
magnet with a magnetic dipole moment µ given by:
µ = I × Area = −(e/T ) × (πr2 )
(1.24)
where T is the period of the orbit. Now T = 2πr/v, and so we obtain
µ=−
ev
e
e
πr2 = −
me vr = −
L,
2πr
2me
2me
(1.25)
where we have substituted L for the orbital angular momentum mvr. The same relationship can be
derived more rigorously for non-circular orbits.
Equation 1.25 shows us that the orbital angular momentum is directly related to atomic magnetism.
The quantity e/2me that appears is called the gyromagnetic ratio. It specifies the proportionality
constant between the angular momentum of an electron and its magnetic moment.
We can see from Eqs. 1.1 or 1.22 that |L| ∼ h̄. Hence the magnitude of the atomic magnetic dipoles
is given by
|µ| ∼ µB ,
(1.26)
where µB is the Bohr magneton defined by:
µB =
eh̄
= 9.27 × 10−24 JT−1 .
2me
(1.27)
6
CHAPTER 1. THE INGREDIENTS OF AN ATOM
non-uniform
magnetic field
ms
atom
beam
+½
l=0
-½
Figure 1.5: The Stern-Gerlach experiment. A beam of atoms with l = 0 is deflected in two discrete ways
by a non-uniform magnetic field. The force on the atoms arises from the interaction between the field
and the magnetic moment due to the electron spin.
1.5
Spin
Spin is a completely quantum mechanical property with no classical equivalent. The Stern-Gerlach
experiment (figure 1.5) showed that electrons with l = 0 still possess angular momentum, even though
Eq. 1.22 shows us that the orbital angular momentum L is zero if l = 0. For want of a better name, this
angular momentum was called “spin”. Paul Dirac at Cambridge successfully accounted for electron spin
when he produced the relativistic wave equation that bears his name in 1928.
The magnitude of the spin angular momentum is given by
p
S = s(s + 1)h̄
(1.28)
and the component along the z axis is given by
Sz = ms h̄ .
(1.29)
The spin quantum number s is equal to 12 and ms increases in integer steps between −s and +s, ie
ms = ± 12 . 2
The deflections measured in the Stern-Gerlach experiment enabled the magnitude of the magnetic
moment due to the spin angular momentum to be determined. The component along the z axis was
found to obey:
µz = −gs µB ms ,
(1.30)
where gs is the g-value of the electron. The Dirac equation predicted that gs should be exactly equal
to 2. Modern calculations by quantum electrodynamics (QED) give a value of 2.0023192 for gs , which
agrees very accurately with recent experimental data.
1.6
Spin-orbit coupling and fine structure
Spin-orbit coupling is a relativistic effect that arises from the interaction between the spin magnetic
moment of the electron and the magnetic field produced by the orbital motion. We will treat this effect
more rigorously in subsequent chapters. At this stage, we can give a hand-waving argument to predict
the magnitude of the spin-orbit coupling.
We shift the origin from the nucleus to the electron. In this frame, the electron is stationary and the
nucleus is moving in a circular orbit of radius rn . The orbit of the nucleus is equivalent to a current
loop, which produces a magnetic field at the origin. Now the magnetic field produced by a circular loop
of radius r carrying a current I is given by:
Bz =
µ0 I
2r
(1.31)
where z is taken to be the direction perpendicular to the loop. As before, the current I is given by the
charge Ze divided by the orbital period T = 2πr/v. Hence for a Bohr atom we find
µ 4¶
µ0 Zevn
µ0 αce
Z
Bz =
=
.
(1.32)
4πrn2
n5
4πa20
2 The fact that the electron has a half-integer spin makes it a fermion. Fermions obey the Pauli exclusion principle.
Particles with integer values of the spin (eg α-particles) do not obey the Pauli exclusion principle. The Pauli exclusion
principle is very important for the explanation of the periodic table of elements, as we will see in subsequent chapters.
1.7. EXTERNAL MAGNETIC FIELDS
7
For hydrogen with Z = n = 1, we find Bz ≈ 12 Tesla. This is a large field.
The interaction energy is given by
E = −µ · B = −µz Bz = gs µB ms Bz = ±µB Bz ,
(1.33)
where we have used gs = 2 in the last equality. By substituting from Eq. 1.32, we obtain the spin-orbit
interaction energy:
Z2
|Eso | = α2 3 En .
(1.34)
n
For the n = 1 orbit of hydrogen, this gives:
|Eso | = α2 RH = 13.6 eV/1372 = 0.7 meV ≡ 6 cm−1 .
• The small energy shifts due to the spin-orbit interaction give rise to the fine structure in the
absorption and emission spectra of atoms.
• The relative size of the spin-orbit interaction grows as Z 2 . Spin-orbit effects are therefore more
important in heavy atoms.
• We expect the magnitude of relativistic corrections to the Bohr model to be of order (v/c)2 En . We
see from Eq. 1.12 that this is basically the same as Eq. 1.34. This is hardly surprising, because
spin-orbit coupling is in fact a relativistic effect.
1.7
External magnetic fields
The splitting of atomic spectral lines when an external magnetic field is applied is called the Zeeman
effect. The external magnetic field interacts with the magnetic moment of the atom, and shifts the
transition lines by −µ · B. If we neglect spin, and we take the field to be applied along the z axis, the
shift is given by:
E = −µz Bz = µB ml Bz .
(1.35)
The observed splittings are of the right magnitude, but only follow Eq. 1.35 exactly if we neglect electron
spin. We will look at both the “normal” and “anomalous” Zeeman effects in detail in subsequent chapters.
Note that we must apply very large fields > 10 Tesla before the strength of the interaction with the external
field exceeds the spin-orbit interaction.
1.8
Nuclear effects
For most of the time in atomic physics we just take the nucleus to be a heavy charged particle sitting
at the centre of the atom. However, two small effects due to the nucleus are readily observable in the
atomic spectra.
1. Isotope effects: the nuclear mass affects the gross structure slightly. The mass m that enters the
Bohr formulae is the reduced mass, not the bare electron mass me (c.f. eqn 1.5). Changes in the
nuclear mass make small changes to m and hence to the atomic energies. The heavy isotope of
hydrogen, namely deuterium, was discovered this way.
2. Hyperfine structure: The nucleus possess a small magnetic moment of magnitude ∼ µB /2000
due to the nuclear spin. This can interact with the magnetic field due to the orbital motion of the
electron just as in spin-orbit coupling. This gives rise to “hyperfine” shifts in the atomic energies
that are about 2000 times smaller than the fine structure shifts. The well-known 21 cm line of radio
astronomy is caused by transitions between the hyperfine levels of atomic hydrogen. The photon
energy in this case is 6 × 10−6 eV, or 0.05 cm−1 .
1.9
Energy units
Atomic energies are frequently quoted in electron volts (eV). 1 eV is the energy acquired by an electron
when it is accelerated by a voltage of 1 Volt. Thus 1 eV = 1.6 × 10−19 J. This is a convenient unit,
because the energies of the electrons in atoms are typically a few eV.
8
CHAPTER 1. THE INGREDIENTS OF AN ATOM
Atomic energies are also often expressed in wave number units (cm−1 ). The wave number ν is the
reciprocal of the wavelength of the photon with energy E. It is defined as follows:
ν=
1
ν
E
= =
.
λ (in cm)
c
hc
Note that the wavelength should be worked out in cm. Thus 1 eV = (e/hc) cm−1 = 8066 cm−1 .
Reading
Haken, H. and Wolf, H.C., The Physics of Atoms and Quanta, chapter 8.
Beisser, A., Concepts of Modern Physics, chapter 4.
French, A.P. and Taylor, E.F., An introduction to quantum physics, chapter 1.
Eisberg, R. and Resnick, R., Quantum Physics, chapter 4.
(1.36)