Chapter 7 Atkins Physical Chemistry (10th Edition) Peter Atkins and Julio de Paula Chapter 7 Atkins Physical Chemistry (10th Edition) Peter Atkins and Julio de Paula Chapter 7 Atkins Physical Chemistry (10th Edition) Peter Atkins and Julio de Paula Chapter 7 Atkins Physical Chemistry (10th Edition) Peter Atkins and Julio de Paula Chapter 7 Atkins Physical Chemistry (10th Edition) Peter Atkins and Julio de Paula
Trang 1chaPter 7
Introduction to quantum theory
It was once thought that the motion of atoms and subatomic
particles could be expressed using ‘classical mechanics’, the
laws of motion introduced in the seventeenth century by Isaac
Newton, for these laws were very successful at explaining the
motion of everyday objects and planets However, a proper
description of electrons, atoms, and molecules requires a
differ-ent kind of mechanics, ‘quantum mechanics’, which we
intro-duce in this chapter and then apply throughout the remainder
of the text
Experimental evidence accumulated towards the end of the
nineteenth century showed that classical mechanics failed
when it was applied to particles as small as electrons More
spe-cifically, careful measurements led to the conclusion that
par-ticles may not have an arbitrary energy and that the classical
concepts of particle and wave blend together In this Topic we
see how these observations set the stage for the development of
the concepts and equations of quantum mechanics through the
early twentieth century
In quantum mechanics, all the properties of a system are
expressed in terms of a wavefunction which is obtained by
solving the ‘Schrödinger equation’ In this Topic we see how to interpret wavefunctions
7C the principles of quantum theory
This Topic introduces some of the mathematical techniques
of quantum mechanics in terms of operators We also see that quantum theory introduces the ‘uncertainty principle’, one of the most profound departures from classical mechanics
What is the impact of this material?
In Impact I7.1 we highlight an application of quantum
mechan-ics that still requires much research before it becomes a useful technology It is based on the speculation that through ‘quan-tum computing’ calculations can be carried out on many states
of a system simultaneously, leading to a new generation of very fast computers
To read more about the impact of this material, scan the QR code, or go to bcs.whfreeman.com/webpub/chemistry/pchem10e/impact/pchem-7-1.html
Trang 27A the origins of quantum mechanics
The basic principles of classical mechanics are reviewed
in Foundations B In brief, they show that classical physics
(1) predicts a precise trajectory for particles, with precisely specified locations and momenta at each instant, and (2) allows the translational, rotational, and vibrational modes of motion
to be excited to any energy simply by controlling the forces that are applied These conclusions agree with everyday experience Everyday experience, however, does not extend to individ-ual atoms, and careful experiments have shown that classical mechanics fails when applied to the transfers of very small energies and to objects of very small mass
We also investigate the properties of light The classical view,
discussed in Foundations C, is of light as an oscillating
electro-magnetic field that spreads as a wave through empty space with
a wavelength, λ (lambda), a frequency, ν (nu), and a constant speed, c (Fig C.1) Again, a number of experimental results are
not consistent with this interpretation
This Topic describes the experiments that revealed tions of classical physics The remaining Topics of the Chapter show how a new picture of light and matter led to the formu-lation of an entirely new and hugely successful theory called
limita-quantum mechanics.
Here we outline three experiments conducted near the end of the nineteenth century and which drove scientists to the view that energy can be transferred only in discrete amounts
(a) Black-body radiation
A hot object emits electromagnetic radiation At high tures, an appreciable proportion of the radiation is in the visible region of the spectrum and a higher proportion of short-wave-length blue light is generated as the temperature is raised This behaviour is seen when a heated metal bar glowing red hot becomes white hot when heated further The dependence is illustrated in Fig 7A.1, which shows how the energy output varies with wavelength at several temperatures The curves are
tempera-those of an ideal emitter called a black body, which is an object
capable of emitting and absorbing all wavelengths of radiation uniformly A good approximation to a black body is a pinhole
in an empty container maintained at a constant temperature: any radiation leaking out of the hole has been absorbed and re-emitted inside so many times as it reflected around inside
Contents
(a) Black-body radiation 282
example 7a.1: using the Planck distribution 284
(b) Heat capacities 285
brief illustration 7a.1: the debye formula 286
(c) Atomic and molecular spectra 286
brief illustration 7a.2: the bohr frequency condition 287
(a) The particle character of electromagnetic radiation 287
example 7a.2: calculating the number of photons 288
example 7a.3: calculating the maximum
wavelength capable of photoejection 289
(b) The wave character of particles 289
example 7a.4: estimating the de broglie wavelength 290
➤
➤ Why do you need to know this material?
You should know how experimental results motivated
the development of quantum theory, which underlies
all descriptions of the structure of atoms and molecules
and pervades the whole of spectroscopy and chemistry
in general.
➤
➤ What is the key idea?
Experimental evidence led to the conclusions that energy
cannot be continuously varied and that the classical
concepts of a ‘particle’ and a ‘wave’ blend together when
applied to light, atoms, and molecules.
➤
➤ What do you need to know already?
You should be familiar with the basic principles of classical
mechanics, which are reviewed in Foundations B The
discussion of heat capacities of solids formally makes use
of material in Topic 2A but is introduced independently
here.
Trang 37A The origins of quantum mechanics 283
the container that it has come to thermal equilibrium with the
walls (Fig 7A.2)
The approach adopted by nineteenth-century scientists
to explain black-body radiation was to calculate the energy
density, dE, the total energy in a region of the
electromag-netic field divided by the volume of the region (units: joules
per metre-cubed, J m−3), due to all the oscillators
correspond-ing to wavelengths between λ and λ + dλ This energy density
is proportional to the width, dλ, of this range, and is written
where ρ (rho), the constant of proportionality between d E
and dλ, is called the density of states (units: joules per metre4,
J m−4) A high density of states at the wavelength λ and
tem-perature T simply means that there is a lot of energy associated
with wavelengths lying between λ and λ + dλ at that
tempera-ture The total energy density in a region is the integral over all
a region of volume V is this energy density multiplied by the
volume:
The physicist Lord Rayleigh thought of the electromagnetic field as a collection of oscillators of all possible frequencies He regarded the presence of radiation of frequency ν (and there-
fore of wavelength λ = c/ν, eqn C.3) as signifying that the
elec-tromagnetic oscillator of that frequency had been excited (Fig 7A.3) Rayleigh knew that according to the classical equipar-
tition principle (Foundations B), the average energy of each oscillator, regardless of its frequency, is kT On that basis, with
minor help from James Jeans, he arrived at the Rayleigh–Jeans
law for the density of states:
ρ λ( , )T =8πλ kT4 rayleigh–Jeans law (7A.4)
where k is Boltzmann’s constant (k = 1.381 × 10−23 J K−1)
Although the Rayleigh–Jeans law is quite successful at long wavelengths (low frequencies), it fails badly at short wave-
lengths (high frequencies) Thus, as λ decreases, ρ increases
without going through a maximum (Fig 7A.4) The equation therefore predicts that oscillators of very short wavelength (corresponding to ultraviolet radiation, X-rays, and even
γ-rays) are strongly excited even at room temperature The total
energy density in a region, the integral in eqn 7A.2, is also dicted to be infinite at all temperatures above zero This absurd result, which implies that a large amount of energy is radiated
pre-in the high-frequency region of the electromagnetic spectrum,
is called the ultraviolet catastrophe According to classical
Wavelength, λ
Figure 7A.1 The energy distribution in a black-body cavity at
several temperatures Note how the spectral density of states
increases in the region of shorter wavelength as the temperature
is raised, and how the peak shifts to shorter wavelengths
Detected radiation
Pinhole
Container
at a
temperature T
Figure 7A.2 An experimental representation of a black body
is a pinhole in an otherwise closed container The radiation
is reflected many times within the container and comes to
thermal equilibrium with the walls Radiation leaking out
through the pinhole is characteristic of the radiation within the
container
(a)
(b)
Figure 7A.3 The electromagnetic vacuum can be regarded
as able to support oscillations of the electromagnetic field When a high-frequency, short-wavelength oscillator (a) is excited, that frequency of radiation is present The presence of low-frequency, long-wavelength radiation (b) signifies that an oscillator of the corresponding frequency has been excited
Trang 4284 7 Introduction to quantum theory
physics, even cool objects should radiate in the visible and
ultraviolet regions, so objects should glow in the dark; there
should in fact be no darkness
In 1900, the German physicist Max Planck found that he
could account for the experimental observations by proposing
that the energy of each electromagnetic oscillator is limited to
discrete values and cannot be varied arbitrarily This proposal
is contrary to the viewpoint of classical physics in which all
possible energies are allowed and every oscillator has a mean
energy kT The limitation of energies to discrete values is called
the quantization of energy In particular, Planck found that he
could account for the observed distribution of energy if he
sup-posed that the permitted energies of an electromagnetic
oscil-lator of frequency ν are integer multiples of hν:
where h is a fundamental constant now known as Planck’s
constant On the basis of this assumption, Planck was able to
derive what is now called the Planck distribution:
ρ λ( , )T =λ5(e8hc kTπ/λ hc−1) Planck distribution (7A.6)
This expression fits the experimental curve very well at all
wavelengths (Fig 7A.5), and the value of h, which is an
unde-termined parameter in the theory, may be obtained by varying
its value until a best fit is obtained The currently accepted value
for h is 6.626 × 10−34 J s
As usual, it is a good idea to ‘read’ the content of an equation:
• The Planck distribution resembles the Rayleigh–
Jeans law (eqn 7A.4) apart from the all-important
exponential factor in the denominator For short
wavelengths, hc/νkT ≫ 1 and e hc/λkT → ∞ faster than
λ5 → 0; therefore ρ → 0 as λ → 0 or ν → ∞ Hence, the
energy density approaches zero at high frequencies,
in agreement with observation
• For long wavelengths, hc/λkT ≪ 1, and the denominator in the Planck distribution can be
replaced by (see Mathematical background 1)
we find that the Planck distribution reduces to the Rayleigh–Jeans law
• As we should infer from the graph in Fig 7A.5, the total energy density (the integral in eqn 7A.2 and therefore the area under the curve) is no longer infinite, and in fact
0
5 4 3
π
πλ
λ λ
Example 7A.1 Using the Planck distribution
Compare the energy output of a black-body radiator (such as
an incandescent lamp) at two different wavelengths by lating the ratio of the energy output at 450 nm (blue light) to that at 700 nm (red light) at 298 K
calcu-Method Use eqn 7A.6 At a temperature T, the ratio of the spectral density of states at a wavelength λ1 to that at λ2 is
ρ λ
ρ λ λ λ
λ λ
( , )( , ) (( ))
/ / 1
2
2 1
5 2 1
11
T T
Wavelength, λ
Rayleigh–Jeans law
Experimental
Figure 7A.4 The Rayleigh–Jeans law (eqn 7A.4) predicts an
infinite spectral density of states at short wavelengths This
approach to infinity is called the ultraviolet catastrophe
Trang 57A The origins of quantum mechanics 285
It is easy to see why Planck’s approach was successful whereas
Rayleigh’s was not The thermal motion of the atoms in the
walls of the black body excites the oscillators of the
electromag-netic field According to classical mechanics, all the oscillators
of the field share equally in the energy supplied by the walls,
so even the highest frequencies are excited The excitation of
very high frequency oscillators results in the ultraviolet
catas-trophe According to Planck’s hypothesis, however, oscillators
are excited only if they can acquire an energy of at least hν
This energy is too large for the walls to supply in the case of the
very high frequency oscillators, so the latter remain unexcited
The effect of quantization is to reduce the contribution from
the high frequency oscillators, for they cannot be significantly
excited with the energy available
(b) Heat capacities
In the early nineteenth century, the French scientists
Pierre-Louis Dulong and Alexis-Thérèse Petit determined the heat
capacities, C V = (∂U/∂T) V (Topic 2A), of a number of
mona-tomic solids On the basis of some somewhat slender
experi-mental evidence, they proposed that the molar heat capacities
of all monatomic solids are the same and (in modern units)
close to 25 J K−1 mol−1
Dulong and Petit’s law is easy to justify in terms of
classi-cal physics in much the same way as Rayleigh attempted to
explain black-body radiation If classical physics were valid,
the equipartition principle could be used to infer that the
mean energy of an atom as it oscillates about its mean
posi-tion in a solid is kT for each direcposi-tion of displacement As
each atom can oscillate in three dimensions, the average
energy of each atom is 3kT; for N atoms the total energy is 3NkT The contribution of this motion to the molar internal
energy is therefore
Um=3N kTA =3RT (7A.8a)
because NAk = R, the gas constant The molar constant volume
heat capacity is then predicted to be
monatomic solids are lower than 3R at low temperatures, and that the values approach zero as T → 0 To account for these
observations, Einstein (in 1905) assumed that each atom lated about its equilibrium position with a single frequency ν
oscil-He then invoked Planck’s hypothesis to assert that the energy
of oscillation is confined to discrete values, and specifically to
nhν, where n is an integer Einstein discarded the
equiparti-tion result, calculated the vibraequiparti-tional contribuequiparti-tion of the atoms
to the total molar internal energy of the solid (by a method described in Topic 15E), and obtained the expression now
known as the Einstein formula:
T T
,
/ /
The Einstein temperature, θE = hν/k, is a way of expressing
the frequency of oscillation of the atoms as a temperature and allows us to be quantitative about what we mean by ‘high tem-
perature’ (T ≫ θE) and ‘low temperature’ (T ≪ θE) in this text Note that a high vibrational frequency corresponds to a high Einstein temperature
con-As before, we now ‘read’ this expression:
• At high temperatures (when T ≫ θE) the exponentials
in fE can be expanded as 1 + θE/T + … and higher
terms ignored The result is
Consequently, the classical result (C V,m = 3R) is
obtained at high temperatures
• At low temperatures (when T ≪ θE) and eθE/T1,
E E
Insert the data to evaluate this ratio
Answer With λ1 = 450 nm and λ2 = 700 nm:
.
e
At room temperature, the proportion of short wavelength
radiation is insignificant
Self-test 7A.1 Repeat the calculation for a temperature of
13.6 MK, which is close to the temperature at the core of the
Sun
Answer: 5.85
Trang 6286 7 Introduction to quantum theory
The strongly decaying exponential function goes to zero
more rapidly than 1/T goes to infinity; so fE → 0 as T → 0,
and the heat capacity therefore approaches zero too
We see that Einstein’s formula accounts for the decrease of
heat capacity at low temperatures The physical reason for this
success is that at low temperatures only a few oscillators possess
enough energy to oscillate significantly so the solid behaves as
though it contains far fewer atoms than is actually the case At
higher temperatures, there is enough energy available for all
the oscillators to become active: all 3N oscillators contribute,
many of their energy levels are accessible, and the heat capacity
approaches its classical value
Figure 7A.6 shows the temperature dependence of the heat
capacity predicted by the Einstein formula The general shape
of the curve is satisfactory, but the numerical agreement is in
fact quite poor The poor fit arises from Einstein’s assumption
that all the atoms oscillate with the same frequency, whereas in
fact they oscillate over a range of frequencies from zero up to a
maximum value, νD This complication is taken into account by
averaging over all the frequencies present, the final result being
the Debye formula:
x x
where θD = hνD/k is the Debye temperature The integral in eqn
7A.11 has to be evaluated numerically, but that is simple with
mathematical software The details of this modification, which,
as Fig 7A.7 shows, gives improved agreement with
experi-ment, need not distract us at this stage from the main
conclu-sion, which is that quantization must be introduced in order to
explain the thermal properties of solids
(c) Atomic and molecular spectra
The most compelling and direct evidence for the
quantiza-tion of energy comes from spectroscopy, the detecquantiza-tion and
analysis of the electromagnetic radiation absorbed, emitted,
or scattered by a substance The record of the intensity of light intensity transmitted or scattered by a molecule as a function
of frequency (ν), wavelength (λ), or wavenumber ( = c is / )
called its spectrum (from the Latin word for appearance).
A typical atomic spectrum is shown in Fig 7A.8, and a typical molecular spectrum is shown in Fig 7A.9 The obvious feature of both is that radiation is emitted or absorbed at a series of discrete frequencies This observation can be understood if the energy
of the atoms or molecules is also confined to discrete values, for then energy can be discarded or absorbed only in discrete amounts (Fig 7A.10) Then, if the energy of an atom decreases by
ΔE, the energy is carried away as radiation of frequency ν, and an
emission ‘line’, a sharply defined peak, appears in the spectrum
We say that a molecule undergoes a spectroscopic transition, a change of state, when the Bohr frequency condition
Brief illustration 7A.1 The Debye formula
The Debye temperature for lead is 105 K, corresponding to a vibrational frequency of 2.2 × 1012 Hz As we see from Fig 7A.7,
fD ≈ 1 for T > θD and the heat capacity is almost classical For
lead at 25 °C, corresponding to T/θD = 2.8, fD = 0.99 and the heat capacity has almost its classical value
Self-test 7A.2 Evaluate the Debye temperature for diamond (νD = 4.6 × 1013 Hz) What fraction of the classical value of the heat capacity does diamond reach at 25 °C?
Answer: 2230 K; 15 per cent
Figure 7A.6 Experimental low-temperature molar heat
capacities and the temperature dependence predicted on the
basis of Einstein’s theory His equation (eqn 7A.10) accounts for
the dependence fairly well, but is everywhere too low
C V,m
T/θE or T/ θD
Debye Einstein
Figure 7A.7 Debye’s modification of Einstein’s calculation (eqn 7A.11) gives very good agreement with experiment For
copper, T/ θD= 2 corresponds to about 170 K, so the detection of deviations from Dulong and Petit’s law had to await advances
in low-temperature physics
(7A.11)
debye formula
Trang 77A The origins of quantum mechanics 287
is fulfilled We develop the principles and applications of atomic spectroscopy in Topics 9A–9C and of molecular spectroscopy
in Topics 12A–14D
At this stage we have established that the energies of the tromagnetic field and of oscillating atoms are quantized In this section we see the experimental evidence that led to the revision of two other basic concepts concerning natural phe-nomena One experiment shows that electromagnetic radia-tion—which classical physics treats as wave-like—actually also displays the characteristics of particles Another experiment shows that electrons—which classical physics treats as parti-cles—also display the characteristics of waves
elec-(a) The particle character of electromagnetic radiation
The observation that electromagnetic radiation of frequency
ν can possess only the energies 0, hν, 2hν, … suggests (and
at this stage it is only a suggestion) that it can be thought of
as consisting of 0, 1, 2, … particles, each particle having an
energy hν Then, if one of these particles is present, the energy
is hν, if two are present the energy is 2hν, and so on These
particles of electromagnetic radiation are now called photons
The observation of discrete spectra from atoms and molecules can be pictured as the atom or molecule generating a photon
Brief illustration 7A.2 The Bohr frequency condition
Atomic sodium produces a yellow glow (as in some street lamps) resulting from the emission of radiation of 590 nm The spectroscopic transition responsible for the emission involves electronic energy levels that have a separation given
−−19JThis energy difference can be expressed in a variety of ways For instance, multiplication by Avogadro’s constant results in
an energy separation per mole of atoms, of 203 kJ mol−1, parable to the energy of a weak chemical bond The calculated
com-value of ΔE also corresponds to 2.10 eV (Foundations B).
Self-test 7A.3 Neon lamps emit red radiation of wavelength
736 nm What is the energy separation of the levels in joules, kilojoules per mole, and electronvolts responsible for the emission?
Answer: 2.70 × 10 −19 J, 163 kJ mol −1 , 1.69 eV
Wavelength, λ/nm
Figure 7A.8 A region of the spectrum of radiation emitted by
excited iron atoms consists of radiation at a series of discrete
wavelengths (or frequencies)
Wavelength, λ/nm
Rotational transitions Vibrational
transitions
Figure 7A.9 When a molecule changes its state, it does so by
absorbing radiation at definite frequencies This spectrum is
part of that due to the electronic, vibrational, and rotational
excitation of sulfur dioxide (SO2) molecules This observation
suggests that molecules can possess only discrete energies,
not an arbitrary energy
Figure 7A.10 Spectroscopic transitions, such as those shown
above, can be accounted for if we assume that a molecule
emits electromagnetic radiation as it changes between
discrete energy levels Note that high-frequency radiation is
emitted when the energy change is large
Trang 8288 7 Introduction to quantum theory
of energy hν when it discards an energy of magnitude ΔE, with
ΔE = hν.
So far, the existence of photons is only a suggestion
Experimental evidence for their existence comes from the
measurement of the energies of electrons produced in the
pho-toelectric effect This effect is the ejection of electrons from
metals when they are exposed to ultraviolet radiation The
experimental characteristics of the photoelectric effect are as
follows:
• No electrons are ejected, regardless of the intensity of the
radiation, unless its frequency exceeds a threshold value
characteristic of the metal
• The kinetic energy of the ejected electrons increases
linearly with the frequency of the incident radiation but
is independent of the intensity of the radiation
• Even at low light intensities, electrons are ejected
immediately if the frequency is above the threshold
Figure 7A.11 illustrates the first and second characteristics.These observations strongly suggest that the photoelectric effect depends on the ejection of an electron when it is involved
in a collision with a particle-like projectile that carries enough energy to eject the electron from the metal If we suppose that
the projectile is a photon of energy hν, where ν is the frequency
of the radiation, then the conservation of energy requires that the kinetic energy of the ejected electron (Ek=1me 2)
2 v should obey
Ek=1mev2= −h Φ Photoelectric effect (7A.13)
In this expression, Φ (uppercase phi) is a characteristic of the
metal called its work function, the energy required to remove
an electron from the metal to infinity (Fig 7A.12), the analogue
of the ionization energy of an individual atom or molecule We
Example 7A.2 Calculating the number of photons
Calculate the number of photons emitted by a 100 W yellow
lamp in 1.0 s Take the wavelength of yellow light as 560 nm
and assume 100 per cent efficiency
Method Each photon has an energy hν, so the total number
of photons needed to produce an energy E is E/hν To use this
equation, we need to know the frequency of the radiation
(from ν = c/λ) and the total energy emitted by the lamp The
latter is given by the product of the power (P, in watts) and the
time interval for which the lamp is turned on (E = PΔt).
Answer The number of photons is
N E=h=h c( / )P t∆λ =λ P t hc∆
Substitution of the data gives
N =( 6 626 10( 5 60 10×× −−734mJs) () ( ××1002 998 10Js−×1) ( )×81 0msss−1)=2 8 10 × 20
Note that it would take the lamp nearly 40 min to produce
1 mol of these photons
A note on good practice To avoid rounding and other
numerical errors, it is best to carry out algebraic
calcu-lations first, and to substitute numerical values into a
single, final formula Moreover, an analytical result may
be used for other data without having to repeat the entire
calculation
Self-test 7A.4 How many photons does a monochromatic
(sin-gle frequency) infrared rangefinder of power 1 mW and
Increasing work function
Figure 7A.11 In the photoelectric effect, it is found that
no electrons are ejected when the incident radiation has a frequency below a value characteristic of the metal, and, above that value, the kinetic energy of the photoelectrons varies linearly with the frequency of the incident radiation
Figure 7A.12 The photoelectric effect can be explained if
it is supposed that the incident radiation is composed of photons that have energy proportional to the frequency of the radiation (a) The energy of the photon is insufficient to drive
an electron out of the metal (b) The energy of the photon is more than enough to eject an electron, and the excess energy
is carried away as the kinetic energy of the photoelectron
Trang 97A The origins of quantum mechanics 289
can now see that the existence of photons accounts for the three
observations we have summarized:
• Photoejection cannot occur if hν < Φ because the photon
brings insufficient energy
• Equation 7A.13 predicts that the kinetic energy of an
ejected electron should increase linearly with frequency
• When a photon collides with an electron, it gives up all
its energy, so we should expect electrons to appear as
soon as the collisions begin, provided the photons have
sufficient energy
A practical application of eqn 7A.13 is that it provides a
tech-nique for the determination of Planck’s constant, for the slopes
of the lines in Fig 7A.11 are all equal to h.
(b) The wave character of particles
Although contrary to the long-established wave theory of light, the view that light consists of particles had been held before, but discarded No significant scientist, however, had taken the view that matter is wave-like Nevertheless, experi-ments carried out in 1925 forced people to consider that possibility The crucial experiment was performed by the American physicists Clinton Davisson and Lester Germer, who observed the diffraction of electrons by a crystal (Fig
7A.13) Diffraction is the interference caused by an object in
the path of waves Depending on whether the interference is constructive or destructive, the result is a region of enhanced
or diminished intensity of the wave Davisson and Germer’s success was a lucky accident, because a chance rise of temper-ature caused their polycrystalline sample to anneal, and the ordered planes of atoms then acted as a diffraction grating
At almost the same time, G.P Thomson, working in Scotland, showed that a beam of electrons was diffracted when passed through a thin gold foil
The Davisson–Germer experiment, which has since been repeated with other particles (including α particles and molecular hydrogen), shows clearly that particles have wave-like properties, and the diffraction of neutrons is a well-established technique for investigating the structures and dynamics of condensed phases (Topic 18A) We have also seen that waves of electromagnetic radiation have particle-like properties Thus we are brought to the heart of modern
Example 7A.3 Calculating the maximum wavelength
capable of photoejection
A photon of radiation of wavelength 305 nm ejects an electron
from a metal with a kinetic energy of 1.77 eV Calculate the
maximum wavelength of radiation capable of ejecting an
elec-tron from the metal
Method Use eqn 7A.13 rearranged into Φ = hν − Ek with ν = c/λ
to calculate the work function of the metal from the data The
threshold for photoejection, the frequency able to remove the
electron but not give it any excess energy, then corresponds
to radiation of frequency νmin = Φ/h Use this value of the
frequency to calculate the maximum wavelength capable of
photoejection
Answer From the expression for the work function Φ = hν − Ek
the minimum frequency for photoejection is
Self-test 7A.5 When ultraviolet radiation of wavelength
165 nm strikes a certain metal surface, electrons are ejected with a speed of 1.24 Mm s−1 Calculate the speed of electrons ejected by radiation of wavelength 265 nm
Answer: 735 km s −1
Electron beam
Diffracted electrons
Ni crystal
Figure 7A.13 The Davisson–Germer experiment The scattering of an electron beam from a nickel crystal shows a variation of intensity characteristic of a diffraction experiment
in which waves interfere constructively and destructively in different directions
Trang 10290 7 Introduction to quantum theory
physics When examined on an atomic scale, the classical
concepts of particle and wave melt together, particles taking
on the characteristics of waves, and waves the characteristics
of particles
Some progress towards coordinating these properties had
already been made by the French physicist Louis de Broglie
when, in 1924, he suggested that any particle, not only
pho-tons, travelling with a linear momentum p = mv (with m the
mass and v the speed of the particle) should have in some
sense a wavelength given by what is now called the de Broglie
relation:
That is, a particle with a high linear momentum has a short
wavelength (Fig 7A.14) Macroscopic bodies have such high
momenta even when they are moving slowly (because their
mass is so great), that their wavelengths are undetectably small,
and the wavelike properties cannot be observed This
unde-tectability is why, in spite of its deficiencies, classical mechanics
can be used to explain the behaviour of macroscopic bodies
It is necessary to invoke quantum mechanics only for
micro-scopic systems, such as atoms and molecules, in which masses
are small
We now have to conclude that not only has electromagnetic radiation the character classically ascribed to particles, but electrons (and all other particles) have the characteristics clas-sically ascribed to waves This joint particle and wave character
of matter and radiation is called wave − particle duality.
Example 7A.4 Estimating the de Broglie wavelength
Estimate the wavelength of electrons that have been ated from rest through a potential difference of 40 kV
acceler-Method To use the de Broglie relation, we need to know the
linear momentum, p, of the electrons To calculate the linear
momentum, we note that the energy acquired by an electron
accelerated through a potential difference Δφ is eΔφ, where
e is the magnitude of its charge At the end of the period of
acceleration, all the acquired energy is in the form of kinetic
energy, Ek me 2 p m2
e
/2
=1 =
2 v , so we can determine p by setting
p2/2me equal to eΔφ As before, carry through the calculation
algebraically before substituting the data
Answer The expression p2/2me = eΔφ solves to p = (2meeΔφ)1/2;
then, from the de Broglie relation λ = h/p,
λ=(2m e h φ)1 2/
e ∆Substitution of the data and the fundamental constants (from inside the front cover) gives
λ ={ ( 2 9 109 10× × −31kg) ( 6 626 10×. 1 609 10× ×−34−Js19C) ( ×4 0 10× 4V))}
/
1 2 12
6 1 10
= × − mFor the manipulation of units we have used 1 V C = 1 J and
1 J = 1 kg m2 s−2 The wavelength of 6.1 pm is shorter than typical bond lengths in molecules (about 100 pm) Electrons acceler-ated in this way are used in the technique of electron diffraction
for the visualization of biological systems (Impact I7.1) and the
determination of the structures of solid surfaces (Topic 22A)
Self-test 7A.6 Calculate the wavelength of (a) a neutron with a
translational kinetic energy equal to kT at 300 K, (b) a tennis
ball of mass 57 g travelling at 80 km h−1
Answer: (a) 178 pm, (b) 5.2 × 10 −34 m
Checklist of concepts
☐ 1 A black body is an object capable of emitting and
absorbing all wavelengths of radiation uniformly ☐ 2 The vibrations of atoms can take up energy only in
dis-crete amounts
Short wavelength,
high momentum Long wavelength,low momentum
Figure 7A.14 An illustration of the de Broglie relation
between momentum and wavelength The wave is
associated with a particle A particle with high momentum
corresponds to a wave with a short wavelength, and
vice versa
Trang 117A The origins of quantum mechanics 291
☐ 3 Atomic and molecular spectra show that atoms and
molecules can take up energy only in discrete amounts
☐ 4 The photoelectric effect establishes the view that
elec-tromagnetic radiation, regarded in classical physics as
wavelike, consists of particles (photons)
☐ 5 The diffraction of electrons establishes the view that trons, regarded in classical physics as particles, are wave-
elec-like with a wavelength given by the de Broglie relation.
☐ 6 Wave–particle duality is the recognition that the
con-cepts of particle and wave blend together
Checklist of equations
Photoelectric effect Ek= 1me 2 =h
2 v −Φ Φ is the work function 7A.13
Trang 127B dynamics of microscopic systems
Wave–particle duality (Topic 7A) strikes at the heart of
clas-sical physics, where particles and waves are treated as entirely
distinct entities Experiments have also shown that the
ener-gies of electromagnetic radiation and of matter cannot be
var-ied continuously, and that for small objects the discreteness of
energy is highly significant In classical mechanics, in contrast,
energies can be varied continuously Such total failure of
clas-sical physics for small objects implied that its basic concepts
are false A new mechanics—quantum mechanics—had to be
devised to take its place
A new mechanics can be constructed from the ashes of sical physics by supposing that, rather than travelling along a definite path, a particle is distributed through space like a wave This remark may seem mysterious: it will be interpreted more fully shortly The mathematical representation of the wave that
clas-in quantum mechanics replaces the classical concept of
trajec-tory is called a wavefunction, ψ (psi), a function that contains
all the dynamical information about a system, such as its tion and momentum
In 1926, the Austrian physicist Erwin Schrödinger proposed an
equation for finding the wavefunction of any system The
time-independent Schrödinger equation for a particle of mass m
moving in one dimension with energy E in a system that does
not change with time (for instance, its volume remains stant) is
con-−2m x V x2 dd2ψ2 + ( )ψ=E ψ
The factor V(x) is the potential energy of the particle at the point x; because the total energy E is the sum of potential and
kinetic energies, the first term must be related (in a manner
we explore later) to the kinetic energy of the particle; ħ = h/2π (which is read h-cross or h-bar) is a convenient modification
of Planck’s constant with the value 1.055 × 10−34 J s Three ple but important general forms of the potential energy are (the explicit forms are found in the corresponding Topics):
sim-• For a particle moving freely in one dimension the
potential energy is constant, so V(x) = V It is often convenient to write V = 0 (Topic 8A).
• For a particle free to oscillate to-and-fro near a point x0,
V(x) ∝ (x − x0)2 (Topic 8B)
• For two electric charges Q1 and Q2 separated by a
distance x, V(x) ∝ Q1Q2/x (Foundations B).
The following Justification shows that the Schrödinger
equa-tion is plausible and the discussions later in the chapter will help to overcome its apparent arbitrariness For the present, we
➤
➤ Why do you need to know this material?
Quantum theory provides the essential foundation for
understanding of the properties of electrons in atoms and
molecules.
➤
➤ What is the key idea?
All the dynamical properties of a system are contained
in the wavefunction, which is obtained by solving the
Schrödinger equation.
➤
➤ What do you need to know already?
You need to be aware of the shortcomings of classical
physics that drove the development of quantum theory
(Topic 7A).
time-independent schrödinger
Contents
7b.2 The Born interpretation of the wavefunction 293
example 7b.1: Interpreting a wavefunction 294
(a) Normalization 295
example 7b.2: normalizing a wavefunction 296
(b) Constraints on the wavefunction 296
example 7b.3: determining a probability 298
Trang 137B Dynamics of microscopic systems 293
shall treat the equation simply as a quantum-mechanical
pos-tulate that replaces Newton’s pospos-tulate of his apparently equally
arbitrary equation of motion (that force = mass × acceleration)
Various ways of expressing the Schrödinger equation, of
incor-porating the time-dependence of the wavefunction, and of
extending it to more dimensions, are collected in Table 7B.1 In
the Topics of Chapter 8 we solve the equation for a number of
important cases; in this chapter we are mainly concerned with
its significance, the interpretation of its solutions, and seeing
how it implies that energy is quantized
wavefunction
A central principle of quantum mechanics is that the
wavefunc-tion contains all the dynamical informawavefunc-tion about the system it describes Here we concentrate on the information it carries
about the location of the particle
The interpretation of the wavefunction in terms of the location of the particle is based on a suggestion made by Max Born He made use of an analogy with the wave theory
of light, in which the square of the amplitude of an magnetic wave in a region is interpreted as its intensity and therefore (in quantum terms) as a measure of the probability
electro-of finding a photon present in the region The Born
interpre-tation of the wavefunction focuses on the square of the
wave-function (or the square modulus, |ψ|2 = ψ *ψ, if ψ is complex; see Mathematical background 3) For a one-dimensional sys-
tem (Fig 7B.1):
If the wavefunction of a particle has the value ψ at some point x, then the probability of finding the particle between x and x + dx is proportional to
|ψ|2dx.
Thus, |ψ|2 is the probability density, and to obtain the
prob-ability it must be multiplied by the length of the infinitesimal
region dx The wavefunction ψ itself is called the probability
amplitude For a particle free to move in three dimensions
(for example, an electron near a nucleus in an atom), the
Justification 7B.1 The plausibility of the Schrödinger
equation
The Schrödinger equation can be seen to be plausible by
not-ing that it implies the de Broglie relation (eqn 7A.14, p = h/λ)
for a freely moving particle After writing V(x) = V, we can
rearrange eqn 7B.1 into
treated in Mathematical background 4 at the end of Chapter 8;
we need only the simplest procedures in this Topic In this case a solution is
/
kx k 2m E V2
1 2
We now recognize that cos kx is a wave of wavelength λ = 2π/k,
as can be seen by comparing cos kx with the standard form
of a harmonic wave, cos(2πx/λ) (Foundations C) The tity E − V is equal to the kinetic energy of the particle, Ek, so
quan-k = (2mEk/2)1/2, which implies that Ek = k22/2m Because
Ek = p2/2m (Foundations B), it follows that p = k Therefore,
the linear momentum is related to the wavelength of the function by
wave-p=2λπ×2hπ=h λwhich is the de Broglie relation
Table 7B.1 The Schrödinger equation
2 2
Two dimensions
2 2
Alternative forms
Legendrian
φ θ θ θ θ θ
2 2
2 2
Trang 14294 7 Introduction to quantum theory
wavefunction depends on the point r with coordinates x, y, and
z, and the interpretation of ψ(r) is as follows (Fig 7B.2):
If the wavefunction of a particle has the value ψ at some
point r, then the probability of finding the particle in an
infinitesimal volume dτ = dxdydz at that point is
proportional to |ψ|2dτ.
The Born interpretation does away with any worry about the
significance of a negative (and, in general, complex) value of ψ
because |ψ|2 is real and never negative There is no direct
sig-nificance in the negative (or complex) value of a wavefunction:
only the square modulus, a positive quantity, is directly
physi-cally significant, and both negative and positive regions of a
wavefunction may correspond to a high probability of finding a
particle in a region (Fig 7B.3) However, later we shall see that
the presence of positive and negative regions of a wavefunction
is of great indirect significance, because it gives rise to the
possi-bility of constructive and destructive interference between
dif-ferent wavefunctions
Example 7B.1 Interpreting a wavefunction
In Topic 9A it is shown that the wavefunction of an electron in the lowest energy state of a hydrogen atom is proportional to
e−r a/ 0, with a0 a constant and r the distance from the nucleus
Calculate the relative probabilities of finding the electron
inside a region of volume δV = 1.0 pm3, which is small even on the scale of the atom, located at (a) the nucleus, (b) a distance
a0 from the nucleus
Method The region of interest is so small on the scale of the
atom that we can ignore the variation of ψ within it and write the probability, P, as proportional to the probability density (ψ 2; note that ψ is real) evaluated at the point of interest mul- tiplied by the volume of interest, δV That is, P ∝ ψ 2δV, with
Therefore, the ratio of probabilities is 1.0/0.14 = 7.1 Note that
it is more probable (by a factor of 7) that the electron will be found at the nucleus than in a volume element of the same
size located at a distance a0 from the nucleus The negatively charged electron is attracted to the positively charged nucleus, and is likely to be found close to it
A note on good practice The square of a wavefunction is
a probability density, and (in three dimensions) has the dimensions of 1/length3 It becomes a (unitless) prob-ability when multiplied by a volume In general, we have
to take into account the variation of the amplitude of the wavefunction over the volume of interest, but here we are supposing that the volume is so small that the variation of
ψ in the region can be ignored.
Figure 7B.1 The wavefunction ψ is a probability amplitude in
the sense that its square modulus (ψ*ψ or |ψ|2) is a probability
density The probability of finding a particle in the region
dx located at x is proportional to | ψ|2dx We represent
the probability density by the density of shading in the
superimposed band
dx dy dz
z
r
Figure 7B.2 The Born interpretation of the wavefunction
in three-dimensional space implies that the probability of
finding the particle in the volume element dτ = dxdydz at some
location r is proportional to the product of d τ and the value of
|ψ|2 at that location
Figure 7B.3 The sign of a wavefunction has no direct physical significance: the positive and negative regions of this wavefunction both correspond to the same probability distribution (as given by the square modulus of ψ and depicted
by the density of the shading)
Trang 157B Dynamics of microscopic systems 295
(a) Normalization
A mathematical feature of the Schrödinger equation is that if
ψ is a solution, then so is Nψ, where N is any constant This
feature is confirmed by noting that ψ occurs in every term in
eqn 7B.1, so any constant factor can be cancelled This
free-dom to vary the wavefunction by a constant factor means that
it is always possible to find a normalization constant, N, such
that the proportionality of the Born interpretation becomes an
equality
We find the normalization constant by noting that, for a
nor-malized wavefunction Nψ, the probability that a particle is in the
region dx is equal to (Nψ *)(Nψ)dx (we are taking N to be real)
Furthermore, the sum over all space of these individual
prob-abilities must be 1 (the probability of the particle being
some-where is 1) Expressed mathematically, the latter requirement is
N2 ψ ψ* d =x 1
−∞
∞
Wavefunctions for which the integral in eqn 7B.2 exists (in the
sense of having a finite value) are said to be ‘square-integrable’
Therefore, by evaluating the integral, we can find the value of N
and hence ‘normalize’ the wavefunction From now on, unless
we state otherwise, we always use wavefunctions that have been
normalized to 1; that is, from now on we assume that ψ already
includes a factor which ensures that (in one dimension)
where dτ = dxdydz and the limits of this definite integral are
not written explicitly: in all such integrals, the integration is over all the space accessible to the particle For systems with spherical symmetry it is best to work in spherical polar coor-
dinates (The chemist’s toolkit 7B.1), so the explicit form of eqn
7B.4c is
ψ ψ* r r2 sinθ θ φ
0
2 0
Self-test 7B.1 The wavefunction for the electron in its lowest
energy state in the ion He+ is proportional to e−2r a/ 0 Repeat the
calculation for this ion Any comment?
Answer: 55; more compact wavefunction
The chemist’s toolkit 7B.1 Spherical polar coordinatesFor systems with spherical symmetry it is best to work in
spherical polar coordinates r, θ, and ϕ (Sketch 1)
x r= sin cos ,θ φ y r= sin sinθ φ,z r= cos θ
where:
r, the radius, ranges from 0 to ∞
θ, the colatitude, ranges from 0 to π
ϕ, the azimuth, ranges from 0 to 2πThat these ranges cover space is illustrated in Sketch 2 Standard manipulations then yield
Sketch 2 The surface of a sphere is covered by allowing θ
to range from 0 to π, and then sweeping that arc around a complete circle by allowing ϕ to range from 0 to 2π
spherical polar coordinates
Trang 16296 7 Introduction to quantum theory
(b) Constraints on the wavefunction
The Born interpretation puts severe restrictions on the
accept-ability of wavefunctions The principal constraint is that ψ must
not be infinite over a finite region If it were, it would not be square-integrable, and the normalization constant would be zero The normalized function would then be zero everywhere, except where it is infinite, which would be unacceptable (the particle must be somewhere) Note that infinitely sharp spikes are acceptable provided they have zero width
The requirement that ψ is finite everywhere rules out many
possible solutions of the Schrödinger equation, because many mathematically acceptable solutions rise to infinity and are therefore physically unacceptable We could imagine a solu-tion of the Schrödinger equation that gives rise to more than
one value of |ψ|2 at a single point The Born interpretation implies that such solutions are unacceptable, because it would
be absurd to have more than one probability that a particle is at the same point This restriction is expressed by saying that the
wavefunction must be single-valued; that is, have only one value
at each point of space
The Schrödinger equation itself also implies some ematical restrictions on the type of functions that can occur Because it is a second-order differential equation, the second
math-derivative of ψ must be well-defined if the equation is to be
applicable everywhere We can take the second derivative of a function only if it is continuous (so there are no sharp steps in
it, Fig 7B.4) and if its first derivative, its slope, is continuous (so there are no kinks in the wavefunction)
There are cases, and we shall meet them, where acceptable wavefunctions have kinks These cases arise when the poten-tial energy has peculiar properties, such as rising abruptly to infinity When the potential energy is smoothly well-behaved and finite, the slope of the wavefunction must be continuous;
if the potential energy becomes infinite, then the slope of the wavefunction need not be continuous There are only two cases
Example 7B.2 Normalizing a wavefunction
Normalize the wavefunction used for the hydrogen atom in
Example 7B.1.
Method We need to find the factor N that guarantees that the
integral in eqn 7B.4c is equal to 1 Because the system is
spheri-cal, it is most convenient to use spherical coordinates (The
chemist’s toolkit 7B.1) and to carry out the integrations
speci-fied in eqn 7B.4d Relevant integrals are found in the Resource
Note that because a0 is a length, the dimensions of ψ are
1/length3/2 and therefore those of ψ2 are 1/length3 (for
instance, 1/m3) as is appropriate for a probability density
If Example 7B.1 is now repeated, we can obtain the actual
probabilities of finding the electron in the volume element at
each location, not just their relative values Given (from inside
the front cover) that a0 = 52.9 pm, the results are (a) 2.2 × 10−6,
corresponding to 1 chance in about 500 000 inspections of
finding the electron in the test volume, and (b) 2.9 × 10−7,
cor-responding to 1 chance in 3.4 million
Self-test 7B.2 Normalize the wavefunction given in Self-test
7B.1
Answer: N=( 8/ πa0) 1 2 /
In these coordinates, the integral of a function f(r,θ,φ) over all
space takes the form
where the limits on the first integral sign refer to r, those on
the second to θ, and those on the third to φ.
Trang 177B Dynamics of microscopic systems 297
of this behaviour in elementary quantum mechanics, and the
peculiarity will be mentioned when we meet them
At this stage we see that ψ:
• must not be infinite over a non-infinitesimal
The restrictions just noted are so severe that acceptable
solu-tions of the Schrödinger equation do not in general exist for
arbitrary values of the energy E In other words, a particle may
possess only certain energies, for otherwise its wavefunction
would be physically unacceptable That is, as a consequence of
the restrictions on its wavefunction, the energy of a particle is
quantized We can find the acceptable energies by solving the
Schrödinger equation for motion of various kinds, and
select-ing the solutions that conform to the restrictions listed above
That task is taken forward in Chapter 8
Once we have obtained the normalized wavefunction, we can
then proceed to determine the probability density As an
exam-ple, consider a particle of mass m free to move parallel to the
x-axis with zero potential energy The Schrödinger equation is
obtained from eqn 7B.1 by setting V = 0, and is
−2m2 d2dψ x( )2x =E x ψ( ) (7B.5)
As shown in the following Justification, the solutions of this
equation have the form
ψ ( ) x A= eikx+Be− ikx E k m= 2 2
where A and B are constants (See Mathematical background 3
at the end of this chapter for more on complex numbers.)
We see in Topic 8A what determines the values of A and B;
here we can treat them as arbitrary constants that we can vary
at will Suppose that B = 0 in eqn 7B.6, then the wavefunction is
This probability density is independent of x; so, wherever we
look in a region of fixed length located anywhere along the
x-axis, there is an equal probability of finding the particle (Fig
7B.5a) In other words, if the wavefunction of the particle is given by eqn 7B.7, then we cannot predict where we will find
it The same would be true if the wavefunction in eqn 7B.6 had
A = 0; then the probability density would be |B|2, a constant
Now suppose that in the wavefunction A = B Then, because
coskx=1 eikx+e− ikx
2( ) (Mathematical background 3), eqn 7B.6
becomes
The probability density now has the form
| ( )| (ψ x 2= 2 cosA kx)*(2 cosA kx)=4A2cos2kx (7B.10)
Justification 7B.2 The wavefunction of a free particle in
one dimension
To verify that ψ(x) in eqn 7B.6 is a solution of eqn 7B.5, we
simply substitute it into the left-hand side of the equation and
show that E = k22/2m To begin, we write
−2 2 = − + −
2
2 2 2