1. Trang chủ
  2. » Giáo án - Bài giảng

Chapter 8 Atkins Physical Chemistry (10th Edition) Peter Atkins and Julio de Paula

40 788 0

Đang tải... (xem toàn văn)

Tài liệu hạn chế xem trước, để xem đầy đủ mời bạn chọn Tải xuống

THÔNG TIN TÀI LIỆU

Thông tin cơ bản

Định dạng
Số trang 40
Dung lượng 1,41 MB

Nội dung

Chapter 8 Atkins Physical Chemistry (10th Edition) Peter Atkins and Julio de Paula Chapter 8 Atkins Physical Chemistry (10th Edition) Peter Atkins and Julio de Paula Chapter 8 Atkins Physical Chemistry (10th Edition) Peter Atkins and Julio de Paula Chapter 8 Atkins Physical Chemistry (10th Edition) Peter Atkins and Julio de Paula Chapter 8 Atkins Physical Chemistry (10th Edition) Peter Atkins and Julio de Paula

Trang 1

chaPter 8

the quantum theory of motion

The three basic modes of motion—translation (motion

through space), vibration, and rotation—all play an important

role in chemistry because they are ways in which molecules

store energy Gas-phase molecules, for instance, undergo

trans-lational motion and their kinetic energy is a contribution to

the total internal energy of a sample Molecules can also store

energy as rotational kinetic energy and transitions between

their rotational energy states can be observed spectroscopically

Energy is also stored as molecular vibration, and transitions

between vibrational states also give rise to spectroscopic

signa-tures In this Chapter we use the principles of quantum theory

to calculate the properties of microscopic particles in motion

In this Topic we see that, according to quantum theory, a

parti-cle constrained to move in a finite region of space is described

by only certain wavefunctions and their corresponding

ener-gies Hence, quantization emerges as a natural consequence of

solving the Schrödinger equation and the conditions imposed

on it The solutions also bring to light a number of non-classical

features of particles, especially their ability to tunnel into and

through regions where classical physics would forbid them to

be found

This Topic introduces the ‘harmonic oscillator’, a simple

but very important model for the description of molecular

vibrations We see that the energies of oscillator are quantized The acceptable wavefunctions also show that the oscillator may

be found at extensions and compressions that are forbidden by classical physics

The energy of a rotating particle is quantized, but in this Topic

we see that its angular momentum is also restricted to tain values The quantization of angular momentum is a very important aspect of the quantum theory of electrons in atoms and of rotating molecules

cer-What is the impact of this material?

‘Nanoscience’ is the study of atomic and molecular blies with dimensions ranging from 1 nm to about 100 nm and ‘nanotechnology’ is concerned with the incorporation of such assemblies into devices We encounter several concepts

assem-of nanoscience throughout the text In Impact I8.1 we explore

quantum mechanical effects that render the properties of a nanometre-sized assembly dependent on its size

To read more about the impact of this material, scan the QR code, or go to bcs.whfreeman.com/webpub/chemistry/pchem10e/impact/pchem-8-1.html

Trang 2

be found.

8A.1 Free motion in one dimension

The Schrödinger equation for a particle of mass m moving

freely in one dimension is (Topic 7B)

2m2 2 2x =

d

dψ( ) ψ( )and the solutions are (as in eqn 7B.6)

for all values of k Because the energy of the particle is portional to k2, all non-negative values, including zero, of the

pro-energy are permitted It follows that the translational pro-energy of a

free particle is not quantized.

➤ Why do you need to know this material?

The application of quantum theory to translation reveals

the origin of quantization and other non-classical features

of physical and chemical phenomena This material is

important for the discussion of atoms and molecules that

are free to move within a restricted volume, such as a gas

in a container.

➤ What is the key idea?

The translational energy levels of a particle confined to

a finite region of space are quantized, and under certain

conditions particles can pass into and through classically

forbidden regions.

➤ What do you need to know already?

You should know that the wavefunction is the solution of the Schrödinger equation (Topic 7B), and be familiar with the techniques of deriving dynamical properties from the wavefunction by using operators corresponding to the observables (Topic 7C).

Free motion

in one dimension

schrödinger equation (8A.1)

Free motion

in one dimension

functions and energies

wave-(8A.2)

Contents

brief illustration 8a.1: the wavefunction

of a freely-moving particle 318

(a) The acceptable solutions 318

brief illustration 8a.2: the energy of a particle

(b) The properties of the wavefunctions 320

example 8a.1: determining the probability

of finding the particle in a finite region 320

(c) The properties of observables 321

example 8a.2: estimating an absorption

8a.3 Confined motion in two or more dimensions 322

(a) Separation of variables 322

brief illustration 8a.3: the distribution of a

particle in a two-dimensional box 323

brief illustration 8a.4: degeneracies in a

brief illustration 8a.5: transmission probabilities

Trang 3

318 8 The quantum theory of motion

The values of the constants A and B depend on how the state

of motion of the particle is prepared:

If it is shot towards positive x, then its linear

momentum is +k (Topic 7C), and its wavefunction

is proportional to eikx In this case B = 0 and A is a

normalization factor

• If the particle is shot in the opposite direction,

towards negative x, then its linear momentum is −k

and its wavefunction is proportional to e−ikx In this

case, A = 0 and B is the normalization factor.

The probability density |ψ|2 is uniform if the particle is in

either of the pure momentum states eikx or e−ikx According to

the Born interpretation (Topic 7B), nothing further can be said

about the location of the particle That conclusion is

consist-ent with the uncertainty principle, because if the momconsist-entum

is certain, then the position cannot be specified (the operators

corresponding to x and p do not commute and thus correspond

to complementary observables, Topic 8C)

8A.2 Confined motion in one

dimension

Consider a particle in a box in which a particle of mass m is

confined to a finite region of space between two

impenetra-ble walls The potential energy is zero inside the box but rises

abruptly to infinity at the walls at x = 0 and x = L (Fig 8A.1)

When the particle is between the walls, the Schrödinger tion is the same as for a free particle (eqn 8A.1), so the general solutions given in eqn 8A.2 are also the same However, it will prove convenient to use e±ikx = cos kx ± i sin kx (Mathematical

ψ k( )x C= sinkx D+ coskx general solution for 0 ≤ x ≤ L (8A.3)

Outside the box the wavefunctions must be zero as the cle will not be found in a region where its potential energy is infinite:

parti-Forx<0andx L> , ψ k( )x =0 (8A.4)

At this point, there are no restrictions on the value of k and all

solutions appear to be acceptable

(a) The acceptable solutions

The requirement of the continuity of the wavefunction (Topic

7B) implies that ψ k (x) as given by eqn 8A.3 must be zero at the

walls, for it must match the wavefunction inside the material of the walls where the functions meet That is, the wavefunction

must satisfy the following two boundary conditions, or

con-straints on the function at certain locations:

ψ k( )0 0= andψ k( )L =0

As we show in the following Justification, the requirement

that the wavefunction satisfy these boundary conditions implies that only certain wavefunctions are acceptable and

Brief illustration 8A.1 The wavefunction of a

freely-moving particle

An electron at rest that is shot out of an accelerator towards

positive x through a potential difference of 1.0 V acquires a

kinetic energy of 1.0 eV or 0.16 aJ (1.6 × 10−19 J) The

wavefunc-tion for such a particle is given by eqn 8A.3 with B = 0 and

k given by rearranging the expression for the energy in eqn

/

m

or 5.1 nm−1 (with 1 nm = 10−9 m) Therefore the wavefunction is

ψ(x) = Ae 5.1ix/nm

Self-test 8A.1 Write the wavefunction for an electron

travel-ling to the left (negative x) after being accelerated through a

Figure 8A.1 A particle in a one-dimensional region with

impenetrable walls Its potential energy is zero between x = 0

and x = L, and rises abruptly to infinity as soon as it touches the

walls

Trang 4

where C is an as yet undetermined constant Note that the

wavefunctions and energy are now labelled with the

dimen-sionless integer n instead of the quantity k.

We conclude that the energy of the particle in a

one-dimen-sional box is quantized and that this quantization arises from

the boundary conditions that ψ must satisfy This is a general

conclusion: the need to satisfy boundary conditions implies that

only certain wavefunctions are acceptable, and hence restricts

observables to discrete values So far, only energy has been

quan-tized; shortly we shall see that other physical observables may

also be quantized

We need to determine the constant C in eqn 8A.6a To do

so, we normalize the wavefunction to 1 by using a standard

integral from the Resource section Because the wavefunction is

zero outside the range 0 ≤ x ≤ L, we use

n n

where the energies and wavefunctions are labelled with the

quantum number n A quantum number is an integer (in some

cases, as we see in Topic 9B, a half-integer) that labels the state

of the system For a particle in a one-dimensional box there is

an infinite number of acceptable solutions, and the quantum

number n specifies the one of interest (Fig 8A.2) As well as

acting as a label, a quantum number can often be used to culate the energy corresponding to the state and to write down the wavefunction explicitly (in the present example, by using the relations in eqn 8A.7)

cal-Justification 8A.1 The energy levels and wavefunctions

of a particle in a one-dimensional box

From the boundary condition ψ k(0) = 0 and the fact that, from

eqn 8A.3, ψ k (0) = D (because sin 0 = 0 and cos 0 = 1), we can

conclude that D = 0 It follows that the wavefunction must be

of the form ψ k (x) = C sin kx From the second boundary

con-dition, ψ k (L) = 0, we know that ψ k (L) = C sin kL = 0 We could

take C = 0, but doing so would give ψ k (x) = 0 for all x, which

would conflict with the Born interpretation (the particle must

be somewhere) The alternative is to require that kL be chosen

so that sin kL = 0 This condition is satisfied if

kL n= π n=1 2, ,…

The value n = 0 is ruled out, because it implies k = 0 and

ψ k (x) = 0 everywhere (because sin 0 = 0), which is

unaccep-table Negative values of n merely change the sign of sin kL

(because sin(−x) = −sin x) and do not result in new solutions

The wavefunctions are therefore

ψ n( )x C= sin (n x Lπ / ) n=1 2, ,…

as in eqn 8A.6a At this stage we have begun to label the

solu-tions with the index n instead of k Because k and E k are related

by eqn 8A.2, and k and n are related by kL = nπ, it follows that

the energy of the particle is limited to E n = n2h2/8mL2, as in eqn

8A.6b

One­dimensional box energy levels (8A.7b)

Brief illustration 8A.2 The energy of a particle in a box

A long carbon nanotube can be modelled as a one-dimensional structure and its electrons described by particle-in-a-box wavefunctions The lowest energy of an electron in a carbon

nanotube of length 100 nm is given by eqn 8A.7b with n = 1:

kg 1100 10× − 9m)2=6 02 10. × −24J

or 0.00602 zJ and its wavefunction is ψ1(x) = (2/L)1/2sin(πx/L).

Self-test 8A.2 What are the energy and wavefunction for the next higher energy electron of the system described in this

functions (8A.7a)

100

Classically allowed energies

12 3

5 4 6 7 8 9 10

n

Figure 8A.2 The allowed energy levels for a particle in a

box Note that the energy levels increase as n2, and that their separation increases as the quantum number increases Classically, the particle is allowed to have any value of the energy in the continuum shown as a shaded area

Trang 5

320 8 The quantum theory of motion

(b) The properties of the wavefunctions

Figure 8A.3 shows some of the wavefunctions of a particle in a

one-dimensional box We see that:

• The wavefunctions are all sine functions with the

same amplitude but different wavelengths

• Shortening the wavelength results in a sharper

average curvature of the wavefunction and therefore

an increase in the kinetic energy of the particle (its

only source of energy because V = 0 inside the box).

The number of nodes also increases as n increase;

the wavefunction ψ n has n − 1 nodes.

• Increasing the number of nodes between walls of a

given separation increases the average curvature of

the wavefunction and hence the kinetic energy of the

• and varies with position The non-uniformity in the

probability density is pronounced when n is small

(Fig. 8A.4) The most probable locations of the particle correspond to the maxima in the probability density

The probability density ψ n2( ) becomes more uniform as n x

increases provided we ignore the fine detail of the increasingly rapid oscillations (Fig 8A.5) The probability density at high quantum numbers reflects the classical result that a particle bouncing between the walls spends, on the average, equal times

at all points That the quantum result corresponds to the cal prediction at high quantum numbers is an illustration of the

classi-correspondence principle, which states that classical

mechan-ics emerges from quantum mechanmechan-ics as high quantum bers are reached

num-Example 8A.1 Determining the probability of finding the particle in a finite region

The wavefunctions of an electron in a conjugated polyene can

be approximated by particle-in-a-box wavefunctions What is

the probability, P, of locating the electron between x = 0 (the left-hand end of a molecule) and x = 0.2 nm in its lowest energy

state in a conjugated molecule of length 1.0 nm?

Method According to the Born interpretation, ψ(x)2dx is the probability of finding the particle in the small region dx located at x; therefore, the total probability of finding the elec- tron in the specified region is the integral of ψ(x)2dx over that

region, as given in eqn 7B.11 The wavefunction of the electron

is given in eqn 8A.7a with n = 1 The integral you need is in the Resource section:

Answer The probability of finding the particle in a region

Now set n = 1, L = 1.0 nm, and l = 0.2 nm, which gives P = 0.05

The result corresponds to a chance of 1 in 20 of finding the

electron in the region As n becomes infinite, the sine term, which is multiplied by 1/n, makes no contribution to P and the classical result for a uniformly distributed particle, P = l/L, is

obtained

Self-test 8A.3 Calculate the probability that an electron in the

state with n = 1 will be found between x = 0.25L and x = 0.75L in

a conjugated molecule of length L (with x = 0 at the left-hand

end of the molecule)

Figure 8A.3 The first five normalized wavefunctions of a

particle in a box Each wavefunction is a standing wave;

successive functions possess one more half wave and a

correspondingly shorter wavelength

Figure 8A.4 (a) The first two wavefunctions, (b) the

corresponding probability densities, and (c) a representation of

the probability density in terms of the darkness of shading

Trang 6

8A Translation  321

(c) The properties of observables

The linear momentum of a particle in a box is not well defined

because the wavefunction sin kx is not an eigenfunction of the

linear momentum operator However, each wavefunction is a

linear combination of the linear momentum eigenfunctions eikx

and e−ikx Then, because sin x = (e ix − e−ix)/2i, we can write

It follows from the discussion in Topic 7C that half the

meas-urements of the linear momentum will give the value +k and

−k for the other half This detection of opposite directions

of travel with equal probability is the quantum mechanical

version of the classical picture that a particle in a

one-dimen-sional box rattles from wall to wall and in any given period

spends half its time travelling to the left and half travelling to

the right

Because n cannot be zero, the lowest energy that the

parti-cle may possess is not zero (as would be allowed by classical

mechanics, corresponding to a stationary particle) but

= Particle in a box Zero-point energy (8A.10)

This lowest, irremovable energy is called the zero-point energy

The physical origin of the zero-point energy can be explained

in two ways:

• The Heisenberg uncertainty principle requires a

particle to possess kinetic energy if it is confined to a

finite region: the location of the particle is not

completely indefinite (Δx ≠ ∞), so the uncertainty in

its momentum cannot be precisely zero (Δp ≠ 0)

Because Δp = (〈p2〉 − 〈p〉2)1/2 = 〈p2〉1/2 in this case, Δp ≠ 0

implies that 〈p2〉 ≠ 0, which implies that the particle must always have nonzero kinetic energy

• If the wavefunction is to be zero at the walls, but smooth, continuous, and not zero everywhere, then

it must be curved, and curvature in a wavefunction implies the possession of kinetic energy

The separation between adjacent energy levels with quantum

numbers n and n + 1 is

E n+ 1−E n= +n mL hn h mL = n+ mL h

2 2 2

2 2 2

2 2

Example 8A.2 Estimating an absorption wavelength

β-Carotene (1) is a linear polyene in which 10 single and 11

double bonds alternate along a chain of 22 carbon atoms If

we take each CeC bond length to be about 140 pm, then the

length L of the molecular box in β-carotene is L = 2.94 nm

Estimate the wavelength of the light absorbed by this molecule from its ground state to the next higher excited state

1 β-Carotene

Method For reasons that will be familiar from introductory chemistry, each C atom contributes one p electron to the π-orbitals Use eqn 8A.11 to calculate the energy separation between the highest occupied and the lowest unoccupied levels, and convert that energy to a wavelength by using the Bohr frequency relation (eqn 7A.12)

Answer There are 22 C atoms in the conjugated chain; each contributes one p electron to the levels, so each level up to

n = 11 is occupied by two electrons The separation in energy

between the ground state and the state in which one electron is

mJ

)

Figure 8A.5 The probability density ψ 2(x) for large quantum

number (here n = 50, blue, compared with n = 1, red) Notice

that for high n the probability density is nearly uniform,

provided we ignore the fine detail of the increasingly rapid

oscillations

Trang 7

322 8 The quantum theory of motion

8A.3 Confined motion in two or more

dimensions

Now consider a rectangular two-dimensional region of a

sur-face with length L1 in the x-direction and L2 in the y-direction;

the potential energy is zero everywhere except at the walls,

where it is infinite (Fig 8A.6) As a result, the particle is never

found at the walls and its wavefunction is zero there and

every-where outside the two-dimensional region Between the walls,

because the particle has contributions to its kinetic energy from

its motion in both the x and y directions, the Schrödinger

equa-tion has two kinetic energy terms, one for each axis For a

parti-cle of mass m the equation is

This is a partial differential equation (Mathematical background 4), and the resulting wavefunctions are functions of both x and

y, denoted ψ(x,y) This dependence means that the

wavefunc-tion and the corresponding probability density depend on the location in the plane, with each position specified by the values

of the coordinates x and y.

(a) Separation of variables

A partial differential equation of the form of eqn 8A.12

can be simplified by the separation of variables technique

(Mathematical background 4), which divides the equation into

two or more ordinary differential equations, one for each

vari-able We show in the Justification below using this technique

that the wavefunction can be written as a product of functions,

one depending only on x and the other only on y:

ψ ( , ) x y X x Y y= ( ) ( ) (8A.13a)

and that the total energy is given by

where E X is the energy associated with the motion of the

parti-cle parallel to the x-axis, and likewise for E Y and motion parallel

to the y-axis.

or 0.160 aJ It follows from the Bohr frequency condition

(ΔE = hν) that the frequency of radiation required to cause this

λ = 1240 nm The experimental value is 603 THz (λ = 497 nm),

corresponding to radiation in the visible range of the

electro-magnetic spectrum Considering the crudeness of the model

we have adopted here, we should be encouraged that the

com-puted and observed frequencies agree to within a factor of 2.5

Self-test 8A.4 Estimate a typical nuclear excitation energy

in electronvolts (1 eV = 1.602 × 10−19 J; 1 GeV = 109 eV) by

cal-culating the first excitation energy of a proton confined to a

one-dimensional box with a length equal to the diameter of a

nucleus (approximately 1 × 10−15 m, or 1 fm)

Answer: 0.6 GeV

Justification 8A.2 The separation of variables technique applied to the particle in a two-dimensional box

We follow the procedure in Mathematical background 4

and apply it to eqn 8A.12 The first step to confirm that the Schrödinger equation can be separated and the wavefunction

can be factored into the product of two functions X and Y is to note that, because X is independent of y and Y is independent

of x, we can write

2 2

2 2

2 2

2 2

2 2

2 2

Figure 8A.6 A two-dimensional square well The particle is

confined to the plane bounded by impenetrable walls As soon

as it touches the walls, its potential energy rises to infinity

Trang 8

8A Translation  323

Each of the two ordinary differential equations in eqn

8A.14 is the same as the one-dimensional particle-in-a-box

Schrödinger equation (Section 8A.2) The boundary

condi-tions are also the same, apart from the detail of requiring

X(x) to be zero at x = 0 and L1, and Y(y) to be zero at y = 0

and L2 We can therefore adapt eqn 8A.7a without further

calculation:

X x n1 L2 n x L 0 x L

1

1 2 1

Similarly, because E = E X + E Y, the energy of the particle is

lim-ited to the values

E n n1 2 n L n L h m

1

1

2 2

with the two quantum numbers taking the values n1 = 1, 2, …

and n2 = 1, 2, … independently The state of lowest energy is

(n1 = 1, n2 = 1) and E1,1 is the zero-point energy

Some of the wavefunctions are plotted as contours in Fig 8A.7 They are the two-dimensional versions of the wavefunc-tions shown in Fig 8A.3 Whereas in one dimension the wave-functions resemble states of a vibrating string with ends fixed,

in two dimensions the wavefunctions correspond to vibrations

of a rectangular plate with fixed edges

The first term on the left, (1/X)(d2X/dx2), is independent of y,

so if y is varied only the second term on the left, (1/Y)(d2Y/dy2),

can change But the sum of these two terms is a constant,

2mE/2, given by the right-hand side of the equation Therefore,

if the second term did change, then the right-hand side could

not be constant Consequently, even the second term cannot

change when y is changed In other words, the second term,

(1/Y)(d2Y/dy2), is a constant, which we write −2mE Y/2 By a

similar argument, the first term, (1/X)(d2X/dx2), is a constant

when x changes, and we write it −2mE X/2, with E = E X + E Y

Therefore, we can write

These expressions rearrange into the two ordinary (that is,

single-variable) differential equations

−2m2 dd2x X2 =E X X −2m2 dd2y Y2 =E Y Y (8A.14)

Two­dimensional box

energy levels (8A.15b)

Brief illustration 8A.3 The distribution of a particle in a two-dimensional box

Consider an electron confined to a square cavity of side L, and

in the state with quantum numbers n1 = 1, n2 = 2 Because the probability density is

ψ1 22

, (x y, )=L sin πL xsin  Lπy

the most probable locations correspond to sin2(πx/L) = 1 and

sin2(2πx/L) = 1, or (x,y) = (L/2, L/4) and (L/2, 3L/4) The least

probable locations (the nodes, where the wavefunction passes through zero) correspond to zeroes in the probability density

within the box, which occur along the line y = L/2.

Self-test 8A.5 Determine the most probable locations of an

electron in a square cavity of side L when it is in the state with quantum numbers n1 = 2, n2 = 3

Answer: points (x = L/4 and 3L/4; y = L/6, L/2, and 5L/6)

(8A.15a)

Two­

dimensional box

functions

Trang 9

324 8 The quantum theory of motion

We treat a particle in a three-dimensional box in the same

way The wavefunctions have another factor (for the

z-depend-ence), and the energy has an additional term in n L3/ Solution 3

of the Schrödinger equation by the separation of variables

tech-nique then gives

1 2 3

1 2 1 1

2 2 , ,

3 3

πfor

Three­dimensional box wavefunctions (8A.16a)

E n n n1 2 3 n L n L n L h m

1

1

2 2

3 3

The quantum numbers n1, n2, and n3 are all positive

inte-gers 1, 2, … that can be varied independently The system has a

zero-point energy (E1,1,1 = 3h2/8mL2 for a cubic box)

(b) Degeneracy

A special feature of the solutions arises when a

two-dimen-sional box is not merely rectangular but square, with L1 = L2 = L

Then the wavefunctions and their energies are

8

, =( + ) Square box energy levels (8.17b)

Consider the cases n1 = 1, n2 = 2 and n1 = 2, n2 = 1:

2 2

Although the wavefunctions are different, they have the same

energy The technical term for different wavefunctions

corres-ponding to the same energy is degeneracy, and in this case we

say that the state with energy 5h2/8mL2 is ‘doubly degenerate’

In general, if N wavefunctions correspond to the same energy,

then we say that the state is ‘N-fold degenerate’.

The occurrence of degeneracy is related to the symmetry

of the system Figure 8A.8 shows contour diagrams of the two

degenerate functions ψ1,2 and ψ2,1 Because the box is square,

one wavefunction can be converted into the other simply by

rotating the plane by 90° Interconversion by rotation through

90° is not possible when the plane is not square, and ψ1,2 and

ψ2,1 are then not degenerate Similar arguments account for the degeneracy of states in a cubic box Other examples of degener-acy occur in quantum mechanical systems (for instance, in the hydrogen atom, Topic 9A), and all of them can be traced to the symmetry properties of the system

8A.4 Tunnelling

If the potential energy of a particle does not rise to infinity

when it is in the wall of the container, and E < V, the

wavefunc-tion does not decay abruptly to zero If the walls are thin (so that the potential energy falls to zero again after a finite dis-tance), then the wavefunction oscillates inside the box, varies

Three­

dimensional box

energy levels (8A.16b)

Brief illustration 8A.4 Degeneracies in a dimensional box

two-The energy of a particle in a two-dimensional square box of

side L in the state with n1 = 1, n2 = 7 is

E1 7 2 2 mL h2 mL h

2

2 2

, =( + ) =

This state is degenerate with the state with n1 = 7 and n2 = 1

Thus, at first sight the energy level 50h2/8mL2 is doubly ate However, in certain systems there may be states that are not apparently related by symmetry but are ‘accidentally’

degener-degenerate Such is the case here, for the state with n1 = 5 and

n2 = 5 also has energy 50h2/8mL2 Accidental degeneracy is also encountered in the hydrogen atom (Topic 9A)

Self-test 8A.6 Find a state (n1, n2) for a particle in a rectangular

box with sides of length L1 = L and L2 = 2L that is accidentally

degenerate with the state (4,4)

Answer: (n1 = 2, n2 = 8)

Square box wave-functions (8.17a)

+–

Figure 8A.8 The wavefunctions for a particle confined to a square well Note that one wavefunction can be converted into the other by rotation of the box by 90° The two functions correspond to the same energy True degeneracy is a

consequence of symmetry

Trang 10

8A Translation  325

smoothly inside the region representing the wall, and oscillates

again on the other side of the wall outside the box (Fig 8A.9)

Hence the particle might be found on the outside of a container

even though according to classical mechanics it has insufficient

energy to escape Such leakage by penetration through a

classi-cally forbidden region is called tunnelling.

The Schrödinger equation is used to calculate the probability

of tunnelling of a particle of mass m incident from the left on

a rectangular potential energy barrier that extends from x = 0

to x = L On the left of the barrier (x < 0) the wavefunctions are

those of a particle with V = 0, so from eqn 8A.2 we can write

ψ = Aeikx+Be− ikx k=(2mE)1 2 /

The Schrödinger equation for the region representing the

bar-rier (0 ≤ x ≤ L), where the potential energy is the constant V, is

2m2 d2dψ x( )2x +V x E x ψ( )= ψ( ) (8A.19)

We shall consider particles that have E < V (so, according to

classical physics, the particle has insufficient energy to pass

through the barrier), and therefore for which V − E > 0 The

general solutions of this equation are

ψ=Ceκ x+De−κ x κ={ (2m V E− )}1 2 /

as can be verified by differentiating ψ twice with respect to x

The important feature to note is that the two exponentials in

eqn 8A.20 are now real functions, as distinct from the complex,

oscillating functions for the region where V = 0 To the right of

the barrier (x > L), where V = 0 again, the wavefunctions are

ψ = A′eikx k=(2mE)1 2 /

Note that to the right of the barrier, the particle can only be moving to the right and therefore terms of the form e−ikx do not contribute to the wavefunction in eqn 8A.21

The complete wavefunction for a particle incident from the left consists of (Fig 8A.10):

an incident wave (Ae ikx corresponds to positive momentum);

a wave reflected from the barrier (Be −ikx corresponds

to negative momentum, motion to the left);

• the exponentially changing amplitudes inside the barrier (eqn 8A.20);

• an oscillating wave (eqn 8A.21) representing the propagation of the particle to the right after tunnelling through the barrier successfully

The probability that a particle is travelling towards positive x (to the right) on the left of the barrier (x < 0) is proportional to

|A|2, and the probability that it is travelling to the right on the

right of the barrier (x > L) is |A′|2 The ratio of these two

prob-abilities, |A′|2/|A|2, which reflects the probability of the particle

tunnelling through the barrier, is called the transmission

prob-ability, T.

To determine the relationship between |A′|2 and |A|2, we need to investigate the relationships between the coefficients

A, B, C, D, and A′ Since the acceptable wavefunctions must

be continuous at the edges of the barrier (at x = 0 and x = L,

function inside barrier

wave-(8A.20)

(8A.18)

Particle in a rectangular barrier

function left of barrier

wave-(8A.21)

Particle in a rectangular barrier

function right of barrier

Incident wave

Transmitted wave Reflected

wave

x

Figure 8A.10 When a particle is incident on a barrier from the left, the wavefunction consists of a wave representing linear momentum to the right, a reflected component representing momentum to the left, a varying but not oscillating component inside the barrier, and a (weak) wave representing motion to the right on the far side of the barrier

Figure 8A.9 A particle incident on a barrier from the left

has an oscillating wave function, but inside the barrier there

are no oscillations (for E < V) If the barrier is not too thick,

the wavefunction is nonzero at its opposite face, and so

oscillates begin again there (Only the real component of the

Trang 11

326 8 The quantum theory of motion

After straightforward but lengthy algebraic manipulations of

the above set of equations 8A.22 (see Problem 8A.6), we find

where ε = E/V This function is plotted in Fig 8A.12 The

trans-mission probability for E > V is shown there too The

transmis-sion probability has the following properties:

T ≈ 0 for E ≪ V;

T increases as E approaches V: the probability of

tunnelling increases;

T approaches, but is still less than, 1 for E > V: there

is still a probability of the particle being reflected by the barrier even when classically it can pass over it;

T ≈ 1 for E ≫ V, as expected classically.

For high, wide barriers (in the sense that κL ≫ 1), eqn 8A.23a simplifies to

T≈16 (1 )eεε − 2κ L

The transmission probability decreases exponentially with the

thickness of the barrier and with m1/2 It follows that particles of low mass are more able to tunnel through barriers than heavy ones (Fig 8A.13) Tunnelling is very important for electrons

and muons (mµ ≈ 207 me), and moderately important for

pro-tons (mp ≈ 1840me); for heavier particles it is less important

A number of effects in chemistry (for example, the dependence of some reaction rates) depend on the ability of the proton to tunnel more readily than the deuteron The very rapid equilibration of proton transfer reactions is also a mani-festation of the ability of protons to tunnel through barriers and transfer quickly from an acid to a base Tunnelling of protons between acidic and basic groups is also an important feature of the mechanism of some enzyme-catalysed reactions

isotope-Rectangular potential barrier

transmission probability (8A.23a)

Rectangular potential barrier; κL ≫1

transmission probability (8A.23b)

con-from the acid is computed using 8A.23a, with ε = E/V = 1.995 eV/2.000 eV = 0.9975 and V − E = 0.005 eV (corresponding to

Figure 8A.11 The wavefunction and its slope must be

continuous at the edges of the barrier The conditions for

continuity enable us to connect the wavefunctions in the three

zones and hence to obtain relations between the coefficients

that appear in the solutions of the Schrödinger equation

0 0.2 0.4

Incident energy, E/V

2

10

0.6 0.8 1

Figure 8A.12 The transmission probabilities for passage

through a rectangular potential barrier The horizontal axis

is the energy of the incident particle expressed as a multiple

of the barrier height The curves are labelled with the value

of L(2mV)1/2/ The graph on the left is for E < V and that on

the right for E > V Note that T > 0 for E < V whereas classically

T would be zero However, T < 1 for E > V, whereas classically T

Trang 12

8A Translation  327

A problem related to tunnelling is that of a particle in a

square-well potential of finite depth (Fig 8A.14) In this kind

of potential, the wavefunction penetrates into the walls, where

it decays exponentially towards zero, and oscillates within the

well The wavefunctions are found by ensuring, as in the

discus-sion of tunnelling, that they and their slopes are continuous at

the edges of the potential Some of the lowest energy solutions

are shown in Fig 8A.15 A further difference from the solutions

for an infinitely deep well is that there is only a finite number

of bound states Regardless of the depth and length of the well,

however, there is always at least one bound state Detailed

con-sideration of the Schrödinger equation for the problem shows

that in general the number of levels is equal to N, with

where V is the depth of the well and L is its length We see that

the deeper and wider the well, the greater the number of bound states As the depth becomes infinite, so the number of bound states also becomes infinite, as we have already seen

Checklist of concepts

quantized

only certain wavefunctions are acceptable and

there-fore restricts observables to discrete values

half-integer) that labels the state of the system

irremovable minimum energy

mechanics emerges from quantum mechanics as high

quantum numbers are reached

three-dimensional box is the product of wavefunctions for the particle in a one-dimensional box

box is the sum of energies for the particle in two or three one-dimensional boxes

two-dimen-sional box corresponds to the state with quantum

num-bers (n1 = 1, n2 = 1); for three dimensions, (n1 = 1, n2 = 1,

n3 = 1)

cor-respond to the same energy

The larger the value of L(2mV)1/2/ (here, 31) the smaller is the

value of T for energies close to, but below, the barrier height.

Self-test 8A.7 Suppose that the junction between two

semi-conductors can be represented by a barrier of height 2.00 eV

and length 100 pm Calculate the probability that an electron

of energy 1.95 eV can tunnel through the barrier

Trang 13

328 8 The quantum theory of motion

symmetry of the system

region is called tunnelling.

in the height and width of the potential barrier

than heavy ones

Checklist of equations

Free-particle wavefunctions and

Transmission probability T= +{ (1 eκ L−e−κ L) /2 16 1ε(−ε)}−1 Rectangular potential barrier 8A.23a

T = 16ε(1 − ε)e −2κL High, wide rectangular barrier 8A.23b

Trang 14

8B Vibrational motion

Atoms in molecules and solids vibrate around their mean tions as bonds stretch, compress, and bend Here we consider one particular type of vibrational motion, that of ‘harmonic motion’ in one dimension

posi-8B.1 The harmonic oscillator

A particle undergoes harmonic motion, and is said to be a harmonic oscillator, if it experiences a restoring force propor-

tional to its displacement:

F= −k xf Harmonic motion restoring force (8B.1)

where kf is the force constant: the stiffer the ‘spring’, the greater

the value of kf Because force is related to potential energy by

F = −dV/dx (see Foundations B), the force in eqn 8B.1

corre-sponds to the particle having a potential energy

V x( )=1k x

2 f 2 Parabolic potential energy (8B.2)

when it is displaced through a distance x from its equilibrium

position This expression, which is the equation of a ola (Fig 8B.1), is the origin of the term ‘parabolic potential energy’ for the potential energy characteristic of a harmonic

parab-oscillator The Schrödinger equation for the particle of mass m

is therefore

Contents

brief illustration 8b.1: the vibration of a

example 8b.1: confirming that a wavefunction

is a solution of the schrödinger equation 332

example 8b.2: normalizing a harmonic oscillator

example 8b.3: calculating properties of a

example 8b.4: calculating the tunnelling

probability for the harmonic oscillator 335

➤ Why do you need to know this material?

The detection and interpretation of vibrational frequencies

is the basis of infrared spectroscopy (Topics 12D and 12E)

Molecular vibration plays a role in the interpretation

of thermodynamic properties, such as heat capacities

(Topics 5E and 15F), and of the rates of chemical reactions

(Topic 21C).

➤ What is the key idea?

The quantum mechanical treatment of the simplest model

of vibrational motion, the harmonic oscillator, reveals

that the energy is quantized and the wavefunctions are

products of a polynomial and a Gaussian (bell-shaped)

function.

➤ What do you need to know already?

You should know how to formulate the Schrödinger

equation given a potential energy function You should

also be familiar with the concepts of tunnelling (Topic 8A)

and the expectation value of an observable (Topic 7B).

a harmonic oscillator, where x is the displacement from

equilibrium The narrowness of the curve depends on the force

constant kf: the larger the value of kf, the narrower the well

Trang 15

330 8 The quantum theory of motion

2 2

2mddψ x( )x k xf ψ( )x E x ψ( )

We can anticipate that the energy of an oscillator will be

quan-tized because the wavefunction has to satisfy boundary

con-ditions (as in Topic 8A for a particle in a box): it will not be

found with very large extensions because its potential energy

rises to infinity there That is, when we impose the

bound-ary conditions ψ = 0 at x = ±∞, we can expect to find that only

certain wavefunctions and their corresponding energies are

possible

(a) The energy levels

Equation 8B.3 is a standard equation in the theory of

differen-tial equations and its solutions are well known to

mathemati-cians.1 The permitted energy levels are

where ω (omega) is the frequency of oscillation of a classical

harmonic oscillator of the same mass and force constant Note

that ω is large when the force constant is large and the mass

small It follows that the separation between adjacent levels is

which is the same for all v Therefore, the energy levels form

a uniform ladder of spacing ω (Fig 8B.2) The energy

sepa-ration ω is negligibly small for macroscopic objects (with

large mass) for which classical mechanics is adequate for

describing vibrational motion; however, the energy tion is of great importance for objects with mass similar to that of atoms

separa-Because the smallest permitted value of v is 0, it follows from

eqn 8B.4 that a harmonic oscillator has a zero-point energy

E0 1 2

= ω Harmonic oscillator Zero-point energy (8B.6)

The mathematical reason for the zero-point energy is that

v cannot take negative values, for if it did the wavefunction would not obey the boundary conditions The physical reason

is the same as for the particle in a box (Topic 8A): the cle is confined, its position is not completely uncertain, and therefore its momentum, and hence its kinetic energy, cannot

parti-be exactly zero We can picture this zero-point state as one in which the particle fluctuates incessantly around its equilib-rium position; classical mechanics would allow the particle to

be perfectly still

Atoms vibrate relative to one another in molecules with the bond acting like a spring The question then arises as to what mass to use to predict the frequency of the vibration

In general, the relevant mass is a complicated combination

of the masses of all the atoms that move, with each bution weighted by the amplitude of the atom’s motion That amplitude depends on the mode of motion, such as whether the vibration is a bending motion or a stretching motion, so each mode of vibration has a characteristic ‘effective mass’ For a diatomic molecule AB, however, for which there is only one mode of vibration, corresponding to the stretching and

contri-compression of the bond, the effective mass, μ, has a very

simple form:

μ = m m m mA+ B

A B Diatomic molecule effective mass (8B.7)

When A is much heavier than B, mB can be neglected in the

denominator and the effective mass is μ ≈ mB, the mass of the lighter atom This result is plausible, for in the limit of the heavy atom being like a brick wall, only the lighter atom moves and hence determines the vibrational frequency

1 For details, see our Molecular quantum mechanics, Oxford University

8 v

Potential energy

Figure 8B.2 The energy levels of a harmonic oscillator are

evenly spaced with separation ħω, with ω = (kf /m)1/2 Even in its

lowest energy state, an oscillator has an energy greater than

which is close to the mass of the proton The force constant of

the bond is kf = 516.3 N m−1 It follows from eqn 8B.4, with μ in place of m, that

(8B.4)

Harmonic oscillator energy levels

(8B.3)

Harmonic oscillator schrödinger equation

Trang 16

8B Vibrational motion  331

The result in Brief illustration 8B.1 implies that excitation

requires radiation of frequency ν = ΔE/h = 90 THz and

wave-length λ = c/ν = 3.3 µm It follows that transitions between

adja-cent vibrational energy levels of molecules are stimulated by or

emit infrared radiation (Topics 12D and 12E)

(b) The wavefunctions

Like the particle in a box (Topic 8A), a particle undergoing

har-monic motion is trapped in a symmetrical well in which the

potential energy rises to large values (and ultimately to infinity)

for sufficiently large displacements (compare Figs 8A.1 and

8B.1) However, there are two important differences:

• Because the potential energy climbs towards infinity

only as x2 and not abruptly, the wavefunction

approaches zero more slowly at large displacements

than for the particle in a box

• As the kinetic energy of the oscillator depends on

the displacement in a more complex way (on account

of the variation of the potential energy), the

curvature of the wavefunction also varies in a more

complex way

The detailed solution of eqn 8B.3 confirms these points

and shows that the wavefunctions for a harmonic oscillator

have the form

where N is a normalization constant A Gaussian function is a

bell-shaped function of the form e−x2

(Fig 8B.3) The precise form of the wavefunctions is

The factor H v (y) is a Hermite polynomial; the form of these

poly-nomials and some of their properties are listed in Table 8B.1 Hermite polynomials, which are members of a class of functions called ‘orthogonal polynomials’, have a wide range of important properties which allow a number of quantum mechanical cal-culations to be done with relative ease Note that the first few

Hermite polynomials are very simple: for instance, H0(y) = 1 and

(8B.9a)

Harmonic oscillator ground-state wavefunction

or 563.4 THz (We have used 1 N = 1 kg m s−2.) Therefore the

separation of adjacent levels is (eqn 8B.5)

E v+1− =E v ( 1 54 57 10 × 0− 34Js) ( ×5 634×1014s) =5 941 1× 0− 2 0J

or 59.41 zJ, about 0.37 eV This energy separation corresponds

to 36 kJ mol−1, which is chemically significant The zero-point

energy, eqn 8B.6, of this molecular oscillator is 29.71 zJ, which

corresponds to 0.19 eV, or 18 kJ mol−1

Self-test 8B.1 Suppose a hydrogen atom is adsorbed on the

surface of a gold nanoparticle by a bond of force constant

855 N m−1 Calculate its zero-point vibrational energy

6 64y6 − 480y4 + 720y2 − 120

* The Hermite polynomials are solutions of the differential equation

H v″− 2yH v′+ 2v H v= 0 where primes denote differentiation They satisfy the recursion relation

Trang 17

332 8 The quantum theory of motion

and the corresponding probability density is

( )x N= e−y =N e−x/

The wavefunction and the probability density are shown in Fig

8B.4 Both curves have their largest values at zero displacement

(at x = 0), so they capture the classical picture of the zero-point

energy as arising from the ceaseless fluctuation of the particle

about its equilibrium position

The wavefunction for the first excited state of the oscillator,

the state with v = 1, is

This function has a node at zero displacement (x = 0), and the

probability density has maxima at x = ±α (Fig 8B.5).

The shapes of several wavefunctions are shown in Fig 8B.6 and the corresponding probability densities are shown in Fig 8B.7 At high quantum numbers, harmonic oscillator wave-functions have their largest amplitudes near the turning points

of the classical motion (the locations at which V = E, so the

kinetic energy is zero) We see classical properties emerging in the correspondence principle limit of high quantum numbers (Topic 8A), for a classical particle is most likely to be found at the turning points (where it travels most slowly) and is least likely to be found at zero displacement (where it travels most rapidly)

Figure 8B.4 The normalized wavefunction and probability

density (shown also by shading) for the lowest energy state of a

Figure 8B.5 The normalized wavefunction and probability

density (shown also by shading) for the first excited state of a

Method Substitute the wavefunction given in eqn 8B.9a

into eqn 8B.3 Use the definition of α given in eqn 8B.8

to determine the energy on the right-hand side of eqn 8B.3 and confirm that it matches the zero-point energy given in eqn 8B.6

Answer We need to evaluate the second derivative of the ground-state wavefunction:

2 2

1 2 0

The second and third terms on the left-hand side (in blue)

can-cel and we obtain E=1 k m

ground-state probability density

(8B.10)

Harmonic oscillator

First state wavefunction

Trang 18

excited-8B Vibrational motion  333

Note the following features of the wavefunctions:

• The Gaussian function goes very strongly to zero as

the displacement increases (in either direction,

extension or compression), so all the wavefunctions

approach zero at large displacements

The exponent y2 is proportional to x2 × (mkf)1/2, so

the wavefunctions decay more rapidly for large

masses and stiff springs

As v increases, the Hermite polynomials become

larger at large displacements (as x v), so the

wavefunctions grow large before the Gaussian

function damps them down to zero: as a result, the

wavefunctions spread over a wider range as v

increases (Fig 8B.7)

8B.2 The properties of oscillators

The average value of a property is calculated by evaluating the expectation value of the corresponding operator (eqn 7C.11,

〈 〉 = ∫ψ Ωψ τ*  d ) Now that we know the wavefunctions of the harmonic oscillator, we can start to explore its properties by evaluating integrals of the type

Figure 8B.6 The normalized wavefunctions for the first

five states of a harmonic oscillator Note that the number

of nodes is equal to v and that alternate wavefunctions

are symmetrical or asymmetrical about y = 0 (zero

displacement)

v

0 1 2 3 4

18

Probability

density

Displacement

Figure 8B.7 The probability densities for the first five

states of a harmonic oscillator and the state with v = 18

Note how the regions of highest probability density move

towards the turning points of the classical motion as v

Method Normalization is carried out by evaluating the

inte-gral of |ψ|2 over all space and then finding the normalization factor from eqn 7B.3 (N = ∫1/(ψ ψ τ* d ) )1 2 / The normalized

wavefunction is then equal to Nψ In this one-dimensional problem, the volume element is dx and the integration is

from −∞ to +∞ The wavefunctions are expressed in terms of

the dimensionless variable y = x/α, so begin by expressing the integral in terms of y by using dx = αdy The integrals required

are given in Table 8B.1

Answer The unnormalized wavefunction is

a box, for a harmonic oscillator N v is different for each value

of v.

Self-test 8B.3 Confirm, by explicit evaluation of the integral,

that ψ0 and ψ1 are orthogonal

Answer: Show that ∫−∞∞ ψ ψ0 1* dx= 0 by using the information in

Table 8B.1

Trang 19

334 8 The quantum theory of motion

(a) Mean values

We show in the following example that the mean displacement,

〈x〉, and the mean square displacement, 〈x2〉, of the oscillator

when it is in the state with quantum number v are

〈 〉x =0 Harmonic oscillator mean displacement (8B.12a)

The result for 〈x〉 shows that the oscillator is equally likely

to be found on either side of x = 0 (like a classical oscillator)

The result for 〈x2〉 shows that the mean square displacement

increases with v This increase is apparent from the probability

densities in Fig 8B.7, and corresponds to the classical

ampli-tude of swing increasing as the oscillator becomes more highly

(v+1)

2 ω, it follows that

〈 〉V =1E

2 v Harmonic oscillator mean potential energy (8B.13b)

The total energy is the sum of the potential and kinetic gies, so it follows at once that the mean kinetic energy of the oscillator is (as could also be shown using the kinetic energy operator)

ener-〈 〉Ek =1E

2 v Harmonic oscillator mean kinetic energy (8B.13c)

The result that the mean potential and kinetic energies of a harmonic oscillator are equal (and therefore that both are

equal to half the total energy) is a special case of the virial theorem:

If the potential energy of a particle has the form

V = ax b, then its mean potential and kinetic energies are related by

Example 8B.3 Calculating properties of a harmonic

oscillator

Consider the harmonic oscillator motion of the HeCl

mol-ecule in Brief illustration 8B.1 Calculate the mean

displace-ment of the oscillator when it is in a state with quantum

number v.

Method Normalized wavefunctions must be used to

calcu-late the expectation value The operator for position along x is

multiplication by the value of x (Topic 7C) The resulting

inte-gral can be evaluated either

• by inspection (the integrand is the product of an odd

and an even function), or

• by explicit evaluation using the formulas in Table 8B.1

The former procedure makes use of the definitions that an

even function is one for which f(−x) = f(x) and an odd

func-tion is one for which f(−x) = −f(x) Therefore, the product of an

odd and even function is itself odd, and the integral of an odd

function over a symmetrical range about x = 0 is zero The

lat-ter procedure using explicit integration is illustrated here to

give practice in the calculation of expectation values We shall

need the relation x = αy, which implies that dx = αdy.

Answer The integral we require is

Both integrals are zero (See Table 8B.1), so 〈x〉 = 0 The mean

displacement is zero because displacements on either side of the equilibrium position occur with equal probability

Self-test 8B.4 Calculate the mean square displacement, 〈x2〉,

of the H − Cl bond distance from its equilibrium position by using the recursion relation in Table 8B.1 twice

Answer: (v+ × 1 )

2 115pm ; eqn 9.12b, with μ in place of m2

(8B.12b)

Harmonic oscillator mean square displacement

(8B.13a)

Harmonic oscillator mean potential energy

Trang 20

8B Vibrational motion  335

For a harmonic oscillator b = 2, so 〈Ek〉 = 〈V〉, as we have found

The virial theorem is a short cut to the establishment of a

num-ber of useful results, and we use it elsewhere (for example, in

Topic 9A)

(b) Tunnelling

An oscillator may be found at extensions with V > E, which are

forbidden by classical physics, for they correspond to negative

kinetic energy; this is an example of the phenomenon of

tun-nelling (Topic 8A) As shown in Example 8B.4, for the lowest

energy state of the harmonic oscillator, there is about an 8 per

cent chance of finding the oscillator stretched beyond its

clas-sical limit and an 8 per cent chance of finding it with a

classi-cally forbidden compression These tunnelling probabilities are

independent of the force constant and mass of the oscillator

Example 8B.4 Calculating the tunnelling probability for

the harmonic oscillator

Calculate the probability that the ground-state harmonic

oscillator will be found in a classically forbidden region

Method Find the expression for the classical turning point,

xtp, where the kinetic energy vanishes, by equating the

poten-tial energy to the total energy E of the harmonic oscillator The

probability of finding the oscillator stretched beyond a

dis-placement xtp is the sum of the probabilities ψ 2dx of finding

it in any of the intervals dx lying between xtp and infinity, so

evaluate the integral

x

=∫∞ψ v2d

tp

The variable of integration is best expressed in terms of y = x/α

and the integral to be evaluated is a special case of the error

function, erf z, defined as

z

= −1 21 2∫∞ − 2

and evaluated for some values of z in Table 8B.2 (this function

is commonly available in mathematical software packages)

By symmetry, the probability of being found stretched into a

classically forbidden region is the same as that of being found

compressed into a classically forbidden region

Answer According to classical mechanics, the turning point,

xtp, of an oscillator occurs when its kinetic energy is zero,

which is when its potential energy 1

2k xf 2 is equal to its total

energy E This equality occurs when

For the state of lowest energy (v = 0), ytp = 1 and the probability

of being beyond that point is

tp

The normalization constant N0 is calculated from the

expres-sion for N v in Example 8B.2 ( N v=1/(α π1 2 /2v v!)1 2 /):

N0 1 2 0

1 2

1 2

1 21

The integral in the expression for P is written in terms of the

error function erf(1) as

1

1 2 1

In 7.9 per cent of a large number of observations, any

oscil-lator in the state with quantum number v = 0 will be found

stretched to a classically forbidden extent There is the same probability of finding the oscillator with a classically forbid-den compression The total probability of finding the oscilla-tor tunnelled into a classically forbidden region (stretched or compressed) is about 16 per cent

Self-test 8B.5 Calculate the probability that a harmonic

oscil-lator in the state with quantum number v = 1 will be found at

a classically forbidden extension (Follow the argument given

in Example 8B.4 and use the method of integration by parts (Mathematical background 1) to obtain an integral that can be

expressed in terms of the error function.)

Answer: P = 0.056

Ngày đăng: 16/07/2017, 13:23

TỪ KHÓA LIÊN QUAN

TÀI LIỆU CÙNG NGƯỜI DÙNG

TÀI LIỆU LIÊN QUAN

w