lecture 10 example: particle in a box

13
Lecture #10 1 Lecture 10 Objectives: 1. Be able to solve the particle in a box problem 2. Compare the classical and quantum harmonic oscillators 3. Compare the classical and quantum rotors Example: Particle in a box Consider a particle trapped in a one-dimensional box, of length L. The Schr¨odinger equation for this system (considering only the spatial function) is - h 2 8π 2 m d 2 ψ dx 2 + = Eψ, where U = 0 when x is inside the box and U = for x outside the box. This is a second order ordinary differential equation. The general solution to this equation is ψ = α sin γx + β cos γx with γ = (8π 2 m(E - U )/h 2 ) 1/2 We need two boundary conditions to fix α and β . What are they? ψ(0) = 0, and ψ(L) = 0. These BCs give ψ(0) = β =0 and ψ(L)= α sin γL + β cos γL =0 hence, either α = 0(trivial solution) or sin γL = 0. Taking the latter we find γL = γ = L Hence, ψ = α n sin nπx L The values of α n can be computed by recalling that 0 ψ * ψdx = 1 so that 0 ψ * ψdx = L 0 ψ * ψdx = α * n α n L 0 sin 2 nπx L =1 Recalling that sin 2 udu = u/2 - sin(2u)/4+ C we have that α * n α n L u 2 - sin(2u) 4 0 = α * n α n L 2 - sin(2) 4 =1 Therefore, α n = i 2 L 1/2 . We can now solve for the energy levels by from the equation L 0 ψ * Hψdx = L 0 ψ * Eψdx = E L 0 ψ * ψdx = E L 0 ψ * Hψdx = - h 2 8π 2 m - L 2 2 L L 0 sin 2 nπx L dx = E n Therefore, E n = nh L 2 1 8m = n 2 h 2 8mL 2 .

Upload: dangdung

Post on 11-Feb-2017

215 views

Category:

Documents


1 download

TRANSCRIPT

Page 1: Lecture 10 Example: Particle in a box

Lecture #10 1

Lecture 10

Objectives:

1. Be able to solve the particle in a box problem

2. Compare the classical and quantum harmonic oscillators

3. Compare the classical and quantum rotors

Example: Particle in a box

Consider a particle trapped in a one-dimensional box, of length L. The Schrodinger equation forthis system (considering only the spatial function) is

−h2

8π2m

d2ψ

dx2+ Uψ = Eψ,

where U = 0 when x is inside the box and U = ∞ for x outside the box.This is a second order ordinary differential equation. The general solution to this equation is

ψ = α sin γx+ β cos γx

with γ = (8π2m(E − U)/h2)1/2 We need two boundary conditions to fix α and β. What are they?ψ(0) = 0, and ψ(L) = 0. These BCs give

ψ(0) = β = 0

and

ψ(L) = α sin γL+ β cos γL = 0

hence, either α = 0(trivial solution) or sin γL = 0. Taking the latter we find

γL = nπ → γ =nπ

L

Hence,

ψ = αn sin

(

nπx

L

)

The values of αn can be computed by recalling that∫

0 ψ∗ψdx = 1 so that∫

0ψ∗ψdx =

∫ L

0ψ∗ψdx = α∗

nαn

∫ L

0sin2

(

nπx

L

)

= 1

Recalling that∫

sin2udu = u/2− sin(2u)/4 + C we have that

α∗

nαnL

[

u

2−

sin(2u)

4

]∣

0= α∗

nαnL

[

2−

sin(2nπ)

4

]

= 1

Therefore, αn = i(

2L

)1/2. We can now solve for the energy levels by from the equation

∫ L

0ψ∗Hψdx =

∫ L

0ψ∗Eψdx = E

∫ L

0ψ∗ψdx = E

∫ L

0ψ∗Hψdx =

(

−h2

8π2m

)

(

−nπ

L

)2 ( 2

L

)∫ L

0sin2

(

nπx

L

)

dx = En

Therefore, En =(

nhL

)218m = n2h2

8mL2 .

Page 2: Lecture 10 Example: Particle in a box

Lecture #10 2

The Harmonic Oscillator

The Classical Harmonic Oscillator

A vibrating body subject to a restoring force, which increases in proportion to the displacementfrom equilibrium, will undergo harmonic motion at constant frequency and is called a harmonicoscillator. Figure 1(a) shows one example of a harmonic oscillator, where a body of mass m isconnected to a fixed support by a spring with a force constant, k. We will assume that gravitationalforces are absent.

Figure 1: (a) Harmonic Oscillator Consisting of a Mass Connected by a Spring to a Fixed Support;(b) Potential Energy, V ,and Kinetic Energy, EK For the Harmonic Oscillator.

When the system is at equilibrium, the mass will be at rest, and at this point the displacement,x, from equilibrium has the value zero. As the mass is pulled to the right, there will be a restoringforce, f , which is proportional to the displacement. For a spring obeying Hooke’s law,

f = −kx = md2x

dt2(1)

The minus sign in equation 1 is related to the fact that the force will be negative, since the masswill tend to be pulled toward the −x direction when the force is positive. From Newton’s secondlaw, the force will be equal to the mass multiplied by the acceleration, d2x

dt2. The equation of motion

is a second order ordinary differential equation, obtained by rearranging equation 1 as

d2x

dt2+k

mx = 0, (2)

and has a general solution given by

x(t) = A sinωt+B cosωt, (3)

where

ω = (k/m)1/2. (4)

The units of ω are radians s−1, and since there are 2π radians/cycle, the frequency ν = ω/2π cycless−1. [Note: The student should check this solution by substituting equation 3 back into equation 2].

We again require two boundary conditions to specify the constants A and B. We choose themass to be at x = 0 moving with a velocity v0 at time= 0. The first condition gives

x(t = 0) = A · 0 +B · 1 = B = 0 (5)

Page 3: Lecture 10 Example: Particle in a box

Lecture #10 3

so that

x(t) = A sinωt. (6)

Using this result, the second boundary condition can be written as

v0 = v(t = 0) =dx

dt

t=0= Aω cosωt|t=0 = Aω, (7)

from which we see that A = v0/ω. As the spring stretches, or contracts, when the mass is undergoingharmonic motion, the potential energy of the system will rise and fall, as the kinetic energy of themass falls and rises. The change in potential energy, dV , is

dV = −fdx = kxdx, (8)

so that upon integrating,

V =1

2kx2 + constant. (9)

The constant of integration may be set equal to zero. This potential energy function is shown asthe parabolic line in Figure 1(b). The kinetic energy of this harmonic oscillator is given by

EK =1

2mv2 =

1

2m

(

dx

dt

)2

=m

2(Aω)2 cos2 ωt (10)

This function is also plotted in Figure 1(b). The total constant energy, E, of the system is givenby

E = V + EK =1

2kx2 +

m

2(Aω)2 cos2 ωt =

1

2kA2 sin2 ωt+

m

2(Aω)2 cos2 ωt, (11)

where we have substituted equation 6 for x. Substituting equation 4 for ω2, we can write

E =1

2A2k

(

sin2 ωt+ cos2 ωt)

=1

2A2k (12)

The total energy is thus a constant and is shown as a horizontal line in Figure 1(b). At the limits ofoscillation, where the mass is reversing its direction of motion, its velocity will be momentarily equalto zero, and its momentary kinetic energy will therefore also be zero, meaning that the potentialenergy will be maximized and equal to the total energy of the system at the two turning points.As the oscillator begins to undergo acceleration away from the turning points, the kinetic energywill increase, and the potential energy will decrease along the curve, V , as shown in Figure 1(b).

If the spring constant, k, should not be constant, but should vary slightly from a constantvalue as x changes, we would be dealing with an anharmonic oscillator . In most cases, ananharmonic oscillator may be closely approximated by the harmonic oscillator equations for smalldisplacements from equilibrium, x.

Soon we will be comparing the quantum harmonic oscillator with the classical harmonic oscil-lator, and the probability of finding the mass at various values of x will be of interest. We nowcalculate this probability for the classical harmonic oscillator.

The probability of finding the mass, m, at any given value of x is inversely proportional to thevelocity, v, of the mass. This is reasonable, since the faster the mass is moving, the less likely itis to observe the mass. Hence, we expect that the probability of observing the mass will be have

Page 4: Lecture 10 Example: Particle in a box

Lecture #10 4

a minimum at x = 0, where the velocity is at a maximum, and conversely will exhibit maximawhen x = ±A. From equation 7 we see that

v(x) =dx

dt= Aω cosωt, (13)

so that

P(x)dx ∝dx

Aω cosωt, (14)

where P(x)dx is the probability of finding the mass between x and x + dx. Note that since x isa continuous variable that we define P(x)dx as the probability and P(x) is called the probabilitydensity , which is the probability per unit length, in this case. We note that at the turning pointsof the oscillation, when 1/4 or 3/4 of a cycle has occurred, that t = π

2ω or t = 3π2ω and at these points

cosωt goes to zero, with 1v going to infinity. We know that the probability of finding the mass at

the end points must be a maximum, but must not be infinite . The reason that P(x) remainsfinite is that dx in equation 14 always has finite width, therefore, P(x) is not defined exactly at apoint. For example, at t = π

2ω , P(x) is evaluated over the range −A < x < −A+ dx.This probability density function as a function of the x-coordinate, P(x), is plotted along with

the velocity, v in Figure 2. The probability density is a smooth function over the range of xavailable to the oscillator and has exactly one minimum at x = 0. We will soon find that thisintuitive classical behavior is not obeyed by the quantized harmonic oscillator. In fact, for thequantum oscillator in the ground state we will find that P(x) has a maximum at x = 0.

0 +A-A

x

P(x)

v(x)

0

Figure 2: Probability Density, P(x), for Classical Harmonic Oscillator at Various Displacements,x. P(x) is plotted as the dashed line and the velocity, v(x) is plotted as the solid curve. The twovertical lines give the limits of the oscillator motion. Note that P(x) ∝ 1

v(x) .

The Quantum Harmonic Oscillator

The quantum harmonic oscillator is a very important problem in quantum mechanics. For example,it serves as a first-order approximation for the bond vibrational problem in diatomic and (with

Page 5: Lecture 10 Example: Particle in a box

Lecture #10 5

coupling) polyatomic molecules. We will examine the quantum harmonic oscillator in some detail,comparing it with what we know about the classical harmonic oscillator from the previous section.The potential energy function for the quantum harmonic function is the same as for the classicalharmonic oscillator, namely, V = 1/2kx2. Thus, in the quantum Hamiltonian is

Hop = EKop + Vop (15)

and we may write the Schrodinger equation as

−h2

2m

∂2ψ

∂x2+

1

2kx2ψ = Eψ (16)

The general solution to this problem (which we will not derive) can be written as

ψn(x) =

(

α

π

)1/4 ( 1

2nn!

)1/2

Hn(y)e−y2/2, (17)

where n = 0, 1, 2, . . . is the quantum number, α = mωh , y = α1/2x, and Hn(y) is a Hermite

polynomial of order n. The first few Hermite polynomials are

H0(y) = 1 (18)

H1(y) = 2y (19)

H2(y) = 4y2 − 2 (20)

H3(y) = 8y3 − 12y (21)

H4(y) = 16y4 − 48y2 + 12 (22)

Hermite polynomials of any order can be calculated from the recursion relation

Hn+1(y) = 2yHn(y)− 2nHn−1(y). (23)

The allowed energies (eigen energies) for the quantum harmonic oscillator are

En =

(

n+1

2

)

hω (24)

and since ω = 2πν,

En =

(

n+1

2

)

hν. (25)

Using equation 4 for ω, we may write

En =

(

n+1

2

)

h

(

k

m

)1/2

. (26)

The ground state wavefunction for the quantum harmonic oscillator can be obtained by substitutingH0(y) from equation 18, using y = α1/2x, into equation 17,

ψ0(x) =

(

α

π

)1/4

e−αx2/2 =

(

)1/4

e−mωx

2

2h . (27)

Likewise, the first and second excited state wavefunctions are

ψ1(x) =

(

4α3

π

)1/4

xe−αx2/2 (28)

ψ2(x) =

(

α

)1/4 (

2αx2 − 1)

e−αx2/2. (29)

Page 6: Lecture 10 Example: Particle in a box

Lecture #10 6

0

x

n=2; E=5/2hn

n=1; E=3/2hn

n=0; E=1/2hn

0

Y2

|Y2|2

|Y1|2

Y1

|Y0|2

Y0

Figure 3: Allowed Energy Levels, ψn, and |ψn|2 for the Quantum Harmonic Oscillator with Quan-

tum Numbers n = 0, 1, 2. The harmonic potential, V (x) = 1/2kx2, is shown, along with the wavefunctions and probability densities for the first three energy levels.

Figure 3 shows the first few allowed energy levels for the quantum harmonic oscillator. Alsoshown are the wavefunctions, ψn and the probability densities, |ψn|

2 for the levels n = 0, 1, 2. Theequally spaced set of allowed vibrational energy levels observed for a quantum harmonic oscillatoris not expected classically, where all energies would be possible. The quantization of theenergy levels of the harmonic oscillator is similar in spirit to the quantization of the energy levelsfor the particle in a box, except that for the harmonic oscillator, the potential energy varies in aparabolic manner with the displacement from equilibrium, and the walls of the “box” thereforeare not vertical. We might say, in comparison to the “hard” vertical walls for a particle in a box,that the walls are “soft” for the harmonic oscillator. In addition, the spacing between the allowedenergy levels for the harmonic oscillator is a constant value, hν, whereas for the particle in a box,the spacing between levels rises as the quantum number, n, increases.

There is another interesting feature seen in Figure 3. For the lowest allowed energy, when n = 0,we see that the quantum harmonic oscillator possesses a zero-point energy of 1

2hν This too isreminiscent of the particle in a box, which displays a finite zero-point energy for the first allowedquantum number, n = 1. This lowest allowed zero-point energy is unexpected on classical grounds,since all vibrational energies, down to zero, are possible in the classical oscillator case.

Recall that we developed an expression for the probability of observing a classical harmonicoscillator between x and x + dx and found that this probability is inversely proportional to thevelocity of the oscillator (see equation 14). The corresponding probability for a quantum harmonicoscillator in state n is proportional to ψ∗

nψn = |ψn|2. We now compare the probability densities of

classical and quantum harmonic oscillators. Recall that the ground state for the classical oscillatorhas zero energy (and zero motion), whereas the quantum oscillator in the ground state has anenergy of 1

2hν. Therefore, it does not make sense to compare classical and quantum oscillators intheir ground states. We can however directly compare probability densities by comparing quantumand classical oscillators having the same energy. Equating the quantum and classical oscillator

Page 7: Lecture 10 Example: Particle in a box

Lecture #10 7

energies, we have

En =

(

n+1

2

)

hν =1

2A2k, (30)

where A is the classical amplitude, or limit of motion. Solving for A, we have

A =

[

2hν

k(2n+ 1)

]1/2

(31)

That is, a classical oscillator with energy En will oscillate between x = ±A, with A given byequation 31. The quantum and classical probability densities for n = 0, 2, 5, and 20 are plotted inFigure 4.

We see from equation 27 that |ψ0|2 is a Gaussian function function with a maximum at x = 0.

This is plotted in the upper left panel of Figure 4. Contrast this behavior with the classical harmonicoscillator, which has a minimum in the probability at x = 0 and maxima at the turning points.Also note that the limits of oscillation are strictly obeyed for the classical oscillator, shown by thevertical lines. In contrast, the probability density for the quantum oscillator “leaks out” beyond thex = ±A classical limits. The quantum harmonic oscillator penetrates beyond the classicalturning point! This phenomenon is akin to the quantum mechanical penetration of afinite barrier seen previously. Thus, the probability densities for the quantum and classicaloscillators for n = 0 have almost opposite shapes and very different behavior. Next, we compare theclassical and quantum oscillators for n = 2 (top right panel in Figure 4). Note that the probabilitydensity for the quantum oscillator now has three peaks. In general, the quantum probability densitywill have n+1 peaks. In addition to having n+1 maxima, the probability density also has nminima.Remarkably, these minima correspond to zero probability! This means that for a particularquantum state n there will be exactly n forbidden locations where the wave functiongoes to zero (nodes). This is very different from the classical case, where the mass can be at anylocation within the limits −A ≤ x ≤ A. Note also that the middle peak centered at x = 0 has asmaller amplitude that the outer two peaks. Thus, for n = 2 we are beginning to see behavior that iscloser in spirit to the classical probability density, that is, the probability of observing the oscillatorshould be greater near the turning points than in the middle. The classical probability density isessentially the same for all energies, but is just “stretched” to larger amplitudes for higher energies.For n = 5, shown in the lower left panel of Figure 4, we see the continued trend that the peaks nearx = 0 are smaller than the peaks near the edges. Note that the probability densities continue toextend past the classical limits of motion, but die off exponentially. Finally, for n = 20 note that thegaps between the peaks in the probability density become very small. At large energies the distancebetween the peaks will be smaller than the Heisenberg uncertainty principle allows for observation.In other words, we will no longer be able to distinguish the individual peaks. The probabilitywill be smeared out. You should be able to see that for n = 20 an appropriate the average ofthe quantum probability density closely approximates the classical behavior probability density.The region of non-zero probability outside the classical limits drops very quickly for high energies,so that this region will also be unobservable as a result of the uncertainty principle. Thus, thequantum harmonic oscillator smoothly crosses over to become a classical oscillator. This crossingover from quantum to classical behavior was called the “correspondence principle” by Bohr.

Page 8: Lecture 10 Example: Particle in a box

Lecture #10 8

00

00

00

00

|0|2

n=0

Pro

babi

lity

Den

sity

Position

|2|2

n=2

|5|2

n=5

|20

|2

n=20

Figure 4: Probability Densities for Quantum and Classical Harmonic Oscillators. The probabilitydensities for quantum harmonic oscillators, |ψn(x)|

2, are plotted as solid lines for the quantumnumbers n = 0, 2, 5, 20. The probability densities of the classical harmonic oscillators having thesame energies as the quantum oscillators are plotted as dashed lines. The classical limits of motionare shown by the vertical lines.

Page 9: Lecture 10 Example: Particle in a box

Lecture #10 9

The Rigid Rotor

The Classical Rigid Rotor—1 Dimensional

We will consider a mass, m, coupled to a fixed point by a weightless rigid rod of length, r, whichmoves in a fixed plane and is not influenced by gravity. Figure 5 shows this mechanical situationand the physical parameters that are of importance. At all values of the angle, φ, the potentialenergy for the rotor is zero, i.e., V (φ) = 0. The moment of inertia is I = mr2. The kinetic energyof the rotor is Erot = Iω2/2, where ω is the angular speed expressed in radians s−1. Classically, allvalues of Erot are allowed. The angular momentum of the rotor is pφ = Iω.

Figure 5: The Rigid Rotor—Planar. The coordinate appropriate to describe the motion of thissystem is the angle, φ

The Quantized Rigid Rotor—Planar

We may write the Schrodinger equation for the planar rigid rotor in exactly the fashion we usedpreviously for problems involving one coordinate, except that the coordinate of importance is theangle, φ, and instead of the mass, m, we employ the moment of inertia, I.

−h2

2I

∂2ψrot

∂φ2= Eψrot. (32)

One solution to this problem is

ψ = Nroteimℓφ (33)

where Nrot is the normalization constant and mℓ is a quantum number, as will be shown later whenwe consider the periodicity of the rotor.

Substituting equation 33 into equation 32, we find that

−h2

2I

(

−m2ℓe

imℓφ)

= Emℓeimℓφ (34)

or,

Emℓ=m2

ℓ h2

2I, mℓ = 0,±1,±2, . . . (35)

Page 10: Lecture 10 Example: Particle in a box

Lecture #10 10

Note that if we solve equation 35 for mℓ, we obtain

mℓ = ±(2IEmℓ

)1/2

h(36)

where mℓ has both positive and negative values. The two signs for the quantum number refer tothe two directions of rotation of the rigid rotor. Note that in this quantum mechanical problem, incontrast to the previous problems we have considered, the zero point energy for the rotor is zero.

The normalized wave functions for the planar rigid rotor may be obtained as shown before.Thus,

N2rot

∫ 2π

0ψ∗

rotψrotdφ = 1 (37)

Using the wave function given in equation 33, and integrating, we find that

N2rot · 2π = 1; Nrot =

(

1

)1/2

(38)

and the normalized wave function is therefore

ψrot =

(

1

)1/2

eimℓφ (39)

The quantization feature of this problem may be found by considering the periodic behavior of thewave function. We recognize that

eimℓφ = eimℓ(φ+2π) (40)

Converting equation 40 to the transcendental form through Euler’s formula, we obtain

cosmℓφ+ i sinmℓφ = cosmℓ(φ+ 2π) + i sinmℓ(φ+ 2π) (41)

This equality will hold only for integer values of mℓ, as may be seen in the following equationsderived from equation 41.

mℓ = 0 ⇒ cos 0 = cos 0 = 1 (42)

mℓ = +1 ⇒ cosφ+ i sinφ = cos(φ+ 2π) + i sin(φ+ 2π) (43)

mℓ = −1 ⇒ cosφ− i sinφ = cos(φ+ 2π)− i sin(φ+ 2π) (44)

Therefore, from these periodic equations, we see that mℓ = 0,±1,±2, . . . will satisfy the periodiccharacter of the wave function.

[The student should examine equations 42–44 to show that non-integral values of mℓ will notsatisfy these equalities].

It is instructive to examine the normalized plane rotor wave functions as a function of thequantum number, mℓ. These are shown for the first three allowed quantum states of the rotor inFigure 6. These wave functions are plotted together with the quantized energy levels for the rotor.

Let us now consider the value of ψ∗ψ for the planar rigid rotor, by considering the normalized

wave function ψ =(

12π

)1/2eimℓφ. We note that ψ∗(φ)ψ(φ) = 1

2πe−imℓφ+imℓφ = 1

2π , which is a

constant for all values of mℓ and all values of φ! This means that the probability of findingthe rotor in the range φ to φ + dφ for any angle, φ, is constant and is dφ

2π , no matterwhat mℓ may be.

Page 11: Lecture 10 Example: Particle in a box

Lecture #10 11

Figure 6: Wave Functions and Allowed Energy Levels for the Rigid Rotor—Planar

Page 12: Lecture 10 Example: Particle in a box

Lecture #10 12

The Quantized Rigid Rotor—Three-Dimensional

Our reason for studying the idealized rigid rotor is to ultimately apply our understanding tomolecules that rotate in a quantized fashion. At this stage of our development, the student shouldbe thinking about a diatomic molecule that is rotating like a dumbbell in space. A schematic of acarbon monoxide molecule is shown in Figure 7. This diatomic molecule possesses a single momentof inertia, I = m1r

21 +m2r

22, where m1 and m2 are the masses of the atoms, and r1 and r2 are the

distances of these masses from the center of mass of the molecule. In the three dimensional case,the diatomic molecule can tumble in space, or more specifically its plane of rotation can occur inany plane in space. The physical condition we have previously considered in which rotation occursin a fixed plane (planar rotor) does not apply. We will not derive the equations related to thethree-dimensional rigid rotor but will give the result in equation 45

E3Drot = J(J + 1)

h2

2I(45)

where the quantum number J = 0, 1, 2, . . .

rO

rC

Figure 7: Schematic of a Carbon Monoxide Molecule. The distance of each atom from the centerof mass is shown, where rC is the distance from the carbon atom (light sphere) to the center ofmass and rO is the distance from the oxygen atom (dark sphere) to the center of mass.

A plot of the allowed rotational energy levels for a diatomic molecule is shown in Figure 8 asa function of the quantum number, J . It may be seen that as J increases, the spacing betweenallowed rotational quantum states increases. If we let J be the rotational quantum number of aparticular state, with (J − 1) the rotational quantum number of the next lower allowed state, then

one may calculate that the spacing between adjacent levels increases by the amount 2Jh2

2I , whichwill be called 2JB where B is the rotational constant, given by

B =h2

2I. (46)

This will become important when we discuss spectroscopic transitions between neighboring rota-tional states. [The student should prove ∆E = EJ − EJ−1 = 2JB].

Page 13: Lecture 10 Example: Particle in a box

Lecture #10 13

Figure 8: Allowed Energy Levels for the Rotations of a Diatomic Molecule.