2025-05-30
Solutions to Principles of Quantum Mechanics
00

目录

Chapter 7 The Harmonic Oscillator
7.1 Why Study the Harmonic Oscillator?
7.2 Review of the Classical Oscillator
7.3 Quantization of the Oscillator (Coordinate Basis)
7.4 The Oscillator in the Energy Basis
7.5 Passege from the Energy Basis to the $X$ Basis

Chapter 7 The Harmonic Oscillator

7.1 Why Study the Harmonic Oscillator?

7.2 Review of the Classical Oscillator

7.3 Quantization of the Oscillator (Coordinate Basis)

Exercise 7.3.1 Consider the question why we tried a power-series solution for Eq. (7.3.11) but not Eq. (7.3.8). By feeding in a series into the latter, verify that a three-term recursion relation between Cn+2C_{n+2}, CnC_{n}, and Cn2C_{n-2} obtains, from which the solution does not follow so readily. The problem is that ψ\psi^{\prime\prime} has two powers of yy less than 2εψ2 \varepsilon \psi, while the y2-y^{2} piece has two more powers of yy. In Eq. (7.3.11) on the other hand, of the three pieces uu^{\prime\prime}, 2yu-2yu^{\prime}, and (2ε1)u(2\varepsilon-1)u, the last two have the same powers of yy.

Exercise 7.3.2 Verify that H3(y)H_{3}(y) and H4(y)H_{4}(y) obey the recursion relation, Eq. (7.3.15).

Exercise 7.3.3 If ψ(x)\psi(x) is even and ϕ(x)\phi(x) is odd under xxx \rightarrow-x, show that

ψ(x)ϕ(x)dx=0\int_{-\infty}^{\infty} \psi(x) \phi(x) \mathrm{d}x=0

Use this to show that ψ2(x)\psi_2(x) and ψ1(x)\psi_1(x) are orthogonal. Using the values of Gaussian integrals in Appendix A.2 verify that ψ2(x)\psi_2(x) and ψ0(x)\psi_0(x) are orthogonal.

Exercise 7.3.4 Using Eqs. (7.3.23)-(7.3.25), show that

nXn=(2mω)1/2[δn,n+1(n+1)1/2+δn,n1n1/2]nPn=(mω2)1/2i[δn,n+1(n+1)1/2δn,n1n1/2]\begin{aligned} & \left\langle n^{\prime}\right| X|n\rangle=\left(\frac{\hbar}{2 m \omega}\right)^{1 / 2}\left[\delta_{n^{\prime}, n+1}(n+1)^{1 / 2}+\delta_{n^{\prime}, n-1} n^{1 / 2}\right] \\ & \left\langle n^{\prime}\right| P|n\rangle=\left(\frac{m \omega \hbar}{2}\right)^{1/2} \mathrm{i}\left[\delta_{n^{\prime}, n+1}(n+1)^{1 / 2}-\delta_{n^{\prime}, n-1} n^{1 / 2}\right] \end{aligned}

Exercise 7.3.5 Using the symmetry arguments from Exercise 7.3.3 show that nXn=nPn=0\langle n|X|n\rangle=\langle n|P|n\rangle=0 and thus that X2=(ΔX)2\left\langle X^2\right\rangle=(\Delta X)^2 and P2=(ΔP)2\left\langle P^2\right\rangle=(\Delta P)^2 in these states. Show that 1X21=3/2mω\langle 1| X^2|1\rangle=3 \hbar / 2 m \omega and 1P21=32mω\langle 1| P^2|1\rangle=\dfrac{3}{2} m \omega \hbar. Show that ψ0(x)\psi_0(x) saturates the uncertainty bound ΔXΔP/2\Delta X \cdot \Delta P \geqslant \hbar / 2.

Exercise 7.3.6 Consider a particle in a potential

V(x)=12mω2x2,x>0=,x0\begin{aligned} V(x) & =\frac{1}{2} m \omega^2 x^2, & & x>0 \\ & =\infty, & & x \leqslant 0 \end{aligned}

What are the boundary conditions on the wave functions now? Find the eigenvalues and eigenfunctions.

Exercise 7.3.7 (The Oscillator in Momentum Space) By setting up an eigenvalue equation for the oscillator in the PP basis and comparing it to Eq. (7.3.2), show that the momentum space eigenfunctions may be obtained from the ones in coordinate space through the substitution xpx \rightarrow p, mω1/mωm \omega \rightarrow 1 / m \omega. Thus, for example,

ψ0(p)=(1mπω)1/4ep2/2mω\psi_0(p)=\left(\frac{1}{m \pi \hbar \omega}\right)^{1 / 4} e^{-p^2 / 2 m \hbar \omega}

There are several other pairs, such as ΔX\Delta X and ΔP\Delta P in the state n|n\rangle, which are related by the substitution mω1/mωm \omega \rightarrow 1 / m \omega. You may wish to watch out for them. (Refer back to Exercise 7.3.5.)

7.4 The Oscillator in the Energy Basis

Exercise 7.4.1 Compute the matrix elements of XX and PP in the n|n\rangle basis and compare with the result from Exercise 7.3.4.

Exercise 7.4.2 Find X\langle X\rangle, P\langle P\rangle, X2\left\langle X^2\right\rangle, P2\left\langle P^2\right\rangle, ΔXΔP\Delta X \cdot \Delta P in the state n|n\rangle.

Exercise 7.4.3 (Virial Theorem) The virial theorem in classical mechanics states that for a particle bound by a potential V(r)=arkV(r)=a r^k, the average (over the orbit) kinetic and potential energies are related by

T=c(k)V\overline{T}=c(k) \overline{V}

when c(k)c(k) depends only on kk. Show that c(k)=k/2c(k)=k / 2 by considering a circular orbit. Using the results from the previous exercise show that for the oscillator (k=2)(k=2)

T=V\langle T\rangle=\langle V\rangle

in the quantum state n|n\rangle.

Exercise 7.4.4 Show that nX4n=(/2mω)2[3+6n(n+1)]\langle n\mid X^{4}\mid n\rangle=(\hbar/2m\omega)^2[3+6n(n+1)].

Exercise 7.4.5 At t=0t=0 a particle starts out in ψ(0)=1/21/2(0+1)|\psi(0)\rangle=1/2^{1/2}(0\rangle+|1\rangle).

(1)Find ψ(t)|\psi(t)\rangle;

(2)Find X(0)=ψ(0)Xψ(0),P(0),X(t),P(t)\langle X(0)\rangle=\langle\psi(0)| X|\psi(0)\rangle,\langle P(0)\rangle,\langle X(t)\rangle,\langle P(t)\rangle;

(3)Find X˙(t)\langle\dot{X}(t)\rangle and P˙(t)\langle\dot{P}(t)\rangle using Ehrenfest's theorem and solve for X(t)\langle X(t)\rangle and P(t)\langle P(t)\rangle and compare with part (2).

Exercise 7.4.6 Show that a(t)=eiωta(0)\langle a(t)\rangle=\mathrm{e}^{-\mathrm{i}\omega t}\langle a(0)\rangle and that a(t)=eiωta(0)\left\langle a^{\dagger}(t)\right\rangle=\mathrm{e}^{\mathrm{i} \omega t}\left\langle a^{\dagger}(0)\right\rangle.

Exercise 7.4.7 Verify Eq. (7.4.40) for the case

(1)Ω=X\Omega=X, Λ=X2+P2\Lambda=X^{2}+P^{2}.

(2)Ω=X2\Omega=X^{2}, Λ=P2\Lambda=P^{2}.

The second case illustrates the ordering ambiguity.

Exercise 7.4.8 Consider the three angular momentum variables in classical mechanics:

lx=ypzzpyly=zpxxpzlz=xpyypx\begin{aligned} & l_{x}=yp_{z}-zp_{y} \\ & l_{y}=zp_{x}-xp_{z} \\ & l_{z}=xp_{y}-yp_{x} \end{aligned}

(1)Construct LxL_{x}, LyL_{y}, and LzL_{z}, the quantum counterparts, and note that there are no ordering ambiguities.

(2)Verify that {lx,ly}=lz\left\{l_{x}, l_{y}\right\}=l_{z} [see Eq. (2.7.3) for the definition of the Poisson Bracket].

(3)Verify that [Lx,Ly]=iLz\left[L_{x}, L_{y}\right]=\mathrm{i} \hbar L_{z}.

Exercise 7.4.9 Consider the unconventional (but fully acceptable) operator choice

XxPiddx+f(x)\begin{aligned} & X \rightarrow x \\ & P \rightarrow -\mathrm{i} \hbar \frac{\mathrm{d}}{\mathrm{d} x}+f(x) \end{aligned}

in the XX basis.

(1)Verify that the canonical commutation relation is satisfied.

(2)It is possible to interpret the change in the operator assignment as a result of a unitary change of the XX basis:

xx~=eig(X)/x=eig(x)/x|x\rangle \rightarrow|\tilde{x}\rangle=\mathrm{e}^{\mathrm{i} g(X) / \hbar}|x\rangle=\mathrm{e}^{\mathrm{i} g(x) / \hbar}|x\rangle

where

g(x)=xf(x)dxg(x)=\int^x f\left(x^{\prime}\right) d x^{\prime}

First verify that

x~Xx~=xδ(xx)\langle\tilde{x}| X\left|\tilde{x}^{\prime}\right\rangle=x \delta\left(x-x^{\prime}\right)

i.e.,

Xnew X basisxX \xrightarrow[\text{new}~X~\text{basis}]{} x

Next verify that

x~Px~=[iddx+f(x)]δ(xx)\langle\tilde{x}| P\left|\tilde{x}^{\prime}\right\rangle=\left[-\mathrm{i} \hbar \frac{d}{d x}+f(x)\right] \delta\left(x-x^{\prime}\right)

i.e.,

Pnew x basisiddx+f(x)P \xrightarrow[\text{new}~x~\text{basis}]{}-\mathrm{i}\hbar \frac{\mathrm{d}}{\mathrm{d} x}+f(x)

Exercise 7.4.10 Recall that we always quantize a system by promoting the Cartesian coordinates x1,,xNx_1, \ldots, x_N; and momenta p1,,pNp_1, \ldots, p_N to operators obeying the canonical commutation rules. If non-Cartesian coordinates seem more natural in some cases, such as the eigenvalue problem of a Hamiltonian with spherical symmetry, we first set up the differential equation in Cartesian coordinates and then change to spherical coordinates (Section 4.2). In Section 4.2 it was pointed out that if H\mathscr{H} is written in terms of non-Cartesian but canonical coordinates q1qNq_1 \ldots q_N; p1pNp_1 \ldots p_N; H(qiqi,pii/qi)\mathscr{H}\left(q_i \rightarrow q_i, p_i \rightarrow-\mathrm{i}\hbar \partial / \partial q_i\right) does not generate the correct Hamiltonian HH, even though the operator assignment satisfies the canonical commutation rules. In this section we revisit this problem in order to explain some of the subtleties arising in the direct quantization of non-Cartesian coordinates without the use of Cartesian coordinates in intermediate stages.

(1) Consider a particle in two dimensions with

H=px2+py22m+a(x2+y2)1/2\mathscr{H}=\frac{p_x^2+p_y^2}{2 m}+a\left(x^2+y^2\right)^{1 / 2}

which leads to

H22m(2x2+2y2)+a(x2+y2)1/2H \rightarrow \frac{-\hbar^2}{2 m}\left(\frac{\partial^2}{\partial x^2}+\frac{\partial^2}{\partial y^2}\right)+a\left(x^2+y^2\right)^{1 / 2}

in the coordinate basis. Since the problem has rotational symmetry we use polar coordinates

ρ=(x2+y2)1/2,ϕ=tan1(y/x)\rho=\left(x^2+y^2\right)^{1 / 2}, \quad \phi=\tan ^{-1}(y / x)

in terms of which

H coordinate  basis 22m(2ρ2+1ρρ+1ρ22ϕ2)+aρ(7.4.41)H \xrightarrow[\substack{\text { coordinate } \\ \text { basis }}]{ } \frac{-\hbar^2}{2 m}\left(\frac{\partial^2}{\partial \rho^2}+\frac{1}{\rho} \frac{\partial}{\partial \rho}+\frac{1}{\rho^2} \frac{\partial^2}{\partial \phi^2}\right)+a \rho\tag{7.4.41}

Since ρ\rho and ϕ\phi are not mixed up as xx and yy are [in the (x2+y2)1/2\left(x^2+y^2\right)^{1 / 2} term] the polar version can be more readily solved.

The question we address is the following: why not start with H\mathscr{H} expressed in terms of polar coordinates and the conjugate momenta

pρ=eρp=xpx+ypy(x2+y2)1/2p_\rho=\mathbf{e}_\rho \cdot \mathbf{p}=\frac{x p_x+y p_y}{\left(x^2+y^2\right)^{1 / 2}}

(where eρ\mathbf{e}_\rho is the unit vector in the radial direction), and

pϕ=xpyypx( the angular momentum, also called lz)p_\phi=x p_y-y p_x \quad\left(\text { the angular momentum, also called } l_z\right)

i.e.,

H=pρ22m+pϕ22mρ2+aρ (verify this) \mathscr{H}=\frac{p_\rho^2}{2 m}+\frac{p_\phi^2}{2 m \rho^2}+a \rho \quad \text { (verify this) }

and directly promote all classical variables ρ\rho, pρp_\rho, ϕ\phi, and pϕp_\phi to quantum operators obeying the canonical commutations rules? Let's do it and see what happens. If we choose operators

PρiρPϕiϕ\begin{aligned} & P_\rho \rightarrow-i \hbar \frac{\partial}{\partial \rho} \\ & P_\phi \rightarrow-i \hbar \frac{\partial}{\partial \phi} \end{aligned}

that obey the commutation rules, we end up with

H coordinate  basis 22m(2ρ2+1ρ22ϕ2)+aρ(7.4.42)H \xrightarrow[\substack{\text { coordinate } \\ \text { basis }}]{} \frac{-\hbar^2}{2 m}\left(\frac{\partial^2}{\partial \rho^2}+\frac{1}{\rho^2} \frac{\partial^2}{\partial \phi^2}\right)+a \rho\tag{7.4.42}

which disagrees with Eq. (7.4.41). Now this in itself is not serious, for as seen in the last exercise the same physics may be hidden in two different equations. In the present case this isn't true: as we will see, the Hamiltonians in Eqs. (7.4.41) and (7.4.42) do not have the same eigenvalues. (What we will see is that Pρ=id/dρP_{\rho}=-\mathrm{i}\hbar\mathrm{d}/\mathrm{d}\rho, and hence the HH constructed with it, are non-Hermitian.) We know Eq. (7.4.41) is the correct one, since the quantization procedure in terms of Cartesian coordinates has empirical support. What do we do now?

(2) A way out is suggested by the fact that although the choice Pρi/ρP_\rho \rightarrow -\mathrm{i} \hbar \partial / \partial \rho leads to the correct commutation rule, it is not Hermitian! Verify that

ψ1Pρψ2=002πψ1(iψ2ρ)ρdρdϕ002π(iψ1ρ)ψ2ρdρdϕ=Pρψ1ψ2\begin{aligned} \left\langle\psi_1\right| P_\rho\left|\psi_2\right\rangle & =\int_0^{\infty} \int_0^{2 \pi} \psi_1^*\left(-\mathrm{i} \hbar \frac{\partial \psi_2}{\partial \rho}\right) \rho \mathrm{d} \rho \mathrm{d} \phi \\ & \neq \int_0^{\infty} \int_0^{2 \pi}\left(-\mathrm{i} \hbar \frac{\partial \psi_1}{\partial \rho}\right)^* \psi_2 \rho \mathrm{d} \rho \mathrm{d} \phi \\ & =\left\langle\mathbf{P}_\rho \psi_1 \mid \psi_2\right\rangle \end{aligned}

(You may assume ρψ1ψ20\rho \psi_1^* \psi_2 \rightarrow 0 as ρ0\rho \rightarrow 0 or \infty. The problem comes from the fact that ρdρdϕ\rho \mathrm{d} \rho \mathrm{d} \phi and not dρdϕ\mathrm{d} \rho \mathrm{d} \phi is the measure for integration.)

Show, however, that

Pρi(ρ+12ρ)P_\rho \rightarrow -\mathrm{i} \hbar\left(\frac{\partial}{\partial \rho}+\frac{1}{2 \rho}\right)

is indeed Hermitian and also satisfies the canonical commutation rule. The angular momentum Pϕi/ϕP_\phi \rightarrow -\mathrm{i} \hbar \partial / \partial \phi is Hermitian, as it stands, on single-valued functions: ψ(ρ,ϕ)=ψ(ρ,ϕ+2π)\psi(\rho, \phi)=\psi(\rho, \phi+2 \pi).

(3) In the Cartesian case we saw that adding an arbitrary f(x)f(x) to i/x-\mathrm{i} \hbar \partial / \partial x didn't have any physical effect, whereas here the addition of a function of ρ\rho to i/ρ-\mathrm{i} \hbar \partial / \partial \rho seems important. Why? [Is f(x)f(x) completely arbitrary? Mustn't it be real? Why? Is the same true for the i/2ρ-\mathrm{i} \hbar / 2 \rho piece?]

(4) Feed in the new momentum operator PρP_\rho and show that

Hcoordinatebasis22m(2ρ2+1ρρ14ρ2+1ρ22ϕ2)+aρH\xrightarrow[\substack{\text{coordinate}\\ \text{basis}}]{}\frac{-\hbar^2}{2 m}\left(\frac{\partial^2}{\partial \rho^2}+\frac{1}{\rho} \frac{\partial}{\partial \rho}-\frac{1}{4 \rho^2}+\frac{1}{\rho^2} \frac{\partial^2}{\partial \phi^2}\right)+a \rho

which still disagrees with Eq. (7.4.41). We have satisfied the commutation rules, chosen Hermitian operators, and yet do not get the right quantum Hamiltonian. The key to the mystery lies in the fact that H\mathscr{H} doesn't determine HH uniquely since terms of order \hbar (or higher) may be present in HH but absent in H\mathscr{H}. While this ambiguity is present even in the Cartesian case, it is resolved by symmetrization in all interesting cases. With non-Cartesian coordinates the ambiguity is more severe. There are ways of constructing HH given H\mathscr{H} (the path integral formulation suggests one) such that the substitution Pρi(/ρ+1/2ρ)P_\rho \rightarrow-i \hbar(\partial / \partial \rho+1 / 2 \rho) leads to Eq. (7.4.41). In the present case the quantum Hamiltonian corresponding to

H=pρ22m+pϕ22mρ2+aρ\mathscr{H}=\frac{p_\rho^2}{2 m}+\frac{p_\phi^2}{2 m \rho^2}+a \rho

is given by

HcoordinatebasisH(ρρ,pρi[ρ+12ρ];ϕϕ,pϕiϕ)28mρ2(7.4.44)H \xrightarrow[\substack{\text{coordinate}\\ \text{basis}}]{} \mathscr{H}\left(\rho \rightarrow \rho, p_\rho \rightarrow -\mathrm{i} \hbar\left[\frac{\partial}{\partial \rho}+\frac{1}{2 \rho}\right] ; \phi \rightarrow \phi, p_\phi \rightarrow -\mathrm{i} \hbar \frac{\partial}{\partial \phi}\right)-\frac{\hbar^2}{8 m \rho^2}\tag{7.4.44}

Notice that the additional term is indeed of nonzero order in \hbar.

7.5 Passege from the Energy Basis to the XX Basis

Exercise 7.5.1 Project Eq. (7.5.1) on the PP basis and obtain ψ0(p)\psi_{0}(p).

Exercise 7.5.2 Project the relation

an=n1/2n1a\mid n\rangle=n^{1/2}\mid n-1\rangle

on the XX basis and derive the recursion relation

Hn(y)=2nHn1(y)H_{n}^{\prime}(y)=2nH_{n-1}(y)

using Eq. (7.3.22).

Exercise 7.5.3 Starting with

a+a=21/2ya+a^{\dagger}=2^{1/2}y

and

(a+a)n=n1/2n1+(n+1)1/2n+1(a+a^{\dagger})\mid n\rangle=n^{1/2}\mid n-1\rangle+(n+1)^{1/2}\mid n+1\rangle

and Eq. (7.3.22), derive the relation

Hn+1(y)=2yHn(y)2nHn1(y)H_{n+1}(y)=2yH_{n}(y)-2nH_{n-1}(y)

Exercise 7.5.4 Thermodynamics of Oscillators. The Boltzmann formula

P(i)=eβE(i)/ZP(i)=\mathrm{e}^{-\beta E(i)}/Z

where

Z=ieβE(i)Z=\sum_{i}\mathrm{e}^{-\beta E(i)}

gives the probability of finding a system in a state ii with energy E(i)E(i), when it is in thermal equilibrium with a reservoir of absolute temperature T=1/βkT=1/\beta k, k=1.4×1016 ergs/Kk=1.4\times 10^{-16}~\mathrm{ergs}/^{\circ}\mathrm{K}; being Boltzmann's constant. (The “probability” referred to above is in relation to a classical ensemble of similar systems and has nothing to do with quantum mechanics.)

(1) Show that the thermal average of the system's energy is

E=iE(i)P(i)=βlnZ\overline{E}=\sum_i E(i) P(i)=\frac{-\partial}{\partial \beta} \ln Z

(2) Let the system be a classical oscillator. The index ii is now continuous and corresponds to the variables xx and pp describing the state of the oscillator, i.e.,

ix,pi \rightarrow x, p

and

idxdp\sum_i \rightarrow \iint d x d p

and

E(i)E(x,p)=p22m+12mω2x2E(i) \rightarrow E(x, p)=\frac{p^2}{2 m}+\frac{1}{2} m \omega^2 x^2

Show that

Zcl=(2πβmω2)1/2(2πmβ)1/2=2πωβZ_{\mathrm{cl}}=\left(\frac{2 \pi}{\beta m \omega^2}\right)^{1 / 2}\left(\frac{2 \pi m}{\beta}\right)^{1 / 2}=\frac{2 \pi}{\omega \beta}

and that

Ecl=1β=kT\overline{E}_{\mathrm{cl}}=\frac{1}{\beta}=k T

Note that EclE_{\mathrm{cl}} is independent of mm and ω\omega.

(3) For the quantum oscillator the quantum number nn plays the role of the index ii. Show that

Zqu=eβω/2(1eβω)1Z_{\mathrm{qu}}=e^{-\beta \hbar \omega / 2}\left(1-e^{-\beta \hbar \omega}\right)^{-1}

and

Equ=ω(12+1eβω1)\overline{E}_{\mathrm{qu}}=\hbar \omega\left(\frac{1}{2}+\frac{1}{e^{\beta \hbar \omega}-1}\right)

(4) It is intuitively clear that as the temperature TT increases (and β=1/kT\beta=1 / k T decreases) the oscillator will get more and more excited and eventually (from the correspondence principle)

EquTEcl\overline{E}_{\mathrm{qu}} \xrightarrow[T \rightarrow \infty]{ } \overline{E}_{\mathrm{cl}}

Verify that this is indeed true and show that “large TT” means Tω/kT \gg \hbar \omega / k.

(5) Consider a crystal with N0N_0 atoms, which, for small oscillations, is equivalent to 3N03 N_0 decoupled oscillators. The mean thermal energy of the crystal Ecrystal\overline{E}_{\text{crystal}} is Ecl\overline{E}_{\text{cl}} or Equ\overline{E}_{\text{qu}} summed over all the normal modes. Show that if the oscillators are treated classically, the specific heat per atom is

Ccl(T)=1N0EcrystalT=3kC_{\mathrm{cl}}(T)=\frac{1}{N_0} \frac{\partial \overline{E}_{\text{crystal}}}{\partial T}=3 k

which is independent of TT and the parameters of the oscillators and hence the same for all crystals. (More precisely, for crystals whose atoms behave as point particles with no internal degrees of freedom.) This agrees with experiment at high temperatures but not as T0T \rightarrow 0. Empirically,

C(T)3k(T large)0(T0)\begin{aligned} C(T) & \rightarrow 3 k & & (T~\text{large}) \\ & \rightarrow 0 & & (T \rightarrow 0) \end{aligned}

Following Einstein, treat the oscillators quantum mechanically, assuming for simplicity that they all have the same frequency ω\omega. Show that

Cqu(T)=3k(θET)2eθE/T(eθE/T1)2C_{\mathrm{qu}}(T)=3 k\left(\frac{\theta_E}{T}\right)^2 \frac{\mathrm{e}^{\theta_E / T}}{\left(\mathrm{e}^{\theta_E / T}-1\right)^2}

where θE=ω/k\theta_E=\hbar \omega / k is called the Einstein temperature and varies from crystal to crystal. Show that

Cqu(T)TθE3kCqu(T)TθE3k(θET)2eθE/T\begin{aligned} & C_{\mathrm{qu}}(T) \xrightarrow[T \gg \theta_E]{ } 3k \\ & C_{\mathrm{qu}}(T) \xrightarrow[T \ll \theta_E]{ } 3 k\left(\frac{\theta_E}{T}\right)^2 \mathrm{e}^{-\theta_E / T} \end{aligned}

Although Cqu(T)0C_{\mathrm{qu}}(T) \rightarrow 0 as T0T \rightarrow 0, the exponential falloff disagrees with the observed C(T)T0T3C(T) \xrightarrow[T \rightarrow 0]{} T^3 behavior. This discrepancy arises from assuming that the frequencies of all normal modes are equal, which is of course not generally true. [Recall that in the case of two coupled masses we get ωI=(k/m)1/2\omega_{\mathrm{I}}=(k / m)^{1 / 2} and ωII=(3k/m)1/2\omega_{\mathrm{II}}=(3 k / m)^{1 / 2}.] This discrepancy was eliminated by Debye.