Quantum Concepts - Basic Macroscopic and Microscopic Concepts: Thermodynamics, Classical, and Quantum Mechanics - Liquid-State Physical Chemistry: Fundamentals, Modeling, and Applications (2013)

Liquid-State Physical Chemistry: Fundamentals, Modeling, and Applications (2013)

2. Basic Macroscopic and Microscopic Concepts: Thermodynamics, Classical, and Quantum Mechanics

2.3. Quantum Concepts

Although atomic and molecular phenomena require quantum mechanics for their description, for our discussion we need only some principles thereof and a few exactly soluble single-particle problems.

2.3.1 Basics of Quantum Mechanics

Since the discovery of quantum mechanics we know that in the Schrödinger representation the state of a system, that is, for the atoms or molecules at hand, is given by the wave function c2-math-5083, where r stands for the full set of coordinates and t for time. This function and its gradient must be single-valued, finite, and continuous for all values of its arguments. A function satisfying these conditions is said to be well-behaved. It also must have a finite integral of its quadrate if integrated over the complete range of coordinates. Generally, we require that c2-math-5084 is normalized, that is,

c2-math-5085

where the asterisk denotes the complex conjugate. Schrödinger proposed that the time development of c2-math-5086 for a free particle with position15) r and mass m is given by

(2.70) c2-math-0070

where ħ = h/2π and h denotes Planck's constant. Using the de Broglie relation p = ħk with k = |k| = 2π/λ, where p is the momentum and k is the wave vector, we can write the wave function as exp(ik·rω t) with the circular frequency ω = 2πν, and, as usual, λ and ν representing the wavelength and frequency, respectively. Since the classical energy for a free particle is E = p2/2m, this suggests the operators

c2-math-5087

and these relations are accepted for all systems. In fact, every physical observable is represented by a linear Hermitian operator16) acting on the wave function, for example, the position operator c2-math-5088. The only possible values, which a measurement of the observable with operator c2-math-5089 can yield, are the eigenvalues a of the eigenvalue equation17) c2-math-5090 with c2-math-5091 as the eigenfunction. For the coordinate operator c2-math-5204 the eigenvalues ξ of a single particle i are the values for which the equation

c2-math-5092

which is an ordinary algebraic equation, possesses solutions. Rewriting this as

c2-math-5093

it is evident that for this expression to be true either r = ξ or c2-math-5094. This corresponds exactly to the definition of the Dirac δ-function (see Appendix B), which is thus the eigenfunction δ(rξ) associated with the operator c2-math-5205 for the eigenvalue ξ.

We note that the energy E for conservative systems is given by the Hamilton function c2-math-5240 in terms of the coordinates r and conjugated momenta p, that is, c2-math-5206. We can therefore form the Hamilton operator c2-math-5095(using a symmetrized function of r and p) leading to the time-dependent Schrödinger equation

(2.71) c2-math-0071

For stationary solutions we may write

c2-math-5096

which, upon substitution in Eq. (2.71), leads to

c2-math-5097

As both sides are independent of each other, they both must be equal to a (so-called separation) constant, say E. Solving the time-dependent part leads to

c2-math-5098

while the space-part yields the time-independent Schrödinger equation

(2.72) c2-math-0072

The latter is an eigenvalue equation that shows that the eigenvalue E represents the allowed quantum energy of the system. The constant C may be chosen to normalize Ψ(r), if necessary. In general a set of solutions is obtained where each solution characterized by a wave function Ψk(r) and associated energy Ek. The various wave functions Ψk(r) form, if normalized, a complete, orthonormal set, that is,

c2-math-5099

where the Dirac bra-ket notation18) introduced. This is also true for the eigenfunctions of any operator, for example, the momentum operator −iħ∇. Here, we have

(2.73) c2-math-0073

Normalization can be done in two ways. First, by periodic box normalization where we assume that the particle is contained in an arbitrary box of volume L3 and that the potential at x = y = z = 0 equals that x = y = z = L. We obtain, using ⟨Ψk|Ψl ⟩ = δkl = δkx,lxδky,lyδkz,lz, C = L−3/2. Second, by δ-function normalization where we employ a particular representation of the Dirac δ-function (see Appendix B), namely

c2-math-5100

The function sin(gx)/πx has a value of g/π at x = 0, oscillates with decreasing amplitude and period 2π/g with increasing |x| and is normalized, independent of the value of g. Hence,

c2-math-5101

Consequently, C = 2π−3/2 using exp(ik·r) or C = h−3/2 using exp(ip·r).

When a system is in an arbitrary state Φ the expected mean of a sequence of measurements of the observable with operator c2-math-5102 is given by

(2.74) c2-math-0074

where c2-math-5103 is referred to as the expectation value and is to be interpreted as c2-math-5104, where ai denotes the measured value and ρi its probability of occurrence. If we expand Φ in terms of Ψk's, we have c2-math-5239 and the expectation value becomes

c2-math-5105

Hence, when the system is in a state Φ, a measurement of the observable A yields the value ak with a probability c2-math-5207, where ck is the expansion coefficient (or probability amplitude) in the expansion of Φ in terms of Ψks, or

(2.75) c2-math-0075

The probability amplitude may be expressed in terms of Φ and Ψk as

(2.76) c2-math-0076

If Φ is one of the eigenfunctions Ψl of c2-math-5106, we obviously have ρk = ⟨Ψk |Ψl⟩ = δkl. In this case, the expectation value c2-math-5107 is thus equal to the eigenvalue al. The above considerations lead directly to the interpretation of the wave function. Consider the probability that a measurement of the position of a single particle i will give the value ξ. The eigenfunction corresponding to the coordinate operator r(i) for the eigenvalue ξ has shown to be Ψξ = δ(rξ). From Eqs (2.75) and (2.76), we obtain

(2.77) c2-math-0077

Hence, the probability of finding the particle i at r(i) = ξ is given by the square of its wave function evaluated at ξ. Generalization to many particles is straightforward.

An important next item is that conjugated variables, like the position coordinate r and momentum p or energy E and time t, cannot be determined precisely at the same time or, in other words, if one of the conjugated variables is exactly known, the other is fully undetermined. For the energy this is already clear: in a stationary state at any time t the energy is exactly E. In general any process that shortens the life time, broadens the energy level. Since transitions to other states are required for a change in time, there is a limited residence time Δt and this leads to a small uncertainty in energy ΔE given by ΔEΔt = ½ħ. More generally, if we define the variance by c2-math-5108we have for any pair of conjugated variables a and b

(2.78) c2-math-0078

which are known as the Heisenberg uncertainty relations. The implication for the pair coordinate r and momentum p is that a precisely localized particle has an indeterminate momentum, and vice versa.

Another important point is that particles also possess spin angular momentum, for which no classical analog is available, but which obeys the same quantization rules as orbital angular momentum. For an electron the eigenvalue for the spin operator is either ½ħ or −½ħ with the eigenfunctions generally denoted by α and β, respectively.

Finally, we focus our attention on systems with a Hamilton operator with additive properties. For many-particle systems the Hamilton operator c2-math-5208 is a function of the co-ordinates and the momenta of all the particles i,j, … , that is, c2-math-5109. In a number of cases, however, the Hamilton operator is the sum (or can be approximated as the sum) of the operators c2-math-5209 for single particles i, that is, c2-math-5110. These single-particle operators satisfy the (single particle) Schrödinger wave equation

(2.79) c2-math-0079

where ϕk(i) and εk denote single-particle wave functions and eigenvalues (particle energies), respectively. The total wave function Ψ for the N-particle system can be taken as the product of the individual particle wave functions c2-math-5210, so that the N-particle Schrödinger equation reads

(2.80) c2-math-0080

and the total energy is given by c2-math-5211. However, we note that the individual particles in a system are indistinguishable. This implies that the wave function must be either symmetrical or anti-symmetrical with respect to exchange of particle coordinates, including spin. Electrons are particles with a half-integer spin, the so-called fermions (or Fermi–Dirac particles), for which the wave function must be anti-symmetric in the coordinates of all electrons. A direct consequence, usually referred to as Pauli's principle, is that each system state can be occupied with only one electron. For an electron with spin ½ this implies either spin up (½ħ) or with spin down (−½ħ). Making allowance for Pauli's principle, a many-electron wave function can be expanded in anti-symmetrized products of one-electron wave functions (or spin orbitals) ϕj, each ϕj consisting of a spatial part (or orbital) and a spin function (σj = α or β). For this the permutation operator c2-math-5111 is used, which permutes the coordinates r of the wave function. A permutation is either odd or even, dependent on the number of binary switches in coordinates c2-math-5212 that is required to realize that permutation. A completely anti-symmetrized wave function Ψ(A) is obtained by combining all N! permutations c2-math-5213 of the wave function Ψ = Ψ(r1, r2, … ,rN), that is, by taking

(2.81) c2-math-0081

So we have c2-math-5112 or 1 for an odd or even permutation, respectively. For future reference we include the prefactor c2-math-5113 in the operator c2-math-5214 by writing c2-math-5114. A convenient form for such an anti-symmetrized product is the Slater determinant, in shorthand written as

(2.82) c2-math-0082

where (i) denotes the coordinates of an electron including spin and N the number of electrons. Similarly, for integer spin particles, the so-called bosons (or Bose–Einstein particles), the wave function must be symmetric in all the coordinates of the particles and the system states may contain any number of particles. This symmetry can be realized by taking the permanent of individual particle wave functions indicated by

(2.83) c2-math-0083

The permanent is constructed similarly as the determinant but upon expanding one takes all signs as positive. Hence, c2-math-5115 always. In order to obtain a normalized wave function Ψ, if all ϕ's are orthonormal, one has to include a normalization factor C = [Ni(ni!)]−1/2

c2-math-5116

for fermions and bosons, respectively. Note that for Fermi–Dirac particles C becomes C = (N!)−1/2 because all ni are either 1 or 0. The factor C is often incorporated in the definition of the (anti-)symmetrized product function so that we write for the wave function Ψ = |ϕa(1) ϕb(2) … ϕn(N)| or even, assuming a fixed order of coordinates Ψ = |ϕa ϕbϕn|. The same convention is used for the symmetrized product function. A straightforward calculation shows that for the Hamilton operator with additive properties c2-math-5117 the N-particle system eigenvalue is still

c2-math-5118

When c2-math-5119, the energy E is no longer the sum of the particle energies and the product function and determinant (or permanent) no longer yield the same energy.

To conclude this section we mention that the number of exact solutions for realistic systems is very limited in quantum mechanics19). Therefore, as has been emphasized many times before, the Schrödinger equation needs drastic approximations in almost all cases, even to obtain approximate solutions. We indicate here only the solutions of a few, exactly solvable single-particle problems, namely the particle-in-a-box, the harmonic oscillator, and the rigid rotator. Thereafter, we briefly discuss some approximation methods. For our discussion of liquids and solutions, the basic concepts and examples discussed will more than suffice. In fact, the existence of energy levels is the most important aspect for the basis of statistical thermodynamics.

Problem 2.16

Show that the eigenvalues of a Hermitian operator are real.

Problem 2.17

Prove the orthogonality of the eigenfunctions of a Hermitian operator.

Problem 2.18

Show that the squared deviation of the energy from its expectation value for a stationary state is zero, that is, c2-math-5120.

Problem 2.19

Write explicitly the expression for two-electron determinant wave function. Show that this expression can be separated in a spatial and a spin factor.

Problem 2.20*

Verify that the total energy for an operator c2-math-5121 for a determinant wave function.

Problem 2.21*

Calculate the energy for a two-electron product and determinant wave function using the Hamilton operator c2-math-5215. Show that the total energy is no longer the sum of the single particle energies for the determinant.

2.3.2 The Particle-in-a-Box

For a particle in a one-dimensional box the potential energy is given by

(2.84) c2-math-0084

The kinetic energy operator is

(2.85) c2-math-0085

where m is the mass of the particle, so that the Schrödinger equation reads

(2.86) c2-math-0086

The solutions are

(2.87) c2-math-0087

where n = 1,2, … is a positive integer, the so-called quantum number, arising since only for discrete wavelengths the solution obeys the boundary conditions. These wavelengths are given by λ = 2w/1, 2w/2, … , 2w/n with allowed energies

(2.88) c2-math-0088

where k is the wave vector and p the momentum. In the ground state with n = 1, the energy is still not zero and the wave function represents a standing wave. For a three-dimensional box with potential energy Φ(r) = 0 for 0 < r < wand Φ(r) = ∞; otherwise, the energy levels are

(2.89) c2-math-0089

If one of the dimensions of the box is equal to another, the energy levels become degenerate, that is, there are two (or more) states with the same energy. For example, if w1 = w2 = w3 the energies for states n = (1,2,2), n = (2,1,2) and n = (2,2,1) are the same, and the system is said to be three-fold degenerate. Using periodic boundary conditions Φ(0) = Φ(w) = constant, instead of the fixed box boundary conditions given by Eq. (2.84), a similar procedure will lead to the same expression for En but with the factor 8 replaced by 2. In this case, n = (n1,n2,n3) ∈ (±1,±2, …) and the wave functions exp(ip·r/ħ) = exp(ik·r) represent running waves.

2.3.3 The Harmonic Oscillator

For a one-dimensional harmonic oscillator the potential energy with force constant a and the kinetic energy operator are given by

(2.90) c2-math-0090

respectively, so that the Schrödinger equation reads

(2.91) c2-math-0091

The solutions for the wave functions are

(2.92) c2-math-0092

c2-math-5122

and where Hn(x) are known as Hermite functions defined by

c2-math-5123

Explicitly the first few Hermite functions are

c2-math-5124

c2-math-5125

The energy levels are given by

(2.93) c2-math-0093

For a three-dimensional oscillator with potential energy (shifting the origin with x0, y0 and z0)

(2.94) c2-math-0094

the energy levels are given by

(2.95) c2-math-0095

Again, if one of the frequencies is equal to another, the system is degenerate. For example, if ωx = ωy = ωz, for states n = (1,1,2), n = (1,2,1) and n = (2,1,1) the energy levels are the same and given by

c2-math-5126

and the system is three-fold degenerate. Finally, we note that the presence of the zero-point energy ½ħω for the ground state n = 0 is in accord with the uncertainty relations. If the oscillator had no zero-point energy, it would have zero momentum and be located exactly at the minimum of Φ(r). The necessary uncertainties in position and momentum thus give rise to the zero-point energy. Finally, for completeness, we mention that the selection rule for transitions between levels is given by Δn = ±1.

2.3.4 The Rigid Rotator

For the rigid two-particle (atom) rotator there is no potential in the conventional sense, but there is the constraint that the distance between the two particles of the rotator should remain constant. Hence, the Schrödinger equation reads

(2.96) c2-math-0096

where μ indicates the reduced mass of the rotator given by 1/μ = 1/m1 + 1/m2, m1 and m2 are the masses of the two particles. The constraint is taken into account as follows. Rigid rotation is realized if we suppose that the two rigidly connected particles rotate freely over a sphere. For this problem it is convenient to think in terms of polar coordinates, where the radial variable or radius r is fixed but the angular coordinates θ and φ are free. The Laplace operator ∇2, expressed in polar coordinates and acting on a scalar x, is expressed by

(2.97) c2-math-0097

c2-math-5127

The Legendre operator Λ2 is the angular part of the Laplace operator. The constraint r = constant implies that the radial part of the Laplace operator can be skipped. Hence, the Schrödinger equation becomes

(2.98) c2-math-0098

where I = μr2 is the moment of inertia. The solutions for this equation are the spherical harmonics, conventionally indicated by Ylm(θ,φ), and characterized by the two quantum numbers [9] l and m. The number l = 0,1,2, … indicates the overall angular momentum, given by [l(l + 1)]1/2ħ, while the number m = −l,−l + 1, … ,l − 1,l indicates the angular momentum in a specified direction, often taken as x, and given by . The energy expression is independent of the quantum number m and becomes

(2.99) c2-math-0099

The above notation is the conventional one in quantum mechanics for the solution of the angular part of the Laplace operator. However, in spectroscopy the symbol for the quantum number l is usually replaced by J, and we will also do so. Each energy level is thus characterized by the quantum number J corresponding to 2J + 1 quantum states, that is, the degeneracy is 2J + 1. Again for completeness, we mention that the selection rule for transitions between levels is ΔJ = ±1.