You are on page 1of 15

CH2422 Quantum Chemistry Although classical physics is often applied within chemistry to explain interactions on a visible scale, the

e same approximations are not as applicable when describing phenomena on a much smaller scale i.e. particles at the molecular scale and during very small energy transfers. Classical physics assumes that; The precise position and momentum of a particle can be precisely determined simultaneously The motion of an object can assume an arbitrary energy Waves and particles are distinct

Quantum mechanical theories are necessary, therefore, to make up for the shortcomings of these assumptions and to develop an understanding of several phenomena, e.g. molecular and atomic structures, chemical bonding, spectroscopy. Quantum Effects in Chemistry Spectroscopy and the Quantisation of Energy Levels Spectroscopy as a technique depends upon quantum effects to provide information about chemical species, for example different molecules will provide different and characteristic UV spectra. These spectra show absorptions and emissions at very particular frequencies, thus disproving the notion that a system may possess an arbitrary energy. Instead, atoms and molecules are said to exist within quantum states, these states have a defined energy and so energy levels are quantised and only certain transitions are permitted. There are exceptions, e.g. ionisation, here the combined ion-electron system can take any energy above the ionisation limit, this energy level is quantised, however, beyond this limit the kinetic energy of the free electron is continuously varied and is therefore not subject to quantisation. The Particle Nature of Light The evidence for the particle nature of light is best described by the photo-electric effect. Upon irradiation with UV light a metal surface will only eject electrons once a characteristic frequency threshold has been surpassed. The kinetic energy of the emitted electrons increases as the frequency increases while increasing the intensity of the light will have no effect. This shows that electromagnetic radiation is transferring energy in strict quanta and that light itself must therefore be able to behave as a particle, these light particles are called photons and can be considered a discrete packet of information. This phenomenon, in conjunction with the quantisation of molecular energy levels, is the basis of absorption and emission spectroscopy. The Wave Nature of Matter While studying the scattering of electrons from the surface of a nickel sample C.J. Davisson and L. Germer, having accidentally melted their sample and reformed it as a single crystal, were able to confirm a theory proposed by de Broglie two years prior that matter is able to possess wave-like

characteristics, completing the quantum picture of wave-matter duality. The electrons scattering from the nickel surface displayed a clear interference pattern, the arranged crystal lattice was able to act as a diffraction grating thus proving that electrons, classically considered a form of matter, were able to behave as waves, similar diffraction patterns are also displayed for objects of much higher masses including protons and neutrons. Molecular theory, in particular our understanding of bonding and anti-bonding interactions are now very heavily reliant on this quantum understanding of the electron, destructive and constructive interference led to the current molecular orbital theory that has provided explanations for realworld phenomena for decades. Wave-Particle Duality Therefore on a microscopic scale it has been shown that particles may take on the characteristics of waves, and that waves may take on the characteristics of particles. This duality was first theorised by de Broglie in the form of the de Broglie relationship (Equation 1). The presence of Plancks constant sets the scale on which wave-particle duality becomes a significant effect as anything with a macroscopic momentum will have an overwhelmingly large denominator, thus leading to a negligible wavelength. This explains why classical physics seemed to provide an accurate fit before the discovery of the constituents of the nucleus, electrons and other incredibly small particles.
Equation 1 - the de Broglie relationship
1

Wavefunctions and the Schrdinger Equation In quantum mechanics the wave nature of a system is expressed quantitatively through the wavefunction, . The wavefunction is defined everywhere in 3N dimensional space as a singlevalued and finite number, the physical properties of a system are dependent on it. The wavefunction is determined by the Schrdinger equation (Equation 2) set within boundary conditions both in space and time. The Hamiltonian operator, , acts on and transforms the mathematical function it prefaces and allows for multiple answers to exist for any one form of the equation. It is the boundary conditions that determine which mathematical solutions are realistically possible. It is often expressed as the total energy of a system and can, for many systems, be considered a combination of the kinetic energy operator, T, and the potential energy operator, V.
Equation 2 - The time-dependent Schrdinger Equation
2

Using the Schrdinger equation in a practical context it is often easier to consider stationary states; these are time-independent systems, i.e. they can be considered to provide a consistent average
1 2

=wavelength, h=Plancks constant (6.626 x 10-34 J.s), =momentum (=mv=mass x velocity) =the Hamiltonian operator, =the wavefunction, i=-1, =h/2, t=time

wavefunction called a standing-wave, atomic and molecular orbitals are examples of standingwaves. Deriving the Time-Independent Schrdinger Equation In one dimension the Schrdinger equation takes the form;
Equation 3 - One dimensional time-dependent Schrdinger equation
3

Equations of this form can be solved by separation of variables, a trial solution (Equation 4) is formed which can then be substituted to obtain Equation 5.
Equation 4 - The trial solution used in separating the variables within the Schrdinger equation
4

Equation 5 - The result of substituting the trial solution into the Schrdinger equation

By dividing both sides of Equation 5 by Equation 6 can be obtained.


Equation 6

Since only the right-hand side of Equation 6 is a function of x only the left side will change upon altering the value of x, since both sides are equal and the left-hand side is unaffected by variation of x, the right-hand side must be equal to a constant. This constant is given the symbol, E, as the dimensions are those of an energy (the same as those of V). The time-dependent equation (Equation 3) can therefore be split into the two differential equations, Equation 7 and Equation 8.
Equation 7 - The differential equation formed by the right-hand side of Equation 6
5

Equation 8 - The differential equation formed by the left-hand side of Equation 6

In principle any value of E is allowed, however with the introduction of typical boundary conditions there are only certain values of E that give a non-zero value of , this is the mathematical origin of quantisation. Quantisation can therefore be rationalised as the requirement to fit an integral

3 4

m=the mass of the particle, V=the potential energy operator and are fractions of the total wavefunction 5 E=the total, conserved energy of the system

number of wavelengths with wavefunctions sufficing the Schrdinger equation and within the potential energy boundaries of a system as any non-integral sum of waveforms would lead to destructive interference. Equation 8 has the solution shown in Equation 9 and so the complete wavefunction (=) has the form given in Equation 10. This wavefunction satisfies Equation 7 and may be written as shown in Equation 11.
Equation 9 - The solution to Equation 8

Equation 10 - The form of the complete wavefunction

Equation 11 - The time-independent Schrdinger equation

Since Equation 7 takes the form of a standing-wave equation it is legitimate, when only considering the spatial dependence of the wavefunction, to consider Equation 11 a wave equation. For as long as the potential energy is independent of time and system is in a state of energy, E, it is possible to construct the time-dependent wavefunction from the time-independent wavefunction by multiplying the latter by e-iEt/. Construction of the Hamiltonian operator, When constructing a Hamiltonian operator there are two steps; 1) Consider the expression that defines the energy as if it were obeying classical mechanics by forming an equation that includes information on the coordinates (x,y and z in Cartesian format) and momentum, px, for each particle 2) Replace each value of px with i d/dx. This comes from the following proof; Firstly, consider the constituents of the Hamiltonian (Equation 12).
Equation 12 - The Hamiltonian shown as the sum of potential and kinetic energies
7

Substituting in for T with the classical equation for kinetic energy (Equation 13) and further substituting for the quantum expression for momentum (Equation 14) will give an expression for T (Equation 15) as defined in step 2 above.
Equation 13 - The classical mechanics expression for kinetic energy
8

6 7

The constant of proportionality in Equation 9 is absorbed into the normalisation constant for T=the kinetic energy operator 8 v=velocity

Equation 14 - The quantum mechanical expression for momentum

Equation 15 - The final form of T according to quantum mechanics

The Born Interpretation and Normalisation The wavefunction contains all information for a system it is not directly measurable, nor does it have any physical meaning. The Born interpretation states that (often simply 2) will provide a probability density. The probability of finding an object between x and dx in one dimension is therefore equal to 2(x)dx, in 3 dimensions the probability becomes 2dV(10). If this is to be true then it must also be true that the overall probability of finding the particle, for example in one dimension beteen the points x=0 and x=L (Equation 16), must be equal to 1.
Equation 16 - A mathematical representation of the Born interpretation for the wavefunction of a particle in one dimension in a box of length, L

Since the wavefunction term appears on both sides of the Schrdinger equation (Equation 11) it is possible to multiply it by any value without any effect on the overall solution. A value is therefore chosen that ensures the probability of finding the particle within the entire system is equal to 1, this function is then considered normalised.

9 10

=Laplace operator, a differential operator given by the divergence of a function on Euclidean space V here is not the potential energy operator but simply denotes volume

Constructing and Normalising the Hamiltonian Operator, , for Conceptual Systems A Particle in a 1D Box Imagine a particle confined within two walls to a region of space of length, L. This system is called a one-dimensional square well, or a particle in a box, and can be represented graphically (Figure 1). The two walls have a potential energy that rises abruptly to infinity, thus providing boundary conditions for the permissible wavefunctions and ensuring that all acceptable energies are subsequently quantised.

Potential energy, V(x)

Position, x

Figure 1 - A graphical representation of a particle in a box

The particle in this box will have a potential energy of 0, and so the first step is to construct a Hamiltonian in which the only term is that of kinetic energy (Equation 17).
Equation 17 - The Hamiltonian for a particle in a box
11

This equation is the same as the Hamiltonian for free translational motion. Again, it is due to the imposed boundary conditions that the number of acceptable wavefunctions for this system is so many fewer than there are for free translational motion. Wavefunctions must be everywhere continuous, and so the biggest implication of the boundary conditions is that at x=0 and x=L, the wavefunction must be equal to 0. From here it is possible to form a general solution (Equation 18) into which the two boundary conditions can be tested (Equation 19 and Equation 20).
Equation 18 - The general solution for a particle in a box

Equation 19 - The wavefunction at x=0

12

11 12

The substitutions described in Construction of the Hamiltonian (page 4) are shown here stepwise cos 0=1 and sin 0=0

Since Equation 19 shows that the wavefunction is equal to C at x=0, it is sensible to define C as being equal to 0, eliminating the cos term from the equation and beginning the process of fitting the wavefunction into the boundary conditions.
Equation 20 - The wavefunction at x=L

It would be possible, using Equation 20, to define D as being equal to 0 as well. This would, however, mean that the wavefunction is 0 everywhere and so an alternative approach must be considered in order to make the sin function in the equation disappear at x=L. This is achieved if kL is equal to an integer multiple of 13 (Equation 21).
Equation 21 - The allowed values for k

The energy of the system can be determined (Equation 22) and is found to also be dependent on n, n is therefore a quantum number and may be used to label the state of a system and is responsible for the quantisation of energy.
Equation 22 - Stepwise calculation of the formula for the allowed energy values

The difference in energy levels for this system can therefore be given by Equation 23.
Equation 23 - Calculating the difference between energy levels

Equation 23 also shows that as L increases, E decreases and so at a large enough distance (beyond the scale of Plancks constant, present in the numerator) the change between energy levels can be treated as done in classical terms. E also decreases as the mass, m, of the object increases and so again, at larger masses the quantisation of energy is unnecessary. The only thing yet to be calculated is D, this constant plays the role of normalising the wavefunction (Equation 24) so as to satisfy the Born interpretation.

13

Sin()=0 in radians, since sin nx = n sin x it follows that any multiple of will give a result of 0, as required by the boundary conditions

Equation 24 - Calculation of the normalisation constant, D, for the wavefunction of a particle in a box14

For a particle in a box an increase in energy level will also lead to an increase in nodal points, positions where the probability of finding the particle is 0. There are n-1 nodes (excluding the endpoints) and it is the presence of these nodes that all for states to be mutually orthogonal and prevents the mixing of all levels (and in turn the falling apart of quantisation).

14

Note that;

The Hamiltonian for Other Systems In the following equations the terms for kinetic energy are in blue, and the terms for potential energies in green. A Particle in a 3D Box ( )

Note that while the Hamiltonian is additively separable, the wavefunction is multiplicatively separable, i.e. lmn=l(x) m(y) n(z), this gives nodal structure in 3 dimensions. The energy is the sum of energies in each degree of freedom;

Harmonic Oscillator
Equation 25 - Hamiltonian for a harmonic oscillator
15

The Schrdinger equation arising from Equation 25 is shown in Equation 26.


Equation 26 - The Schrdinger equation for a harmonic oscillator

The solution for this Schrdinger equation gives an equation for the energy values shown in
Equation 27 - The allowed energy levels of the harmonic oscillator
16

According to classical physics the motion of an oscillator ought to be restricted by turning points (points at the edge of the oscillating motion on which the particle momentarily stops before changing direction) given by x= . In quantum mechanics, however, 2 is non-zero (Figure 2) beyond these classical barriers. This has the implication that the potential energy is in fact greater than the total energy, in turn implying that the kinetic energy must be negative. This is the basis for quantum tunnelling, a phenomena utilised by scanning electron microscopy (SEM).

15 16

k= the force constant, m=mass ( )

=the vibrational quantum number

Figure 2 - The forms of the first 7 wavefunctions found upon application of the equations for a harmonic oscillator

It is worth noting in Figure 2 that as the energy level increases, the wave takes a shape more in alignment with classical predictions, i.e. the probability of finding the particles at the classical turning points is the most noticeable cf. the probability of finding the particles at the classical turning points is the most noticeable cf. 0 where the probability density is almost entirely within the parabola and not on the line of the curve.

Hydrogenic Atom
Equation 28 - The Hamiltonian for a hydrogenic atom

The second term of Equation 28 is the motion of the nucleus, this is not entirely necessary as the electron can be considered to move around the nucleus while the nucleus remains stationary. By replacing the terms for nuclear and electronic motion with those of centre-of-mass coordinates and the relative electronic position it is possible to reach Equation 29. The solution to the Schrdinger equation is then shown by Equation 30.
Equation 29 - The Hamiltonian of a hydrogenic molecule in which the nuclei is considered a stationary centre of mass
17

(
Equation 30 - Solution to the Schrdinger equation for a hydrogenic atom

)]

The allowed energy levels are therefore;


17

The reduced mass, =memn/me+mn

Equation 31 - The energies allowed by the Schrdinger equation (Equation 30)

18

The energies given by Equation 31 have negative values; this implies that the electron has a lower energy than it would were it free, as would be predicted. Hydrogen Molecule Ion (H2+)19
Equation 32 - The Hamiltonian for a hydrogen molecule ion

Hydrogen Molecule20
Equation 33 - The Hamiltonian for a hydrogen molecule

Particle on a Ring The particle on a ring is a particularly useful description as it can apply to both particles moving in a circular motion as well as the gyration of a particle around its own centre of mass. The Hamiltonian for a particle on a ring (Equation 34) can be simplified if polar coordinates are adopted and x and y are replaced by (r cos ) and (r sin ) respectively, where varies from 0 to 2 (Equation 35).
Equation 34 - The Hamiltonian for a particle on a ring, considered to have no potential energy and motion on the xy plane

Equation 35 - The Hamiltonian after substitution of polar coordinates

If r is constant the derivatives with respect to r can be disregarded, giving Equation 36.
Equation 36 - The final Hamiltonian for a particle on a ring
21

18 19

RH is the Rydberg constant A and B denote the two hydrogen nuclei 20 Where A and B are defined above, 1 and 2 denote the two electrons 21 Inertia = I = mr2

Equation 37 - The energy levels of a particle on a ring

22

Equation 37 allows for the calculation of energy levels for a particle on a ring. m l is a dimensionless number with integral values (both positive and negative). This expression defines four characteristic features of the energy levels for this scenario; The separation of neighbouring levels increases as ml increases (the ml term is squared and so the relationship is non-linear) The separation of energy levels is small for systems with large moments of inertia, i.e. those beyond a quantum scale There is no zero-point energy (at E0 the energy is 0) All energy levels other than the ground state are double degenerate i.e. m l=1 will have the same energy as ml=-1, ml essentially represents the direction of travel around the ring

22

ml=0, 1, 2

Penetration of Potential Barriers Imagine, again, a barrier of infinite width that has a potential energy labelled V(x) (Figure 3). Potential energy, V(x)

Position, x

Figure 3 - A barrier of infinite width

The potential of a particle inside the barrier is equal to 0 at x<0 and V at x0 and the two Hamiltonians are shown below (Equation 38 and Equation 39).
Equation 38 - Hamiltonian for the particle where x<0

Equation 39 - Hamiltonian for the particle where x0

The general solutions to these equations are shown by Equation 40 and Equation 41. By considering only the case in which E<V, where classically it would be impossible for the particle to be found in the barrier, and applying this to Equation 41 it is implied that k is imaginary. k can therefore be redefined as i where kappa is real, giving Equation 42.
Equation 40 - Solution to the Hamiltonian for x<0

(
Equation 41 - Solution to the Hamiltonian for x0

Equation 42 - Modification to the solution for the Hamiltonian at x0

This wavefunction is therefore a mixture of decaying and increasing exponentials, i.e. it does not oscillate when E<V. The increasing exponential must be dismissed as, across an infinite barrier, this implies an infinite amplitude. The wavefunction is therefore an exponentially decaying function of the form exp(-x). Most importantly, at x0 the wavefunction is not equal to zero, the particle can exist within the classically forbidden region, this is penetration. Kappa is a measure of how quickly the exponential decays to zero and the penetration depth, 1/, decreases with increasing particle

mass and increasing height of the barrier (i.e. V-E). Again this effect is limited to the smaller particles, protons and electrons are able to penetrate forbidden zones to an appreciable extent. Crossing a Barrier of Finite Width - Quantum Mechanical Tunnelling Imagine a barrier, of potential energy V(x), similar to that imagined for the particle in a box but with a defined width,0<x< L (Figure 4). At x<0 and xL , V(x)=0 and at 0x<L, V(x)=V. Potential energy, V(x)

Position, x

Figure 4 - A barrier of finite width

The general solutions to the time-independent Schrdinger equations can easily be written (Equation 43,Equation 44 and Equation 45).
Equation 43 - The wavefunction at x<0 for a finite width barrier

Equation 44 - The wavefunction at 0x<L for a finite width barrier

(
Equation 45 - The wavefunction at Lx for a finite width barrier

A and B represent incoming and outgoing waves, an incoming wave contributes to the total wavefunction with a component of linear momentum towards the target whereas an outgoing wave has a contribution with a component of linear momentum away from the target. It can be seen that either side of the barrier the wave acts as before, it is able to move in both directions and oscillates as a function of both a positive and negative exponential. For the same reasons as discussed above, the B term is the only of significance (as the A term would imply an infinite amplitude) and as a result the wave decays exponentially in this region. As a result, the emerging wave will be of a lower amplitude than the original wave on the left.

To consider a simple, workable example, imagine the wave is formed from a cos function 23. In this case;
Equation 46 - The form of the wavefunction on the left of the barrier when representing the wave as a cos function

Equation 47 - The form of the wavefunction in the barrier when representing the wave as a cos function

Equation 48 - The form of the wavefunction on the right of the barrier when representing the wave as a cos function

The transmission coefficient is given the label T and accounts for the decay across the barrier and is given by;
Equation 49 - The value of the transmission coefficient, T

Expectation Values

23

A sin wave would also work but if this were to be an expansion of the particle in a box (a useful approximation in worked examples) the sin wave would be 0 at the barrier and so tunnelling could not occur, a cos wave will have an amplitude of 1 at the barrier and so could decay across the barrier as required

You might also like