wavemech.htm(Ó R. Egerton)
Figure references are to the second edition of Modern Physics by Serway, Moses and Moyer (Saunders, 1989).
Wave Mechanics
Wave mechanics is a branch of quantum physics in which the behaviour of objects (such as elementary particles) is described in terms of their wavelike properties. The simplest form of wave is a travelling sine or cosine wave, moving in the x-direction :
Y
= A cos (kx - w t) ................ (1)
where A is the amplitude of the wave, k = 2
p /l is known as the wavenumber (l = wavelength) and w = 2p f is an angular frequency, measured in radians/sec (f being the frequency in Hz). The phase velocity of the wave is defined to be vp = f l = w /k .For mechanical waves, such as sound or water waves,
Y (Greek letter psi) would represent the physical displacement of atoms at position x and time t . In the case of an electromagnetic wave, it would represent electric or magnetic field. In the case of matter waves, Y in Eq.(1) is known as a wavefunction and must satisfy a wave equation known as the Schrodinger equation.
Unfortunately, Eq.(1) cannot serve as a wave-mechanical description of a single particle, such as an electron (or a single photon of electromagnetic radiation) since the wavefunction has no localization; the cosine function continues oscillating over an infinite range of x . However, if we add two cosine waves, with slightly different wavenumbers k1 and k2 and angular frequencies (
w 1 and w 2), the result is an amplitude-modulated wave; see Fig.4.11. This is like a standing wave, with nodes and antinodes, except that each "bunch" of waves moves forward in the x-direction with a group velocity vg given byvg = (
w 2 - w 1) / (k2 - k1) = D w / D k ................ (2)Eq.(2) might represent a beam of uniformly spaced electrons travelling in the x-direction with a speed vg .
To represent a single particle, we can extend the above process, adding in cosine waves over a continuous range of frequency and wavelength. (This is sometimes called harmonic synthesis and is the principle behind the musical synthesizer, except that this electronic device uses only frequencies which are integer multiples of a fundamental frequency). The phase of each component must be chosen such that the waves interfere destructively to give zero amplitude except over the small region of x containing the particle. The result is a wave packet description of the particle.
Heisenberg's Uncertainty Principle
In developing a branch of quantum mechanics called matrix mechanics, Werner Heisenberg discovered his famous uncertainty principles, one of which may be written as:
D
p D x > h/2p ................ (3)
Here
D p represents the uncertainty in our knowledge of the momentum (p=h/l =hk/2p ) of a particle and D x is the corresponding uncertainty in our knowledge of its position. This relationship follows directly from the wave description of a particle. For example, the length of each wave packet in Fig. 4.11 can be shown to be D x = 2p /D k, so that the product D k D x = 2p , equivalent to D p D x = h , which is in agreement with Eq.(3).
As an example of this principle in operation, consider the Bohr model of the hydrogen atom in its ground state. Our uncertainty in knowledge of the position of the orbiting electron might be taken as the diameter of its orbit:
D x = 2 a0. Since at any one instant the electron might have an x-component of velocity anywhere between v1 and -v1, we can say that D v = 2 v1, giving D p D x = (mD v) D x = 2m(h/[2p ma0])(2 a0) = 2h/p , again in agreement with Eq.(3).
The most famous example of Eq.(3) is the Heisenberg microscope, a Gedanken (thought) experiment where we try to image a single particle by focussing the electromagnetic radiation which it scatters. If the optical system consists of a single lens which subtends a semiangle
q at the object (as in Fig. 4.26). the image resolution is limited (by diffraction at the lens edges) to D x = 0.6l / sinq . However, a photon initially travelling along the optic axis) which is scattered through an angle q transfers a horizontal momentum of magnitude px = p sinq = (h/l ) sin q to the particle. This momentum might be in the +x or -x direction, so the uncertainty in our knowledge of the momentum of the particle (after it has scattered the photon) is D p = 2(h/l ) sinq , leading to D p D x = [2(h/l ) sinq ][0.6l /sinq ] = 1.2 h , again validating Eq.(3). If we choose a small wavelength l in an attempt to minimize D x, we 'disturb' the particle more because the short-wavelength particle transfers a larger momentum. Therefore Eq.(3) is often interpreted by saying that by making a measurement on an object, we necessarily change it; the wavelike properties of the object deny us the option of performing measurements on an 'isolated' system.
Because the product
D pD x often comes out a factor of 5 or 6 larger than the lower limit prescribed by Eq.(3), the Heisenberg uncertainty principle is sometimes written as D p D x » h .
Another version of the uncertainty principle states:
D
E D t > h/2p ................ (4)
where
D E is the uncertainty in our knowledge of the energy of a particle, if its measurement is carried out over a time interval D t . Eq.(4) allows us to estimate the width of the emission lines emitted by atoms, for example in a gas-discharge lamp. The lifetime in the excited state is commonly of the order of 10^-8 seconds, so D E cannot be less than approximately 10^-34/10^-8 = 10^-26 J, or about 10^-7 eV. For visible light, the photon energy is two or three eV, so the fractional width D l /l must be at least three parts in 10^8 . If the pressure in the lamp is increased, collisions between gas atoms tend to reduce the lifetime D t and the spectral lines become broader.
The Wave-Particle Duality
Given two alternative descriptions of the same object (such as an electron) how do we know which is applicable in a given situation ? Answers to this question begin to emerge when we investigate simple situations such as diffraction of electrons by a double slit, sometimes called the Young's slit experiment. This experiment can be performed fairly easily with visible light (i.e. photons) and, with some technical difficulty, with electrons. The result is a series of parallel fringes, recorded on a plane behind the double slit; see Fig. 4.28. Using a wave picture for the incident radiation (electrons, photons, ... ) maxima in intensity should occur where there is constructive interference, corresponding to a path difference which is equal to an integral number of wavelengths: n
l = D sinq , where D is the separation of the two narrow slits. The experimental results are in agreement with this prediction, showing that the wave picture correctly describes the overall effect.
However, particle-like properties begin to emerge if we reduce the flux (intensity) of the incident radiation. The results for incident electrons are shown in Fig. 4.29. At low intensity (Figs. 4.18d), the recorded pattern takes on a spotty appearance, confirming that the electrons arrive as particles (each spot corresponds to the arrival of a single electron at the photographic emulsion or electron detector). However, the overall pattern still shows that the electrons tend to fall in bands, rather than uniformly across the detection plane, indicating that wavelike (diffraction) effects are still present. In fact, a simple calculation of the particle flux under low-intensity conditions reveals that the interference effects are occurring when only one particle is present within the apparatus at any one time. Using the particle view alone, this makes no sense: how can a single electron pass through both slits, as required to interfere with itself to give a nonuniform intensity distribution ?
The usual answer to this question is that the probability of arrival of a particle (at a particular location on the detector plane) is determined by the resultant intensity of two interfering matter waves, one (
Y 1) representing an electron passing through the upper slit, the other (Y 2) representing an electron passing through the lower slit. This is consistent with Max Born's probabilistic interpretation of the electron wavefunction Y . Strictly speaking, Y must be represented as a vector (or a complex number with real and imaginary parts), indicating that it contains phase as well as amplitude information. However, if we measure just the length of the vector, represented by | Y | , we obtain a scalar quantity whose square is proportional to the probability of finding the electron at a particular point in time and space.
It should be clear from the above interpretation that the wavefunction of a particle is a property of the particle and its surroundings (the apparatus, in the double-slit experiment). If we cover up one of the diffracting slits during the course of the experiment, we observe no interference fringes (only a broad "diffraction" pattern characteristic of a single slit) since we remove one of the matter waves. Likewise, if we have an oscillating shutter which alternately covers the upper and lower slits but leaves neither uncovered at the same time, we would see no interference.
According to the quantum mechanics, it is meaningless to say that an electron passes through one or other of the slits unless we measure it doing so. Therefore, we can retain the idea that the electron passes through both slits and is able to create an interference pattern. One way of measuring an electron passing through one of the slits might be to direct a light beam at the slit and observe the presence of an electron by detecting the deflection of a single photon due to its interaction with the electron (Compton scattering). However, such scattering transfers momentum to the electron (in a direction which is not predictable) and would alter the position at which the electron arrives at the detection plane. It is possible to estimate the average momentum transfer and to show that the interference fringes would be blurred out if we tried to detect the electrons in this way. Other ingenious ways of detecting the electron have also been analysed and found to destroy the interference effect. So we again see that it is not possible to do measurements on an isolated system; the act of measurement determines the properties which we observe.
The above discussion illustrates a general conclusion of quantum mechanics: that objects do not interact in a completely deterministic way. They sometimes exhibit particle-like properties (as when detected in a Geiger counter) and sometimes wavelike properties. Max Born described this dual nature as complementarity; both of our common-sense notions, of a particle and a wave, are needed to describe what matter actually is. We also conclude that, at the atomic level, it is not possible to separate the properties of an object from those of the measuring instrument.
Confined particles
The wavelike properties of an electron also show up when the particle is confined within a small region of space. Usually this confinement is by electrostatic forces, which can be visualised in terms of the corresponding electrostatic potential energy. The simplest case to consider is a one-dimensional square potential well, in which the potential energy rises abruptly to infinity at x=0 and x=L ; see Fig. 5.8c. Since an electron of finite kinetic energy (and wavenumber k) could not exist in the regions x<0 and x>L , its wavefunction must be zero at these boundaries. This places constraints on the wavefunction
Y , if represented by a simple sine or cosine function as in Eq.(1). An integral number of half-wavelengths must fit inside the well (Fig. 5.11a) or n l /2 = L , where the integer n is a quantum number. Fig. 5.11b shows the square of the magnitude of the wavefunction and indicates that the probability of finding an electron inside the well varies with its position x , in a way which depends on the value of n .If we define the potential energy to be zero inside the well, the total energy E of the electron is equal to its kinetic energy, given by:
E = (1/2) m v^2 = (p^2)/2m = (1/2m)(h/
l )^2 = n^2 h^2 /(8 m L^2) .............. (5)Wave mechanics tells us that the energy of the electron is quantised and that this energy increases (and the energy levels get more widely spaced) with increasing quantum number. The possibility n=0 can be ignored (it would correspond to zero wave function, i.e. no electron), so the lowest electron energy is not zero (as expected from classical physics) but E1 = h^2 /(8 m L^2), the zero-point energy. Even at T = 0 (in the absence of any thermal energy), the electron must have at least this much energy, which we might think of as vibrational energy (the electron travelling back and forth between reflections at x=0 and x=L).
A more realistic situation is represented in Fig. 5.12, where the potential energy rises to a value U outside the region 0<x<L . We assume E<U, since the electron is confined by the potential. Even so, the Schrodinger equation for
Y has solutions outside the well, although these decay rapidly (exponentially) away from the boundaries x=0 and x=L (Fig. 5.13). The wavefunction decays by a factor of e (2.718) over a distance called the penetration depth, whose value depends on U but is typically a fraction of a nanometer.
The conduction electrons in a metal are confined by an electrostatic potential not unlike that of Fig. 5.12. They are contained by a potential barrier at each surface whose height is U =
j , the work function. The electrons can escape if given energy in excess of j , for example by interaction with photons (photoelectric effect). However, the electrons can also "leak out" if a second solid is brought up very close to the first, with the gap between them less than about 1 nm (several times the penetration depth). This behaviour is called tunnelling and is a direct result of the wavelike nature of electrons.Tunnelling is utilized in several types of electronic device in which the electrons pass from one region to another, in defiance of classical physics. It is also the basic of the scanning tunnelling microscope (STM), a device invented in 1980 and since developed considerably, which is able to image surfaces with atomic resolution. A sharp metal tip is scanned aalong the surface of the specimen, while being held about 1 nm away. A small potential difference between the specimen and tip causes electrons to tunnel between them (see Fig. 3, p.257). The tunnelling current is very sensitive to the gap between tip and specimen, so that individual atoms can be seen in favourable circumstances (absence of vibration etc.).
Another situation where electrons are confined is in the vicinity of the nucleus of an atom. For a hydrogen atom, the potential is a hyperbolic-shaped well: U = k(-e)(Ze)/r where k is the Coulomb constant. Solution of the Schrodinger equation yields wavefunctions of somewhat complicated mathematical form, known as orbitals. Their three-dimensional shape depends on the values of several quantum numbers (n, l, m) but the corresponding energy levels depend only on the principal quantum number n . The energies turn out to be identical to those predicted by the Bohr model of the hydrogen atom, and are therefore in agreement with spectral data. For the ground state (n=1) the electron orbital is a spherical cloud surrounding the nucleus, for which the probability density is a maximum (Fig. 7.10) at a radius r = a0, which is just the radius of the first Bohr orbit. But in place of the precise electron orbits envisaged by the Bohr model, we have delocalised orbitals which represent the probability of finding the electron at a particular point in space.
Electrons in atoms other than hydrogen have similar properties, the wavefunction of an electron determining its spatial localization. The Pauli exclusion principle states that no two electrons can share the same wavefunction, so the electrons in a multielectron atom can be classified in of their different energies. Those having the highest energy are furthest (on average) from the nucleus and their orbitals can overlap with those of neighbouring atoms, forming chemical bonds. In fact, quantum mechanics provides a complete description for the chemical properties of the different elements, including their position in the periodic table.