10.3 Wave Nature of Matter

Learning Objectives

By the end of this section, you will be able to:

  • Extend the concept of wave–particle duality that was observed in electromagnetic radiation to matter as well

Bohr’s model explained the experimental data for the hydrogen atom and was widely accepted, but it also raised many questions. Why did electrons orbit at only fixed distances defined by a single quantum number n = 1, 2, 3, and so on, but never in between? Why did the model work so well describing hydrogen and one-electron ions, but could not correctly predict the emission spectrum for helium or any larger atoms? To answer these questions, scientists needed to completely revise the way they thought about matter.

Wave Behaviour of Matter from a Microscopic Perspective

We know how matter behaves in the macroscopic world – objects that are large enough to be seen by the naked eye follow the rules of classical physics. A billiard ball moving on a table will behave like a particle: It will continue in a straight line unless it collides with another ball or the table cushion, or is acted on by some other force (such as friction). The ball has a well-defined position and velocity (or a well-defined momentum, p = mv, defined by mass m and velocity v) at any given moment. In other words, the ball is moving in a classical trajectory. This is the typical behaviour of a classical object.

When waves interact with each other, they show interference patterns that are not displayed by macroscopic particles such as the billiard ball. For example, interacting waves on the surface of water can produce interference patterns similar to those shown on Figure 10.3a. This is a case of wave behaviour on the macroscopic scale, and it is clear that particles and waves are very different phenomena in the macroscopic realm.

A photograph of ripples in water as two waves meet and interfere with one another. The water has concentric rings of lighter and darker areas.
Figure 10.3a: Interference patterns on water surface: an interference pattern on the water surface is formed by interacting waves. The waves are caused by the reflection of water from the rocks. (credit: work by Daniela_deGol, CC BY 4.0)

As technological improvements allowed scientists to probe the microscopic world in greater detail, it became increasingly clear by the 1920s that very small pieces of matter follow a different set of rules from those we observe for large objects. The unquestionable separation of waves and particles was no longer the case for the microscopic world.

One of the first people to pay attention to the special behaviour of the microscopic world was Louis de Broglie. He asked the question: If electromagnetic radiation can have particle-like character, can electrons and other submicroscopic particles exhibit wavelike character? In his 1925 doctoral dissertation, de Broglie extended the wave–particle duality of light that Einstein used to resolve the photoelectric-effect paradox to material particles. He predicted that a particle with mass m and velocity v (that is, with linear momentum p) should also exhibit the behaviour of a wave with a wavelength value λ, given by this expression in which h is the familiar Planck’s constant:

[latex]\lambda = \frac{h}{mv} = \frac{h}{p}[/latex]

This is called the de Broglie wavelength. Unlike the other values of λ discussed in this chapter, the de Broglie wavelength is a characteristic of particles and other bodies, not electromagnetic radiation (note that this equation involves velocity [v, m/s], not frequency [ν, Hz]. Although these two symbols are identical, they mean very different things). Where Bohr had postulated the electron as being a particle orbiting the nucleus in quantized orbits, de Broglie argued that Bohr’s assumption of quantization can be explained if the electron is considered not as a particle, but rather as a circular standing wave such that only an integer number of wavelengths could fit exactly within the orbit (Figure 10.3b).

This figure includes a circle formed from a dashed line. A sinusoidal wave pattern indicated with a solid red line is wrapped around the circle, centered about the edge of the circle. Line segments extend outward from the circle extending through 2 wave crests along the circle. A double ended arrow is drawn between these segments and is labeled, “wavelength, lambda.” A dashed double headed arrow is drawn from the center to the edge of the circle and is labeled, “radius r.”
Figure 10.3b: de Broglie’s electron circulating standing wave theory: if an electron is viewed as a wave circling around the nucleus, an integer number of wavelengths must fit into the orbit for this standing wave behaviour to be possible (credit: Chemistry (OpenStax), CC BY 4.0).

For a circular orbit of radius r, the circumference is 2πr, and so de Broglie’s condition is:

[latex]2\pi r = n\lambda , n = 1, 2, 3, \cdots[/latex]

Since the de Broglie expression relates the wavelength to the momentum and, hence, velocity, this implies:

[latex]2 \pi r = n\lambda = \frac{nh}{p} = \frac{nh}{mv} = \frac{nhr}{mvr} = \frac{nhr}{L}[/latex]

This expression can be rearranged to give Bohr’s formula for the quantization of the angular momentum:

[latex]L = \frac{nh}{2\pi} = n{\hbar}[/latex]

Classical angular momentum L for a circular motion is equal to the product of the radius of the circle and the momentum of the moving particle (Figure 10.3c).

[latex]L = rp = rmv \;\text{(for a circular motion)}[/latex]
The diagram shows a blue circle. At the center, there is an arrow labeled, “L,” which points upward. Another arrow labeled, “r,” points from the center to the edge of the circle. Another arrow labeled, “m times v” extends from the point where the r-labeled arrow reaches the edge of the circle.
Figure 10.3c: Angular momentum for circular motion: the diagram shows angular momentum for a circular motion. L = rp = rmv (credit: Chemistry (OpenStax), CC BY 4.0).

Shortly after de Broglie proposed the wave nature of matter, two scientists at Bell Laboratories, C. J. Davisson and L. H. Germer, demonstrated experimentally that electrons can exhibit wavelike behaviour by showing an interference pattern for electrons travelling through a regular atomic pattern in a crystal. The regularly spaced atomic layers served as slits, as used in other interference experiments. Since the spacing between the layers serving as slits needs to be similar in size to the wavelength of the tested wave for an interference pattern to form, Davisson and Germer used a crystalline nickel target for their “slits,” since the spacing of the atoms within the lattice was approximately the same as the de Broglie wavelengths of the electrons that they used. Figure 10.3d shows an interference pattern.

This figure has two parts. Part a shows a diagram of an electron source emitting waves that pass through two narrow slits in a barrier. A wave interference pattern results on the opposite side of the barrier that result in evenly spaced, relatively wide horizontal grey lines on a black surface. The black surface is placed at a distance beyond the barrier. Part b shows an electron source at the far left. Two gold arrows point toward two narrow horizontal slits in a barrier. Around these arrows are many small golden dots. On the right side of the barrier, the gold dots are more widely dispersed. A black surface a distance beyond the barrier shows evenly, yet widely dispersed gold dots. An arrow is present below this black surface and is labeled, “Time.” A second black surface is shown to the right that has many more golden dots that appear to be organizing into a pattern of evenly spaced horizontal line segments. A third black surface is shown even further to the right in which many more gold dots are shown in a very clearly established, evenly spaced pattern of horizontal line segments.
Figure 10.3d: Experiments whowing interference pattern for electrons: (a) The interference pattern for electrons passing through very closely spaced slits demonstrates that quantum particles such as electrons can exhibit wavelike behaviour. (b) The experimental results illustrated here demonstrate the wave–particle duality in electrons. The electrons pass through very closely spaced slits, forming an interference pattern, with increasing numbers of electrons being recorded from the left image to the right. With only a few electrons recorded, it is clear that the electrons arrive as individual localized “particles,” but in a seemingly random pattern. As more electrons arrive, a wavelike interference pattern begins to emerge. Note that the probability of the final electron location is still governed by the wave-type distribution, even for a single electron, but it can be observed more easily if many electron collisions have been recorded (credit: Chemistry (OpenStax), CC BY 4.0).

The wave–particle duality of matter can be seen in Figure 10.3d by observing what happens if electron collisions are recorded over a long period of time. Initially, when only a few electrons have been recorded, they show clear particle-like behaviour, having arrived in small localized packets that appear to be random. As more and more electrons arrived and were recorded, a clear interference pattern that is the hallmark of wavelike behaviour emerged. Thus, it appears that while electrons are small localized particles, their motion does not follow the equations of motion implied by classical mechanics, but instead it is governed by some type of a wave equation that governs a probability distribution even for a single electron’s motion. Thus the wave–particle duality first observed with photons is actually a fundamental behaviour intrinsic to all quantum particles.

Watch Dr. Quantum – Double Slit Experiment (5min 12s)

Video Source: Angel Art. (2010, December 27). Dr. Quantum – Double slit experiment [Video]. YouTube.

Example 10.3a

Calculating the Wavelength of a Particle

If an electron travels at a velocity of 1.000 × 107 m s–1 and has a mass of 9.109 × 10–28 g, what is its wavelength?

Solution

We can use de Broglie’s equation to solve this problem, but we first must do a unit conversion of Planck’s constant. You learned earlier that 1 J = 1 kg m2/s2. Thus, we can write h = 6.626 × 10–34 J s as 6.626 × 10–34 kg m2/s.

[latex]\lambda = \frac{h}{mv}[/latex]
[latex]\begin{array}{l} = \frac{6.626 \times 10^{-34} \;\text{kg m}^2 \text{/s}}{(9.190 \times 10^{-31} \;\text{kg})(1.000 \times 10^7 \;\text{m/s})} \\[1em] = 7.274 \times 10^{-11} \;\text{m} \end{array}[/latex]

This is a small value, but it is significantly larger than the size of an electron in the classical (particle) view. This size is the same order of magnitude as the size of an atom. This means that electron wavelike behaviour is going to be noticeable in an atom.

Exercise 10.3a

Calculate the wavelength of a softball with a mass of 100 g traveling at a velocity of 35 m s–1, assuming that it can be modelled as a single particle.

Check Your Answer[1]

We never think of a thrown softball having a wavelength, since this wavelength is so small it is impossible for our senses or any known instrument to detect (strictly speaking, the wavelength of a real baseball would correspond to the wavelengths of its constituent atoms and molecules, which, while much larger than this value, would still be microscopically tiny). The de Broglie wavelength is only appreciable for matter that has a very small mass and/or a very high velocity.

Werner Heisenberg considered the limits of how accurately we can measure properties of an electron or other microscopic particles. He determined that there is a fundamental limit to how accurately one can measure both a particle’s position and its momentum simultaneously. The more accurately we measure the momentum of a particle, the less accurately we can determine its position at that time, and vice versa. This is summed up in what we now call the Heisenberg uncertainty principle: It is fundamentally impossible to determine simultaneously and exactly both the momentum and the position of a particle. For a particle of mass m moving with velocity vx in the x direction (or equivalently with momentum px), the product of the uncertainty in the position, Δx, and the uncertainty in the momentum, Δpx , must be greater than or equal to [latex]\frac{\hbar}{2}[/latex] (recall that [latex]\hbar = \frac{h}{2 \pi}[/latex], the value of Planck’s constant divided by 2π).

[latex]\Delta x \times \Delta p_x = (\Delta x)(m \Delta v) \ge \frac{\hbar}{2}[/latex]

This equation allows us to calculate the limit to how precisely we can know both the simultaneous position of an object and its momentum. For example, if we improve our measurement of an electron’s position so that the uncertainty in the position (Δx) has a value of, say, 1 pm (10–12 m, about 1% of the diameter of a hydrogen atom), then our determination of its momentum must have an uncertainty with a value of at least

[latex][\Delta p = m\Delta v = \frac{h}{(2 \Delta x )}] = \frac{(1.055 \times 10^{-34} \;\text{kg m}^2/\text{s})}{(2 \times 1 \times 10^{-12} \;\text{m})} = 5 \times 10^{-23} \;\text{kg m/s}[/latex]

The value of ħ is not large, so the uncertainty in the position or momentum of a macroscopic object like a baseball is too insignificant to observe. However, the mass of a microscopic object such as an electron is small enough that the uncertainty can be large and significant.

It should be noted that Heisenberg’s uncertainty principle is not just limited to uncertainties in position and momentum, but it also links other dynamical variables. For example, when an atom absorbs a photon and makes a transition from one energy state to another, the uncertainty in the energy and the uncertainty in the time required for the transition are similarly related, as [latex]\Delta E \; \Delta t \ge \frac{\hbar}{2}[/latex]. Even the vector components of angular momentum cannot all be specified exactly simultaneously.

Watch The Basics of Quantum Mechanics: What is the Heisenberg Uncertainty Principle? (4min 43s).

Heisenberg’s principle imposes ultimate limits on what is knowable in science. The uncertainty principle can be shown to be a consequence of wave–particle duality, which lies at the heart of what distinguishes modern quantum theory from classical mechanics. Recall that the equations of motion obtained from classical mechanics are trajectories where, at any given instant in time, both the position and the momentum of a particle can be determined exactly. Heisenberg’s uncertainty principle implies that such a view is untenable in the microscopic domain and that there are fundamental limitations governing the motion of quantum particles. This does not mean that microscopic particles do not move in trajectories, it is just that measurements of trajectories are limited in their precision. In the realm of quantum mechanics, measurements introduce changes into the system that is being observed.

The modern model for the electronic structure of the atom is based on recognizing that an electron possesses particle and wave properties, the so-called wave–particle duality.

Attribution & References

Except where otherwise noted, this page is adapted by Jackie MacDonald from:

  1. 1.9 × 10–34 m
definition

License

Icon for the Creative Commons Attribution 4.0 International License

Enhanced Introductory College Chemistry Copyright © 2023 by Gregory Anderson; Caryn Fahey; Jackie MacDonald; Adrienne Richards; Samantha Sullivan Sauer; J.R. van Haarlem; and David Wegman is licensed under a Creative Commons Attribution 4.0 International License, except where otherwise noted.

Share This Book