Você está na página 1de 12

16.

Duality of Matter
When Louis-Victor-Pierre-Raymond de Broglie (1892-1987) wrote his doctoral thesis in 1923, he proposed a radical new idea with implications he, himself, did not fully appreciate. De Broglie, a graduate student of noble French background, was guided by personal intuition and mathematical analogies rather than by any experimental evidence when he posed what turned out to be a very crucial question: Is it possible that particles, such as electrons, exhibit wave characteristics? De Broglies arguments led to the conclusion that electrons had a wavelength given by wavelength Plancks constant . (mass of an electron)(speed of an electron)

De Broglie Wave Nucleus Allowed Bohr Orbit

Indeed, Plancks constant is a very small number and, for reasonable speeds of an electron, the equation always results in a very short wavelength (of the order of atomic dimensions). In order for a wave to exhibit the characteristic behavior of interference and diffraction, it must pass through openings of about the same dimension as the wavelength, i.e., in the case of electrons, through openings about as wide as an atom or the spacing between atoms. If one were to calculate from de Broglies relationship the wavelength associated with a baseball, one would find a prediction of the order of 1034 meters. This small number follows from de Broglies relationship because it has the mass of the object in the denominator (very large for a baseball) and Plancks constant in the numerator (a very small number). Waves so short would exhibit diffraction and interference only if they went through very narrow openings (about 1034meters). Since the present limit of measurability is about 1015 meters, we can conclude that for macroscopic objects like baseballs, the waves might just as well not exist. The next step was to somehow bring together Bohrs atom (with its success at predicting the atomic spectra) and de Broglies matter waves. The success of Bohrs atom hinged on the idea that there were certain discrete orbits with associated discrete energies. But what determines which orbits are allowed and which are not? Perhaps, someone thought, it is de Broglies wave. Perhaps the only orbits allowed are those that an integral number of de Broglie electron wavelengths would fit into! (See Fig.

De Broglie Wave

Nucleus Disallowed Bohr Orbit

16.1) Figure 16.1. De Broglies matter waves explain why only certain orbits are possible in Bohrs atom. The waves must just fit the orbit. This idea was a great success. It predicted just the right orbits and just the right energy levels to explain the light spectrum for hydrogen (though it was less successful for helium, lithium, etc.). Certainly it was mathematically equivalent to the idea that Bohr had originally used (which did not use waves), but it was also radically different because it introduced a powerful new idea: the electrons surrounding the nucleus of the atom are some kind of wave whose wavelength depends on the mass and speed of the electron. But waves of what? When we think of the waves of our everyday experience, we think of disturbances propagating in a medium as, for example, waves on the

143

ocean moving through the water. We have introduced waves, wavelengths, and frequencies without ever addressing the fundamental question: waves of what? The Two-Slit Experiment There is probably no stronger evidence for the wave nature of light and electrons than the so-called two-slit experiment in which electrons (or photons) pass through a double slit arrangement to produce an interference pattern on the screen behind. The experiment is simple and straightforward and, if we can understand that, then we ought to be able to understand wave-particle duality (if it can be understood at all). One of the most enlightening explanations of quantum mechanics (the name given to the general area of atomic modeling that we are discussing) was given in a series of lectures by Nobel laureate Richard Feynman and reproduced in his book, The Character of Physical Law. He discusses the two-slit experiment by contrasting the experiment using indisputable particles, then using indisputable waves, and finally using electrons. To better understand the two-slit experiment, imagine a shaky machine gun that fires bullets at a two-slit arrangement fashioned out of battleship steel (Fig. 16.2). Bullets are clearly particles. They are small, localized structures that can be modeled as tiny points. In our experiment we shall assume that the bullets do not break up. The machine gun is a little shaky, so the bullets sometimes go through one slit, sometimes the other. They can ricochet from the edges of the slits, so they can arrive at various positions behind them.

Positioned behind the slits is a bucket of sand to catch the bullets. It is placed first in one position and the number of bullets arriving in a given amount of time is counted. The bucket is then moved to an adjacent position, and the process is repeated until all the possible arrival positions behind the slits have been covered. Then a piece of paper is used to make a graph of the number of bullets which arrive as a function of the position of the bucket. The result is a double-peaked curve that reflects the probability of catching bullets at various positions behind the slits. The two peaks correspond to the two regions of high probability that lie directly behind the open slits. When the experiment is repeated with one of the slits closed, the curve has only one peak. We will refer to the plotted curves as probability curves. Now imagine an analogous experiment performed with waves (Fig. 16.3). Visualize long straight waves moving along the length of a pan of water. Into the path of the waves we will place an obstacle with two slits. In doing so we set up the classical demonstration of wave interference. Behind the slits and along a straight line paralleling the barrier but some distance behind, we will observe the waves. Yet waves are not particles; it doesnt make any sense to measure the probability of arrival of a wave at some particular point. In fact, the wave arrives spread out over many points along the backdrop. So rather than even trying to measure a probability curve, we will observe the amplitude of the wave at various positions along the backdrop by placing a cork in the water and observing the extent of the vertical motion of the cork. Squaring the amplitude gives the intensity of the wave. The result is a multipeaked curve. The peaks mark the regions of constructive interference; the valleys mark the regions of destructive interference. If one slit is closed, the interference largely disappears and

Figure 16.2. A two-slit experiment using particles (bullets) fired by a rickety machine gun. The probability of arrival of the bullets at the backdrop is described by a double-peaked curve. Each peak is roughly behind one of the open slits.

Figure 16.3. A two-slit experiment using waves. The graph with the peaks represents the intensity of the waves along the backdrop as might be observed by watching a cork at various positions. The peaks in the curve are points of constructive interference; the valleys are points of destructive interference.

144

a single peaked curve reminiscent of the single-peaked curve for the bullets is observed. We refer to these curves as interference curves. Now we can try the experiment with electrons (Fig. 16.4). But which experiment? If the electrons are particles, then we try to measure the arrival of little lumps at spots behind the screen and try to plot a probability curve. If the electrons are waves, we try to measure the amplitude of some disturbance and plot an interference curve. What, then, is the electron? Some experiments do indicate that electrons behave like little particles. So we proceed to set up a two-slit experiment for electrons and design a detector to place behind the slits to play the role the bucket of sand played for bullets. Note, however, that the wavelengths of electrons (according to de Broglies formula) are very short and the slits will have to be similar to the spacing between layers of atoms in a crystal. As Richard Feynman expressed, this is where nature provides us with a very strange and unexpected result: That is the phenomenon of nature, that she produces the curve which is the same as you would get for the interference of waves. She produces this curve for what? Not for the energy in a wave but for the probability of arrival of one of these lumps (Richard Feynman, The Character of Physical Law, p. 137).

arrival of electrons that must be coming through the remaining open slit. Now open the closed slit. The detector stops counting. How do electrons going through one slit know whether the other slit is open or closed? If the electrons are little lumps, how do the lumps get canceled so that nothing arrives at the detector? You can slow down the rate of the electrons so that only one electron at a time is firedand they still know. The interference curve, in fact, becomes a composite of all the arrivals of the individual electrons. It is very puzzling. Waves of Probability The presently accepted explanation is that the waves associated with electrons are waves of probability. They are not real disturbances in a medium. The waves themselves are nothing more than mathematical descriptions of probability of occurrence. And perhaps we wouldnt take them too seriously if Erwin Schrdinger (1887-1961) hadnt devised the equation that bears his name. Without some mathematical sophistication, Schrdingers equation doesnt mean much. But it is an equation that describes how waves move through space and time. The waves it describes may have peaks and valleys that correspond to the high or low probability of finding an electron at a particular place at a given time. The peaks and valleys move and the equation describes their movements. When we imagine firing an electron at the two slits, we visualize such a wave being produced, and we imagine it propagating toward and through the two slits and approaching the screen behind. All of this motion of the probability wave is described by Schrdingers equation. The motion of the wave of probability is deterministic. Now imagine the wave of probability positioned just in front of the detecting screen with its peaks and valleys spread out like an interference curve. At this point something almost mystical happens that no one can predict. Out of all the possibilities represented by the spread-out probability curve, one of them becomes reality and the electron is seen to strike the screen at a particular, small spot. No one knows how this choice is made. Theories that have tried to incorporate some way of deciding how the probabilities become realities have always failed to agree with experiments. It is as if some giant dice-roller in the sky casts the dice and reality rests on the outcome. Electron Microscope Yet the diffraction and interference patterns displayed by electrons seem to be real enough. If they are real in at least some sense, they offer the potential to solve a very important problem. Microscopes are limit-

Figure 16.4. A two-slit experiment using electrons. The probability of arrival of the particles (electrons) takes the form of the graph that describes the intensity of waves. As Feynman wrote, this is very strange indeed. Look at one of the valleys in the probability curve, a point wherewith both slits openthere are no detected electrons (see point A in Fig. 16.4). Then imagine slowly closing one of the slits so that the valley of probability goes away and the detector begins to count the

145

ed in what they can see by the effects of diffraction. Typical microscopes use visible light. When the wavelength of the light being used is about the same measure as the size of the objects being viewed, the diffraction of the light around the edges of the objects becomes substantial and the images get fuzzy. The microscope cannot resolve the detail and has reached its fundamental limit (see Fig. 16.5). The only way the problem can be solved is to use shorter wavelengths of viewing light. Short-wavelength x-rays would work, but nothing is able to bend or focus short-wavelength x-rays in the way that a glass lens bends and focuses rays of visible light.

electrons could be accelerated to shorten the wavelength, and if they could be focused, we would have the makings of a very high-resolution microscopean electron microscope. The electron accelerator (or electron gun) is made of two parts: an electrically heated wire and a grid, which is basically a small piece of window screen. When the grid is positively charged, it attracts electrons from the hot wire. The electrons are accelerated toward the grid, and most of them pass through its holes, as diagramed in Figure 16.6. The electrons achieve more speed (shorter wavelengths) as the grid is given more positive charge. In the microscope itself, the electron beam is focused and its diameter is magnified by magnets, as shown in Figure 16.7. The image is made visible when the electrons strike a glass plate (the screen) that is coated with a material that glows when struck by energetic particles. Because the wavelengths of the electrons can be made very short, the resolution of the electron microscope is much better than the resolution that can be achieved by visible light microscopes. (See Fig. 16.8; see also Fig. 5.7 and Color Plates 1 and 2.)

Figure 16.5. Upper: Two small objects that are to be viewed by a microscope. Lower: When the wavelength of the viewing light is the same size as the objects, diffraction causes a loss of resolution. This is where electron waves come to the rescue. According to de Broglies equation, electrons have very short wavelengths. Indeed, they can be made even shorter by increasing the speed of the electrons. If the

Figure 16.7. Diagram of an electron microscope. Actual height is about 2 meters. The (Heisenberg) Uncertainty Principle Figure 16.6. Electron source. through the grid. Most electrons pass The wave nature of matter raises another interesting problem. To what extent is it possible to determine

146

Figure 16.8. Scanning electron micrograph of a human whisker at the position of a particle? The probability of locating a particle at a particular point is high where the wave function is large and low where the wave function is small. Imagine such a wave as depicted in Figure 16.9. This kind of wave might be called a wave packet. If

500 magnification. (Courtesy of W. M. Hess)

Figure 16.9. A moving electron might be represented as a localized wave or wave packet. In each case, the electron can only be specified as being somewhere inside the region bounded by the dashed lines.

we were to look, we would expect to find the particle within the space occupied by the packet, but we dont know where with certainty. The uncertainty of the position, which we will designate as x, is roughly the length of the region in which there are large peaks and valleys. If the region is broad (x is large), then the position is quite uncertain; if the region is narrow (x is small), then the position is less uncertain. In other words, the position of particles is specified more precisely by narrow waves than by broad waves. This limiting precision is decreed by nature. Since the packet can be no shorter than the wavelength, the wavelength itself sets the limit of precision. The speeds of particles have probabilities and uncertainties in just the same way that positions have probabilities and uncertainties. For technical reasons we will multiply the uncertainty in speed by the mass of the electron and call the resulting quantity p. The two uncertainties are related through the Schrdinger equation in a relationship called the (Heisenberg) Uncertainty Principle: (x) (p) is greater than Plancks constant .

147

Figure 16.10. A beam of electrons diffracts as it passes through a narrow slit. How would the pattern change if the slit were even narrower? What the relationship means is that experiments designed to reduce the uncertainty in the position of a particle always result in loss of certainty about the speed of the particle. The opposite is also true: Experiments designed to reduce the uncertainty in the speed of a particle always result in a loss of certainty about the position of the particle. Because Plancks constant is small, these two uncertainties can still be small by everyday standards; but when we get down to observations of the atomic world, the Uncertainty Principle becomes an important factor. Consider, for example, the electrons passing through a single slit as in Figure 16.10. The arrival pattern of the many electrons passing through the slit is broad. Just as for light, the narrower the slit, the broader the pattern. When we make the slit narrow, we also make the uncertainty in the horizontal position of the electron small, as it passes through the slit. We know the position of the electron to within the width of the slit as it passes through. But making the slit narrower means that the uncertainty in the horizontal speed of the electron must get larger so that the Uncertainty Principle is satisfied. We no longer know precisely where it is headed or how much horizontal motion it has acquired by interacting with the slit. It is this large uncertainty in the horizontal speed that makes it impossible to predict precisely where the electron will land. This causes the pattern to spread. We can think of the spreading as a consequence of the Uncertainty Principle. What is Reality? The Newtonian clockwork was a distressingly deterministic machine. The Second Law of Thermodynamics said the clock was running down. Quantum mechanics threatens to make the world more a slot machine than a clock. The Uncertainty Principle prevents us from predicting with certainty the future of an individual particle. In the Newtonian view the future was exactly predictable. If the position, speed and direction of motion of a particle were known, the Newtonian laws of motion would predict the future. Indeed, by following Newtonian ideas and using computer programs, scientists can predict the motions of planets for thousands of years into the future. But to make Newtonian physics work for electrons, we have to know exactly where the electron is and, simultaneously, its speed and direction of motion. This is precisely what the Uncertainty Principle says we fundamentally cannot know. We can know one or the other, but not both together. Thus, we cannot predict the future of the electron. Therefore, when an electron is fired at the two slits, we cannot predict exactly where it will land on the screen behind. The best we can do is to know the probabilities associated with the wave function. Bohr saw the position and the speed as complementary descriptors of the electron. But no one could know both with precision at the same time. He, therefore, denied reality to a description that specifies both at the same time. Recall the two-slit experiment. Imagine again a wave function that describes an electron fired by a gun toward the slits and a screen behind. Imagine the wave function with its hills and valleys undulating through space. The wave function (probability curve) contains all the information about the possibilities of where the electron might land for any electron fired through the slits to a screen behind. Not only does the wave function contain the possibilities about where the electron should finally land on the screen, but it also contains a description of the probabilities to assign to those possibilities. Now imagine that the wave function reaches the screen. What we would see is a tiny spot where the electron strikes the screen. The spot is much smaller than the space over which the wave was actually extended. The wave is spread over the entire pattern of spots that eventually becomes the interference pattern. When the electron is revealed as it strikes the screen, the pattern of probabilities changes drastically and immediately. Suddenly the probability becomes 1.0 at the spot of observation and 0.0 everywhere else. Of all the possibilities, only one has become reality. For

148

the next electron through the slits, a different possibility will become reality. This sudden change is called the collapse of the wave function. The collapse of the wave function is a great mystery. One interpretation (by Max Born in 1920) is that the collapse is not a physical change but, rather, our knowledge of the system suddenly changes. Humans, in their capacity as observers, become an essential feature of physics. In science, reality is what we see or detect with the help of instruments. As long as the wave function moves unobserved through space with all of its possibilities and probabilities, we cannot attribute reality to the electrons position, its speed, or even its existence. Is it the seeing and observing itself which turns the possibilities into a single reality? If this view is correct (and it is surely debatable), then reality is literally created by the act of conscious observation! Without the observation, the world remains unrealized possibility and probability. In a sense, we can say that the electron moves from place to place as a wave of probability. Yet it is always observed as a particle. When we dont observe it, we cannot say what its nature is. We cannot ascribe reality to any of the usual descriptions of position, speed, energy, and so forth. Niels Bohr defended and enlarged the above viewpoint into what has come to be called the Copenhagen Interpretation of quantum mechanics. To Einstein, who believed in a world that existed completely independently of mans consciousness, the interpretation was anathema; he argued and fought against it with tenacity in what came to be a losing battle. The whole idea is in violent contradiction to the rigidly deterministic Newtonian view we discussed earlier. The Role of the Newtonian Laws The Newtonian view has merit even though we now know it is incomplete. Its amazing success when applied to a wide variety of phenomena should indicate that it is at least close to being an accurate description of nature. How can we reconcile the quantum mechanical view of change in the universe with the older, but successful, exact determinism of Newtonian physics? Quantum mechanics and Newtonian laws give exactly the same results in the domain in which the Newtonian laws have been successful. Suppose we were to use the usual laws of motion and gravitation to predict the position of the earth in its motion about the sun. We would assign a specific location and motion for the earth at the present time, and then predict exact locations and motions for later times. Now suppose we were to use quantum mechanics and the modern view to describe the same motion. We could then predict probabilities for the arrival of the earth at certain locations at particular times. The result would be a high probabili-

ty, essentially 100 percent, of finding the earth at the locations predicted by the Newtonian laws. We would find low probability, nearly zero percent, of finding it anywhere else. The two views predict exactly the same future locations. However, the situation is quite different if we try to predict the motion of an electron inside an atom. Here we cannot even begin by using the Newtonian laws. We cannot assign an exact position to the electron because of its wave nature. Since its position is not known, there is no way to calculate accurately the force acting on it and thus no way to predict its acceleration or its future position. In this domain, quantum mechanics provides the only way to predict the motion and future course of the electron wave. Consider Figure 16.11. The sizes of things are listed along a horizontal axis; their speeds are listed along the vertical axis. The things we normally encounter would occur near the middle of the range of possible sizes and near the low end of possible speeds. Such objects and such motions form the basis of our intuitive feelings about nature. The Newtonian laws accurately describe the motions of such objects that seem reasonable and proper. Since we rarely deal directly with small things, such as atoms and their constituent particles, we should not be surprised that their rules of behavior are different from those to which we are accustomed. The laws governing atoms represent refinements of the Newtonian laws. We could use any or all of the laws of quantum mechanics to calculate and predict the motions of objects we normally encounter and find the same answer as if we had used Newtons laws. The additional sophistication is simply not necessary for an accurate description of the things with which we ordinarily deal. Only when we wish to understand the behavior of things exceedingly small or fast do we need the additional insights of the extended laws. In this sense, the Newtonian laws are approximations of the more complete laws; however, as approximations, they are totally adequate when limited to objects with appropriate size and speed. Summary Matter has a dual natureparticle and wave. Its particulate nature dominates the behavior of large-scale objects; the wave nature dominates small-scale (submicroscopic) behavior. The wave properties of submicroscopic matter limit the motion of particles in such a way that the particles cannot have unique values for both position and speed simultaneously. This limitation is formalized by the Uncertainty Principle. The wave nature of matter also implies that changes in any part of the universe are statistical processes, and the laws we have discovered allow us to predict the probability that certain kinds of changes will occur. Newtons laws of

149

Figure 16.11. The Newtonian laws are adequate for describing the motions of objects whose size and speed are within our range of experience. More precise laws must be used outside that range. motion are ways to calculate with high probability the kinds of motion that large objects will undergo. Electrons in atoms behave almost entirely as waves. These wave properties determine the structure of atoms and the ways that each atom interacts with its neighbors. Electron wave interactions thus regulate the kinds of collections of atoms that can occur, and also determine the properties of bulk matter. Historical Perspectives Some scoffed at de Broglies thesis, labeled it a French comedy, and expressed reservations about accepting it. But by chance de Broglies thesis came to the attention of Albert Einstein, who in 1905 had proposed the revolutionary idea of the dual nature of light in order to explain the photoelectric effect. Therefore, it was natural for Einstein to be drawn to de Broglies proposal for a dual nature of matter. By 1923 Einstein was quite famous, and the scientific world hung on his every word. Einstein drew attention to de Broglies thesis in one of his own papers, restated the argument in a forceful way, and bolstered the conjecture with additional arguments. Thus, the idea came to the attention of Erwin Schrdinger, who developed a wave model of the atom based on the wave equation that bears his name and is synonymous with quantum mechanics. Yet de Broglie still had not one shred of solid experimental evidence that electrons behaved like waves. By 1925 the crucial experiments had already been done by an American, Clinton Davisson, and, independently, by George Thomson. (Ironically, George Thomsons father, J. J. Thomson, received a Nobel Prize [1906] for crucial experiments in demonstrating the particle nature of the electron, and his son, George, later received a Nobel Prize [1937] for showing the entirely opposite wave nature of the electron.) But in 1925, Davisson had not heard of de Broglies idea and was struggling to explain puzzling results that he had acquired quite by accident. In his experiments he was observing the scattering of electrons as they encountered a piece of nickel.

150

However, quite by accident, the surface of his piece of nickel became oxidized, and he was forced to interrupt the experiment to heat the piece of metal to restore its condition. In doing so, and without knowing, some areas of the nickel on the surface crystallized, forming the regular layered structure of a crystal. The spaces between the layers became perfect slits of just the right dimension for electron waves to diffract and interfere; this demonstrated their wave nature. But for Davisson the interference pattern that appeared was a puzzle, an irritating failure as he put it. Nevertheless, he was alert to a possible discovery and tried with theory after theory to explain the results, until he was led by discussions with European physicists to de Broglies work. De Broglie won the Nobel Prize for his French comedy in 1929 and Davisson for elaborations and refinements of his irritating failure in 1937. (Adapted from Barbara Lovett Cline, Men Who Made A New Physics, pp. 152-156.) STUDY GUIDE Chapter 16: Duality of Matter A. FUNDAMENTAL PRINCIPLES 1. Wave-Particle Duality of Matter: Matter in its finest state is observed as particles (electrons, protons, quarks), but when unobserved (such as moving from place to place) is described by waves of probability. 2. Wave-Particle Duality of Electromagnetic Radiation: See Chapter 14. 3. The (Heisenberg) Uncertainty Principle: The product of the uncertainty in the position of an object and the uncertainty in its momentum is always larger than Plancks constant. (The momentum of an object is its mass times its speed; thus, uncertainty in momentum of an object of fixed mass is an uncertainty in speed.) B. MODELS, IDEAS, QUESTIONS, OR APPLICATIONS 1. What was the de Broglie hypothesis? 2. What pattern would be observed if a rickety machine gun fired bullets through two closely spaced slits in a metal sheet? 3. What pattern would be observed if water waves were allowed to pass through two openings in a dike? 4. What pattern would be observed if electrons were allowed to pass through two very closely spaced slits? What would the pattern be like if the electrons were sent through the device one at a time? If one of the holes were closed, what pattern would develop? 5. Is there evidence to support the position that matter has a particle aspect?

6. 7.

Is there evidence to support the position that matter has a wave aspect? What is uncertain in the Uncertainty Principle, and can the uncertainty be eliminated with more careful experiments?

C. GLOSSARY 1. Interference Curve: In the context of this chapter, we mean a mathematical graph which is a measure of the squared amplitude of waves in a region where wave interference is taking place. 2. Plancks Constant: See Chapter 14. 3. Probability Curve: In the context of this chapter, we mean a mathematical graph (like the famed bell-shaped curve that describes the probability of having a particular IQ) which describes the probability of finding an electron (or other particle) at various positions in space. 4. Probability Wave: A probability curve which is changing in time and space in a manner that is like the movement of a wave in space and time. 5. Quantum Mechanics: The set of laws and principles that govern wave-particle duality. 6. Uncertainty: For a quantity that is not known precisely, the uncertainty is a measure of the bounds within which the quantity is known with high probability. If you knew that your friend was on the freeway somewhere between Provo and Orem, the uncertainty in your knowledge of exactly where he/she was on the freeway might be about 10 miles since the two towns are about 10 miles apart. D. FOCUS QUESTIONS 1. A single electron is sent toward a pair of very closely spaced slits. The electron is later detected by a screen placed on the opposite side. Then, a great many electrons are sent one at a time through the same device. a. Describe the pattern produced on the screen by the single electron and later the total pattern of the many electrons. b. Name and state a fundamental principle that can account for all of these observations. c. Explain the observations in terms of the fundamental principle. 2. A single photon is sent toward a pair of closely spaced slits. The photon is later detected by a screen placed on the opposite side. Then, a great many photons are sent one at a time through the same device. a. Describe the pattern produced on the screen by the single photon, and later the total pattern of the many photons. b. Name and state in your own words the fundamental principle that can account for the observations.

151

3.

4.

c. Explain the observations in terms of the fundamental principle. A single electron is sent through a tiny slit. Later it is detected by a screen placed on the opposite side. It is possible to change the width of the slit. a. What is observed on the screen? b. Is it possible to predict exactly where the electron will be seen when it arrives at the screen? c. State the Heisenberg Uncertainty Principle. d. If the slit is made narrower in an attempt to know exactly where the electron is when it passes through the slit, what else will happen? Will we now be able to predict where the electron will be seen on the screen? Explain in terms of the Heisenberg Uncertainty Principle. How does the Newtonian Model of motion of things in the world differ from the Uncertainty Principle? To what extent is the future determined by the present according to the Newtonian Model? To what extent is the future determined according to a model that includes wave-particle duality?

mechanics? 16.9. Why are electron microscopes used for viewing atoms instead of regular light microscopes? 16.10. Why must a particle have a high speed if it is to be confined within a very small region of space? 16.11. What does the Uncertainty Principle say about simultaneous measurements of position and speed? 16.12. How does the Newtonian model differ from the Uncertainty Principle? 16.13. Explain the meaning of the Uncertainty Principle. 16.14. Why is it that the Uncertainty Principle is important in dealing with small particles such as electrons, but unimportant when dealing with ordinarysized objects such as billiard balls and automobiles? 16.15 Explain why the Uncertainty Principle does not permit objects to be completely at rest, even when at the temperature of absolute zero. 16.16. To what extent is the future determined by the present according to (a) Newtonian physics and (b) quantum physics? 16.17. Illustrate the statistical nature of physical processes by describing the motion of individual particles in the one- or two-slit experiments. 16.18. How does the Uncertainty Principle modify our view that the universe is deterministic? 16.19. Why should we not be surprised when the rules governing very small or very fast objects do not seem reasonable? 16.20. In what situations would you expect both the Newtonian laws and wave mechanics to accurately predict the motions of objects? In what situations would the two predictions be significantly different? 16.21. Which of the following would form an interference pattern? (a) electrons (b) blue light (c) radio waves (d) sound waves (e) all of the above

E. EXERCISES 16.1. Why do we think of matter as particles? Carefully describe some experimental evidence that supports this view. 16.2. Why do we think of matter as waves? Carefully describe some experimental evidence that supports this view. 16.3. Is matter wavelike or particlelike? Carefully describe some experimental evidence that supports your conclusion. 16.4. What is meant by the term wave-particle duality? Does it apply to matter, or to electromagnetic radiation, or to both? 16.5. One possible explanation of the interference effects of electrons is to presume that the wavelike behavior is due to the cooperative effect of groups of electrons acting together. What experimental evidence is there for believing the opposing view that electron waves are associated with individual particles? 16.6. Consider an experiment to test the diffraction of electrons as illustrated in Figure 16.10. Why would it be important to place a charged rod or a magnet near the beam between the diffracting hole and photographic film? 16.7. How does one describe the motion of electrons when their wave properties must be taken into account? 16.8. What is the meaning of the term quantum

152

16.22. The Uncertainty Principle (a) is an outcome of Newtonian mechanics (b) applies mainly to subatomic particles (c) conflicts with wave-particle duality (d) supports strict determinism (e) all of the above

153

154

Você também pode gostar