Sei sulla pagina 1di 10

Quantum Mechanics

Welcome to the quantum zone. You hop in your car one morning and start heading north
toward school. For some reason, the smooth acceleration you are accustomed to during the
trip does not happen. The speedometer needle locks on 10 mi/h, then suddenly jumps to 20
mi/h, then to 30 mi/h, and so on as you depress the accelerator pedal The scenery drifts by at
10 mi/h, then suddenly at double that rate, then suddenly at triple that rate, and so forth.
Another strange thing happens. There is no way you can get out of first gear when you are
going 10 mi/h. After the jump to 20 mi/h, however, you have a choice of two gears, first and
second. After another leap to 30 mi/h, three gears are available. You notice that the needle on
your car's tachometer (engine-revolutions-per-minute gauge) jumps from one specific value
to another specific value whenever you change gears but you do not change the car's speed.
You find that the crankshaft of your engine can turn only at specific rates (or revolutions per
minute) and that no intermediate rates of engine speed are possible.
And here's one more strange thing. At 10 mi/h, your car acts as if it were affixed to trolley
tracks in the roadway; your car can move only north or south. At 20 mi/h, your car is able to
turn, but only very sharply, onto any street going east or west. From those streets, abrupt
turns onto any north-south street are possible. At 30 mi/h, your options are broadened; you
can now move in any of 8 directions: north, northeast, east, southeast, south, southwest, west,
and northwest. The wheels of your car simply refuse to spin in all other directions besides
these 8 points on the compass. At 40 mi/h, there are 16 directions to choose from.
There is more strangeness in the quantum zone. Your car itself is a kind of "fuzzy beast." It is
like a long series of waves rippling down the road. Waves are difficult to locate precisely in
space, so it is hard to say just where your car is at any moment within the train of waves. It
might be on or near the leading edge, on the trailing edge, or somewhere in the middle of the
wave train.
You are approaching a traffic light that is changing from yellow to red, and you are probably
exceeding the speed limit, too. A traffic officer on the curb, seeking to enrich the coffers of
the municipal treasury, wants to issue you two tickets: one for speeding and one for running a
red light. Alas, he finds he cannot gather definitive evidence for both of these infractions at
the same time. He has to settle for one only. If he measures the speed of the entire wave train
and that speed is in excess of the speed limit, then he has caught you speeding. However, at
the same time, he cannot know whether or not you ran a red light because he does not know
exactly where your car is within the wave train. On the other hand, if the officer is able to
precisely locate your car entering the intersection on a red light, then he really does not know
your car's speed. To determine speed (the distance between two positions divided by time),
he needs a later measurement of your car's position, and by that time, your car could be
located somewhere else (ahead of or behind the place where it was before) within the wave
train.
These imaginary quantum zone experiences, which are based loosely on the behavior of
electrons, may give you some insight into several aspects of quantum mechanical theory.
Together, they are a far from complete or even accurate analogy of what really goes on inside
the atom-but then again, quantum mechanics by its very nature is abstract. It always resists
being compared to ordinary experience.
Quantum mechanics is a mathematical description of the behavior of particles far too small to
see. Time and time again, it has proved its worth through the success of its predictions. As
1

such, scientists accept its validity. We need not fully understand something (in an
experiential sense) to make use of it. In fact, well over half of the world's entire economy is
linked to inventions that have stemmed from the applied principles of quantum mechanics.
Let us now investigate, in a more formal manner, the aspects of quantum theory that pertain
to tiny bits of matter.
Light has a wave-particle duality: its behavior incorporates wavelike features and particle
like features. Might not pieces of matter, which we have traditionally thought of as particles,
act like waves? If so, then a certain kind of symmetry would encompass both matter and pure
energy (for example, light). In other words, both would exhibit wave-particle duality.
Figure 1

Figure 1 (a) When waves diffract through two slits and interfere, a pattern of many interference
fringes appears on the screen. The waves must be of relatively large wavelength compared with the
slit width in order to diffract significantly. (b) Particles, assuming that they have a wavelength too
small to diffract significantly, have trajectories that carry them to only two spots on the screen.

Matter Waves
These thoughts and some of their implications were first expressed by the French physicist
Louis de Broglie in his doctoral dissertation, completed in 1923. Mindful of Einstein's
conclusions that mass and light energy are equivalent (E = mc2) and that there is an intimate
relationship between space and time, de Broglie deduced that particles of matter could indeed
be thought of as waves of matter, or matter waves. De Broglie predicted that the matter
waves associated with a moving particle of matter should have a wavelength of
A = h / p = h / mv

Equation 1

where h is Planck's constant (6.63 X 10-34 J*s), and p is the particle's momentum (which, as
we learned in the notes on Laws of Motion, is the particle's mass multiplied by its velocity,
or mv).
Equation 1 says that a particle of matter somehow propagates as moving waves with a
wavelength that depends on both the mass and the speed of the moving particle.
= 10-35 m
The meaning of matter waves is difficult to interpret, but the matter (or de Broglie)
wavelength A of a particle can be somewhat concretely described as being related to the
size of the region of influence of the particle.

Planck's Constant and the Energy of a Photon


In 1900, Max Planck was working on the problem of how the radiation an object
emits is related to its temperature. He came up with a formula that agreed very
closely with experimental data, but the formula only made sense if he assumed
that the energy of a vibrating molecule was quantized - that is, it could only take
on certain values. The energy would have to be proportional to the frequency of
vibration, and it seemed to come in little "chunks" of the frequency multiplied by
a certain constant. This constant came to be known as Planck's constant, or h,
and it has the value 6.63 x 10-34 J*s.
J stands for Joule, in honor of James Joule who examined the relationship
between heat and energy in the 19th Century. The Joule is the SI unit for
measuring energy. A joule is related to the fundamental units in that a joule is 1
kg m2/s2. By no means is that obvious, but it is similar to measuring area in m2
(distance times distance) or speed in m/s (distance divided by time).
So J*s is a Joule second. What would that be? That's not quite energy, not quite
time.
The fancy name for what a Joule represents is angular momentum, but all that
you need to know now is that it involves things that spin, like bicycle wheels and
tops.
Matter Waves and the Bohr Atom
It was not clear in the beginning just what matter waves might be, but de Broglie went ahead
and applied the matter-as-moving-waves idea to the electrons of Bohr's model of the
hydrogen atom. Figure 2.
Figure 2

The speed of the electron in each of the Bohr orbits is known. For each orbit the speed is just
sufficient to keep the electron orbiting around the proton in the nucleus to which it is
attracted. (Similarly, a planet must move around the sun at a particular speed if it is to remain
in the same orbit.) The radii of the various Bohr orbits were known as well; they were related
to the potential energies of the electron at different distances from the nucleus. Since these
orbits are all circles, the distance around the rim of each orbit (circumference) is just 27Tr.
Knowing the mass of the electron and its speed in each of the Bohr orbits, de Broglie used
Equation 1 to calculate the electron's wavelength in each case. The amazing result,
schematically illustrated in Figure 3, is that the electron's matter waves fit in a standing wave
pattern around the circumference of each orbit. In other words, the electron, as a traveling
wave, reinforces itself (experiences constructive interference) every time it completes an
orbit. One complete wave fits into the first Bohr orbit (n = I, the ground state). Two complete
waves fit into the second Bohr orbit (n = 2, the "first excited state). Three complete waves
fit into the n = 3 orbit, and so on. By this way of thinking, the electron cannot exist in any
orbit other than those described by n = 1, 2, 3, and so on, because those orbits would involve
a mismatch of wave "crests and "troughs as the wave repeats its path around the orbit.
These situations of destructive interference make it impossible for an electron to persist in
intermediate orbits.
The electron "waves sketched in Figure 2 resemble the fundamental and harmonic
frequencies of a vibrating string (see Figure 8.21), except that they are wrapped around
circles. It seems that in the hydrogen atom, the electron can exist in only one mode of
vibration at a time; and it may jump from one discrete mode (or orbit) to another, absorbing
or giving up energy as it does.
Figure 3

The perfect fit of hypothetical matter waves around the circumference of electron orbits was
either an incredible coincidence or a true description of what really happens. Soon, a host of
experiments were demonstrating that matter waves are indeed real. When a beam of X-rays is
sent through a thin metal foil, many are diffracted (bent) when passing through the crystalline
lattice of metal atoms. The X-rays emerge in a geometrical pattern somewhat similar to that
produced by light when it reflects from a diffraction grating or the pitted surface of a
4

compact disc. These X-rays are simply exhibiting normal wave behavior. When electrons,
having a matter wavelength equal to that of the X-rays, are used in the same experiments, the
diffraction effect is exactly the same (Figure 4). Somehow, the electrons act just like waves.
The wave nature of speeding electrons has been put to good use in electron microscopes
which use beams of electrons instead of light waves. The faster the electrons move in a beam,
the shorter their wavelength is and the better they are able to produce high-resolution images
of tiny objects.
Figure 4

The Uncertainty Principle


If, on the quantum level, matter is significantly smeared out into waves, then there must be a
substantial amount of imprecision involved in measurements made at very small scales.
Werner Heisenberg, one of the many contributors to the theory of quantum mechanics,
proposed in 1927 that it is impossible to know both the exact position and the exact
momentum of a particle at the same time. Heisenberg's mathematical statement of this
postulate, which is called the uncertainty principle, can be written as
(p)( X) = h

Equation 2

where p is the momentum of a particle, x represents its position, and h is Planck's constant.
The quantity p represents the uncertainty in the measurement of momentum, and x
represents the uncertainty in the measurement of position. This state of affairs was alluded to
in the imaginary tale that opened this chapter: The traffic officer could not issue two tickets
for two infractions at the same time-one based on speeding (which involves knowing the car's
momentum or speed precisely) and the other based on running a red light (which involves
measuring the car's position precisely).
Two things are important to keep in mind about the uncertainty principle. One is that a
relatively good knowledge of a particle's momentum (or essentially, velocity, as long as the
particle's mass is constant) implies a relatively poor knowledge of the particle's position at
the same time. The opposite is true as well. A relatively good knowledge of a particle's
position implies a relatively poor knowledge of the particle's momentum at the same time.
The second important thing to realize is that the two uncertainties, multiplied together, are
approximately equal to h, which has an incredibly small value. Thus, any trade-offs between
p and x are significant only for particles with tiny momenta and tiny sizes. These are the
same particles (particularly atoms and subatomic particles) for which all the other quantum
effects are significant. Because the value of h is very small, quantum uncertainty has almost
no effect on macroscopic bodies. Only if h (which, like c, the speed of light, is a fundamental
constant of nature) were very much larger than it is would it be impossible to simultaneously
get good measurements of both the speed and the position of a car nearing an intersection.
5

To get some feeling for how the uncertainty principle works for small particles, imagine
trying to simultaneously locate and measure the momentum of some very small particle. To
"see" this particle, you must send at least one other "particle" (a photon of light or an
electron) that will bounce off the particle you are trying to observe and then come back to
you. Photons carry momentum, and certainly moving electrons do, too. So when the probing
photon or electron returns to you, it has already interacted with and disturbed the particle you
are trying to observe in a way that cannot be known unless you succeed in locating the
particle a second time. The very act of observing the particle disturbed it in an unpredictable
way. The returning electron or photon told you the particle's position at the moment of
encounter, but the particle skittered away at an unknown speed and direction after the
encounter, and that made it impossible for you to determine the particle's momentum.
Some interesting analogies have been drawn between the study of social behavior and the
uncertainty principle. For example, suppose that an anthropologist parachutes into a jungle
village whose inhabitants have had no previous contact with the modern world. By the very
act of the anthropologist's godlike arrival and her interaction with the villagers, the culture
might very well change in an unpredictable way. It is therefore quite likely that the
anthropologist will never be able to formulate an objective and accurate description of all
aspects of that culture as it was before she arrived.
Despite such comparisons, it is important to realize that only at the quantum level does the
act of making a physical measurement significantly affect the object being measured. Even
with this restriction in mind, however the philosophical implications of the uncertainty
principle are enough to destroy the notion of the universe running like a perfect machine. If it
is in principle, impossible to measure all the properties of a particle or system of particles at
anyone instant of time, then it is impossible to make confident predictions concerning all
later instants of time.

Quantum Numbers
In the older (pre-quantum mechanics or classical physics) way of picturing the atom, an
electron moving about the nucleus is described as having energy, momentum, and a certain
orbit (path of travel). In the newer quantum mechanical way of picturing the atom, the
uncertainty principle comes into play, and our knowledge of an electron's position and movement lacks the crispness and apparent certainty with which we view the larger world. In the
quantum world, an atomic electron still has energy and momentum, but these quantities can
change only in discrete steps; they are quantized just as everything else is in the realm of the
very small.
We may be uncertain about exactly where an atomic electron is and where it is headed at any
instant of time, but we can certainly identify trends, or probabilities associated with its
position. These probabilities were given a firm mathematical description in the work of
Erwin Schrodinger, another giant in the development of quantum mechanics. Schrodinger's
wave equation (which because of its mathematical complexity will remain unwritten here)
describes particles, such as electrons, as waves occupying all three dimensions of space and
(if need be) changing with time.
If an electron in an atom remains in a stationary state, then its wavelike nature does not
change with time, and Schrodinger's wave equation can be used to describe the probability of
6

its position in three-dimensional space. For an electron in the ground state (n = 1) of a


hydrogen atom, the electron ends up "looking" as if it occupied a fuzzy sphere around the
nucleus (see Figure 5). In other words, rather than following the simplistic circular path of a
Bohr model orbit, the electron follows some unknown random path at any instant. Over time,
however, the pattern of probability of the electron's position (called the electron's probability
cloud) spreads out radially from the nucleus so that the negative charge carried by the
electron occupies a fuzzy sphere. If you choose to measure the exact position of the electron
at some instant, it will most likely be found at or near the same distance from the nucleus as
the radius of the n = 1 Bohr orbit. (Of course, then you don't know its momentum very well.
You are uncertain of its speed and direction.) There is also some chance that your
instantaneous positional measurement might locate the electron inside or outside the most
probable radius.
The number n, which increases in integer steps in a hydrogen atom (as well as in all other
atoms), denotes different quantized energy states in which an electron may be found when it
is bound to an atom. It can be thought of as a code number symbolizing those various energy
states. It is also used as a multiplier in formulas that determine just how much overall
potential energy the electron has in each state and how far the electron is : on average) from
the nucleus in each state.
Figure 5

It turns out that electrons, when bound in some way to atoms, possess properties apart from
overall potential energy, which is symbolized by n. These other properties, which involve the
electron's angular momentum, are only three in number. (Recall that angular momentum is a
quantity associated with bodies that revolve or spin. An atomic electron can be thought of as
capable of doing both.) Each of these other three properties s also quantized. Thus, a
complete description of an atomic electron's state of being (its "quantum state") can be
symbolized by a total of four code numbers, called quantum numbers. We will now briefly
describe the meaning of each of these quantum numbers.
The Four Quantum Numbers
Principal quantum number, (n) describes the electron's average distance from the nucleus as
well as its overall potential energy. Within a given quantum state n, however, the exact
potential energy of the electron may vary somewhat owing to different possible values of its
angular momentum. As we have seen, n takes on integer values: n = 1, 2, 3, The overall
effect of n is on the size of the electron probability cloud; the larger n is, the farther out from
the nucleus an electron is most likely to be found.
7

The quantum number 1 (el in italics), which is called the orbital angular momentum quantum
number, describes the magnitude of the electron's angular momentum as it moves about the
nucleus. It, too, is quantized. (Macroscopically, this would be like a bicycle wheel that could
spin only at certain speeds with no intermediate speeds possible.)
The possible angular momentum magnitudes are limited in a way that depends on the
electron's n state; this is codified by the rule l = 0, 1, 2, ..., n - 1. In other words, if an electron
is in the n = 1 state, then 1 = 0 is the only possible angular momentum the electron can havewhich is to say that its angular momentum is zero. That particular state is represented by the
electron occupying a spherical fuzzy cloud, as in Figure 3. An 1 = 0 electron has not stopped
moving; rather, it moves around the nucleus in all possible directions, so that its angular
momentum sums to zero. In the n = 2 state, however, an electron may assume either of two
possible angular momentum magnitude states, 1 = 0 (zero angular momentum) and 1 = 1 (a
nonzero angular momentum). The overall effect of l is on the shape of the electron
probability cloud. An electron with 1 = 0 forms a spherical cloud, an electron with 1 = 1
forms a dumbbell shaped cloud and electrons with higher 1 values form probability clouds of
greater complexity.
Figure 6
Probability clouds

Figure 7
Angular Momentum

Angular momentum is a vector that has direction as well as magnitude. The quantum number
ml, the orbital magnetic quantum number determines the quantized direction of the electron's
angular momentum as it moves about the nucleus. (Macroscopically, this would be like the
axis of a spinning bicycle wheel that could be oriented only in specific directions.) The term

magnetic is used to describe ml because an electron with an angular momentum, like a bunch
of electrons circulating in a coil, acts just like a tiny bar magnet.
The fourth and final quantum number ms the spin magnetic quantum number is a description
of an atomic electron's spin. Regardless of what other quantum numbers a given electron has,
it either spins in one particular direction or in the opposite direction. These spin states are
denoted ms = +1/2 and ms = - 1/2. There is some question as to whether or not electrons
actually spin, because we cannot see them spinning. But that is really beside the point,
because even nonmoving electrons act like very weak bar magnets that either point "up" or
"down." This is exactly how they would behave if they were spheres of negative charge
spinning either one way or the other.
The Exclusion Principle
The electron in a hydrogen atom has a strong tendency to fall into its lowest energy state.
So far, we have limited our discussion of atomic electrons to the single electron in a
hydrogen atom. What about all the other atoms that make up the chemical elements heavier
than hydrogen. They have more than one electron. (Uranium, in fact, has 92!) Where do
these many electrons fit? Do they all crowd into the same n = 1 probability cloud associated
with the hydrogen atom's ground state? Or do they somehow "stack up" in some sort of
organized pattern? The chemical behavior of the various elements, which is widely divergent
in most instances, suggests that electrons arrange themselves differently in the atoms
belonging to different elements.
In 1925 Wolfgang Pauli found the key to electron arrangements in multielectron atoms. His
exclusion principle says that no two electrons in the same atom can have the same set of four
quantum numbers. This means that there is a unique set of attributes for every electron in a
given atom; no two electrons in a given atom can behave in exactly the same way. So how do
the electrons arrange themselves? Just remember that each of the electrons in an undisturbed
(unexcited) atom strives always toward a state of lower energy.

Summary
Light and matter exhibit both wavelike properties and particle like properties. This behavior
is called wave-particle duality.
A particle of matter has a wavelength that increases with decreasing momentum (mass and
speed). Because the wavelength of a particle also depends on Planck's constant h, which has
an extremely small value, only particles with very small momenta exhibit measurable
wavelike properties. In the atomic realm, the wavelike properties of particles are significant.
The wavelike properties of tiny particles can be tested by subjecting them to the same
experiments that have been used to confirm the wave nature of light. Particles and photons
behave similarly when their wavelengths are the same.
It is impossible to know both the exact position and the exact momentum of a particle at the
same time. The relative uncertainty involved is large only for small particles.
Because uncertainty is so prevalent in the realm of the atom, atomic electrons are best
described as behaving like standing waves positioned around the nucleus. Each electron in an
9

atom can be completely described by a set of four quantum numbers. Furthermore, each
electron in an atom differs from all the others in the sense that no two electrons in the same
atom can have exactly the same set of four quantum numbers.

10

Potrebbero piacerti anche