You are on page 1of 16

Chapter n: Atomic Structure

n.1: Introduction
Newton falls, not unlike his apple.

At the close of the 19th century, a number of “old guys” in physics were
proclaiming that “physics was over.” By that they meant that every large discovery in
physics had been made and all that was required was increasing the precision of
measurements… that is, to report the values of physical constants to a greater number
of decimal points. At the same time, younger scientists were making several very
disturbing observations that were about to topple the entire way that physics (and
chemistry) was viewed. Isaac Newton’s view of the universe (tested, accepted and
dominate for well over 200 years) was unraveling, and by 1913 would be totally
displaced. There are three classic experiments that are usually blamed for the fall of
Newtonian mechanics: the blackbody radiator, the photoelectric effect and the hydrogen
emission spectrum. While all three are important, we will focus only on the third. After
a brief mention of the first two.

The blackbody radiator and the photoelectric effect.


When an object is heated, it begins to glow, first in the long-wavelength region of
the spectrum (humans emit radio, microwave and infrared photons), then eventually
red, orange, yellow, white and so forth. Blacksmiths used the color of hot iron to gauge
heat for ages; one could work dull red iron, but could not make welds until the metal
was an intense straw yellow. In the summer of 1900, Lord Rayleigh (John W. Strutt)
presented (with a bit of help from Jean James) his derivation of the blackbody spectrum
used accepted mechanics. Although his work was brilliant application of mathematics, it
produced an absurd result, suggesting that, even a room temperature, objects would
emit strongly in the blue and ultraviolet region of the spectrum! In short, Rayleigh’s
blackbody radiator (developed using classical mechanics) produces a universe with no
darkness. In the fall of 1900, Max Plank introduced in a series of papers and empirical
equation that reproduced the expected spectrum of a blackbody radiator, then a
philosophical model. Plank’s model was revolutionary: a blackbody could only generate
a proper emission spectrum if the individual parts of the body (what we now call atoms
or molecules) could release or absorb very small but discrete amounts of energy, like
little packets which could not be subdivided. These small packets were dubbed
“quanta.” It is worth noting that Plank had a very conservative disposition, and his
model deeply disturbed him; he spent a large amount of time following the
announcement of his own equation working on a model which did not required this
quantization. He never succeeded. Gradually, the idea that matter can only gain or
lose energy is these very small amounts was accepted. Plank’s “quantum” is
unimaginably small, on the order of 10-20 to 10-30 J. For reference, a baseball (145 g)
thrown at 90 mph (40 m/s) has a kinetic energy of 3 J, roughly 20 – 30 orders of
magnitude larger. Up until Plank’s work, the quantum unit was so small, no one had
noticed it before; so everyone though that energy releases or absorptions were
continuous. Understanding the blackbody radiator problem was the first serious blow to
the classical view.
About the same time that Plank was working on the blackbody radiator problem,
the photoelectric effect was noticed. The photoelectric effect is seen when light is
impinged on a metal and, if the light has sufficient energy, an electron will be released
from that metal with some kinetic energy (given by ½mv2, where m is the mass of the
electron: 9.12×10-28 g). For an arbitrary metal, let us say that the minimum energy to
eject an electron is delivered by a photon of 500 nm (green-blue). A photon of 500 nm
will have energy of

A much more useful unit for photons and electrons is the electron-volt (eV). One eV =
1.602×10-19 J, so this 500 nm photon has an energy of 2.48 eV. If a higher-energy
photon (shorter wavelength), say 485 nm (2.56 eV) strikes the metal, an electron will
first absorb 2.48 eV of that delivered energy, then leave with a kinetic energy of 0.08 eV
of kinetics energy.
What’s the big deal, you ask? To this point, light was best understood as a
wave. The intensity of light (brightness) was interpreted as the amplitude of the wave,
while the color of the light was understood as the wavelength. If the wave interpretation
of light was correct, and the metal needed 2.48 eV of energy to leave, any color of light
should be able to deliver that energy provided it was bright enough, or if the metal was
irritated long enough. Albert Einstein postulated that light (in addition to its well-known
wave character) must also have some particle character. Light was packaged into little
“photons,” and individual photons must have sufficient energy to dislodge an electron.
Interestingly, though the photoelectric effect was empirically established by the end of
1900, and Einstein has explained it by 1905, it was not until 11 years later than R. A.
Millikan was able to get a sheet of metal clean enough for long enough to perform the
experiments that verified Einstein’s equation.
Problems:
1. The visible spectrum runs between (about) 700 nm and 400 nm. Calculate the
energy in electron-volts for a 400 and 700 nm photon.
2. Determine the energy of a mole of 700 nm photons, in kJ/mol.
3. If a photon ( = 472 nm) strikes a metal surface and ejects an electron (energy of
1.89 eV)
a. How much energy was absorbed by the metal to dislodge the electron?
b. How fast is the ejected electron moving?

n.2: The Atomic Emission Spectrum


“The proof may be in the
pudding, but electrons sure
aren’t.”

In the 1880’s J. J.
Thomson’s work with
Cathode rays (now known
to be “streams” of
electrons) lead Thomson to
develop his “plum pudding”
model of the atom (Fn.1). I
have heard, but was not
able to verify, the term
“plum pudding” was a
somewhat cruel joke made
by Millikan… working on the stereotype that all Brits love plum pudding. I really don’t
know if Thomson liked plum pudding or not, but he did not approve of the name of the
model. As it turned out,
experiments done by Marsden
and Geiger (the gold foil
experiment) and interpreted by
Rutherford demanded the
existence of a small, extremely
dense and positively charged
nucleus (Fn.2), no more
pudding.
While Rutherford’s
model accounted for the
results of the gold foil
experiment, it raised several
serious questions, the answers to which would finally overthrow Newton and establish
quantum mechanics as a better view of the microscopic world. The first question asked
was: “if the nucleus is positive and the electrons are negative, why don’t the electrons
crash into the nucleus?” The second was a bit more complex. If an electron is attracted
to the nucleus, then moving the electron closer should lower the potential energy of the
electron-nucleus system. That change in energy would be released as a photon. In the
Rutherford model, atoms should “glow” white as electrons moved toward the nucleus by
arbitrary amounts
and release energy
as photons were
emitted of different
wavelengths.
However, figure n.3
clearly shows that is
not the case. For
Hydrogen (and every other element), when excited, it’s “emission” is a series discrete
lines. Only wavelengths of specific energy are released.
Hydrogen’s emission is simple, more complex atoms have more complex
emission spectra. As far back as 1885, Balmer (a high school teacher) introduced an
empirical equation that correctly predicted the emission spectrum of hydrogen; but was
unable to develop a coherent theory based on his equation. Balmer’s problem was (not
unlike the blackbody radiator issue) that he needed to introduced a series of integer
value “steps”:

It was later observed that Balmer’s lines were repeated in the UV (Lyman’s series) and
the IR (Paschen, Brackett and others’ series).
In 1913, the Danish physicist N. Bohr proposed a revolutionary idea. Rather than
trying to develop a theory from Newton’s ideas and then push the theory to fit observed
data, Bohr worked backward (from the known data back to a model). Bohr used the
known data to formulate five postulates, two of which were later shown to be false. The
exact postulates are beyond the scope of this discussion, but you can check the Levine
(or any other) text if you are interested. In essence, Bohr’s model had electrons
“orbiting” the nuclei much like planets orbit a star. The electrons mostly obeyed the
laws of classical mechanics expect for one wild idea: the electrons were restricted to
particular distances from the nucleus. No longer could one electron move some
arbitrary distance closer to the nuclei, but only fixed discrete distances. The electrons
were now confined to a series of “stair-steps” of distance from the nuclei. This fixed-
distance movement is reflected the Balmer-Ryberg equation, with the stairs for the “n”
numbers. Bohr also applied his equation to “pseudo hydrogen” like He+ or Li2+. As long
as he accounted for the increasing nuclear charge, his predicted atomic emission
spectrum matched experimental data very well. According to Bohr, atomic emission
occurs when an “excited” electron moves from an orbit far from the nucleus to one
closer, the electron is now at a lower potential energy state. That change in energy is
released as a photon with a wavelength equal to the change in energy of the atom. Or,
E = h = hc/.
Where the E is the change in energy for the atom, and it will emit a photon of energy:
Ephoton = Eatom.
Bohr’s final equation for the energy state of a particular orbit is:

Where the following constants are: me: mass of the electron, e is the fundamental
charge (the charge in coulombs of a proton), 𝜖 is Coulomb’s constant, h is Plank’s
constant, and n is the energy level (orbit) of the electron. Atomic absorption “reverses”
the emission processes, and is illustrated in figure n.3.
Bohr’s model, though a huge step in the right direction, was totally wrong. In
fact, it was something of a happy accident that it worked at all. Applied to any atom with
more than two bodies (that is, more than the nucleus and a single electron), Bohr’s
model fails to predict the emission spectrum. To develop a realistic view of the
microscopic world, one more insight was needed, and it came for the Frenchman, L. de
Broglie.
de Broglie noted
that electrons (known to
have mass) could be
diffracted by passing
them through a very thin
slit. He proposed in 1923
that the electron—like a
photon—must have some
wave-like behavior. de
Broglie’s hypothesis was
later confirmed via
experiment, Figure n.4.
The physics community,
by 1924, had all the
pieces in place to
describe (as far as we
know) the nature of the
atom. The “old quantum theory” of Plank and Bohr was replaced by quantum
mechanics (QM): a theory the treats the electron has a wave… with mass (whatever
that means).

Problems:
4. Comment on the biggest problem with Bohr’s model.
5. The energy carried away by an electron when atomic emission occurs must, by
the law of conservation of energy, equal to Ef – Ei.
a. Use Bohr’s equation to write an expression for E.
b. What is the energy of ionization of a hydrogen atom? Hint, nf = ∞.
c. Modify the expression to solve for the wavelength of the photon emitted.
d. Compare Ryberg’s equation with Bohr’s. Calculate Ryberg’s constant
from fundamental physical constants (me, h etc.) and compare with
Ryberg’s empirically derived constant (which you need to look up).
6. Look up the ‘de Broglie equation (any decent general chemistry or physical
chemistry text). Use it to calculate the wavelength of a proton traveling at 500
m/s and a baseball (145 g) traveling at 18.2 m/s. Comment on the differences &
their physical meaning.

n.3. Quantum Mechanics (QM)


“Our best understanding of the atomic world is based on the ideas of a womanizer, a
confessing Nazi and some mathematicians.”

Actually, two forms of QM exist. The most common is “wave mechanics”


postulated by E. Schrödinger in early 1926. Wave mechanics is actually based on
earlier work by W. Hamilton in the mid 1800s that reformulated Newton’s laws to apply
to waves (very few people know this, so it pays to read the professor’s quantum
ramblings). Schrödinger lifted Hamilton’s work, and added terms to account for the
mass and spatial location of the electron. His overall approach was to reason that,
because the electron [particles] had wave-like character, one should view them as
waves, though this makes no sense to us. This wave mechanics model is the most
widely accepted and first-taught from of QM.
Just before Schrödinger presented his model, W. Heisenberg presented a
mathematical model based on using matrices to describe everything we can measure
about electrons (or other quantum particles) Unfortunately (for Heisenberg and
hundreds of physical chemistry students every year) when Heisenberg introduced his
theory, most physicists didn’t understand matrix algebra; and his work was largely
ignored. This is a major shame, because matrix algebra is super easy compared to
Schrödinger second order differential equation. However, that is history. To stay
consistent with everyone else, we will focus mostly on Schrödinger’s interpretation. In
fact, for small atoms (hydrogen through about platinum) both forms of quantum
mechanics give identical results. It should be remember, however, that all basic tenants
of modern quantum mechanics are postulates, and cannot be derived from any more
basic or fundamental models. A postulate is a “guess” that we accept as correct, not
because it makes sense, but because it allows us to correctly predict observed
behavior.

n.3.1: The Schrödinger Wave Equation (SWE):

The wave equation is often given the stupidly easy-looking form:


Ĥ = E
And it looks like we could just divide psi () out and all our troubles would be over.
Except the little “hat” on top of the Ĥ signifies an “operator.” You know about operators,
though you may not have called them that (yet). Things like the cosine (cos), logarithm
(log or ln) or even plus (+) are operators. They need something (a function or number)
to operate on or they are meaningless. Before we get into the guts of this equation, we
should take a look at it and ask ourselves, what is this equation trying to do? First, we
see that Ĥ does “something” to psi and returns a scalar (E) and psi, unchanged. Psi is
called “the wavefunction” and is postulated to be a function that contains all the
knowable information about a quantum waveparticle (perhaps an electron). The
Hamiltonian (Ĥ) operator works on psi and extracts the energy of the electron; and also
gives us the wavefunction back. Both entities on the right-hand side of the equation are
valuable. The scalar, E, tells us about the energy of the wave-particle, the wavefunction
itself () tells us about where the electron is “located.”
But what is Ĥ, I hear you ask. As I stated above, it was work of the 19th century
genius William R. Hamilton. It is a second-order differential equation, and I produce it
here because we will use it, though I will not expect you to reproduce the derivation that
will follow. Remember again that Schrödinger postulated this equation as the basis for
quantum mechanics, and there is no fundamental reason why this equation should
describe the behavior of an electron.

Don’t panic! Let’s unpack this operator. The first component, ħ (h-bar) is Plank’s
constant divided by 2, and has the value 1.055×10-34 Js. The 2 is a constant, me is the
mass of an electron, and dell squared (𝛻2) is called the gradient, or Laplacian, operator
(yes, we have an operator within an operator):

The Laplacian operator will take the second derivative of the wavefunction with respect
to each of the three Cartesian coordinates (madness!). Finally, the V-term is a potential
energy function whose identity depends on the system in question. For instance, if we
consider an electron attracted to a proton, we would used columbic potential energy
equation for electrostatic attraction.
To actually solve the wave equation is a challenge, given that it’s a second order
differential equation (in fact, it is only possible to analytically solve it for hydrogen-like
atoms), but we have the good fortune of 80 years of really smart people before us that
have already solved the equation for a number of useful (and some useless) models.
To give you some idea how this equation actually works, we are going to derive a very
simple model, called the “particle in a box” model in a single dimension (x) and the
move on to give the results only of a few more complicated models, notably the
harmonic oscillator and the hydrogen atom. Again, for clarity, I will not expect you to
reproduce this on any exam, but it is so incredibly cool, we need to take a look at it.

n.3.2: The Particle-in-a-box (PiB) Model:


“Donuts come in boxes too, but they don’t show quantum behavior.”

Electrons are “confined”


in conjugated polyenes to the
length of the alternating double
and single bonds. If we
envision 1,4-dipentene (Figure
Fn.5), we should remember
from organic chemistry that the
 electrons actually live across
Fn.5: 1,4-dipentene inside a 1D “box” of length “l.”
the whole length of the molecule.
We are going to try to figure out how long 1,4-dipentene should be, construct a linear,
1D “box” (or tube) based on the energy required to move a  electron from the ground
to an excited state. Figure Fn.6 is a “restatement” of figure Fn.5; but with the added
restrictions that; box within which a quantum particle has no potential field exerted on it,
and the particle cannot exist outside the box (by setting the potential energy of the
particle outside the box to be infinity). This is a totally foolish model, but you will see at
the end it works surprisingly well.
Based on our
boundary conditions any
particle (an electron, for
instance) will, inside the box,
not be attracted or repelled
by anything. For the
moment, please don’t try to
stick any molecule in that box,
Fn.6: 1,4-dipentene inside a 1D “box” of length “l,” with
the appropriate potential energy functions included.
despite the fact I drew a molecule in our box. However, at the edges (x = 0; = l) trying
to move the particle outside the box would require you deliver infinite energy to the
particle, a non-starter. The particle must remain in the box.
Inside the box, the SWE is easy, remember that there is no potential energy in
the box:

But, because our box is only 1D,

So, within the box:

Or upon expansion:

Or:

At this point, I suppose some mathematician could solve this second-order differential
equation: to give us some function of psi, but let’s think about what the math is telling us
here. What we need for psi is something that you can take the second derivative of,
and you get the same thing back, multiplied by a constant (which could even be 1).
There are two functions that I can think of that act this way: sine and cosine.
Remember:
But, we can’t quite be sure which would work. Mathematicians far smarter than I have
suggested we use the following equation, and we’ll prove that it works!

Where A, B, s, r are all constants, and one of them might even equal 0 (happiness!)
Taking the derivate of psi twice with respect to x yields:

We must now decide on something for r and s to equal to make this equality hold.
Remember, whatever they might be, they must be all constants. If we select that r = s =
(2meE)1/2ħ-1, then the equality will hold.
Substitution of our solution for r and s back into the original “guess” of a function
for  gives an equation for psi. However, because we cannot find the particle “outside”
our box, we need   0 as x  0 and   0 as x  l. So, in the above equation, set
both psi and x to 0. Then you can solve out:

 = A sin[(2mE)1/2ħ-1x] 0≤x≤l

We can solve this for an energy function, if we work with setting x = l, then we know that
 = 0, so:
(2mE)1/2ħ-1a = ±n
Substitution yields:
 = A sin[±nx/l]
We must rule out n = 0, thought that is mathematically “legal.” If n = 0, then  is
everywhere zero inside the box, and the particle will not exist. Also, as A is some
constant (and could therefore be positive or negative), there is no need to consider n <
0. Therefore:
 = A sin(nx/a) n = 1, 2, 3, …

And we can also find:

E = n2h2/(8ml2) n = 1, 2, 3, …

It turns out that A is useful as a “normalization constant” that can be solved to make
sure that ∫inside = 1. The solution to A is beyond the scope of this text but it turns out to
be:
A = (2/l)1/2
So the final solution to the wavefunction for a particle in a 1D box is:

For 0 ≤ x ≤ l; where n = 1, 2, 3, …
Outside the box, the potential energy part of the Hamiltonian goes to infinity, and hence
the partial would require infinite energy to exist there.
The “primary quantum number” “n” is very important here. It refers the energy
level our current system is in. When n = 1, the system is at its lowest energy, or
“ground,” state. For n = 2, the system is in its “first excited state” and so on. To move
the system from its ground to first excited state requires the input of a specific amount of
energy, any less energy input, and nothing will happen (!). Truthfully, something will
happen, because of the conservation of energy, but our model is quite crude right now.

Problems:

7. Determine the value of  and 2 for an electron in a box 6.00Å long for x = 3.00Å
for n=1 and n=2. Remember to get your calculator into radians.
8. Now, returning to our 1,4-dipenenete, we know from experiment that the [light]
absorption maxima (to move an electron from a  to a * orbital) is located at:
180 nm. We are going to use a lot of what was just derived to estimate the
length of the molecule based on the energy absorption maxima.
a. We are going to need the energy input of that wavelength of photon. Do
you remember how to do that?
b. Find E, using our result of E for the 1D PiB model.
c. Solve your expression of E for L.
d. If the molecule started at its ground state, ni = ____?
e. The first absorption band will move it to its first excited state, n f = 2.
f. Now, crank through the arithmetic.
If you did it right, you should get around 4Å. If you mentally estimate
the length you would expect our molecule to be, this is a surprisingly
good ballpark value (given all the bonds present in this molecule are
around 1 Å, and considering the bonds are not R). It is a bit on the
short side, but not too bad. I said this is “surprisingly” good because
think about the total crudity of this model.

n.3.3: What does the PiB


model tell us?
“QM says a lot, but I don’t think it
gets us any closer to the secrets of
the Old One.”
-Albert Einstein.
There are a few interesting things to note here. First, we can move electrons
from one wavefunction to another by energy input (this is, in fact, the basis of all
spectroscopy), the energy required is related to some very simple and measurable
variables, and we can calculate the wavefunction. What good the wavefunction? It
turns out that the wavefunction is the square root of the probability of finding the
electron at any given point; therefore, 2 is a measure of how likely you are to find the
waveparticle at any given point. The value of y alone does not have a good physical
interpretation, as far as I know. Figure n.7 illustrates. While a physical interpretation of
 may elude us, note that when n > 1, it make take on a positive or negative value.
While this does not really concern us for the PiB model, it will become very important
later when we discuss chemical bonding, so keep it in the back of your mind.
Now, let’s first consider the plot when n=1. Notice that the particle is most likely
going to be found at x = L/2. That is very
strange, and totally different from what we
usually see. We can approximate a PiB model
by considering a marble in a long thin tube. If
you were to shake the tube and then allow the
marble to come to rest, we expect the same
chance of finding the marble everywhere in the
tube. However, if we have a quantum
waveparticle, and it’s at its ground state, then we
will most likely find the waveparticle right in the
middle of the tube.
When n = 2 things get even stranger.
Compare the plots in figure n.7 again. No longer

is the particle most likely in the middle,


the particle is most likely found at x =
L/4 and = 3L/4 (!!) In fact, at x = L/2,
we have a region where you will not
find the waveparticle ever. Odd as that
may be, it is partly related to the fact
that our particle is a wave-particle, and
that node (where the probably drops to
zero) is the wave part of the
waveparticle is being very apparent.
Still, our particle is not behaving as
Newton said it should.
As we move to n = 14 we start to
see one final interesting thing happen.
Now we have multiple locations where the waveparticle might be found, all equally
likely. Imagine if we were at n = 1000. You begin to see, as n becomes large, we start
to have a nearly equal probability of finding the electron anywhere in the tube, akin to
what Newton told us would happen. This is an example of the correspondence
principle. The correspondence principle states that, within certain limits, the
predictions of quantum mechanics are indistinguishable from the predictions of classical
mechanics. The correspondence principle is one of the concepts that helped win early
acceptance for quantum mechanics, as its results (when applied to electron-sized stuff)
are—well—absurd.
Problems:
9. How many nodes would you expect to find when n=3? Where would you expect
the maxima to be?
10. What is the minimum n value where you could start finding  with a value < 0?
11. What is the minimum n value where you could start finding 2 with a value < 0?

n.3.4: Extending PiB model to 2 and 3D


“At least we don’t need to wear 3D glasses for this to work.”

Thankfully, extending the PiB model to multidimensional systems is actually quite


easy. This model has several applications (for instance, benzene may be approximated
as a 2D box, the hydrogen atom as a 3D box and DNA as a long 3D rectangular box).
We can follow a derivation very similar to what we just did for the 1D box, but one needs
to employ the separation of variables, where we treat any change in  based on a
change in the x-direction to have no effect on changes in the y or z coordinate. The
final equations for the 3D box are:

Where a, b, c are the lengths of our box in the x, y, z directions, respectively.

For a 2D box, simply remove the c and z containing terms, and change the
normalization constant on y to (4/ab)1/2.

Problems:
12. Attempt to sketch the contour plots of  and2 in a 2D box for:
a. nx = 1; ny = 2
b. nx = 2; ny = 2
13. Consider Benzene as fitting inside a square box of length a = b = 5.00Å.
c. Where would the maximum adsorption be for this molecule? Look up the
actual max. How do they compare?
d. Choose two things that are totally wrong with using the PiB model on
benzene, that is to say, what idea do we have about benzene, and it’s
bonding, that is totally inconsistent with our model.
14. A sample of DNA has a absorption maxima at 265 nm. If our DNA fits into a box
of a = b = 20Å. Determine the length needed.
15. Can you think of an [annoying and difficult] experimental technique that could be
done away with if we had really precise UV-visible spectroscopy, based on the
answer to the previous question?

n.3.5: The harmonic oscillator as an approximation of the Chemical Bond.


“Mathematicians often joke that the only problem physicists know how to solve is the
simple harmonic oscillator, but of course this is only true to first order.”

The Chemical Bond is well approximated as a very stiff spring. While a detailed
discussion of the quantum-mechanical harmonic oscillator is [way] beyond the scope of
this text, I do want to draw attention to this model because (a)it is the basis of the entire
field of IR spectroscopy, (b)it introduces the principle of zero-point energy, and (c)it will
introduce a phenomenon called “tunneling.” Zero-point energy is a quantum
mechanical phenomena wherein, even at absolute zero, there is still some “residual”
energy—motion—that cannot be removed from the system. The Heisenberg
uncertainty principle is partly based on zero point energy. Tunneling is another
quantum mechanical phenomena wherein waveparticle can exist where they do not
have the energy to be, for a short time. It is anaglous to a basketball sitting on the floor
suddenly being on top of a table, just because. Tunneling is of critical importance in
technology (tunneling electron microscopy) and the very high-energy processes
(nuclear fusion).
We approximate the bond as a spring, so we use Hooke’s law (Frestoring = -k,  is
the displacement from the equilibrium position). To develop an expression for the
potential energy function (V()) in the SWE. The other part of the Hamiltonian is
unchanged. The derivation is challenging, and shall be skipped (in fact, much like the
PiB model, we guess at an answer, but the calculus is a bit more challenging).
However, we finally arrive at:
E = ( + ½)h = 0, 1, 2, 3, …
Where  is the quantum number and  is the frequency of the vibration. Notice here,
unlike “n”  can have a value of zero. Yet, even when  is 0, there is nonzero energy.
The wavefunction for the harmonic oscillator is less import (to us), and it turns out to
have two forms, depending on if  is even or odd, which we will not reproduce here.
However, the value of  only exponentially approaches zero as   ∞, much like if  =
1/2 (which, in fact, is part of the equation for the wavefunction). As a result, the value
of , and therefore, 2, never reaches zero; or in other words, there is a tiny, but finite,
chance that you will find the nuclei (the masses at the ends of a vibrating spring) at very
large values of (to infinite rho, actually). One final note, for this section: we have been
developing a model in which one side of the spring is fixed. Clearly, in a real bond, both
ends of the spring have mobile nuclei attached. In the next chapter, we will allow for
both nuclei to move in space.

Problems:
16. In a TEM (tunneling electron microscopy) setup, electrons are forced to jump
across a very small distance (no more than a few Ǻ) from a charged sample to a
sharp tungsten probe tip in a near-absolute vacuum. A true vacuum cannot
conduct electricity, as there are no charged species to transmit electrons. What
property, then, allows the electrons to move from sample to tip?
17. Determine the frequency of vibration (in Hz) for a sample that has a vibrational
energy of 3210 cm-1 during the 01 excitation (fundamental) and the 02
excitation (also called the first overtone).

n.4: The Hydrogen & Hydrogen-Like Atom

To solve the SWE for an atom, we start with the simplest atom, and invoke a
rather simple model; that of columbic attraction between the (positively-charge) nucleus
and the negatively-charged electron. If we consider the nucleus to be fixed (a
reasonable assumption, given that the mass of the proton is about 10 3 times greater
than that of the electron), the only thing we need to worry about in the SWE is the
potential energy fuction, which takes the form of Columb’s law.
However, trying to solve the SWE using cartisean coordinates (x, y, z) is
impossible, so we resort to polar spherical coordinates: r (the distance from the
nucleus); , theta (the angle off the polar axis); , phi (the angle off an arbitrary 0 line in
the equatorial plane). To convert the laplacian operator into polar spherical coordinates,
we need to use:
x = r sin  cos 
y = r sin  sin 
z = r cos 
As you solve the SWE, one needs to introduce 3 numbers, now called “quantum
numbers.” In much the same way, we did this with the PiB model when we had to get
the sine term equal to zero, so we realized that:
sin w = 0 = ±n
But now, we need three numbers:
n ≡ 1, 2, 3, … |principle quantum number – defines what energy level the electron is in
(how “far away” from the nucleus it is.
l ≡ 0, 1, 2, … (n-1) |angular quantum number – defines what the area of space, within a
given “n” that the electron has a high density (what shape the orbital has).
ml ≡ (-l) …, 0, …, (l) |magnetic quantum number- defines what area of space, within a
given “l” that the electron has a high density (what direction in space an electronic
orbital is pointed).

Recall that the angular quantum numbers are related to letters:


l=0=s
l=1=p
l=2=d
l=3=f
l = 4 = g (thought there are no known elements that have “g” electrons in their ground
state, yet).

You are urged to review (check you General Chemistry, Organic Chemistry or the
internet) the “shapes” of the s, p, d and f orbitals. However, as you review, please keep
in mind that the “lobe” shapes usually associated with p, d and f orbitals are stylized
forms, and a more realistic picture of a “p” orbital (for instance) is two spheres extending
for the origin.
These orbital “shapes” are all derived from the so-called hydrogen-like atom.
That is, if an electron in hydrogen’s electron is excited from the ground state (n=1, l=0,
ml=0) to the first excited state (n=2, l=1, ml=-1) then that orbital’s shapes was
determined by solving 2 for 90% probability, and you get one of the “p” orbitals. Using
the hydrogen-like orbitals is done for two reasons. First, calculating an orbital from the
hydrogen atom is (relatively speaking) easy—in fact, it is impossible to calculate
(precisely) the orbitals for any atom other than hydrogen. Look up “the three body
problem” and you will be introduced to something mathematicians have had trouble with
since before Newton.

You might also like