Вы находитесь на странице: 1из 24

VII.

Nuclear Chemistry (Chapter 18)


A. Modes of radioactive decay, nuclear reactions, fission, fusion
The standard notation for atoms that includes atomic number and mass is:

A
Z

12
6

where
E = symbol of element
Z = atomic number (number of protons)
A = mass number (similar to the number of protons plus number of neutrons)
The masses are:
neutron 1.67493 10-24 g
proton 1.67263 10-24 g
electron 9.1094 10-28 g
In atomic mass units (amu or just u) the 126 C atom mass is 12 u. Thus a mole of 126 C is defined as
having a mass of 12 g. Given Avogadros number, we have 1u = 1.66054 10-24 g.
In amu units,
neutron = 1.0086649 u
proton = 1.007276 u
electron = 0.000549 u
hydrogen atom = 1.0078205 u
Note: 1.007276+0.000549=1.007825
However what we really should have is: (since E=mc2)
mpc2 + mec2= mHc2 + binding energy
The binding energy of the electron to the proton, 13.6 eV, has an even smaller effect, amounting to only
1.4 10-8 u.
Normally we will include electrons in defining atom masses.
Isotopes: Although the atomic mass of H is 1.00794 overall, this arises from a mixture of the isotopes:
Hydrogen

1
1

H 1.0078

deuterium

2
1

H 2.014

tritium

3
1

H 3.016

12 .011
6

C is similarly a statistical mixture of carbon isotopes, mostly carbon 12.

11
6

C = 11.011433 u

12
6

C = 12 exactly

13
6

C = 13.003354 u

14
6

C = 14.003242 u

These masses are not exactly integer multiples due to nuclear interactions between the protons and
neutrons (which are collectively called nucleons). The mass variation thus reflects binding energy
differences via the famous formula E = mc2.

43

Nuclear Fusion!
If you combine 2 protons and two neutrons to give 42 He, there is a mass defect as follows:

m m 42 He 2 m 11 H 2 m 01 n

= 4.0026033 - 2(1.0078205) -2(1.0086649)


=-0.0303766 u
From this, the binding energy of

4
2

He

is:

E = m c2 = (-0.0303766)(1.660540 10-27 kg)(2.9979246 108 m/s)2


= -4.53346 10-12 J per He atom
= -2.73 1012 J/mol
= -2.73 109 kJ/mol
This is a much larger number (by 7 orders of magnitude) than is associated with typical chemical bond
energies (4 x 102 kJ/mol). This is the principle of nuclear fusion reactions!!
Radioactive decay
Many nuclei have neutron/proton combinations that make them unstable to nuclear decay, i.e., m >0 for
a nuclear transformation. The dominant modes of decay are:
1) particles: emission of a ??? nucleus. This is only important for Z > 60 and is most important for Z >
83. This leads to reduction in Z by 2 and in A by 4. For example, the most important decay pathway
starting from
238
92 U

is

238
92

U 23490Th 42

emission doesnt penetrate very far, can be stopped by paper or skin, but its ionizing capability is
substantial.
2) particles: emission of an ???????. This is common in lighter elements, but it also plays a role in
heavy elements. It raises Z by one with no change in A. An example is:
24
11

24
Na 12
Mg 01 e

emission can penetrate 3 mm of Al.


Neutrinos: emission is always accompanied by emission of an antineutrino (massless and
uncharged), and in fact this is how neutrinos were discovered. Neutrino production was invoked because
the energy of the electron is less than (or equal to) what would be expected by the m calculation. (The
observation of neutrinos from the sun was the subject of the 2002 Nobel Prize in physics.)
Charge Neutralization: emission initially produces a positive ion of the Z+1 nucleus, but neutralization
will (usually) quickly occur by electron transfer from nearby atoms. To calculate m, one needs to
determine A for the positive ion, rather than the neutral (minor change).

44

3) emission: emission of a ???? of very short wavelength/high energy. This often happens during or
decay, but it can also happen by itself. There is no change in Z or A. An important example is:
60
27

60
Co* 27
Co

emission is commonly used for cancer treatment, as the rays can ionize cells, killing them. rays can
penetrate lead and concrete.
Other modes of decay:
4) electron capture: electron in orbital near nucleus is captured by nucleus, decreasing Z by 1, with no
change in A.
5) positron emission: emission of an antielectron, decreasing Z by 1, with no change in A
6) proton emission: decreases Z and A by 1.
7) neutron emission: decrease A but not Z by 1.
8) fission: only important for U or higher

Nuclear stability
The protons and neutrons in the nucleus are held together by the strong force. The precise
details of this are not understood, but many qualitative features are known:
1. Stability requires A/Z to be roughly 2 for Z<20 and greater than 2 for larger Z.
2. Nuclear stability is enhanced for even numbers of protons and neutrons
3. There are magic numbers of nucleons: 2, 8, 20, 50, 82, 114, 126, 184 which correspond to shell
closings in the nuclear structure.
4. All nuclei with Z > 83 are unstable, and therefore decay (typically to Pb) but the lifetimes can be
billions of years. Highest stable element is Bi-209, highest naturally occurring is U-238, highest element
on chart is Uub (Z=112, A=277). Web site of the chart:
http://www.nndc.bnl.gov/chart/reCenter.jsp?z=118&n=176
Nuclear reactions can also be induced artificially by bombarding the atom with particles that have
the right charge and mass combination (or close to) for the desired transformation. For example, Tc is
produced by bombarding Mo with deuterium nuclei:
97
42

Mo 21H 9743Tc 2 01 n

45

Nuclear decay laws


In radioactive decay, the rate of decay is proportional to the number of nuclei N:
Rate = dN/dt = -kN
where k is a rate constant for decay, which has units of number of disintegrations per second.
This expression can be rearranged to give
N(t) = N0 e-kt
This shows exponential decay of number of nuclei N with time.

Halflife
The halflife is obtained from
N0 = N0 exp(-kt)
from which one finds:

t1/ 2

ln 2 0.693

k
k

Example: The Chernobyl nuclear plant accident is estimated to


have released 10 million Curies of Iodine-131 into the
atmosphere. This decays via:
131
53

I 131
54 Xe

The specific activity of I-131 is 1.2 105 Curies/g. Its half life is
8.05 days. How many grams of I-131 would exist after 38 days?
k = 0.693/8.05 = 0.08619 day-1
m0 = 10 106 / 1.2 105 = 83.33 g
m = m0 e-kt = 83.33 e-(0.08619) (38)=3.2 g
Uses of Radioisotopes
14

1. Carbon-14 dating: based on the fact that a living organism exchanges 6 C with its environment, but
this stops when it dies. So by measuring Carbon-14 activity, one can determine how long ago that the
organism died. This has a half life of 5700 years. Tritium dating works the same way, but its half life is
only 12.3 y, so the time scale that can be accessed is much shorter.
2. Dating of rocks is done with Uranium-238
3. Medical uses include 99Tc for bone scans ( emitter), 18F for PET (image tumors), 60Co for tumor
irradiation, and many others.
4. Americium-243 is used in smoke detectors to ionize smoke particles.

46

5. Powering spacecraft for long missions (plutonium)


Units of radiation
1. Curie (commonly used) = 3.7x1010 nuclear disintegrations/s.
The SI unit is the Becquerel = 1 disintegration/s
2. Dose: rad (commonly used) = exposure to radiation which deposits 10-2 J of energy per kilogram of
tissue. The SI unit is the gray (Gy) = 100 rad.
Note: if h = 106 eV = 1.6 x 10-13 J, 1 Curie implies 0.006 J/s = 0.6 rad/s.
Dose/Dose Equivalent
Unfortunately Curies or rads arent very informative, as the ionizing ability of radiation depends on a lot
more than just the amount of energy released. Because of this, there is something called the dose
equivalent which is Q times the dose, where Q is the biological effectiveness. Q is 1 for and
radiation but is 20 for .
Units of dose equivalent are:
roentgen (same unit as rem (roentgen equivalent man))
sievert (Sv, same as gray).
Intensity
Curie
Becquerel

Dose
rad
gray

Dose equiv.
roentgen (rem)
sievert

How much ionizing radiation is dangerous?


0.3-0.6 mSv/yr is a typical range of dose rates from artificial sources of radiation, mostly medical. (1mSv
= 0.001 J/kg)
3 mSv/yr (approx) is the normal background radiation from natural sources in North America, including
an average of almost 2 mSv/yr from radon in air.
50 mSv/yr is, conservatively, the lowest dose rate where there is any evidence of cancer being caused. It
is also the dose rate which arises from natural background levels in several places. Above this, the
probability of cancer occurrence (rather than the severity) increases with dose.
As a dose accumulated over some time, 1000 mSv would probably cause a fatal cancer many years later
in 5 of every 100 persons exposed to it (i.e. if the normal incidence of fatal cancer were 25%, this dose
would increase it to 30%).
1,000 mSv (1 sievert) in a short term dose would probably cause (temporary) illness such as nausea and
decreased white blood cell count, but not death. Above this, severity of illness increases with dose.
Between 2 and 10 sieverts in a short-term dose would cause severe radiation sickness with increasing
likelihood that this would be fatal.
Gases (Chapter 5)

47

Having spent a lot of time talking about the molecule-specific chemical properties, let us now turn to
physical properties that are often not very much dependent of chemical identity.
Much of what we think about gases is based on measurements of atmospheric pressure at 25C and sea
level. This pressure is:
1atm = 1.01 105 Pa, where 1 Pa =1 Pascal = 1N/m2 =1 kg m-1s-2.
Pressure is the force per unit area exerted on the surface of a container which holds the gas. Pressure is
also measured in atmospheres (1 atm), in mm of Hg (760 Torr), and in pounds per square inch (14.7 psi).
For all gases at a constant temperature, pressure varies inversely with volume (to very high precision).
This is ????s law (1662):
Pressure ~ 1/Volume
For a gas at constant pressure, the volume of the gas is proportional to the absolute temperature. This is
??????s Law.
Volume ~ temperature
Definition of absolute temperature
Note that Charles law actually defines what we mean by absolute temperature.
For other temperature scales (like Celsius), the volume varies linearly with temperature, but with an
intercept. This intercept, which corresponds to V=0 occurs at -273.15C. Since negative Vs are
unphysical, this implies that -273.15C is the lowest possible temperature, i.e., absolute zero.
A related law to Charless law applies to the case of constant volume. Here one finds that
Pressure ~ temperature
The physical interpretation of this is that as temperature increases, the kinetic energy of the gas increases,
increasing the velocity and therefore increasing the force or pressure on the surface.
A third principle of importance is ???????s principle:
A given number of gas molecules occupy the same volume, regardless of their chemical identity.
This means that the volume per mole is a constant, independent of chemical identity. This volume turns
out to be 22.41 liters at 273K (0C) and 1 atm pressure. This combination of temperature and pressure is
called standard temperature and pressure (STP) so 22.41 is a good number to remember.
Ideal gas law: The combination of the formulas given above is the ideal gas law:
pV = nRT
where R is a constant called the gas constant. This equals 8.31451 J/K mol = 0.0820578 L atm/K mol.

48

The pV=nRT formula is called an equation of state, for it tells us how to connect all the thermodynamic
variables that are needed to describe an ideal gas.
There are many ways to use the ideal gas law:
1) At constant temperature: p1V1=p2V2
2) At constant volume: p1/T1 = p2/T2
3) For any transformation of a given amount of gas:
p1V1/T1 = p2V2/T2.
Mixtures of gases: Another important principle is ??????s law:
The total pressure of a mixture of ideal gases is the sum of the partial pressures of its components.
Thus if we combine in a tank a species A having nA moles with species B having nB moles:
pA = nART/V
pB = nBRT/V
pA+pB = (nA + nB) RT/V = nRT/V = p
Another useful way to express partial pressures:
mole fraction: xA = pA/p.
From the above results it is easy to see that xA + xB = 1.
As an example of this, in air we have 0.78 atm of N2, 0.22 atm of O2, so the mole fractions of N2 and O2
(neglecting the Ar) are 0.78 and 0.22, respectively.
Diffusion and Effusion
Diffusion and effusion are two processes where the velocity of the atoms in a gas can be determined.
??fusion = gradual dispersal of one substance through another
??ffusion = escape of a gas through a pinhole into vacuum
Grahams law of effusion: the rate of effusion is inversely proportional to the square root of the molar
mass of the substance.

49

If we assume that the molecules in a gas only have kinetic energy, and that all gas molecules have the
same average kinetic energy, then KE = Mc2, where M is the molar mass and c is the velocity. This
means that

2KE
M

This explains Grahams law, assuming that KE doesnt depend on mass. Also, if we assume that KE ~ T,
then c ~ T which is another observed result. But why is KE independent of M, and why is KE ~ T? To
explain this we need to learn about the Kinetic Theory of Gases.
Kinetic Theory (Model) of Gases
The text gives a detailed derivation of this on pages 170-174. This theory assumes:
1. The gas consists of a collection of molecules that move independently
2. The molecules are treated as point particles (no volume)
3. The molecules move in straight lines until they collide
4. The molecules do not interact, except for collisions
Here is a brief version of the derivation. This is not required for the class, although the final
formulas are.
We will use Newtons second law to calculate the pressure that the molecules exert on the surface of the
container that encloses them. This law states that
Force = mass acceleration.
Since acceleration = time rate of change of velocity, and
mass velocity = momentum, we can also write Newtons law as:
Force = time rate of change of momentum
Now we want to calculate the momentum change of molecules that hit each wall. Lets consider motions
in the x direction only, and assume that each molecule strikes the wall and bounces back elastically. If all
the molecules have the same velocity vx, then the momentum change is 2mvx.
We consider all the collisions with the wall in a time interval t. In this interval, any molecule with x
component of velocity vx, and that is initially within a distance vxt of the wall will collide. If the wall
has an area A, then the total volume containing molecules that will collide is Avxt.
The fraction of all the molecules in the volume V that will collide is:

fraction

Av x t
2V

Here the 2 arises because on average only half the molecules will be moving towards the wall; the other
half will be moving away from it.
If there are N molecules in the container, the number that will strike the wall in t will be:

number

NAv x t
2V

and the total momentum change will be that number times 2mvx.

momentum change (2mv x )

NAv x t
2V 50

The rate of change of the momentum will therefore be:

NAv x t NmAv 2x
momentum change
(2mv x )

t
2Vt
V
This will equal the force on the wall, which is the pressure times the area A. Thus:

F NmAv 2x Nmv 2x

A
VA
V

Of course not all the molecules will have the same velocity vx. Instead we expect to have a distribution of
velocities, and the pressure will be determined by an average over the distribution, i.e.,

Nm v 2x
V

In addition, we need to generalize the derivation from one dimension to three. This can be done by
realizing
<vx2> = <vy2> = <vz2>.
Also,

<vx2> + <vy2> + <vz2> = c2=3<vx2>

where c is the root mean square speed (the square root of the average of the square of the speed of each
molecule). Thus we can write:

Nmc 2 nMc 2

3V
3V

where we have used Nm = nM.


Note that pV

1
nMc 2 = quantity independent of p or V.
3

If we now invoke the ideal gas law, pV = nRT, we find

1
pV nRT nMc 2
3
or

3RT
M

In addition, the energy per mole associated with translational motions of the molecules is:

1
3
Mc 2 RT
2
2

This is an example of the ?????????? theorem (a fundamental theorem of statistical mechanics) which
says that the average energy per mole of molecules is RT per degree of freedom. For a monatomic
ideal gas (rare gas) with three degrees of freedom, this implies that the average KE = 3/2 RT

Consequences of Kinetic theory:


1. Kinetic theory provides a derivation of the ??????? theorem

51

2. Kinetic theory leads to Grahams Law:


This shows that H2 moves at 1930 m/s (4300 mph), which is faster than a speeding bullet, though not fast
enough to escape from the atmosphere. In the atmosphere, there are 1010 collisions per second, so the
poor H2 only moves 2 10-7m = 200 nm before colliding with another molecule and changing directions.
3. The average velocity of N2, O2 and other gases that are common in the atmosphere is 500 m/s (1100
mph) which is much smaller than that for H2. Still, these velocities are larger than the speed of sound
(around 350 m/s at sea level).
4. Kinetic theory tells us that the velocity c that we have calculated is the rms (root mean square) velocity,
i.e., <c2>.
The mean velocity c is:

8RT
M

so it is only slightly smaller than c.

???????????? distribution
An extension of kinetic theory also gives us the distribution of velocities of molecules in a gas. This is
the ??????? distribution:

M
f (v) 4

2RT

3/ 2

v 2 e Mv

/ 2RT

where v2 = vx2+vy2+vz2. Note: <v2>=c2, <v>=

.c

Here f(v) is the probability of finding a speed v between v and v+dv, with the total probability of all
possible speeds normalized to unity.
Note that:
f(v)~ v2 for small v, and
f(v) ~exp(-Mv2/2RT) for large v .
The distribution peaks at v=vmp

2RT
, at a velocity that is close to c
M

52

Real gases: The ideal gas law predicts that the quantity
Z = pV/nRT
(known as the compression factor) should be 1 independent of pressure or temperature. However for real
gases it is found that Z depends on both p and T.
One way to think about this result is that at low pressure, the molecules are far apart, and ideal gas
behavior is observed
(Z 1).

At intermediate pressure, the interactions between molecules are attractive at long range, so the
molecules attract, and Z is smaller than the ideal gas value.
At high p, the molecules repel (due to the short range repulsion) and Z is greater than 1.

A consequence of Z<1 is that it is possible to cool gases by expanding them. This is the process used to
produce liquid N2, O2 etc that was discussed earlier.
Van der Waals equation: There have been many attempts to improve the ideal gas equation so as to be
able to describe real gas behavior. The most well known of these is the van der Waals equation:

(p

an 2
)(V nb) nRT
V2

What this does is to replace the pV term in the ideal gas equation a corrected pressure and volume that
take into account the effects on interactions between the gas molecules.

(p

an 2
)(V nb) nRT
V2

(1) The volume term is the easiest to understand: the nb term is designed to subtract ???????????? of the
gas from V so that what remains is the effective volume of empty space occupied by the gas.
(2) The pressure term adds in a correction for the reduction in pressure that occurs because of attractive
interactions between the gas molecules. Note that the dependence on n/V = density tells us that pairwise
interactions are dominant.
The parameters a and b are adjusted to fit the observed behavior of the gas as well as one can. Specific
values of a and b are in Table 5.5.
For water, b = 30.49 cm3/mol (compare to 18 cm3/mol)
a (n/V)2= 5.536 atm L2/(22.4L)2 =0.011 atm
(compare to 1 atm)
Liquids and Solids (Chapter 6)
Intermolecular Forces
Weve already encountered the subject of intermolecular forces in discussing real gases; however,
if we are going to describe liquids or solids, we need to learn more about how molecules interact.
The common intermolecular forces are:
1. Coulomb: The strongest of these are ion-ion interactions, such as arises in molten salts. The form of
this is:
Vc ~ q1q2/r

53

although in reality the bare Coulomb force is screened by intervening ions.


2. Ion-dipole: Ion-dipole forces are, of course, the dominant forces when ions are dissolved in water. The
form of this is:
Vid ~ -|Z|/r2
where Z is the charge of the ion, is the dipole moment of the molecule and r is the ion-molecule
separation.
3.Dipole-dipole: Dipole-dipole forces arise in solids composed of polar molecules. Nominally these have
the form:
Vdd ~ -12/r3
where the dipoles are assumed to be oriented such that attractive interactions are more favorable.
In gases, and in liquids composed of small molecules, the dipole-dipole interaction is orientation
averaged, and almost vanishes. What is left is a weak attractive interaction:
Vdda ~ -1222/r6
Note that Vdda is much weaker than Vdd. In a gas of polar molecules it is Vdda that controls the
interactions; however if the same molecules are condensed to a solid, the orientation averaging doesnt
occur, and Vdd is relevant. Liquids composed of large rod-like molecules where rotation is slow also
experience Vdd rather than Vdda.
4. London dispersion forces: The picture just given would not explain liquefaction of a rare gas like Ar, so
there must be interactions between nonpolar molecules. This is called the London dispersion force.
To understand this force we need to imagine that as the electrons circulate around the nuclei, there is an
instantaneous dipole present, even on a rare gas atom. This gives rise to a dipole-dipole interaction that
would be averaged, like Vdda.
The form of the force is:

VL ~- 12/r6

where 1 and 2 are the polarizabilities of the two interacting species.


The polarizability measures the ability of the electron cloud to distort in the presence of an applied field.
Thus if a constant field E is applied to a rare gas atom, an induced dipole = E is produced.
VL reflects the mutual ability of the oscillating charge clouds to induce dipoles in the other species.
Polarizabilities scale with the ?????? of the atom or molecule.
London forces are always attractive, and they exist between all atoms and molecules. However they are
weak forces, so they tend to control liquefaction between nonpolar molecules like rare gases and
hydrocarbons.
Note that for gases and liquids composed of small molecules, the interactions (Vdda and VL) scale as r-6.
There are also dipole-induced dipole interactions, Vdind, which have the same dependence on r.
The sum of these interactions is called the van der Waals interaction:

54

Vvdw = Vdda + VL + Vdind


5. Hydrogen bonds: We have already introduced this concept. There is no simple formula for the
hydrogen bond force, and it is special to water, ammonia, HF and the O-H...O bonds in organics and
biomolecules. Such bonds play an important role in living systems.
6. Short range repulsion: If you push two molecules close enough to each other there is always repulsion
(which decays as e-r) when the electron clouds overlap. This is sometimes called ??????????, as it is the
Pauli principle which forces the electrons to occupy different states.
7. Chemical bonding: This is another short-ranged effect. In this case one needs to have empty or halfoccupied energy levels so that the interacting molecules can form a bond.

Properties of liquids
A liquid is a substance in which the distances between molecules are similar to atomic or ionic radii, but
where there is no long range order. The molecules in a liquid are always in motion, and indeed a frozen
liquid is called a glass.
Most liquids are nearly ??????????? (volume is independent of pressure), and the molar volume is only
weakly dependent on temperature.
The theory of liquids is beyond the scope of this course, but there are some points that we need to be
aware of concerning the gas-liquid and liquid-solid transitions:
1. Triple point: is the point where the gas/liquid/solid phases all
coexist. There is a single combination of T, p, V where a triple
point can occur.
If T>Ttp, then there is a liquid/gas coexistence line.
If T<Ttp, there is a solid/gas line.
The liquid/solid line can have a negative or positive slope,
depending on the substance (water=negative slope; most other
substances have positive slopes).

55

2. ???????? point: this is a T, p, V combination where the gas/liquid coexistence line ends. Above Tc,
there is no difference between the liquid and gas phases.

Structure of atomic solids


Solids may be crystalline or amorphous (as in a glass). In this course we will only consider crystalline
solids. There are four basic types of crystalline solids:
1. metallic solids - both s and d block metals, where the basic lattice building block are the atoms
2. ionic solids - NaCl, composed of interleaved lattices of positive and negative ions
3. network solids - diamond, where there a covalent bonds throughout the solid
4. molecular solids - ice, sugar, where the building blocks are the molecules and for intermolecular forces
listed above hold things together.

Simple solids: Here we consider metallic bonding where the bonding is not directional, and packing
effects dominate.
To describe the structure of closest packed solids, we refer to Fig. 6.26 and 6.27. This shows the first,
second and third layers of a hexagonal closest packed solid. Atoms are pushed together as close as
possible, with the second layer atoms fitting in the dimples formed by the first layer atoms.

Fig. 6.26
Fig. 6.27
Fig. 6.28
Note that the third layer is the same as the first, so the structure is ABAB... Figure 6.28 shows the basic
building block, which is given the name hexagonal closest packed (hcp).
Figure 6.29 shows another way to arrange the solid, in which the third layer is shifted relative to the first
to form an ABCABC structure.

56

Fig. 6.29
Figure 6.30 shows the basic building block of the ABCABC solid. This is a cube with atoms in the faces
of the cube. This is either called cubic closed packed (ccp) or face centered cubic (fcc).

Fig. 6.30
Both hcp and fcc structures are found in nature. The coordinate number around each atom in each
tructure is 12. These structures have one important difference: fcc can be cleaved along the faces of the
cube, while hcp has no such planes. This makes fcc structures more maleable. Thus copper is fcc, but
zinc is hcp.
Tetrahedral and octahedral holes: Fig. 6.32 shows that there are both octahedral and tetrahedral holes in a
fcc or hcp lattice. These will turn out to be important later on when we talk about ionic solids, and they
are also important to adding impurities to metals.

57

Fig. 6.32
Other simple structures: Two other common solid structures are primitive cubic and body centered cubic.
See figs. 6.34 and 6.35.

Figure 6.34

Figure 6.35

Unit cells:
The smallest unique unit within a lattice is called the unit cell. Fig. 6.36 shows unit cells for (a) fcc and
(b) bcc lattices. Here a dot indicates the location of atoms, with corner atoms being shared by eight cells,
face atoms by two cells and central atoms by just one cell.

Fig. 6.36
Filling factor calculation: An important property of crystalline solids is the ratio of the volume occupied
by the atoms or ions (imagined to be hard spheres) to the total volume; this ratio is the filling factor.
The figure below shows how this is calculated for a fcc lattice:

58

1. The face diagonal is the same as 4 times the radius.


2. If a is the cell dimension, then 2a2=(4r)2 so that a = (8) r. This means that the lattice cell volume is
a3 = (8)3/2 r3.
3. The total volume of spheres in the cell is the sum of the eight 1/8's that are at the corners, and the six
's that are on the faces. This makes the total sphere volume be 8/8 + 6/2 = 4.

59

4. Multiply this by 4/3 r3 (the volume of a sphere), and we get:

4 r 3

0.74
filling factor = (8)3/ 2 r 3
3 2
The same filling factor applies to the hcp lattice. This filling factor is the highest possible, as all other
lattices are less dense than these.

Bravais lattices:
There are ??? possible unit cell structures. These are called the Bravais lattices, and they are pictured in
Fig. 6.37. Weve already seen the first three, and most of the rest involve distortions relative to these
three.

Fig. 6.37
The assignment of cell types is normally done by doing X-ray diffraction on the crystal. The density
mass/volume of the crystal is also dependent on which cell type is involved.
Ionic solids:
The rock salt structure (Fig. 6.41) is appropriate to sodium chloride and other ionic solids where the ratio
of radii is in the range 0.4-0.7. Here the anions form a fcc lattice, and the cations fill the octahedral sites.
This is called (6,6)-coordination as the (cation, anion) coordination numbers are both 6.

60

Fig 6.41
When the radius ratio is greater than 0.7, we get the cesium chloride structure (Fig. 6.43). This is built
out of two interpenetrating primitive cubic lattices. The coordination number of anion and cation are both
eight, so this is an example of (8,8) coordination.

Fig. 6.43
When the radius ratio is less than 0.4, we get the Zinc blend structure (Fig. 6.45). This is a form of ZnS,
and it has the large anions in a fcc lattice, and the small Zn ions occupying half of the tetrahedral sites.
This gives rise to (4,4) coordination.

Fig. 6.45

61

Molecular solids:
In molecular solids, each molecule occupies a lattice site. A good example is sucrose. Here the
molecules are held together by hydrogen bond forces.
A more complicated example is ice. As shown in Fig. 6.19, the water molecules occupy lattice sites in
such a way that each oxygen interacts with four hydrogens, two via covalent bonds and two via hydrogen
bonds. This leads to a more open structure than occurs in liquid water, which makes it possible for liquid
water to have higher density.

Fig. 6.19
Note that ice has many different possible crystalline forms (but only ice I is important at atmospheric
pressure). These correspond to slight variations in the hydrogen bond distances.

Growth Processes in Ice Crystals


Crystal faceting

Branching instability
Fig. 6.21 shows density as a function of temperature for water, and for a more normal liquid, CCl4. The
density in the solid is 0.92 g/cm3 while that in the liquid is 1.00 g/cm3.

62

Fig. 6.21
Network solids
Network solids are essentially ?????????????, where all the atoms in the crystal are held together by
strong covalent (or other) forces.
A good example is diamond, which forms a tetrahedrally coordinated lattice (Fig. 6.22).

Fig 6.22

Graphite forms a two dimensional network (Fig. 6.23), with the planes held together by weak van der
Waals interactions.

63

Fig. 6.23

Special topic: high temperature superconductors.


Superconductivity is the complete absence of resistance, typically for temperatures below a transition
temperature. It was discovered in 1911 by the Dutch physicist, Heike Kammerlingh Onnes, for mercury
at 4K (liq He temp) and below. Onnes received the 1913 Nobel Prize for this.
In 1933, the ???????? effect was discovered. This is the exclusion of magnetic fields from a
superconductor.
Until 1986, the highest superconducting transition temperature was 23K. Most of the research in
superconductors was confined to metals and alloys.
In 1986 a new class of superconducting ceramic materials was discovered [J. G. Bednorz, K. A. Mueller,
Z. Phys. 64, 189 (1986)], and by 1988, materials with Tc above 100K were identified (the current record
for these materials is 133K). These materials have the general formula YBa2Cu3O6.5-7.0. The structure
(see figure) consists of a planar array of Cu atoms, surrounded by O2-. The superconducting current is
transported primarily within this array.
The figure for YBa2Cu3O7 (often known as YBCO), shows
conductivity as a function of temperature. Here the transition
temperature is 92K.

64

MgB2 is an important new class of superconducting materials that was discovered in 2001. The critical
temperature is 39K.
Liquid Crystals
Liquid crystals are an example of a state of matter that is in-between the solid and liquid phases, i.e., a
mesophase. Typically liquid crystals are formed from molecules that are long and rod-like. The rod-like
molecules allow for crystalline ordering in one or two dimensions, and then liquid-like behavior in the
remaining dimensions.
Types of liquid crystals:
nematic - molecules display a preferred orientation in a particular direction, but their centers are
distributed at random. Shake spaghetti in a box and you have a nematic liquid crystal.
smectic - similar to nematic, but ????????????????????. Within each layer, the center locations are
randomly distributed. Example: membranes
cholesteric - rod-like molecules form sheets that are rotated relative to each other into a helical structure.

Classification according to method of preparation:


thermotropic - refers to liquid crystals that are made by melting a solid. These can be nematic, smectic or
cholesteric. Applications include watches, computer screens, thermometers.
lyotropic- refers to liquid crystals that are made by dissolving (or mixing) a liquid or solid in a solvent.
Again these may be nematic, smectic or cholesteric. Applications: detergents and membranes.

65

Applications
Liquid Crystal Displays (first developed in 1970)

Biological membranes (an example of a smectic liquid crystal)


Basic building block: a phospholipid
This has both hydrophobic and hydrophilic parts, and thus is called amphiphilic

66

Вам также может понравиться