Sie sind auf Seite 1von 146

Chapter 5

Gases

Chemical Principles 7th Edition


Steven S. Zumdahl

5.1 Early Experiments


5.2 The Gas Laws of Boyle, Charles, and Avogadro
5.3 The Ideal Gas Law
5.4 Gas Stoichiometry
5.5 Daltons Law of Partial Pressures
5.6 The Kinetic Molecular Theory of Gases
5.7 Effusion and Diffusion
5.8 Collisions of Gas Particles with the Container Walls
5.9 Intermolecular Collisions
5.10 Real Gases
5.11 Characteristics of Real Gases
5.12 Chemistry in the Atmosphere

Hot air balloon taking off from the ski resort of Chateau dOex in the
Swiss Alps.

Matter exists in three distinct physical states: gas, liquid, and


solid. Of these, the gaseous state is the easiest to describe
both experiment and theoretically.

5.1 Early Experiments


In 1643 an Italian physicist named Evangelista Torricelli (16081647), who had been a student of Galileo, performed
experiments that showed that the air in the atmosphere
exerts pressure. (In fact, as we will see all gases exert
pressure.) Torricelli designed the first barometer by filling a
tube that was closed at one end with mercury and then
inverting it in a dish of mercury (see Fig. 5.1).

FIGURE 5.1
A torricellian barometer. The tube, completely filled with mercury, is
inverted in a dish of mercury. Mercury flows out of the tube until the
pressure of the column of mercury (shown by black arrow) standing
on the surface of the mercury in the dish is equal to the pressure of
the air (shown by green arrows) on the rest of the surface of the
mercury in the dish.

Units of Pressure
Because instruments used for measuring pressure, such as the
manometer (see Fig. 5.2), often use columns of mercury
because of its high density, the most commonly used units for
pressure are based on the height of the mercury column (in
millimeters) the gas pressure can support.
The unit millimeters of mercury (mm Hg) is called the torr in
honor of Torricelli.
A related unit for pressure is the standard atmosphere:
1 standard atmosphere = 1 atm = 760 mm Hg = 760 torr

1 atm: 760 mm Hg, 760 torr, 101.325 Pa (N/m2), 1.01325 bar;


29.92 in Hg, 14.7 lb/in2 (Psi).

FIGURE 5.2
A simple manometer, a device for measuring the pressure of a gas in
a container. The pressure of the gas is given by h (the difference in
mercury levels) in units of torr (equivalent to mm Hg). (a) Gas
pressure = atmospheric pressure h. (b) Gas pressure = atmospheric
pressure + h.

However, since pressure is defined as force per unit area,


Pressure = force/area

In the SI system the unit of force is the Newton (N) and the
unit of area is meters squared (m).
Thus the unit if pressure in the SI system is newtons per meter
squared (N/m), called the pascal (Pa). In terms of pascals the
standard atmosphere is
1 standard atmosphere = 101,325 Pa

The International Union of Pure and Applied Chemist (IUPAC)


has adopted 1 bar (100,000 Pa) as the standard pressure
instead of 1 atm (101,325 Pa). Both standards are now widely
used.

5.2 The Gas Laws of Boyle, Charles, and Avogadro


Boyles Law
The first quantitative experiments on gases were performed
by an Irish chemist, Robert Boyle (1627-1691). Using a Jshaped tube closed at one end (Fig. 5.3), which he reportedly
set up in the multistory entryway of his house, Boyle studied
the relationship between the pressure of the trapped gas and
its volume.

FIGURE 5.3
A J-tube similar to the one used by Boyle.

Respective values from Boyles experiments are given in Table


5.1.

These data show that the product of the pressure and volume
for the trapped air sample is constant within the accuracies of
Boyles measurements (note the third column in Table 5.1).
This behavior can be represented by the equation
PV = k

which is called Boyles law, where k is a constant at a specific


temperature for a given sample of air.
Figure 5.4(a) shows a plot of P versus V, which produces a
hyperbola. Notice that as the pressure drops by half, the
volume doubles. Thus there is an inverse relationship
between pressure and volume.

FIGURE 5.4
Plotting Boyles data from Table 5.1. (a) A plot of P versus V shows
that the volume doubles as the pressure is halved. (b) A plot of V
versus 1/P gives a straight line. The slope of this line equals the value
of the constant k.

The second type of plot can be obtained by rearranging


Boyles law to give
V = k/p

which is the equation for a straight line of the type


y = mx + b

where m represents the slope and b is the intercept of the


straight line. In this case, y = V, x = 1/P, m = k, and b = 0. Thus
a plot of V versus 1/P using Boyles data gives a straight line
with an intercept of zero, as shown in Fig. 5.4(b).
Highly accurate measurements in various gases at a constant
temperature have shown that the product PV is not quite
constant but changes with pressure.
Results for several gases are shown in Fig. 5.5.

FIGURE 5.5
A plot of PV versus P for several gases. An ideal gas is expected to
have a constant value of PV, as shown by the dashed line. Carbon
dioxide shows the largest change in PV, and this change is actually
quite small: PV changes from approximately 22.39 L atm at 0.25 atm
to 22.26 L atm at 1.00 atm. Thus Boyles law is a good approximation
at these relatively low pressures.

Note the small changes that occur in the product PV as the


pressure is varied.
Boyles law: V 1/P at constant temperature.
A gas that obeys Boyles law is called an ideal gas.

Example 5.1
In a study to see how closely gaseous ammonia obeys Boyles
law, several volume measurements were made at various
pressures, using 1.0 mol of NH3 gas at a temperature of 0C.
Using the results listed blow, calculate the Boyles law
constant for NH3 at the various pressure.
Experiment
1
2
3
4
5
6

Pressure (atm)
0.1300
0.2500
0.3000
0.5000
0.7500
1.000

Volume (L)
172.1
89.28
74.35
44.49
29.55
22.08

Solution
To determine how closely NH3 gas follows Boyles law under
these conditions, we calculate the value of k (in L atm) for
each set of values:
Experiment
k = PV

1
22.37

2
22.32

3
22.31

4
22.25

5
22.16

6
22.08

Although the deviations from true Boyles law behavior are


quite small at these low pressures, the value of k changes
regularly in one direction as the pressure is increased. Thus, to
calculate the ideal value of k for NH3, plot PV versus P, as
shown in Fig. 5.6, and extrapolate (extend the line beyond the
experimental points) back to zero pressure, where, for
reasons we will discuss later, a gas behaves most ideally. The
value of k obtained by this extrapolation is 22.41 L atm. This is
the same value obtained from similar plots for the gases CO2,
O2, and Ne at 0C, as shown in Fig. 5.5.

FIGURE 5.6
A plot of PV versus P for 1 mol of ammonia. The dashed line shows
the extrapolation of the data to zero pressure to give the ideal
value of PV of 22.41 L atm.

Charless Law
Charles found in 1787 that the volume of a gas at constant
pressure increases linearly with the temperature of the gas.
That is, a plot of the volume of a gas (at constant pressure)
versus its temperature (C) gives a straight line.
This behavior is shown for several gases in Fig. 5.7.
As with Boyles law, Charless law is obeyed exactly only at
relatively low pressures.
One very interesting feature of these plots is that the volumes
of all the gases extrapolate to zero at the same temperature,
273.2C.
On the Kelvin temperature scale this point is defined as 0 K,
which leads to the following relationship between the Kelvin
and Celsius scales:
Temperature (K) = 0C + 273

FIGURE 5.7
Plots of V versus T (C) for several gases. The solid lines represent
experimental measurements on gases. The dashed lines represent
extrapolation of the data into regions where these gases would
become liquids or solids. Note that the samples of the various gases
contain different numbers of moles.

When the volumes of the gases shown in Fig. 5.7 are plotted
versus temperature on the Kelvin scale, the plots in Fig. 5.8
result.
In this case the volume of each gas is directly proportional to
temperature and extrapolates to zero when the temperature
is 0 K.
This behavior is represented by the equation known as
Charless law,
V = bT

where T is the temperature (in Kelvins) and b is a


proportionality constant.
0 K is called absolute zero, and there is much evidence to
suggest that this temperature cannot be attained.
Temperature of approximately 106 K have been produced in
laboratories, but 0 K has never been reached.

FIGURE 5.8
Plots of V versus T as in Fig. 5.7 except that here the Kelvin scale is
used for temperature.

Avogadros Law
Equal volumes of gases at the same temperature and pressure
contain the same number of particles.
This observation is called Avogadros law, which can be stated
mathematically as
V = an

where V is the volume of the gas, n is the number of moles,


and a is a proportionality constant.
This equation states that for a gas at constant temperature
and pressure the volume is directly proportional to the number
of moles of gas.
This relationship is obeyed closely by gases at low pressures.

Cold Atoms
The latest low-temperature of Colorado in Boulder when a team
of scientists led by Carl Wieman reported that they had cooled a
sample containing 2 107 cesium atoms to 1.1 106 K, about
one-millionth of a degree above absolute zero.
This record-low temperature was achieved by a technique know
as laser cooling, in which a laser beam is directed against a beam
individual atoms, dramatically slowing the movement of the
atoms. The atoms are then further cooled in an optical molasses
produced by intersection of field, the Colorado scientists were
able to hold the supercold cesium atoms for about 1 s, and the
possibility exists that the cold atoms could be trapped for much
longer periods of time with improvements in the apparatus.

5.3 The Ideal Gas Law


We have considered three laws that describe the behavior of
gases as revealed by experimental observations:
Boyles law:
Charless law:
Avogadros law:

V = k/P
V = bT
V = an

(at constant T and n)


(at constant P and n)
(at constant T and P)

Number of moles of gas present can be combined as follows:


V = R(Tn/P)

where R is the combined proportionality constant called the


universal gas constant.
R has the value of 0.08206 L atm K1 mol1. The preceding
equation can be rearranged to the more familiar form of the
ideal gas law:
PV = nRT

The ideal gas law is an equation of state for a gas, where the
state of the gas in its condition at a given time. A particular
state of a gas is described by its pressure, volume,
temperature, and number of moles.
It is important to recognize that the ideal gas law is an
empirical equation it is based on the experimental
measurements of the properties of gases. A gas that obeys
this equation is said to behave ideally.
The ideal gas equation is best regarded as a limiting law it
expresses behavior that real gasses approach at low pressure
and high temperature.
The idea gas law applied best at pressure below 1 atm.

Example 5.2
A sample of hydrogen gas (H2) has a volume of 8.56 L at a
temperature of 0C and a pressure of 1.5 atm. Calculate the
moles of H2 present in this gas sample.

Solution
Solving the ideal gas law for n gives
n = PV/RT

In this case P = 1.5 atm, V = 8.56 L, T = 0C + 273 = 273 K, and R


= 0.08206 L atm K1 mol1. Thus
n = [(1.5 atm)(8.56 L)]/[(0.08206 L atm/K mol)(273 K)] = 0.57 mol

Example 5.3
Suppose we have a sample of ammonia gas with a volume of
3.5 L at a pressure of 1.68 atm. The gas is compressed to a
volume of 1.35 L at a constant temperature. Use the ideal gas
law to calculate the final pressure.

Solution
The basic assumption we make when using the ideal gas law
to describe a change in state for a gas is that the equation
applies equally well to both the initial and the final states. In
dealing with a change in state, we always place the variables
on one side of the equals sign and the constants on the other.
In this case the pressure and volume change, while the
temperature and the number of moles remain constant (as
does R, by definition). Thus we write the ideal gas law as
PV = nRT
Change Remain constant

Since n and T remain the same in this case, we can write P1V1
= nRT and P2V2 = nRT. Combining these equations gives
P1V1 = nRT = P2V2

or

P1V1 = P2V2

We are giver P1 = 1.68 atm, V1 = 3.5 L, V2 = 1.35 L. Solving for


P2 gives
P2 = (V1/V2)P1 = (3.5 L/1.35 L) 1.68 atm = 4.4 atm

Check: Does this answer make sense? The volume decreased


(at constant temperature), so the pressure should increase, as
the result of the calculation indicates. Note that the calculated
final pressure is 4.4 atm. Because most gases do not behave
ideally above 1 atm, we might find that if we measured the
pressure of this gas sample, the observed pressure would
differ slightly from 4.4 atm.

Example 5.4
A sample of methane gas that has a volume of 3.8 L at 5C is
heated to 86C at constant pressure. Calculate its new volume.

Solution
To solve this problem, we take the ideal gas law and segregate
the changing variables and the constants by placing them on
opposite sides of the equation. In this case volume and
temperature change, and number of moles and pressure (and
of course R) remain constant. Thus PV = nRT becomes
V/T = nR/P

which leads to
V1/T1 = nR/P

and

V2/T2 = nR/P

Combining these equations gives


V1/T1 = nR/P = V2/T2

or

V1/T1 = V2/T2

We are given
T1 = 5C + 273 = 278 K
V1 = 3.8 L

T2 = 86C + 273 = 359 K


V2 = ?

Thus
V2 = T2V1/T1 = (359 K)(3.8 L)/278 K = 4.9 L

Check: Is the answer sensible? In this case the temperature


was increased (at constant pressure), so the volume should
increase. Thus the answer makes sense.

Example 5.5
A sample of diborane gas (B2H6), a substance that bursts into
flames when exposed to air, has a pressure of 345 torr at a
temperature of 15C and a volume of 3.48 L. If conditions are
changed so that the temperature is 36C and the pressure is
468 torr, what will be the volume of the sample?

Solution
Since, for this sample, pressure, temperature, and volume all
change while the number of moles remains constant, we use
the ideal gas law in the form
PV/T = nR

which leads to
P1V1/T1 = nR = P2V2/T2

or

P1V1/T1 = P2V2/T2

Then
V2 = (T2P1V1)/(T1P2)

We have
P1 = 345 torr
T1 = 15C + 273 = 258 K
V1 = 3.48 L

P2 = 468 torr
T2 = 36C + 273 = 309 K
V2 = ?

Thus
V2 = [(309 K)(345 torr)(3.48 L)]/[(258 K)(468 torr)] = 3.07 L

Always convert temperature to the Kelvin scale when applying


the ideal gas law.

5.4 Gas Stoichiometry


Suppose we have 1 mole of an ideal gas at 0C (273.2 K) and 1
atm. From the ideal gas law, the volume of the gas is given by
V = nRT/P
= [(1.000 mol)(0.08206 L atm K1 mol1)(273.2 K)]/1.000 atm
= 22.42 L

This volume of 22.42 liters is called the molar volume of an


ideal gas.

The measured molar volumes of several gases are listed in


Table 5.2.

The conditions 0C and 1 atm, called standard temperature


and pressure (abbreviated STP) are common reference
conditions for the properties of gasses. For example the molar
volume of an ideal gasses is 22.42 liter at STP.

Example 5.6
Quicklime (CaO) is produced by the thermal decomposition of
calcium carbonate (CaCO3). Calculate the volume of CO2
produced at STP from the decomposition of 152 g of CaCO3
according to the reaction.

Solution
We use the same strategy we used in the stoichiometry
problems earlier in the book. That is, we compute the number
of moles of CaCO3 consumed and the number of moles of CO2
produced. The moles of CO2 can then be converted to volume
by using the molar volume of an ideal gas.
Using the molar of CaCO3, we can calculate the number of
moles of CaCO3:
152 g CaCO3 (1 mole CaCO3/100.1 g CaCO3) = 1.52 mol CaCO3

Since each mole of CaCO3 produces 1 mol of CO2, 1.52 mol of


CO2 will be formed. We can compute the volume of CO2 at STP
by using the molar volume:
1.52 mol CaCO3 22.42 L CO2/1 mol CO2 = 34.1 L CO2

Thus the decomposition of 152 g of CaCO3 will produce 34.1 L


of CO2 at STP.

Remember that the molar volume of an ideal gas is 22.42 L at


STP.

Molar Mass
One very important use of the ideal gas law is in the
calculation of the molar mass (molecular weight) of a gas from
its measured density. To understand the relationship between
gas density and molar mass, note that the number of moles of
gas n can be expressed as
n = (grams of gas/molar mass) = (mass/molar mass) = (m/molar mass)

Substitution into the ideal gas equation gives


P = nRT/V = [(m/molar mass)RT]/V = [m(RT))/(V(molar mass)]

But m/V is the gas density, d, in units of grams per liter. Thus
or

P = dRT/molar mass
Molar mass = dRT/P

Thus, if the density of a gas at a given temperature and


pressure is known, its molar mass can be calculated.

5.5 Daltons Law of Partial Pressures


In 1803 Dalton summarized his observations as follows: For a
mixture of gases in a container, the total pressure exerted is
the sum of the pressures that each gas would exert if it were
alone.
This statement, known as Daltons law of partial pressures,
can be expressed as follows:
PTotal = P1 + P2 + P3 +

The pressures P1, P2, P3, and so on, are called partial
pressures.
Assuming that each gas behaves ideally, the partial pressure
of each gas can be calculated from the ideal gas law:
P1 = n1RT/V, P2 = n2RT/V, P3 = n3RT/V,

The total pressure of the mixture, PTotal, can be represented as


PTotal = P1 + P2 + P3 + = n1RT/V + n2RT/V + n3RT/V
= (n1 + n2 + n3 +)(RT/V) = nTotal(RT/V)

where nTotal is the sum of numbers of moles of the various


gases.
Thus, for a mixture of ideal gases, it is the total number of
moles of particles that is important, not the identity or
composition of the individual gas particles.
This idea is illustrated in Fig. 5.9.
At this point we need to define the mole fraction: the ratio of
the number of moles of a given component in a mixture to the
total number of moles in the mixture.
The Greek letter chi () is used to symbolize the mole fraction.
For a given component in a mixture, the mole fraction 1 is
1 = n1/nTotal = n1/(n1 + n2 + n3 +)

FIGURE 5.9
The partial pressure of each gas in a mixture of gases depends on the
number of moles of that gas. The total pressure is the sum of the
partial pressures and depends on the total moles of gas particles
present, no matter what their identities.

Since
n = P(V/RT)

That is, for each component in the mixture,


n1 = P1(V/RT), n2 = P2(V/RT),

Therefore, we can represent the mole fraction in terms of


pressures:
1 = n1/nTotal = P1(V/RT)/[P1(V/RT) + P2(V/RT) + P3(V/RT) + ]
n1

n1

n2

n3

= [(V/RT)P1]/[(V/RT)(P1 + P2 + P3 + )]
= P1/(P1 + P2 + P3 + ) = P1/PTotal

Similarly,
2 = n2/nTotal = P2/PTotal

and so on. Thus the mole fraction of a particular component in


a mixture of ideal gases is directly related to its partial
pressure.

The expression for the mole fraction,


1 = P1 / PTotal

can be rearranged:
P1 = 1 PTotal

That is, the partial pressure of a particular component of a


gaseous mixture is equal to the mole fraction of that
component times the total pressure.
Fig. 5.10 shows the collection of oxygen gas produced by the
decomposition of solid potassium chlorate.

FIGURE 5.10
The production of oxygen by thermal decomposition of KClO3. The
MnO2 catalyst is mixed with the KClO3 to make the reaction faster.

The Chemistry of Air Bags


Most experts agree that air bags represent a very important
advance in automobile safety.
Air bag is activated when a serve deceleration (an impact) causes
a steel ball to compress a spring and electrically ignite detonator
cap, which, in turn, causes sodium azide (NaN3) to decompose
explosively, forming sodium and nitrogen gas:
2NaN3(s) 2Na(s) + 3N2(g)

This system works very well and required only a relatively small
amount of sodium azide [100 g yield 56 L of N2(g) at 25C and 1
atm].

Sodium azide, besides being explosive, has a toxicity roughly


equal to that of sodium cyanide. It also forms hydrazoic acid
(HN3), a toxic and explosive liquid, when treated with acid.

NaN3

Example 5.7
The mole fraction of nitrogen in air is 0.7808. Calculate the
partial pressure of N2 in air when the atmospheric pressure is
760 torr.

Solution
The partial pressure of N2 can be calculated as follows:
PN = N PTotal = 0.7808 760 torr = 593 torr
2

Example 5.8
A sample of solid potassium chlorate (KClO3) was heated in a
test tube (see Fig. 5.10) and decomposed according to the
following reaction:
2KClO3(s) 2KCl(s) + 3O2(g)

The oxygen produced was collected by displacement of water


at 22C at a total pressure of 754 torr. The volume of the gas
collected was 0.650 L, and the vapor pressure of water at 22C
is 21 torr. Calculate the partial pressure of O2 in the gas
collected and the mass of KClO3 in the sample that was
decomposed.

Solution
First, we find the partial pressure of O2 from Daltons law of
partial pressure:
PTotal = PO + PH O = PO + 21 torr = 754 torr
2

Thus
PO = 754 torr 21 torr = 733 torr
2

Now we use the ideal gas law to find the number of moles of
O2:
nO = (PO V)/RT
2

In this case
PO = 733 torr = 733 torr/(760 torr/atm) = 0.964 atm
2
V = 0.650 L
T = 22C + 273 = 295 K
R = 0.08206 L atm K1 mol1

Thus
nO = [(0.964 atm)(0.650 L)]/[(0.08206 L atm K1 mol1)(295 K)]
2
= 2.59 102 mole

Next, we calculate the moles of KClO3 needed to produce this


quantity of O2, using the mole ratio from the balanced
equation for the decomposition of KClO3:
2.59 102 mole O2 (2 mol KClO3/3 mol KClO3) = 1.73 102 mole KClO3

Using the molar mass of KClO3, we calculate the grams of


KClO3:
1.73 102 mole KClO3 (122.6 g KClO3/1 mol KClO3) = 2.12 g KClO3

Thus the original sample contained 2.12 g of KClO3.

5.6 The Kinetic Molecular Theory of Gases


On the basis of observations from different types of
experiments, we know that at pressure less than 1 atm most
gases closely approach the behavior described by the ideal gas
law. Now we want to construct a model to explain this
behavior.
However, although laws summarize observed behavior, they
do not tell us why nature behaves in the observed fashion.
The model in chemistry consist of speculations about what
individual atoms or molecules might be doing to cause the
observed behavior of the macroscopic system.

A model is consisted successful if it explains the observed


behavior in question and predicts correctly the results of
future experiments. Note that a model can never be proved to
be absolutely true.
In fact, any model is an approximation by its very nature and
is bound to fail at some point.
An example of this type of model is the kinetic molecular
theory, a simple model that attempts to explain the
properties of an ideal gas. This model is based on speculations
about the behavior of the individual gas particles (atoms or
molecules).

Kinetic Molecular Theory


The particles are so small compared with the distances
between them that the volume of the individual particles can
be assumed to be negligible (zero).
The particles are in constant motion. The collisions of the
particles with the walls of the container are the cause of the
pressure exerted by the gas.
The particles are assumed to exert no forces on each other;
they are assumed to neither attract nor repel each other.
The average kinetic energy of a collection of gas particles is
assumed to be directly proportional to the Kelvin temperature
of the gas.

Of course, real gas particle do have a finite volume and do


exert forces on each other. Thus they do not conform exactly
to these assumptions. But we will see that these postulates do
indeed explain ideal gas behavior.
The true test of a model is how well its predictions fit the
experimental observations. The postulates of the kinetic
molecular model picture an ideal gas as consisting of particles
have no volume and no attraction for each other, and the
model assumes that the gas produces pressure on its
container by collisions with the walls.

Separating Gases
Assume you work for an oil company that owns a huge natural
gas reservoir containing a mixture of methane and nitrogen
gases.
Your job is to separate the nitrogen (N2 size 430 pm) from the
methane (CH4 size 410 pm). How might you accomplish this
task?
They produced a molecular sieve in which the pore (passage)
sizes can be adjusted precisely enough to separate N2 molecules
from CH4 molecules. The material involved is a special hydrate
trianosilicate (contains H2O, Ti, Si, O, and Sr) compounds and
known as ETS-4.

The researchers has shown that the material can be used to


separate N2 ( 410 pm) from O2 ( 390 pm). They have also shown
that it is possible to reduce the N2 content of natural gas from 18%
to less than 5% with a 90% recovery of CH4.

Molecular sieve framework of titanium (blue), silicon (green), and


oxygen (red) atoms contract on heatingat room temperature (left)
d = 4.27 ; at 250C (right) d = 3.94 .

The Quantitative Kinetic Molecular Model


Suppose there are n moles of an ideal gas in a cubical
container with sides each of length L in meters Assume each
gas particle has a mass m and that it is in rapid, random,
straight-line motion colliding with the walls, as shown in Fig.
5.11.
The collisions will be assumed to be elastic no loss of kinetic
energy occur.
Each particle in the gas has a particular velocity u, which can
be broken into components ux, uy, and uz, as shown in Fig.
5.12.
uxy2 = ux2 + uy2
Hypotenuse of
right triangle

Sides of
right triangle

FIGURE 5.11
An ideal gas particle in a cube whose sides are of length L (in meters).
The particle collides elastically with the walls in a random, straightline motion.

FIGURE 5.12
(a) The Cartesian coordinate axes. This can be represented as a
rectangular solid with sides ux, uy, and uz and body diagonal u. (b) The
velocity u of any gas particle can be broken down into three mutually
perpendicular components, ux, uy, and uz. (c) In the xy plane, ux2 + uy2
= uxy2 by the Pythagorean theorem. Since uxy and uz are also
perpendicular, u2 = uxy2 + uz2= ux2 + uy2 + uz2.

Then, constructing another triangle as shown in Fig. 5.12, we


find
u2 = uxy2 + uz2
u2 = ux2 + uy2 + uz2
Note that only the x component of the velocity affects the
particles impacts on these two walls, as shown in Fig. 5.13.
The larger the x component of the velocity, the faster the
particle travels between these two walls, thus producing more
impacts per unit of time on these walls.
Remember that the pressure of the gas is caused by these
collisions with the walls.

FIGURE 5.13
(a) Only the x component of the gas particles velocity affects the
frequency of impacts on the shaded walls, the walls that are
perpendicular to the x axis. (b) For an elastic collision, there is an
exact reversal of the x component of the velocity and of the total
velocity. The change in momentum (final initial) is then mux
mux= 2mux

The collision frequency (collisions per unit of time) with the


two walls that are perpendicular to the x axis is given by
(Collision frequency)x
= velocity in the x direction/distance between the walls = ux/L

Next , what is the force of a collision? Force is defined as mass


times acceleration:
F = ma = m(u/t)

where F represents forces, a represents the acceleration, u


represents a change in velocity, t represents a given length
of time.
Since we assume that the particle has constant mass, we can
write
F = (mu)/t = (mu)/t

Remember that an elastic collision means that there is no


change in the magnitude of the velocity. The change in
momentum in the x direction is
Change in momentum = (mux)
= final momentum initial momentum
= mux mux
Final
momentum
in x diretion

Iinal
momentum
in x diretion

= 2mux

Recall that since force is the change in momentum per unit of


time, then
Forcex = (mux)/t

This expression can be obtained by multiplying the change in


momentum per impact by the number of impacts per unit of
time:
Forcex = (2mux)(ux/L) = change in momentum per unit of time
Change in
Momentum per impact

Impacts per
unit of time

That is,
Forcex = 2mux2/L
Forcey = 2muy2/L
Forcez = 2muz2/L

Since we have shown that


u2 = ux2 + y2 + uz2

The total force on the box is


ForceTotal = forcex + forcey + forcez
= 2mux2/L + 2muy2/L + 2muz2/L
= (2m/L)(ux2 + y2 + uz2) = (2m/L)(u2)

We use the average of the square of the velocity u2 to obtain


ForceTotal = (2m/L)(u2)

Next, we need to calculate the pressure:


Pressure caused by average particle = forceTotal / areaTotal
= (2mu2/L)/6L2 = mu2/3L3
The 6 sides
of the cube

Area of each
side

Since the volume V of the cube is equal to L3, we can write


Pressure = P = mu2/3V

The number of particles in a given gas sample can be


expressed as follows:
Number of gas particles = nNA

The total pressure on the box caused by n moles of a gas is


therefore
P = nNA(mu2/3V)

We have
P = (2/3)[nNA(1/2mu2)/V)]

KE = 1/2mu2, the energy caused by the motion of particle.


We get the average kinetic energy for a mole of gas particles:
(KE)avg = NA(1/2mu2)

Use this definition, we can rewrite the expression for pressure


as
P = (2/3)[n(KE)avg/V)]

or

PV/n = (2/3)(KE)avg

The fourth postulate of the kinetic molecular theory is that


the average kinetic energy of the particles in the gas sample is
directly proportional to the temperature in Kelvins. Thus,
since (KE)avg T, we can write
PV/n = (2/3)(KE)avg T

or

PV/n T

Compare the ideal gas law,


PV/n = RT From experiment

With the result from the kinetic molecular theory,


PV/n T From theory

These expression have exactly the same form of R, the


universe gas constant, is considered the proportionally
constant in the second case.

(Left) A balloon filled with air at room temperature. (Center) The


balloon is dipped into liquid nitrogen at 77 K. (Right) The balloon
collapses at the molecules inside slow down because of the decrease
temperature. Slower molecules produce a lower pressure.

Room temperature

77 K

The Meaning of Temperature


The exact relationship between temperature and average
kinetic energy can be obtained by combining the equations
PV/n = RT = (2/3)(KE)avg

which yields the expression


(KE)avg = (3/2)RT

The Kelvin temperature is an index of the random motions of


the particles of a gas, with higher temperature meaning
greater motion.

Root Mean Square Velocity


The square root of u is called the root mean square velocity
and is symbolized urms:
urms = u

We can obtain an expression for urms from the equations


(KE)avg = NA[(1/2)mu2)]

and

(KE)avg = (3/2)RT

Combination of these equations gives


NA[(1/2)mu2)] = (3/2)RT

or

u2 = (3RT)/(NAm)

Taking the square root of both sides of the last equation


produce
u = urms = (3RT)/(NAm)

In these expression m represents the mass in kilograms of a


single gas particle.

When NA, the number of particles in a mole, is multiplied by


m, the product is the mass of a mole of gas particles in
kilograms. We will call this quantity M. Substituting M for NAm
in the equation for urms, we obtain
urms = (3RT)/(M)

Before we can use this equation, we need to consider the


units for R. So far we have used 0.08206 L atm K1 mol1 as the
value for R.
A joule (J) is defined as a kilogram meter squared per second
squared (kg m2/s2). When R is converted from liter
atmospheres to joules, it has the value 8.3145 J K1 mol1.
When R with these units is used in the expression (3RT)/(M),
urms has units of meters per second, as described.
If the path of a particular gas particle could be monitored, it
would probably look very erratic, something like that shown in
Fig. 5.14.

FIGURE 5.14
Path of particle in a gas. Any given particle will continuously change
its course as a result of collisions with other particles, as well as with
the walls of the container.

The average distance a particle travels between collisions in a


particular gas sample is called the mean free path.
It is typically a very small distance (1 107 m for O2 at STP).
The actual distribution of molecular velocities for oxygen gas
at STP is shown in Fig. 5.15.
Figure 5.16 shows the velocity distribution for nitrogen gas at
three temperature.
The distribution of velocities of the particles in an ideal gas is
described by the MaxwellBoltzmann distribution law:
f(u) = 4(m/2kBT)(3/2)u2e(mu

2/2k T)
B

where u = velocity in m/s


m = mass of a gas particle in kg
kB = Boltzmanns constant = 1.38066 1023 J/K
T = temperature in K

FIGURE 5.15
A plot of the relative number of O2 molecules that have a given
velocity at STP.

FIGURE 5.16
A plot of the relative number of N2 molecules that have a given
velocity at three temperatures. Note that as the temperature
increases, both the average velocity (reflected by the curves peak)
and the spreads of velocities increase.
273 K
1273 K
2273 K

Analysis of the expression for f(u) yields the following


equation for the most probable velocity ump (the velocity
possessed by the greatest number of gas particles):
ump = (2kBT)/(m) = (2RT)/(M)

where
M = molar mass of the gas particles in kg = 6.022 1023 m
R = gas constant = 6.022 1023 kB
Another type of velocity that can be obtained from f(u) is the
average velocity uavg (sometimes written u), which is given by
equation
uavg = u = (8kBT)/(m) = (8RT)/(M)

ump: uavg: urms stand in the ratios:


ump: uavg: urms = 1.000 : 1.128 : 1.225

This relationship is shown for nitrogen gas at 0C in Fig. 5.17.

FIGURE 5.17
A velocity distribution for nitrogen gas at 273 K, with the values of
most probable velocity (ump, the velocity at the curve maximum), the
average velocity (uavg), and the root mean square velocity (urms)
indicated.

5.7 Effusion and Diffusion


Diffusion is the term used to describe the mixing of gases. The
rate of diffusion is the rate of the mixing of gases.
Effusion is the term used to describe the passage of a gas
through a tiny orifice into an evacuated chamber, as shown in
Fig. 5.18. The rate of effusion measures the rate at which the
gas is transferred into the chamber.

FIGURE 5.18
The effusion of a gas into an evacuated chamber. The rate of effusion
(the rate at which the gas is transferred across the barrier through
the pin hole) is inversely proportional to the square root of the mass
of the gas molecules.

Effusion
The relative rates of effusion of two gases at the same
temperature and pressure are given by the inverse ratio of the
square roots of the masses of the gas particles:
Rate of effusion for gas 1/Rate of effusion for gas 2 = M2/M1

where M1 and M2 represent the molar masses of the gases.


This equation is called Grahams law of effusion.
In Grahams law the units for molar mass can be g/mol or
kg/mol, since the units cancel in the ratio M2/M1.
This reasoning leads to the following prediction for two gases
at the same temperature T:
Effusion rate for gas 1/Effusion rate for gas 2
= (uavg for gas 1)/(uavg for gas 2) = (8RT)/(M1)/(8RT)/(M2)
= M2/M1

Diffusion
Diffusion is frequently illustrated by the lecture
demonstration represented in Fig. 5.19, in which two cotton
plugs, one soaked in ammonia and the other hydrochloric
acid, are simultaneously placed at the ends of a long tube. A
white ring of ammonium chloride (NH4Cl) forms where the
NH3 and HCl molecules meet several minutes later:
NH3(g) + HCl(g) NH4Cl(s)
White solid

The progress of the gases through the tube is surprisingly slow


in light of the fact that the velocities of the HCl and NH3
molecules at 25C are approximately 450 and 660 meter per
second, respectively. Why does it take several minutes for the
NH3 and HCl molecules to meet?

FIGURE 5.19
(a) A demonstration of the relative diffusion rates of NH3 and HCl
molecules through air. Two cotton plugs, one dipped in HCl(aq) and
one dipped in NH3(aq), are simultaneously inserted into the ends of
the tube. Gaseous NH3 and HCl vaporizing from the cotton plugs
diffuse toward each other and, where they meet, react to form
NH4Cl(s). (b) When HCl(g) and NH3(g) meet in the tube, a white ring
of NH4Cl(s) forms.

The answer is that the tube contains air and thus the NH3 and
HCl molecules undergo many collision with O2 and N2
molecules as they travel through the tube.
dNH /dHCl = distance traveled by NH3/distance traveled HCl
3
= uavg(NH )/uavg(HCl) = MHCl/MNH = 36.5/17 = 1.5
3

However, careful experiments show that this prediction is not


borne out-the observed ratio of distances is 1.3, not 1.5 as
predicted by Grahams law.
If no air were present in the tube, the ratio of distances would
be 1.5 as predicted from Grahams law.
Natural uranium is mostly 92238U, which cannot be fissioned to
produce energy. It contains only about 0.7% of the fissionable
nuclide 92235U must be increased to about 3%. In the gas
diffusion enrichment process, the natural uranium (containing
238U and a small amount of 235U) react with fluorine to
92
92
form a mixture of 238UF6 and 235UF6.

Because these molecules have slightly different masses, which


allows them to be separated by a multistage diffusion process.
(238UF6 99.3% v.s 235UF6 0.7%)
Although the process is called gaseous diffusion, because the
chambers are separated by barriers that effectively allow only
individual UF6 molecules to pass through, it behaves like an
effusion process.
Thus we can find the actual ratio of the two types of UF6 in
chamber 2 from Grahams law:
Diffusion rate for 235UF6/Diffusion rate for 238UF6
= Mass(238UF6)/Mass(235UF6)
= (352.05 g/mol)/(349.03 g/mol)
= 1.0043

We can use this factor to calculate the ratio of 235UF6 / 238UF6


in chamber 2:
235UF

6/

238UF

= 1.0043 (235UF6/238UF6) = 1.0043(7/993)

Chamber 2

Chamber 1

= 1.0043(7.0493 103)
= 7.0797 103

This enrichment process (in 235UF6) continues as the slight


enriched gas in chamber 2 diffuses into chamber 3 and is
again enriched by a factor of 1.0043.
To calculate the number of steps required to enrich from
0.700% 235U to 3.00% 235U, we have the following equation:
(0.700 235UF6/99.3 238UF6) (1.0043)N = 3.00 235UF6/97.0 238UF6
Original ratio

where N represents the number of states.

Desired ratio

Thus
Origin ratio 1.0043 1.0043 1.0043 = final ratio
First
stage

Second
stage

Third
stage

Solving this equation for N yields 345.


A photo of actual diffusion cells is shown in Fig. 5.20.

FIGURE 5.20
Uranium-enrichment converters from the Paducah gaseous diffusion
plant in Kentucky.

5.8 Collisions of Gas Particles with the Container


Walls
We will define the quantity we are looking for as ZA, the
collision rate (per second) of the gas particles with a section of
wall that has an area A (in m).
We expect ZA to depend on the following factors:
1. The average velocity of the gas particles
2. The size of the area being considered
3. The number of particles in the container
How is the ZA expected to depend on the average velocity of
the gas particles? For example, if we double the average
velocity, we double the number of wall impacts, so ZA should
double. Thus, ZA depends directly on uavg:
ZA uavg

Similarly, ZA depends directly on A, the area of the wall under


consideration. Thus
ZA A

Thus ZA is expected to depend directly on N/V. That is, ZA


N/V.
In summary, ZA should be directly proportional to uavg, A, and
N/V:
ZA uavg A (N/V)

Note that the units for ZA expected from this relationship are
(m/s) m2 (particles)/m3 (particles)/s or (collisions)/s

The correct units for ZA are 1/s, or s1.


Substituting the expression for uavg gives
ZA (N/V)A [(8RT)/(M)]

A more detail analysis of this situation shows that


proportionality constant is 1/4. Thus the exact equation for ZA
is
ZA = (1/4) (N/V)A((8RT)/(M)) = A(N/V)((RT)/(2M))

Example 5.9
Calculate the impact rate in a 1.00-cm section of a vessel
containing oxygen gas at a pressure of 1.00 atm and 27C.

Solution
To calculate ZA, we must identify the values of the variables in
the equation
ZA = A(N/V)[(RT)/(2M)]

In this case A is given as 1.00 cm. However, to be inserted


into the expression for ZA, A must have the unit m. The
appropriate conversion gives A = 1.00 104 m.
The quantity N/V can be obtained from the ideal gas law by
solving for n/V and then converting to the appropriate units:
n/V = P/RT = 1.00 atm/[(0.08206 L atm/K mol)(300 K)]
= 4.06 102 mol/L

To obtain N/V, which has the units (molecules)/m, from n/V,


we make the following conversion:
N/V = (4.06 102 mol/L) [6.022 1023 (molecules/mole)] (1000 L/m3)
= 2.44 1025 molecules/m3

The quantity M represents the molar mass of O2 in kg. Thus


M = (32.0 g/mol) (1 kg/1000 g) = 3.20 102 kg/mol

Next, we insert these quantities into the expression for ZA:


ZA = A(N/V)[(RT)/(2M)] = (1.00 104 m2)(2.44 1025 m3)
[(8.3145 J/K mol)(300 K)]/[(2)(3.14)(3.20 102 kg/mol)]
= 2.72 1023 s1

That is, in this gas 2.72 10 collisions per second occur on


each 1.00-cm area of the container.

5.9 Intermolecular Collisions


In this section we will consider the collision frequency of the
particles in a gas. We will start by considering a single
spherical gas particle with diameter d (in meters) that moving
with velocity uavg. As this particle moves through the gas in a
straight line, it will collide with another particle only if the
other particle has its center in a cylinder with radius d, as
shown in Fig. 5.21.
Thus our particle sweeps out a cylinder of radius d and
length uavg 1 second during every second of it flight.
Therefore, the volume of the cylinder swept out per second is
V = volume = (d)2(uavg)(1 s)
Area of Length of
cylinder cylinder
slice

FIGURE 5.21
The cylinder swept out by a gas particle of diameter d.

To specify the number of gas particles, we use the number


density of the gas N/V, which indicates the number of gas
particles per unit volume. Thus we can write
Number of collisions = (volume swept out) N/V = d2(uavg)(N/V)
per second
= d2[(8RT)/(M)](N/V) = (N/V)d2[(8RT)/(M)]

What about the motions of the other particles?


They are moving in various directions with various velocities,
When the motions of the particles are accounted for, the
relative velocity of the primary particle becomes 2uavg rather
than the value uavg that we have been using. Thus the
expression for the collision rate becomes
Collision rate (per second) = Z = 2 (N/V) d2 [(8RT)/(M)]
= 4 (N/V) d2 [(RT)/(M)]

Example 5.10
Calculate the collision frequency for an oxygen molecule in a
sample of pure oxygen gas at 27C and 1.0 atm. Assume that
the diameter of an O2 molecule is 300 pm.

Solution
To obtain the collision frequency, we must identify the
quantities in the expression
Z = 4(N/V)d2(RT)/(M)

that are appropriate to this case. We can obtain the value of


N/V for this sample of oxygen by assuming ideal behavior.
From the ideal gas law
n/V = P/RT = 1.0 atm/[(0.08206 L atm/K mol)(300 K)]
= 4.1 102 mol/L

Thus
N/V = (4.1 102 mol/L)(6.022 1023 molecules/mol)(1000 L/m3)
= 2.5 1025 molecules/m3

From the given information we know that


d = 300 pm = 300 1012 m or 3 1010 m

Also, for O2, M = 3.20 10- kg/mol. Thus


Z = 4(2.5 1025 m3)(3 1010 m)
[(8.314 J K1 mol1)(300 K)]/(3.20 102 kg/mol)
= 4 109 (collisions)/s = 4 109 s1

Notice how large this number is. Each O2 molecule undergoes


approximately 4 billion collisions per second in this gas
sample.

Mean Free Path


If we multiply 1/Z by the average velocity, we obtain the mean
free path :
= (1/Z) uavg = distance between collisions
Time between Distance traveled
collisions (s)
per second

Substituting the expressions for 1/Z and uavg gives


= [1/(4(N/V)d2(RT)/(M)][(8RT)/(M)] = 1/(2(N/V)d2)

Example 5.11
Calculate the mean free path in a sample of oxygen gas at
27C and 1.0 atm.

Solution
Using data from the previous example, we have
= 1/[2(2.5 1025 m3)()(2(3 1010 m)2] = 1 107 m

Note that an O2 molecule travels only a very short distance


before it collides with another O2 molecule. This produces a
path for a given O2 molecule like the one represented in Fig.
5.14, where the length of each straight line is 107 m.

5.10 Real Gases


An ideal gas is a hypothetical concept. No gas exactly follows
the ideal gas law, although many gases come very close at low
pressures and/or high temperatures. Thus ideal gas behavior
can best be thought of as the behavior approached by real
gases under certain conditions.
Plots of PV/nRT versus P are shown for several gases in Fig.
5.22. For an ideal gas PV/nRT equals 1 under all conditions,
but notice that for real gases PV/nRT approaches 1 only at low
pressures (typically at 1 atm).
To illustrate the effect of temperature, we have plotted
PV/nRT versus P for nitrogen gas at several temperatures in
Fig. 5.23.

FIGURE 5.22
Plots of PV/nRT versus P for several gases (200 K). Note the
significant deviations from ideal behavior (PV/nRT = 1). The behavior
is close to ideal only at low pressure (less than 1 atm).

FIGURE 5.23
Plots of PV/nRT versus P for nitrogen gas at three temperatures. Note
that, although nonideal behavior is evident in each case, the
deviations are smaller at the higher temperatures.

The most important conclusion to be drawn from these plots


is that a real gas typically exhibits behavior that is closet to
ideal behavior at low pressures and high temperatures.
To follow van der Waals analyses, we start with the ideal gas
law,
P = nRT/V

Remember that this equation describes the behavior of a


hypothetical gas consisting of volumeless entities that do not
interact with each other.
P = nRT/V is also 1 at high pressures for many gases because
of a canceling of nonideal effects.
P is corrected for the finite volume of the particles. The
attractive forces have not yet been taken into account.

In contrast, a real gas consists of atoms or molecules that have


finite volumes. Thus the volume available to a given particle in
a real gas is less than the volume of container because the gas
particle themselves take up some of the space.
Pobs is usually called just P.
We have now corrected for both the finite volume and the
attractive forces of the particles.
The volume actually available to a given gas molecule is given
by the difference V nb.
This modification of the ideal gas equation leads to the
expression
P = nRT/(V nb)

The volume of the gas particles has now been taken into
account.

The next step is to account for the attractions that occur


among the particles in a real gas. The effect of these
attractions is to make the observed pressure Pobs smaller that
it would be if the gas particles did not interact:
Pobs = (P correction factor) = (nRT/(V nb) correction factor)

When gas particles come close together, attractive force


occur, which cause the particles to hit the wall slightly less
often than they would in the absence of these interactions
(see Fig. 5.24).
The size of the correction factor depends on the
concentration of gas molecules defined in terms of moles of
gas particles per liter (n/V).
In a gas sample containing N particles, there are N 1
partners available for each particle, as shown in Fig. 5.25.

FIGURE 5.24
(a) Gas at low concentration-relatively few interactions between
particles. The indicated gas particle exerts a pressure on the wall
close to that predicted for an ideal gas. (b) Gas at high concentrationmany more interactions between particles. Because of these
interactions the collision frequency with the walls is lowered, thus
causing the observed pressure to be smaller than if the gas were
behaving ideally.

FIGURE 5.25
Illustration of pairwise interactions among gas particles. In a sample
with 10 particles, each particle has 9 possible partners, to give
10(9)/2 = 45 distinct pairs. The factor of arises, because when
particle 1 is the particle of interest, we count the pair; and
when particle is the particle of interest, we count the pair.
However, and are the same pair, which we thus have
counted twice. Therefore, we must divide by 2 to get the correct
number of pairs.

Since the 1 2 pair is the same as the 2 1 pair, this


analysis counts each pair twice. Thus for N particles there are
N(N 1)/2 pairs. If N is a very large number, N 1
approximately equals N, given N2/2 possible pairs.
Thus the correction to ideal pressure for the attractions of the
particles has the form
Pobs = P a(n/V)2

where a is a proportionality constant (which includes the


factor of 1/2 from N2/2)
Inserting the corrections for both the volume and of the
particle and the attractions of the particles give the equation
Pobs =
Observed
Pressure

[nRT/( V nb)] a(n/V)2


Volume
Volume
of the correction
container

Pressure
correction

This equation can be rearranged to give the van der Waals


equation:
[Pobs + a(n/V)2] (V nb) = nRT
Corrected pressure

Corrected volume

Pideal

Videal

The values of a and b for various gases are given in Table 5.3.
These observations are illustrated in Fig. 5.26.
The fact that a real gas tends to behavior more ideally at high
temperatures can also be explained in terms of the van der
Waals model. At high pressure the particles are moving so
rapidly that the effects of interparticle interactions are not
very important.

a
H2 < N2 < CH4 < CO2

FIGURE 5.26
The volume occupied by the gas particles themselves is less
important at (a) large container volumes (low pressure) than at (b)
small container volumes (high pressure).

Cold Sounds
In view of the 1996 prohibition on the production of the ozonedamaging chloroflurocarbons (CFCs) formerly used in many air
conditioners and refrigerators, innovative methods for cooling
without CFCs are being considered.
Because of the forces that exist among its molecules, a real gas
becomes hot when it is compressed and cools when it expands.
A novel idea is to use sound waves to power a refrigerator. The
apparatus is fairly simple: Loudspeakers are mounted at each
end of a U-shaped tube filled with an inert gas. The sound waves
produced by the speakers cause pressure variations in the tube.
In low-pressure areas the gas cools; in high-pressure areas the
gas become hot. The hot areas are cooled by a heat exchanger
and over time the gas in the tube becomes very cold.

Acoustic refrigeration units has some real advantages: They have


no moving parts (except the speaker drivers) and they are
environmentally safe. Presently the sound-fridge is neither as
cheap nor as efficient as conventional cooling devices, but
Garrett believes that the unit can be made competitive within
several years.

5.11 Characteristic of Several Real Gases


The plot for H2(g) (Fig. 5.22) never drops below the ideal value
(1.0) in contrast to all the other gases. What is special about
H2 compared to these other gases?
Recall from Section 5.8 that the reason that the
compressibility of a real gas falls below 1.0 is that the actual
(observed) pressure is lower than occur in real gases. This
must mean that H2 molecules have very low attractive forces
for each other.
Note that H2 has the lowest value of a among the gases H2, N2,
CH4, and CO2 (Table 5.3).

Remember that the value of a reflects how much of a


correction must be made to adjust the observed pressure up
to the expected ideal pressure:
Pideal = Pobs + a(n/V)2

A low value of a reflects weak intermolecular forces among


the gas molecules.
Also notice that although the compressibility for N2 dips below
1.0, it dose not show as much deviation as that for CH4, which
in turn does not show as much deviation as the
compressibility for CO2.
Based on this behavior we can surmise that the importance of
intermolecular interactions increases in this order:
H2 < N2 < CH4 < CO2

This order is reflected by the relative a values for these gases


in Table 5.3.

5.12 Chemistry in the Atmosphere


The gases that are most important to us are located in the
atmosphere that surrounds the earths surface.
The average composition of the earths atmosphere near sea
level, with the water vapor removed, is shown in Table 5.4.
The atmosphere is a highly complex and dynamic system, but
for convenience we divide it into several layers based on the
way the temperature changes with altitude. (The lowest layer,
called the troposphere, is shown in Fig. 5.27.)
The chemistry occurring in the troposphere is strongly
influenced by human activities. Millions of tons of gases and
particulates are released into the troposphere by our highly
industrial civilization.

FIGURE 5.27
The variation of temperature and pressure with altitude. Note that
the pressure steadily decreases with increasing altitude but that the
temperature does not change monotonically.

Actually, it is amazing that the atmosphere can absorb so


much material with relatively small permanent changes.
Significant changes, however, are occurring. Severe air
pollution is found around many large cities, and it is probable
that long-range changes in the planets weather are taking
place.
The two main sources of pollution are transportation and the
production of electricity. The combustion of petroleum in
vehicles produced CO, CO2, NO, and NO2, along with unburned
molecules from petroleum.
The complex chemistry of polluted air appears to center on
ozone and the nitrogen oxides (NOx).
At the high temperatures in the gasoline and diesel engines of
cars and trucks, N2 and O2 react to form a small quantity of
NO, which is emitted into the air with the exhaust gases (see
Fig. 5.28).

FIGURE 5.28
Concentration (in molecules per million molecules of air) of some
smog components versus time of day. After P.A. Leightons classic
experiment, Photochemistry of Air Pollution, in Physical Chemistry:
A Series of Monographs, ed. Eric Hutchinson and P. Van
Rysselberghe, Vol. IX, New York: Academic Press, 1961.

This NO is oxidized in air to NO2, which in turn absorbs energy


from sunlight and breaks up into nitric oxide and free oxygen
atoms:
Radiant
energy

NO2(g) NO(g) + O(g)

Oxygen atoms are very reactive and can be combine with O2


to form ozones:
O(g) + O2(g) O3(g)

Although represented here as O2, the actual oxidant is an


organic peroxide such as CH3COO, formed by reaction of O2
with organic pollutants.
Ozone in also very reactive. It can react with the unburned
hydrocarbons in the polluted air to produce chemicals that
cause the eyes to water and burn and are harmful to the
respiratory system.

The end product of this whole process is often referred to as


photochemical smog, so called because light is required to
initiate some of the reactions.
NO2(g) NO(g) + O(g)
O(g) + O2(g) O3(g)
NO(g) + 1/2O2(g) NO2(g)
Net reaction:
3/2O2(g) O3(g)

We can observe this process by analyzing polluted air at


various times during a day (see Fig. 5.28).
Current efforts to combat the formation of photochemical
smog are focused on cutting down the amounts of molecules
from unburned fuel in automobile exhaust and designing
engines that produce less nitric oxide (see Fig. 5.29).

FIGURE 5.29
Our various modes of transportation produce large amounts of
nitrogen oxides, which facilitate the formation of photochemical
smog.

The other major source of pollution results from burning coal


to produce electricity. Much of the coal found in Midwest
contains significant quantities of sulfur, which, when burned,
produces sulfur oxide:
S(In coal) + O2(g) SO2(g)

A further oxidation reaction occurs when sulfur dioxide is


changed to sulfur trioxide in the air:
2SO2(g) + O2(g) 2SO3(g)

The production of sulfur trioxide is significant because it can


combine with droplet of water in the air to form sulfuric acid:
SO3(g) + H2O(l) H2SO4(aq)

Sulfuric acid is very corrosive to both living things and building


materials. Another result of this type of pollution is acid rain
(see Fig. 5.30).

FIGURE 5.30
A helicopter dropping lime in a lake in Sweden to neutralize excess
acid from acid rain.

One way to use high-sulfur coal without further harming the


air quality is to remove the sulfur oxide from the exhaust gas
by means of a system called a scrubber before it is emitted
from the power plant stock. A common method called
scrubbing involves blowing powdered limestone (CaCO3) into
the combustion camber, where it is decomposed to lime and
carbon dioxide:
CaCO3(s) CaO(s) + CO2(g)

The lime then combines with the sulfur dioxide to form


calcium sulfite:
CaO(s) + SO2(g) CaSO3(s)

The calcium sulfite and any remaining unreacted sulfur


dioxide are removed by injecting an aqueous suspension of
lime into the combustion chamber and the stack, producing a
slurry (a thick suspension), as shown in Fig. 5.31.

FIGURE 5.31
Diagram of the process for scrubbing sulfur dioxide from stack gases
in power plants.

Unfortunately, there are many problems associated with


scrubbing. The systems are complicated and expensive and
consume a great deal of energy. The large quantities of
calcium sulfite produced in the process present a disposal
problem.

The Important of Oxygen


Oxygen has been present only for the last half of the earths
history appearing about 2.5 billion years ago when organisms
began to use chlorophyll to convert sunlight to stored energy.
The burial of plants to begin the formation of coal and petroleum
also released oxygen into the atmosphere. In contrast, oxygen
was removed from the atmosphere by mountain building and
erosion as freshly exposed rocks combined with oxygen to form
oxygen-containing minerals.

Atmospheric Concentration of Oxygen

Acid Rain: An Expensive Problem


Rainwater, even in pristine wildness areas, is slightly acidic
because some of the carbon dioxide present in the atmosphere
dissolves in the raindrops to produce H+ ions by the following
reaction:
H2O(l) + CO2(g) H+(aq) + HCO3(aq)

This process produces only very small concentrations of H+ ions


in the rainwater. However, gases such as NO2 and SO2, which are
by-products of energy use, can produce significantly higher H+
concentrations. Nitrogen dioxide reacts with water to give a
mixture of nitrous acid and nitric acid:
2NO2(g) + H2O(l) HNO2(aq) + HNO3(aq)

Sulfur dioxide is oxidized to sulfur trioxide, which then reacts


with water to form sulfuric acid:
2SO2(aq) + O2(g) 2SO3(g)
SO3(g) + H2O(l) H2SO4(aq)

The damage cause by the acid formed in polluted air is a growing


worldwide problem. Lakes are dying in Norway, the forests are
sick in Germany, and buildings and status are deteriorating all
over the world.
Both marble and limestone react with sulfuric acid to form
calcium sulfate. The process can be represented most simply as
CaCO3(s) + H2SO4(aq) Ca2+(aq) + SO42(aq) + H2O(l) + CO2(g)

What can be done to protect limestone and marble structures


from this kind of damage? Of course, one approach is to lower
sulfur dioxide and nitrogen oxide emissions from power plant. In
addition, scientists are experimenting with coatings to protect
marble from the acidic atmosphere. However, a coating can do
more harm than good unless it breathes. If moisture trapped
beneath the coating freeze, the expanding ice can fracture the
marble. Needless to say, it is difficult to find a coating that will
allow water to pass but not allow acid to pass, so the search
continues.

The damaging effects of acid rain can be seen by comparing these


photos of a decorative statue on the Field Museum in Chicago. The
photo on the left was taken c. 1920; the photo on the right was taken
in 1990. Recent renovation has since replaced the deteriorating
marble.

Homework
25, 39, 45, 67, 69, 81, 97, 109, 113, 125.

146

Das könnte Ihnen auch gefallen