Добавил:
Upload Опубликованный материал нарушает ваши авторские права? Сообщите нам.
Вуз: Предмет: Файл:
Kinetic Theory of Gases.doc
Скачиваний:
9
Добавлен:
28.09.2019
Размер:
27.23 Mб
Скачать

26

Kinetic Theory of Gases

Molecules and intermolecular forces. Solids, liquids and gases

The evidence for molecular structure is based on such observations as Brownian motion, diffusion of gases and X-ray diffraction.

All familiar matter is made up of molecules. For any specific chemical compound, all the molecules are identical. The smallest molecules contain one atom each and are of the order of 10-10 m in size; the largest contain many atoms and are at least 10,000 times larger. In gases the molecules move nearly independently; in liquids and solids they are held together by intermolecular forces that are electrical in nature, arising from interactions of the electrically charged particles that make up the molecules. Gravitational forces between molecules are negligible in comparison with electrical forces.

The interaction of two point electric charges is described by a force (repulsive for like charges, attractive for unlike charges) with a magnitude proportional to , where is the distance between the points. Molecules are not point charges but complex structures containing both positive and negative charges, and their interactions are more complex. The force between molecules in a gas varies with the distance between molecules somewhat as shown in Fig. 1, where a positive corresponds to a repulsive force and a negative to an attractive force. When molecules are far apart, the intermolecular forces are very small and usually attractive. As a gas is compressed and its molecules are brought closer together, the attractive forces increase. The intermolecular force becomes zero at an equilibrium spacing , corresponding roughly to the spacing between molecules in the liquid and solid states. In liquids and solids, relatively large pressures are needed to compress the substance appreciably. This shows that at molecular distances slightly less than the equilibrium spacing, the forces become repulsive and relatively large.

Figure 1

Figure 1 also shows the potential energy as a function of . This function has a minimum at , where the force is zero. The two curves are related by . Such a potential energy function is often called a potential well. A molecule at rest at a distance from a second molecule would need an additional energy , the "depth" of the potential well, to "escape" to an indefinitely large value of .

Molecules are always in motion; their kinetic energies usually increase with temperature. At very low temperatures the average kinetic energy of a molecule may be much less than the depth of the potential well. The molecules then condense into the liquid or solid phase with average intermolecular spacing of about . But at higher temperatures the average kinetic energy becomes larger than the depth of the potential well. Molecules can then escape the intermolecular force and become free to move independently, as in the gaseous phase of matter.

In solids, molecules vibrate about more or less fixed points. In a crystalline solid these points are arranged in a recurring crystal lattice. The vibration of molecules in a solid about their equilibrium positions may be nearly simple harmonic if the potential well is approximately parabolic in shape at distances close to . But if the potential-energy curve rises more gradually for than for , as in Fig. 1, the average position shifts to larger with increasing amplitude. This is the basis of thermal expansion.

In a liquid, the intermolecular distances are usually only slightly greater than in the solid phase of the same substance, but the molecules have much greater freedom of movement. Liquids show regularity of structure only in the immediate neighborhood of a few molecules. This is called short-range order, in contrast with the long-range order of a solid crystal.

The molecules of a gas are usually widely separated and so have only very small attractive forces. A gas molecule moves in a straight line until it collides with another molecule or with a wall of the container. In molecular terms, an ideal gas is a gas whose molecules exert no attractive forces on each other and therefore have no potential energy.

At low temperatures, most common substances are in the solid phase. As the temperature rises, a substance melts and then vaporizes. From a molecular point of view, these transitions are in the direction of increasing molecular kinetic energy. Thus temperature and molecular kinetic energy are closely related.

The Mole and the Avogadro Constant

The unit for the amount of a substance is the mole. It is defined as the amount of a substance that contains the same number of particles (atoms or molecules) as the number of atoms in 12 g of the isotope carbon-12. This number is known as the Avogadro constant ( ) and is particles per mole.

The ratio of the mass of one mole of the substance to one-twelfth of the mass of one mole of carbon-12 is called the relative molecular mass of the substance - it is 32 for oxygen, 2 for hydrogen and so on. The molar mass of a compound is the mass of 1 mole. It is equal to the mass of a single molecule multiplied by Avogadro’s number: .

A knowledge of the Avogadro constant enables us to calculate the number of molecules in any mass of a substance and therefore to get an idea of the size of one molecule. For example, a drop of water of volume =1.0 cm3 has a mass of =1 g. The relative molecular mass of water is 18. To obtain the number of molecules we express mass of the water as product of mass of one molecular and the total number of molecules : and . As molar mass is , then , and therefore this drop of water must contain molecules.

The average volume of a water molecule must therefore be cm3, and if we assume the molecules to be spherical the diameter of a water molecule is about cm or m, a result confirmed by X-ray diffraction.

Example 1 A sample of 2 mg of polonium is found to emit 2.90 x 1018 alpha-particles during one half-life. If the relative atomic mass of polonium is 210, calculate a value for the Avogadro constant.

Solution During a half-life one half of the polonium atoms will have decayed. Therefore 2 mg of polonium would emit 5.80 x 1018 alpha-particles before it decayed completely.

Therefore number of moles in 2 mg

Therefore one mole contains

particles

a good approximation to the Avogadro constant.

A gas is a collection of a large number of molecules, which are in continuous rapid motion. The gas molecules have a size (molecular diameter) of the order of 10-10 m. Ordinarily, the distance between the gas molecules is of the order 10-9 m, i.e. about 10 times as large as their size. Therefore, the gas molecules can be considered as to be moving freely with respect to each other. However, during motion, when they come close to each other, they suffer a change in their velocities due to the inter-molecular forces. Therefore, such changes in velocities of the molecules of a gas are due to molecular collision.

Rudolf Claussius and James Clark Maxwell developed the kinetic theory of gases in order to explain gas laws in terms of the motion of the gas molecules. The theory is based on the following assumptions as regards to the motion (molecules and the nature of the gases:

Assumptions:

1. All gases consist of molecules. The molecules of a gas are all alike and differ from those of other gases.

  1. The molecules of a gas are very small in size as compared to the distance between them.

  2. The molecules of a gas behave as perfect elastic spheres.

  3. The molecules are always in random motion. They have velocities in all directions ranging from zero to infinity.

  4. During their random motion, the molecules collide against one another and the walls of the containing vessel. The collisions of the molecules with one another and with the walls of the vessel are perfectly elastic.

  5. Between two collisions, a molecule moves along a straight line and the distance covered between two successive collisions is called the free path of the molecules.

  6. The collisions are almost instantaneous i.e. the time during which a collision takes place is negligible as compared to the time taken by the molecule to cover the free path.

  7. The molecules do not exert any force on each other except during collisions.

Equations of State

The conditions in which a particular material exists are described by physical quantities such as pressure, volume, temperature, and amount of substance. These variables describe the state of the material and are called state variables.

The volume of a substance is usually determined by its pressure , temperature , and amount of substance, described by the mass or number of moles . Ordinarily, we can't change one of these variables without causing a change in another. When the tank of oxygen gets hotter, the pressure increases. If the tank gets too hot, it explodes; this happens occasionally with overheated steam boilers.

The Ideal-Gas Equation

Experiments show that, at low enough densities, all real gases tend to obey the relation

or (ideal-gas equation) (1)

where is a proportionality constant. Equation (1) is called the ideal gas equation. Provided the gas density is low, this law holds for any single gas or for any mixture of different gases. (For a mixture, is the total number of moles in the mixture.) An ideal gas is one for which Eq. (1) holds precisely for all pressures and temperatures. This is an idealized model; it works best at very low pressures and high temperatures, when the gas molecules are far apart and in rapid motion. It is reasonably good (within a few percent) at moderate pressures (such as a few atmospheres) and at temperatures well above those at which the gas liquefies.

The numerical value of depends on the units of , , and . In SI units, in which the unit of is Pa (l Pa = l N/m2) and the unit of is m3, the current best numerical value of is

= 8.314472(15) j/mol • K

Note that the units of pressure times volume are the same as the units of work or energy (for example, N/m2 times m3); that's why has units of energy per mole per unit of absolute temperature.

We can express the ideal-gas equation, Eq. (1), in terms density of the gas, using :

.

From this we can get an expression for the density of the gas:

(2)

For a constant mass (or constant number of moles) of an ideal gas the product is constant, so the quantity pV/T is also constant. If the subscripts 1 and 2 refer to any two states of the same mass of a gas, then

.

Van der Waals Equation

The ideal gas equation can be obtained from a simple molecular model that ignores the volumes of the molecules themselves and the attractive forces between them.

Meanwhile, for real gases there is another equation of state, the van der Waals equation, that makes approximate corrections for these two omissions. This equation was developed by the 19th-century Dutch physicist J. D. van der Waals; the interaction between atoms was named the van der Waals interaction after him. The van der Waals equation is:

(3)

The constants and are empirical constants, different for different gases. Roughly speaking, represents the volume of a mole of molecules; the total volume of the molecules is then , and the net volume available for the molecules to move around in is . The constant depends on the attractive intermolecular forces, which reduce the pressure of the gas for given values of , , and by pulling the molecules together as they push on the walls of the container. The decrease in pressure is proportional to the number of molecules per unit volume in a layer near the wall (which are exerting the pressure on the wall) and is also proportional to the number of molecules per unit volume in the next layer beyond the wall (which are doing the attracting). Hence the decrease in pressure due to intermolecular forces is proportional to .

When is small (that is, when the gas is dilute), the average distance between molecules is large, the corrections in the van der Waals equation become insignificant, and Eq. (3) reduces to the ideal-gas equation.

Dalton's Law of Partial Pressures

If two or more gases are mixed in a container then the final pressure can be found from Dalton's law, which states that:

The pressure in a container is the sum of the partial pressures of the gases that occupy the container.

The partial pressure of a gas is that pressure that the gas would exert if it alone occupied the container. For example, consider a container with a volume and a temperature . Let the container be occupied by two different gases, there being \ moles of gas 1 and moles of gas 2 (Fig. 2). Let the partial pressures of the gases be and .

Now and . But the total pressure in the container is , where for , the total number of moles of whatever type occupying the container. Therefore

; and ,

and since , we have

or

which is Dalton's law.

A useful application of this law is in cases where volumes of the same gas occupy separate but connected containers at differing temperatures.

Pressure Exerted by a Gas

A gas exerts pressure on the walls of the containing vessel due to continuous collisions of the molecules against the wall. During collisions the molecules exert forces on the walls of the container; this is the origin of the pressure the gas exerts.

Consider a gas contained in a vessel. Suppose that the edges OA, OC and OE of the vessel are along X, Y and Z-axis respectively [Fig.3]. Let n be the number of the gas molecules per unit volume inside the vessel and be the mass of each molecule. Let be the velocity of any gas molecule at any instant. If , , components of the velocity along , and -axis respectively, then

.

The molecule moves with momentum along X-axis and strikes against the face ABGF. Since the collision is perfectly elastic in nature, the molecule rebounds back with same speed . As the direction of motion is reversed, the momentum of the molecule after the rebound becomes . Therefore, change in momentum of the molecule along -axis

Let us find the change in momentum of the gas molecules hitting the area of the face ABGF in time . Now, in time , all those molecules will hit the area of the face ABGF, which lie in a cylinder of length and area of cross-section .

Therefore, such molecules lie in volume . Since the number of molecules per unit volume is n, the number of such molecules is . In fact, on the average, half of this number is expected to move along negative -axis and the other half towards the face ABGF along positive -axis. Therefore, the number of molecules hitting the area in time along positive -axis

.

Therefore, the total change in momentum of molecules in time along -axis,

.

Now, different molecules or even the same molecule at different times may possess different velocities. Therefore, in the above expression for , it is reasonable to replace by , where is mean square component of the velocities of the molecules along -axis, Therefore, in the above equation, replacing by , we have

.

Now, the rate of change of momentum is called force (Second Newton’s law, ). Therefore, force exerted by walls of the vessel on the gas molecules along -axis is given by

.

According to Newton's third law of motion, the gas molecules also exert an equal but opposite force on the walls of the vessel. Therefore, force exerted by the gas molecules on the walls of container along -axis,

If is pressure exerted by gas molecules along -axis, then

.

If and are pressures along -axis and -axis respectively, then

and .

Since gas molecules have same properties in all directions, the choice of -axis, -axis and -axis is arbitrary. In other words, pressure exerted by the gas molecules must be same in all directions i.e.

(say)

In other words,

, or .

Or , (4)

where is mean square velocity of the gas molecules.

Now, , the density of the gas. Therefore,

. (5)

The expression for the pressure exerted by a gas can be put in many other useful forms. From the equation (4), we have

Now, = average kinetic energy of a gas molecule, ; .

If is total number of gas molecules in volume , then . Therefore, equation (4) becomes

or

(6)

Also, the equation (5) may be written as

.

Now, = average kinetic energy of the gas per unit volume. And, Finally,

. (7)

The assumption that individual molecules undergo perfectly elastic collisions with the container wall is actually a little too simple. More detailed investigation has shown that in most cases, molecules actually adhere to the wall for a short time and then leave again with speeds that are characteristic of the temperature of the wall. However, the gas and the wall are ordinarily in thermal equilibrium and have the same temperature. So there is no net energy transfer between gas and wall, and this discovery does not alter the validity of our conclusions.

Kinetic Energy of a Molecule and of One Mole of a Gas

Consider one mole of a monatomic gas. If , and are the pressure, volume and temperature, respectively, of the gas, then according to the ideal gas equation,

.. (8)

Let be molar mass of the gas. Then, density of the gas is . In the equation (5), substituting for , we have

or . (9)

From the equations (8) and (9), we have

. (10)

Therefore, average kinetic energy of one mole of the gas,

.

The equation (10) gives mean kinetic energy of one mole of the gas. Now, one mole of the gas contains molecules equal to Avogadro number. If is mass of one molecule and , the Avogadro number, then

Therefore, the equation (10) becomes

Therefore, average kinetic energy of one molecule,

,

Now, is equal to the universal gas constant per molecule of the gas and called Boltzmann's constant. It is denoted by and its value is J/(molecule x K). Therefore, the above equation may be written as

(11)

The equation (11) gives average kinetic energy per molecule of the gas and it may be noted that it does not depend upon the mass of the molecule of the gas.

Relation between pressure and temperature of a gas

Let’s return to perfect gas equation: . The molar mass can be expressed as product of mass of one molecular and number of molecules in one mole, that is Avogadro number : / Then . But , hence . The ratio of total mass and mass of one molecular gives the total number of molecules in given volume: and . Then, if we divide the total number of molecules in given volume by this volume, we obtain the number of molecules per unit volume, that is, number density of molecules in given volume . And, finally, we obtain very important result: pressure of gas depends on temperature as

. (12)

Degrees of freedom

So far we have only been considering monatomic gases. We must now extend the ideas to cover gases of higher atomicity, that is, molecules with more than one atom per molecule. As we know, kinetic energy of one monatomic molecule = . We have considered the motion of these molecules to be in three directions; we say that the molecule has three degrees of freedom, . It is therefore sensible to suppose that one-third of the total energy is associated with each degree of freedom, and this is known as Boltzmann's law of equipartition of energy. Thus each degree of freedom has an amount of energy associated with it.

Now consider a diatomic molecule. In addition to three translational degrees of freedom it can also rotate about three axes , and (Figure 7).

Figure 7

The energy associated with axis is very small, however, and so we say that the molecule has five degrees of freedom, . If we assume that the energy associated with each rotational degree of freedom is the same as that for each translational degree of freedom then the total energy of the molecule will be .

(The vibrational energy of the molecule is insignificant except at very high temperatures.)

For more complex molecules we have that (Three translational and three rotational)

Root-mean-square speed

From Eq. (11) we can obtain expressions for the square root of called the root-mean-square speed (or rms speed)

(13)

It might seem more natural to characterize molecular speeds by their average value rather than by , but we see that follows more directly from Eqs. (11). To compute the rms speed, we square each molecular speed, add, divide by the number of molecules, and take the square root; vrms is the root of the mean of the squares.

Equations (11) shows that at a given temperature T, gas molecules of different mass m have the same average kinetic energy but different root-mean-square speeds. On average, the nitrogen molecules (M = 28 g/mol) in the air around you are moving faster than are the oxygen molecules (M = 32 g/mol). Hydrogen molecules (M = 2 g/mol) are fastest of all; this is why there is hardly any hydrogen in the earth's atmosphere, despite its being the most common element in the universe. A sizable fraction of any H2 molecules in the atmosphere would have speeds greater than the earth's escape speed of 1.12 x 104 m/s and would escape into space. The heavier, slower-moving gases cannot escape so easily, which is why they predominate in our atmosphere.

The Distribution of Molecular Speeds

Thus far we have neglected the fact that not all molecules in a gas have the same speed and energy. In reality, their motion is extremely chaotic. Any individual molecule is colliding with others at an enormous rate - typically, a billion times per second. Each collision results in a change in the speed and direction of motion of each of the participant molecules. From Equation 21.7, we see that average molecular speeds increase with increasing temperature. What we would like to know now is the relative number of molecules that possess some characteristic, such as a certain percentage of the total energy or speed. The ratio of the number that have the desired characteristic to the total number of molecules is the probability that a particular molecule has that characteristic.

The root-mean-square speed gives us a general idea of molecular speeds in a gas at a given temperature. We often want to know more. For example, what fraction of the molecules have speeds greater than the rms value? Greater than twice the rms value? To answer such questions, we need to know how the possible values of speed are distributed among the molecules. Figure 4a shows this distribution for oxygen molecules at room temperature (T = 300 K); Fig.4b compares it with the distribution at T = 80 K.

In 1852, Scottish physicist James Clerk Maxwell first solved the problem of finding the speed distribution of gas molecules. His result, known as Maxwell's speed distribution law, is

(14)

Here is the molecular speed, T is the gas temperature, M is the molar mass of the gas, and R is the gas constant. It is this equation that is plotted in Fig. 4-a,b. The quantity P(v) in Eq. 14 and Fig. 4 is a probability distribution function: For any speed , the product (a dimensionless quantity) is the fraction of molecules whose speeds lie in the interval of width centered on speed .

As Fig. 4a shows, this fraction is equal to the area of a strip with height and width . The total area under the distribution curve corresponds to the fraction of the molecules whose speeds lie between zero and infinity. All molecules fall into this category, so the value of this total area is unity; that is,

Fig. 4 (a) The Maxwell speed distribution for oxygen molecules at T = 300 K. The three characteristic speeds are marked, (b) The curves for 300 K and 80 K. Note that the molecules move more slowly at the lower temperature. Because these are probability dis­tributions, the area under each curve has a numerical value of unity.

The fraction (frac) of molecules with speeds in an interval of, say, , to is then

centered on .Then we add up all these values of . The result is .

Average, RMS. and Most Probable Speeds

In principle, we can find the average speed of the molecules in a gas with the following procedure: We weight each value of in the distribution; that is, we multiply it by the fraction of molecules with speeds in a differential interval; centered on . Then we add up all these values of .The result is . In practice, we do all this by evaluating

(15)

Substituting for from Eq. 14 and using integral from the list o integrals, we find

,

, Hence

(average speed). (16)

Similarly, we can find the average of the square of the speeds with

(17)

Substituting for from Eq. 14 and using generic integral from the list of integrals, we find

.

The square root of is the root-mean-square speed vms. Thus,

(18)

which agrees with Eq. 12.

The most probable speed is the speed at which is maximum (see Fig. 4a). To calculate , we set (the slope of the curve in Fig 4 is zero at the maximum of the curve) and then solve for . Doing so, we find

(19)

A molecule is more likely to have speed than any other speed, but some molecules will have speeds that are many times . These molecules lie in the high-speed tail of a distribution curve like that in Fig. 4. We should be thankful for these few, higher speed molecules because they make possible both rain and sunshine (without which we could not exist). We next see why.

Rain: The speed distribution of water molecules in, say, a pond at summertime temperatures can be represented by a curve similar to that of Fig. 4a. Most of the molecules do not have nearly enough kinetic energy to escape from the water through its surface. However, small numbers of very fast molecules with speeds far out in the tail of the curve can do so. It is these water molecules that evaporate, making clouds and rains a possibility.

As the fast water molecules leave the surface, carrying energy with them, the temperature of the remaining water is maintained by heat transfer from the surroundings. Other fast molecules - produced in particularly favorable collisions-quickly take the place of those that have left, and the speed distribution is maintained.

Sunshine: Let the distribution curve now refer to protons in the core of the Sun. The Sun's energy is supplied by a nuclear fusion process that starts with the merging of two protons. However, protons repel each other because of their electrical charges, and protons of average speed do not have enough kinetic energy to overcome the repulsion and get close enough to merge. Very fast protons with speeds in the tail of the distribution curve can do so, however, and for that reason the Sun can shine.

Collisions between Molecules, Mean Free Path

Most of us are familiar with the fact that the strong odor associated with a gas such as ammonia may take a fraction of a minute to diffuse throughout a room. However, because average molecular speeds are typically several hundred meters per second at room temperature, we might expect a diffusion time much less than 1 s. But, molecules collide with one other because they are not geometrical points. Therefore, they do not travel from one side of a room to the other in a straight line. Between collisions, the molecules move with constant speed along straight lines. The average distance between collisions is called the mean free path. The path of an individual molecule is random and resembles that shown in Figure 5. As we would expect from this description, the mean free path is related to the diameter of the molecules and the density of the gas.

Figure 5

We now describe how to estimate the mean free path for a gas molecule. For this calculation, we assume that the molecules are spheres of diameter . We see from Figure 5a that no two molecules collide unless their centers are less than a distance apart as they approach each other. An equivalent way to describe the collisions is to imagine that one of the molecules has a diameter and that the rest are

Fig. 6

geometrical points. Let us choose the large molecule to be one moving with the average speed . In a time t, this molecule travels a distance . In this time interval, the molecule sweeps out a cylinder having a cross-sectional area and a length (Fig. 6). Hence, the volume of the cylinder is . If is the number of molecules per unit volume, then the number of point-size molecules in the cylinder is . The molecule of equivalent diameter collides with every molecule in this cylinder in the time t. Hence, the number of collisions in the time t is equal to the number of molecules in the cylinder, .

The mean free path equals the average distance traveled in a time t divided by the number of collisions that occur in that time:

(20)

Because the number of collisions in a time is , the number of collisions per unit time, or collision frequency , is

. (21)

The inverse of the collision frequency is the average time between collisions, known as the mean free time . (22)

Our analysis has assumed that molecules in the cylinder are stationary. When the motion of these molecules is included in the calculation, the correct results are

, (23)

. (24)

The Boltzmann Distribution Law

The Exponential Atmosphere

We begin by considering the distribution of molecules in our atmosphere. Let us determine how the number of molecules per unit volume varies with altitude. Our model assumes that the atmosphere is at a constant temperature T. (This assumption is not entirely correct because the temperature of our atmosphere decreases by about 2°C for every 300-m increase in altitude. However, the model does illustrate the basic features of the distribution.)

According to the ideal gas law, a gas containing N molecules in thermal equilibrium obeys the relationship . It is convenient to rewrite this equation in terms of the number density , which represents the number of molecules per unit volume of gas. This quantity is important because it can vary from one point to another. In fact, our goal is to determine how changes in our atmosphere. We can express the ideal gas law in terms of as . Thus, if the number density is known, we can find the pressure, and vice versa. The pressure in the atmosphere decreases with increasing altitude because a given layer of air must support the weight of all the atmosphere above it - that is, the greater the altitude, the less the weight of the air above that layer, and the lower the pressure. To determine the variation in pressure with altitude, let us consider an atmospheric layer of thickness dy and cross-sectional area , as shown in Figure 21.9. Because the air is in static equilibrium, the magnitude PA of the upward force exerted on the bottom of this layer must exceed the magnitude of the downward force on the top of the layer, , by an amount equal to the weight of a gas molecule in the layer , and if a total number of N molecules are in the layer, then the weight of the layer is given by . Thus, we see that

This expression reduces to

Because and T is assumed to remain constant, we see that . Substituting this result into the previous expression for dP and rearranging terms, we have

Integrating this expression, we find that

(25)

where the constant is the number density at y = 0. This result is known as the Boltzmann distribution law.

According7 to Equation 21.23, the number density decreases exponentially with increasing altitude when the temperature is constant. The number density of our atmosphere at sea level is about n0 = 2.69 X 1025 molecules/m3. Because the pressure is , we see from Equation 21.23 that the pressure of our atmos­phere varies with altitude according to the expression

Law of atmosphere (26)

Соседние файлы в предмете [НЕСОРТИРОВАННОЕ]