ART

.


The kinetic theory describes a gas as a large number of submicroscopic particles (atoms or molecules), all of which are in constant rapid motion that has randomness arising from their many collisions with each other and with the walls of the container.

Translational motion

The temperature of an ideal monatomic gas is proportional to the average kinetic energy of its atoms. The size of helium atoms relative to their spacing is shown to scale under 1950 atmospheres of pressure. The atoms have a certain, average speed, slowed down here two trillion fold from room temperature.

Kinetic theory explains macroscopic properties of gases, such as pressure, temperature, viscosity, thermal conductivity, and volume, by considering their molecular composition and motion. The theory posits that gas pressure is due to the impacts, on the walls of a container, of molecules or atoms moving at different velocities.

Kinetic theory defines temperature in its own way, not identical with the thermodynamic definition.[1]

Under a microscope, the molecules making up a liquid are too small to be visible, but the jittery motion of pollen grains or dust particles can be seen. Known as Brownian motion, it results directly from collisions between the grains or particles and liquid molecules. As analyzed by Albert Einstein in 1905, this experimental evidence for kinetic theory is generally seen as having confirmed the concrete material existence of atoms and molecules.

Assumptions

The theory for ideal gases makes the following assumptions:

The gas consists of very small particles known as molecules. This smallness of their size is such that the total volume of the individual gas molecules added up is negligible compared to the volume of the smallest open ball containing all the molecules. This is equivalent to stating that the average distance separating the gas particles is large compared to their size.
These particles have the same mass.
The number of molecules is so large that statistical treatment can be applied.
These molecules are in constant, random, and rapid motion.
The rapidly moving particles constantly collide among themselves and with the walls of the container. All these collisions are perfectly elastic. This means, the molecules are considered to be perfectly spherical in shape, and elastic in nature.
Except during collisions, the interactions among molecules are negligible. (That is, they exert no forces on one another.)

This implies:

1. Relativistic effects are negligible.
2. Quantum-mechanical effects are negligible. This means that the inter-particle distance is much larger than the thermal de Broglie wavelength and the molecules are treated as classical objects.
3. Because of the above two, their dynamics can be treated classically. This means that the equations of motion of the molecules are time-reversible.

The average kinetic energy of the gas particles depends only on the absolute temperature of the system. The kinetic theory has its own definition of temperature, not identical with the thermodynamic definition.
The time during collision of molecule with the container's wall is negligible as compared to the time between successive collisions.
Because they have mass, the gas molecules will be affected by gravity.

More modern developments relax these assumptions and are based on the Boltzmann equation. These can accurately describe the properties of dense gases, because they include the volume of the molecules. The necessary assumptions are the absence of quantum effects, molecular chaos and small gradients in bulk properties. Expansions to higher orders in the density are known as virial expansions.

An important book on kinetic theory is that by Chapman and Cowling.[1] An important approach to the subject is called Chapman–Enskog theory.[2] There have been many modern developments and there is an alternative approach developed by Grad based on moment expansions.[3] In the other limit, for extremely rarefied gases, the gradients in bulk properties are not small compared to the mean free paths. This is known as the Knudsen regime and expansions can be performed in the Knudsen number.
Properties
Pressure and kinetic energy

Pressure is explained by the kinetic theory as arising from the force exerted by molecules or atoms when they hit the walls of a container. Consider a gas of N molecules, each of mass m, enclosed in a cuboid of volume V=L3. When a gas molecule collides with the wall of the container perpendicular to the x axis and bounces off in the opposite direction with the same speed (an elastic collision), the change in momentum is given by:

\( {\displaystyle \Delta p=p_{i,x}-p_{f,x}=p_{i,x}-(-p_{i,x})=2p_{i,x}=2mv_{x}\,} \)

where p is the momentum, i and f indicate initial and final momentum (before and after collision), x indicates only the x-direction is being considered and v is the speed of the particle (which is the same before and after the collision).

If one considers the average speed of the particle in the container is v ¯ x {\displaystyle {\bar {v}}_{x}} , one can then consider the average change in momentum due to the collision of n particles:

\( {\displaystyle \Delta {\bar {p}}_{tot}=n\times 2m{\bar {v}}_{x}} \)

The force on the wall is given by:

\( {\displaystyle {\bar {F}}={\frac {\Delta {\bar {p}}_{tot}}{\Delta t}}} \)

or what is the same: the average change in momentum per second.

The question is then to calculate how many particles n will collide with the wall during a one-second interval (on average).

Since the particle travels \( {\displaystyle v_{x}} \) metres per second, if the particle is at less than \( {\displaystyle v_{x}} \) metres away from the wall, it will hit it before a second has elapsed. Hence, any particle in a volume \( {\displaystyle {\bar {v}}_{x}\times L^{2}}\) (where \( {\displaystyle L^{2}} \) is the area of the wall) will hit the end of the container. However, for this calculation only one of the ends of the container is of interest (not the two walls perpendicular to x).

Half the particles will be travelling in the positive x direction and the other half will be travelling in the negative x direction such that only half the particles in the volume defined will hit the wall of interest ( a factor of a half is thus required). Now, to know how many particles are in that volume we need to multiply by the density :

\( {\displaystyle n={\frac {1}{2}}\times {\bar {v}}_{x}\times L^{2}\times {\frac {N}{L^{3}}}} \)

\( {\displaystyle n={\frac {{\bar {v}}_{x}N}{2L}}} \)

The total force on the wall is then

\( {\displaystyle F={\frac {{\bar {v}}_{x}N}{2L}}\times 2m{\bar {v}}_{x}} \)

\( {\displaystyle F={\frac {Nm{\overline {v}}_{x}^{2}}{L}}} \)

Since the motion of the particles is random and there is no bias applied in any direction, the average speed in each direction is identical:

\( {\displaystyle {\bar {v}}_{x}^{2}={\bar {v}}_{y}^{2}={\bar {v}}_{z}^{2}} \)

The total speed v is given by:

\( {\displaystyle {\bar {v}}^{2}={\bar {v}}_{x}^{2}+{\bar {v}}_{y}^{2}+{\bar {v}}_{z}^{2}} \)

\( {\displaystyle {\bar {v}}^{2}=3{\bar {v}}_{x}^{2}} \)

Therefore:

\( {\displaystyle {\bar {v}}_{x}^{2}={\frac {{\bar {v}}^{2}}{3}}} \)

and the force can be written as:

\( {\displaystyle F={\frac {Nm{\overline {v^{2}}}}{3L}}.} \)

This force is exerted on an area L2. Therefore, the pressure of the gas is

\( {\displaystyle P={\frac {F}{L^{2}}}={\frac {Nm{\overline {v^{2}}}}{3V}}} \)

where V=L3 is the volume of the box.

In terms of the kinetic energy K.E.:

\( {\displaystyle P={\frac {2}{3}}\times {\frac {N}{V}}\times {K.E.}} \)

This is a first non-trivial result of the kinetic theory because it relates pressure, a macroscopic property, to the average (translational) kinetic energy per molecule \( {\displaystyle {1 \over 2}m{\overline {v^{2}}} \) which is a microscopic property.
Temperature and kinetic energy

Rewriting the above result for the pressure as P V = N m v 2 ¯ 3 {\displaystyle PV={Nm{\overline {v^{2}}} \over 3}} , we may combine it with the ideal gas law

\( {\displaystyle \displaystyle PV=Nk_{B}T,} \)

(1)

where \( {\displaystyle \displaystyle k_{B}} \)is the Boltzmann constant and \( {\displaystyle \displaystyle T} \) the absolute temperature defined by the ideal gas law, to obtain

\( {\displaystyle k_{B}T={m{\overline {v^{2}}} \over 3}} \),

which leads to the expression of the average kinetic energy of a molecule,

\( {\displaystyle \displaystyle {\frac {1}{2}}m{\overline {v^{2}}}={\frac {3}{2}}k_{B}T} \).

The kinetic energy of the system is N times that of a molecule, namely \( {\displaystyle K={\frac {1}{2}}Nm{\overline {v^{2}}}} \) . Then the temperature \( {\displaystyle \displaystyle T}\) takes the form

T = m v 2 ¯ 3 k B {\displaystyle \displaystyle T={m{\overline {v^{2}}} \over 3k_{B}}}
(2)

which becomes

\( {\displaystyle \displaystyle T={\frac {2}{3}}{\frac {K}{Nk_{B}}}.} \)

(3)

Eq.(3) is one important result of the kinetic theory: The average molecular kinetic energy is proportional to the ideal gas law's absolute temperature. From Eq.(1) and Eq.(3), we have

\( {\displaystyle \displaystyle PV={\frac {2}{3}}K.}

(4)

Thus, the product of pressure and volume per mole is proportional to the average (translational) molecular kinetic energy.

Eq.(1) and Eq.(4) are called the "classical results", which could also be derived from statistical mechanics; for more details, see:[4]

Since there are \( {\displaystyle \displaystyle 3N} \) degrees of freedom in a monatomic-gas system with \( {\displaystyle \displaystyle N} \) particles, the kinetic energy per degree of freedom per molecule is

\( {\displaystyle \displaystyle {\frac {K}{3N}}={\frac {k_{B}T}{2}}} \)

(5)

In the kinetic energy per degree of freedom, the constant of proportionality of temperature is 1/2 times Boltzmann constant. In addition to this, the temperature will decrease when the pressure drops to a certain point. This result is related to the equipartition theorem.

As noted in the article on heat capacity, diatomic gases should have 7 degrees of freedom, but the lighter gases act as if they have only 5.

Thus the kinetic energy per kelvin (monatomic ideal gas) is:

per mole: 12.47 J
per molecule: 20.7 yJ = 129 μeV.

At standard temperature (273.15 K), we get:

per mole: 3406 J
per molecule: 5.65 zJ = 35.2 meV.....

Collisions with container

One can calculate the number of atomic or molecular collisions with a wall of a container per unit area per unit time.

Assuming an ideal gas, a derivation[5] results in an equation for total number of collisions per unit time per area:

\( {\displaystyle A={\frac {1}{4}}{\frac {N}{V}}v_{avg}={\frac {n}{4}}{\sqrt {\frac {8k_{B}T}{\pi m}}}.\,} \)

This quantity is also known as the "impingement rate" in vacuum physics.
Speed of molecules

From the kinetic energy formula it can be shown that

\( {\displaystyle v_{\mathrm {rms} }={\sqrt {{3k_{B}T} \over {m}}}} \)

with v in m/s, T in kelvins, and m is the molecular mass (kg). The most probable speed is 81.6% of the rms speed, and the mean speeds 92.1% (isotropic distribution of speeds).

See: Root-mean-square speed
Transport properties

The kinetic theory of gases deals not only with gases in thermodynamic equilibrium, but also very importantly with gases not in thermodynamic equilibrium. This means considering what are known as 'transport properties', such a viscosity and thermal conductivity.
History

In approximately 50 BCE, the Roman philosopher Lucretius proposed that apparently static macroscopic bodies were composed on a small scale of rapidly moving atoms all bouncing off each other.[6] This Epicurean atomistic point of view was rarely considered in the subsequent centuries, when Aristotlean ideas were dominant.
Hydrodynamica front cover

In 1738 Daniel Bernoulli published Hydrodynamica, which laid the basis for the kinetic theory of gases. In this work, Bernoulli posited the argument, still used to this day, that gases consist of great numbers of molecules moving in all directions, that their impact on a surface causes the gas pressure that we feel, and that what we experience as heat is simply the kinetic energy of their motion. The theory was not immediately accepted, in part because conservation of energy had not yet been established, and it was not obvious to physicists how the collisions between molecules could be perfectly elastic.[7]:36–37

Other pioneers of the kinetic theory (which were neglected by their contemporaries) were Mikhail Lomonosov (1747),[8] Georges-Louis Le Sage (ca. 1780, published 1818),[9] John Herapath (1816)[10] and John James Waterston (1843),[11] which connected their research with the development of mechanical explanations of gravitation. In 1856 August Krönig (probably after reading a paper of Waterston) created a simple gas-kinetic model, which only considered the translational motion of the particles.[12]

In 1857 Rudolf Clausius, according to his own words independently of Krönig, developed a similar, but much more sophisticated version of the theory which included translational and contrary to Krönig also rotational and vibrational molecular motions. In this same work he introduced the concept of mean free path of a particle. [13] In 1859, after reading a paper by Clausius, James Clerk Maxwell formulated the Maxwell distribution of molecular velocities, which gave the proportion of molecules having a certain velocity in a specific range. This was the first-ever statistical law in physics.[14] In his 1873 thirteen page article 'Molecules', Maxwell states: "we are told that an 'atom' is a material point, invested and surrounded by 'potential forces' and that when 'flying molecules' strike against a solid body in constant succession it causes what is called pressure of air and other gases."[15] In 1871, Ludwig Boltzmann generalized Maxwell's achievement and formulated the Maxwell–Boltzmann distribution. Also the logarithmic connection between entropy and probability was first stated by him.

In the beginning of the twentieth century, however, atoms were considered by many physicists to be purely hypothetical constructs, rather than real objects. An important turning point was Albert Einstein's (1905)[16] and Marian Smoluchowski's (1906)[17] papers on Brownian motion, which succeeded in making certain accurate quantitative predictions based on the kinetic theory.
See also
Statistical mechanics
Increasing disorder.svg

Thermodynamics
Kinetic theory

Particle Statistics

Bogoliubov-Born-Green-Kirkwood-Yvon hierarchy of equations
Boltzmann equation
Collision theory
Critical temperature
Gas laws
Heat
Magnetohydrodynamics
Maxwell–Boltzmann distribution
Mixmaster dynamics
Thermodynamics
Vlasov equation

Notes

Chapman, S., Cowling, T.G. (1939/1970).
Kauzmann, W. (1966). Kinetic Theory of Gases, W.A. Benjamin, New York, pp. 232–235.
Grad 1949
Configuration integral (statistical mechanics)
Collisions With a Surface
Maxwell, J. C. (1867). "On the Dynamical Theory of Gases". Philosophical Transactions of the Royal Society of London 157: 49. doi:10.1098/rstl.1867.0004.
L.I Ponomarev; I.V Kurchatov (1 January 1993). The Quantum Dice. CRC Press. ISBN 978-0-7503-0251-7.
Lomonosov 1758
Le Sage 1780/1818
Herapath 1816, 1821
Waterston 1843
Krönig 1856
Clausius 1857
Mahon 2003
Maxwell 1875
Einstein 1905

Smoluchowski 1906

References

Clausius, R. (1857), "Ueber die Art der Bewegung, welche wir Wärme nennen", Annalen der Physik 176 (3): 353–379, Bibcode:1857AnP...176..353C, doi:10.1002/andp.18571760302

de Groot, S. R., W. A. van Leeuwen and Ch. G. van Weert (1980), Relativistic Kinetic Theory, North-Holland, Amsterdam.
Einstein, A. (1905), "Über die von der molekularkinetischen Theorie der Wärme geforderte Bewegung von in ruhenden Flüssigkeiten suspendierten Teilchen" (PDF), Annalen der Physik 17 (8): 549–560, Bibcode:1905AnP...322..549E, doi:10.1002/andp.19053220806

Grad, Harold (1949), "On the Kinetic Theory of Rarefied Gases.", Communications on Pure and Applied Mathematics 2 (4): 331–407, doi:10.1002/cpa.3160020403

Herapath, J. (1816), "On the physical properties of gases", Annals of Philosophy (Robert Baldwin): 56–60

Herapath, J. (1821), "On the Causes, Laws and Phenomena of Heat, Gases, Gravitation", Annals of Philosophy (Baldwin, Cradock, and Joy) 9: 273–293

Krönig, A. (1856), "Grundzüge einer Theorie der Gase", Annalen der Physik 99 (10): 315–322, Bibcode:1856AnP...175..315K, doi:10.1002/andp.18561751008

Le Sage, G.-L. (1818), "Physique Mécanique des Georges-Louis Le Sage", in Prévost, Pierre, Deux Traites de Physique Mécanique, Geneva & Paris: J.J. Paschoud, pp. 1–186

Liboff, R. L. (1990), Kinetic Theory, Prentice-Hall, Englewood Cliffs, N. J.
Lomonosov, M. (1970) [1758], "On the Relation of the Amount of Material and Weight", in Henry M. Leicester, Mikhail Vasil'evich Lomonosov on the Corpuscular Theory, Cambridge: Harvard University Press, pp. 224–233

Mahon, Basil (2003), The Man Who Changed Everything – the Life of James Clerk Maxwell, Hoboken, New Jersey: Wiley, ISBN 0-470-86171-1

Maxwell, James Clerk (1873), "Molecules", Nature 417 (6892): 903, Bibcode:2002Natur.417..903M, doi:10.1038/417903a, PMID 12087385, archived from the original (– Scholar search) on February 9, 2007
Smoluchowski, M. (1906), "Zur kinetischen Theorie der Brownschen Molekularbewegung und der Suspensionen", Annalen der Physik 21 (14): 756–780, Bibcode:1906AnP...326..756V, doi:10.1002/andp.19063261405

Waterston, John James (1843), Thoughts on the Mental Functions (reprinted in his Papers, 3, 167, 183.)

Williams, M. M. R. (1971), Mathematical Methods in Particle Transport Theory, Butterworths, London.

Further reading

Sydney Chapman and T.G. Cowling (1939/1970). The Mathematical Theory of Non-uniform Gases: An Account of the Kinetic Theory of Viscosity, Thermal Conduction and Diffusion in Gases, (first edition 1939, second edition 1952), third edition 1970 prepared in co-operation with D. Burnett, Cambridge University Press, London.
J.O. Hirschfelder, C.F. Curtiss, and R.B. Bird (1964). Molecular Theory of Gases and Liquids, second edition (Wiley).
R.L. Liboff (2003). Kinetic Theory: Classical, Quantum, and Relativistic Descriptions, third edition (Springer).

External links

Early Theories of Gases
Thermodynamics - a chapter from an online textbook
Temperature and Pressure of an Ideal Gas: The Equation of State on Project PHYSNET.
Introduction to the kinetic molecular theory of gases, from The Upper Canada District School Board
Java animation illustrating the kinetic theory from University of Arkansas
Flowchart linking together kinetic theory concepts, from HyperPhysics
Interactive Java Applets allowing high school students to experiment and discover how various factors affect rates of chemical reactions.
https://www.youtube.com/watch?v=47bF13o8pb8&list=UUXrJjdDeqLgGjJbP1sMnH8A A demonstration apparatus for the thermal agitation in gases.

Physics Encyclopedia

Retrieved from "http://en.wikipedia.org/"
All text is available under the terms of the GNU Free Documentation License

Home - Hellenica World