Sie sind auf Seite 1von 19

Boyle's law

Boyle's law (sometimes referred to as the Boyle-Mariotte law) is one of many gas laws and a special case of the ideal gas law. Boyle's law describes the inversely proportional relationship between the absolute pressure and volume of a gas, if the temperature is kept constant within aclosed system.[1][2] The law was named after chemist and physicist Robert Boyle, who published the original law in 1662.[3] The law itself can be stated as follows: For a fixed amount of an ideal gas kept at a fixed temperature, P [pressure] and V [volume] are inversely proportional (while one doubles, the other halves). [2]

History
This relationship between pressure and volume was first noted by two amateur scientists, Richard Towneley and Henry Power. Boyle confirmed their discovery through experiments and published the results. According to Robert Gunther and other authorities, it was Boyle's assistant, Robert Hooke, who built the experimental apparatus. Boyle's law is based on experiments with air, which he considered to be a fluid of particles at rest in between small invisible springs. At that time, air was still seen as one of the four elements, but Boyle disagreed. Boyle's interest was probably to understand air as an essential element of life; [4] he published e.g. the growth of plants without air.[5] The French physicist Edme Mariotte (16201684) discovered the same law independently of Boyle in 1676, but Boyle had already published it in 1662. Thus this law may, improperly, be referred to as Mariotte's or the Boyle-Mariotte law. Later, in 1687 in the Philosophi Naturalis Principia Mathematica, Newton showed mathematically that if an elastic fluid consisting of particles at rest, between which are repulsive forces inversely proportional to their distance, the density would be directly proportional to the pressure,[6] but this mathematical treatise is not the physical explanation for the observed relationship. Instead of a static theory a kinetic theory is needed, which was provided two centuries later by Maxwell and Boltzmann. Definition

Relation to kinetic theory and ideal gases Boyles law states that at constant temperature for a fixed mass, the absolute pressure and the volume of a gas are inversely proportional. The law can also be stated in a slightly different manner, that the product of absolute pressure and volume is always constant. Most gases behave like ideal gases at moderate pressures and temperatures. The technology of the 17th century could not produce high pressures or low temperatures. Hence, the law was not likely to have deviations at the time of publication. As improvements in technology permitted higher pressures and lower temperatures, deviations from the ideal gas behavior became noticeable, and the relationship between pressure and volume can only be accurately described employing real gas theory.[7] The deviation is expressed as the compressibility factor. Robert Boyle (and Edme Mariotte) derived the law solely on experimental grounds. The law can also be derived theoretically based on the presumed existence of atoms and molecules and assumptions about motion and perfectly elastic collisions (see kinetic theory of gases). These assumptions were met with enormous resistance in the positivist scientific community at the time however, as they were seen as purely theoretical constructs for which there was not the slightest observational evidence. Daniel Bernoulli in 1737-1738 derived Boyle's law using Newton's laws of motion with application on a molecular level. It remained ignored until around 1845, when John Waterston published a paper building the main precepts of kinetic theory; this was rejected by the Royal Society of England. Later works of James Prescott Joule, Rudolf

Clausius and in particular Ludwig Boltzmann firmly established the kinetic theory of gases and brought attention to both the theories of Bernoulli and Waterston.[8] The debate between proponents of Energetics and Atomism led Boltzmann to write a book in 1898, which endured criticism up to his suicide in 1906.[8] Albert Einstein in 1905 showed how kinetic theory applies to the Brownian motion of a fluid-suspended particle, which was confirmed in 1908 by Jean Perrin.[8] Equation The mathematical equation for Boyle's law is:

where: P denotes the pressure of the system. V denotes the volume of the gas. k is a constant value representative of the pressure and volume of the system. So long as temperature remains constant the same amount of energy given to the system persists throughout its operation and therefore, theoretically, the value of k will remain constant. However, due to the derivation of pressure as perpendicular applied force and the probabilistic likelihood of collisions with other particles through collision theory, the application of force to a surface may not be infinitely constant for such values of k, but will have a limit when differentiating such values over a given time. Forcing the volume V of the fixed quantity of gas to increase, keeping the gas at the initially measured temperature, the pressure p must decrease proportionally. Conversely, reducing the volume of the gas increases the pressure. Boyle's law is used to predict the result of introducing a change, in volume and pressure only, to the initial state of a fixed quantity of gas. The before and after volumes and pressures of the fixed amount of gas, where the before and after temperatures are the same (heating or cooling will be required to meet this condition), are related by the equation:

Boyle's law, Charles's law, and Gay-Lussac's law form the combined gas law. The three gas laws in combination with Avogadro's law can be generalized by the ideal gas law.

Charles's law
Charles's law (also known as the law of volumes) is an experimental gas law which describes how gases tend to expand when heated. It was first published by French natural philosopherJoseph Louis Gay-Lussac in 1802,
[1]

although he credited the discovery to unpublished work from the 1780s by Jacques Charles. The law was

independently discovered by British natural philosopher John Dalton by 1801, although Dalton's description was less thorough than Gay-Lussac's.[2] The basic principles had already been described a century earlier by Guillaume Amontons. Taylor Buchanan was the first to demonstrate that the law applied generally to all gases, and also to the vapours of volatile liquids if the temperature was more than a few degrees above the boiling point.[citation needed] His statement of the law can be expressed mathematically as:

where V100 is the volume occupied by a given sample of gas at 100 C; V0 is the volume occupied by the same sample of gas at 0 C; and k is a constant which is the same for all gases at constant pressure. Gay-Lussac's value for k was 12.6666, remarkably close to the present-day value of 12.7315. A modern statement of Charles' law is: At constant pressure, the volume of a given mass of an ideal gas increases or decreases by the same factor as its temperature on the absolute temperature scale (i.e. the gas expands as the temperature increases).[3] which can be written as:

where V is the volume of the gas; and T is the absolute temperature. The law can also be usefully expressed as follows:

The equation shows that, as absolute temperature increases, the volume of the gas also increases in proportion.

Charles's Law
When a given mass of a gas is heated at constant pressure, the volume V of given mass of a gas is directly proportional to its absolute temperature.

Relation to the ideal gas law


French physicist mile Clapeyron combined Charles's law with Boyle's law in 1834 to produce a single statement which would become known as the ideal gas law.[4] Claypeyron's original statement was:

where t is the Celsius temperature; and p0, V0 and t0 are the pressure, volume and temperature of a sample of gas under some standard state. The figure of 267 came directly from Gay-Lussac's work: the modern figure would be 273.15. For any given sample of gas, p0V0267+t0 is a constant (Clapeyron denoted this constant R, and it is closely related to the modern gas constant); if the pressure is also constant, the equation simplifies to

as required. The modern statement of the ideal gas law is:

where n is the amount of substance of the gas sample; and R is the gas constant. The amount of substance is constant for any given gas sample so, at constant pressure, the equation rearranges to:

where nRp is the constant of proportionality.

An ideal gas is defined as a gas which obeys the ideal gas law, so Charles's law is only expected to be followed exactly by ideal gases. Nevertheless, it is a good approximation to the behaviour of real gases at relatively high temperatures and relatively low pressures.

Relation to absolute zero


Charles' law appears to imply that the volume of a gas will descend to zero at a certain temperature (266.66 C according to Gay-Lussac's figures) or -273C. Gay-Lussac was clear in his description that the law was not applicable at low temperatures: but I may mention that this last conclusion cannot be true except so long as the compressed vapors remain entirely in the elastic state; and this requires that their temperature shall be sufficiently elevated to enable them to resist the pressure which tends to make them assume the liquid state.[1] Gay-Lussac had no experience of liquid air (first prepared in 1877), although he appears to believe (as did Dalton) that the "permanent gases" such as air and hydrogen could be liquified. Gay-Lussac had also worked with the vapours of volatile liquids in demonstrating Charles's law, and was aware that the law does not apply just above the boiling point of the liquid: I may however remark that when the temperature of the ether is only a little above its boiling point, its condensation is a little more rapid than that of atmospheric air. This fact is related to a phenomenon which is exhibited by a great many bodies when passing from the liquid to the solid state, but which is no longer sensible at temperatures a few degrees above that at which the transition occurs.[1] The first mention of a temperature at which the volume of a gas might descend to zero was by William Thomson (later known as Lord Kelvin) in 1848:[5] This is what we might anticipate, when we reflect that infinite cold must correspond to a finite number of degrees of the air-thermometer below zero; since if we push the strict principle of graduation, stated above, sufficiently far, we should arrive at a point corresponding to the volume of air being reduced to nothing, which would be marked as 273 of the scale (100/.366, if .366 be the coefficient of expansion); and therefore 273 of the air-thermometer is a point which cannot be reached at any finite temperature, however low. However, the "absolute zero" on the Kelvin temperature scale was originally defined in terms of the second law of thermodynamics, which Thomson himself described in 1852.[6] Thomson did not assume that this was equal to the "zero-volume point" of Charles's law, merely that Charles's law provided the minimum temperature which could be attained. The two can be shown to be equivalent by Ludwig Boltzmann'sstatistical view of entropy (1870).

Relation to kinetic theory


The kinetic theory of gases relates the macroscopic properties of gases, such as pressure and volume, to the microscopic properties of the molecules which make up the gas, particularly the mass and speed of the molecules. In order to derive Charles's law from kinetic theory, it is necessary to have a microscopic definition of temperature: this can be conveniently taken as the temperature being proportional to the average kinetic energy of the gas molecules, Ek:

Under this definition, the demonstration of Charles's law is almost trivial. The kinetic theory equivalent of the ideal gas law relates pV to the average kinetic energy:

where N is the number of molecules in the gas sample. If the pressure is constant, the volume is directly proportional to the average kinetic energy (and hence to the temperature) for any given gas sample. ... absolute zero is in attainable in gases because most of the gases turn to liquids i.e. they leave the state of gas thus the law is not valid . This is only a theoretical limitation and thus works in practice. Applications of Charles's Law Bursting of hydrogen balloon Making of chappathi

Gay-Lussac's law
The expression Gay-Lussac's law is used for each of the two relationships named after the French chemist Joseph Louis Gay-Lussac and which concern the properties of gases, though it is more usually applied to his law of combining volumes, the first listed here. One law relates to volumes before and after a chemical reaction while the other concerns the pressure and temperature relationship for a sample of gas.

Law of combining volumes


The law of combining volumes states that, when gases react together to form other gases, and all volumes are measured at the same temperature and pressure: The ratio between the volumes of the reactant gases and the products can be expressed in simple whole numbers. For example, Gay-Lussac found that 2 volumes of Hydrogen and 1 volume of Oxygen would react to form 2 volume of gaseous water. In addition to Gay-Lussac's results, Amedeo Avogadro theorized that, at the same temperature and pressure, equal volumes of gas contain equal numbers of molecules (Avogadro's law). This hypothesis meant that the previously stated result 2 volumes of Hydrogen + 1 volume of Oxygen = 2 volume of gaseous water could also be expressed as 2 molecules of Hydrogen + 1 molecule of Oxygen = 2 molecule of water. The law of combining gases was published by Joseph Louis Gay-Lussac in 1808.[1] Avogadro's hypothesis, however, was not initially accepted by chemists until the Italian chemist Stanislao Cannizzaro was able to convince them at the First International Chemical Congress in 1860.[2]

Pressure-temperature law
Gay-Lussac's name is also associated erroneously with another gas law, the so-called pressure law, which states that: The pressure of a gas of fixed mass and fixed volume is directly proportional to the gas' absolute temperature. Simply put, if a gas' temperature increases then so does its pressure, if the mass and volume of the gas are held constant. The law has a particularly simple mathematical form if the temperature is measured on an absolute scale, such as in kelvins. The law can then be expressed mathematically as:

or

where: P is the pressure of the gas (measured in ATM). T is the temperature of the gas (measured in Kelvin). k is a constant. This law holds true because temperature is a measure of the average kinetic energy of a substance; as the kinetic energy of a gas increases, its particles collide with the container walls more rapidly, thereby exerting increased pressure. For comparing the same substance under two different sets of conditions, the law can be written as:

Amontons' Law of Pressure-Temperature: The pressure law described above should actually be attributed to Guillaume Amontons, who, in the late 17th century (more accurately between 1700 and 1702[3][4]), discovered that the pressure of a fixed mass of gas kept at a constant volume is proportional to the temperature. Amontons discovered this while building an "air thermometer". Calling it Gay-Lussac's law is simply incorrect as Gay-Lussac investigated the relationship between volume and temperature (i.e. Charles' Law), not pressure and temperature. Charles' Law was also known as the Law of Charles and Gay-Lussac, because GayLussac published it in 1802 using much of Charles's unpublished data from 1787. However, in recent years the term has fallen out of favor, and Gay-Lussac's name is now generally associated with the law of combining volumes. Amontons' Law, Charles' Law, and Boyle's law form the combined gas law. The three gas laws in combination with Avogadro's Law can be generalized by the ideal gas law.

Avogadro's law
Definition: Under the same condition of temperature and pressure, equal volumes of all gases contain the same number of molecules. Avogadro's law (sometimes referred to as Avogadro's hypothesis or Avogadro's principle) is a gas law named after Amedeo Avogadrowho, in 1811,[1] hypothesized that two given samples of an ideal gas, of the same volume and at the same temperature and pressure, contain the same number of molecules. Thus, the number of molecules or atoms in a specific volume of gas is independent of their size or the molar mass of the gas. As an example, equal volumes of molecular hydrogen and nitrogen contain the same number of molecules when they are at the same temperature and pressure, and observe ideal gas behavior. In practice, real gases show small deviations from the ideal behavior and the law holds only approximately, but is still a useful approximation for scientists.

Mathematical definition Avogadro's law is stated mathematically as:

Where: V is the volume of the gas. n is the amount of substance of the gas. k is a proportionality constant. The most significant consequence of Avogadro's law is that the ideal gas constant has the same value for all gases. This means that:

Where: p is the pressure of the gas T is the temperature in kelvin of the gas

Ideal gas law


A common rearrangement of this equation is by letting R be the proportionality constant, and rearranging as follows: pV = nRT This equation is known as the ideal gas law.

Molar volume
Taking STP to be 101.325 kPa and 273.15 K, we can find the volume of one mole of a gas:

For 100.000 kPa and 273.15 K, the molar volume of an ideal gas is 22.414 dm3mol-1.

Ideal gas law


The ideal gas law is the equation of state of a hypothetical ideal gas. It is a good approximation to the behavior of many gases under many conditions, although it has several limitations. It was first stated by mile Clapeyron in 1834 as a combination of Boyle's law and Charles's law.[1] It can also be derived from kinetic theory, as was achieved (apparently independently) by August Krnig in 1856[2] and Rudolf Clausius in 1857.[3] The state of an amount of gas is determined by its pressure, volume, and temperature. The modern form of the equation is:

where P is the absolute pressure of the gas measured in atmospheres; V is the volume (in this equation the volume is expressed in liters); N is the number of particles in the gas; k isBoltzmann's constant relating temperature and energy; and T is the absolute temperature.

In SI units, P is measured in pascals; V in cubic metres; N is a dimensionless number; and T in kelvin. k has the value 1.381023 JK1 in SI units. Sometimes this is expressed as

where n is the amount of substance of gas (also known as number of moles) and R is the ideal, or universal, gas constant, equal to the product of Boltzmann's constant and Avogadro's constant. In SI units, n is measured in moles, and T in kelvin. R has the value 8.314 JK1mol1or 0.08206 Latmmol-1K. The temperature used in the equation of state is an absolute temperature: in the SI system of units, kelvins; in the Imperial system, degrees Rankine.[4]

Deviations from real gases


The equation of state given here applies only to an ideal gas, or as an approximation to a real gas that behaves sufficiently like an ideal gas. There are in fact many different forms of the equation of state for different gases. Since it neglects both molecular size and intermolecular attractions, the ideal gas law is most accurate for monatomic gases at high temperatures and low pressures. The neglect of molecular size becomes less important for lower densities, i.e. for larger volumes at lower pressures, because the average distance between adjacent molecules becomes much larger than the molecular size. The relative importance of intermolecular attractions diminishes with increasing thermal kinetic energy, i.e., with increasing temperatures. More detailed equations of state, such as the van der Waals equation, allow deviations from ideality caused by molecular size and intermolecular forces to be taken into account. A residual property is defined as the difference between a real gas property and an ideal gas property, both considered at the same pressure, temperature, and composition.

Alternative forms
Molar form As the amount of substance could be given in mass instead of moles, sometimes an alternative form of the ideal gas law is useful. The number of moles (n) is equal to the mass (m) divided by the molar mass (M):

By replacing n, with density = m/V, we get:

Defining the specific gas constant Rspecific as the ratio R/M,

This form of the ideal gas law is very useful because it links pressure, density, and temperature in a unique formula independent of the quantity of the considered gas. Alternatively, the law may be written in terms of the specific volume v, the reciprocal of density, as

It is common, especially in engineering applications, to represent the ideal gas constant by the symbol R. In such cases, the universal gas constant is usually given a different symbol such as R to distinguish it. In any case, the context and/or units of the gas constant should make it clear as to whether the universal or specific gas constant is being referred to.[5] Statistical mechanics In statistical mechanics the following molecular equation is derived from first principles:

Here k is the Boltzmann constant, and N is the actual number of molecules, in contrast to the other formulation, which uses n, the number of moles. This relation implies that Nk = nR, and the consistency of this result with experiment is a good check on the principles of statistical mechanics. From here we can notice that for an average particle mass of times the atomic mass constant mu (i.e., the mass is u)

and since = m/V, we find that the ideal gas law can be rewritten as:

Applications to thermodynamic processes


The table below essentially simplifies the ideal gas equation for a particular processes, thus making this equation easier to solve using numerical methods. A thermodynamic process is defined as a system that moves from state 1 to state 2, where the state number is denoted by subscript. As shown in the first column of the table, basic thermodynamic processes are defined such that one of the gas properties (P, V, T, or S) is constant throughout the process. For a given thermodynamics process, in order to specify the extent of a particular process, one of the properties ratios (listed under the column labeled "known ratio") must be specified (either directly or indirectly). Also, the property for which the ratio is known must be distinct from the property held constant in the previous column (otherwise the ratio would be unity, and not enough information would be available to simplify the gas law equation). In the final three columns, the properties (P, V, or T) at state 2 can be calculated from the properties at state 1 using the equations listed.

Process

Constant

Known ratio

P2

V2

T2

Isobaric process

Pressure

V2/V1

P2 = P1

V2 = V1(V2/V1)

T2 = T1(V2/V1)

T2/T1

P2 = P1

V2 = V1(T2/T1)

T2 = T1(T2/T1)

P2/P1 Isochoric process Volume T2/T1

P2 = P1(P2/P1)

V2 = V1

T2 = T1(P2/P1)

P2 = P1(T2/T1)

V2 = V1

T2 = T1(T2/T1)

P2/P1 Isothermal process Temperature V2/V1

P2 = P1(P2/P1)

V2 = V1/(P2/P1)

T2 = T1

P2 = P1/(V2/V1)

V2 = V1(V2/V1)

T2 = T1

P2/P1 Isentropic process Entropy[a] V2/V1

P2 = P1(P2/P1)

V2 = V1(P2/P1)(1/)

T2 = T1(P2/P1)(1 1/)

(Reversible adiabatic process)

P2 = P1(V2/V1)

V2 = V1(V2/V1)

T2 = T1(V2/V1)(1 )

T2/T1

P2 = P1(T2/T1)/( 1) V2 = V1(T2/T1)1/(1 ) T2 = T1(T2/T1)

P2/P1

P2 = P1(P2/P1)

V2 = V1(P2/P1)(-1/n)

T2 = T1(P2/P1)(1 - 1/n)

Polytropic process

P Vn

V2/V1

P2 = P1(V2/V1)n

V2 = V1(V2/V1)

T2 = T1(V2/V1)(1n)

T2/T1

P2 = P1(T2/T1)n/(n 1) V2 = V1(T2/T1)1/(1 n) T2 = T1(T2/T1)

a. In an isentropic process, system entropy (S) is constant. Under these conditions, P1 V1 = P2 V2, where is

defined as the heat capacity ratio, which is constant for an ideal gas. The value used for is typically 1.4 for diatomic gases like nitrogen(N2) and oxygen (O2), (and air, which is 99% diatomic). Also is typically 1.6 for monatomic gases like the noble gases helium(He), and argon (Ar). In internal combustion engines varies between 1.35 and 1.15, depending on constitution gases and temperature.

Derivations
Empirical The ideal gas law can be derived from combining two empirical gas laws: the combined gas law and Avogadro's law. The combined gas law states that

where C is a constant which is directly proportional to the amount of gas, n (Avogadro's law). The proportionality factor is theuniversal gas constant, R, i.e. C = nR. Hence the ideal gas law

Theoretical The ideal gas law can also be derived from first principles using the kinetic theory of gases, in which several simplifying assumptions are made, chief among which are that the molecules, or atoms, of the gas are point masses, possessing mass but no significant volume, and undergo only elastic collisions with each other and the sides of the container in which both linear momentum and kinetic energy are conserved. From statistical mechanics Let q = (qx, qy, qz) and p = (px, py, pz) denote the position vector and momentum vector of a particle of an ideal gas, respectively. Let F denote the net force on that particle. Then the time average momentum of the particle is:

where the first equality is Newton's second law, and the second line uses Hamilton's equations and the equipartition theorem. Summing over a system of N particles yields

By Newton's third law and the ideal gas assumption, the net force on the system is the force applied by the walls of their container, and this force is given by the pressure P of the gas. Hence

where dS is the infinitesimal area element along the walls of the container. Since the divergence of the position vector q is

the divergence theorem implies that

where dV is an infinitesimal volume within the container and V is the total volume of the container. Putting these equalities together yields

which immediately implies the ideal gas law for N particles:

where n = N/NA is the number of moles of gas and R = NAkB is the gas constant. The readers are referred to the comprehensive article Configuration integral (statistical mechanics) where an alternative statistical mechanics derivation of the ideal-gas law, using the relationship between the Helmholtz free energy and the partition function, but without using the equipartition theorem, is provided.

Graham's law
Graham's law, known as Graham's law of effusion, was formulated by Scottish physical chemist Thomas Graham in 1846. Graham found experimentally that the rate of effusion of a gas is inversely proportional to the square root of the mass of its particles. This formula can be written as:

where: Rate1 is the rate of effusion of the first gas (volume or number of moles per unit time). Rate2 is the rate of effusion for the second gas. M1 is the molar mass of gas 1 M2 is the molar mass of gas 2. Graham's law states that the rate of effusion of a gas is inversely proportional to the square root of its molecular weight. Thus, if the molecular weight of one gas is four times that of another, it would diffuse through a porous plug or escape through a small pinhole in a vessel at half the rate of the other. A complete theoretical explanation of Graham's law was provided years later by the kinetic theory of gases. Graham's law provides a basis for separating isotopes by diffusion a method that came to play a crucial role in the development of the atomic bomb. Graham's law is most accurate for molecular effusion which involves the movement of one gas at a time through a hole. It is only approximate for diffusion of one gas in another or in air, as these processes involve the movement of more than one gas.

History
Graham's research on the diffusion of gases was triggered by his reading about the observation of German chemist Johann Dbereiner that hydrogen gas diffused out of a small crack in a glass bottle faster than the surrounding air diffused in to replace it. Graham measured the rate of diffusion of gases through plaster plugs, through very fine tubes, and through small orifices. In this way he slowed down the process so that it could be studied quantitatively. He first stated the law as we know it today in 1831. Graham went on to study the diffusion of substances in solution and in the process made the discovery that some apparent solutions actually are suspensions of particles too large to pass through a parchment filter. He termed these materials colloids, a term that has come to denote an important class of finely divided materials. At the time Graham did his work the concept of molecular weight was being established, in large part through measurements of gases. Italian physicist Amadeo Avogadro had suggested in 1811 that equal volumes of different

gases contain equal numbers of molecules. Thus, the relative molecular weights of two gases are equal to the ratio of weights of equal volumes of the gases. Avogadro's insight together with other studies of gas behaviour provided a basis for later theoretical work by Scottish physicistJames Clerk Maxwell to explain the properties of gases as collections of small particles moving through largely empty space. Perhaps the greatest success of the kinetic theory of gases, as it came to be called, was the discovery that for gases, the temperature as measured on the Kelvin (absolute) temperature scale is directly proportional to the average kinetic energy of the gas molecules. The kinetic energy of any object is equal to one-half its mass times the square of its velocity. Thus, to have equal kinetic energies, the velocities of two different molecules would have to be in inverse proportion to the square roots of their masses. The rate of effusion is determined by the number of molecules entering an aperture per unit time, and hence by the average molecular velocity. Graham's law for diffusion could thus be understood as a consequence of the molecular kinetic energies being equal at the same temperature.

Example
Let gas 1 be H2 and gas 2 be O2.

Therefore, hydrogen molecules effuse four times faster than those of oxygen. Graham's Law can also be used to find the approximate molecular weight of a gas if one gas is a known species, and if there is a specific ratio between the rates of two gases (such as in the previous example). The equation can be solved for either one of the molecular weights provided the subscripts are consistent.

Graham's law was the basis for separating 235U from 238U found in natural uraninite (uranium ore) during the Manhattan project to build the first atomic bomb. The United States government built a gaseous diffusion plant at the then phenomenal cost of $100 million in Clinton, Tennessee. In this plant, uranium from uranium ore was first converted to uranium hexafluorideand then forced repeatedly to diffuse through porous barriers, each time becoming a little more enriched in the slightly lighter235U isotope.

Dalton's law
In chemistry and physics, Dalton's law (also called Dalton's law of partial pressures) states that the total pressure exerted by a gaseous mixture is equal to the sum of the partial pressures of each individual component in a gas mixture. This empiricallaw was observed by John Dalton in 1801 and is related to the ideal gas laws. Mathematically, the pressure of a mixture of gases can be defined as the summation

or where represent the partial pressure of each component.

It is assumed that the gases do not react with each other.

where

the mole fraction of the i-th component in the total mixture of n components .

The relationship below provides a way to determine the volume based concentration of any individual gaseous component.

where:

is the concentration of the ith component expressed in ppm.

Dalton's law is not exactly followed by real gases. Those deviations are considerably large at high pressures. In such conditions, the volume occupied by the molecules can become significant compared to the free space between them. Moreover, the short average distances between molecules raises the intensity of intermolecular forces between gas molecules enough to substantially change the pressure exerted by them. Neither of those effects are considered by the ideal gas model.

Henry's law
In physics, Henry's law is one of the gas laws formulated by William Henry in 1803. It states that: At a constant temperature, the amount of a given gas that dissolves in a given type and volume of liquid is directly proportional to the partial pressure of that gas in equilibrium with that liquid. An equivalent way of stating the law is that the solubility of a gas in a liquid at a particular temperature is proportional to the pressure of that gas above the liquid. Henry's law has since been shown to apply for a wide range of dilute solutions, not merely those of gases. An everyday example of Henry's law is given by carbonated soft drinks. Before the bottle or can is opened, the gas above the drink is almost pure carbon dioxide at a pressure slightly higher than atmospheric pressure. The drink itself contains dissolved carbon dioxide. When the bottle or can is opened, some of this gas escapes, giving the characteristic hiss (or "pop" in the case of a sparkling wine bottle). Because the pressure above the liquid is now lower, some of the dissolved carbon dioxide comes out of solution as bubbles. If a glass of the drink is left in the open, the concentration of carbon dioxide in solution will come into equilibrium with the carbon dioxide in the air, and the drink will go "flat". Note that the pressure acting above the drink in the sealed container must come from the partial pressure of carbon dioxide. If the gas is only air it would not produce the same effect even if the pressure value is the same. A slightly more exotic example of Henry's law is in decompression and decompression sickness of divers. Formula and the Henry's law constant Henry's law can be put into mathematical terms (at constant temperature) as

where p is the partial pressure of the solute in the gas above the solution, c is the concentration of the solute and kH is a constant with the dimensions of pressure divided by concentration.[1] The constant, known as the Henry's law constant, depends on the solute, the solvent and the temperature. Some values for kH for gases dissolved in water at 298 K include:

oxygen (O2) : 769.2 Latm/mol carbon dioxide (CO2) : 29.41 Latm/mol hydrogen (H2) : 1282.1 Latm/mol There are other forms of Henry's Law, each of which defines the constant kH differently and requires different dimensional units.[2] In particular, the "concentration" of the solute in solution may also be expressed as a mole fraction or as a molality.[1] Other forms of Henry's law There are various other forms of Henry's Law which are discussed in the technical literature.[2][3][4]

Table 1: Some forms of Henry's law and constants (gases in water at 298.15 K), derived from [4]

equation:

units:

dimensionless

O2

769.23

1.3103

4.259104

3.181102

H2 CO2 N2 He Ne Ar CO where:

1282.05 29.41 1639.34 2702.7 2222.22 714.28 1052.63

7.8104 3.4102 6.1104 3.7104 4.5104 1.4103 9.5104

7.099104 0.163104 9.077104 14.97104 12.30104 3.955104 5.828104

1.907102 0.8317 1.492102 9.051103 1.101102 3.425102 2.324102

c = amount concentration of gas in solution (in mol/L) p = partial pressure of gas above the solution (in atm) x = mole fraction of gas in solution (dimensionless)

As can be seen by comparing the equations in the above table, the Henry's law constant kH,pc is simply the inverse of the constant kH,cp. Since all kH may be referred to as Henry's law constants, readers of the technical literature must be quite careful to note which version of the Henry's Law equation is being used.[2] It should also be noted the Henry's Law is a limiting law that only applies for sufficiently dilute solutions. The range of concentrations in which it applies becomes narrower the more the system diverges from ideal behavior. Roughly speaking, that is the more chemically different the solute is from the solvent. It also only applies simply for solutions where the solvent does not react chemically with the gas being dissolved. A common example of a gas that does react with the solvent is carbon dioxide, which forms carbonic acid (H2CO3) to a certain degree with water. Temperature dependence of the Henry constant When the temperature of a system changes, the Henry constant will also change. [2] This is why some people prefer to name it Henry coefficient. There are multiple equations assessing the effect of temperature on the constant. These forms of the van 't Hoff equation are examples:[4]

where kH for a given temperature is the Henry's Law constant (as defined in the first section of this article). Notice that the correct sign of C depends on whether kH,pc or kH,cp is used. T is the thermodynamic temperature i.e. value in K], T o refers to the standard temperature (298 K). This equation is only an approximation, and should be used only when no better, experimentally-derived formula is known for a given gas. The following table lists some values for constant C (in kelvins) in the equation above:

Table 2: Values of C

Gas O2

H2 CO2 N2

He Ne Ar

CO

C(K) 1700 500 2400 1300 230 490 1300 1300

Because solubility of permanent gases usually decreases with increasing temperature at around the room temperature, the partial pressure a given gas concentration has in liquid must increase. While heating water (saturated with nitrogen) from 25 C to 95 C the solubility will decrease to about 43% of its initial value. This can be verified when heating water in a pot: small bubbles evolve and rise, long before the water reaches boiling temperature. Similarly, carbon dioxide from a carbonated drink escapes much faster when the drink is not cooled because the required partial pressure of CO2 to achieve the same solubility increases in higher temperatures. Partial pressure of CO2 in the gas phase in equilibrium with seawater doubles with every 16 K increase in temperature.[5] The constant C may be regarded as:

where solvH is the enthalpy of solution R is the gas constant. The solubility of gases does not always decrease with increasing temperature. For aqueous solutions, the Henry-law constant usually goes through a maximum (i.e., the solubility goes through a minimum). For most permanent gases, the minimum is below 120 C. It is often observed that the smaller the gas molecule (and the lower the gas solubility in water), then the lower the temperature of the maximum of the Henry-law constant. Thus, the maximum is about 30 C for helium, 92 to 93 C for argon, nitrogen and oxygen, and 114 C for xenon.[6]

In geophysics
In geophysics, a version of Henry's law applies to the solubility of a noble gas in contact with silicate melt. One equation used is

where: C = the number concentrations of the solute gas in the melt and gas phases = 1/kBT, an inverse temperature scale: kB = the Boltzmann constant E = the excess chemical potentials of the solute gas in the two phases.

Comparison to Raoult's law


For a dilute solution, the concentration of the solute is approximately proportional to its mole fraction x, and Henry's law can be written as:

This can be compared with Raoult's law:

where p* is the vapor pressure of the pure component. At first sight, Raoult's law appears to be a special case of Henry's law where kH = p*. This is true for pairs of closely related substances, such as benzene and toluene, which obey Raoult's law over the entire composition range: such mixtures are called "ideal mixtures". The general case is that both laws are limit laws, and they apply at opposite ends of the composition range. The vapor pressure of the component in large excess, such as the solvent for a dilute solution, is proportional to its mole fraction, and the constant of proportionality is the vapor pressure of the pure substance (Raoult's law). The vapor pressure of the solute is also proportional to the solute's mole fraction, but the constant of proportionality is different and must be determined experimentally (Henry's law). In mathematical terms:

Raoult's law: Henry's law: Raoult's law can also be related to non-gas solutes.

Standard chemical potential


Henry's law has been shown to apply to a wide range of solutes in the limit of "infinite dilution" (x0), including non-volatile substances such as sucrose or even sodium chloride. In these cases, it is necessary to state the law in terms of chemical potentials. For a solute in an ideal dilute solution, the chemical potential depends on the concentration:

, where

for a volatile solute; co = 1 mol/L.

For non-ideal solutions, the activity coefficient c depends on the concentration and must be determined at the concentration of interest. The activity coefficient can also be obtained for non-volatile solutes, where the vapor pressure of the pure substance is negligible, by using the GibbsDuhem relation:

By measuring the change in vapor pressure (and hence chemical potential) of the solvent, the chemical potential of the solute can be deduced. The standard state for a dilute solution is also defined in terms of infinite-dilution behavior. Although the standard concentration co is taken to be 1 mol/L by convention, the standard state is a hypothetical solution of 1 mol/L in which the solute has its limiting infinite-dilution properties. This has the effect that all nonideal behavior is described by the activity coefficient: the activity coefficient at 1 mol/L is not necessarily unity (and is frequently quite different from unity). All the relations above can also be expressed in terms of molalities b rather than concentrations, e.g.:

, where

for a volatile solute; bo = 1 mol/kg.

The standard chemical potential mo, the activity coefficient m and the Henry's law constant kH,b all have different numerical values when molalities are used in place of concentrations.

Das könnte Ihnen auch gefallen