Heat capacity ratio

From Wikipedia, the free encyclopedia
Heat capacity ratio for various gases[1][2]
Temp. Gas γ   Temp. Gas γ   Temp. Gas γ
−181 °C H2 1.597 200 °C Dry air 1.398 20 °C NO 1.400
−76 °C 1.453 400 °C 1.393 20 °C N2O 1.310
20 °C 1.410 1000 °C 1.365 −181 °C N2 1.470
100 °C 1.404 15 °C 1.404
400 °C 1.387 0 °C CO2 1.310 20 °C Cl2 1.340
1000 °C 1.358 20 °C 1.300 −115 °C CH4 1.410
2000 °C 1.318 100 °C 1.281 −74 °C 1.350
20 °C He 1.660 400 °C 1.235 20 °C 1.320
20 °C H2O 1.330 1000 °C 1.195 15 °C NH3 1.310
100 °C 1.324 20 °C CO 1.400 19 °C Ne 1.640
200 °C 1.310 −181 °C O2 1.450 19 °C Xe 1.660
−180 °C Ar 1.760 −76 °C 1.415 19 °C Kr 1.680
20 °C 1.670 20 °C 1.400 15 °C SO2 1.290
0 °C Dry air 1.403 100 °C 1.399 360 °C Hg 1.670
20 °C 1.400 200 °C 1.397 15 °C C2H6 1.220
100 °C 1.401 400 °C 1.394 16 °C C3H8 1.130

In thermal physics and thermodynamics, the heat capacity ratio, also known as the adiabatic index, the ratio of specific heats, or Laplace's coefficient, is the ratio of the heat capacity at constant pressure (CP) to heat capacity at constant volume (CV). It is sometimes also known as the isentropic expansion factor and is denoted by γ (gamma) for an ideal gas[note 1] or κ (kappa), the isentropic exponent for a real gas. The symbol γ is used by aerospace and chemical engineers.

where C is the heat capacity, the molar heat capacity (heat capacity per mole), and c the specific heat capacity (heat capacity per unit mass) of a gas. The suffixes P and V refer to constant-pressure and constant-volume conditions respectively.

The heat capacity ratio is important for its applications in thermodynamical reversible processes, especially involving ideal gases; the speed of sound depends on this factor.

To understand this relation, consider the following thought experiment. A closed pneumatic cylinder contains air. The piston is locked. The pressure inside is equal to atmospheric pressure. This cylinder is heated to a certain target temperature. Since the piston cannot move, the volume is constant. The temperature and pressure will rise. When the target temperature is reached, the heating is stopped. The amount of energy added equals CV ΔT, with ΔT representing the change in temperature. The piston is now freed and moves outwards, stopping as the pressure inside the chamber reaches atmospheric pressure. We assume the expansion occurs without exchange of heat (adiabatic expansion). Doing this work, air inside the cylinder will cool to below the target temperature. To return to the target temperature (still with a free piston), the air must be heated, but is no longer under constant volume, since the piston is free to move as the gas is reheated. This extra heat amounts to about 40% more than the previous amount added. In this example, the amount of heat added with a locked piston is proportional to CV, whereas the total amount of heat added is proportional to CP. Therefore, the heat capacity ratio in this example is 1.4.

Another way of understanding the difference between CP and CV is that CP applies if work is done to the system, which causes a change in volume (such as by moving a piston so as to compress the contents of a cylinder), or if work is done by the system, which changes its temperature (such as heating the gas in a cylinder to cause a piston to move). CV applies only if , that is, no work is done. Consider the difference between adding heat to the gas with a locked piston and adding heat with a piston free to move, so that pressure remains constant. In the second case, the gas will both heat and expand, causing the piston to do mechanical work on the atmosphere. The heat that is added to the gas goes only partly into heating the gas, while the rest is transformed into the mechanical work performed by the piston. In the first, constant-volume case (locked piston), there is no external motion, and thus no mechanical work is done on the atmosphere; CV is used. In the second case, additional work is done as the volume changes, so the amount of heat required to raise the gas temperature (the specific heat capacity) is higher for this constant-pressure case.

Ideal-gas relations[]

For an ideal gas, the heat capacity is constant with temperature. Accordingly, we can express the enthalpy as H = CPT and the internal energy as U = CVT. Thus, it can also be said that the heat capacity ratio is the ratio between the enthalpy to the internal energy:

Furthermore, the heat capacities can be expressed in terms of heat capacity ratio (γ) and the gas constant (R):

where n is the amount of substance in moles.

Mayer's relation allows to deduce the value of CV from the more commonly tabulated value of CP:

Relation with degrees of freedom[]

The heat capacity ratio (γ) for an ideal gas can be related to the degrees of freedom (f) of a molecule by

Thus we observe that for a monatomic gas, with 3 degrees of freedom:

while for a diatomic gas, with 5 degrees of freedom (at room temperature: 3 translational and 2 rotational degrees of freedom; the vibrational degree of freedom is not involved, except at high temperatures):

For example, the terrestrial air is primarily made up of diatomic gases (around 78% nitrogen, N2, and 21% oxygen, O2), and at standard conditions it can be considered to be an ideal gas. The above value of 1.4 is highly consistent with the measured adiabatic indices for dry air within a temperature range of 0–200 °C, exhibiting a deviation of only 0.2% (see tabulation above).

For a non-colinear triatomic gas as water vapor having 6 degrees of freedom :

For a co-linear triatomic molecule such as CO2, there are only 5 degrees of freedom, assuming vibrational modes are not excited. In general however, as mass increases and the frequency of vibrational modes decreases, vibrational degrees of freedom start to enter into the equation at far lower temperatures. For example, it requires a far larger temperature to excite vibrational modes for H2, for which one quantum of vibration is a much larger energy, than for CO2.

Real-gas relations[]

As temperature increases, higher-energy rotational and vibrational states become accessible to molecular gases, thus increasing the number of degrees of freedom and lowering γ. For a real gas, both CP and CV increase with increasing temperature, while continuing to differ from each other by a fixed constant (as above, CP = CV + nR), which reflects the relatively constant PV difference in work done during expansion for constant pressure vs. constant volume conditions. Thus, the ratio of the two values, γ, decreases with increasing temperature. For more information on mechanisms for storing heat in gases, see the gas section of specific heat capacity. While at 273 K (0 °C), Monatomic gases such as the noble gases He, Ne, and Ar all have the same value of γ, that being 1.664.

Thermodynamic expressions[]

Values based on approximations (particularly CPCV = nR) are in many cases not sufficiently accurate for practical engineering calculations, such as flow rates through pipes and valves. An experimental value should be used rather than one based on this approximation, where possible. A rigorous value for the ratio CP/CV can also be calculated by determining CV from the residual properties expressed as

Values for CP are readily available and recorded, but values for CV need to be determined via relations such as these. See relations between specific heats for the derivation of the thermodynamic relations between the heat capacities.

The above definition is the approach used to develop rigorous expressions from equations of state (such as Peng–Robinson), which match experimental values so closely that there is little need to develop a database of ratios or CV values. Values can also be determined through finite-difference approximation.

Adiabatic process[]

This ratio gives the important relation for an isentropic (quasistatic, reversible, adiabatic process) process of a simple compressible calorically-perfect ideal gas:

is constant

Using the ideal gas law, :

is constant
is constant

where P is pressure in Pa, V is the volume of the gas in and T is the temperature in K.

In gas dynamics we are interested in the local relations between pressure, density and temperature, rather than considering a fixed quantity of gas. By considering the density as the inverse of the volume for a unit mass, we can take in these relations. Since for constant entropy, , we have , or , it follows that

For an imperfect or non-ideal gas, Chandrasekhar[3] defined three different adiabatic indices so that the adiabatic relations can be written in the same form as above; these are used in the theory of stellar structure:

All of these are equal to in the case of a perfect gas.

See also[]

References[]

  1. ^ White, Frank M. (October 1998). Fluid Mechanics (4th ed.). New York: McGraw Hill. ISBN 978-0-07-228192-7.
  2. ^ Lange, Norbert A. (1967). Lange's Handbook of Chemistry (10th ed.). New York: McGraw Hill. p. 1524. ISBN 978-0-07-036261-1.
  3. ^ Chandrasekhar, S. (1939). An Introduction to the Study of Stellar Structure. Chicago: University of Chicago Press. p. 56. ISBN 978-0-486-60413-8.

Notes[]

  1. ^ γ first appeared in an article by the French mathematician, engineer, and physicist Siméon Denis Poisson:
    • Poisson (1808). "Mémoire sur la théorie du son" [Memoir on the theory of sound]. Journal de l'École Polytechnique (in French). 7 (14): 319–392. On p. 332, Poisson defines γ merely as a small deviation from equilibrium which causes small variations of the equilibrium value of the density ρ.
    In Poisson's article of 1823 –
    • Poisson (1823). "Sur la vitesse du son" [On the speed of sound]. Annales de chimie et de physique. 2nd series (in French). 23: 5–16.
    γ was expressed as a function of density D (p. 8) or of pressure P (p. 9).
    Meanwhile, in 1816 the French mathematician and physicist Pierre-Simon Laplace had found that the speed of sound depends on the ratio of the specific heats. However, he didn't denote the ratio as γ.
    In 1825, Laplace stated that the speed of sound is proportional to the square root of the ratio of the specific heats:
    • Laplace, P.S. (1825). Traité de mecanique celeste [Treatise on celestial mechanics] (in French). vol. 5. Paris, France: Bachelier. pp. 127–137. |volume= has extra text (help) On p. 127, Laplace defines the symbols for the specific heats, and on p. 137 (at the bottom of the page), Laplace presents the equation for the speed of sound in a perfect gas.
    In 1851, the Scottish mechanical engineer William Rankine showed that the speed of sound is proportional to the square root of Poisson's γ: It follows that Poisson's γ is the ratio of the specific heats — although Rankine didn't state that explicitly.
Retrieved from ""