Specific heat ratio: Difference between revisions

From Citizendium
Jump to navigation Jump to search
imported>Milton Beychok
m (→‎Determination of  C_v  values: Revised P to p to be consistent with Paul's revisons in the rest of the article)
imported>John Stephenson
m (moved Specific heat ratio/Draft to Specific heat ratio: citable version policy)
 
(18 intermediate revisions by 4 users not shown)
Line 104: Line 104:
|}
|}


The '''specific heat ratio''' of a [[gas]] is the ratio of the [[specific heat]] at constant [[pressure]], <math>C_p</math>, to the specific heat at constant volume, <math>C_v</math>. It is sometimes  referred to as the '''adiabatic index''' or the '''heat capacity ratio''' or the '''isentropic expansion factor'''.
The '''specific heat ratio''' of a [[gas]] is the ratio of the [[specific heat]] at constant [[pressure]], <math>C_p</math>, to the specific heat at constant volume, <math>C_v</math>. It is sometimes  referred to as the '''adiabatic index''' or the '''heat capacity ratio''' or the '''isentropic expansion factor''' or the '''adiabatic exponent''' or the '''isentropic exponent'''.


Either <math>k</math> ([[K (letter)|Roman letter k]]), <math>\gamma</math> ([[Gamma (Greek letter) |gamma]]) or <math>\kappa</math> ([[Kappa (Greek letter) |kappa]]) may be used to denote the specific heat ratio:
Either <math>\kappa</math> ([[Kappa (Greek letter) |kappa]]), <math>k</math> ([[K (letter)|Roman letter k]]) or <math>\gamma</math> ([[Gamma (Greek letter) |gamma]]) may be used to denote the specific heat ratio:


:<math>k = \gamma = \kappa = \frac{C_p}{C_v}</math>
:<math>\kappa = k = \gamma = \frac{C_p}{C_v}</math>


where:
where:
Line 119: Line 119:
== Ideal gas relations ==
== Ideal gas relations ==


For an [[Ideal gas law|ideal gas]], the [[heat capacity]] is constant with [[temperature]]. Accordingly we can express the [[enthalpy]]  as <math>H = C_p T</math> and the [[internal energy]] as <math>U = C_v T</math>. Thus, it can also be said that the specific heat ratio of an ideal gas is the ratio between the enthalpy to the internal energy:<ref name=Turns/>
In [[Thermodynamics|thermodynamic]] terminology, <math>C_p</math> and <math>C_V</math> may be expressed as:  


:<math> k = \frac{H}{U}</math>  
:<math>C_p=\left(\frac{\partial H}{\partial T}\right)_p </math> &nbsp; &nbsp; and &nbsp; &nbsp; <math>C_V=\left(\frac{\partial U}{\partial T}\right)_V</math>
 
where <math>T</math> stands for [[temperature]], <math>H</math> for the [[enthalpy]] and <math>U</math> for the [[internal energy]]. For an [[Ideal gas law|ideal gas]], the [[heat capacity]] is constant with temperature. Accordingly, we can express the enthalpy as <math>H = C_p\, T</math> and the internal energy as <math>U = C_v\, T</math>. Thus, it can also be said that the specific heat ratio of an ideal gas is the ratio of the enthalpy to the internal energy:<ref name=Turns/>
 
:<math>\kappa = \frac{H}{U}</math>  


The specific heats at constant pressure, <math>C_p</math>, of various gases are relatively easy to find in the technical literature. However, it can be difficult to find values of the specific heats at constant volume, <math>C_v</math>. When needed,  given <math>C_p</math>, the following equation can be used to determine <math>C_v</math> :<ref name=Turns/>
The specific heats at constant pressure, <math>C_p</math>, of various gases are relatively easy to find in the technical literature. However, it can be difficult to find values of the specific heats at constant volume, <math>C_v</math>. When needed,  given <math>C_p</math>, the following equation can be used to determine <math>C_v</math> :<ref name=Turns/>
Line 129: Line 133:
where <math>R</math> is the [[molar gas constant]] (also known as the [[Universal gas constant]]). This equation can be re-arranged to obtain:
where <math>R</math> is the [[molar gas constant]] (also known as the [[Universal gas constant]]). This equation can be re-arranged to obtain:


:<math> C_p = \frac{k R}{k - 1} \qquad \mbox{and} \qquad C_v = \frac{R}{k - 1}</math>
:<math> C_p = \frac{\kappa R}{\kappa - 1} \qquad \mbox{and} \qquad C_v = \frac{R}{\kappa - 1}</math>


=== Relation with degrees of freedom ===
=== Relation with degrees of freedom ===
Line 135: Line 139:
The specific heat ratio ( <math>k</math> ) for an ideal gas can be related to the [[degrees of freedom (physics)|degrees of freedom]] ( <math>f</math> ) of a molecule by:
The specific heat ratio ( <math>k</math> ) for an ideal gas can be related to the [[degrees of freedom (physics)|degrees of freedom]] ( <math>f</math> ) of a molecule by:


:<math> k = \frac{f+2}{f}</math>
:<math> \kappa = \frac{f+2}{f}</math>


Thus for a [[monatomic]] gas, with three degrees of freedom:
Thus for a [[monatomic]] gas, with three degrees of freedom:


:<math> k = \frac{5}{3} = 1.67</math>
:<math> \kappa = \frac{5}{3} = 1.67</math>


and for a [[diatomic]] gas, with five degrees of freedom (at room temperature):
and for a [[diatomic]] gas, with five degrees of freedom (three translational and two rotational degrees of freedom, the vibrational degree of freedom is not involved except at high temperatures):  


:<math> k = \frac{7}{5} = 1.4</math>.
:<math> \kappa = \frac{7}{5} = 1.4</math>.


Earth's atmospheric [[air]] is primarily made up of [[diatomic]] gases with a composition of ~78% [[nitrogen]] (N<sub>2</sub>) and ~21% [[oxygen]] (O<sub>2</sub>). At 20 °C and an absolute pressure of 101.325 kPa, the atmospheric air can be considered to be an ideal gas. A diatomic molecule has five degrees of freedom (three translational and two rotational degrees of freedom, the vibrational degree of freedom is not involved except at high temperatures). This results in a value of:
Earth's atmospheric [[air]] is primarily made up of [[diatomic]] gases with a composition of ~78% [[nitrogen]] (N<sub>2</sub>) and ~21% [[oxygen]] (O<sub>2</sub>). At 20 °C and an absolute pressure of 101.325 kPa, the atmospheric air can be considered to be an ideal gas. This results in a value of:
   
   
: <math> k = \frac{5 + 2}{5} = \frac{7}{5} = 1.4</math>  
: <math> \kappa = \frac{5 + 2}{5} = \frac{7}{5} = 1.4</math>  
 
which is consistent with the value of 1.40 listed for oxygen in the above table.


The specific heats of real gases (as differentiated from ideal gases) are not constant with temperature. As temperature increases, higher energy rotational and vibrational states become accessible to molecular gases, thus increasing the number of degrees of freedom and lowering <math>k</math>.
The specific heats of real gases (as differentiated from ideal gases) are not constant with temperature. As temperature increases, higher energy rotational and vibrational states become accessible to molecular gases, thus increasing the number of degrees of freedom and lowering <math>\kappa</math>.
   
   
For a real gas, <math>C_p</math> and <math>C_v</math> usually increase with increasing temperature and <math>k</math> decreases. Some correlations exist to provide values of <math>k</math> as a function of the temperature.
For a real gas, <math>C_p</math> and <math>C_v</math> usually increase with increasing temperature and <math>\kappa</math> decreases. Some correlations exist to provide values of <math>\kappa</math> as a function of the temperature.<ref>
{{cite book|
author=Richard W. Miller
|title=Flow Measurement Engineering Handbook
|edition=3rd Edition
|publisher=McGraw-Hill
|date=1996
|pages=pp 2.110-2.115
|url=http://books.google.de/books?id=lxd_Fcv7mHAC&pg=PT118&vq=%22specific+heat+ratio%22&dq=%22Flow+Measurement+Engineering+Handbook%22&lr=&hl=en&source=gbs_search_s&cad=0
|id=ISBN 0-07-042366-0
}}</ref>


== Isentropic compression or expansion of ideal gases ==
== Isentropic compression or expansion of ideal gases ==


The specific heat ratio plays an important part in the [[isentropic]] process of an ideal gas (i.e., a process that occurs at constant [[entropy]]):<ref name=Turns/>
The specific heat ratio plays an important part in the [[isentropic process]] of an ideal gas (i.e., a process that occurs at constant [[entropy (thermodynamics)|entropy]]):<ref name=Turns/>


:(1) &nbsp; &nbsp;<math> p_1{V_1}^k = p_2{V_2}^k </math>
:(1) &nbsp; &nbsp;<math> p_1{V_1}^\kappa = p_2{V_2}^\kappa </math>


where, <math>p</math> is the absolute pressure and <math>V</math> is the volume. The subscripts 1 and 2  refer to conditions before and after the process, or at any time during that process.  
where, <math>p</math> is the absolute pressure and <math>V</math> is the volume. The subscripts 1 and 2  refer to conditions before and after the process, or at any time during that process.  
Line 165: Line 177:
Using the ideal gas law, <math>pV = nRT</math>, equation (1) can be re-arranged to:
Using the ideal gas law, <math>pV = nRT</math>, equation (1) can be re-arranged to:


:(2) &nbsp; &nbsp;<math>\frac{p_1}{p_2} = \left(\frac{V_2}{V_1}\right)^k = \left(\frac{T_2}{T_1}\right)^k \left(\frac{p_1}{p_2}\right)^k</math>
:(2) &nbsp; &nbsp;<math>\frac{p_1}{p_2} = \left(\frac{V_2}{V_1}\right)^\kappa = \left(\frac{T_2}{T_1}\right)^\kappa \left(\frac{p_1}{p_2}\right)^\kappa</math>


where <math>T</math> is the absolute temperature. Re-arranging further:
where <math>T</math> is the absolute temperature. Re-arranging further:


:(3) &nbsp; &nbsp;<math>\left(\frac{T_2}{T_1}\right)^k = \left(\frac{p_1}{p_2}\right)\left(\frac{p_2}{p_1}\right)^k = \left(\frac{p_2}{p_1}\right)^{k-1}</math>
:(3) &nbsp; &nbsp;<math>\left(\frac{T_2}{T_1}\right)^\kappa = \left(\frac{p_1}{p_2}\right)\left(\frac{p_2}{p_1}\right)^\kappa = \left(\frac{p_2}{p_1}\right)^{\kappa-1}</math>


we obtain the equation for the temperature change that occurs when an ideal gas is isentropically compressed or expanded:<ref name=Turns/><ref>{{cite book|author=Don. W. Green, James O Maloney and Robert H. Perry (Editors)|title=Perry's Chemical Engineers' Handbook|edition=Sixth Edition|publisher=McGraw-Hill|pages=page 6-17|year=1984|id=ISBN 0-07-049479-7}}</ref>
we obtain the equation for the temperature change that occurs when an ideal gas is isentropically compressed or expanded:<ref name=Turns/><ref>{{cite book|author=Don. W. Green, James O Maloney and Robert H. Perry (Editors)|title=Perry's Chemical Engineers' Handbook|edition=Sixth Edition|publisher=McGraw-Hill|pages=page 6-17|year=1984|id=ISBN 0-07-049479-7}}</ref>


:(4) &nbsp; &nbsp; <math>\frac{T_2}{T_1} = \left(\frac{p_2}{p_1}\right)^{(k-1)/k}</math>
:(4) &nbsp; &nbsp; <math>\frac{T_2}{T_1} = \left(\frac{p_2}{p_1}\right)^{(\kappa-1)/\kappa}</math>


Equation (4) is widely used to model ideal gas compression or expansion processes in [[internal combustion engine]]s, [[gas compressor]]s and [[gas turbines]].
Equation (4) is widely used to model ideal gas compression or expansion processes in [[internal combustion engine]]s, [[gas compressor]]s and [[gas turbines]].
Line 185: Line 197:
:<math>C_v = C_p + T \frac{{\; \left( {\frac{\part p}{\part T}} \right) }^2_V} {\left( \frac{\part p}{\part V} \right)_T} </math>
:<math>C_v = C_p + T \frac{{\; \left( {\frac{\part p}{\part T}} \right) }^2_V} {\left( \frac{\part p}{\part V} \right)_T} </math>


[[Equations of state]] (EOS) (such as the Peng-Robinson equation of state) can be used to solve this relation and to provide values of <math>C_v</math> that match experimental values very closely.
[[Equations of state]] (EOS) (such as the [[Peng-Robinson equation of state]]) can be used to solve this relation and to provide values of <math>C_v</math> that match experimental values very closely.  
 
== Definitions of specific heat and heat capacity ==
 
The ''specific heat'' (or ''specific heat capacity''), is the amount of heat energy required to increase the temperature of a unit quantity of a substance by a certain temperature interval. For example, the heat required to raise the temperature of water by 1 [[Kelvin (unit)|kelvin]] is 4.184 [[Joule (unit)|joule]]s. The specific heat capacity is usually expressed as Jg<sup>-1</sup>K<sup>-1</sup>. It may also be expressed on a molar basis as Jmol<sup>-1</sup>K<sup>-1</sup>.
 
The ''heat capacity'' (as distinct from ''specific heat'') is the amount of heat energy required to increase the temperature of a substance by a certain temperature interval. Heat capacity is an extensive property because its value is proportional to the amount of the substance. For example, a kilogram of water has a greater heat capacity than 100 grams of water. The heat capacity is usually expressed as JK<sup>-1</sup>.
 
''Specific heat capacities'' and ''heat capacities'' have the same symbols of <math>C_p</math> and <math>C_v</math>. The specific heat ratio, <math>k</math>, has the same numeric value whether based on ''specific heats'' or ''heat capacities''.


== References ==
== References ==
<references/>
<references/>

Latest revision as of 16:10, 23 September 2013

This article has a Citable Version.
Main Article
Discussion
Related Articles  [?]
Bibliography  [?]
External Links  [?]
Citable Version  [?]
 
This editable Main Article has an approved citable version (see its Citable Version subpage). While we have done conscientious work, we cannot guarantee that this Main Article, or its citable version, is wholly free of mistakes. By helping to improve this editable Main Article, you will help the process of generating a new, improved citable version.
Specific heat ratio of various gases[1][2][3]
Gas °C k   Gas °C k
H2 −181 1.597 Dry
Air
20 1.40
−76 1.453 100 1.401
20 1.41 200 1.398
100 1.404 400 1.393
400 1.387 CO2 0 1.310
1000 1.358 20 1.30
2000 1.318 100 1.281
He 20 1.66 400 1.235
N2 −181 1.47 NH3 15 1.310
15 1.404 CO 20 1.40
Cl2 20 1.34 O2 −181 1.45
Ar −180 1.76 −76 1.415
20 1.67 20 1.40
CH4 −115 1.41 100 1.399
−74 1.35 200 1.397
20 1.32 400 1.394

The specific heat ratio of a gas is the ratio of the specific heat at constant pressure, , to the specific heat at constant volume, . It is sometimes referred to as the adiabatic index or the heat capacity ratio or the isentropic expansion factor or the adiabatic exponent or the isentropic exponent.

Either (kappa), (Roman letter k) or (gamma) may be used to denote the specific heat ratio:

where:

= the specific heat of a gas
= refers to constant pressure conditions
= refers to constant volume conditions



Ideal gas relations

In thermodynamic terminology, and may be expressed as:

    and    

where stands for temperature, for the enthalpy and for the internal energy. For an ideal gas, the heat capacity is constant with temperature. Accordingly, we can express the enthalpy as and the internal energy as . Thus, it can also be said that the specific heat ratio of an ideal gas is the ratio of the enthalpy to the internal energy:[3]

The specific heats at constant pressure, , of various gases are relatively easy to find in the technical literature. However, it can be difficult to find values of the specific heats at constant volume, . When needed, given , the following equation can be used to determine  :[3]

where is the molar gas constant (also known as the Universal gas constant). This equation can be re-arranged to obtain:

Relation with degrees of freedom

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

Thus for a monatomic gas, with three degrees of freedom:

and for a diatomic gas, with five degrees of freedom (three translational and two rotational degrees of freedom, the vibrational degree of freedom is not involved except at high temperatures):

.

Earth's atmospheric air is primarily made up of diatomic gases with a composition of ~78% nitrogen (N2) and ~21% oxygen (O2). At 20 °C and an absolute pressure of 101.325 kPa, the atmospheric air can be considered to be an ideal gas. This results in a value of:

The specific heats of real gases (as differentiated from ideal gases) are not constant with temperature. As temperature increases, higher energy rotational and vibrational states become accessible to molecular gases, thus increasing the number of degrees of freedom and lowering .

For a real gas, and usually increase with increasing temperature and decreases. Some correlations exist to provide values of as a function of the temperature.[4]

Isentropic compression or expansion of ideal gases

The specific heat ratio plays an important part in the isentropic process of an ideal gas (i.e., a process that occurs at constant entropy):[3]

(1)    

where, is the absolute pressure and is the volume. The subscripts 1 and 2 refer to conditions before and after the process, or at any time during that process.

Using the ideal gas law, , equation (1) can be re-arranged to:

(2)    

where is the absolute temperature. Re-arranging further:

(3)    

we obtain the equation for the temperature change that occurs when an ideal gas is isentropically compressed or expanded:[3][5]

(4)    

Equation (4) is widely used to model ideal gas compression or expansion processes in internal combustion engines, gas compressors and gas turbines.

Determination of    values

Values for are readily available, but values for are not as available and often need to be determined. Values based on the ideal gas relation of   are in many cases not sufficiently accurate for practical engineering calculations. If at all possible, an experimental value should be used.

A rigorous value can be calculated by determining from the residual property functions (also referred to as departure functions)[6][7][8] using this relation:[9]

Equations of state (EOS) (such as the Peng-Robinson equation of state) can be used to solve this relation and to provide values of that match experimental values very closely.

References

  1. Frank M. White (1999). Fluid Mechanics, Fourth Edition. McGraw-Hill. ISBN 0-07-0697167. 
  2. Norbert A. Lange (Editor) (1969). Lange's Handbook of Chemistry, 10th Edition. McGraw-Hill, page 1524. 
  3. 3.0 3.1 3.2 3.3 3.4 Stephan R. Turns (2006). Thermodynamics: Concepts and Application, First Edition. Cambridge University Press. ISBN 0-521-85042-8. 
  4. Richard W. Miller (1996). Flow Measurement Engineering Handbook, 3rd Edition. McGraw-Hill, pp 2.110-2.115. ISBN 0-07-042366-0. 
  5. Don. W. Green, James O Maloney and Robert H. Perry (Editors) (1984). Perry's Chemical Engineers' Handbook, Sixth Edition. McGraw-Hill, page 6-17. ISBN 0-07-049479-7. 
  6. K.Y. Narayanan (2001). A Textbook of Chemical Engineering Thermodynamics. Prentice-Hall India. ISBN 81-203-1732-7. 
  7. Y.V.C. Rao (1997). Chemical Engineering Thermodynamics, First Edition. Universities Press. ISBN 81-7371-048-1. 
  8. Thermodynamics of Pure Substances Lecture by Mark Gibbs, University of Edinburgh, Scotland.
  9. Isochoric heat capacity (pdf page 61 of 308)