fbpx
Wikipedia

Clausius–Clapeyron relation

The Clausius–Clapeyron relation, in chemical thermodynamics specifies the temperature dependence of pressure, most importantly vapor pressure, at a discontinuous phase transition between two phases of matter of a single constituent. It's named after Rudolf Clausius[1] and Benoît Paul Émile Clapeyron.[2] Its relevance to meteorology and climatology is the increase of the water-holding capacity of the atmosphere by about 7% for every 1 °C (1.8 °F) rise in temperature.

James Thomson and William Thomson confirmed the relation experimentally in 1849-50, and it was historically important as a very early successful application of theoretical thermodynamics.[3]

Definition Edit

Exact Clapeyron equation Edit

On a pressuretemperature (PT) diagram, for any phase change the line separating the two phases is known as the coexistence curve. The Clapeyron relation[4] gives the slope of the tangents to this curve. Mathematically,

 

where   is the slope of the tangent to the coexistence curve at any point,   is the specific latent heat,   is the temperature,   is the specific volume change of the phase transition, and   is the specific entropy change of the phase transition.

Clausius–Clapeyron equation Edit

The Clausius–Clapeyron equation[5]: 509  applies to vaporization of liquids where vapor follows ideal gas law using the ideal gas constant   and liquid volume is neglected as being much smaller than vapor volume V. It is often used to calculate vapor pressure of a liquid.[6]

 
 

The equation expresses this in a more convenient form just in terms of the latent heat, for moderate temperatures and pressures.

Derivations Edit

 
A typical phase diagram. The dotted green line gives the anomalous behavior of water. The Clausius–Clapeyron relation can be used to find the relationship between pressure and temperature along phase boundaries.

Derivation from state postulate Edit

Using the state postulate, take the specific entropy   for a homogeneous substance to be a function of specific volume   and temperature  .[5]: 508 

 

The Clausius–Clapeyron relation describes a Phase transition in a closed system composed of two contiguous phases, condensed matter and ideal gas, of a single substance, in mutual thermodynamic equilibrium, at constant temperature and pressure. Therefore,[5]: 508 

 

Using the appropriate Maxwell relation gives[5]: 508 

 

where   is the pressure. Since pressure and temperature are constant, the derivative of pressure with respect to temperature does not change.[7][8]: 57, 62, 671  Therefore, the partial derivative of specific entropy may be changed into a total derivative

 

and the total derivative of pressure with respect to temperature may be factored out when integrating from an initial phase   to a final phase  ,[5]: 508  to obtain

 

where   and   are respectively the change in specific entropy and specific volume. Given that a phase change is an internally reversible process, and that our system is closed, the first law of thermodynamics holds:

 

where   is the internal energy of the system. Given constant pressure and temperature (during a phase change) and the definition of specific enthalpy  , we obtain

 
 
 

Given constant pressure and temperature (during a phase change), we obtain[5]: 508 

 

Substituting the definition of specific latent heat   gives

 

Substituting this result into the pressure derivative given above ( ), we obtain[5]: 508 [9]

 

This result (also known as the Clapeyron equation) equates the slope   of the coexistence curve   to the function   of the specific latent heat  , the temperature  , and the change in specific volume  . Instead of the specific, corresponding molar values may also be used.

Derivation from Gibbs–Duhem relation Edit

Suppose two phases,   and  , are in contact and at equilibrium with each other. Their chemical potentials are related by

 

Furthermore, along the coexistence curve,

 

One may therefore use the Gibbs–Duhem relation

 

(where   is the specific entropy,   is the specific volume, and   is the molar mass) to obtain

 

Rearrangement gives

 

from which the derivation of the Clapeyron equation continues as in the previous section.

Ideal gas approximation at low temperatures Edit

When the phase transition of a substance is between a gas phase and a condensed phase (liquid or solid), and occurs at temperatures much lower than the critical temperature of that substance, the specific volume of the gas phase   greatly exceeds that of the condensed phase  . Therefore, one may approximate

 

at low temperatures. If pressure is also low, the gas may be approximated by the ideal gas law, so that

 

where   is the pressure,   is the specific gas constant, and   is the temperature. Substituting into the Clapeyron equation

 

we can obtain the Clausius–Clapeyron equation[5]: 509 

 

for low temperatures and pressures,[5]: 509  where   is the specific latent heat of the substance. Instead of the specific, corresponding molar values (i.e.   in kJ/mol and R = 8.31 J/(mol⋅K)) may also be used.

Let   and   be any two points along the coexistence curve between two phases   and  . In general,   varies between any two such points, as a function of temperature. But if   is approximated as constant,

 
 
 

or[8]: 672 [10]

 

These last equations are useful because they relate equilibrium or saturation vapor pressure and temperature to the latent heat of the phase change without requiring specific-volume data. For instance, for water near its normal boiling point, with a molar enthalpy of vaporization of 40.7 kJ/mol and R = 8.31 J/(mol⋅K),

 .

Clapeyron's derivation Edit

In the original work by Clapeyron, the following argument is advanced.[11] Clapeyron considered a Carnot process of saturated water vapor with horizontal isobars. As the pressure is a function of temperature alone, the isobars are also isotherms. If the process involves an infinitesimal amount of water,  , and an infinitesimal difference in temperature  , the heat absorbed is

 

and the corresponding work is

 

where   is the difference between the volumes of   in the liquid phase and vapor phases. The ratio   is the efficiency of the Carnot engine,  .[a] Substituting and rearranging gives

 

where lowercase   denotes the change in specific volume during the transition.

Applications Edit

Chemistry and chemical engineering Edit

For transitions between a gas and a condensed phase with the approximations described above, the expression may be rewritten as

 

where   is the pressure,   is the specific gas constant (i.e., the gas constant R divided by the molar mass),   is the absolute temperature, and   is a constant. For a liquid–gas transition,   is the specific latent heat (or specific enthalpy) of vaporization; for a solid–gas transition,   is the specific latent heat of sublimation. If the latent heat is known, then knowledge of one point on the coexistence curve, for instance (1 bar, 373 K) for water, determines the rest of the curve. Conversely, the relationship between   and   is linear, and so linear regression is used to estimate the latent heat.

Meteorology and climatology Edit

Atmospheric water vapor drives many important meteorologic phenomena (notably, precipitation), motivating interest in its dynamics. The Clausius–Clapeyron equation for water vapor under typical atmospheric conditions (near standard temperature and pressure) is

 

where

  is saturation vapor pressure,
  is temperature,
  is the specific latent heat of evaporation of water,
  is the gas constant of water vapor.

The temperature dependence of the latent heat   (and of the saturation vapor pressure  ) cannot be neglected in this application. Fortunately, the August–RocheMagnus formula provides a very good approximation:[12][13]

 

where   is in hPa, and   is in degrees Celsius (whereas everywhere else on this page,   is an absolute temperature, e.g. in kelvins).

This is also sometimes called the Magnus or Magnus–Tetens approximation, though this attribution is historically inaccurate.[14] But see also the discussion of the accuracy of different approximating formulae for saturation vapour pressure of water.

Under typical atmospheric conditions, the denominator of the exponent depends weakly on   (for which the unit is degree Celsius). Therefore, the August–Roche–Magnus equation implies that saturation water vapor pressure changes approximately exponentially with temperature under typical atmospheric conditions, and hence the water-holding capacity of the atmosphere increases by about 7% for every 1 °C rise in temperature.[15]

Example Edit

One of the uses of this equation is to determine if a phase transition will occur in a given situation. Consider the question of how much pressure is needed to melt ice at a temperature   below 0 °C. Note that water is unusual in that its change in volume upon melting is negative. We can assume

 

and substituting in

  (latent heat of fusion for water),
  K (absolute temperature),
  (change in specific volume from solid to liquid),

we obtain

 

To provide a rough example of how much pressure this is, to melt ice at −7 °C (the temperature many ice skating rinks are set at) would require balancing a small car (mass ~ 1000 kg[16]) on a thimble (area ~ 1 cm2). This shows that ice skating cannot be simply explained by pressure-caused melting point depression, and in fact the mechanism is quite complex.[17]

Second derivative Edit

While the Clausius–Clapeyron relation gives the slope of the coexistence curve, it does not provide any information about its curvature or second derivative. The second derivative of the coexistence curve of phases 1 and 2 is given by[18]

 

where subscripts 1 and 2 denote the different phases,   is the specific heat capacity at constant pressure,   is the thermal expansion coefficient, and   is the isothermal compressibility.

See also Edit

References Edit

  1. ^ Clausius, R. (1850). "Ueber die bewegende Kraft der Wärme und die Gesetze, welche sich daraus für die Wärmelehre selbst ableiten lassen" [On the motive power of heat and the laws which can be deduced therefrom regarding the theory of heat]. Annalen der Physik (in German). 155 (4): 500–524. Bibcode:1850AnP...155..500C. doi:10.1002/andp.18501550403. hdl:2027/uc1.$b242250.
  2. ^ Clapeyron, M. C. (1834). "Mémoire sur la puissance motrice de la chaleur". Journal de l'École polytechnique [fr] (in French). 23: 153–190. ark:/12148/bpt6k4336791/f157.
  3. ^ Pippard, Alfred B. (1981). Elements of classical thermodynamics: for advanced students of physics (Repr ed.). Cambridge: Univ. Pr. p. 116. ISBN 978-0-521-09101-5.
  4. ^ Koziol, Andrea; Perkins, Dexter. "Teaching Phase Equilibria". serc.carleton.edu. Carleton University. Retrieved 1 February 2023.
  5. ^ a b c d e f g h i Wark, Kenneth (1988) [1966]. "Generalized Thermodynamic Relationships". Thermodynamics (5th ed.). New York, NY: McGraw-Hill, Inc. ISBN 978-0-07-068286-3.
  6. ^ Clausius; Clapeyron. "The Clausius-Clapeyron Equation". Bodner Research Web. Purdue University. Retrieved 1 February 2023.
  7. ^ Carl Rod Nave (2006). "PvT Surface for a Substance which Contracts Upon Freezing". HyperPhysics. Georgia State University. Retrieved 2007-10-16.
  8. ^ a b Çengel, Yunus A.; Boles, Michael A. (1998) [1989]. Thermodynamics – An Engineering Approach. McGraw-Hill Series in Mechanical Engineering (3rd ed.). Boston, MA.: McGraw-Hill. ISBN 978-0-07-011927-7.
  9. ^ Salzman, William R. (2001-08-21). . Chemical Thermodynamics. University of Arizona. Archived from the original on 2007-06-07. Retrieved 2007-10-11.
  10. ^ Masterton, William L.; Hurley, Cecile N. (2008). Chemistry : principles and reactions (6th ed.). Cengage Learning. p. 230. ISBN 9780495126713. Retrieved 3 April 2020.
  11. ^ Clapeyron, E (1834). "Mémoire sur la puissance motrice de la chaleur". Journal de l ́École Polytechnique. XIV: 153–190.
  12. ^ Alduchov, Oleg; Eskridge, Robert (1997-11-01), Improved Magnus' Form Approximation of Saturation Vapor Pressure, NOAA, doi:10.2172/548871 Equation 25 provides these coefficients.
  13. ^ Alduchov, Oleg A.; Eskridge, Robert E. (1996). "Improved Magnus Form Approximation of Saturation Vapor Pressure". Journal of Applied Meteorology. 35 (4): 601–609. Bibcode:1996JApMe..35..601A. doi:10.1175/1520-0450(1996)035<0601:IMFAOS>2.0.CO;2. Equation 21 provides these coefficients.
  14. ^ Lawrence, M. G. (2005). "The Relationship between Relative Humidity and the Dewpoint Temperature in Moist Air: A Simple Conversion and Applications" (PDF). Bulletin of the American Meteorological Society. 86 (2): 225–233. Bibcode:2005BAMS...86..225L. doi:10.1175/BAMS-86-2-225.
  15. ^ IPCC, Climate Change 2007: Working Group I: The Physical Science Basis, "FAQ 3.2 How is Precipitation Changing?". 2018-11-02 at the Wayback Machine.
  16. ^ Zorina, Yana (2000). "Mass of a Car". The Physics Factbook.
  17. ^ Liefferink, Rinse W.; Hsia, Feng-Chun; Weber, Bart; Bonn, Daniel (2021-02-08). "Friction on Ice: How Temperature, Pressure, and Speed Control the Slipperiness of Ice". Physical Review X. 11 (1): 011025. doi:10.1103/PhysRevX.11.011025.
  18. ^ Krafcik, Matthew; Sánchez Velasco, Eduardo (2014). "Beyond Clausius–Clapeyron: Determining the second derivative of a first-order phase transition line". American Journal of Physics. 82 (4): 301–305. Bibcode:2014AmJPh..82..301K. doi:10.1119/1.4858403.

Bibliography Edit

  • Yau, M. K.; Rogers, R. R. (1989). Short Course in Cloud Physics (3rd ed.). Butterworth–Heinemann. ISBN 978-0-7506-3215-7.
  • Iribarne, J. V.; Godson, W. L. (2013). "4. Water-Air systems § 4.8 Clausius–Clapeyron Equation". Atmospheric Thermodynamics. Springer. pp. 60–. ISBN 978-94-010-2642-0.
  • Callen, H. B. (1985). Thermodynamics and an Introduction to Thermostatistics. Wiley. ISBN 978-0-471-86256-7.

Notes Edit

  1. ^ In the original work,   was simply called the Carnot function and was not known in this form. Clausius determined the form 30 years later and added his name to the eponymous Clausius–Clapeyron relation.

clausius, clapeyron, relation, clapeyron, equation, clapeyron, equation, redirect, here, state, equation, ideal, chemical, thermodynamics, specifies, temperature, dependence, pressure, most, importantly, vapor, pressure, discontinuous, phase, transition, betwe. Clapeyron equation and Clapeyron s equation redirect here For a state equation see ideal gas law The Clausius Clapeyron relation in chemical thermodynamics specifies the temperature dependence of pressure most importantly vapor pressure at a discontinuous phase transition between two phases of matter of a single constituent It s named after Rudolf Clausius 1 and Benoit Paul Emile Clapeyron 2 Its relevance to meteorology and climatology is the increase of the water holding capacity of the atmosphere by about 7 for every 1 C 1 8 F rise in temperature James Thomson and William Thomson confirmed the relation experimentally in 1849 50 and it was historically important as a very early successful application of theoretical thermodynamics 3 Contents 1 Definition 1 1 Exact Clapeyron equation 1 2 Clausius Clapeyron equation 2 Derivations 2 1 Derivation from state postulate 2 2 Derivation from Gibbs Duhem relation 2 3 Ideal gas approximation at low temperatures 2 4 Clapeyron s derivation 3 Applications 3 1 Chemistry and chemical engineering 3 2 Meteorology and climatology 4 Example 5 Second derivative 6 See also 7 References 8 Bibliography 9 NotesDefinition EditExact Clapeyron equation Edit On a pressure temperature P T diagram for any phase change the line separating the two phases is known as the coexistence curve The Clapeyron relation 4 gives the slope of the tangents to this curve Mathematically d P d T L T D v D s D v displaystyle frac mathrm d P mathrm d T frac L T Delta v frac Delta s Delta v nbsp where d P d T displaystyle mathrm d P mathrm d T nbsp is the slope of the tangent to the coexistence curve at any point L displaystyle L nbsp is the specific latent heat T displaystyle T nbsp is the temperature D v displaystyle Delta v nbsp is the specific volume change of the phase transition and D s displaystyle Delta s nbsp is the specific entropy change of the phase transition Clausius Clapeyron equation Edit The Clausius Clapeyron equation 5 509 applies to vaporization of liquids where vapor follows ideal gas law using the ideal gas constant R displaystyle R nbsp and liquid volume is neglected as being much smaller than vapor volume V It is often used to calculate vapor pressure of a liquid 6 d P d T P L T 2 R displaystyle frac mathrm d P mathrm d T frac PL T 2 R nbsp V R T P displaystyle V frac RT P nbsp The equation expresses this in a more convenient form just in terms of the latent heat for moderate temperatures and pressures Derivations Edit nbsp A typical phase diagram The dotted green line gives the anomalous behavior of water The Clausius Clapeyron relation can be used to find the relationship between pressure and temperature along phase boundaries Derivation from state postulate Edit Using the state postulate take the specific entropy s displaystyle s nbsp for a homogeneous substance to be a function of specific volume v displaystyle v nbsp and temperature T displaystyle T nbsp 5 508 d s s v T d v s T v d T displaystyle mathrm d s left frac partial s partial v right T mathrm d v left frac partial s partial T right v mathrm d T nbsp The Clausius Clapeyron relation describes a Phase transition in a closed system composed of two contiguous phases condensed matter and ideal gas of a single substance in mutual thermodynamic equilibrium at constant temperature and pressure Therefore 5 508 d s s v T d v displaystyle mathrm d s left frac partial s partial v right T mathrm d v nbsp Using the appropriate Maxwell relation gives 5 508 d s P T v d v displaystyle mathrm d s left frac partial P partial T right v mathrm d v nbsp where P displaystyle P nbsp is the pressure Since pressure and temperature are constant the derivative of pressure with respect to temperature does not change 7 8 57 62 671 Therefore the partial derivative of specific entropy may be changed into a total derivative d s d P d T d v displaystyle mathrm d s frac mathrm d P mathrm d T mathrm d v nbsp and the total derivative of pressure with respect to temperature may be factored out when integrating from an initial phase a displaystyle alpha nbsp to a final phase b displaystyle beta nbsp 5 508 to obtain d P d T D s D v displaystyle frac mathrm d P mathrm d T frac Delta s Delta v nbsp where D s s b s a displaystyle Delta s equiv s beta s alpha nbsp and D v v b v a displaystyle Delta v equiv v beta v alpha nbsp are respectively the change in specific entropy and specific volume Given that a phase change is an internally reversible process and that our system is closed the first law of thermodynamics holds d u d q d w T d s P d v displaystyle mathrm d u delta q delta w T mathrm d s P mathrm d v nbsp where u displaystyle u nbsp is the internal energy of the system Given constant pressure and temperature during a phase change and the definition of specific enthalpy h displaystyle h nbsp we obtain d h T d s v d P displaystyle mathrm d h T mathrm d s v mathrm d P nbsp d h T d s displaystyle mathrm d h T mathrm d s nbsp d s d h T displaystyle mathrm d s frac mathrm d h T nbsp Given constant pressure and temperature during a phase change we obtain 5 508 D s D h T displaystyle Delta s frac Delta h T nbsp Substituting the definition of specific latent heat L D h displaystyle L Delta h nbsp gives D s L T displaystyle Delta s frac L T nbsp Substituting this result into the pressure derivative given above d P d T D s D v displaystyle mathrm d P mathrm d T Delta s Delta v nbsp we obtain 5 508 9 d P d T L T D v displaystyle frac mathrm d P mathrm d T frac L T Delta v nbsp This result also known as the Clapeyron equation equates the slope d P d T displaystyle mathrm d P mathrm d T nbsp of the coexistence curve P T displaystyle P T nbsp to the function L T D v displaystyle L T Delta v nbsp of the specific latent heat L displaystyle L nbsp the temperature T displaystyle T nbsp and the change in specific volume D v displaystyle Delta v nbsp Instead of the specific corresponding molar values may also be used Derivation from Gibbs Duhem relation Edit Suppose two phases a displaystyle alpha nbsp and b displaystyle beta nbsp are in contact and at equilibrium with each other Their chemical potentials are related by m a m b displaystyle mu alpha mu beta nbsp Furthermore along the coexistence curve d m a d m b displaystyle mathrm d mu alpha mathrm d mu beta nbsp One may therefore use the Gibbs Duhem relation d m M s d T v d P displaystyle mathrm d mu M s mathrm d T v mathrm d P nbsp where s displaystyle s nbsp is the specific entropy v displaystyle v nbsp is the specific volume and M displaystyle M nbsp is the molar mass to obtain s b s a d T v b v a d P 0 displaystyle s beta s alpha mathrm d T v beta v alpha mathrm d P 0 nbsp Rearrangement gives d P d T s b s a v b v a D s D v displaystyle frac mathrm d P mathrm d T frac s beta s alpha v beta v alpha frac Delta s Delta v nbsp from which the derivation of the Clapeyron equation continues as in the previous section Ideal gas approximation at low temperatures Edit When the phase transition of a substance is between a gas phase and a condensed phase liquid or solid and occurs at temperatures much lower than the critical temperature of that substance the specific volume of the gas phase v g displaystyle v text g nbsp greatly exceeds that of the condensed phase v c displaystyle v text c nbsp Therefore one may approximate D v v g 1 v c v g v g displaystyle Delta v v text g left 1 frac v text c v text g right approx v text g nbsp at low temperatures If pressure is also low the gas may be approximated by the ideal gas law so that v g R T P displaystyle v text g frac RT P nbsp where P displaystyle P nbsp is the pressure R displaystyle R nbsp is the specific gas constant and T displaystyle T nbsp is the temperature Substituting into the Clapeyron equation d P d T L T D v displaystyle frac mathrm d P mathrm d T frac L T Delta v nbsp we can obtain the Clausius Clapeyron equation 5 509 d P d T P L T 2 R displaystyle frac mathrm d P mathrm d T frac PL T 2 R nbsp for low temperatures and pressures 5 509 where L displaystyle L nbsp is the specific latent heat of the substance Instead of the specific corresponding molar values i e L displaystyle L nbsp in kJ mol and R 8 31 J mol K may also be used Let P 1 T 1 displaystyle P 1 T 1 nbsp and P 2 T 2 displaystyle P 2 T 2 nbsp be any two points along the coexistence curve between two phases a displaystyle alpha nbsp and b displaystyle beta nbsp In general L displaystyle L nbsp varies between any two such points as a function of temperature But if L displaystyle L nbsp is approximated as constant d P P L R d T T 2 displaystyle frac mathrm d P P cong frac L R frac mathrm d T T 2 nbsp P 1 P 2 d P P L R T 1 T 2 d T T 2 displaystyle int P 1 P 2 frac mathrm d P P cong frac L R int T 1 T 2 frac mathrm d T T 2 nbsp ln P P P 1 P 2 L R 1 T T T 1 T 2 displaystyle ln P Big P P 1 P 2 cong frac L R cdot left frac 1 T right T T 1 T 2 nbsp or 8 672 10 ln P 2 P 1 L R 1 T 2 1 T 1 displaystyle ln frac P 2 P 1 cong frac L R left frac 1 T 2 frac 1 T 1 right nbsp These last equations are useful because they relate equilibrium or saturation vapor pressure and temperature to the latent heat of the phase change without requiring specific volume data For instance for water near its normal boiling point with a molar enthalpy of vaporization of 40 7 kJ mol and R 8 31 J mol K P vap T 1 bar exp 40 700 K 8 31 1 T 1 373 K displaystyle P text vap T cong 1 text bar cdot exp left frac 40 700 text K 8 31 left frac 1 T frac 1 373 text K right right nbsp Clapeyron s derivation Edit In the original work by Clapeyron the following argument is advanced 11 Clapeyron considered a Carnot process of saturated water vapor with horizontal isobars As the pressure is a function of temperature alone the isobars are also isotherms If the process involves an infinitesimal amount of water d x displaystyle mathrm d x nbsp and an infinitesimal difference in temperature d T displaystyle mathrm d T nbsp the heat absorbed is Q L d x displaystyle Q L mathrm d x nbsp and the corresponding work is W d p d T d T V V displaystyle W frac mathrm d p mathrm d T mathrm d T V V nbsp where V V displaystyle V V nbsp is the difference between the volumes of d x displaystyle mathrm d x nbsp in the liquid phase and vapor phases The ratio W Q displaystyle W Q nbsp is the efficiency of the Carnot engine d T T displaystyle mathrm d T T nbsp a Substituting and rearranging gives d p d T L T v v displaystyle frac mathrm d p mathrm d T frac L T v v nbsp where lowercase v v displaystyle v v nbsp denotes the change in specific volume during the transition Applications EditChemistry and chemical engineering Edit For transitions between a gas and a condensed phase with the approximations described above the expression may be rewritten as ln P L R 1 T c displaystyle ln P frac L R frac 1 T c nbsp where P displaystyle P nbsp is the pressure R displaystyle R nbsp is the specific gas constant i e the gas constant R divided by the molar mass T displaystyle T nbsp is the absolute temperature and c displaystyle c nbsp is a constant For a liquid gas transition L displaystyle L nbsp is the specific latent heat or specific enthalpy of vaporization for a solid gas transition L displaystyle L nbsp is the specific latent heat of sublimation If the latent heat is known then knowledge of one point on the coexistence curve for instance 1 bar 373 K for water determines the rest of the curve Conversely the relationship between ln P displaystyle ln P nbsp and 1 T displaystyle 1 T nbsp is linear and so linear regression is used to estimate the latent heat Meteorology and climatology Edit Atmospheric water vapor drives many important meteorologic phenomena notably precipitation motivating interest in its dynamics The Clausius Clapeyron equation for water vapor under typical atmospheric conditions near standard temperature and pressure is d e s d T L v T e s R v T 2 displaystyle frac mathrm d e s mathrm d T frac L v T e s R v T 2 nbsp where e s displaystyle e s nbsp is saturation vapor pressure T displaystyle T nbsp is temperature L v displaystyle L v nbsp is the specific latent heat of evaporation of water R v displaystyle R v nbsp is the gas constant of water vapor The temperature dependence of the latent heat L v T displaystyle L v T nbsp and of the saturation vapor pressure e s displaystyle e s nbsp cannot be neglected in this application Fortunately the August Roche Magnus formula provides a very good approximation 12 13 e s T 6 1094 exp 17 625 T T 243 04 displaystyle e s T 6 1094 exp left frac 17 625T T 243 04 right nbsp where e s displaystyle e s nbsp is in hPa and T displaystyle T nbsp is in degrees Celsius whereas everywhere else on this page T displaystyle T nbsp is an absolute temperature e g in kelvins This is also sometimes called the Magnus or Magnus Tetens approximation though this attribution is historically inaccurate 14 But see also the discussion of the accuracy of different approximating formulae for saturation vapour pressure of water Under typical atmospheric conditions the denominator of the exponent depends weakly on T displaystyle T nbsp for which the unit is degree Celsius Therefore the August Roche Magnus equation implies that saturation water vapor pressure changes approximately exponentially with temperature under typical atmospheric conditions and hence the water holding capacity of the atmosphere increases by about 7 for every 1 C rise in temperature 15 Example EditOne of the uses of this equation is to determine if a phase transition will occur in a given situation Consider the question of how much pressure is needed to melt ice at a temperature D T displaystyle Delta T nbsp below 0 C Note that water is unusual in that its change in volume upon melting is negative We can assume D P L T D v D T displaystyle Delta P frac L T Delta v Delta T nbsp and substituting in L 3 34 10 5 J kg displaystyle L 3 34 times 10 5 text J text kg nbsp latent heat of fusion for water T 273 displaystyle T 273 nbsp K absolute temperature D v 9 05 10 5 m 3 kg displaystyle Delta v 9 05 times 10 5 text m 3 text kg nbsp change in specific volume from solid to liquid we obtain D P D T 13 5 MPa K displaystyle frac Delta P Delta T 13 5 text MPa text K nbsp To provide a rough example of how much pressure this is to melt ice at 7 C the temperature many ice skating rinks are set at would require balancing a small car mass 1000 kg 16 on a thimble area 1 cm2 This shows that ice skating cannot be simply explained by pressure caused melting point depression and in fact the mechanism is quite complex 17 Second derivative EditWhile the Clausius Clapeyron relation gives the slope of the coexistence curve it does not provide any information about its curvature or second derivative The second derivative of the coexistence curve of phases 1 and 2 is given by 18 d 2 P d T 2 1 v 2 v 1 c p 2 c p 1 T 2 v 2 a 2 v 1 a 1 d P d T 1 v 2 v 1 v 2 k T 2 v 1 k T 1 d P d T 2 displaystyle begin aligned frac mathrm d 2 P mathrm d T 2 amp frac 1 v 2 v 1 left frac c p2 c p1 T 2 v 2 alpha 2 v 1 alpha 1 frac mathrm d P mathrm d T right amp frac 1 v 2 v 1 left v 2 kappa T2 v 1 kappa T1 left frac mathrm d P mathrm d T right 2 right end aligned nbsp where subscripts 1 and 2 denote the different phases c p displaystyle c p nbsp is the specific heat capacity at constant pressure a 1 v d v d T P displaystyle alpha 1 v mathrm d v mathrm d T P nbsp is the thermal expansion coefficient and k T 1 v d v d P T displaystyle kappa T 1 v mathrm d v mathrm d P T nbsp is the isothermal compressibility See also EditVan t Hoff equation Antoine equation Lee Kesler methodReferences Edit Clausius R 1850 Ueber die bewegende Kraft der Warme und die Gesetze welche sich daraus fur die Warmelehre selbst ableiten lassen On the motive power of heat and the laws which can be deduced therefrom regarding the theory of heat Annalen der Physik in German 155 4 500 524 Bibcode 1850AnP 155 500C doi 10 1002 andp 18501550403 hdl 2027 uc1 b242250 Clapeyron M C 1834 Memoire sur la puissance motrice de la chaleur Journal de l Ecole polytechnique fr in French 23 153 190 ark 12148 bpt6k4336791 f157 Pippard Alfred B 1981 Elements of classical thermodynamics for advanced students of physics Repr ed Cambridge Univ Pr p 116 ISBN 978 0 521 09101 5 Koziol Andrea Perkins Dexter Teaching Phase Equilibria serc carleton edu Carleton University Retrieved 1 February 2023 a b c d e f g h i Wark Kenneth 1988 1966 Generalized Thermodynamic Relationships Thermodynamics 5th ed New York NY McGraw Hill Inc ISBN 978 0 07 068286 3 Clausius Clapeyron The Clausius Clapeyron Equation Bodner Research Web Purdue University Retrieved 1 February 2023 Carl Rod Nave 2006 PvT Surface for a Substance which Contracts Upon Freezing HyperPhysics Georgia State University Retrieved 2007 10 16 a b Cengel Yunus A Boles Michael A 1998 1989 Thermodynamics An Engineering Approach McGraw Hill Series in Mechanical Engineering 3rd ed Boston MA McGraw Hill ISBN 978 0 07 011927 7 Salzman William R 2001 08 21 Clapeyron and Clausius Clapeyron Equations Chemical Thermodynamics University of Arizona Archived from the original on 2007 06 07 Retrieved 2007 10 11 Masterton William L Hurley Cecile N 2008 Chemistry principles and reactions 6th ed Cengage Learning p 230 ISBN 9780495126713 Retrieved 3 April 2020 Clapeyron E 1834 Memoire sur la puissance motrice de la chaleur Journal de l Ecole Polytechnique XIV 153 190 Alduchov Oleg Eskridge Robert 1997 11 01 Improved Magnus Form Approximation of Saturation Vapor Pressure NOAA doi 10 2172 548871 Equation 25 provides these coefficients Alduchov Oleg A Eskridge Robert E 1996 Improved Magnus Form Approximation of Saturation Vapor Pressure Journal of Applied Meteorology 35 4 601 609 Bibcode 1996JApMe 35 601A doi 10 1175 1520 0450 1996 035 lt 0601 IMFAOS gt 2 0 CO 2 Equation 21 provides these coefficients Lawrence M G 2005 The Relationship between Relative Humidity and the Dewpoint Temperature in Moist Air A Simple Conversion and Applications PDF Bulletin of the American Meteorological Society 86 2 225 233 Bibcode 2005BAMS 86 225L doi 10 1175 BAMS 86 2 225 IPCC Climate Change 2007 Working Group I The Physical Science Basis FAQ 3 2 How is Precipitation Changing Archived 2018 11 02 at the Wayback Machine Zorina Yana 2000 Mass of a Car The Physics Factbook Liefferink Rinse W Hsia Feng Chun Weber Bart Bonn Daniel 2021 02 08 Friction on Ice How Temperature Pressure and Speed Control the Slipperiness of Ice Physical Review X 11 1 011025 doi 10 1103 PhysRevX 11 011025 Krafcik Matthew Sanchez Velasco Eduardo 2014 Beyond Clausius Clapeyron Determining the second derivative of a first order phase transition line American Journal of Physics 82 4 301 305 Bibcode 2014AmJPh 82 301K doi 10 1119 1 4858403 Bibliography EditYau M K Rogers R R 1989 Short Course in Cloud Physics 3rd ed Butterworth Heinemann ISBN 978 0 7506 3215 7 Iribarne J V Godson W L 2013 4 Water Air systems 4 8 Clausius Clapeyron Equation Atmospheric Thermodynamics Springer pp 60 ISBN 978 94 010 2642 0 Callen H B 1985 Thermodynamics and an Introduction to Thermostatistics Wiley ISBN 978 0 471 86256 7 Notes Edit In the original work 1 T displaystyle 1 T nbsp was simply called the Carnot function and was not known in this form Clausius determined the form 30 years later and added his name to the eponymous Clausius Clapeyron relation Retrieved from https en wikipedia org w index php title Clausius Clapeyron relation amp oldid 1175021803, wikipedia, wiki, book, books, library,

article

, read, download, free, free download, mp3, video, mp4, 3gp, jpg, jpeg, gif, png, picture, music, song, movie, book, game, games.