You are on page 1of 6

Beyond Clausius–Clapeyron: Determining the second derivative of a first-order phase

transition line
Matthew Krafcik and Eduardo Sánchez Velasco

Citation: American Journal of Physics 82, 301 (2014); doi: 10.1119/1.4858403


View online: http://dx.doi.org/10.1119/1.4858403
View Table of Contents: http://scitation.aip.org/content/aapt/journal/ajp/82/4?ver=pdfcov
Published by the American Association of Physics Teachers

Articles you may be interested in


Conserving energy in physics and society: Creating an integrated model of energy and the second law of
thermodynamics
AIP Conf. Proc. 1513, 114 (2013); 10.1063/1.4789665

Relaxation and Prigogine–Defay ratio of compressed glasses with negative viscosity-pressure dependence
J. Chem. Phys. 130, 204506 (2009); 10.1063/1.3141382

Computational study of the melting-freezing transition in the quantum hard-sphere system for intermediate
densities. I. Thermodynamic results
J. Chem. Phys. 126, 164508 (2007); 10.1063/1.2718523

Determination of the entropy changes in the compounds with a first-order magnetic transition
Appl. Phys. Lett. 90, 032507 (2007); 10.1063/1.2425033

Thermodynamic functions of water and ice confined to 2 nm radius pores


J. Chem. Phys. 122, 104712 (2005); 10.1063/1.1862244

This article is copyrighted as indicated in the article. Reuse of AAPT content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:
68.37.176.114 On: Tue, 08 Apr 2014 23:12:37
Beyond Clausius–Clapeyron: Determining the second derivative
of a first-order phase transition line
Matthew Krafcika) and Eduardo Sa
nchez Velascob)
Department of Physics, Truman State University, Kirksville, Missouri 63501
(Received 19 August 2013; accepted 12 December 2013)
We obtain an expression for the second derivative of the line in a PT diagram denoting a first-order
phase transition for a pure hydrostatic system. Our result goes beyond the classical
Clausius–Clapeyron equation, which provides only the first derivative of the pressure with respect to
the temperature along the transition line. We present two pedagogical derivations suitable for an
undergraduate thermodynamics class; the first one uses derivatives of the entropy while the second one
uses derivatives of the enthalpy. The final expression for the second derivative involves only standard
thermodynamic quantities such as the specific heats, the isothermal compressibilities, and the
coefficients of thermal expansion of the two phases at the transition line. As an illustration, we
compute the second derivatives of the freezing and vaporization lines of water at atmospheric pressure,
and show that at this pressure the freezing line is concave down (negative second derivative) while the
vaporization line is concave up (positive second derivative). VC 2014 American Association of Physics Teachers.
[http://dx.doi.org/10.1119/1.4858403]

I. INTRODUCTION dP s2  s1
¼ : (4)
dT v2  v1
First-order phase transitions for pure substances subject
only to pressure forces (sometimes called1 pure hydrostatic
systems), such as melting or vaporization, are usually repre- The Clausius–Clapeyron equation is remarkable because it
sented as lines in a PT diagram. Let us consider a transition allows one to approximate linearly the transition line using
line in a PT diagram that separates a low-temperature phase, only thermodynamic data from a single point. Interestingly,
which we will call phase 1, from a high-temperature phase, it is also possible to compute the second derivative of the
which we will call phase 2 (see Fig. 1). transition line using only standard thermodynamic data—iso-
In this paper, we are interested in obtaining local properties baric specific heat, isothermal compressibility, and coeffi-
of the transition line at a particular point in the curve (for cient of thermal expansion—from that point on the curve.
example, the circled point in Fig. 1). At this point, the slope of Knowledge of the second derivative provides a clearer pic-
the transition line is given by the Clausius–Clapeyron ture of the local properties of the transition line at the point
equation,2–14 which uses only standard thermodynamic prop- of interest; it gives information about the curvature and con-
erties of the system. The Clausius–Clapeyron equation can be cavity of the transition line at that point, and allows one to
written in several forms, the most common of which is extend the Taylor expansion of the transition line to second
order. There are several equations that relate the second de-
dP l12 rivative of the transition line to other thermodynamic quanti-
¼ : (1) ties, but most of them involve non-standard thermodynamic
dT Tðv2  v1 Þ
data15 or are only valid in the vicinity of the critical
Here, v1 and v2 are the specific volumes in each phase point.16,17 We could find only one reference, Ref. 18, that
and l12 is the heat of transformation, or latent heat mentions the identity we have in mind, but it does not offer a
(the amount of energy per unit of substance needed to go derivation of the result. In addition, we could not find any
from the low-temperature phase to the high-temperature textbook containing a discussion of our result. We hope that
phase). our paper will fill this gap while providing a derivation sim-
Because the transition from one phase to the other ple enough to be suitable for an undergraduate thermody-
happens at constant pressure and temperature, the heat of namics class as an illustration of the standard techniques
transformation l12 is equal to the difference between the used to transform thermodynamic identities.19–23
specific enthalpies h2 and h1 of the two phases; it is also Our starting point will be the Clausius–Clapeyron equa-
equal to the absolute temperature times the difference tion, from which we obtain our main result: an expression
between the specific entropies s2 and s1 of the two phases, for the second derivative of the PT curve. We provide two
so that different derivations of our final result. The first derivation,
presented in Sec. II, uses the Clausius–Clapeyron equation in
l12 ¼ h2  h1 ¼ Tðs2  s1 Þ: (2) the form of Eq. (4); the second one, presented in Sec. III,
uses the Clausius–Clapeyron equation in the form of Eq. (3).
Using Eq. (2), one can write the Clausius–Clapeyron equa- Although both derivations are straightforward, we believe it
tion in terms of the discontinuity of the specific enthalpy, is pedagogically useful to present, when possible, alternative
approaches to a given problem. In Sec. IV, we provide a sim-
dP h2  h1 ple application of our expression by applying it to the freez-
¼ ; (3)
dT Tðv2  v1 Þ ing and vaporization of water at ordinary pressure.
In the derivations of Secs. II and III, we will use molar
or in terms of the discontinuity of the specific entropy, specific quantities, such as hi, si, and vi (i ¼ 1,2), although

301 Am. J. Phys. 82 (4), April 2014 http://aapt.org/ajp C 2014 American Association of Physics Teachers
V 301

This article is copyrighted as indicated in the article. Reuse of AAPT content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:
68.37.176.114 On: Tue, 08 Apr 2014 23:12:37
  
dsi @si @si dP
¼ þ ; (8a)
dT @T P @P T dT
   
dvi @vi @vi dP
¼ þ : (8b)
dT @T P @P T dT

Our next task is to reduce the four partial derivatives that


appear on the right-hand sides of these two equations into
standard thermodynamic quantities
Starting with the first partial derivative in Eq. (8a), we
have
     
@si 1 @Si 1 T @Si cpi
¼ ¼ ¼ ; (9)
@T P n @T P;n T n @T P;n T

where cpi is the molar specific heat at constant pressure (iso-


baric specific heat) in each phase. The second partial deriva-
tive in Eq. (8a) is
Fig. 1. A schematic drawing of a phase transition curve separating a low-
temperature phase (phase 1) from a high-temperature phase (phase 2). The    
circled point is where we want to obtain local properties of the transition @si 1 @Si
¼ : (10)
curve. @P T n @P T;n

Here the derivative of Si with respect to P can be reduced


one could use quantities per unit mass and obtain the same by using one of Maxwell’s relations,21 obtained from the
results; the ratio of two specific quantities per unit mass is differential of the Gibbs free energy dGi ¼ Si dT
the same as the ratio of the same quantities per mole. In Sec. þ Vi dP þ li dn,
IV, it will be more practical to use quantities per unit mass.    
In each phase (not only along the transition line), we will use @Si @Vi
¼ ¼ Vi bi ; (11)
as independent variables the temperature T, pressure P, and @P T;n @T P;n
the number of moles n. Intensive quantities will be functions
of only the temperature and pressure. The entropy Si, en- where
thalpy Hi, and volume Vi in each phase are extensive quanti-  
ties and are given as functions of T, P, and n by 1 @Vi
bi ¼ (12)
Vi @T P;n
Si ðT; P; nÞ ¼ nsi ðT; PÞ; (5a)
Hi ðT; P; nÞ ¼ nhi ðT; PÞ; (5b) is the coefficient of thermal expansion in phase i.
Substituting Eq. (11) into Eq. (10), we then get
Vi ðT; P; nÞ ¼ nvi ðT; PÞ: (5c)  
@si Vi
¼  bi ¼ vi bi : (13)
@P T n
II. DERIVATION USING THE ENTROPY
DISCONTINUITY Next, we reduce the partial derivatives on the right-hand-
side of Eq. (8b). Beginning with the first, we have
Our starting point is the Clausius–Clapeyron equation in      
terms of the discontinuity of the specific entropy, Eq. (4). @vi 1 @Vi Vi 1 @Vi
Taking the derivative of that equation with respect to T along ¼ ¼ ¼ vi bi : (14)
@T P n @T P;n n Vi @T P;n
the transition line, we get
Finally, the second partial derivative in Eq. (8b) is
d2 P 1 d s2  s1 d
¼ ðs2  s1 Þ  ðv2  v1 Þ:
dT 2 v2  v1 dT ðv2  v1 Þ2 dT      
@vi 1 @Vi Vi 1 @Vi
(6) ¼ ¼ ¼ vi jTi ;
@P T n @P T;n n Vi @P T;n
In the last term, ðs2  s1 Þ=ðv2  v1 Þ is precisely the slope of (15)
the transition line given by the Clausius–Clapeyron equation;
therefore, we can write where
 
d2 P 1 d 1 dP d 1 @Vi
2
¼ ðs2  s1 Þ  ðv2  v1 Þ: jTi ¼  (16)
dT v2  v1 dT v2  v1 dT dT Vi @P T;n
(7)
is the isothermal compressibility in phase i.
To evaluate the derivatives on the right-hand side, we note Substituting Eqs. (9), (13)–(15) into Eqs. (8a) and (8b), we
that along the transition line (for either phase) we have get

302 Am. J. Phys., Vol. 82, No. 4, April 2014 M. Krafcik and E. Sanchez Velasco 302

This article is copyrighted as indicated in the article. Reuse of AAPT content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:
68.37.176.114 On: Tue, 08 Apr 2014 23:12:37
dsi cpi dP From the differential of the enthalpy dHi ¼ TdSi þ Vi dP
¼  vi bi ; (17a)
dT T dT þ li dn, we have
dvi dP    
¼ vi bi  vi jTi : (17b) @Hi @Si
dT dT ¼T þ Vi : (24)
@P T;n @P T;n
Then, using these expressions in Eq. (7) and rearranging
terms, we arrive at our final result In this equation, the derivative of Si with respect to P can be
rewritten using the same Maxwell relation we obtained in
d2 P cp2  cp1 2ðv2 b2  v1 b1 Þ dP Eq. (11), giving
2
¼   
dT Tðv2  v1 Þ v2  v1 dT @Hi
 2 ¼ TVi bi þ Vi ¼ Vi ð1  Tbi Þ: (25)
v2 jT2  v1 jT1 dP @P T;n
þ : (18)
v2  v1 dT
Substituting this result into Eq. (23) leads to
 
III. DERIVATION USING THE ENTHALPY @hi Vi
¼ ð1  Tbi Þ ¼ vi ð1  Tbi Þ; (26)
DISCONTINUITY @P T;n n

The procedure here is similar to the one in the previous sec- and using Eqs. (22) and (26) in Eq. (21) gives
tion, but we start with the Clausius–Clapeyron equation writ-
ten in terms of the enthalpy discontinuity, Eq. (3). Taking its dhi dP
derivative with respect to T along the transition line, we get ¼ cpi þ vi ð1  Tbi Þ : (27)
dT dT
d2 P h2  h1 1 d Finally, substituting Eqs. (17b) and (27) into Eq. (20) we get,
¼ 2 þ ðh2  h1 Þ after rearranging terms, the same result obtained Eq. (18).
dT 2 T ðv2  v1 Þ Tðv2  v1 Þ dT
h2  h1 d
 ðv2  v1 Þ: (19) IV. APPLICATION TO THE FREEZING AND
Tðv2  v1 Þ2 dT VAPORIZATION OF WATER
In the first and last terms on the right-hand side, we use the As an illustration of the final result, Eq. (18), we will con-
Clausius–Clapeyron equation to get sider the cases of freezing and vaporization of water at ordi-
nary pressure (P ¼ 1 atm ¼ 101, 325 Pa), which occur at the
d2 P 1 dP 1 d temperatures 273.15 K (0  C) and 373.15 K (100  C),
2
¼ þ ðh2  h1 Þ respectively.
dT T dT Tðv2  v1 Þ dT
First we consider freezing, with ice as phase 1 and liquid
1 dP d
 ðv2  v1 Þ: (20) water as phase 2. The heat of transformation for melting (l12)
v2  v1 dT dT and the densities of ice and liquid water (q1 ; q2 ) at 0  C
are27,28
The derivative of vi with respect to T along the transition line
that appears in the last term of this equation was written in l12 ¼ 333:6 kJ=kg; q1 ¼ 916:7 kg=m3 ;
terms of standard thermodynamic quantities in the previous
section [Eq. (17b)]. What remains is to do the same for the q2 ¼ 999:84 kg=m3 : (28)
derivative of hi with respect to T along the transition line,
    The volume per kilogram in each phase is simply the inverse
dhi @hi @hi dP of the corresponding density (vi ¼ 1=qi ). With these values,
¼ þ : (21) the Clausius–Clapeyron equation gives a slope of
dT @T P @P T dT
dP
We will need to reduce the two partial derivatives that ¼ 1:347  107 Pa=K : (29)
appear on the right-hand side of Eq. (21) into standard ther- dT
modynamic quantities
Two things are noteworthy about this slope. The first is that it
At constant pressure, the thermal energy given to a system
is negative, which is due to the fact that the density of ice is
is the same as the change in enthalpy. Therefore, the first par-
smaller than the density of liquid water (ice floats). The sec-
tial derivative on the right-hand-side of Eq. (21) can be written
ond is that it is a very steep slope. In the 0.01  C temperature
directly in terms of the specific heat at constant pressure,24–26
interval that separates the melting point of ice at 1 atm from
    the triple point of water (Ptp ¼ 611:73 Pa ¼ 0:0060373 atm),
@hi 1 @Hi
¼ ¼ cpi : (22) this slope corresponds to a drop in pressure of more than 1
@T P n @T P;n atm (by some 33.4 kPa).
To find the second derivative using Eq. (18), we need the
The second partial derivative on the right-hand-side of values of the isobaric specific heats, isothermal compressibil-
Eq. (21) is ities, and coefficients of thermal expansion for both
    phases,27–32
@hi 1 @Hi
¼ : (23)
@P T n @P T;n cp1 ¼ 2:108 kJ=kg K; cp2 ¼4:2194 kJ=kg K; (30a)

303 Am. J. Phys., Vol. 82, No. 4, April 2014 M. Krafcik and E. Sanchez Velasco 303

This article is copyrighted as indicated in the article. Reuse of AAPT content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:
68.37.176.114 On: Tue, 08 Apr 2014 23:12:37
b1 ¼ 1:59  104 K1 ; b2 ¼ 0:677  104 K1 ;
(30b)
jT1 ¼ 1:3  1010 Pa1 ; jT2 ¼ 5:086  1010 Pa1 :
(30c)

Using this data in Eq. (18), we find that the first and last
terms give negative contributions, and the second term a
positive contribution, with the last term being dominant. The
final value of the second derivative is thus negative so that
the melting curve is concave down with a value of

d2 P
¼ 7:5  105 Pa=K2 : (31)
dT 2
Next we consider the vaporization of water, also at 1 atm.
In this case, phase 1 will be the liquid and phase 2 will be
the vapor. The heat of transformation for vaporization (l12) Fig. 2. Comparison of the linear (dashed) and quadratic (solid) terms of a
and the densities of the liquid and the vapor (q1 ; q2 ) at Taylor expansion around 100  C (circled point) for the vaporization curve
100  C are27,28,33 for water. The dots represent experimental data.

l12 ¼ 2256:472 kJ=kg; q1 ¼ 958:367 kg=m3 ; entire vaporization curve. Rather, we just want to look at the
q2 ¼ 0:5977 kg=m3 ; (32) local properties of the curve at that pressure and temperature.
For the entire curve, the usual approach13 of neglecting the
and using the Clausius–Clapeyron equation, we obtain the specific volume of the liquid, treating the vapor as an ideal
slope gas, and integrating the Clausius–Clapeyron equation by fur-
ther assuming that the heat of transformation l12 is constant,
dP gives a much better approximation.37
¼ 3617 Pa=K: (33) Figure 2 shows the first terms of a Taylor expansion of the
dT
vaporization curve around 100  C (circled point). The graph
To compute the second derivative using Eq. (18), we shows the linear approximation using the slope obtained
need the values of the isobaric specific heats, isothermal from the Clausius–Clapeyron equation [Eq. (31)] and the
compressibilities, and coefficients of thermal expansion for quadratic approximation by using the value of the second de-
both phases.27,29,33,34 For water vapor, we could not find rivative given in Eq. (35). The dots in the figure represent
any tables that list the coefficient of thermal expansion. the experimental values of the vaporization curve.38 As
However, a reasonable approximation is to use the coeffi- expected, the quadratic approximation provides a much bet-
cient of thermal expansion for a van der Waals gas35 with ter fit to the experimental data.
the appropriate coefficients for water vapor.36 We then
have
V. SUMMARY
cp1 ¼ 4:2156 kJ=kgK; cp2 ¼ 2:0799 kJ=kgK; (34a)
By using standard techniques to transform thermodynamic
b1 ¼ 7:52  10 4 1
K ; identities, we have presented two different methods to obtain
3
the second derivative of a first-order phase transition line in
b2 ¼ 2:74  10 K1 ðvan der WaalsÞ; (34b) a PT diagram at a particular transition point. The second de-
rivative allows for a better understanding of the local proper-
jT1 ¼ 4:902  1010 Pa1 ; jT2 ¼ 1:002  105 Pa1 : ties of the curve at that point, including its curvature and
(34c) concavity, and gives a better idea of what the curve looks
like beyond what is provided by the Clausius–Clapeyron
For the vaporization of water, the first and second terms in equation alone. The final expression for the second deriva-
Eq. (18) are negative, while the last term, which is by far the tive, Eq. (18), uses only standard thermodynamic properties
dominant one, is positive (its size is about 37 times the mag- of the substance at the transition point, and can be obtained
nitude of the first term, and about six times the magnitude of from either the entropy or the enthalpy formulation of the
the second term). Thus, the second derivative is positive, so Clausius–Clapeyron equation. The derivation of Eq. (18)
the vaporization curve is concave up with a value of does not require advanced mathematical techniques and it is
suitable for presentation in a typical undergraduate thermo-
d2 P dynamics course.
¼ 108 Pa=K2 : (35) As an application of the final result, we analyzed the freez-
dT 2
ing and vaporization curves of water at atmospheric pressure.
To illustrate the local properties of the vaporization line We found a negative second derivative for the freezing
that we can obtain from the first and second derivatives of curve, indicating that the curve is concave down at that
the curve, we can compare the linear and quadratic Taylor point; for the vaporization of water we found a positive sec-
expansions of the curve at the boiling point with experimen- ond derivative, indicating that the curve is concave up. We
tal data. Note that we are not trying to find a good fit to the illustrated the local properties of the vaporization line by

304 Am. J. Phys., Vol. 82, No. 4, April 2014 M. Krafcik and E. Sanchez Velasco 304

This article is copyrighted as indicated in the article. Reuse of AAPT content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:
68.37.176.114 On: Tue, 08 Apr 2014 23:12:37
18
comparing the second-order Taylor expansion of the curve W. Wagner, A. Saul, and A. Pruß, “International equations for the pressure
with experimental data. along the melting and along the sublimation curve of ordinary water sub-
stance,” J. Phys. Chem. Ref. Data 23(3), 515–527 (1994).
19
F. Reif, Fundamentals of Statistical and Thermal Physics (McGraw-Hill,
ACKNOWLEDGMENT New York, 1965), pp. 161–184.
20
M. W. Zemansky and R. H. Dittman, Heat and Thermodynamics, 7th edi-
The authors are grateful to one of the reviewers of an ear- tion (McGraw-Hill, New York, 1997), pp. 249–276.
lier version of this manuscript for pointing out a numerical 21
H. B. Callen, Thermodynamics and an Introduction to Thermostatics, 2nd
error in the calculation of the second derivative of the freez- ed. (John Wiley & Sons, New York, 1985), pp. 181–202.
22
ing curve of water. W. Greiner, L. Neise, and H. St€ ocker, Thermodynamics and Statistical
Mechanics (Springer-Verlag, Berlin, 1995), pp. 108–118.
23
E. Sanchez Velasco, Fundamentals of Thermodynamics and Statistical
a)
Electronic mail: mkrafcik@purdue.edu; Present address: Purdue Mechanics—Lecture Notes, 2nd ed. (CreateSpace, 2010), pp. 235–246.
24
University, School of Materials Engineering, Neil Armstrong Hall of E. Sanchez Velasco, Fundamentals of Thermodynamics and Statistical
Engineering, 701 West Stadium Avenue, West Lafayette, Indiana 47907- Mechanics—Lecture Notes, 2nd ed. (CreateSpace, 2010), pp. 109–110.
25
2045. C. Kittel and H. Kroemer, Thermal Physics, 2nd ed. (W. H. Freeman and
b)
Electronic mail: velasco@truman.edu Company, New York, 1980), pp. 284–285.
1 26
M. W. Zemansky and R. H. Dittman, Heat and Thermodynamics, 7th ed. M. W. Zemansky and R. H. Dittman, Heat and Thermodynamics, 7th edi-
(McGraw-Hill, New York, 1997), pp. 32–35. tion (McGraw-Hill, New York, 1997), p. 255.
2 27
Some authors, such as Callen, and Dondepudi and Prigogine, refer to the W. Wagner and A. Pruß, “The IAPWS formulation 1995 for the thermo-
Clausius–Clapeyron equation as just the Clapeyron equation, reserving the dynamic properties of ordinary water substance for general and
name Clausius–Clapeyron for an approximate expression of the equation scientific use,” J. Phys. Chem. Ref. Data 31(2), 387–535 (2002), avail-
for the vaporization line, a common naming convention in the chemistry able online at <http://www.nist.gov/data/PDFfiles/jpcrd617.pdf>; see
literature. Other authors, such as Landau and Lifshitz, and Reiss, refer to pp. 497–498.
28
the Clausius–Clapeyron equation as the Clapeyron–Clausius equation. We D. R. Lide, Handbook of Chemistry and Physics, 89th ed. (CRC Press,
opted for the name Clausius–Clapeyron equation because it seems to be Boca Raton, 2008), pp. 6-4–6-8.
29
the most commonly used in the physics literature. Water—Thermal properties, Data located in a table on The Engineering
3
D. V. Schroeder, An Introduction to Thermal Physics (Addison-Wesley, ToolBox website, <http://www.engineeringtoolbox.com/water-thermal-
Reading, 2000), pp. 172–179. properties-d_162.html>.
4 30
D. Kondepudi and I. Prigogine, Modern Thermodynamics: From Heat Table II—Physical properties of ice Ih at 0  C. This table can be found on
Engines to Dissipative Structures (John Wiley & Sons, New York, 1998), Caltech’s website at <http://glaciology.caltech.edu/ice.table2.html>.
31
pp. 178–180. Water Thermal Expansion Coefficient, <http://physchem.kfunigraz.ac.at/
5
H. Reiss, Methods of Thermodynamics (Dover, New York, 1996), pp. 137–139. sm/Service/Water/H2Othermexp.htm>.
6 32
F. Reif, Fundamentals of Statistical and Thermal Physics (McGraw-Hill, Water isothermal compressibility, <http://physchem.kfunigraz.ac.at/sm/
New York, 1965), pp. 304–306. service/water/H2Obetat.htm>.
7 33
L. E. Reichl, A Modern Course in Statistical Physics, 2nd ed. (John Wiley Table of properties of water at varying temperatures and saturation pres-
& Sons, New York, 1998), pp. 105–110. sure, RoyMech web site, <http://www.roymech.co.uk/Related/Fluids/
8
M. W. Zemansky and R. H. Dittman, Heat and Thermodynamics, 7th edi- Fluids_Water_Props.html>.
34
tion (McGraw-Hill, New York, 1997), pp. 286–293. I. S. Grigoriev and E. Z. Meilikhov, Handbook of Physical Quantities
9
G. Carrington, Basic Thermodynamics (Oxford U. P., New York, 1994), (CRC Press, Boca Raton, 1997), p. 271.
35
pp. 294–295. E. Sanchez Velasco, Fundamentals of Thermodynamics and Statistical
10
H. B. Callen, Thermodynamics and an Introduction to Thermostatics, 2nd Mechanics—Lecture Notes, 2nd ed. (CreateSpace, 2010), p. 81.
36
ed. (John Wiley & Sons, New York, 1985), pp. 228–233. D. R. Lide, Handbook of Chemistry and Physics, 89th ed. (CRC Press,
11
W. Greiner, L. Neise, and H. St€ ocker, Thermodynamics and Statistical Boca Raton, 2008), pp. 6–36.
37
Mechanics (Springer-Verlag, Berlin, 1995), pp. 64–66. Similar approximations can be made to Eq. (18). If, in addition to neglect-
12
L. D. Landau and E. M. Lifshitz, Statistical Physics Part 1, 3rd ed. ing the specific volume of the liquid and approximating the vapor as an
(Pergamon Press, Oxford, 1980), pp. 255–256. ideal gas, it is further assumed that the discontinuity in the specific heats
13
C. Kittel and H. Kroemer, Thermal Physics, 2nd ed. (W.H. Freeman and across the transition is constant (instead of assuming constant heat of
Company, New York, 1980), pp. 276–283. transformation), then Eq. (18) can be integrated and provides for water a
14
R. K. Pathria and P. D. Beale, Statistical Mechanics, 3rd ed. (Elsevier, better approximation to the vaporization curve than the usual approach.
Amsterdam, 2011), pp 109–110. This approximation is equivalent to assuming in the original
15
V. V. Sychev, The Differential Equations of Thermodynamics (Mir, Clausius–Clapeyron equation that the latent heat is not constant, but has a
Moscow, 1983), pp. 186–188. linear dependence in T given by l12 ¼ l12;0 þ ðcp2  cp1 ÞðT  T0 Þ. Of
16
S. W. Van Sciver, Helium Cryogenics, 2nd ed. (Springer-Verlag, Berlin, course, these types of approximations cannot be used for solid-liquid or
2012), p. 178. solid-solid transitions.
17 38
H. Eugene Stanley, Introduction to Phase Transitions and Critical I. S. Grigoriev and E. Z. Meilikhov, Handbook of Physical Quantities
Phenomena (Oxford U. P., New York, 1971), p. 51. (CRC Press, Boca Raton, 1997), pp. 323–324.

305 Am. J. Phys., Vol. 82, No. 4, April 2014 M. Krafcik and E. Sanchez Velasco 305

This article is copyrighted as indicated in the article. Reuse of AAPT content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:
68.37.176.114 On: Tue, 08 Apr 2014 23:12:37

You might also like