You are on page 1of 6

ARTICLE

pubs.acs.org/jchemeduc

Complexities of One-Component Phase Diagrams


Andrea Ciccioli and Leslie Glasser*,

Department of Chemistry, Sapienza - University of Rome, p.le Aldo Moro 5 00185 Roma, Italy

Nanochemistry Research Institute, Department of Applied Chemistry, Curtin University, GPO Box U1987,
Perth, WA 6845, Australia

ABSTRACT: For most materials, the solid at and near the triple-point
temperature is denser than the liquid with which it is in equilibrium.
However, for water and certain other materials, the densities of the
phases are reversed, with the solid being less dense. The profound
consequences for the appearance of the pVT diagram of one-component
materials resulting from such anomalous volume changes in solid-
liquid transitions are discussed. We discuss and illustrate how the 3D
pVT phase diagram changes for this case. A more complex case occurs in
systems where the solid liquid eld displays continuous density
reversal at high pressure, making the phase diagrams of some elements
unexpectedly complex. The controversial case of graphite is presented as
an example of the diculties of interpretation. A current version of the
carbon pT phase diagram is provided, in a 2D pT representation as well as
in a virtual 3D version. The phase diagram of sodium, newly determined
to extremely high pressures and illustrated here, shows both melting
maxima and minima as well as a number of phase transitions as pressure
increases.
KEYWORDS: Upper-Division Undergraduate, Graduate Education/Research, Physical Chemistry, Materials Science, Phases/
Phase Transitions/Diagrams, Thermodynamics

STANDARD AND ANOMALOUS VOLUME


P hase diagrams depict the relations among the various phases
of a material, generally within the space of pressure, volume,
and temperature (pVT); that is, phase diagrams depict the
CHANGES UNDER MELTING
To lend clarity to the complexity of the region around the
equilibrium pVT conditions under which phase transitions occur. triple point, where solid, liquid, and vapor or gas are in
Since pVT is a space of three dimensions, it is most appropriate to equilibrium, it has been customary to oer diagrams of general-
display the diagram either as a physical model or in virtual 3D ized models, such as that in Figure 1. Generalized models such as
space,1 but the diculties inherent in such constructions have led this depict the properties of standard materials, for which the
to a general preference for projections onto the pT, pV, and VT density of the solid is greater than that of the equilibrium liquid.
planes. Most introductory textbooks of physical chemistry tend However, water (together with germanium, silicon and carbon,3
to present thorough discussions of pT planes, where a clear which are also tetrahedrally coordinated as in ice, together with a
correspondence between degrees of freedom and dimensionality few other materials) is anomalous in that the solid, ice, is less
emerges. Owing to their nature as thermodynamic potentials, the dense than its liquid, water, at the triple-point temperature. This
p and T variables assume the same value in phases coexisting at results in such pVT diagrams as Figure 1 being nonrepresentative
equilibrium, and so invariant points, monovariant lines, and and potentially confusing.
bivariant areas are displayed in the pT plane. pV representations Consider now how the diagram in Figure 1 is altered in the
are more often presented as experimental isothermal condensa- anomalous density case. Thermodynamic phase transition beha-
tion curves (that is, isothermal sections of the 3D diagram), vior is described by the Clapeyron equation
whereas little attention is usually devoted to the VT plane. Recent  
dp trans H trans S
computational developments, however, permit extension to
interactive virtual 3D representation,2 making it easier for dT trans Ttrans trans V trans V
students to appreciate the shape of the pVT surfaces and their where transH, transS, and transV represent the enthalpy, en-
relation to 2D projections. In this article, we consider some of the tropy, and volume change, respectively, of an equilibrium phase
complexities that may occur in one-component phase diagrams. transition at the absolute temperature, Ttrans. This equation
These are seldom mentioned, even beyond the educational
literature, although they are of considerable chemical and
technical concern. Published: March 11, 2011

Copyright r 2011 American Chemical Society and


Division of Chemical Education, Inc. 586 dx.doi.org/10.1021/ed100638r | J. Chem. Educ. 2011, 88, 586591
Journal of Chemical Education ARTICLE

Figure 1. pVT phase diagram model of a generalized material, empha-


sizing the region (yellow) surrounding the triple-point line (the line
connecting the molar volumes of the three phases in equilibrium at the
triple point) and showing the usual case where the solid is denser than its
equilibrium liquid, so that the slope of the melting curve is positive
according to the Clapeyron equation (see text). Three surface types are
depicted in white: T, p, V (solid) surface; T, p, V (liquid) surface,
continuing into the T, p, V (vapor or gas) surface.

describes how pressure (temperature) changes as a result of


temperature (pressure) alterations when two phases are in mutual
equilibrium; that is, it describes the slope of the pressure-
temperature curve. At the equilibrium condition, transG = trans
H - TtranstransS = 0, hence, transS = transH/Ttrans.
Because Ttrans is necessarily positive, the sign of the pressure-
temperature slope (whether it is positive or negative) depends on
the ratio of the signs of the enthalpy (entropy) and volume
changes involved in the phase transition. More than a century
ago, Gustav Tammann4 discussed in detail all the possible cases
associated with the positive and negative sign of both numerator
and denominator in the Clapeyron equation, in melting and
amorphization processes, resulting in a loop-shaped melting line Figure 2. pVT phase diagram model of a generalized material, empha-
in the pT plane (also termed a Kauzmann curve5), with the solid sizing the region surrounding the triple-point line and showing the
inside and the liquid surrounding the loop.6 In principle at least, anomalous case where the liquid is denser than its equilibrium solid, so
melting processes with negative H and S values could not be that the slope of the melting curve is negative. (B) The temperature axis
ruled out. Although a few examples of exothermic melting have of the diagram is reversed with respect to panel A to better display the s
l vapor areas (and thus corresponds to a mirror image of panels A).
actually been presented in the literature in more recent times,7 in
the overwhelming majority of cases enthalpies (and entropies) of
materials increase in the sequence solid < liquid , gas as the to face the increasing temperature regime, whereas the phase
molecules in the phases become successively increasingly inde- equilibrium lines surrounding the gap now slope toward higher
pendent of each other. Hence, in a melting process, where solid is pressure as the temperature decreases, that is, they are reversed
converted to liquid, the enthalpy change is usually positive. For from the standard situation and are not easily visualized. To have
most materials, the volume change on melting is also positive the s l ribbon face the viewer for easier comprehension, the
(because densities generally decrease in the reversed sequence, temperature axis should be reversed, as shown in a mirror image
solid > liquid . gas), so that the pressure-temperature slope is version in Figure 2B.
positive. This results in the solid liquid two-phase ribbon in In contrast with early texts, discussion and presentation of
Figure 1 (labeled s l) facing the viewer. But ice is less dense similar 3D representations are seldom found in recent textbooks
than its equilibrium liquid at the triple-point temperature, so that of physical chemistry, the focus being now preferentially directed
the pressure-temperature slope for water (and the other toward pT projections only. However, these 3D diagrams have
above-mentioned anomalous materials) is negative. considerable value from a pedagogical point of view. Not only do
This reversed slope in the pT projection is simply the 2D they show the source of the pT projections depicted in the 2D
image of a profound change in the appearance of the full pVT phase diagram, but consideration of their isothermal and isobaric
phase diagram near the triple point (Figure 2A). The solid plus sections permits discussion of the role of the compressibility8 and
liquid phase gap in this circumstance has become twisted around, thermal expansion coecients (and their respective derivatives) in
587 dx.doi.org/10.1021/ed100638r |J. Chem. Educ. 2011, 88, 586591
Journal of Chemical Education ARTICLE

Figure 3. The conventional pT phase diagram of carbon, displaying a


maximum in the melt curve of graphite, shown as the broken convex
solid-liquid equilibrium curve. The diamond-liquid curve is shown
with a metastable extension into the graphite region. (Adapted from ref 3c.
See additional note for the data.) The solid near-vertical graphite melt
line is currently considered to be more plausible (see text).
Figure 4. pT phase diagram of sodium from ambient conditions to high
determining the slopes (and curvatures) of solid and liquid single pressures. As pressure increases, the crystal structure commences as
phase pVT domains. For example, the negative thermal expansion body-centerd cubic (bcc), passes through a melting-point maximum at
coecients of hexagonal ice and water in some temperature about 31 GPa and 990 K, and converts to face-centered cubic (fcc) at
ranges can be visualized on a properly expanded T scale. about 65 GPa and 700 K (with virtually no volume change; see Figure 6).
As pressure increases, the melting point decreases to about 300 K at 118
GPa, while the crystal structures change through a progression of
MELTING CURVES WITH MAXIMA AND MINIMA complex structures19a,c,d,21 (see text and Table 1).
For some materials, the phase diagrams can display interesting
behavior if the pressure scale is extended to higher pressures. 2, where P is the number of phases in thermodynamic equilibrium
This is particularly useful because students are often intrigued by and C is the number of independent components, determines that
the fate of the melting line at high pressure, especially as to beyond the critical point the one-phase uid system has two
whether it ends with a critical point, as is the case for the pT degrees of freedom, F = 2, to be represented on the phase diagram
liquid-vapor line (i.e., the liquid vapor eld in the 3D by a surface, whereas the liquid-gas curve terminates at a critical
diagram) or otherwise. Consider the situation when the solid- point with zero variance.9
liquid equilibrium curve turns over from a positive slope to a The unusual behavior of an apparent continuation for carbon
negative slope (as in the case of graphite in the current standard of the phase equilibrium curve through the melt maximum has
carbon phase diagram;3 broken convex curve in Figure 3) or vice evoked an alternative suggestion that there is actually an ob-
versa, from negative to positive. This corresponds to a maximum scured liquid-liquid phase transition (LLPT) at this point,
(or minimum when the slope changes from negative to positive) where one form of liquid carbon is converted into another,10
in the melt curve with pressure; for graphite, such a maximum is each in equilibrium with solid. If correct, then F = 1 - 3 2 = 0,
reported to occur at about 6 GPa and 4300 K, as depicted by the and this corresponds to an invariant, triple point. While this is an
broken convex curve in Figure 3.a unusual condition, phosphorus under pressure does undergo
This behavior may be read in the light of the Clapeyron such an LLPT transition.11
equation. Indeed, the change of sign in the slope of a solid-liquid High-pressure melt experiments are exceedingly dicult to
equilibrium curve can reect the fact that, with changing pressure perform, and the reported maximum for carbon has only been
and temperature along the curve, the volumes of the two phases found indirectly. Moreover, no direct studies have found this
change progressively to dierent extents, so that their volume invariant point for graphite melt, while careful simulations show a
dierence becomes zero at some point, and the dp/dT slope small positive slope throughout the melting range.12 It is currently
becomes innite (corresponding to division by zero in the considered that the melt maximum is an experimental artifact that
Clapeyron equation) or, equivalently, the dT/dp slope becomes should be rejected, whereas the graphite melt curve correctly has a
zero, corresponding to a maximum or minimum in the Tp plane. slight positive slope, as depicted by the solid line in Figure 3. This
With some caution, comparison may be made with the more would result in a standard pVT diagram in the high-pressure
familiar, but profoundly dierent, situation occurring at the range for carbon (cf. Figure 1). For diamond, other simulations
critical point where the liquid-vapor line terminates. Here, with have suggested a maximum in the melt curve13 while very recent
increasing temperature and pressure along the two-phase line, the experimental measurements14 indicate a negative Clapeyron
liquid and vapor molar volumes tend to approach each other and slope between 0.60 and 1.05 TPa with a complex, conducting
eventually become equal at the critical point. However, the uid state produced at higher pressures and temperatures.
analogy between the two situations begins and ends here because, There are a number of reports in the literature of other
at the critical point, the two phases merge into a single super- materials having maximum melting points with pressure, such
critical phase. Not only the volumes, but also the enthalpies as the early report by Bridgman on the transition between HgI2
assume the same value, so the Clapeyron slope in this case solids I and II15 and the melting behavior of several block s
becomes indeterminate (0/0). Gibbss phase rule, F = C - P (rubidium, cesium, barium) and block p (bismuth, tellurium)
588 dx.doi.org/10.1021/ed100638r |J. Chem. Educ. 2011, 88, 586591
Journal of Chemical Education ARTICLE

Figure 5. pVT phase diagram model of a generalized material, showing


the situation when the initial slope of the s l curve is positive, passes
through the vertical, and then becomes negative. This progressively
developing situation is shown as a rising ribbon, with width initially Figure 7. pVT phase diagram of sodium, solid and liquid only, from
decreasing to zero to the point of equality of solid and liquid molar ambient conditions to high pressures. The molar volume of the liquid
volumes, then expanding again as the solid molar volume progressively has been exaggerated for clarity of presentation. The broken line
exceeds that of the liquid. represents the transition between the bcc and fcc crystalline phases;
the dots represent further crystalline phase transitions into a series of
complex solids. The pressure ranges from 0 to 120 GPa, the temperature
from 200 to 1000 K, and the molar volume from 0 to 40 3.

In the case of an unidentied phase change, the triple point


appears as a maximum or minimum in the melting temperature,
with the derivative passing discontinuously from positive to
negative or vice versa. This behavior appears as a turnover point
in the pVT surface, as may readily be seen1b,c for water at the
liquid-ice I-ice III transition point (251.165 K, 0.2099 GPa),
where the molar volume of liquid is 16.52 cm3 mol-1, that of ice I
is greater at 19.4 cm3 mol-1 (hence a negative slope before the
transition), and that of ice III is less at 15.7 cm3 mol-1 (yielding a
positive slope after the transition).

PHASE DIAGRAMS OF ALKALI METALS


The alkali metals are of technological interest as liquid cool-
ants for nuclear reactors and of chemical interest as they are
generally anticipated to be simple representatives of metals.
However, it turns out that they exhibit quite complex phase
Figure 6. pV relationship of the solid phases of sodium at T = 298 K
from ambient to high pressure; the volume dierence between fcc and behavior as pressure increases,16 undergoing a number of melting
bcc phases is not discernible on the scale of this diagram. There is a maxima and minima together with changes of crystal structure,
more-or-less continuous change in volume even as the crystal structure from the simple cubic ambient structure to increasingly complex
undergoes a number of increasingly complex phase transitions with structures.
increasing pressure.20. As an example of the case of a smooth, genuine maximum in the
melting curve, let us consider sodium. According to recent rst-
elements.16 Although some of these observations may be asso- principles simulations,18 the maximum displayed by the melting
ciated with obscured triple points,17 where a new solid or liquid temperature of sodium (about 990 K at 31 GPa; Figure 4) would
phase becomes stable at high pressure without detection, others be caused by the larger compressibility of the liquid phase
may indeed involve the volumes of liquid and solid phases compared to the solid. With increasing pressure, the (positive)
passing through a point of equal density (i.e., equal volume).18 dierence Vliquid - Vsolid is reduced by this eect and eventually
589 dx.doi.org/10.1021/ed100638r |J. Chem. Educ. 2011, 88, 586591
Journal of Chemical Education ARTICLE

passes through zero to negative values at some pressure. This to rise, the melting point falls further to a minimum of about
causes a smooth maximum in the pT plane, without any accom- 300 K at 118 GPa, accompanied by a series of phase changes to
panying rst-order liquid-liquid phase transition. Figure 5 illus- increasingly complex crystal structures.19a,c,d,21 The initial max-
trates how the width of the solid-liquid phase gap in the pVT imum in the melting point is not accompanied by a phase
diagram passes through zero as the Clapeyron slope goes from transition, either in the solid or liquid. The phase transition
positive to negative through the point of volume (density) equality. points are collected in Table 1.
Figure 4 displays the experimental pT phase diagram of Similarly complex is the behavior shown by cesium, where two
sodium in all its complexity,19 whereas Figure 6 shows the simple distinct maxima are reported in the T versus p melting curve.22
pV behavior of the phases at T = 298 K as determined by a rst- Much of the evidence for anomalous and unusual phase diagram
principles density functional theory study.20 In this gure, the behaviors for one-component systems at high pressures can be
bcc-fcc phase change is not distinguishible owing to the very found among the elements16,23 and in polymer materials.24
small volume change accompanying the transition, so that a
single continuous curve ts the pV data for both solid phases AUTHOR INFORMATION
rather well.
The 3D pVT diagram of sodium, liquid and solid only, is Corresponding Author
reported in Figure 7. This shows that at low pressures the less *E-mail: l.glasser@curtin.edu.au.
dense liquid is more compressible than the solid, with the two
phases reaching equality of density at a maximum melting point
ACKNOWLEDGMENT
at about 31 GPa and 990 K, whereafter the melting point
decreases to a bcc-fcc phase transition. As pressure continues We acknowledge helpful discussions with Nigel Marks
(Curtin) on the properties of carbon and the possibilities of
Table 1. pVT Data on the High-Pressure Phase Relations of liquid-liquid phase transitions. The 3D phase and other dia-
Sodium grams depicted were prepared with the aid of the free scalable
vector graphics program Inkscape.29
Crystal Structurea,b or Transition Point p/GPac T/Kc V/3 (estimated)

Hexagonald ambient <36 37.8 ADDITIONAL NOTE


a
hP2 Data for the carbon phase diagram: liquid-graphite-diamond
liquid ambient 300 40.6e triple point: ptr = 16.4 GPa; Ttr = 4250 K, VG = 4.57 cm3 mol-1;
body-centered cubic ambient 300 40.0 VD = 3.50 cm3 mol-1; VL = 4.97 cm3 mol-1 12b and ptr = 10.8 GPa;
bcc, cI2 Ttr = 4300 K, VG = 5.29 cm3 mol-1; VD = 3.42 cm3 mol-1; VL =
melting maximum 31 990 14.4 8.76 cm3 mol-1.25
face-centered cubic (variable)
fcc, cF4 REFERENCES
bcc/fcc 65 700 11.3 (1) (a) Glasser, L. J. Chem. Educ. 2009, 86 (12), 14571458. (b)
fcc/cI16 105 307 Glasser, L. J. Chem. Educ. 2004, 81 (3), 414418. (c) Glasser, L. J. Chem.
Educ. 2004, 81 (5), 645645. (d) Glasser, L. J. Chem. Educ. 2002, 79 (7),
cI16 9.771
874876.
tI50 9.622 (2) Herraez, A., Biomodel Web site, http://biomodel.uah.es/Jmol/
mP512 9.606 plots/phase-diagrams/ (accessed Dec 2010).
oC120 9.577 (3) (a) Bundy, F. P. J. Chem. Phys. 1964, 41, 38093814. (b) Bundy,
cI16/oP8 118 300 F. P. Carbon 1996, 34, 141153.(c) McHale, A. E. Phase Equilibria
Diagrams: Phase Diagrams for Ceramists, Vol. X; American Ceramic
oP8 9.445
Society: Westville, OH, 1994.
aP90 9.397 (4) (a) Tammann, G.; Mehl, R. F. The States of Aggregation; van
oP8/tI19 125 322 9.3 Nostrand: New York, 1925; (b) Tammann, G. Kristallisieren und
tI19 9.210 Schmeltzen; Johann Ambrosius Barth: Leipzig, 1903.
hP4 (transparent)b 200 7.88 (5) Stillinger, F. H.; Debenedetti, P. G. Biophys. Chem. 2003,
105, 211220.
320 6.50
(6) (a) Greer, A. L. Nature 2000, 404. (b) Johari, G. P. Phys. Chem.
600 5.03 Chem. Phys. 2001, 3, 24832487.
a
The crystal structures are described by their Pearson symbols, where (7) Angelini, R.; Ruocco, G.; De Panlis, S. Phys. Rev. E 2008,
the initial italicized lower-case letter describes the crystal class (a = 78, 020502.
triclinic/anorthic; m = monoclinic; o = orthorhombic; t = tetragonal; h = (8) (a) Glasser, L. Inorg. Chem. 2010, 49, 34243427. (b) Glasser, L.
hexagonal and rhombohedral; c = cubic), the second italicized capital J. Phys. Chem. C 2010, 114, 1124811251.
denotes the lattice type (C = side face-centered; F = all face-centered; I = (9) Morrison, G. J. Chem. Educ. 1993, 70, 606.
body-centered; R = rhombohedral; P = primitive). The nal number (10) Glosli, J. N.; Ree, F. H. Phys. Rev. Lett. 1999, 82, 4659.
represents the number of atoms in the unit cell. The Pearson symbol (11) Katayama, Y.; Mizutani, T.; Utsumi, W.; Shimomura, O.;
does not uniquely identify the space group. b Structures are as reported Yamakata, M.; Funakoshi, K.-i. Nature 2000, 403, 170173.
by Gregoryanz, et al.19d except for the high-pressure transparent phase (12) (a) Colonna, F.; Los, J. H.; Fasolino, A.; Meijer, E. J. Phys. Rev. B
and beyond.26 c The pT data are experimental; the V values to one 2009, 80, 134103. (b) Ghiringhelli, L. M.; Los, J. H.; Meijer, E. J.;
decimal place are estimated from a rst-principles density functional Fasolino, A.; Frenkel, D. Phys. Rev. Lett. 2005, 94, 145701.
theory study,20 whereas the values reported to two and three decimal (13) Brygoo, S.; Henry, E.; Loebeyre, P.; Eggert, J.; Koenig, M.;
places are based upon X-ray diraction measurements in situ.19d d Data Loupias, B.; Benuzzi-Mounaix, A.; le Gloahec, M. R. Nat. Mater. 2007,
from ref 27. e Data from ref 28. 6, 274277.

590 dx.doi.org/10.1021/ed100638r |J. Chem. Educ. 2011, 88, 586591


Journal of Chemical Education ARTICLE

(14) Eggert, J. H.; Hicks, D. G.; Celliers, P. M.; Bradley, D. K.;


McWilliams, R. S.; Jeanloz, R.; Miller, J. E.; Boehly, T. R.; Collins, G. W.
Nat. Phys. 2010, 6, 4043.
(15) Bridgman, P. W. Proc. Am. Acad. Sci. 1915, 51, 55124.
(16) Degtyareva, O. High Pressure Res. 2010, 30, 343371.
(17) Lazicki, A.; Fei, Y.; Hemley, R. J. Solid State Commun. 2010,
150, 625627.
(18) Hernandez, E. R.; Iniguez, J. Phys. Rev. Lett. 2007, 98, 055501.
(19) (a) ESRF home page, Spotlight on Science, http://www.esrf.
eu/news/spotlight/spotlight69/ (accessed Dec 2010). (b) Gregoryanz,
E.; Lundegaard, L. F.; McMahon, M. I.; Guillaume, C.; Nelmes, R. J.;
Mezouar, M. Science 2008, 320 (5879), 10541057. (c) Gregoryanz, E.;
Degtyareva, O.; Somayazulu, M.; Hemley, R. J.; Mao, H. K. Phys. Rev.
Lett. 2005, 94 (18), 185502/1185502/4. (d) Gregoryanz, E. Science
2008, 321 (5890), 772772.
(20) Hanand, M.; Loa, I.; Syassen, K. Phys. Rev. B 2002, 65, 184109.
(21) Eshet, H.; Khaliullin, R. Z.; Kuhne, T. D.; Behler, J.; Parrinello,
M. Phys. Rev. B 2010, 81 (18), 184107/1184107/8.
(22) Falconi, S.; Lundegaard, L. F.; Hejny, C.; McMahon, M. I. Phys.
Rev. Lett. 2005, 94 (12), 125507/1125507/4.
(23) (a) Tonkov, E. Y.; Ponyatovsky, E. G. Phase Transformations of
Elements under High Pressure; CRC Press: Boca Raton, London, New
York, and Washington DC, 2004; (b) Young, D. A. Phase Diagrams of the
Elements; Univ. of California Press: Berkeley, Los Angeles and Oxford,
1991. (c) Jayaraman, A.; Cohen, L. H. Phase Diagrams: Materials Science
and Technology; Academic Press: New York and London, 1970; Vol. 1.
(24) Rastogi, S.; Hohne, G. W. H.; Keller, A. Macromolecules 1999,
32, 88978909.
(25) Sekine, T. Carbon 1993, 31, 227233.
(26) Ma, Y. M.; Eremets, M.; Oganov, A. R.; Xie, Y.; Trojan, I.;
Medvedev, S.; Lyakhov, A. O.; Valle, M.; Prakapenka, V. Nature 2009,
458 (7235), 182185.
(27) (a) Barrett, C. S. Acta Crystallogr. 1956, 9, 671677.(b) National
Institute for Materials Science, AtomWork Inorganic Material Database.
http://crystdb.nims.go.jp/index_en.html (accessed Dec 2010).
(28) International Nuclear Safety Center, Argonne National Lab.
http://www.insc.anl.gov/matprop/sodium/section13.pdf (accessed Dec
2010).
(29) Inkscape Home Page. http://inkscape.org (accessed Dec 2010).

591 dx.doi.org/10.1021/ed100638r |J. Chem. Educ. 2011, 88, 586591

You might also like