You are on page 1of 6

MEANINGFUL TEACHING OF PHASE DIAGRAMS'

IT IS the writer's opinion aft,er twenty years of teaching


that the basic principles and utility of phase diagrams
are rarely grasped at the college level and that the
study of heterogeneous equilibria from the standpoint
of the phase rule, treated often in college as one small
facet of physical chemistry tends, in the student's
e n d , to remain so for the rest of his iife without the
realization of its immediate impact on many of the
procedures with which he is concerned as a working
chemist. In undergraduate texts the subject is treated,
perhaps unavoidably, in a chapter by itself, and refer-
ences to it elsewhere are usually lacking. In one widely
used book, for instance, at least seven chapters con-
tain material interpretable in terms of the phase rule
or phase diagrams but in only one of these chapters
is the interpretation as such undertaken. This is
entirely understandable and probably necessary if
clarity is to be preserved, but it is not uncommon to
find that the relevance of phase diagrams t o common
laboratory operations has been missed. How many
college graduates, for example, would dispute a state-
ment that an impurity always lowers the melting point
of a substance? How often is the relation of simple
eutectic diagrams to the phenomenon of steam distil-
lation realized? How often is it seen that the distrihu-
tion law determines the slope of the tie lines under a
binodal curve in ternary liquid systems? It is the
purpose of this article (1) to suggest a few aids in the
teaching of phase diagrams, (2) to emphasize the con-
nection between many seemingly unrelated diagrams
and (3) to voint out some conclusions derivable from
NORMAN 0. SMITH
Fordham University, New York, N. Y.
have already beeu described. The continuous change
from a vapor to a liquid ordinarily dealt with when
considering continuity of state in connection with the
Andrews isothermals can now be demonstrated on the
pressure-temperature diagram by passing from the
field for vapor to that for liquid around the end of the
vaporization line thereby emphasizing the abrupt end
of the latter at a critical point. This gives an oppor-
tunity to remark that the critical phenomenon is ob-
servable only on passing through the critical point to
or from a point on the vaporization curve. The usual
experimental demonstration of this phenomenon in-
volves heating a sealed glass tube containing liquid
and vapor in equilibrium, but it is rarely stated that
only when the relative amounts of liquid and vapor
are initially such as to give an over-all density equal to
the critical density will both liquid and vapor persist up
to the critical temperature and so give the desired effect.
That this condition must obtain is immediately evident
from the fact that on the Andrews diagram only an
isochor for a particular volume of system, namely
that equal to the critical volume, will pass through the
critical point. Incidentally the horizontal portions of
the Andrews isothermals provide an early opportunity,
if so desired, to introduce the rule known variously as
the lever law, the proportioning rule or the law of
mixtures, according to which (in this instance) the
relative amounts of coexisting liquid and vapor for a
given over-all volume can be readily calculated. Fa-
miliarity with this rule is of great assistance later ~vhen
the idea of tie lines is encountered.
them'khich-have a direct bearing on everyday labora-
tory phenomena, Only those aspects of the subject
BINARY SYSTEMS
which seem to the writer to require comment will be The phase behavior of binary systems requires a
referred to. In no sense. therefore. is what follows three-dimensional model for its com~lete revresentation
to be regarded as comvrehensive. in terms of the conventional variables temperature,
pressure, and concentration, and it is only through a
ONE-COMPONENT SYSTEMS
study of such models that complete understanding of
There is rarely difficulty in grasping the meaning heterogeneous equilibria is gained. With some sacri-
of the familiar ~ressure-temnerature diagram for a fice of the over-all viewuoint but considerable gain in
-
one-component system and the application of the
Clapeyron relation to the vaporization, sublimation
and fusion curves. In this connection, however, it is
helpful to recall the pressurevolume isotherms (1)
("Andrews isothermals") commonly used in discussing
the liquefaction of gases, and to remind the student
that, for a given temperature, the vapor pressure of a
liquid is the height of the horizontal line of the cor-
responding Andrews isothermal, which will normally
Presented before the South Jersey Section of the American
Chemical Society, Penns Grove, N. J., January 15, 1957, and
before the Division of Chemical Education a t the 132nd Meeting
of the American Chemical Society, New York, September, 1957.
-
clarity and simplicity it is common, however, to draw
only isobaric or isothermal sections of these models.
In discussing equilibria involving vapor both kinds
of. section are commonly used, whereas in discussing
other types only the isobar is usually of consequence.
Nevertheless, it is helpful to realize that, qualitatively,
the isobar is often a mirror image of the isotherm.
Examples of this will be seen below.
Let us consider a hypothetical "binary system"
A-B in which there is immiscibility in all states of ag-
gregation (solids A and B are not isomorphous, liquids
A and B are immiscible and so are gapes A and B (!)).
Our isobaric "phase diagram" (one atmosphere total
VOLUME 35, NO. 3, MARCH, 1958 125
Figure 1. Isobar for lmmisci- figure 2. l s obu for Complete
bility in i l l States of A g p g a - Miscibility in All Statn of Agmege-
tion tion
pressure) is shown in Figure 1 if the normal melting
points (T) and normal boiling points (T') are in the
order T, < T, < T', < TI,. The rectangular char-
acter of the fields expresses the chief feature of this
system, namely immiscibility. An isotherm for this
system would similarly be so characterized.
Consider, on the other hand, a binary system in which
there is complete miscibility in all states of aggregation.
(The system bromobenzene-chlorobenzene may be
cited as an example.) If we ignore, temporarily, the
possibility of azeotropism, which may be regarded as
incidental to the main diecussion, the phase relations
are given in the isobar Figure 2, where, again, T, <
T, < T', < T',. The rectangular character of Figure
1 has disappeared. The juxtaposition of vaporization
and melting equilibria and their obvious similarity is,
regrettably, rarely shown in textbooks. The relation
of the upper pair of curves to fractional distillation
needs no emphasis. I t may be noted parenthetically
that if one of the components be regarded as an im-
purity in the other then it is clear that (volatile) im-
purities can either raise or lower a boiling point. The
lower pair of curves, similarly, form the basis for frac-
tional crystallization and show that impurities can
raise as well as lower a melting point. Furthermore,
all mixtures of A and B w~ll have a melting range.
It is well, here, to admit to ambiguity in the usage of
the term "melting point" where melting ranges exist.
If the terms melting point and freezing point are
taken to be synonymous, then, of course, the melting
point will not be the temperature at which insipient
melting point occurs, although there would seem to
be much in favor of such a usage.
Both lenticular areas are crossed by (two-point)
tie lines, such as ab, to which the lever law is applicable
for closed systems. (The horizontal axis, be it noted,
measures not only the composition of every phase
present but also the total composition of the system-
this is invariably a trouble spot in teaching the subject.)
With reduction of total pressure the diagram would
gradually alter by the approach of the two lenticular
regions until they would overlap eventually and a
three-phase equilibrium would be possible. (Actually
in this approach, both areas would drop, but the
upper one would drop much faster because of the far
greater sensitivity of liquid-vapor phenomena to pres-
sure changes.) One may predict with some confidence
that an isotherm for the system (over a restricted tem-
perature range) would have the appearance of an ap-
proximate mirror image of Figure 2 with the mirror
126
at the base AB. For ideal behavior in all three states
of aggregation(this is usually assumed anyway in the
gaseous state) equations can be found for all of these
curves. I t is worth noting that, with this restriction,
the lower curve of the upper pair on the isobar is still
curved although its counterpart in the isothermal mirror
image is straight (if concentrations are expressed in
mole fraction), in accordance with Raoult's law.
Deviations from ideal behavior can, but do not
necessarily produce maxima or minima in one or
both pairs giving rise to "azeotropism" in vaporization
and/or melting phenomena, although according to its
etymology, the term should be confined to vaporization.
It may be commented that the generally accepted
statement that in distillation the vapor is richer in the
more volatile component is not necessarily true when
azeotropism is present unless the azeotrope itself is
regarded as a component. Whether maxima or minima
result is determined by the sign of the difference be-
tween the deviations in the two stages of aggregation
involved in the phase change, and not by the sign of
the deviations themselves ( 2) . If, for example, there
are positive deviations in the solid state which are
much greater than the positive deviations in the liquid,
a minimum in the freezing point curve on the isobar
will result. The same will be true if there are negative
deviations in the solid state which are much less than
the negative deviations in the liquid. A reversal of
these conditions will give a maximum. It is not cor-
rect therefore to assume, as is commonly done, that
a minimum in the freezing point curve necessarily im-
plies positive deviations in the solid state and so a
tendency to partial miscibility. The same consider-
ations apply in principle to the liquid-vapor equilibria.
Here, however, one can assume the vapor to be ideal
so that only the deviation in the liquid is the determin-
ing quantity.
In the foregoing discussion of complete miscibility
i t has, of course, been implied that both components
are volatile. I t is hardly necessary to say that if only
one of these is appreciably volatile we have isotherm and
isobar tvoes as in Figures 3 and 4. the line in Figure 3
" A - -
being straight only for ideal systems.
MOLE FRACTI ON
Figure 3. Isotherm for Only
One Volatils component (Ideal Figu=. 4. Isobar Only One
Behavior)
Volatile compon.nt
Let us now arrive at a very common type of diagram
by combining the two preceding categories to give a
two-phase equilibrium in which there is complete mis-
cibility in one phase, say a gas phase, but complete
immiscibility in the other, say the liquid state. Both
curved and rectangular features now appear. By as-
suming ideality in all binary phases it is possible to
JOURNAL OF CHEMICAL EDUCATION
draw both an isotherm and an isobar from readily
available or readily calculated data. It is suggested
that having the student construct such a drawing can
be very instructive.
The isotherm, shown in Figure 5, can be drawn
merely from a knowledge of the vapor pressures of
pure A and pure B (PA, P i ) at the given tempera-
ture. Point E, which represents the composition and
total pressure of the vapor saturated with both (pure)
liquids, has a mole fraction of B equal to PB/(PA +
Pi ) , and a pressure equal to P> + Pi. Any
other point on either curve can be plotted readily as
+
A MOLE FRACTI ON B
MOLE FRACTI ON B
Pig"?* 5. Isotherm for Immis- Figure 6. 1.ob.r 0. Irnmisci-
cib1e Liquid. bl. Liquids
follows: The curve EPL, for example, represents
vapors saturated with liquid B but also containing A.
Any point R on it, of mole fraction of B equal to x, must
have an ordinate, P, equal to the partial pressure of
A, namely (1 - x,) P plus the partial pressure of B,
namely Pi. From P = (1 - xl) P + Pi it is seen
that P = Pilx,. It is, therefore, a simple matter
to plot the whole diagram for given values of PA and
Pi. (Note that the position of the right-hand curve
is quite independent of the nature of the left-hand com-
ponent and vice versa.) Thus we have the genesis of
a diagram type which is prominent in most phase
diagram studies. It is safe to say that the analogy
between the phase behavior just discussed and that of
the familiar eutectic diagram passes quite unnoticed
by showing Figure 3 in textbooks merely as the three
horizontal lines through E, PI and Pi, respectively,
in connection with a discussion of steam distillation.
The isobar (Figure 6) can also be easily plotted,
ignoring changes of vapor pressure with total pressure.
The necessary data are the vapor pressures of both
pure liquid components over a range of temperature.
The immiscible liquid system water-chlorohenzene at
a total pressure of one atmosphere may be chosen for
illustration. The handbook gives the following values
of vapor pressure: water (A) 100, 400, and 760 mm.
at 52, 83, and 100C. respectively; chlorobenzene
(B) 100, 400, and 760 mm. at 71, 110, and 13ZC.
respectively. An equilibrium vapor for which the
partial pressure of B, P,, is 400 mm. will have a mole
fraction of B, x,, of 400/760 or 0.53. If saturated
only with liquid B the temperature must be 110'.
Similarly a vapor for which PA is 400 mm. (x, = 0.47)
and saturated only with A must be at 83". If P, is
100 mm. xB = 0.13 and the temperature must be 71'
for such a vapor saturated with B; if PA is 100 mm.
VOLUME 35, NO. 3, MARCH, 1958
xB = 0.87 and the temperature must be 52' if saturated
with A. Figure 6 is a plot of the data. Point E is
at xB = 0.30 and 91, the vapor composition and
boiling point of the constant boiling mixture for one
atmosphere Again the curves are independent.
Clearly, the bases of the calculation are ideal behavior
in the phase in which there is complete miscibility,
and the Clausius-Clapeyron equation for the hetero-
geneous liquid-vapor equilibria for both of the im-
miscible liquids. The horizontal line through E is
conveniently described as a three-point tie line marking
the juxtaposition of three areas crossed by two-point
tie lines.
We are now in a position to reason that the isobaric
phase diagram for solid-liquid equilibria, in which
there is complete miscibility in the liquid state but
complete immiscibility in the solid state will be en-
tirely analogous to Figure 6, as shown in Figure 7.
Here, however, the Clausius-Clapeyron equations for
both the solid and liquid forms are involved, but by
combining them with Raoult's law in the well-known
way (3) it can be readily shown that the left-hand
curve follows the equation
where AH,, is the molar heat of fusion of A. An ap-
proximate form of this relation, arrived at by confining
x, to small values, is invariably dealt with in discussing
molecular weight determinations by freezing point
depression, hut the connection between the latter and
ideal solubility can easily be missed. It is well to
make clear, too, that the eutectic is marked by the
three-point tie line through E and not merely by the
point E.
Inspection of Figure 7 gives opportunity to discuss
the effect of imuurities on melting points and melting
ranges, and to give the basis
for the use in the laboratory
of "mixed melting point"
determinations, purification
by partial melting, zone-
melting (4, 5), freezing mix-
tures, etc. Incidentally,
the use of an ice-sodium
chloride mixture for a freez- T~
ing mixture does not pro-
vide the best example in
this connection, for the salt
forms a dihydrate at these -
temperatures (6). It may A
be pointed out that a shesp
rieum 7. Isobar for Imrnisci-
ble Salids hut Miscible Liquids
melting uoint is usually,
but ni t always, characteristic of a pure compound,
for intimate mixtures of the eutectic composition,
racemic mixtures, and solid solutions with a minimum
melting point all melt sharply. Erasing one of the
curves of Figure 7 and rotating the whole diagram
through 90 gives the same diagram usually encountered
as a solubility curve. Systems with a eutectic also
raise an interesting question which has not been an-
swered to everyone's satisfaction: Why should an
intimate mixture of A and B begin to melt at a tem-
perature lower than the melting point of both pure
components, even when their volatility is negligible?
I n other words, why should the melting behavior of
solid A depend on the presence or absence of solid B?
We now consider the effect on the phase diagram of
the appearance of partial miscibility. The familiar
parabola-like curve enclosing the miscibility gap and
seen in the phenol-water system now makes its appear-
ance. If the two components are partially miscible
in the liquid state they will probably be immiscible
in the solid state, whereas if they are partially miscible
in the solid state they will probably be completely
miscible in the liquid state. This is because solid
miscibility requires greater similarity of the solid com-
ponents than is required of the liquid components for
liquid miscibility. The superposition of a miscibility
gap on melting phenomena will give the isobar, Figure
8, when the partial miscibility is in the liquid. There
are here two isobaric invariant points at temperatures
T, and T,. Unfortunately, there seems to be no
generally accepted name for the second of these.
(Metallurgists have, on occasion, referred to such a
point as a monotectic, but this term really refers to a
special kind of eutectic-one in which the invariant
liquid composition nearly coincides with that of one
Figws 8. I sobu for Partid Figure 9. hobo. for Miscibility
Mi s ~i bi l i t ~ in the Liquid but in the Liquid but Partid Misci-
lmmissibility in the Solid bility in the Solid
of the solid phases.) The reality of Figure 8 can be
brought out by showing that it describes the phase
behavior of "melting under solvent," which refers to
the phenomena sometimes encountered when attempts
to recrystallize a solid by heating it with excess solvent
show that the solid, rather than dissolving, suddenly
appears to melt, giving two liquid phases. Figure 8
zlso relates to the laboratory technique known as
"wet melting" in which one measures the melting point
depression of a solid produced by deliberately covering
it with a considerable excess of water. A compound
with intramolecular hydrogen bonding (chelation)
may frequently he distinguished from an isomer with
intermolecular hydrogen bonding (association) by the
fact that in the presence of water the latter has its
melting point lowered to a far greater extent than the
former (7, 8).
When the partial miscibility is in the solid state and
complete miscibility in the liquid either Figure 9 (eu-
tectic type) or a similar one (peritectic type) results.
These will not be discussed here (9). The analogy of
Figure 9 to boiling point diagrams of partially miscible
liquids (partial miscibility in the liquid but complete
miscibility in.the gaseous state) is obvious.
The appearance of binary solid compounds with
128
congruent or incongrueut meking points will not be
detailed here except as the subject relates t o hydrates
in equilibrium with water vapor. It is the writer's
opinion that, for instance, the commouly drawn iso-
therm for a system consisting of a salt and water, with
its step-wise character does absolutely nothing t o
relate such a phase diagram to any other kinds. The
Figur. 10. 1~0th-m for water and Salt Whish
Form. Hydrates
steps usually leave off in mid-air with no accompanying
explanation, and the two-and three-point tie lines,
common elsewhere, suddenly seem to have vanished.
A representation as in Figure 10, however, for a system
which forms two hydrates, HI and Hz, although losing
some simplicity, gains in every other way. The im-
miscibility of the solids appears in the rectangular
quality of most of the areas. The upper righehand
portion derives from Figure 3, modified for non-ideal
systems. There are three horizontal three-point tie
l i es. Such a diagram forms an excellent basis for
de~cribing not only efflorescence and hydration in the
usual way, hut also deliquescence, as well as hetero-
geneous equilibria from the standpoint of the mass law.
Deliquescence can occur when the partial pressure of
water in the vapor above the hydrate is larger than that
corresponding to the highest horizontal. The compo-
sition of the resultine laver of saturated solution is
- "
given, of course, by that of point Y.
Figure 11. Isobar for the System Sodium Chlo-
ride-Wster at 240 at-. (Schematic)
JOURNAL OF CHEMICAL EDUCATION
I t may, perhaps, form an interesting conclusion to
a study of binary systems to present an isobar for the
system sodium chloride-water. Figure 11 is the sche-
matic isobar for 240 atmospheres (6, 10, 11); on one
diagram are included melting, boiling and critical
phenomena, miscibility, immiscibility, eutectic and
peritectic phase relations, and the boiling point of a
saturated solution-all for a system of two familiar
components. The dotted lines are only conjecture.
The table gives the coordinates of the salient points
on the graph. (For the condensed equilibria the
coordinates given are actually those for a pressure of
1 atm., hut it is unlikely that they would be greatly
different at 240 atm.)
Cornmsi- Comuosi-
ti& Tempera- ti& Ternpew-
(wt . O/o lure (wt . O/o ture
Point NaC1) ( T . ) Point NaC1) ( "C. )
TERNARY SYSTEMS
The systematization adopted above for binary
systems is not conveniently extended to ternary systems
because of the much larger number of possible com-
binations of phenomena. In any case, there is only
time in a college course for a brief reference to ternary
phase diagrams. It is perhaps regrettable, for instance,
that liquid-vapor equilibria are never described, as
the student is likely, sooner or later, to encounter the
fractionation of a three-component liquid in the labo-
ratory (12). Only a few of the more deserving prac-
tical aspects, however, will be discussed here.
The isothermal behavior of two salts with a common
ion and water is shown in Figure 12 for a temperature
"zO SOLVENT
A 8 A X IMPURITY
rigur. la. 1 ~ 0 t h ~ ~ - for TWO F ~ W ~ ~ I ~ . ~ ~ ~ ~ t h ~ ~ ~ n i ~ . t n t .
Salts with a Common Ion and ing R.cvtalliration Between
Wa t e ~ Two Tsmpo~atures
above the freezing point of water. The course of the
curves JK and MK is usually determined by the phe-
nomena of common ion effect and salt effect. It may
he noted, in passing, that theory requires the extensions
J K and MK both to enter either the triangle AKB or
the fan-shaped areas. Juet as the lever law applies to
the tie lines in the fan-shaped areas, so also there i? a
simple rule by which the relative amount? of solid
A, solid B and liquid K can be calculated for total
compositions lying within the triangle AKB. For
those interested in analytical chemistry the utility of
this diagram in connection with the recent technique
of phase solubility analysis (IS, 14) for determining
purity both qualitatively and quantitatively is worth
mentioning. Moreover, this diagram relates to the
problem of determining whether a given unknown solid,
containing only A and B but of variable composition,
is a mixture or a solid solution. If a mixture, the com-
position of a liquid saturated with it would be inde-
pendent of the solid composition, but not so if a solid
Figure 14. Iaotherm for Salt (S)-Organic
Liquid (0)-wetar (W)
solution. (It is assumed that, if a mixture, not enough
solvent has been added to dissolve either solid phase
completely.)
Figure 13 shows the behavior represented in Figure
12, but at two temperatures, T and T' (T' > T, nor-
mally), and might well be used to describe recrystalliza-
tion between two temperatures. A composition of
impure solid, x, after addition of solvent to bring the
total composition to y, becomes homogeneous below
T' (after filtering off insolubles), but, on cooling to
T, throws down solid A, which is then filtered off.
If mother liquor is completely removed one such
treatment is sufficient for purification, at least in
theory. Clearly, this would be true only when A
and the impurity form no solid solution.
The concluding illustrations concern ternary liquid
syetems zhowing partial miscibility. The classical
example chloroform-water-acetic acid, or a similar
one, can he used as an opportunity to refer to the dis-
tribution law and process of extraction insofar as the
direction of the tie lines under the biuodal curve ie
concerned. Procedures such as "salting out" of an
organic liquid and the precipitation of a soluble salt
by addition of an organic liquid can be illnstrated by
means of Figure 14, which may be imagined t o have
arisen by the superposition of partial miscibihty on the
salt-water behavior shown in Figure 13, but with a
water-soluble organic liquid replacing the "impurity".
I n Figure 14, for the system salt (S) -organic liquid
(0) -water (W), the isothermal solubility curves of the
salt, AB and CD, are separated by a region of partial
miscibility bounded by the binodal BKC. Liquid B
and C are in equilibrium with each other and also with
solid salt. The "salting out" may be thought of as
follows: An appreciable quantity of the 0 present in
an aqueous solution F is obtained by adding salt until
the total composition, always on the line FS, lies wit,hm
the triangle BCS and then separating the 0-rich layer
of composition C. On the other hand, the precipitation
of a ~ohbl e salt from it.s aqueous solution can he visual-
VOLUME 35, NO. 3, MARCH, 1958
ized by imagining liquid 0 to be added to a saturated
(4) PFANN, W. G., J . Metah, 4 , Trans., 747 (1932).
aqueous solution of the salt, namely., A. The total
( 5 ) CHRISTIAN, J. D., J. CnEM. EDUC., 333 32
(6) "International Critical Tables," Val. IV, 1928, p. 235.
composition follows A0 and the precipitation of the
(7) SIDGWrCK, N, V,, W. J. SPURRELL, T. E. .,Y,E S, J .
desired salt is seen to occur immediately. Addition
Chem. Soc., 1915, 1202.
of sufficient 0 gives a total composition in the triangle (8) BAKER, W., J. Chem. Soe., 1934,1684.
BCS and a second liquid layer appears in addition to
(9) For a discussion of these see Pnur r o~, C. F., AND S. H.
the salt-familiar observations t o the practicing
MARON, "Fundamental Principles of Physical Chemis-
try," Revised ed., The Milcmillan Co., New York, 1951,
chemist.
pp. 410-12. Or see Ref. (IS), pp. 164-9.
LITERATURE CITED
(10) KEEVIL, N. B., J. Am. Chem. Sac., 64, 841 (1942).
(11) OLANDER, A,, AND H. LIANDER, Ada Chem. Scandinauica, 4,
(1) See, for example, DANIELS, F., AND R. A. ALBERTY, "Phyai-
1437 (1950).
cal Chemistry," John Wiley & Sans, he. , New York,
(12) FINDMY, A., A. N. CAMPBELL, AND N. 0. SMITH, '#The
1955, p. 16.
Phase Rule and Its Applications," 9th ed., Dover Puh-
(2) See, for example, HILDEBRAND, J. H., AND R. L. SCOTT,
lioations, Inc., 1951, Chap. 15.
"The Solubility of Nanelectrolytes," 3rd ed., A.C.S. (13) WEBB, T. J., Anal. Chem., 20, 100 (1948).
Monograph Series, Reinhold Publishing Corp., New (14) MADER, W. J., "Phase Solubility Analysis" i n WEISSBER-
York, 1950, p. 304. GER, "Organic Analysis," Vol. 4, Interscience Publishers.
(3) Ref. (I), pp. 215-6. Inc., 1954, p. 253.
JOURNAL OF CHEMICAL EDUCATION

You might also like