You are on page 1of 11

J Mater Sci (2014) 49:31353145

DOI 10.1007/s10853-014-8014-9

Phase equilibria in the system SmRhO and thermodynamic


and thermal studies on SmRhO3
K. T. Jacob Preeti Gupta Donglin Han
Tetsuya Uda

Received: 14 October 2013 / Accepted: 2 January 2014 / Published online: 17 January 2014
Springer Science+Business Media New York 2014

Abstract Using isothermal equilibration, phase relations


are established in the system SmRhO at 1273 K. SmRhO3
with GdFeO3-type perovskite structure is found to be the
only ternary phase. Solid-state electrochemical cells, containing calcia-stabilized zirconia as an electrolyte, are used
to measure the thermodynamic properties of SmRhO3
formed from their binary component oxides Rh2O3 (ortho)
and Sm2O3 (C-type and B-type) in two different temperature
ranges. Results suggest that C-type Sm2O3 with cubic
structure transforms to B-type Sm2O3 with monoclinic
structure at 1110 K. The standard Gibbs energy of transformation is Dtr Go 87=J mol1 3763  3:39 T=K.
Standard Gibbs energy of formation of SmRhO3 from binary
component oxides Rh2O3 and Sm2O3 with B-type rare earth
oxide structure can be expressed as Dfox Go 75=
J mol1 64230 6:97T=K. The decomposition temperature of SmRhO3 estimated from the extrapolation of
electrochemical data is 1665 (2) K in air and 1773 (3) K
in pure oxygen. Temperature-composition diagrams at
constant oxygen pressures are constructed for the system
SmRhO. Employing the thermodynamic data for
K. T. Jacob (&)  P. Gupta
Department of Materials Engineering, Indian Institute of
Science, Bangalore 560012, India
e-mail: katob@materials.iisc.ernet.in; ktjacob@hotmail.com
P. Gupta
e-mail: drpreeti@platinum.materials.iisc.ernet.in;
preeti17_nov@yahoo.com
D. Han  T. Uda
Department of Materials Science and Engineering, Kyoto
University, Sakyo-ku, Kyoto 606-8501, Japan
e-mail: han.donglin.8n@kyoto-u.ac.jp
T. Uda
e-mail: uda.tetsuya.5e@kyoto-u.ac.jp

SmRhO3 from emf measurement and auxiliary data for other


phases from the literature, oxygen potential-composition
phase diagram and 3-D chemical potential diagram for the
system SmRhO at 1273 K are developed.

Introduction
Ternary oxides LnRhO3 containing lanthanide elements
display interesting magnetic [1, 2], electrochemical [3, 4]
and catalytic [5] properties. Some of the rare earth perovskite oxides (LaRhO3, DyRhO3, GdRhO3 and LuRhO3)
find application as cathode in the photoelectrolysis of water
[3, 4]. Crystal structures of LnRhO3 have also been
investigated by Macquart et al. [6]. Taniguchi et al. [1]
explored the electrical and magnetic properties of LnRhO3
(Ln = rare earth except Ce and Pm). They reported paramagnetic behaviour of LnRhO3 compounds above 5 K.
Magnetic property of these orthorhombic perovskites arises
from the presence of rare earth ions. Six 4d electrons of
Rh3? ion prefer low spin state with no magnetic moment.
Resistivity of all LnRhO3 compounds exhibits semiconductor-like temperature dependence. As the part of organized study of metalrhodiumoxygen systems [719],
thermodynamic properties and phase relations in the system SmRhO were investigated. This is the first study of
phase relations and thermodynamic properties for the system SmRhO. Thermodynamic data for SmRhO3 are
unavailable in the literature.
This article reports measurement of Gibbs energy of
formation of SmRhO3 using solid-state electrochemical
cells in two different temperature ranges. In the temperature range from 875 to 1110 K, C-type Sm2O3 with cubic
structure is in equilibrium with Rh and SmRhO3, whereas
B-type Sm2O3 is present in the temperature range

123

3136

J Mater Sci (2014) 49:31353145

11101300 K. Different types of two- and three-dimensional phase diagrams are developed using thermodynamic
data to provide comprehensive coverage.

Experimental
Materials
Sm2O3 powder used in this study is obtained from two
sources. Spectroscopic grade Sm2O3 from Johnson Matthey Rare Earths Products, U.K., consisted mainly of the
cubic phase with small amount of the monoclinic variety.
Sm2O3, *99.9 % pure (Indian Rare Earths Ltd., India),
was mainly monoclinic in the as received condition.
Sm2O3, from both sources, was heated at 1473 K for 1 h in
the vacuum furnace to remove H2O and CO2 and then
cooled to room temperature. The product had monoclinic
structure, C2/m(12), with lattice parameters a =
0.1415 nm, b = 0.3640 nm and c = 0.8828 nm, irrespective of the source. By heating the samarium oxide from
Indian Rare Earths Ltd. at 1073 K for 48 h in the dry
synthetic air, the C-form of Sm2O3, Ia3 (206) with lattice
parameter a = 1.093 nm, was obtained.
Powders of Rh (99.95 %), Sm (99.9 %) and Rh2O3
(99.9 %, metal basis) were obtained from Alfa Aesar.
Rh2O3 was heated at 1273 K under high-purity oxygen gas
for 24 h to obtain phase pure b-Rh2O3 or Rh2O3 (III),
which had orthorhombic structure with lattice parameters
a = 0.5148 nm, b = 0.5438 nm and c = 1.4693 nm. The
intermetallic SmRh was prepared by heating Sm and Rh in
the appropriate ratio at 1473 K for 48 h.
SmRhO3 was prepared by heating in pure oxygen gas a
pelletized mixture of Sm2O3 and Rh2O3 in the stoichiometric ratio first at 1273 K for 96 h and subsequently at
1573 K for 48 h with several intermediate grinding steps.
Oxides powders were ground together in an agate mortar,

passed through *325 mesh sieve and pressed into a pellet


at 150 MPa using a steel die. The final product was characterized by X-ray diffraction (XRD). Samarium orthorhodite was found to crystallize in GdFeO3-type distorted
perovskite structure (Pbnm) with lattice parameters
a = 0.5322 nm, b = 0.5759 nm and c = 0.7708 nm. The
obtained lattice parameters are in fair agreement with literature data [6].
Phase equilibrium studies at 1273 K
In order to explore the phase relations in the system Sm
RhO, mixtures of pure metals/intermetallics and binary
oxides were equilibrated at 1273 K, quenched in liquid
nitrogen and the phase identified at room temperature.
Starting materials used for the preparation of each sample,
the average composition and equilibrium phases identified
after quenching are listed in Table 1. The average compositions of the samples used are displayed in Fig. 1.
Components of each sample were thoroughly mixed and
pelletized in a steel die at 150 MPa before equilibration.
Eight samples containing metals/intermetallics and oxides
were pelletized and equilibrated in a closed system under
vacuum. The samples, contained in closed alumina crucibles and kept on a sacrificial pellet of the same composition, were placed in quartz ampules. These ampules were
evacuated to 0.1 MPa pressure and flame sealed. To initiate
the reaction between metals/intermetallics and oxide,
samples were first heated to 1373 K for *8 h. Subsequently, they were equilibrated for a total period of
*120 h at 1273 K. Two mixtures containing only oxide
phases were equilibrated under pure oxygen atmosphere at
standard pressure Po = 0.1 MPa. Samples were first heated
to 1623 K for *8 h to initiate the reaction between the
oxides and then equilibrated for a total period of *144 h at
1273 K. The samples were quenched at regular intervals,
ground to -325 mesh and repelletized six times during this

Table 1 Results for phase equilibrium studies at 1273 K


Sample No.

Starting material used

Average composition of samples

Equilibrium phase identified

XRh

XO

XSm

Sm2O3 ? Rh2O3 ? Rh

0.402

0.403

0.195

Sm2O3 ? SmRhO3 ? Rh

Sm2O3 ? Rh2O3 ? Rh

0.522

0.402

0.076

SmRhO3 ? Rh2O3 ? Rh

Sm ? Rh2O3

0.264

0.401

0.335

SmRh2 ? Rh ? Sm2O3

Sm2O3 ? Rh

0.234

0.351

0.415

SmRh ? SmRh2 ? Sm2O3

Sm ? SmRh ? Sm2O3

0.223

0.297

0.480

Sm5Rh4 ? Sm2O3

Sm ? SmRh ? Sm2O3

0.171

0.296

0.533

Sm7Rh3 ? Sm5Rh3 ? Sm2O3

Sm ? SmRh ? Sm2O3

0.131

0.301

0.568

Sm4Rh ? Sm7Rh3 ? Sm2O3

Sm ? SmRh ? Sm2O3

0.010

0.305

0.685

Sm (s.s.) ? Sm2O3

Sm2O3 ? Rh2O3

0.106

0.601

0.293

Sm2O3 ? SmRhO3

10

Sm2O3 ? Rh2O3

0.297

0.598

0.105

SmRhO3 ? Rh2O3

123

J Mater Sci (2014) 49:31353145

3137

Fig. 1 Isothermal section of the phase diagram for the system Sm


RhO at 1273 K. Plus marks represent the average compositions of
the samples equilibrated at high temperature

period. The phase compositions of the samples were


unchanged by further heat treatment. When there was no
change in phase composition with time, attainment of
equilibrium was assumed. Optical and scanning electron
microscopy (OM/SEM), energy dispersive spectroscopy
(EDS) and XRD were used for phase identification.
Electrochemical measurements
The reversible emfs of two solid-state electrochemical
cells,
Pt13 % Rh; Rh SmRhO3 C  Sm2 O3 ==CaOZrO2 ==
Rh2 O3 Rh; Pt13 %Rh I
Pt13 % Rh; Rh SmRhO3 B  Sm2 O3 ==CaOZrO2 ==
Rh2 O3 Rh; Pt13 %Rh

II

were measured as a function of temperature. In both cells,


the reference electrode on the right-hand side is positive.
The temperature range for the cell I was from 875 to
1125 K and for cell II from 1050 to 1300 K. Calcia-stabilized zirconia (CSZ) tube, which served as solid electrolyte, also provided physical isolation of reference and
measuring electrodes so that transport of oxygen between
the electrodes via the gas phase was prevented. In the
temperature and oxygen partial pressure ranges encountered in this study, CSZ had oxygen ion transport number
larger than 0.999. Wires of Pt13 % Rh alloy were used
as electrical leads. Electrodes were contained in evacuated and sealed enclosures since relatively high oxygen
partial pressures were encountered at high temperatures.

Decomposition of Rh2O3 and SmRhO3 determined an oxygen partial pressure over the reference and working electrodes. As the apparatus used in this study was similar to that
used earlier [7, 8], only a brief description is provided here.
Measuring electrodes were prepared by compacting a
mixture of Sm2O3 (C-type in cell I and B-type in cell II),
SmRhO3 and Rh in the molar ratio 1:1.5:1 against the
closed end of the CSZ tube, with Pt13 % Rh wire
embedded in the mixture. The particle size of the powders
used to prepare electrodes was in the range 525 lm. A
bell-shaped Pyrex tube was attached to the top open end of
the CSZ tube using De-Khotinsky cement. The tube was
evacuated through a side arm and sealed under vacuum. An
excess of SmRhO3 was taken because part of it would
decompose to establish oxygen partial pressure in evacuated and sealed enclosure around the measuring electrode
at high temperatures. The reference electrodes were made
by compacting a mixture of Rh and Rh2O3 in the molar
ratio 1:1.5 in a CSZ crucible with an implanted Pt13 %
Rh wire. The crucible was placed inside an outer silica tube
closed at one end. The CSZ tube enclosing the measuring
electrode and the attached bell jar was placed on the reference electrodes. The outer silica tube was also closed
with a Pyrex cover using De-Khotinsky cement. The
chamber containing the reference electrodes was also
evacuated and sealed. The difference in oxygen partial
pressure at the two electrodes at high-temperature was
measured as emf of the cell.
The cell assembled as described above was suspended in
a vertical resistance furnace with the electrodes located in
the even temperature (1 K) zone. The upper part of the
cell with the cement seals was maintained at room temperature during measurement. In order to minimize the
induced emf, a Faraday cage made from stainless steel foil
was placed between the furnace and the cell assembly. The
furnace temperature was controlled to (1) K and was
measured by a Pt/Pt13 %Rh thermocouple, checked
against the melting point of Au. A high-impedance
([1012 X) digital voltmeter with the sensitivity of
(0.01) mV was used to measure the cell potential.
The reversibility of the cell emf was verified by microcoulometric titration in both directions. A small direct
current (*50 lA) was passed through the cell using an
external potential source for *300 s. The open circuit emf
was monitored as a function of time after each titration.
The cell reversibility was confirmed when the emf returned
to the same value after successive titrations in opposite
directions. The emfs were also reproducible on temperature
cycling of the cell.
The electrodes were examined by XRD, SEM and EDS
before and after emf measurement. The phase composition of the electrodes was found to be essentially unchanged during high-temperature measurements. Only minor

123

3138

J Mater Sci (2014) 49:31353145

variation in the relative concentrations of the components,


consistent with decomposition of one phase to establish
oxygen pressure, was detected. The three-phase electrodes
consisting of Rh, SmRhO3 and Sm2O3 were stable. There
was no evidence of reaction between the electrodes and the
solid electrolyte.
Thermogravimetric studies
In order to confirm data from high-temperature electrochemical measurements, thermogravimetric analysis (TGA)
and differential thermal analysis (DTA) of SmRhO3 powder
were carried out in air in the temperature range from room
temperature to 1774 K at a heating rate of 10 K min-1 using
a standard instrument, TG8120 (Rigaku, Japan).

Results and analysis


Phase relations in the system SmRhO at 1273 K
Phase equilibrium experimental results of ten samples are
summarized in Table 1. In order to establish the isothermal
section of the phase diagram of the system SmRhO at
1273 K, the experimental data are combined with information on the three binary systems available in the literature
[20]. The ternary phase diagram for the system SmRhO at
1273 K is shown in Fig. 1. On the RhO binary, the only
stable oxide is Rh2O3. Along SmO binary edge, only Sm2O3
is found as stable oxide. Along the SmRh binary, five
stoichiometric (Sm4Rh, Sm7Rh3, Sm5Rh3, Sm5Rh4 and
SmRh) and one nonstoichiometric (SmRh2, 0.658 \ XRh
\ 0.678) intermetallic phases are observed [20]. A liquid
phase in the composition range 0.026 \ XRh \ 0.155 is also
identified at 1273 K. There is only one stable ternary oxide,
SmRhO3, in the system SmRhO at 1273 K. Sm2O3 ?
SmRhO3 ? Rh and SmRhO3 ? Rh2O3 ? Rh are two threephase regions involving pure metal Rh and SmRhO3 identified by experiment where all phases are solid. The tie-triangle with Sm2O3, SmRhO3 and Rh as corners is useful for
the measurement of Gibbs energy of formation of SmRhO3.
Stability of SmRhO3 depends upon the oxygen chemical
potential associated with this three-phase equilibrium;
greater the stability, the lower is the oxygen potential.
Sm2O3 is found to be in equilibrium with metal Rh, Smrich liquid alloys and all six intermetallic phases. This is
caused by the more negative value of Gibbs energy of
formation of Sm2O3 than that of Rh2O3. There are eight
three-phase regions involving alloys and intermetallics:
Sm ? SmRh (l) ? Sm2O3, SmRh (l) ? Sm4Rh ? Sm2O3,
Sm4Rh ? Sm7Rh3 ? Sm2O3, Sm7Rh3 ? Sm5Rh3 ? Sm2O3,
Sm5Rh3 ? Sm5Rh4 ? Sm2O3, Sm5Rh4 ? SmRh ? Sm2O3,
SmRh ?SmRh2 ? Sm2O3 and SmRh2 ? Rh ? Sm2O3.

123

Fig. 2 Temperature dependence of the reversible emf of the solidstate electrochemical cells

There are two three-phase fields where oxygen gas coexists


with two condensed phases: O2 ? Sm2O3 ? SmRhO3 and
O2 ? Rh2O3 ? SmRhO3. Two visible two-phase regions
involving alloys of variable composition, SmRh (l) and
SmRh2, in equilibrium with Sm2O3 are seen in Fig. 1. Other
narrow two-phase fields are defined by the boundary lines
separating adjacent three-phase fields.

High-temperature thermodynamic properties


of SmRhO3
The reversible emfs of cell I in the range 8751125 K and
cell II in the span 10501300 K are shown in Fig. 2 as
function of temperature. Within experimental error, the
emfs are linear functions of temperature. Least square
regression analysis gives:
EI 0:48=mV 208:9  0:01236 T=K

2
EII 0:26=mV 221:9  0:02407 T=K
The uncertainty limit represents twice to the standard
deviation (2r). The estimated systematic errors associated
with emf and temperature measurements are significantly
smaller than 2r. The reaction taking place at the right-hand
reference electrode of the two cells is
4=3Rh O2 ! 2=3 Rh2 O3

Reaction occurring at the left-hand measuring electrode


of the cell I composed of three condensed phases, Rh, Ctype Sm2O3 (cubic) and SmRhO3 is
4=3Rh 2=3 Sm2 O3 cubic O2 ! 4=3 SmRhO3

Similarly, at the measuring electrode of cell II, which


consists of Rh, B-type Sm2O3 (monoclinic) and SmRhO3,
the reaction is

J Mater Sci (2014) 49:31353145

3139

4=3 Rh 2=3 Sm2 O3 monoclinic O2 ! 4=3 SmRhO3


5
The virtual cell reaction obtained by combining the two
half-cell reactions (3) and (4) is
1=2 Rh2 O3 ortho 1=2 Sm2 O3 cubic ! SmRhO3

The standard Gibbs energy change for this reaction,


which represents the formation of SmRhO3 from the
component binary oxides with orthorhombic and cubic
crystal structures, is given by:
Dr6 Go 139=J mol1 gFE
60467 3:58 T=K

The standard entropy of Sm2O3 (cubic) (144.77


0.6 J K-1 mol-1) [22] and Rh2O3 (ortho) (75.69
0.5 J K-1 mol-1) [23] at 298.15 K is used in making the
estimate for SmRhO3. The estimated heat capacity of
SmRhO3 at 298.15 K is 102.48 ( 0.3) J K-1 mol-1.

where g = 3 is the number of electrons involved in the


electrode reactions, F is the Faraday constant and E is the
reversible emf of the electrochemical cell. Likewise, the
virtual reaction for cell II obtained by adding reactions (3)
and (5) is
1=2 Rh2 O3 ortho 1=2 Sm2 O3 monoclinic ! SmRhO3
8
The corresponding change in standard Gibbs energy is
Dr8 Go 75=J mol1 gFE 64230 6:97 T=K
9
The temperature-independent term in Eqs. (7 and 9)
gives the enthalpy of formation, and the temperaturedependent term with sign reversed is the entropy of formation of SmRhO3. These are average second-law values
in the temperature range of measurement.
Thermodynamic properties of SmRhO3 at 298.15 K
From Eq. (7), the second-law enthalpy of formation of
SmRhO3 from Rh2O3 with orthorhombic structure and
Sm2O3 with cubic structure is -60.47 (0.53) kJ mol-1 at
a mean temperature of 1000 K. In the absence of heat
capacity data for SmRhO3, thermodynamic properties at
298.15 K can be estimated by invoking the NeumannKopp
rule for the assessment of the heat capacity. The standard
o
enthalpy of formation (Df H298:15
K ) of SmRhO3 at 298.15 K
from elements in their normal standard states thus estimated
is -1176.9 ( 1.7) kJ mol-1. Auxiliary data on standard enthalpy of formation at 298.15 K of cubic Sm2O3
(-1827.4 1.6 kJ mol-1) [21, 22] and orthorhombic Rh2O3
(-405.53 0.26 kJ mol-1) [23] are used in the estimation.
Similarly, from the second-law entropy of formation of
SmRhO3 from Rh2O3 with orthorhombic structure and
Sm2O3 with cubic structure (-3.58 J K-1 mol-1) obtained
from Eq. (7), the standard entropy (So298:15 K ) of SmRhO3 can
be estimated as 106.65 ( 0.84) J K-1 mol-1 at 298.15 K.

Thermodynamics of phase transition of Sm2O3


It is clear from Fig. 2 that the cubic form of Sm2O3 is more
stable than the monoclinic form at lower temperatures. The
emf lines for the two cells intersect at 1110 K, identifying
the equilibrium transformation temperature of Sm2O3. For
the polymorphic transition,
Sm2 O3 cubic ! Sm2 O3 monoclinic

10

The standard Gibbs energy of transition obtained by


combining Eqs. (7 and 9) is
Dtr Go 87=J mol1 3763  3:39 T=K

11

-1

The enthalpy (3763 630 J mol ) and entropy


(3.39 0.6 J K-1 mol-1) of phase transition obtained in
this study are relatively small. From the difference in the
calorimetric enthalpy of formation of the two forms of
Sm2O3, Baker et al. [21] suggested a value of 3766
( 2510) J mol-1 for enthalpy of transition. The results
from two very different sources are in excellent agreement.
In a review of thermodynamics of rare earth sesquioxides,
Zinkevich [24] has suggested transition enthalpy of
3313 J mol-1, which is also in good agreement with the
result obtained in this study. However, the entropy of
transition suggested by Zinkevich [24] is somewhat higher
(4.899 J K-1 mol-1) because of an incorrect choice of
transition temperature (676 K).
Decomposition temperature of SmRhO3
Since monoclinic form of Sm2O3 is more stable at higher
temperatures, the decomposition of SmRhO3 at relatively
high oxygen pressures will produce a mixture of Rh ?
Sm2O3 (monoclinic). The variation of oxygen chemical
potential with temperature for three-phase equilibrium
between Rh, Sm2O3(monoclinic) and SmRhO3 can be
obtained from the emf of cell II and oxygen potential of the
reference electrode (Rh ? Rh2O3). Accurate measurement
of the oxygen potential of the reference electrode has been
reported earlier [9]:
DlrO2 80=J mol1 264243 188 T=K

12

The oxygen chemical potential of measuring electrode


of cell II computed from emf using the Nernst relation
DlrO2  Dlm
O2 4FEII is given by

123

3140

J Mater Sci (2014) 49:31353145

Fig. 3 Time dependence of


a TGA, b DTA and
c temperature signals on heating
SmRhO3 in air at 10 K min-1

1
Dlm
RT ln PO2
O2 128=J mol
349883 197:29 T=K

13

The decomposition temperature of SmRhO3 can be calculated using Eq. (13) at any defined pressure of oxygen. The
decomposition temperature of the ternary oxide in pure
oxygen at standard atmospheric pressure Po = 0.1 M Pa is
1773 (3) K and in air decomposition occurs at 1665
(2) K. Since the computed decomposition temperatures
are well beyond the temperature range of emf measurement,
it would be useful to confirm them by direct measurement.
Thermal analysis of SmRhO3 in air
The results of TGA and DTA analysis of SmRhO3 in air are
displayed in Fig. 3. Heat flow, mass change and temperature
are plotted as a function of time. There is some difference in
the onset temperatures for decomposition indicated in TGA
and DTA. The onset temperature from TGA in Fig. 3 and
differential DTG in Fig. 4 is 1650 (5) K, approximately
15 K below the value calculated from extrapolated thermodynamic data. Although TGA and DTG are generally
more sensitive indicators of decomposition, presence of
oxygen nonstoichiometry in SmRhO3 can bias the results.
More careful study of both cation and anion nonstoichiometry of SmRhO3 is required for resolving the small but
significant differences in data from different techniques.
Temperature-composition phase relations at constant
oxygen partial pressures
Temperature-composition phase diagram for a ternary
system such as SmRhO with a volatile component will

123

be three dimensional even at constant partial pressure of


oxygen; phase composition is represented on a Gibbs
triangle in the XY plain, and temperature is plotted
along the Z-axis. Although such a plot gives a good
picture of the evolution of phase relations with temperature, it is difficult to read either the temperature or
composition accurately from a 3-D diagram. Therefore, it
is often more useful to present temperaturecomposition
relations in 2-D. This can be accomplished by defining a
single modified composition parameter. The composition
can be represented by metallic or cationic fraction, gRh/
(gSm ? gRh), where gi denotes the number of moles of
component i. A disadvantage of this choice is that oxygen
nonstoichiometry cannot be shown since oxygen is not
included in the composition parameter. The phase diagrams for the system SmRhO, obtained by plotting
temperature as a function of the modified composition
parameter, at constant oxygen partial pressures
(PO2 =Po 1, PO2 =Po 0:212 and PO2 =Po 106 ) are
shown in Figs. 5, 6 and 7, where Po = 0.1 MPa is the
standard pressure. At these oxygen partial pressures, Sm
component is fully oxidized to Sm2O3 in the whole range
of temperature. Only three-phase equilibria associated
with relatively high oxygen potentials fall in the parametric domain of the diagram. It is seen that the diagram
is sensitive to oxygen partial pressure, although the
sequence of phase stability remains essentially the same
at each oxygen partial pressure. Decomposition temperatures of all the oxides are reduced when the oxygen
partial pressure is reduced. As a second-order effect, the
difference between the decomposition temperatures of
SmRhO3 and Rh2O3 also decreases with decreasing oxygen partial pressure.

J Mater Sci (2014) 49:31353145

3141

Fig. 4 Differential
thermogravimetric analysis
(DTG) of SmRhO3 as a function
of temperature in air

Fig. 5 Temperature-composition phase diagram for the system Sm


RhO in pure oxygen, oxygen partial pressure PO2 =Po 1,
Po = 0.1 MPa

Fig. 6 Temperature-composition phase diagram for the system Sm


RhO in air, oxygen partial pressure PO2 =Po 0:212, Po = 0.1 MPa

Thermodynamic data for alloys and intermetallics

Baker et al. [21]; for Rh2O3, data have already been


cited as Eq. (12) [9, 23]. Gibbs energy of formation for
SmRhO3 is obtained from this study. Using calorimetric
data of Guo and Kleppa [25] for enthalpy of formation
of Sm5Rh4 and SmRh2, Miedemas model [26] and

In general, thermodynamic data for all the phases are


required to compute a ternary phase diagram. Most
reliable thermodynamic data for Sm2O3 are given by

123

3142

J Mater Sci (2014) 49:31353145

Fig. 8 Gibbs free energy of mixing as a function of composition for


the system SmRh at 1273 K. Phase boundaries are indicated by
vertical dotted lines

Fig. 7 Temperature-composition phase diagram for the system Sm


RhO at low oxygen partial pressure PO2 =Po 106 , Po = 0.1 MPa

SmRh phase diagram [20], thermodynamic data for


intermetallic phases in the binary system SmRh are
evaluated. The Gibbs energy of mixing for the liquid
alloy along the SmRh binary system can be approximated by a subregular solution model [27] at 1273 K.
Dmix G=J mol1 RTXSm ln XSm XRh ln XRh
XSm XRh XSm XSm XRh XRh
461XSm 11450XRh

14

where XSm and XRh are mole fractions of Sm and Rh,


respectively, R is gas constant and T is temperature in
Kelvin. XSm and XRh are the two subregular solution
parameters. The values for XSm and XRh are -225 and
-200 kJ mol-1, respectively. The last two terms on the
right-hand side of Eq. (14) represent the Gibbs energy of
pure liquid metals relative to the corresponding solids at
1273 K. As suggested by Hardy [27], partial Gibbs
energies Dli (i = Sm, Rh) can be calculated as:
2
DlSm =J mol1 RT ln XSm XRh
2XSm  XRh
3
2XRh XRh  XSm 461

15

2
DlRh =J mol1 RT ln XRh XSm
2XRh  XSm
3
2XSm XSm  XRh 11450

16

123

Fig. 9 Composition dependence of the evaluated chemical potentials


(Dli) of Rh and Nd in the system SmRh at 1273 K

Assessed Gibbs energy of mixing is plotted as a function


of composition in Fig. 8. The entropy of formation of
intermetallics from metals in the solid state is assumed to
be zero. The estimated values are consistent with the phase
diagram and calorimetric data. Tangent intercept method is
used to derive the chemical potentials of Sm and Rh, which
are plotted as a function of composition in Fig. 9. The
chemical potentials are constant in two-phase regions.
They are presumed to vary linearly with composition in
intermetallic compounds. Chemical potentials of liquid
alloys vary nonlinearly with composition following Eqs.
(15) and (16).

J Mater Sci (2014) 49:31353145

Fig. 10 Oxygen potential-composition diagram for the system Sm


RhO at 1273 K

Oxygen potential-composition diagram at 1273 K


The chemical potential of oxygen is displayed as a function
of the modified composition parameter, gRh/(gSm ? gRh),
in Fig. 10 for the system SmRhO at 1273 K. At very low
oxygen potentials at the lower end of the diagram, only
metallic phases are present and the sequence of phases
matches that in the binary SmRh phase diagram at the
same temperature. As the oxygen potential is increased, Sm
gets progressively oxidized to Sm2O3 (monoclinic)
accompanied by the formation of adjacent alloy phases
depleted in Sm. When three condensed phases and a gas
phase (oxygen) coexist at equilibrium in a ternary system
such as SmRhO, the system is mono-variant. Thus, at a
fixed temperature, three condensed phases are in equilibrium at a unique partial pressure of oxygen. The threephase equilibria are therefore represented by horizontal
lines in the diagram. The diagram obeys the same topological rules of construction as the conventional temperature-composition phase diagrams. For example, two-phase
regions separate three-phase horizontals.
Entire oxygen potential-composition diagram can be
divided into two discrete regions. Large difference in oxygen
potential between these two regions is shrunk by a break in
the axis of Fig. 10. Three-phase equilibria in the low oxygen
potential region involve two neighbouring alloy/

3143

intermetallic phases and Sm2O3. Information on the variation of chemical potential of Sm as a function of composition
is required for the calculation of the oxygen potential corresponding to two- and three-phase equilibria. Proceeding
from the lower end of the diagram, with gradual increase in
oxygen potential pure Sm at unit activity gets oxidized first to
Sm2O3. Thus, the first horizontal line at the bottom of the
oxygen potential diagram represents equilibrium between
Sm, SmRh (l) and Sm2O3. The area between the first and
second horizontal lines signifies two-phase region involving
Sm2O3 and liquid alloy. In two-phase region, oxygen
potential increases with progressive oxidation and consequent decrease in the activity of Sm. The second horizontal
represents three-phase equilibrium between SmRh
(l) ? Sm4Rh ? Sm2O3. In a similar way, with progressive
oxidation of Sm in each intermetallic, a series of two- and
three-phase equilibria are generated. When all the Sm present in SmRh2 is oxidized, then metallic Rh is seen to coexist
with Sm2O3. At significantly higher chemical potential of
oxygen, Rh combines with Sm2O3 and oxygen gas to form
ternary oxide SmRhO3. The oxidation of Rh to Rh2O3 occurs
at an oxygen potential higher than that for the formation of
SmRhO3. Thus, the high oxygen potential region of the
diagram involves the phases Rh, Rh2O3, Sm2O3 and
SmRhO3.
Construction of a three-dimensional (3-D) oxygen
potential-composition diagram with Gibbs triangle representing composition at the base (XY plane) and oxygen
potential along the vertical (Z) axis is possible. However, it
will be cumbersome to read data accurately from such a
diagram. Hence, a two-dimensional (2-D) representation of
chemical potential is preferred. A disadvantage of such
diagrams is that the composition of alloys and oxides with
the same value for metallic fraction, gRh/(gSm ? gRh),
cannot be differentiated since they fall on the same vertical
line. However, as the stability fields of alloys and oxides
are well differentiated along the oxygen potential axis, this
is not generally a major hurdle. Another disadvantage is
that information on the oxygen nonstoichiometry cannot be
provided on the diagram because oxygen is not included in
the modified composition parameter. Nevertheless, oxygen
potential-composition diagram exhibits useful information
regarding the oxygen potential range for the stability of the
various phases. The diagram is complementary to the
conventional Gibbs triangle representation of the phase
relations in the ternary system (Fig. 1), where the composition of each phase can be clearly depicted.
Three-dimensional chemical potential diagram
of the system SmRhO at 1273 K
The isothermal three-dimensional chemical potential diagram for the system SmRhO at 1273 K is presented in

123

3144

J Mater Sci (2014) 49:31353145

potential diagram gives an unambiguous geometric picture


of the stability domains of different phases in the chemical
potential space.

Conclusions

Fig. 11 3-D chemical potential diagram for the system SmRhO at


1273 K

In this investigation, phase relations in the system SmRh


O at 1273 K are established by the isothermal equilibration
technique and thermodynamic properties of SmRhO3
determined using a solid-state electrochemical method. All
six SmRh alloy phases coexist with Sm2O3 with monoclinic structure. Experimental results confirm the two threephase fields (Sm2O3 ? SmRhO3 ? Rh and SmRhO3 ?
Rh2O3 ? Rh) in the ternary phase diagram. The standard
Gibbs energies of formation for SmRhO3 from component
binary oxides [C-type Sm2O3 (cubic) or B-type Sm2O3
(monoclinic) and Rh2O3 (orthorhombic)] can be represented as:
1=2 Rh2 O3 ortho 1=2 Sm2 O3 cubic ! SmRhO3

Fig. 11. Chemical potentials of the three components Sm,


Rh and O2 are plotted along the orthogonal axis. It is to be
noted that the axis range for DlRh is substantially less than
those for the other two components. A plane represents the
stability domain of each stoichiometric phase. This follows
from the fact that the sum of chemical potentials of the
three components weighted by the appropriate stoichiometric coefficients is equal to the Gibbs energy of formation
of the phase. Stoichiometry of compound characterizes
slope of the plane. A plane will have curvature only if it
exhibits significant nonstoichiometry or variable composition. For a binary compound, slope is zero or infinite in one
direction. A line, defined by the intersection of the two
planes or surfaces, represents a two-phase field. A point of
intersection of three planes or surfaces represents a threephase region.
It is apparent from the Fig. 10 that the dominant phase
over a wide range of chemical potentials is Sm2O3 with a
slopes dDlO2 =dDlSm 1:333 and dDlO2 =dDlRh 0.
The intermetallics are represented by vertical planes at low
oxygen potentials. The width of the plane is a measure of the
stability of the intermetallic. The plane representing
SmRhO3 is restricted to the right top corner of the diagram
where chemical potential of Rh and O2 is relatively high and
chemical potential of Sm is very low. The slopes of the plane
representing SmRhO3 are dDlO2 =dDlSm 0:667 and
dDlO2 =dDlRh 0:667. The stability domain of Rh2O3
lies above that of the ternary oxide SmRhO3. The plane
representing Rh2O3 has slopes dDlO2 =dDlRh 1:333
and dDlO2 =dDlsm 0. The alloy phases are represented
by vertical planes. The width of the plane is a measure of the
relative stability of the phase. Thus, the 3-D chemical

123

Dfox Go 139=J mol1 60467 3:58T=K


1=2 Rh2 O3 ortho 1=2 Sm2 O3 monoclinic ! SmRhO3
Dfox Go 75=J mol1 64230 6:97T=K
Electrochemical measurement studies suggest that cubic
form of Sm2O3 converts to monoclinic form at 1110 K.
The enthalpy and entropy of phase transition are
(3763 630 J mol-1) and (3.39 0.6 J K-1 mol-1),
respectively. Using the NeumannKopp rule for heat
capacity, the standard enthalpy of formation of SmRhO3
at 298.15 K from elements in their normal standard states
1
o
is estimated as Df H298:15
1176:9.
K 1:7=kJ mol
The standard entropy of SmRhO3 at 298.15 K is
So298:15 K 0:84=J K1 mol1 106:65. The variation of
oxygen potential with temperature for the decomposition of
ternary oxide SmRhO3 to monoclinic form of Sm2O3 and
metal Rh, obtained from emf measurements, is given by
1
Dlm
349883 197:29T=K
O2 128=J mol

Two-dimensional temperature-composition phase diagrams at different partial pressures of oxygen and oxygen
potential-composition diagrams at 1273 K for the system
SmRhO are computed from experimental data obtained
in this study and auxiliary data from the literature. An
isothermal three-dimensional chemical potential diagram
for the system SmRhO at 1273 K is also presented. The
four types of phase diagrams presented here provide a
comprehensive description of the system. Accurate calorimetric measurements of heat capacity as a function of
temperature and enthalpy of formation can further refine
the thermodynamic data obtained in this study.

J Mater Sci (2014) 49:31353145


Acknowledgements K. T. Jacob is grateful to the Indian National
Academy of Engineering for support as INAE Distinguished Professor. Preeti Gupta thanks the University Grants Commission, India, for
the award of Dr. D.S. Kothari Postdoctoral Fellowship.

References
1. Taniguchi T, Iizuka W, Nagata Y, Uchida T, Samata H (2003)
Magnetic properties of RRhO3 (R = rare earth). J Alloys Compd
350:2429
2. Yi W, Liang Q, Matsushita Y, Tanaka M, Xiao H, Belik AA
(2013) Crystal structure and properties of high-pressure-synthesized BiRhO3, LuRhO3, and NdRhO3. J Solid State Chem
200:271278
3. Jarrett HS, Kung HHC, Sleight AW (1979) Photolysis of water
using rhodate semiconductive electrodes. US Patent 4,144,147
4. Jarrett HS, Sleight AW, Kung HH, Gillson JL (1980) Photoelectrochemical and solid-state properties of LuRhO3. J Appl
Phys 51:39163925
5. Gysling HJ, Monnier JR, Apai G (1987) Synthesis, characterization, and catalytic activity of LaRhO3. J Catal 103:407418
6. Macquart RB, Smith MD, Loye H-CZ (2006) Crystal growth and
single-crystal structures of RERhO3 (RE = La, Pr, Nd, Sm, Eu,
Tb) orthorhodites from a K2CO3 flux. Cryst Growth Des 6:
13611365
7. Jacob KT, Shekhar C, Waseda Y (2009) Phase relations in the
system TaRhO and thermodynamic properties of TaRhO4.
Mater Chem Phys 116:289293
8. Jacob KT, Shekhar C, Waseda Y (2009) Phase relations in the
system (chromium ? rhodium ? oxygen) and thermodynamic
properties of CrRhO3. J Chem Thermodyn 41:5661
9. Jacob KT, Sriram MV (1994) Phase relations and Gibbs energies
in the system MnRhO. Metall Mater Trans A 25A:13471357
10. Jacob KT, Okabe TH, Uda T, Waseda Y (1990) System CuRh
O: phase diagram and thermodynamic properties of ternary oxides CuRhO2 and CuRh2O4. Bull Mater Sci 22:741749
11. Jacob KT, Waseda Y (1995) Phase relations in the system La
RhO and thermodynamic properties of LaRhO3. J Am Ceram
Soc 78:440444
12. Jacob KT, Priya S, Waseda Y (1998) Alloy-oxide equilibria in the
system PtRhO. Bull Mater Sci 21:99103
13. Jacob KT, Priya S, Waseda Y (1998) Thermodynamic mixing
properties and solid-state immiscibility in the systems PdRh and
PdRhO. J Phase Equilib 19:340350

3145
14. Jacob KT, Waseda Y (2000) State cells with buffer electrodes for
measurement of chemical potentials and Gibbs energies of formation: system CaRhO. J Solid State Chem 150:213220
15. Jacob KT, Uda T, Okabe TH, Waseda Y (2000) Phase equilibria
and thermodynamics of (silver ? rhodium ? oxygen). J Chem
Thermodyn 32:13991408
16. Jacob KT, Gupta P (2013) Electrochemical determination of
thermodynamic properties of DyRhO3 and phase relations in the
system DyRhO. J Solid State Electrochem 17:607615
17. Jacob KT, Dhiman AK, Gupta P (2013) System GdRhO:
thermodynamics and phase relations. J Alloys Compd 546:
185191
18. Jacob KT, Agarawal K, Gupta P (2012) Thermodynamics of
TmRhO3, phase equilibria, and chemical potentials in the system
TmRhO. J Chem Eng Data 57:36773684
19. Jacob KT, Sharma J, Gupta P (2012) System HoRhO: phase
equilibria, chemical potentials and Gibbs energy of formation of
HoRhO3. J Phases Equilib Diff 33:429436
20. Massalski TB, Subramaniam PR, Okamoto H, Kacprza KL (eds)
(1990) Binary alloy phase diagrams, vol 3, 2nd edn. ASM
International, Materials Park, OH
21. Baker FB, George C, Fitzgibbon C, Pavone D, Holley CE,
Hansen LD, Lewis EA (1972) Enthalpies of formation of Sm2O3
(monoclinic) and Sm2O3 (cubic). J Chem Thermodyn 4:621636
22. Pankratz LB (1982) Thermodynamic properties of elements and
oxides, US Department of Interior, Bureau of Mines, Bulletin
672, pp. 394394
23. Jacob KT, Uda T, Okabe TH, Waseda Y (2000) The standard
enthalpy and entropy of formation of Rh2O3a third-law optimization. High Temp Mater Process 19:1116
24. Zinkevich M (2007) Thermodynamics of rare earth sesquioxides.
Prog Mater Sci 52:597647
25. Guo Q, Kleppa OJ (1998) Standard enthalpies of formation for
some samarium alloys, Sm ? Me (Me = Ni, Rh, Pd, Pt), determined by high-temperature direct synthesis calorimetry. Mater
Metall Trans B 29B:815820
26. Niessen AK, de Boer FR, Boom R, de Chatel PF, Mattens WCM,
Miedema AR (1983) Model predictions for the enthalpy of formation of transition metal alloys. Calphad 7:5170
27. Hardy HK (1953) A sub-regular solution model and its
application to some binary alloy systems. Acta Metall 1:202209

123

You might also like