You are on page 1of 10

Powder Metallurgy and Metal Ceramics, Vol. 50, Nos. 7-8, November, 2011 (Russian Original Vol.

50, Nos. 7-8, July-August, 2011)

THERMODYNAMIC AND THERMOCHEMICAL


CHARACTERISTICS OF HOLMIUM SILICIDES

L. V. Goncharuk,1 V. R. Sidorko,1,2 N. P. Gorbachuk,1


and M. V. Bulanova1
UDC 541.11

The Gibbs free energy, enthalpy, and entropy of formation of holmium silicides HoSi2–a (HoSi1.82),
HoSi2–b (HoSi1.67), Ho3Si4, HoSi, HoSi1–x (HoSi0.98), Ho5Si4, and Ho5Si3 are determined by the
electromotive force method. The heat capacity and enthalpy of HoSi1.67 in the temperature range
55–2007 K are measured with adiabatic and mixing calorimetry. The enthalpy and entropy of α → β
polymorphic transformation and melting of HoSi1.67 are established. Temperature dependences of
the heat capacity, enthalpy, and entropy of HoSi and Ho5Si3 are evaluated based on the behavior of
heat capacity of mono- and lower silicides of rare-earth metals and their melting enthalpies in the
series from La to Lu.
Keywords: Gibbs energy of formation, enthalpy of formation, entropy of formation, enthalpy, heat
capacity, entropy, reduced Gibbs energy, holmium silicides.

INTRODUCTION
Silicides of rare-earth metals (REMs) represent an important class of compounds that possess unique
properties and are widely used in various areas, such as electronics and radio engineering, in the manufacture of
magnetic materials, and in the development of new structural corrosion- and heat-resistant materials. The
thermodynamic properties and phase diagram of compounds need to be understood to select the composition of an
alloy with desired properties on a reasoned basis and to perceive the nature of physicochemical interactions between
components over wide temperature ranges. These data may also be efficiently used in calculating the properties of
multicomponent systems that cannot be experimentally studied for some reasons. Systematic studies of the phase
equilibria and thermodynamic properties of REM–Si alloys that were initiated and guided by V. N. Eremenko,
Academician of the Ukrainian Academy of Sciences, have been underway at the Frantsevich Institute for Problems
of Materials Science for almost three decades.
The objective is to experimentally determine the thermodynamic characteristics (Gibbs energy, enthalpies,
and entropies of formation) of holmium silicides using the electromotive force (emf) measurement of galvanic cells
in the temperature range 610–970 K, to establish the heat capacity and enthalpy of HoSi1.67 with calorimetric
methods in the range 55–2007 K, and to evaluate the temperature dependences of heat capacity and melting
enthalpy of HoSi and Ho5Si3 compounds, for which this information is missing.

1Frantsevich Institute for Problems of Materials Science, National Academy of Sciences of Ukraine, Kiev,
Ukraine.
2To whom correspondence should be addressed; e-mail: dep6@ipms.kiev.ua.
Translated from Poroshkovaya Metallurgiya, Vol. 50, No. 7–8 (480), pp. 177–189, 2011. Original article
submitted August 27, 2010.

528 1068-1302/11/0708-0528 ©2011 Springer Science+Business Media, Inc.


t, °C
2200 1:1−x
5:3 5:4 1:1 3:4
1:2−b
1920 1:2−а
1860
1800 1850
1795 47 1820
1620

1474
1400 1305
1255 1290
1210
12.5
82.5

1000

795

600
520

Ho 20 40 60 80 Si
Si, at.%

Fig. 1. The Ho–Si phase diagram

Seven silicides form in the Ho–Si system. Silicide Ho5Si3 has a hexagonal structure of Mn5Si3 type [1] and
Ho5Si4 an orthorhombic structure of Sm5Ge4 type [2]. The formation and structure of Ho5Si3 and Ho5Si4 have been
ascertained may times. Compound HoSi with an orthorhombic structure of FeB type was revealed in [1]. This
structure was later reported in [3–5]. Holmium monosilicide with a CrB-type structure was found in [6] and then
ascertained in [4, 5, 7, 8]. It is believed [3, 4] that the two structures belong to stoichiometric composition and, thus,
can be regarded as polymorphic modifications of the compound: FeB-type structure as a high-temperature
modification and CrB-type structure as a low-temperature modification. The FeB-type structure was found in [5] in
alloys on the silicon side relative to HoSi and the CrB structure in alloys on the holmium side. Therefore, it is
considered in [5] that this polymorphism is due to composition rather than temperature. In constructing the phase
diagram of the system [8], it was established that these crystalline structures are not polymorphic modifications of
monosilicide but two compounds of close composition, one, HoSi1–x with the CrB-type structure, being silicon-
deficient. However, it is thought [9] that this interpretation is not likely and that the structures are two polymorphic
modifications of one compound. A compound was revealed in [8] at silicon content close to 55 at.%, whose
composition was deemed to be Ho4Si5. Later on, the composition of the compound was refined and determined to
be Ho3Si4 and its own structural type was established in [10]. According to [5, 8, 11, 12], hexagonal disilicide
HoSi2–b with an AlB2-type structure has composition close to HoSi1.7. The composition of the compound is
indicated as HoSi2 in [12]. The compound has a homogeneity range at 1000°C near HoSi1.67–1.6 according to [3]
and near HoSi1.8–1.63 according to [13]. In constructing the phase diagram of the system [8, 14, 15], thermal effects
were observed at 795°C in the range at about 60 at.% Si, which were attributed to polymorphic transformation of
HoSi2–b. Higher silicide HoSi2–a with an orthorhombic lattice of α-GdSi2 type is found in [16]. A tetragonal
structure of α-ThSi2 type was revealed in [7] for higher disilicide HoSi2–a. According to [3, 8, 13, 16–18], the
composition of the compound varies between HoSi2 and HoSi1.63.
The lattice parameters of holmium silicides [5] show that they all, except Ho5Si4, have insignificant
homogeneity ranges.
Hence, the composition and ranges of disilicide phases have not been ultimately established. In our study,
we accept the conclusion [7] on the extension of the HoSi2–b homogeneity range at 1000°C to be HoSi1.67–1.6,
which tends to composition HoSi1.67 with decreasing temperature. For HoSi2–a, we also accept the silicon-rich
homogeneity range with the boundary at HoSi1.82 [8, 14, 15]. Figure 1 shows the complete phase diagram of the
system constructed in [8, 14, 15] using electron microprobe analysis, x-ray diffraction, and differential thermal

529
analysis and refines the composition of “Ho4Si5” as Ho3Si4. Compounds Ho5Si3 and HoSi melt congruently at 1920
and 1860°C, respectively. Silicides Ho5Si4, HoSi1–x, HoSi2–b, and HoSi2–a form peritectically at 1850, 1820, 1620,
and 1290°C. Compound Ho3Si4 forms in the solid state at 1305°C.
The thermodynamic properties of Ho–Si alloys are limited to data on the standard formation entropies of
HoSi2–a (a = 0), HoSi, and Ho5Si3 determined with direct synthesis calorimetry [19] and to the mixing enthalpies in
the range of dilute solutions (0 < xHo < 0.04) at 1827 K [20, 21].

DESCRIPTION OF ALLOYS
The test alloys were melted from holmium (99.84 wt.%) and single-crystalline silicon (99.999 wt.%) on a
water-cooled copper hearth in pure argon in an arc furnace with a nonconsumable tungsten electrode.
Argon atmosphere was purged with a molten titanium getter. To complete the reaction and homogenize the
sample, it was tilted and melted three times. Melting losses determined by comparing the weight of the charge and
ingot were no more than 0.3%. To bring alloys to the equilibrium state for electrochemical measurements, the
samples were subjected to homogenizing annealing (in tantalum containers in argon at 1100°C for 200 h). Certified
samples were also used, whose phase and bulk composition was determined in constructing the Ho–Si phase
diagram [8] (Table 1). In making Ho–Si alloys, we did not manage to obtain an alloy from the heterogeneous range
[HoSi + HoSi0.98] because it was two narrow. Hence, the thermodynamic functions for HoSi1–x (x = 0.02) were
calculated using the respective data for HoSi and Ho5Si4.

TABLE 1. Measured EMF of Galvanic Cells (1), (2), and (3)


Galvanic Phase composition of B ⋅ 10–3, Σ (T– T )2,
cell alloys
xSi n T ,K E , mV δ 20 , (mV)2
mV/K K2
(1) HoSi2–a + Si 0.9073 43 736 –20.69 77.09 258952 7.81
(a = 0.18) 0.8500
0.7800
(2) HoSi2–b+HoSi2–a 0.6207 17 847 –0.50 –21.70 313161 17.27
(a = 0.18; b = 0.33)
(2) HoSi2–b + Ho3Si4 0.5794 25
(b = 0.33)
(2) HoSi + Ho3Si4 0.5450 49
(2) Ho5Si4 + HoSi1–x 0.4865 46 825 –108.6 2.06 72099 9.31
(x = 0.02) 0.4580
(2) Ho5Si3 + Ho5Si4 0.4120 49 862 –129.7 –28.32 45234 6.54
0.3900
(3) HoBi + (Bi) xBi = 0.800 15 799 706 –39.25 45883 4.42

Notes: n is the number of experimental data points; B = Σ (Ei – E )(Ti – T ) / Σ (Ti – T )2, where Ei and Ti are results of
individual measurements; and δ20 = Σ (Ei – E)2 / (n – 2), where E is emf calculated by Eq. (4).

The sample of composition HoSi1.67 (62.55 at.% Si) for calorimetric measurements was also arc-melted
from pure components in argon and annealed for 10 h at 1400°C. According to x-ray diffraction, the sample had a
crystalline structure of AlB2 type and lattice parameters a = 381.7 pm and c = 405.3 pm.

EXPERIMENTAL METHODS AND RESULTS


Electrochemical Measurements. The thermodynamic properties of holmium silicides were examined
through measurement of electromotive forces (emf) of the galvanic cells
[HoBi + L] | Ho+3 in molten electrolyte | [HoSi1.82 + Si] (1)

530
Е, mV 650 700 750 800 850 Т, K

−10
−20
−30
Si, at.%: 90.73 85.00 78.00
−40
700 750 800 850 900

10
0
−10
Si, at.%: 62.07 57.94 54.50

700 750 800 850 900


−90
−100 Si, at.%: 48.65 45.80
−110
−120
−130 Si, at.%: 41.20 39.00
−140

Fig. 2. Temperature dependences of emf for galvanic cells for Ho–Si alloys from the ranges
[HoSi1.82 + Si] (cell (1)) and [HoSi1.67 + HoSi1.82], [Ho3Si4 + HoSi1.67], and [HoSi + Ho3Si4] (cell (2))

in the temperature range 610–875 K and


[HoSi1.82 + Si] | Ho+3 in molten electrolyte | [HoxSi1–x] (2)
in the range 710–970 K (xHo > 0.35).
To weaken the interaction of holmium with the salt melt at elevated temperatures, we used a Ho–Bi alloy
from the heterogeneity range [HoBi + (Bi)] as a reference electrode, whose thermodynamic properties we examined
previously [22] in the range 673–873 K using emf measurement of the galvanic cell
Ho | KCl–LiCl–HoCl3 | [HoBi + (Bi)]. (3)
The left electrode of galvanic cell (2) was an alloy of holmium with silicon containing 90.73 at.% Si and
the right electrodes of galvanic cells (1) and (2) were two-phase alloys of holmium with silicon. Depending on the
extension of the two-phase region (Fig. 1), we examined from one to three alloys in each of them. Their phase and
bulk compositions are summarized in Table 1.
Eutectic mixtures of potassium and lithium chlorides with Tmelt = 627 K and potassium, sodium, and zinc
chlorides with Tmelt = 493 K [23], conventional for emf measurements, were used as an electrolyte. To prepare the
electrolyte, we used reagents of chemically pure grade (heavy metal content ≤10–3 wt.%). The preliminary
dehydrated salts were slowly heated in air with excess NH4Cl and the melt was additionally treated with ammonium
chloride to reach full transparency and remelted in a vacuum of about 1 Pa immediately after cooling.
Holmium ions were introduced into the electrolyte by holding pure holmium in molten potassium and
sodium at 750°C for 5 h in high-purity helium (99.985 vol.% He) and then added as a salt alloy to the galvanic cell.
The charge of holmium ions was taken to be +3 [24, 25].
The measurement procedure did not differ from conventional one and is detailed in [26]. The measured emf
values of galvanic cells (1) and (2) are presented in Fig. 2. In processing the experimental results, we took into
account data of experiments satisfying the conditions of a reversible galvanic cell. If emf does not change with time
at constant temperature, does not depend on how the temperature is reached (heating or cooling of the galvanic
cell), and is restored after small currents are passed through the cell, the conditions of the cell are considered to be
close to equilibrium and measured emf values can be used for thermodynamic calculations. The temperature
dependences of emf fitted to linear equations
E = A + BT ≡ E + B (T – T ), (4)

531
where E and T are the average emf and temperature. The parameters of these equations were calculated with the
least-squares method; they are summarized in compact form in Table 1 in accordance with [27]. The same table
provides data for the Ho–Bi alloy used as a reference electrode in cell (1). Using the emf addition rule and
temperature dependences of emf for galvanic cells (1)–(3), we plotted dependences E = f(T) for Ho–Si alloys
relative to pure holmium. It was found that the temperature dependences of emf relative to pure holmium for alloys
from the range [HoSi1.82 + Si] (cell (1)) and subsequent three ranges [HoSi1.67 + HoSi1.82], [Ho3Si4 + HoSi1.67], and
[HoSi + Ho3Si4] (cell (2)) agreed within the error of determination, as evidenced by near-zero values of
dependence E = f (T) for cell (2). Hence, the two dependences (for cells (1) and (2)) relative to pure holmium were
coprocessed using the procedure described in [28]. The overall equation for temperature dependence of emf for
these ranges becomes
⎡ 1 (T − 811) 2 ⎤⎥ mV.
E = 667.0 + 28.8 · 10–3 T ± ⎢ 4.2 +
⎢ 134 931 905 ⎥ (5)
⎣ ⎦

The emf matching values indicate that the chemical potentials of holmium in silicon-rich silicides are close, the
variation in ΔG Ho in transition from one phase range to another is lower than the total error of a single
measurement, and the excess thermodynamic stability of holmium silicides with respect to the coexisting phases is
very small. This apparently explains why the composition and ranges of disilicide phases existing in the Ho–Si
system have not been ultimately established.
The thermodynamic functions of formation of holmium silicide from solid components were calculated
through a combination of emf values for the respective galvanic cells taking into account their potential-determining
reactions. In the temperature range of interest (610–970 K), the Gibbs energies of formation of Ho–Si compounds
(J/mole) are determined by the following equations:
ΔfG° (HoSi1.82) = –193,170 – 8.46T; (6)
ΔfG° (HoSi1.67) = –193,041 – 8.28T; (7)
ΔfG° (Ho3Si4) = –579,600 – 25.20T; (8)
ΔfG° (HoSi) = –193,200 – 8.40T; (9)
ΔfG° (HoSi0.98) = –192,456 – 8.32T; (10)
ΔfG° (Ho5Si4) = –933,300 – 43.21T; (11)
ΔfG° (Ho5Si3) = –902,400 – 32.01T. (12)

The enthalpies and entropies of formation of holmium silicides from solid components are summarized in
Table 2.
Calorimetric Measurements. Calorimetric measurements of the heat capacity and enthalpy of HoSi1.67 were
performed in the temperature range 55–2007 K. The heat capacity at low temperatures was measured adiabatically
with an error no more than 0.4% [30]. The enthalpy was measured with a mixing method with en error no more than
1.5% in the range 300–2007 K using a high-temperature differential calorimeter [31] and above 1200 K using a
high-temperature calorimetric unit [32].
The measured heat capacities were used to calculate the thermochemical characteristics of HoSi1.67 in
standard conditions:
H° (298.15 K) – H° (0 K) = 12,364 ± 62 J/mole,
C op (298.15 K) = 61.87 ± 0.25 J/(mole ⋅ K),
S° (298.15 K) = 86.3 ± 0.7 J/(mole ⋅ K),
Φ′ (298.15 K) = 44.9 ± 0.7 J/(mole ⋅ K).

532
TABLE 2. Thermodynamic Properties of Ho–Si compounds

–ΔfH°, ΔfS°,
Compound Method T, K Reference
kJ/mole atom J/(K ⋅ mole atom)

HoSi2–a:
a=0 58.1 ± 2.4 Calorimetry 298.15 [19]
a = 0.18 67.8 ± 1.9 3.9 ± 2.4 emf (600–875) Our data
68.5 ± 0.7 3.0 ± 0.9 emf 710–930
66 Evaluation [29]
HoSi2–b, 72.3 ± 0.8 3.1 ± 1.0 emf 710–930 Our
b = 0.33 73.7 0.7 emf 298.15 data
Ho3Si4 82.8 ± 0.9 3.6 ± 1.1 emf 710–930 Same
HoSi 80.9 ± 2.2 Calorimetry 298.15 [19]
96.6 ± 1.3 4.2 ± 1.3 emf 710–930
96.5 4.3 emf 298.15
77 Evaluation [29]
HoSi1–x, 97.2 ± 1.4 4.2 ± 1.3 emf 760–970 Our data
x = 0.02
Ho5Si4 103.7 ± 1.6 4.8 ± 1.9 emf 760–970 Same
Ho5Si3 112.8 ± 2.1 4.0 ± 2.1 emf 760–970
113.2 3.2 emf 298.15
74.6 ± 2.1 Calorimetry 298.15 [19]
64 Evaluation [29]

The temperature dependences of the enthalpy (J/mole), heat capacity, entropy, and reduced Gibbs energy of
HoSi1.67 in the solid state are described by the following equations:
H° (T) – H° (298.15 K) = AT 2 + BT + CT – 1 + D, (13)

C op (T) = 2AT + B – CT–2, (14)


S° (T) = 2AT + BlnT + 0.5 CT – 2 + E, (15)
Φ′ (T) = AT + BlnT – DT–1 – 0.5 CT – 2 + (E – B). (16)
The enthalpy of HoSi1.67 above the melting point fits to the linear equation
H° (T) – H° (298.15 K) = aT + b. (17)
The coefficients of fitting dependences of thermochemical functions of HoSi1.67 are summarized in
Table 3.
The temperatures of enthalpy and entropy of the polymorphic transformation and melting of HoSi1.67,
determined from calorimetric measurements of enthalpy, are provided below:

Ttrans, K……………………………………. 1107 ± 25


ΔHtrans, kJ/mole…………………………… 4.2 ± 1.6
ΔStrans, J/mole · K………………………….. 3.8 ± 1.5
Tmelt, K…………………………………….. 1875 ± 40
ΔHmelt, kJ/mole……………………………. 79.7 ± 5.0
ΔSmelt, J/mole · K………………………….. 42.5 ± 2.7

To calculate the standard enthalpies and entropies of formation of holmium silicides, the temperature
dependences of the enthalpy and entropy of the compounds and of holmium and silicon at medium and high

533
TABLE 3. Coefficients in Temperature Dependences of Thermochemical Functions of HoSi1.67

Silicide A ⋅ 103 B C –D –E a –b

α-HoSi1.67 2.279 78.80 1625600 29149 373.21 – –


β-HoSi1.67 3.929 78.52 578804 25701 370.64 136.54 40709

temperatures need to be understood. The standard enthalpy and entropy of formation of HoSi1.67 (Table 2) were
calculated using the experimental results presented in this paper. The required values of enthalpies and entropies of
the elements at corresponding temperatures were taken from [33]. We determined the temperature dependences of
H° (T)–H° (298.15 K) and S (T) for HoSi and Ho5Si3 from analysis of the temperature dependences of heat
capacities of isoformula yttrium-subgroup silicides.
Analysis of variation in the heat capacity of gadolinium [31], erbium [34, 35], and lutetium [36, 37]
silicides shows that it changes monotonically in the range 298.15–Tmelt and, therefore, can be represented as a sum
of the lattice, anharmonic, and electron components and by the Schottky contribution due to multiplet therms of
trivalent REM ions (Cf,m). Since there is no Schottky contribution for gadolinium and lutetium silicides, their
isobaric heat capacity is determined by the first three components. Considering that the physicochemical
characteristics of mono- and lower silicides of REMs in the series from Gd to Lu [38] determining the first three
components of the heat capacity are similar, Cp(T) of other silicides can be represented as a sum of the regular part
of heat capacity and the respective Schottky contribution (Cf,m). The regular part of the heat capacity of HoSi and
Ho5Si3 was calculated as a sum of the parts of isobaric heat capacities of mono- and lower silicides of gadolinium
and lutetium in proportion to the position of Ho in the Gd–Lu series. The values of Cf,m were taken from [39].
Comparison of the heat capacities calculated by Eq. (14) and with the proposed method shows that the maximum
deviation for erbium monosilicide [34] is 6.0% and for Er5Si3 [35] no more than 5%, which is a good result taking
into account the error of determining the heat capacity from enthalpy data (4–5%). The calculated heat capacities of
HoSi and Ho5Si3 (in the range 298.15–Tmelt) are represented as equations (J ⋅ mole–1 ⋅ K–1):
Cp(T) = 14.02 ⋅ 10–3 ⋅ T + 43.75 – 44977 ⋅ T–2 , (18)
Cp(T) = 50.86 ⋅ 10–3 ⋅ T + 188.48 + 206350 ⋅ T–2, (19)
respectively.
The coefficients of Eqs. (18) and (19) were calculated with the least-squares method. The average deviation
of the calculated heat capacities from those fitted using Eqs. (18) and (19) was 0.3 and 0.4%. We integrated (18)
and (19) to find equations for the temperature dependences of enthalpy (J · mole–1) and entropy (J · mole–1 · K–1):
H°(T) – H° (298.15) = 7.01 ⋅ 10–3 ⋅ T 2 + 43.75 ⋅ T + 44977 ⋅ T–1 – 13818, (20)
S (T) = 14.02 ⋅ 10–3 ⋅ T + 43.75 ⋅ lnT + 22489 ⋅ T–2 – 182.71 (21)
for HoSi;
H°(T) – H° (298.15) = 25.43 ⋅ 10–3 ⋅ T 2 + 188.48 ⋅ T – 206350 ⋅ T–1 – 57764, (22)
S (T) = 50.86 ⋅ 10–3 ⋅
T + 188.48 ⋅ lnT – 103175 ⋅ T–2
– 697.38 (23)
for Ho5Si3.
The experimental enthalpies of formation of HoSi and Ho5Si3, Eqs. (20)–(23), and enthalpies and entropies
of individual substances [33] were used to find the standard enthalpies of formation of these compounds (Table 2).
Taking into account regular variations in temperature and melting enthalpies in the series of isoformula mono- and
lower yttrium-subgroup silicides, we assessed the melting enthalpies of HoSi and Ho5Si3. The melting enthalpy of
holmium silicide was determined as a sum of the melting enthalpies of the corresponding gadolinium silicide [31]
and erbium silicide [34, 35] taken in proportions. The melting enthalpies are 71.6 kJ ⋅ mole–1 (HoSi)
and 312.0 kJ ⋅ mole–1 (Ho5Si3). Using the melting temperatures established in [8, 15], we determined the melting
entropies to be 33.5 and 142.0 J/(mole ⋅ K) for HoSi and Ho5Si3, respectively.

534
at.% Si
Ηο 20 40 60 80 Si

−20

ΔfHo, kJ/mole atom


−40
−60
−80 - 1
- 2
−100 - 3
−120

Fig. 3. Formation enthalpies of holmium silicides: 1) 750–970 K, our data; 2) 298.15 K [19];
3) assessment [29]

The enthalpies of formation of HoSi1.82, HoSi, and Ho5Si3 resulting from emf measurement in the range
610–930 K are more exothermic than standard values determined with direct synthesis calorimetry [19]. For
HoSi1.82, the discrepancy may be due not only to the temperature dependence of formation enthalpy but also to the
difference in compositions. In addition, the difference may be because an ideal galvanic cell is hardly possible in
studying such reactive elements as REMs. These difficulties become especially evident when REM content of the
alloy increases and primarily affect the slope of curves E = f(T), i.e., the entropy and enthalpy of formation of
compounds. For this reason, the formation enthalpies determined at low REM contents (xR < 0.5) agree well with
calorimetric data, and the differences become more noticeable at xR > 0.5 (Fig. 3). Nevertheless, the phase diagram
of the Ho–Si system (Fig. 1) leads to a conclusion that the standard enthalpy of formation of Ho5Si3 is
underestimated in [19]. The mixing enthalpies for liquid alloys of holmium and silicon at xHo ≤ 0.04 are determined
calorimetrically in [20, 21] at 1827 K. The maximum enthalpy of dissolution of holmium in molten silicon at
1827 K is –198 kJ/mole. If we assume that this value will remain for solid alloy as well, then, taking into account
that Ho does not dissolve in solid Si, we can calculate the enthalpy of formation of the most silicon-rich compound,
HoSi1.82, which is –70.2 kJ/mole atom. This agrees well with the value resulting from emf measurement (–68.5 ±
± 0.7 kJ/mole atom).
The above experimental data on the phase equilibria and thermodynamic properties of Ho–Si alloys permit
the following conclusion. The Ho–Si system is characterized by strong chemical interaction between components.
The most thermodynamically and thermally stable compounds are in the holmium-rich range: Ho5Si3 (Tmelt =
= 2193 K) and Ho5Si4 (Tmelt = 2123 K) [8].

CONCLUSIONS
The Gibbs energy, enthalpy, and entropy of formation of holmium silicides HoSi1.82, HoSi1.67, Ho3Si4,
HoSi, HoSi0.98, Ho5Si4, and Ho5Si3 are determined with emf measurement of galvanic cells in the temperature
range 610–970 K. The thermodynamic stability of holmium silicides increases with REM content of a compound
and reaches its maximum values at the most refractory silicide, Ho5Si3.
Adiabatic and mixing calorimetry is used to measure the heat capacity and enthalpy of HoSi1.67 in the range
55–2007 K. The enthalpies of HoSi1.67 polymorphic transformation and melting are determined from the
temperature dependence of enthalpy.
The temperature dependences of the heat capacity, enthalpy, and entropy of melting of HoSi and Ho5Si3 are
assessed from the variation in the heat capacity of mono- and lower silicides of REMs and their melting enthalpies
in the series from La to Lu.
The standard enthalpy and entropy of formation of HoSi1.67, HoSi, and Ho5Si3 are calculated.

535
REFERENCES
1. E. I. Gladyshevskii and P. I. Kripyakevich, “Monosilicides of rare-earth metals and their crystal structures,”
Zh. Struct. Khim., 5, 853–859 (1964).
2. F. Holtzberg, R. J. Gambino, and T. R. McGuire, “New ferromagnetic 5:4 compounds in the rare earth
silicon and germanium systems,” J. Phys. Chem. Sol., 28, No. 11, 2283–2289 (1967).
3. J. Roger, V. Babizhetskyy, T. Guizouarn, et al., “The ternary RE–Si–B systems (RE = Dy, Ho, Er and Y) at
1270 K: solid state phase equilibria and magnetic properties of the solid solution REB2–xSix (RE = Dy and
Ho),” J. Alloys Compd., 417, 72–84 (2006).
4. D. Hohnke and E. Parthe, “AB compounds with Sc, Y and rare earth metals. II. FeB and CrB type
structures of monosilicides and germanides,” Acta Cryst., 20, 572–582 (1966).
5. F. Weitzer, J. C. Schuster, J. Bauer, and B. Jounel, “Phase equilibria in ternary RE–Si–N systems (RE⎯Sc,
Ce, Ho),” J. Mater. Sci., 26, 2076–2080 (1991).
6. E. Parthe, D. Hohnke, W. Jeitschko, and O. Schob, “Structure data of new intermetallic compounds,”
Naturwissenshaften, 65, No. 7, 13 (1965).
7. E. Houssay, Thesis, I.N.P.G, Grenoble (1990), [as cited in [13]].
8. V. N. Eremenko, V. E. Listovnichii, S. P. Luzan, et al., “Phase diagram of the holmium-silicon system and
physical properties of holmium silicides up to 1050°C,” J. Alloys Compd., 219, 181–184 (1995).
9. H. Okamoto, “Ho–Si (holmium–silicon),” J. Phase Equilib., 17, No. 4, 370–371 (1996).
10. I. Ijjaali, G. Venturini, and B. Malaman, “X-ray single crystal refinement of the Ho3Si4 structure,” J. Alloys
Compd., 269, L6–L8 (1998).
11. E. I. Gladyshevskii, “Crystal structures of Si-rich silicides of rare-earth elements in the yttrium group,”
Dop. AN URSR, No. 7, 886–888 (1963).
12. I. Mayer, E. Yaniar, and T. Shidlovsky, “Dimorphism of rare earth disilicides,” Inorg. Chem., 6, No. 4,
842–844 (1967).
13. S. Auffret, J. Pierre, B. Lambert-Andron, et al., “Magnetic properties versus crystal structure in heavy rare-
earth silicides RSi2–x,” Physica B, 173, 265–276 (1991).
14. V. N. Eremenko, V. E. Listovnichii, S. P. Luzan, et al., “Phase equilibria and constitution holmium–silicon
alloys,” in: Phase Equilibria, Phase Stability, and Metastable States in Metal Systems [in Russian], Inst.
Probl. Materialoved. AN Ukrainy, Kiev (1993), pp. 32–48.
15. V. N. Eremenko, V. E. Listovnichii, S. P. Luzan, et al.., “Phase diagram and physical properties of
holmium–silicon alloys,” Prots. Lit., No. 4, 66–74 (1994).
16. I. P. Mayer, E. Banks, and B. Post, “Rare earth disilicides,” J. Phys. Chem., 66, No. 4, 337–340 (1962).
17. J. Pierre, B. Lambert-Andron, and J. L. Soubeyrouz, “Magnetic structures of rare earth silicides RSi2–x (R =
Nd, Ho, Dy),” J. Magn. Mater., 81, 39–46 (1989).
18. J. Pierre, E. Siaud, and D. Frachon, “Magnetic properties of rare earth disilicides,” J. Less-Common Met.,
139, No. 2, 321–329 (1988).
19. S. V. Meschel and O. J. Kleppa, “Standard enthalpies of formation of some carbides, silicides, germanides
and borides of holmium by high temperature direct synthesis calorimetry,” J. Alloys Compd., 247, 52–56
(1997).
20. G. I. Batalin, V. S. Sudavtsova, and N. V. Stroganova, “Formation enthalpy of liquids Si–La (Gd, Dy, Ho,
Er) binary alloys,” Ukr. Khim. Zh., 51, No. 7, 775–777 (1985).
21. I. Mateiko, V. Sudavtsova, N. Kotova, and N. Sharkina, “Thermodynamic properties of Si–REM and Si–
Al–REM melts,” Visn. Kyiv. Nats. Univ., Issue 46, 50–52 (2008).
22. R. V. Antonchenko, L. V. Goncharuk, and V. R. Sidorko, “Thermodynamic properties of holmium
bismuthide,” Powder Metall. Met. Ceram., 48, No. 9–10, 578–581 (2009).
23. B. G. Korshunov, V. V. Safonov, and D. V. Drobot, Melting Diagrams of Chloride Systems [in Russian],
Khimiya, Leningrad (1972), p. 384.

536
24. V. I. Goryacheva, Ya. I. Gerasimov, and V. P. Vasil’ev, “Thermodynamic study of holmium and erbium
monoantimonides with electromotive force method,” Zh. Fiz. Khim., 55, No. 4, 1080–1082 (1981).
25. L. F. Yamshchikov, V. A. Lebed’, S. P. Rasponin, and P. A. Arkhipov, “Thermodynamics of formation of
holmium monobismuthide and monoantimonide,” Izv. Vuzov. Tsvet. Metall., No. 6, 106–108 (1985).
26. V. N. Eremenko, G. M. Lukashenko, and V. R. Sidorko, “Thermodynamics of compounds of vanadium,
chromium, and manganese with p-elements,” in: Physical Chemistry of Inorganic Materials [in Russian],
in 3 Vols., Vol. 1, Naukova Dumka, Kiev (1988), pp. 9–70.
27. “Compact presentation of experimental data in publishing thermochemical and thermodynamic research:
Recommendations by the Scientific Council on Chemical Thermodynamics and Thermochemistry of the
USSR Academy of Sciences,” Zh. Fiz. Khim., 46, No. 11, 2975–2979 (1972).
28. A. N. Kornilov and L. B. Stepina, “Some aspects of statistical processing of thermodynamic data. IV. Joint
processing of several linear equations,” Zh. Fiz. Khim., 44, No. 8, 1932–1938 (1970).
29. A. K. Niessen, F. R. de Boer, R. Boom, et al., “Model predictions for the enthalpy of formation of transition
metal alloys,” Calphad, 7, No. 1, 51–70 (1983).
30. A. S. Bolgar, A. I. Kriklya, A. P. Suodis, and A. V. Blinder, “Low-temperature heat capacity of
praseodymium, neodymium, and samarium sesquicarbides,” Zh. Fiz. Khim., 72, No. 4, 439–443 (1998).
31. A. S. Bolgar, N. P. Gorbachuk, and A. V. Blinder, “Enthalpies of Gd5Si3, Gd5Si4, GdSi, and GdSi1.88
silicides in the range 298.15–2200 K. Melting enthalpies,” Teplofiz. Vys. Temp., 34, No. 4, 541–545 (1996).
32. E. A. Guseva, A. S. Bolgar, S. P. Gordienko, et al., “Determining the enthalpy of self-bonded silicon
carbide,” Teplofiz. Vys. Temp., 34, No. 5, 649–653 (1996).
33. R. Hultgren, P. D. Desai, D. T. Hawkins, et al., Selected Values of the Thermodynamic Properties of the
Elements, ASM International, Metals Park, Ohio (1973), p. 631.
34. N. P. Gorbachuk, S. N. Kirienko, V. R. Sidorko, and I. M. Obushenko, “Enthalpy of ErSi and ErSi1.67 in
the range 400–2300 K,” Teplofiz. Vys. Temp., 45, No. 2, 203–207 (2007).
35. N. P. Gorbachuk, S. N. Kirienko, V. R. Sidorko, and I. M. Obushenko, “Thermodynamic properties of
Er5Si3 over a wide temperature range,” Ukr. Khim. Zh., 72, No. 11, 20–24 (2006).
36. N. P. Gorbachuk, S. N. Kirienko, V. R. Sidorko, and I. M. Obushenko, “Thermodynamic properties of LuSi
over a wide temperature range,” in: Proc. 6th Int. Conf. Materials and Coatings in Extreme Conditions [in
Russian], Inst. Probl. Materialoved. NAN Ukrainy, Kiev (2010), p. 278.
37. N. P. Gorbachuk, S. N. Kirienko, V. R. Sidorko, and I. M. Obushenko, “Thermodynamic properties of
Lu5Si3 over a wide temperature range,” Powder Metall. Met. Ceram., 48, No. 7–8, 449–453 (2009).
38. G. V. Samsonov, L. A. Dvorina, and B. M. Rud’, Silicides [in Russian], Metallurgiya, Moscow (1979),
p. 272.
39. A. I. Kriklya, Thermodynamic Properties of Sulfides of Cerium-Subgroup REMs [in Russian], Author’s
Abstract of PhD Thesis, Kiev (1986), p. 22.

537

You might also like