Next Article in Journal
Handling of Human-Geosphere Intersections
Previous Article in Journal
Evaluation of VIIRS Land Surface Temperature Using CREST-SAFE Air, Snow Surface, and Soil Temperature Data
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Gibbs Free Energy of Formation for Selected Platinum Group Minerals (PGM)

by
Spiros Olivotos
and
Maria Economou-Eliopoulos
*
Department of Geology and Geoenvironment, National University of Athens, Panepistimiopolis, Athens 15784, Greece
*
Author to whom correspondence should be addressed.
Geosciences 2016, 6(1), 2; https://doi.org/10.3390/geosciences6010002
Submission received: 6 October 2015 / Revised: 11 December 2015 / Accepted: 14 December 2015 / Published: 6 January 2016

Abstract

:
Thermodynamic data for platinum group (Os, Ir, Ru, Rh, Pd and Pt) minerals are very limited. The present study is focused on the calculation of the Gibbs free energy of formation (ΔfG°) for selected PGM occurring in layered intrusions and ophiolite complexes worldwide, applying available experimental data on their constituent elements at their standard state (ΔG = G(species) − ΔG(elements)), using the computer program HSC Chemistry software 6.0. The evaluation of the accuracy of the calculation method was made by the calculation of (ΔGf) of rhodium sulfide phases. The calculated values were found to be in good agreement with those measured in the binary system (Rh + S) as a function of temperature by previous authors (Jacob and Gupta (2014). The calculated Gibbs free energy (ΔfG°) followed the order RuS2 < (Ir,Os)S2 < (Pt, Pd)S < (Pd, Pt)Te2, increasing from compatible to incompatible noble metals and from sulfides to tellurides.

Graphical Abstract

1. Introduction

The platinumgroup elements Os, Ir, Ru, Rh, Pd and Pt (PGE) are among the most valuable elements in nature with strategic importance, due to their growing use in advanced technologies and automobile catalyst converters. Platinum groupelements (PGE) can be classified into two subgroups: the Os-, Ir-, Ru-rich or IPGE (more refractory) and PPGE (low-melting and more soluble) (Ir group), interpreted to reflect compatible behavior (partition or distribution coefficient between solid and magma, D ≥ 1) during large-degree mantle melting and (Pt, Pd)-rich or PPGE assemblages, reflecting the incompatible behavior (D ≤ 1) of the PPGE, showing enrichment as a function of the differentiation degree [1]. The behavior of PGE during partial melting and crystal fractionation and which minerals are collectors of PGEs have been investigated and reviewed extensively [2,3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20].
Despite the great interest in the PGE, the thermo-chemical basis of their geochemical properties remains unclear. Thermodynamic data for PGE-minerals (PGM) are very limited, due probably to experimental difficulties and very limited data on the activity-composition models and relevant phase diagrams [21,22,23,24,25,26,27,28]. Assuming that the required thermodynamic data (enthalpy, entropy, Gibbs free energy) of the reactants and physical/chemical conditions (P, T) are available, then the mineral Gibbs free energy can be calculated [26,27,28]. This paper focuses on: (1) the calculation of the Gibbs free energy of formation for the selected PGM in PGE-bearing complexes worldwide, at standard state conditions (ΔfG°), using the current state of knowledge from experimental data on the PGE; and (2) the application of the calculated Gibbs free energy (ΔG) values to predict the stability of PGM.

2. Methodology

Minerals have a Gibbs free energy of formation (∆fG°) value, which describes the amount of energy that is released or consumed when a phase is created from its constituent elements in their standard state. It is well known that the ΔfG° of a mineral varies with changes in pressure (P), temperature (T) and mineral composition (X) and that the more stable position is one of lower energy. The HSC Chemistry software 6.0, includes databases and various modules providing different types of chemical/thermodynamic calculation.
The Gibbs free energy is defined [26] by the following equation:
G = H − T × S
The enthalpy and entropy values are available in different databases (HSC 6.0 Thermo-chemical Data Source on PGE [29,30,31,32,33,34,35,36,37]). The standard Gibbs free energy (ΔG°) values of PGM were calculated (Table 1, [21,22,23,24,35]) as the difference between free energy (G) of the products and reactants (Equation (2) using data from the NBS tables of chemical thermodynamic properties [37]).
ΔG = G(species) ΣG(elements)
Table 1. The calculated standard Gibbs free energy (ΔfG°) values of formation for platinumgroup minerals are compared to available literature data.
Table 1. The calculated standard Gibbs free energy (ΔfG°) values of formation for platinumgroup minerals are compared to available literature data.
MineralLiterature DataRef.
FormulaNameΔfΔfΔfCp
kJ/molkJ/molkJ/molJ/mol·KJ/mol·K
RuS2Laurite−204.1−188.1−199.255.266.46[22,23]
OsS2Erlichmanite−135.1−134.1−146.954.8--[21]
IrS2Unnamed iridium disulfide−123.9−131.8−143.172.8--[21]
IrTe2Shuangfengite−68.0--−71.13123.43--[35]
PtTe2Moncheite−52.9−52.33−58.19120.9575.09[24,35]
PdTe2Merenskyite−60.6−47.36−50.40126.6776.30[24,35]
Ir2S3Kashinite−196.4−220.5−241.497.1--[21]
PtSCooperite−76.1−76.22−81.7954.8748.17[24]
PdSVysotskite−66.7−72.25−70.8757.6348.66[24]
OsAs2Omeiite−75.8--−76.57101.32--[35]
PtAs2Sperrylite−193.4--−217.5731.928--[35]
Rh2S3Bowieite−252.4--−262.47125.52--[35]
RuO2Ru-oxide−252.7--−305.0158.158--[35]
It is notable that the ° (“not”) on ΔfG° indicates that the ΔG value is based on the reaction at standard conditions (1 M solution concentration, 1 atm gas pressure). Temperature is not part of the standard conditions, but commonly, a temperature of 298 K is used. If the concentration is different from 1 M or 1 atm gas pressure, the change of the Gibbs free energy is written as ΔG.
Using the Reaction Equations mode, chemical compounds and reactions were analyzed, in order to calculate the change of the Gibbs free energy (ΔfG°) at standard conditions. Calibrating a chemical reaction, ΔrG° is calculated by Equation (3), when the chemical reaction follows Equation (5):
ΔG = G(Products) − G(Reactants)
ΣG(Products) ΣG(Reactants) = (c × GC + d × GD + …) (a × GA + b × GB + …)
aA + bB + … = cC + dD + …
Although the question of the original existence of bonds between metallic PGEs and sulfur, oxygen or other ligands remains unclear, the calculated Gibbs free energy values were calculated assuming the existence of metallic PGEs and sulfur [21,22,38] via the Equation (2).
Although thermodynamic calculations, which are derived by various computational methods based on the available reference data, and experimental methods yield considerable uncertainties [22,24], the calculated thermodynamic values for PGM appear to be in a good agreement with those given in the literature. In addition, to evaluate the accuracy of the method of the calculation of the Gibbs free energy of formation, the (ΔfG) for the rhodium sulfide phases Rh3S4 and Rh2S3were calculated using the HSC program (Table 2 and Table 3). The Reaction Equations mode was used to estimate the free energy of Reactions (6) and (7) through Equation (3), combining data from the HSC main database.
3Rh(s) + 2S2(g) = Rh3S4
4Rh3S4(s) + S2(g) = 6Rh2S3
These values were compared to those obtained using a solid-state electrochemical technique [25]. The calculated values of ΔfG for these rhodium sulfides (Table 2 and Table 3; Figure 1) were found to be in good agreement with those of temperature in the range from 925 to 1275 K.
Figure 1. Comparison of the calculated Gibbs free energy (ΔG) against temperature, for two reactions. (A) 3Rh(s) + 2S2(g) = Rh3S4(s) in this study with values obtained by previous authors [25]. Data for Line (a) from Table 2; data for Line (b): ΔGf(J·mol−1) = −548026 + 304.5 × T (K) from [25]. (B) 4Rh3S4(s) + S2(g) = 6Rh2S3(s) in this study with values obtained by previous authors [25]. Data for Line (a) from Table 3; data for Line (b): ΔGf(J·mol−1)= −230957 + 160.03 × T (K) from [25].
Figure 1. Comparison of the calculated Gibbs free energy (ΔG) against temperature, for two reactions. (A) 3Rh(s) + 2S2(g) = Rh3S4(s) in this study with values obtained by previous authors [25]. Data for Line (a) from Table 2; data for Line (b): ΔGf(J·mol−1) = −548026 + 304.5 × T (K) from [25]. (B) 4Rh3S4(s) + S2(g) = 6Rh2S3(s) in this study with values obtained by previous authors [25]. Data for Line (a) from Table 3; data for Line (b): ΔGf(J·mol−1)= −230957 + 160.03 × T (K) from [25].
Geosciences 06 00002 g001aGeosciences 06 00002 g001b
Table 2. Comparison between ΔGf values of (a) this study and (b) [25] for the reaction: 3Rh(s) + 2S2(g) = Rh3S4(s).
Table 2. Comparison between ΔGf values of (a) this study and (b) [25] for the reaction: 3Rh(s) + 2S2(g) = Rh3S4(s).
T (K)ΔG (a) (kJ/mol)ΔG (b) (kJ/mol)|ΔG (b) − ΔG (a)|
900−289.89−273.9815.92
950−272.63−258.7513.88
1000−255.46−243.5311.93
1050−238.37−228.310.07
1100−221.37−213.088.29
1150−204.46−197.856.60
1200−187.62−182.635.00
1250−170.88−167.43.48
1300−154.22−152.182.04
Table 3. Comparison between ΔGf values of (a) this study and (b) [25] for the reaction: 4Rh3S4(s) + S2(g) = 6Rh2S3(s).
Table 3. Comparison between ΔGf values of (a) this study and (b) [25] for the reaction: 4Rh3S4(s) + S2(g) = 6Rh2S3(s).
T (K)ΔG (a) (kJ/mol)ΔG (b) (kJ/mol)|ΔG (b) − ΔG (a)|
900−96.016−86.939.086
950−86.619−78.9297.691
1000−77.219−70.9276.292
1050−67.812−62.9264.887
1100−58.392−54.9243.468
1150−48.956−46.9232.034
1200−39.499−38.9210.578
1250−30.018−30.9200.901
1300−20.509−22.9182.409
Near 1250 K, there is excellent agreement between the measured and calculated values. At lower temperatures, the calculated values fall below the measured values, and this discrepancy increases with decreasing temperature. The reason for this discrepancy is not obvious, but it is probably due to many experimental difficulties and uncertainties and/or a database collected from different methods.

3. Calculated Change of the Gibbs Free Energy (ΔfG°)

The calculated changes of the free energy of formation (ΔfG°) for selected PGM in layered intrusions and ophiolite complexes are negative (Table 1), supporting their stability compared to the separate elements (spontaneous reaction). They showed the following order: RuS2 < (Ir, Os)S2 < (Pt, Pd)S < (Pd, Pt)Te2; that is, increasing from compatible to incompatible noble metals and from sulfides to tellurides. In addition, due to the common presence of (Os–Ir–Ru)-oxides in nature, the ΔfG° for RuO2 was calculated (Table 1). Using the equations of the HSC program, the calculated ΔfG° values in standard conditions (Table 1) were extrapolated to a wide range of temperature, up to 1300 °C (Table 4), and the variation of the free energy versus temperature is plotted (Figure 2, Figure 3 and Figure 4). Since ∆H and ∆S are essentially constant with temperature unless a phase change occurs, the free energy versus temperature plot can be drawn as a series of straight lines, where ∆S is the slope and ∆H is the y-intercept [39,40]. Furthermore, since the ΔS values are always negative for these reactions, the ΔH-TΔS becomes larger with temperature, and the lines slope upwards (formation reactions are exothermic).
Figure 2. Plot of the Free energy (ΔfG°) against temperature for (A) RuS2 (laurite), OsS2 (erlichmanite) and IrS2 (unnamed iridium disulfide) and (B) Rh2S3 (bowieite) and Ir2S3 (kashinite). Data from Table 4.
Figure 2. Plot of the Free energy (ΔfG°) against temperature for (A) RuS2 (laurite), OsS2 (erlichmanite) and IrS2 (unnamed iridium disulfide) and (B) Rh2S3 (bowieite) and Ir2S3 (kashinite). Data from Table 4.
Geosciences 06 00002 g002
With respect to the recorded difference between the diagrams showing smooth curves (Figure 2A, Band 3B) and diagram for tellurides (PtTe2, PdTe2 and IrTe2 (Figure 3A), showing an abrupt change of direction at 450 °C (Figure 3B)), it may be related to the phase changes, as is exemplified by a phase change at 449.51 °C, for compositions between PtTe2 and Te (melting point of Te metal) in the related phase diagram [41,42].
Table 4. Extrapolation of the calculated free energy values at standard stage (ΔfG°) (Table 1) to a wide range of temperatures.
Table 4. Extrapolation of the calculated free energy values at standard stage (ΔfG°) (Table 1) to a wide range of temperatures.
T (°C)RuS2OsS2IrS2IrTe2PtTe2PdTe2Ir2S3Rh2S3PtSPdSOsAs2PtAs2RuO2
25−204.1−135.1−123.9−68.0−52.9−60.6−196.4−252.4−76.1−66.7−75.76−193.4−252.7
50−202.9−134.0−123.2−67.7−52.4−60.3−195.2−251.5−75.7−66.4−75.6−191.3−248.3
100−200.5−131.8−121.6−66.5−51.4−59.9−192.8−249.8−74.7−65.6−74.9−186.7−239.5
150−197.7−129.1−119.6−64.9−50.5−59.4−189.8−247.3−73.5−64.7−73.7−181.5−230.9
200−194.5−126.2−117.2−62.8−49.4−58.9−186.2−244.3−72.1−63.6−72.0−175.9−222.3
250−191.1−123.0−114.7−60.3−48.4−58.4−182.4−241.1−70.6−62.4−70.0−169.9−213.9
300−187.5−119.7−112.0−57.4−47.4−57.9−178.4−237.6−69.0−61.1−67.5−163.5−205.5
350−183.9−116.2−109.3−54.2−46.3−57.5−174.2−234.0−67.5−59.8−64.7−156.7−197.2
400−180.1−112.7−106.5−50.6−45.2−57.0−170.0−230.3−65.8−58.4−61.6−149.6−189.1
449.5------−46.7−44.1−56.6--------------
450−176.3−109.2−103.6−46.7−44.0−56.5−165.6−226.5−64.1−57.0−58.2−142.2−180.9
500−172.5−105.6−100.7−40.0−40.4−53.6−161.2−222.6−62.4−55.5−54.3−134.6−172.9
550−168.6−101.9−97.8−33.1−36.7−50.7−156.7−218.6−60.7−54.1−50.6−126.7−165.0
600−164.6−98.2−94.8−25.8−32.8−47.6−152.2−214.6−59.0−52.7−46.4−118.4−157.1
626.8----------−46.0------−51.9------
650−160.6−94.4−91.7−18.2−29.0−44.6−147.7−210.5−57.0−51.2−42.0−110.0−149.3
700−156.6−90.7−88.7−10.4−25.0−41.5−143.1−206.3−55.4−49.7−37.3−101.3−141.5
750−152.5−86.9−85.6−2.3−21.0−38.4−138.4−202.1−53.6−48.2−32.4−92.4−133.8
800−148.4−83.0−82.56.0−17.0−35.3−133.8−197.9−51.9−46.8−27.3−83.3−126.2
850−144.2−79.1−79.414.5−12.9−32.2−129.1−193.6−50.0−45.2−20.6−72.5−118.6
876.8------19.2−10.7−30.6--------------
900−140.0−75.3−76.323.3−8.7−29.1−124.4−189.3−48.3−43.8−12.8−60.8−111.1
926.8----−74.628.1----−121.9−187.0----------
950−135.8−71.3−73.132.2−4.5−26.0−119.7−185.0−46.5−42.3−4.9−48.8−103.6
1000−131.6−67.4−70.041.4−0.3−22.9−115.0−180.7−44.7−40.8−3.1−36.7−96.2
1026.8−129.3------2.0------−43.7--−30.1−92.3
1050−127.3−63.5−66.850.73.9−19.8−110.2−176.3−42.9−39.311.4−24.4−88.9
1100−123.1−59.5−63.660.28.3−16.7−105.5−171.9−41.1−37.819.8−11.9−81.5
1150−118.8−55.5−60.469.912.6−13.6−100.7−167.4−39.3−36.328.40.7−74.3
1200−114.4−51.5−57.279.717.0−10.5−96.0−163.0−37.5−34.837.113.5−67.0
1250−110.1−47.5−54.089.821.4−7.4−91.2−158.5−35.7−33.346.026.5−59.9
1300−105.7−43.5−50.899.925.8−4.3−86.4−154.0−34.0−31.855.039.6−52.7
A particular case of the free energy change accompanying a chemical reaction is the standardfree energy of formation, which is the free energy change accompanying the formation of one mole of a compound from the constituent elements, all being in their standard states. The change in the standard free energy with temperature is due to the changes ΔH° and ΔS° (this is related to the standard enthalpy and entropy ΔG° (ΔG° =ΔH°−T·ΔS°) and, of course, to the change in the temperature itself. It has been assumed that the free energies of formation of the elements in their standard states are, by convention, taken to be zero; the enthalpy scale is fixed by defining H = 0, describing the most stable phase of the pure elements, and the enthalpy of compounds contains the enthalpy of formation Hf from elements, which is usually measured calorimetrically [25,26,27,28]. Thus, the enthalpy of the compound is calculated by adding the enthalpy of formation to the experimental enthalpy difference H(T)–H [27].
H ( T ) = H f ( 298.15 ) + 298.15 T C p d T + H t r
where Hf(298.15) is the enthalpy of formation at 298.15 K and Htr is the enthalpy of transformation of the substance.
Absolute entropy values can be calculated from the experimental heat capacity values through the equation:
S = S ( 298.15 ) + 298.15 T ( Cp / dT ) + H t r / T t r
where S(298.15) is the standard entropy of the substance, which can be calculated by integrating Cp/T function from 0 to 298.15 K, T is temperature and Htr is the enthalpy of phase transformation at a temperature Ttr.
Figure 3. Plot of the Free energy (ΔfG°) against temperature for (A) PtTe2 (moncheite), PdTe2 (merenskyite), IrTe2 (shuangfengite) and (B) PtS (cooperite) and PdS (braggite). Data from Table 4.
Figure 3. Plot of the Free energy (ΔfG°) against temperature for (A) PtTe2 (moncheite), PdTe2 (merenskyite), IrTe2 (shuangfengite) and (B) PtS (cooperite) and PdS (braggite). Data from Table 4.
Geosciences 06 00002 g003
Entropies of crystalline substances approach zero at 0 K (−273.15 °C). This fundamental experimental observation is compatible with the third law of thermodynamics [27]. If entropy is understood as a measure of disorder, the disorder reaches its minimum at absolute zero and in perfect crystal structures. The heat capacity values of crystalline substances in equilibrium approaches the zero value at 0 K. Theoretical thermodynamic modeling cannot predict the temperature dependence of heat capacity at elevated temperatures. The term ΔCp is usually small, but seldom zero. Plots of ΔG° versus T are, therefore, slightly curved. The term ΔCp can usually be expressed as a function of temperature by an empirical equation, and the inclusion of this in the expression for ΔG° leads to the following fully-mathematical correlation, which is therefore adopted for fitting experimental heat capacities:
Cp = A + B × 10−3 × T + C × 105 × T−2 + D × 10−6 × T2
where A, B, C and D are coefficients estimated from experimental data [26].
Figure 4. Plot of the free energy (ΔfG°) against temperature for (A) PtAs2 (sperrylite), OsAs2 (Omeiite) and (B) RuO2, and RuS2 (laurite). Data from Table 4.
Figure 4. Plot of the free energy (ΔfG°) against temperature for (A) PtAs2 (sperrylite), OsAs2 (Omeiite) and (B) RuO2, and RuS2 (laurite). Data from Table 4.
Geosciences 06 00002 g004
Since ∆H and ∆S are essentially constant with temperature unless a phase change occurs, the free energy versus temperature plot can be drawn as a series of straight lines, where ∆S is the slope and ∆H is the y-intercept [26,27,28]. The lines slope for selected PGM (Figure 5) upwards because ΔS for these reactions is always negative, therefore ΔH-TΔS becomes larger with temperature because the formation reactions are exothermic (Δ H< 0).

4. Application of (ΔfG) Values to the Origin of PGM

The mineral stability, depending mainly on the availability of PGE/reactants and/or differences in pressure and temperature relative to standard conditions, may have an effect on the equilibrium constants [26,27,28]. Although the stability relationships between various phases can be studied using the experimental method, thermodynamics may give us an approach to the calculation of the Gibbs free energy and phase diagrams [28]. However, the requirements for the PGM stability and re-deposition remain still unclear, since the activity-composition models and relevant phase diagrams for PGM are very limited, and the PGM forming processes in natural systems are complicated [1,2,3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,28].
The calculated Gibbs free energy values may contribute to better understanding of the knowledge provided by geochemical data, mineral chemistry and structure relationships (microscopic scale) as they are obtained from the investigation of PGE/PGM in nature [43,44,45,46,47,48,49,50]. The more negative Gibbs free energy values for the RuO2 compared to those for the RuS2, in the range of temperature lower than 500 °C (Table 4, Figure 5), seems to be in good agreement with the common transformation of laurite to oxides/hydroxides in small chromite occurrences [13,48], in placer deposits [49], in the oxidized zone of the Great Dyke of Zimbabwe layered intrusion [50], in laterite deposits [51,52,53] and elsewhere. The equilibrium constant for the transfer of an ion from one phase to another is directly related to Gibbs free energy and the partition coefficient, which are defined by the behavior of the elements. The apparent epigenetic desulfurization at relatively low (350–500 °C) temperature and partial oxidation of primary laurite crystals has been facilitated by strong brittle deformation and crystal deformation and dislocations, making it possible to remove the sulfur [3,13,48].
Figure 5. A comparative diagram plotting the free energy (ΔfG°) against temperature. Data from Table 4.
Figure 5. A comparative diagram plotting the free energy (ΔfG°) against temperature. Data from Table 4.
Geosciences 06 00002 g005
Thus, a differential stress driving the deformation of ores, in particular those associated with ophiolite complexes, and changes in temperature, pressure, surface energies, solubility and the diffusion process, have a potential effect on PGM reactions [54]. Although the relative stability of RuO2 and RuS2 phases as functions of fO2 and fS2 (a redox reaction) is not constrained, the change of a mineral assemblage into a different assemblage means that the new association at the given pressure, temperature and mineral association has a lower free energy than the initial one [55].
On the basis of the calculated Gibbs free energy of formation for the RuS2 (−204.1 kJ/mol), which is lower compared to that for the OsS2 (−135.1 kJ/mol) (Table 1, Figure 2A), the reaction Os + RuS2 ↔Ru + OsS2 is not favorable thermodynamically (positive ΔfG°), although the substitution reaction of Ru by Os has been recorded in nature: laurite in chromitites associated with ophiolite complexes is commonly characterized by the general formula (Ru > Os > Ir)S2, but in the Othrys complex and other ophiolites of the Balkan Peninsula, it is Os-rich (Os > Ru > Ir)S2, reaching the composition of erlichmanite [4]. Such a composition has been interpreted as Os–Ru substitution during crystal fractionation, while the metal-sulfide equilibrium curves for PGE indicate that at given T, the formation of RuS2 requires fS2 lower than that for OsS2 [56].
Furthermore, the role of fS2, fTe2 and temperature has been emphasized for the composition and crystallization of sulfides and tellurides, scavenging Pt and Pd originally contained in the sulfide melt [41,42]. In addition, the earlier deposition of (Pt, Pd)S compared to (Pd, Pt)Te2, since (Pt, Pd)S is dominant in the UG2chromitite unit (Bushveld), which is located at a lower stratigraphic level compared to the Merensky Reef, where (Pd, Pt)Te2 are dominant PGM [18,43], seems to be consistent with the more negative free energy values in the former than in the latter (Table 1).
The occurrence of (Pd, Pt)-tellurides, such as moncheite (PtTe2), merenskyite (PdTe2), melonite [(Ni, Pd)Te2], kotulskite [(Pd, Ni)Te] and bismuth-tellurides [(Pt, Pd)(Te, Bi)2] in certain porphyry-Cu–Au ± Pd ± Pt systems [44,45,46,47], may be related to fS2, fTe2 and temperature, as well [41,42]. Assuming that the elevated values of the Pd/Pt ratios, the extremely low Cr contents (<1 ppm) in high Cu–Pd–Pt-grade ores, negative correlation between Cr content and the Pd/Pt, δ18O values in porphyry deposits of the Balkan Peninsula suggest genesis from more evolved mineralized fluids in porphyry Cu–Au–Pd–Pt deposits [44,45,46,47], the dominance of (Pd-Pt)-tellurides versus (Pt, Pd)S may be related to fS2, fTe2 and temperature variation. In addition, the common occurrence of the (Pd, Pt ± Bi)Te2, merenskyite (the main PGM in porphyry Cu–Au–Pd–Pt deposits), at the peripheral parts of chalcopyrite [44,45,46,47], may be related to the lower free energy of formation (ΔfG) values for chalcopyrite (−369.6 kJ/mol [57]) compared to that for merenskyite (−70 kJ/mol), suggesting the subsequent deposition of the latter, under appropriate fTe2 and temperature conditions. Thus, the combination of geochemical characteristics of a porphyry deposit indicating an evolved mineralized system, texture relationships of minerals and free energy data for those minerals data may provide evidence for the existence of (Pd, Pt)-mineralization.
The strong effect of As on PGE partitioning has been documented by the formation of stable high-temperature phases between As and Pt or Rh, including sperrylite (PtAs2), hollingworthite (Rh, Pt, Pd)AsS, and platarsite (Pt, Rh, Ru)AsS [5,13,14,15]. The presented (ΔGf) values for sperrylite (PtAs2) (Table 1; Figure 5) seem to be consistent with its common occurrence in chromitites as an interstitial phase between chromite grains [5,13,14,15].
Therefore, although the application of calculated Gibbs free energy values for selected PGM (Table 1) is not clear on the PGE behavior during the evolution of the mineralized systems, the free energy of PGM formation values can be compared to geological and mineralogy data and structure relationships observed in rocks associated with PGM to predict their occurrence and stability.

5. Conclusions

Although there are uncertainties in heat capacity and enthalpy data for PGM and the applications of thermodynamic experiments are limited, the application of thermodynamics to the study of nature systems can be compared to phase relations observed in rocks associated with PGM to predict their occurrence and stability. The preliminary results presented for the Gibbs free energy of the PGM allow us to present the following conclusions:
  • The calculated free energy of formation (ΔGf) for selected PGM were negative values, indicating that these minerals are more stable than the separate elements at standard conditions (the process is spontaneous).
  • The calculated Gibbs free energy of formation in the order RuS2 < OsS2 < (Pt, Pd)S < (Pd, Pt)Te2 is increasing from sulfides to tellurides, and they are consistent with their compatibility.
  • The evaluation of the accuracy of the method of calculation of the Gibbs free energy of formation using the HSC program was made by comparison with those obtained using a solid-state electrochemical technique by previous authors and was found to be in a good agreement.
  • The lower values of the calculated Gibbs free energy of formation (ΔG°f) for the oxide RuO2 compared to that for RuS2 at temperatures lower than 500 °C are consistent with the observed transformation of laurite to oxides/hydroxides at relatively low temperatures.
  • Although much more experimental work is required, the compilation of preliminary Gibbs free energy values with geological, mineralogical data and phase relations in natural systems suggests that a thermodynamic approach on PGM may contribute to the better understanding of the PGM thermodynamic behavior and PGE mineralization.

Acknowledgments

Many thanks are expressed to John Bowles. His constructive criticism and suggestions on an earlier draft of the manuscript, as well as the linguistic improvement of this work are greatly appreciated. The reviewers are greatly acknowledged for their constructive criticism and suggestions for improvement of our manuscript.

Author Contributions

Spiros Olivotos performed all calculations of the Gibbs free energy and diagrams. Both authors contributed to the evaluation of the obtained data and wrote the paper.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Barnes, S.-J.; Picard, C.P. The behaviour of platinum-group elements during partial melting, crystal fractionation, and sulphide segregation: An example from the Cape Smith Fold Belt, northern Quebec. Geochim. Cosmochim. Acta 2013, 57, 59–87. [Google Scholar]
  2. Economou-Eliopoulos, M. Platinum-group element distribution in chromite ores from ophiolite complexes: Implications for their exploration. Ore Geol. Rev. 1996, 11, 363–381. [Google Scholar] [CrossRef]
  3. Prichard, H.; Economou-Eliopoulos, M.; Fisher, P.C. Platinum-group minerals in podiformchromitite in the Pindos ophiolite complex, Greece. Can. Mineral. 2008, 46, 329–341. [Google Scholar] [CrossRef]
  4. Garuti, G.; Zaccarini, F.; Economou-Eliopoulos, M. Paragenesis and composition of laurite from chromitites of Othrys(Greece): Implications for Os-Ru fractionation in ophiolitic upper mantle of the Balkan peninsula. Mineral. Depos. 1999, 34, 312–319. [Google Scholar] [CrossRef]
  5. Ohnenstetter, M.; Johan, Z.; Cocherie, A.; Fouillac, A.M.; Guerrot, C.; Ohnenstetter, D.; Chaussidon, M.; Rouer, O.; Makovicky, E.; Makovicky, M.; et al. New exploration methods for platinum and rhodium deposits poor in base-metal sulphides—NEXTRIM. Trans. Inst. Min. Metall. Sect. B Appl. Earth Sci. 1999, 108, B119–B150. [Google Scholar]
  6. Hattori, K.H.; Cabri, L.J.; Johanson, B.; Zientek, M.L. Origin of placer laurite from Borneo: Se and As contents, and S isotopic compositions. Mineral. Mag. 2004, 68, 353–368. [Google Scholar] [CrossRef]
  7. Augé, T.; Maurizot, P. Stratiform and alluvial platinum mineralization in the New Caledonia ophiolite complex. Can. Mineral. 1995, 33, 1023–1045. [Google Scholar]
  8. Brenan, J.M.; Andrews, D. High-temperature stability of laurite and Ru–Os–Ir alloy and their role in PGE fractionation in mafic magmas. Can. Mineral. 2001, 39, 341–360. [Google Scholar] [CrossRef]
  9. Cabri, L.J. The Geology, Geochemistry, Mineralogy and Mineral Beneficiation of Platinum-Group Elements; Canadian Institute of Mining, Metallurgy and Petroleum: Montreal, QC, Canada, 2002. [Google Scholar]
  10. Weiser, T.W. Platinum-Group Minerals (PGM) in placer deposits. In The Geology, Geochemistry, Mineralogy and Mineral Beneficiation of Platinum-Group Elements; Cabri, L.J., Ed.; Canadian Institute of Mining, Metallurgy and Petroleum: Montreal, QC, Canada, 2002; pp. 721–756. [Google Scholar]
  11. Mungall, J.E. Magmatic geochemistry of the platinum-group elements. In Exploration for Platinum-Group Element Deposits; Mineralogical Association of Canada: Quebec, QC, Canada, 2005; pp. 1–34. [Google Scholar]
  12. Hanley, J.J. The aqueous geochemistry of the Platinum-Group Elements (PGE) in surficial, low-T hydrothermal and high-T magmatic hydrothermal environments. In Exploration for Platinum-Group Element Deposits; Mungall, J.E., Ed.; Mineralogical Association of Canada: Quebec, QC, Canada, 2005; pp. 35–56. [Google Scholar]
  13. Tsoupas, G.; Economou-Eliopoulos, M. High PGE contents and extremely abundant PGE-minerals hosted in chromitites from the Veriaophiolite complex, northern Greece. Ore Geol. Rev. 2008, 33, 3–19. [Google Scholar] [CrossRef]
  14. Kapsiotis, A.; Grammatikopoulos, T.A.; Tsikouras, B.; Hatzipanagiotou, K.; Zaccarini, F.; Garuti, G. Chromian Spinel composition and Platinum-group element mineralogy of Chromitites from the Milia area, Pindosophiolite complex, Greece. Can. Mineral. 2009, 47, 1037–1056. [Google Scholar] [CrossRef]
  15. Prichard, H.M.; Tarkian, M. Platinum and palladium minerals from two PGE-rich localities in the Shetland ophiolite complex. Can. Mineral. 1988, 26, 979–990. [Google Scholar]
  16. Prichard, H.M.; Neary, C.R.; Fisher, P.C.; O’hara, M.J. PGE-rich Podiform Chromitites in the Al AysOphiolite Complex, Saudi Arabia: An Example of Critical Mantle Melting to Extract and Concentrate PGE. Econ. Geol. 2008, 103, 1507–1529. [Google Scholar] [CrossRef]
  17. Locmelis, M.; Pearson, N.J.; Barnes, S.J.; Fiorentini, M.L. Ruthenium in Komatiitic Chromite. Geochim. Cosmochim. Acta 2011, 75, 3645–3661. [Google Scholar] [CrossRef]
  18. Maier, W.D.; Barnes, S.J. Platinum-Group Elements in Silicate Rocks of the Lower, Critical and Main Zones at Union Section, Western Bushveld Complex. J. Petrol. 1999, 40, 1647–1671. [Google Scholar] [CrossRef]
  19. Melcher, F.; Grum, W.; Simon, G.; Thalhammer, T.V.; Stumpfl, E.F. Petrogenesis of the giant ophiolitic chromite deposits of Kempirsai, Kazakhstan: A study of solid and fluid inclusions in chromite. J. Petrol. 1997, 38, 1419–1458. [Google Scholar] [CrossRef]
  20. Helmy, H.M.; Ballhaus, C.; Berndt, J.; Bockrath, C.; Wohlgemuth-Ueberwasser, C. Formation of Pt, Pd and Ni tellurides: Experiments in sulfide-telluride systems. Contrib. Mineral. Petrol. 2007, 153, 577–591. [Google Scholar] [CrossRef]
  21. Westrum, E.F., Jr.; Carlson, H.G.; Gronvold, F.; Kjekshus, A. Low-Temperature Heat Capacitiesand Thermodynamic Functions of some Palladium and Platinum Group Chalcogenides. II. Dichalcogenides; PtS2, PtTe2, and PdTe2. J. Chem. Phys. 1961, 35, 1670–1676. [Google Scholar] [CrossRef]
  22. Svendsen, S.R. High temperature Enthalpy and Decomposition Pressures of RuS2. Acta Chem. Scand. 1979, 33, 601–607. [Google Scholar] [CrossRef]
  23. Ezzoula, H.; Heindi, R.; Parson, R.; Tributsch, H. Studies on the stability of RuS2 single crystals and the photo-oxidation of halides. J. Electroanal. Chem. 1984, 165, 155–166. [Google Scholar] [CrossRef]
  24. Karzhavin, V.K. Sulfides, Selinides and Tellurides of Planinum and Palladium: Estimation of Thermodynamic Properties. Geochem. Int. 2007, 45, 931–937. [Google Scholar] [CrossRef]
  25. Jacob, K.T.; Gupta, P. Gibbs free energy of formation of rhodium sulfides. J. Chem. Ther. 2014, 70, 39–45. [Google Scholar] [CrossRef]
  26. Gupta, C.K. Chemical Metallurgy: Principles and Practice; John Wiley and Sons: Weinheim, Germany, 2003. [Google Scholar]
  27. Powell, R. Equilibrium Thermodynamics in Petrology: An Introduction; Harper & Row: London, UK, 1978. [Google Scholar]
  28. Cemic, C.L. Thermodynamics in Mineral Sciences: An Introduction; Springer: Berlin, Germany, 2005. [Google Scholar]
  29. Nordstrom, D.K.; Munoz, J.L. Geochemical Thermodynamics; Blackburn Press: Caldwell, NJ, USA, 2006. [Google Scholar]
  30. Barin, I. Thermochemical Data of Pure Substances; VCH VerlagsGesellschaft: Weinheim, Germany, 1989. [Google Scholar]
  31. Barin, I. Thermochemical Data of Pure Substances; Part I and part II; VCH Verlags Gesellschaft: Weinheim, Germany, 1993. [Google Scholar]
  32. Barin, I.; Knacke, O. Thermochemical Properties of Inorganic Substances; Springer: Berlin, Germany, 1973. [Google Scholar]
  33. Knacke, O.; Kubaschewski, O.; Hesselman, K. Thermochemical Properties of Inorganic Substances, 2nd ed.; Springer: Berlin, Germany, 1991. [Google Scholar]
  34. Kubaschewski, O.; Slough, W. Recent Progress in Metallurgical Thermochemistry, 1st ed.; Pergamon Press: Oxford, UK, 1969. [Google Scholar]
  35. Barin, I.; Knacke, O.; Kubaschewski, O. Thermochemical Properties of Inorganic Substances, Supplement; Springer: Berlin, Germany, 1977. [Google Scholar]
  36. Dean, J.A. Lange’s Handbook of Chemistry, Thermodynamic Properties; McGraw-Hill: New York, NY, USA, 1992. [Google Scholar]
  37. Bailey, S.M.; Churney, K.L.; Nuttall, R.L. The NBS tables of chemical thermodynamic properties.Selected values for inorganic and C1 and C2 organic substances in SI units. J. Phys. Chem. Ref. Data 1982, 11, 1–392. [Google Scholar]
  38. Bockrath, C.; Ballhaus, C.; Holzheid, A. Stabilities of laurite RuS2 and monosulfide liquid solution at magmatic temperature. Chem. Geol. 2004, 208, 265–271. [Google Scholar] [CrossRef]
  39. Ives, D.J.G. Principles of the Extraction of Metals, Monograph 3; Royal Institute of Chemistry: London, UK, 1960. [Google Scholar]
  40. Wang, C.X.; Yang, G.W. Thermodynamics of metastable phase nucleation at the nanoscale. Mater. Sci. Eng. 2005, 49, 157–202. [Google Scholar] [CrossRef]
  41. Kim, W.S.; Chao, G.Y. Phase relations in the system Pt–Sb–Te. Can. Mineral. 1990, 28, 675–685. [Google Scholar]
  42. Okimoto, H. The Pd–Te System (Palladium-Tellurium). J. Phase Equilib. Diffus. 1992, 34, 72–73. [Google Scholar] [CrossRef]
  43. Olivotos, S. Thermodynamic Controls on the Formation and Stability of Platinum-Group Elements. Master’s Thesis, University of Athens, Athens, Greece, 31 August 2015. [Google Scholar]
  44. Tarkian, M.; Stribrny, B. Platinum-group elements in porphyry copper deposits: A reconnaissance study. Mineral. Petrol. 1999, 65, 161–183. [Google Scholar] [CrossRef]
  45. Eliopoulos, D.G.; Economou-Eliopoulos, M. Platinum-group element and gold contents in the Skouries porphyry-copper deposit, Chalkidiki Peninsula, northern Greece. Econ. Geol. 1991, 86, 740–749. [Google Scholar] [CrossRef]
  46. Economou-Eliopoulos, M. Platinum-Group Element Potential of Porphyry Deposits; Mineralogical Association of Canada: Quebec, QC, Canada, 2005; pp. 203–245. [Google Scholar]
  47. Augé, T.; Petrunov, R.; Bailly, L. On the mineralization of the PGE mineralization in the Elastite porphyry Cu-Au deposit, Bulgaria: Comparison with the Baula-Nuasahi Complex, India, and other alkaline PGE-rich porphyries. Can. Mineral. 2005, 43, 1355–1372. [Google Scholar] [CrossRef]
  48. Garuti, G.; Zaccarini, F. In-situ alteration of platinum-group minerals at low temperature: Evidence from serpentinized and weathered chromitites of the Vourinos complex (Greece). Can. Mineral. 1997, 35, 611–626. [Google Scholar]
  49. Tolstykh, N.D.; Sidorov, E.G.; Krivenko, A.P. Platinum-group element placers associated with Ural-Alaska type complexes. In Exploration for Platinum-Group Element Deposits; Mineralogical Association of Canada: Quebec, QC, Canada, 2005; pp. 113–143. [Google Scholar]
  50. Oberthür, T.; Melcher, F.; Buchholz, P.; Locmelis, M. The oxidized ores of the Main Sulphide Zone, Great Dyke, Zimbabwe: Turning resources into minable reserves-mineralogy is the key. J. South. Afr. Inst. Min. Metall. 2013, 113, 191–201. [Google Scholar]
  51. Bowles, J.F.W.; Gize, A.P.; Vaughan, D.J.; Norris, S.J. Development of platinum-group minerals in laterites-initial comparison of organic and inorganic controls. Trans. Inst. Min. Metall. 1994, 103, B53–B56. [Google Scholar]
  52. Bowles, J.F.W. The development of platinum-group minerals in laterites. Econ. Geol. 1984, 81, 1278–1285. [Google Scholar] [CrossRef]
  53. Aiglsperger, T.; Proenza, J.A.; Zaccarini, F.; Lewis, J.F.; Garuti, G.; Labrador, M.; Longo, F. Platinum Group Minerals (PGM) in the Falcondo Ni-laterite deposit, Loma Caribe peridotite (Dominican Republic). Mineral. Depos. 2015, 50, 105–123. [Google Scholar] [CrossRef]
  54. Van der Weijden, C.H. Cahiers of Geochemistry; Utrecht University: Utrecht, The Netherlands, 2007. [Google Scholar]
  55. Anderson, D.L. Theory of the Earth; Blackwell Scientific Publications: Boston, MA, USA, 1989. [Google Scholar]
  56. Toulmin, P.; Barton, P.B. A thermodynamic study of pyrite and pyrrhotite. Geochim. Cosmochim. Acta 1964, 23, 641–671. [Google Scholar] [CrossRef]
  57. Robie, R.A.; Seal, R.R.; Hemingway, B.S. Heat capacity and entropy of bornite (Cu5FeS4) between 6 and 760 K and the thermodynamic properties of phases in the system Cu–Fe–S. Can. Mineral. 1994, 32, 945–956. [Google Scholar]

Share and Cite

MDPI and ACS Style

Olivotos, S.; Economou-Eliopoulos, M. Gibbs Free Energy of Formation for Selected Platinum Group Minerals (PGM). Geosciences 2016, 6, 2. https://doi.org/10.3390/geosciences6010002

AMA Style

Olivotos S, Economou-Eliopoulos M. Gibbs Free Energy of Formation for Selected Platinum Group Minerals (PGM). Geosciences. 2016; 6(1):2. https://doi.org/10.3390/geosciences6010002

Chicago/Turabian Style

Olivotos, Spiros, and Maria Economou-Eliopoulos. 2016. "Gibbs Free Energy of Formation for Selected Platinum Group Minerals (PGM)" Geosciences 6, no. 1: 2. https://doi.org/10.3390/geosciences6010002

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop