Andradite, ideal end-member formula Ca3Fe3+2Si3O12, is one of the common rock-forming garnets found in the Earth's crust. There are several outstanding questions regarding andradite's thermodynamic and physical property behavior. Three issues are: i) Could there be differences in the thermodynamic properties, namely heat capacity, Cp, between synthetic and natural andradite crystals, as observed in the Ca-garnet grossular, Ca3Al2Si3O12? ii) What is the thermal nature of the low-temperature magnetic-phase-transition behavior of andradite? and iii) How quantitative are older published calorimetric (i.e., adiabatic and DSC) heat-capacity results? In this work, four natural nearly end-member single crystals and two synthetic polycrystalline andradite samples were carefully characterized by optical microscope examination, X-ray powder diffraction, microprobe analysis, and IR and UV/VIS single-crystal spectroscopy. The IR spectra of the different samples commonly show a main intense OH− stretching band located at 3563 cm−1, but other OH− bands can sometimes be observed as well. Structural OH− concentrations, calculated from the IR spectra, vary from about 0.006 to 0.240 wt% H2O. The UV/VIS spectra indicate that there can be slight, but not fully understood, differences in the electronic state between synthetic and natural andradite crystals. The Cp behavior was determined by relaxation calorimetry between 2 and 300 K and by differential scanning calorimetry (DSC) methods between 150/300 and 700/950 K, employing the same andradite samples that were used for the other characterization measurements. The low-temperature Cpresults show a magnetic phase transition with a Néel temperature of 11.3 ± 0.2 K, which could be slightly affected by the precise electronic state of Fe2+/3+ in the crystals. The published adiabatic calorimetry results on andradite do not provide a full and correct thermal description of this magnetic transition. The calorimetric Cp measurements give a best estimate for the standard third-law entropy at 298.15 K for andradite of So ≈ 324 ± 2 J/mol ∙ K vs. the value of 316.4 ± 2.0 J/mol ∙ K, as given in an early adiabatic investigation. Both natural and synthetic crystals give similar So values within experimental uncertainty of about 1.0%, but one natural andradite, richer in OH, may have a very slightly higher value around So ≈ 326 J/mol ∙ K. Low-temperature DSC measurements made below 298 K agree excellently with those from relaxation calorimetry. The DSC measurements above 298 K show a similarity in Cp behavior among natural and synthetic andradites. A Cp polynomial for use above room temperature to approximately 1000 K was calculated from the data on synthetic andradite giving:
Cp (J/mol ∙ K) = 599.09 (±14) − 2709.5 (±480) · T−0.5 − 1.3866 (±0.26) · 107 · T−2 + 1.6052 (±0.42) · 109 · T−3.
Introduction
Andradite, ideal end-member formula Ca3Fe3+2Si3O12, is one of the common silicate garnets found in Earth's crust. Close to end-member andradite can occur in nature, for example, at the well-known locality of Val Malenco, Italy, where it occurs in serpentinite formed during the metamorphism of ultramafic rocks. There, it is yellow-green to green in color and the gem variety name for such crystals is known as demantoid. Other andradite occurrences are known and the crystals can deviate to various degrees from strict end-member andradite composition (e.g., Bocchio et al., 2010). Andradite-rich garnet, commonly in solid solution with grossular, can be found in nature. A common occurrence is in contact metamorphic rocks, especially skarns (Huckenholz & Yoder, 1971). Other types of occurrences have been reported including, for example, low-grade regionally metamorphosed rocks (T = 300 to 400 °C or less) including metabasites and volcanic sandstones (Coombs et al., 1977).
There have been a number of phase-equilibrium studies made to determine the stability of andradite under different P-T-fO2 conditions and also with other solid phases and under different fluid compositions (e.g., Huckenholz & Yoder, 1971; Gustafson, 1973; Suwa et al., 1976; Shoji, 1977; Taylor & Liou, 1978; Wykes et al., 2008). In addition to phase-equilibrium research, calorimetric measurements (i.e., adiabatic calorimetry and differential scanning calorimetry − DSC) were carried out on a synthetic andradite to determine its Cp behavior between 8 K and 1000 K (Robie et al., 1987). Low temperature Cp results allow the standard third-law entropy, So, at 298.15 K to be calculated. These various experimental results have been used to “optimize” the thermodynamic properties of andradite, as has been done in different internally consistent thermodynamic databases (Gottschalk, 1997; Chatterjee et al., 1998; Holland & Powell, 1998; 2011).
The vibrational spectrum of andradite has been studied by Raman polarized single-crystal (Kolesov & Geiger, 1998) and IR reflectance single-crystal (McAloon & Hofmeister, 1993) spectroscopy. The heat capacity and entropy of andradite can be calculated under model assumptions, together with other physical property input data, using such vibrational spectra (i.e., Kieffer, 1979a, b; 1980), as also done later by Madon et al. (1991) and Chopelas (2006).
The crystal structures of both natural (Novak & Gibbs, 1971; Adamo et al., 2011) and synthetic andradite (Armbruster & Geiger, 1993) have been studied by X-ray single-crystal diffraction and the two types are very similar structurally. The magnetic behavior of andradite from Val Malenco, Italy, was studied by low temperature 57Fe-Mössbauer measurements (Murad, 1984), which show that the electron spins of the Fe3+ cations order from the paramagnetic to antiferromagnetic state with a Néel transition temperature of 11.5 ± 0.1 K. The nature of this transition has also been investigated via ab initio calculations (Meyer et al., 2010). Finally, the UV/VIS spectrum of andradite has been measured multiple times (e.g., Manning, 1969; Burns, 1993; Adamo et al., 2009; Taran & Langer, 2000) focusing on the Fe3+-O charge-transfer as well as on the spin-forbidden bands.
One could possibly think that there is little left to research in terms of andradite's key physical properties. However, this is not the case and several important issues still need to be addressed and clarified. They are: i) It needs to be investigated if there could be small differences in thermodynamic, and here Cp, behavior between synthetic and natural andradite crystals as well as among natural crystals themselves. This general question has not been carefully researched in the mineral sciences. In terms of another Ca-bearing garnet, grossular, Ca3Al2Si3O12, which is one of the few silicates to be studied extensively in this regard, there are small but measurable differences in Cp behavior at low temperatures between natural and synthetic crystals (Dachs et al., 2012a). Small but meaningful differences in So result (∼254 vs. ∼260 J/mol ∙ K, respectively). It is not known if this could be the case with andradite, which shares mineralogical and petrological similarities with grossular, because both garnets can occur in similar rock types and crystallize under similar P-T conditions. And here, the issue of careful sample characterization plays an essential role. Most older (i.e., say roughly pre 1990) calorimetric investigations undertook only rudimentary characterization measurements (i.e., optical and X-ray examination and perhaps a microprobe analysis), at least by today's standards, of the mineral(s) to be studied. These results were, moreover, in general, cursorily reported. Thus, it is essentially impossible to be able to make quantitative crystal-chemical and structural interpretations of the macroscopic Cp results; ii) The precise thermal magnetic-phase-transition behavior of andradite needs further study. Here, it must be noted that low-temperature adiabatic calorimetry Cp measurements are/were often not made below about 8 K. This is the case for andradite (Robie et al., 1987) as well as other transition-metal-bearing garnets. They show cooperative magnetic transitions whose thermal signatures extend below 8 K and approach 0 K (Klemme et al., 2005; Dachs et al., 2009; Dachs et al., 2012b), and iii) Improvements in power-compensated DSC methods allow more precise determinations of Cp(T) from roughly 150 or 300 K to 900–1000 K compared to data obtained using first-generation devices and measurements made from the 1970s through the 1980s (see discussion in, e.g., Bosenick et al., 1996, for the case of grossular).
In light of these various issues, this work focuses on the thermophysical Cp behavior of different well-characterized synthetic and natural andradite samples from 2 K to 900 K. The calorimetric results are also considered in terms of the IR and UV/VS spectra and compositional data.
Analytical and experimental methods
Natural and synthetic samples and optical examination
Four natural, nearly end-member composition andradite single crystals from different localities and two synthetic polycrystalline samples were investigated by calorimetry. They are listed and described in Table 1. The much larger natural crystals were prepared as small platelets, polished on both sides, of roughly 0.5 mm thickness for the compositional (microprobe), spectroscopic (some additional platelets were made thinner to allow IR characterization) and calorimetric measurements.
All andradites were examined under binocular and polarizing microscopes. The optical properties (e.g., possible presence of inclusions and possible anisotropy) of the single-crystal platelets were examined and photographed. Photos of the various crystals, excepting the fine-grained polycrystalline sample SD23, are shown in Figs. S1 and S2 (freely available online as supplementary material linked to this article on the GSW website of the journal: https://pubs.geoscienceworld/eurjmin/). The synthesis and characterization of sample And #27 (Fig. 1e and 1f) is described in Armbruster & Geiger (1993). They studied its crystal structure at five different temperatures between 100 K and 500 K, in addition to measuring its 57Fe Mössbauer and IR (3800 to 3200 cm−1) spectra. The second andradite sample, SD23, was synthesized in a manner similar to And #27 from a glass in a piston-cylinder device (in a large Au capsule at P = 15 kbar and T = 900 °C. PtO2 was used in both syntheses in order to maintain a high oxygen fugacity in the charge.
Microprobe analysis
The composition of the natural andradite platelets was determined using a JEOL microprobe at Kiel University, Germany, and WDS methods were applied under conditions of 15 kV and 15 nA with a focused beam size of 1 μm as done in other studies on garnet (e.g., Dachs et al., 2012a, b). Due to the generally small sizes (less than roughly 40 μm) of most crystals and difficulties in obtaining quantitative microprobe results on synthetic garnets (Dachs et al., 2009; Fournelle & Geiger, 2010), the synthetic andradites were not measured. The standards used for analysis were corundum for Al, wollastonite for Ca and Si, and a natural fayalite-rich olivine for Fe, V metal for V, chromite for Cr, tephroite for Mn, rutile for Ti, and forsterite for Mg. The data correction program employed was CITZAF.
X-ray powder diffraction
X-ray powder patterns were recorded for the two synthetic samples. Sample And #27 was measured between 5 and 110° 2θ with a step size of 0.02° with 20 sec/step with a Siemens D-500 diffractometer (Salzburg) having a graphite secondary monochromator with Cu Kα radiation (40 kV, 35 mA). Synthetic andradite SD23 was measured over the 2θ range 10° to 90° using a Siemens D5000 diffractometer (Kiel) with Cu Kα radiation (40 kV, 30 mA). Diffraction patterns were recorded with 0.02° 2θ increments and up to 30 s counting time per step. Silicon powder NBS SRM 640b was used as an external calibration standard. In both cases, we estimate that minor phases, if well crystalline, are detectable above about the 0.5% level.
FTIR spectroscopy
The FTIR spectra were collected on five single-crystal platelets (Fig. 1), doubly polished on both sides, using apertures of either 50 or 200 μm diameter. The main purpose of the IR measurements was to characterize any structural OH- in the garnet. The preparation and experimental measuring conditions are similar to those described in Geiger & Rossman (2018). Spectra were collected between 10,000 and 2,000 cm−1 using a Thermo-Nicolet iS50 FTIR spectrometer (at Caltech). Efforts were made to make measurements on different parts of the natural crystal platelets, and also on those minor areas that appeared to have inclusion phases.
UV/VIS optical absorption single-crystal spectroscopy
The UV/VIS single-crystal spectrum of one of the larger synthetic crystals taken from the sample And #27 was recorded in the 380 to 1100 nm range at about 1.5 nm resolution. This was done using a homebuilt microspectrometer system consisting of a 1024-element Si diode-array detector coupled to a 1/3 meter grating spectrometer system attached via fiber optics to a highly modified NicPlan infrared microscope containing a calcite polarizer (at Caltech). Spectra in the 300 to 480 nm region were also obtained with a homebuilt system with an Ocean Optics spectrometer and a deuterium lamp.
Calorimetric measurements
Low-temperature relaxation calorimetry
Low-temperature (i.e., 2 to 300 K) heat-capacity behavior was investigated with the Physical Properties Measurement System (PPMS) constructed by Quantum Design® (e.g., Lashley et al., 2003; Dachs & Bertoldi, 2005). The relaxation calorimeter and its experimental set-up and method have been described a number of times, including investigations on garnet (e.g., Dachs et al., 2012a; 2012b). Thus, only a brief summary is given here.
Heat capacity was measured at 60 different temperatures and three times at each temperature on cooling from 300 K with a logarithmic spacing. A complete PPMS experiment to determine Cp comprises an ‘addenda run’ and a ‘sample run’. The first measurement determines the heat capacity of the empty sample platform of the calorimeter plus Apiezon N grease that facilitates thermal contact between the platform and the sample ± container. In the second measurement, the sample ± container is included and the heat capacity of the whole ensemble is measured. The net heat capacity of the sample ± container is then given by the difference between the two measurements. An uncertainty of ±0.02 mg for the sample weight was adopted for converting the PPMS data from units of µJ ∙ K−1 to units of J/mol ∙ K.
The natural andradite crystal platelets were directly mounted onto the measurement platform for Cp determination. They weighed between 6 and 36 mg (Table 1). The molar mass of each natural sample was calculated using the microprobe analysis results. The two synthetic samples were measured by enclosing the polycrystalline material/powder in Al foil. Their sample masses were 11.09 mg (And #27) and 13.20 mg (SD23). The use of polished single crystals permits better sample coupling during PPMS measurement (Dachs & Bertoldi, 2005; Benisek & Dachs, 2008). Sample coupling is defined as the ratio 100 Kg/(Kg + Kw), where Kg is the thermal conductance between the sample and the measurement platform and Kw is the thermal conductance of the wires that attach the platform to the puck frame of the calorimeter (see Hwang et al., 1997; Lashley et al., 2003; Dachs & Bertoldi, 2005, for more details). The closer the sample coupling is to 100%, the better the thermal conductance between the sample platform and the sample, which ensures a more robust heat-capacity determination via relaxation calorimetry. Experience indicates that an accuracy of about 0.2 (or even 0.1)% above ∼30 K in Cp and the standard entropy, S°, can be achieved for single crystals (Benisek & Dachs, 2008; Dachs & Geiger, 2018). It is between about 0.2% and 0.5% for measurements on powders (weighing about 10–20 mg) contained in thin Al-metal foil.
DSC measurements
Heat-capacity determinations were made in two different DSC laboratories, namely at Kiel and Salzburg Universities. The labs and experimental set-ups have already been described (for the Kiel laboratory see Bosenick et al., 1996, and for the Salzburg laboratory see Benisek et al., 2010 or Dachs et al., 2012a, 2012b). The Cp measurements were made between about 150 and 950 K in the Kiel lab using a power-compensated Perkin Elmer DSC 7 device and between about 300 and 750 K in the Salzburg lab using a Perkin Elmer Diamond DSC. Briefly, samples were contained in thin Au (Kiel) or Al (Salzburg) pans and lids, with the former containing a loose polycrystalline powder or a single-crystal platelet. A Cp determination usually consisted of three or more separate measurements of a blank, a reference and a sample measurement. Before a sample measurement, the DSC was calibrated and checked with a reference run using a synthetic single-crystal platelet of corundum (31.764 mg). Its heat-capacity values were taken from a National Bureau of Standards Certificate (Ditmars et al., 1982). Each garnet sample was measured three times in this manner. These data were averaged to obtain the final Cp(T).
Calorimetric data evaluation
The standard molar entropy at 298.15 K (i.e., third-law entropy), So, was calculated from the final averaged Cp data set using the Mathematica® functions NIntegrate and Interpolation for linear interpolation between data points via:
,(1)
assuming ST = 0K = 0. The Cp behavior for the minor low-temperature interval from 2 K down to about 0 K, which was not measured, was extrapolated using results from the synthetic sample SD23. This interval corresponds to So–2K = 1.3 J/mol ∙ K and this value was considered the same for all other studied andradites.
The values for volume, V (using ao = 12.054 Å), thermal expansion, α, (20 + 0.01 · T 10−6/K) and bulk modulus, KT (162 GPa), of andradite were taken from Chopelas (2006). This Komada-Westrum model was applied to the Cp data of synthetic SD23 andradite, which is considered to best represent end-member andradite.
Results
Sample description and optical properties
All natural andradites have a similar light yellow-green color (Fig. S1 and Table 1). The garnet crystals can contain very minor included solid phases. Sample 113-102 from Dobrovka, Nizhniy Tagil, Russia, contains very small amounts of chrysotile occurring as fine hair-like fibers (see Phillips & Talantsev, 1996, for a full description), as shown in Fig. S1g. Sample VM-11, from the well-known demantoid locality at Val Malenco, Italy, also appears to contain very minor amounts of a fine hair-like phase (see also Adamo et al., 2009). Sample Bal-1, occurring as euhedral crystals with elongated {1 1 0} faces, from Balochistan, Pakistan, contains, as well, very small amounts of chrysotile and, in addition, tiny Cr-bearing magnetite crystals, as described by Adamo et al. (2015). Sample 4282, whose locality is unknown to us (this andradite has a similar color and optical behavior and a very similar IR OH−-band spectrum to an andradite reported from Madagascar − Pezzotta et al., 2011, and Adamo et al., 2011), appears to be largely free of solid-inclusion phases. Of the different garnets that were studied, only the natural sample 113-102 and the two synthetics are optically isotropic. The other crystals show rather complex anisotropic behavior similar to that observed in many natural grossulars (e.g.Dachs et al., 2012a), as shown in Fig. S2.
Microprobe analysis
The compositions of the natural andradite samples, based on about 30 to 40 point analyses taken from a line traverse across each platelet, which were averaged to obtain the final composition, are given in Table 2a. The crystals showed no major compositional zonation. The crystal-chemical formulae, broken down into various end-member garnet components, were calculated using the program of Locock (2008) and are given in Table 2b. Slight deviations from strict end-member andradite composition are noted. Based on the formulation of Locock (2008), samples VM-11 and Bal-1 show 1.0 to 2.5% of a “remainder component” that could not be assigned precisely to any standard garnet component. Difficulties in obtaining exact garnet stoichiometry are probably due to several factors. These include structural OH− (see below), possible recording of minor solid inclusions in a few of the random point measurements, difficulties in calculating Fe3+/(Fe2+ + Fe3+) ratios precisely using olivine with Fe2+ as a Fe standard, and finally, normal analytical uncertainty. In summary, we consider the microprobe results to be good and all natural samples are considered to be greater than about 99 mol% andradite in composition.
X-ray powder diffraction
The X-ray powder diffractogram of the synthetic sample And #27 showed very small amounts of fine-grained Pt. It originates from the PtO2 that was used in the high P-T synthesis procedure, as described by Armbruster & Geiger (1993). Sample SD23 may have a trace of wollastonite.
FTIR spectroscopy
The FTIR single-crystal spectra of the studied andradite crystals are shown in Fig. 1 in the wavenumber region from 3800 to 3200 cm−1. The spectra presented here are considered to be free of any inclusions and, thus, representative of structural OH− in the andradite. Samples Bal-1, 113-102 and And #27 are characterized by a main asymmetric OH− stretching band located at 3563 ± 1 cm−1. This mode energy agrees perfectly with the IR results in Adamo et al. (2009, 2011), but is different from that stated in Amthauer & Rossman (1998), where a value of 3555 cm−1 is presented (however, our reanalysis of these latter spectra gives a value of 3563 cm−1). Sample VM-11 gives spectra with one notable additional peak at 3605 cm−1 (see also Adamo et al., 2009, 2011). The spectrum of sample 4282 is radically different from the others, as it contains multiple low-intensity OH bands (cf., Amthauer & Rossman, 1998; Adamo et al., 2011). The assignment for the asymmetric band at 3563 cm−1 is to a hydrogarnet substitution (Geiger & Rossman, 2018), whereas the other bands must represent OH− in an another structural site or cluster.
We were also able to record a few spectra on two samples where additional weak, apparently non-garnet-related OH− bands, could be observed. Mostly, but not always, these bands occurred in those small areas of the natural crystals that were more optically turbid and appeared to contain inclusion phases. For example, one measuring spot on the crystal VM-11 produced a spectrum showing an asymmetric band located at about 3678 cm−1. This band can be best assigned to OH− in talc (Petit et al., 2004; Parry et al., 2007 − we note, further, that this band may also be present in the spectrum of sample GRR 1263, San Benito Co., CA, of Amthauer & Rossman, 1998). We did not observe any overt OH− bands that could be assigned to chrysotile in, for example, andradite 113-102, even though we made measurements on regions that appeared to contain fine hair-like inclusions. Chrysotile and lizardite show OH− bands at 3691–3696 cm−1 and 3647–3650 cm−1 and a possible third one around 3685 cm−1 (Post & Borer, 2000). We do not have an explanation for the lack of observable OH− bands that could be related to the possible presence of minor fine, hair-like chrysotile.
The intensities of the OH− bands allow the amount of structural “H2O” in the garnet to be determined. There is no direct IR calibration for OH− in andradite and, thus, the estimated H2O amounts must be based on calibrations made for other garnets, namely grossular, pyrope or spessartine (Rossman, 2006). The total integrated absorbance per cm of the OH− bands of the three andradite samples, along with their respective calculated H2O contents in ppm, based on three reference garnets, are given in Table 3. They range between about 0.006 wt% for And #27 to 0.24 wt% for Bal-1 using the calibration for grossular. Using the calibration for spessartine gives roughly similar H2O concentrations for the different crystals, while that for pyrope would give considerably higher values.
UV/VIS spectroscopy
Figure 2 shows UV/VIS spectra between 380 to 1100 nm of three andradite garnets, namely synthetic And #27 and two representative natural crystals (GRR 48 and GRR 3137) from the collection of G.R. Rossman. The absorption features for the natural crystals are very similar to those described in the literature (e.g., Manning, 1969; Burns, 1993; Adamo et al., 2009; Taran & Langer, 2000). A very intense Fe3+-O charge-transfer band dominates andradite's spectrum at lower wavelengths. Only a part of the low-energy flank of this intense band is observable in Fig. 2. Superimposed on this flank is a sharp asymmetric peak at about 440 nm and at higher wavelengths two weak broad bands are located. These spectroscopic features represent various spin-forbidden transitions, as discussed in Taran & Langer (2000). They analyzed the exact fitting of the absorption features and made band assignments and crystal-chemical interpretations.
What is slightly new and of interest, here, is the spectrum of sample And #27. It shows different absorption behavior that cannot be precisely interpreted. It is notable that the color of this synthetic crystal is also slightly different from those of the natural andradites (Fig. S1), being more golden in color rather than light green. These UV/VIS spectroscopic results are relevant in terms of the magnetic phase transition behavior of andradite as discussed below.
Low-temperature (0 to 300 K) heat-capacity behavior and standard third-law entropy of andradite
The raw Cp data for the two synthetic polycrystalline andradites (SD23 and And #27) and the four natural andradite platelets are given in Appendix 1 (supplementary material). Interpolated Cp behavior is shown in Fig. 3. Table 4 (bottom) summarizes the salient results of the PPMS and DSC measurements. In terms of low-temperature relaxation calorimetry, sample coupling lies in the range of 98–100%. For the natural crystals 4282, Bal-1, VM-11 and the synthetic sample SD23, the agreement between PPMS- and DSC-measured heat capacity around ambient temperature was excellent with maximum deviations of 0.5%. The sample 113-102 showed a slightly larger deviation of 1% and, therefore, its PPMS Cp data was adjusted (see Dachs & Benisek, 2011). No data adjustment was made for the other samples. The Cp values for And #27 are slightly different from those of the other samples because of the presence of a small amount of Pt. These data were not used to calculate S°, but the low-temperature Cp magnetic behavior of the sample is analyzed below in connection with its UV/VIS spectrum.
The individual standard third-law entropy values for the various andradite samples studied are listed in Table 4. They range between 321.8 ± 2.2 and 325.7 ± 2.2 J/mol ∙ K and, thus, overlap within experimental error. We note, however, that andradite Bal-1 has the highest S° with a value of 325.7 ± 2.2 J/mol ∙ K. This very slightly larger value might reflect the larger OH− concentration in this andradite (Table 3) compared to the other crystals. The hydrogarnet substitution should, based on calorimetric work on katoite (Ca3Al2O12H12 - Geiger et al., 2012), lead to higher Cp and S° values in silicate garnet.
High-temperature (300 to 900 K) heat-capacity behavior of andradite
The raw DSC data for the various studied andradite samples are also compiled in Appendix 1 that lists the mean Cp values and standard deviation. Three samples were used for extensive DSC measurements to higher temperatures, namely synthetic SD23, 113-102 and VM-11. Sample SD23 was studied at Kiel and Salzburg and the latter two only in Salzburg. The Cp values for SD23, measured in Kiel, 113-102 and VM-11 are shown in Fig. 4. The agreement among the three data sets is excellent and Cp(T) is the same within ±0.5%. This uncertainty probably approaches the optimal experimental precision that one can obtain with power-compensated DSC methods (at least up to roughly 500–600 K).
We choose the DSC data for synthetic andradite SD23 to fit a Cp polynomial for use above room temperature (Berman & Brown, 1985), because the data set for this sample covers the largest temperature interval and the Cp data are well behaved. Applying the method of non-linear least squares yields:
Cp (J/mol ∙ K) = 599.09 (±14) − 2709.5 (±480) · T−0.5 − 1.3866 (±0.26) · 107 · T−2 + 1.6052 (±0.42) · 109 · T−3 (4).
This polynomial describes all the Cp data within ±0.5% and can be used safely up to about 1000 K.
Discussion
Cp behavior of andradite below 300 K: synthetic vs. natural garnets and a comparison of adiabatic, relaxation and DSC results
Although the important subject matter of possible differences in thermodynamic behavior between synthetic vs. natural minerals has been broached in the past (e.g., Helgeson et al., 1978; Westrum, 1978), surprisingly little quantitative research has been done in this direction. Probably the most extensive investigation in this regard is in terms of calorimetry centered on the garnet grossular Ca3Al2Si3O12 (see discussion in Dachs et al., 2012a). Here, the Cp results on a number of samples and from different investigations are extensive. It was shown that synthetic grossular and natural nearly end-member crystals have small but measurable and meaningful differences in Cp(T) behavior, giving rise to S° values from ∼260 vs. ∼254 J/mol · K, respectively. The Cp of natural crystals shows smaller values at temperatures roughly between 10 and 100 K.
Andradite has similarities to grossular in terms of both its mode of occurrence and crystal chemistry. Both can occur from lower- to higher-grade metamorphic rocks, usually in solid solution with one another. Grossular and andradite are also commonly marked by complex anisotropic optical behavior (Fig. S2), but the precise origin for it is not fully understood (e.g., Hirai et al., 1982). A single-crystal structure refinement on an anisotropic natural andradite yielded triclinic space group symmetry, I-1 (Kingma & Downs, 1989) and grossular can show symmetries lower than cubic (see references in Dachs et al., 2012a). This may indicate possible partial long-range cation ordering. On the other hand, in terms of andradite, Adamo et al. (2011) refined several natural single crystals (including one from Val Malenco, Italy) in cubic symmetry Ia-3d and optically anisotropic andradites have been refined in space group Ia-3d using high-resolution synchrotron radiation (Antao et al., 2015). The cubic space group does not allow for long-range atomic ordering. Unlike the case for grossular, based on the extensive results of this study, we do not observe any distinctly measurable differences in the Cp and S° behavior between isotropic (113-102) and anisotropic (VM-11, 4282, and Bal-1) andradite crystals.
Robie et al. (1987) studied the Cp of a single, synthetic andradite sample using adiabatic calorimetry and an early-generation power-compensated DSC device. We can compare their results with those from our relaxation calorimeter and later-generation DSC made down to roughly 150 K. Consider first Cp behavior below 300 K. Fig. 5 shows the raw adiabatic Cp data from Robie et al. and our results obtained on andradite SD23. The agreement between the lower temperature (<300 K) DSC and relaxation data is excellent. On the other hand, the adiabatic Cp data beginning at about 150 K and with increasing temperature lie at slightly higher values. For example, Robie et al. (1987) report a Cp(298.15 K) value of 351.9 J/mol · K, while the various samples from this study show Cp(298.15 K) from about 345 to 348 J/mol · K. It is to be noted that the synthetic andradite of Robie et al. (1987) contained about 2% wollastonite, which they considered to be related to the synthesis route adopted. They stated that they corrected their Cp results for the presence of wollastonite. However, their Table 3 of smoothed thermodynamic properties, including Cp, show similar values as their raw experimental Cp data. The reason for this is not clear to us.
Based on our more extensive results on both well-characterized natural and synthetic samples, we argue that the Cp data, herein, describe better the thermophysical behavior of andradite at low temperatures. Robie et al. (1987) present a S° value of 316.4 J/mol ∙ K, lower by about 3% compared to our recommended value of S° ≈ 324 J/mol ∙ K. On further closer inspection of the data of Robie et al., other issues come to light and, now, we analyze the low-temperature magnetic phase transition in andradite in detail.
Magnetic phase transition
An early X-ray crystal-structure determination of andradite (Menzer, 1928) and later refinements of both natural (Novak & Gibbs, 1971; Armbruster & Geiger, 1993; Adamo et al., 2011) and synthetic crystals (Armbruster & Geiger, 1993) have been made. The structures of both types of crystals are very similar with no marked differences. The Ca cation occurs in eight-fold coordination at the 24c Wyckoff crystallographic site and Fe3+ is six-fold coordinated at the 16a site. The Fe3+ cation is characterized as having one electron in each of the five different d orbitals. Thus, no spin-allowed electronic transitions are permitted, as shown by the UV/VIS spectra (Fig. 2). The electron spins are thermally disordered (i.e., paramagnetic state) at ambient conditions, but they start to order at very low temperatures giving rise to an antiferromagentic transition at 11.5 ± 0.1 K (i.e., Néel temperature − TN,), as based on 57Fe Mössbauer spectroscopic measurements (Murad, 1984).
The thermal behavior of this magnetic transition is clearly shown by our closely spaced relaxation-calorimetry measurements made down to 2 K (Fig. 6 and Appendix 1). The adiabatic calorimetry measurements (Robie et al., 1987), on the other hand, were only made down to 8 K and with limited data coverage. Of critical importance, here, we note that the two lowest temperature Cp data (i.e., at 8.17 K and 8.40 K) were used in the description of the transition. However, we think that our more extensive low-temperature relaxation measurements down to 2 K give a better description of the thermal behavior (thus, for example, the Cp value at 5 K in Table 3 (Robie et al., 1978) of smoothed and fitted thermodynamic properties is too low).
On a further note, our Cp results show that TN for sample And #27 is very slightly lower in value than the other andradites (Fig. 6). This synthetic garnet shows TN = 10.7 ± 0.3 K, whereas TN for all other andradite samples is 11.3 ± 0.2 K. The latter value is in excellent agreement with the Mössbauer results. The precise physical explanation for the slightly lower TN value for And #27 is not clear, but it is noteworthy that it also shows a different UV/VIS spectrum compared to natural nearly end-member composition andradite crystals. It is quite possible that this synthetic crystal has a slightly different crystal-chemical and/or electronic state than the natural garnets. It may contain, for example, very minor Fe2+ at the 24c and/or 16a sites that are below the detection limits of 57Fe Mössbauer spectroscopy, which do not show measurable Fe2+ (Armbruster & Geiger, 1993).
In our previous Cp investigation on the garnet almandine, Fe3Al2Si3O12, we observed slight differences in TN and thermal peak behavior relating to a low-temperature magnetic transition at 9.2 K (Dachs et al., 2012b). The transition is again caused by an ordering of electron spins but in this case of Fe2+ at the 24c site. The exact TN appears to be affected by small amounts of Fe3+ present at the 16a site. It follows that magnetic phase transition behavior in garnet appears to be sensitive to the precise electronic state. This aspect has not been studied, as best we know, for other common silicates and more work is required.
Standard third-law entropy, S°, of andradite
Calorimetry
Robie et al. (1987) presented a calorimetrically-based standard entropy value for andradite of 316.4 ± 2.0 J/mol ∙ K. This is lower than our best estimate value of S° ≈ 324 ± 2 J/mol ∙ K that is based on results from both natural and synthetic samples. There are several issues concerning the Robie et al. investigation that require careful analysis. These authors state “Because of the large heat capacity at the lowest temperature at which measurements were made and because of the much reduced precision in the measurements below 10 K, the uncertainty in the extrapolation of C°p,m to 0 K introduces a quite sizeable uncertainty in S°m (298.15 K).” Indeed, it is clear that the manner in which the phase transition is treated (see discussion above and Robie et al., 1987 − Fig. 1) will affect S°. We calculated S° using the Robie et al. (1987),Cp data set and by extrapolating the low temperature flank of the magnetic transition towards 0 K using our Cp results below 8 K. Doing this, we obtain S° = 330.8 J/mol ∙ K (Table 4), which is larger than their reported value. A further issue regards the presence of about 2 vol.% wollastonite in their synthetic sample and how it was treated in their data analysis. It is not fully clear how they calculated Cp(T) and S° in their study. We calculate a “wollastonite-corrected” S° from their Cp data, using our revised value of 330.8 J/mol ∙ K, under the assumption of 2 vol.% wollastonite in the sample (see Robie et al., 1987). Converting vol.% to mol%, we obtain 6.7 mol% wollastonite and 93.3 mol% andradite, and taking a value of S° = 163.4 J/mol ∙ K for wollastonite (Robie & Hemingway, 1995), we obtain a final corrected S° value of ∼330 J/mol ∙ K (Table 4). The S(T) behavior for andradite from 0 K to 298 K is given in Dachs & Geiger (2018). It shows that S(T) is underestimated by the Robie et al. (1987) results at about T < 13 K and slightly overestimated at roughly T >50 K. Finally, in another early calorimetric investigation, Kiseleva et al. (1972) reported S° value of 329.28 ± 5.4 J/mol ∙ K (Table 4).
Vibrational spectroscopy and model heat-capacity behavior
The Cvib and Svib behavior of a crystal can be calculated from knowledge of its phonon density of states (DOS). A complete phonon DOS determination, as would be given by neutron diffraction measurements and lattice dynamic calculations, is not at hand for andradite. To circumvent such laborious study, (1979a,b) proposed a simplified lattice-dynamic model, using Raman and IR spectra, along with the input of other physical data, to obtain the DOS and ultimately Cp,vib and Svib of a crystal. Kieffer (1980) thereby calculated a value for S° of 322.67 J/mol ∙ K for andradite. This is in remarkably good agreement with the value proposed here of S° ≈ 324 J/mol ∙ K. Later, Madon et al. (1991), adopting the Kieffer model, suggested a value of S° ≈ 313.6 J/mol ∙ K apparently in their attempt to bring it in line with Robie et al.'s (1987) value of S° = 316.4 J/mol ∙ K. This was achieved by adopting a reduced theoretical value for Smag of 17.3 J/mol ∙ K, for which there are no physical grounds (see discussion above). Indeed, Madon et al.'s model calculations give S°v = 324.7 J/mol ∙ K and S°v(anhar.) = 326.1 J/mol ∙ K (constant volume and with an anharmonic correction, respectively − see Madon et al., 1991), when the full theoretical Smag value of 29.8 J/mol ∙ K is used.
The good agreement between model Cp and S° behavior, based on simple IR and Raman spectroscopic results and experimental calorimetry, is remarkable. We note, however, that in the case of other transition-metal-bearing garnets, such as almandine and spessartine, there appears to be differences between model calorimetric-based and theoretical Smag values (Dachs et al., 2009, 2012b). The reason for this is not understood.
Thermodynamic behavior from phase equilibrium results and in databases
It is also necessary to consider S° values obtained from phase-equilibrium results and those found in various thermodynamic databases (Table 6). An early calculation of S° (Taylor & Liou, 1978) from phase-equilibrium data is clearly too low in value, as is the value in the early thermodynamic database of Helgeson et al. (1978). Improved thermodynamic properties for many rock-forming silicates are given in internally consistent thermodynamic databases (e.g., Gottschalk, 1997; Chatterjee et al., 1998; Holland & Powell, 2011). All their S° values are similar to the S° value of Robie et al. (1987).
High-temperature Cp behavior above 300 K: comparison of DSC results
The DSC is a standard method to measure Cp from about 300 to 1000 K. There has been much discussion in the literature concerning the exact accuracy and precision of the method over the years. We discuss, here, the various DSC results on andradite, as there are now results from three different laboratories (USGS, Kiel, Salzburg), which is an uncommon situation for minerals (Fig. 8). We note that there is excellent agreement in the Cp data sets obtained in the Kiel and Salzburg DSC laboratories using the same andradite sample SD23 and with more recently built DSC devices. On the other hand, the older Cp data from Robie et al. (1987) appear slightly too high especially between 600 and 800 K. We conclude, based on numerous Cp measurements on various minerals in two different DSC laboratories (Kiel and Salzburg) over the years, that it may be possible to achieve an experimental precision in the best cases of roughly ±0.5%. Precision and accuracy are a function of temperature and they decrease with increasing temperature. This is an improvement over older DSC measurements, as Robie et al. (1987) state that their DSC results have a precision of 1.0%. We think uncertainties for the older measurements can be even higher (see Bosenick et al., 1996).
Acknowledgments
R. Boccio (University of Mailand, Italy, sample VM-11) and C. Ferraris (Muséum National d'Histoire Naturelle, Paris, France, sample 113-102) generously donated crystals for study. S. Speziale (Potsdam) kindly made the photos under crossed polars. This research was supported by grants to C.A.G. from the Austrian Science Fund (FWF: P 25597-N20; FWF and P 30977-NBL). D. Belmonte and two anonymous reviewers made constructive and thorough comments that improved the manuscript.