banner

News

Aug 18, 2023

Mg3Al2Si3O12 jeffbenite inclusion in super

Scientific Reports volume 13, Article number: 83 (2023) Cite this article

849 Accesses

1 Citations

2 Altmetric

Metrics details

An Author Correction to this article was published on 17 February 2023

This article has been updated

Jeffbenite (having the same chemical composition of pyrope, ~ Mg3Al2Si3O12, and also known as TAPP phase) is a mineral inclusion only found in diamonds formed between about 300 and 1000 km depth) and is considered a stable phase in the transition zone (410–660 km depth) and/or in the shallowest regions of the lower mantle (around 660–700 km depth). This rare and enigmatic mineral is considered to be a pressure marker for super-deep diamonds and therefore it has a key role in super-deep diamond research. However, the pressure–temperature stability fields for Mg3Al2Si3O12 jeffbenite is unknown and its actual formation conditions remain unexplored. Here we have determined the thermodynamic pressure–temperature stability field for the jeffbenite Mg-end member and surprisingly discovered that it is stable at low pressure–temperature conditions, i.e., 2–4 GPa at 800 and 500 °C. Thus, Mg3Al2Si3O12 jeffbenite is not the high-pressure polymorph of pyrope and is likely a retrogressed phase formed during the late ascent stages of super-deep diamonds to the surface.

Jeffbenite (ideal formula Mg3Al2Si3O12) is a rare mineral that so far was only found as mineral inclusion in super-deep diamonds1. It was discovered in 19972,3 and since then it has been referred to as TAPP, an acronym from “Tetragonal Almandine Pyrope Phase” for its stoichiometry, which is coincident with that of the pyrope-almandine garnet series. However, the crystal structure of jeffbenite is different from that of garnet, thus garnet and jeffbenite are actually polymorphs. In 2016, TAPP was finally given a mineral name approved by IMA, which is “jeffbenite” (IMA 2014–097)1 to honour Jeffrey W. Harris and Ben Harte, two eminent experts in the field of diamond research. From its first discovery 26 years ago, only 23 natural jeffbenites were reported in literature and 7 of them were identified only by chemical analysis; 2 further jeffbenites reported in the literature are synthetic. So, only 16 natural jeffbenite inclusions in super-deep diamonds have been identified by X-ray diffraction and/or micro-Raman spectroscopy (see Table 1, which summarizes all jeffbenites reported so far in the literature). Despite its rarity, jeffbenite inclusions in diamonds are considered to be an indicator of a super-deep origin for the diamond hosts and therefore it is an important mineral. Its super-deep origin is indeed well accepted in literature and jeffbenite is generally considered a transition zone or lower mantle mineral by the diamond research community1,2,3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20.

However, excluding a very Ti–rich synthetic jeffbenite11, at present no pressure–temperature stability fields of jeffbenite are published and this mineral remains a geological enigma: (1) at what depth in the mantle does jeffbenite actually form? (2) is the Mg-end member of jeffbenite a higher- or lower-pressure polymorph of pyrope garnet?

To answer these questions, here we constrain the pressure–temperature stability field of pure jeffbenite, Mg3Al2Si3O12. As no thermodynamic data for jeffbenite are available in literature, we compute the data from first principles, at the hybrid Hartree–Fock/Density Functional Theory (HF/DFT) level, within the limit of the quasi-harmonic approximation and in the framework of statistical thermodynamics. This allows us to comprehend the actual nature of the pure Mg end-member of jeffbenite.

Indeed, the present work was driven not only by the need to constrain the pressure–temperature stability field for a mineral taken to be characteristic of super-deep diamonds, but also because already in 19972 study of the first natural jeffbenite (at that time indicated as TAPP) revealed inconsistencies in terms of volume and density with respect to garnet. In detail, the natural jeffbenite discovered in 19972 (sample 244B, on which the authors refined the crystal structure) had an approximate composition equal to [Mg2.64Fetot0.27(Ca + Na + Mn)0.08][Al1.85 Cr0.15][Si2.91Al0.09]O12 and a Mg# [Mg/(Mg + Fe)] = 0.91; its unit-cell volume was V = 774.35(± 0.77) Å3. Such unit-cell volume can be converted in a molar volume equal to 11.657(± 0.012) J/bar. Comparing this molar volume with that of a garnet with similar composition along the pyrope-almandine series [Mg2.70Fe0.30]Al2Si3O12 and Mg# = 0.91 (e.g., see Table 3 of ref.21), it is evident that the molar volume of the garnet, which is equal to 11.332(± 0.001) J/bar, is significantly smaller than that of jeffbenite.

This first simple calculation in terms of molar volume shows a strong discrepancy: the molar volume of jeffbenite seems significantly larger than its pyrope polymorph, thus jeffbenite should not be the higher-pressure polymorph of pyrope; this said on the basis of simple thermodynamical considerations.

To confirm this unexpected result with respect to what is believed in the super-diamond research, we need a complete set of thermodynamic parameters.

We have determined a Birch-Murnaghan equation of state truncated to the third order (BM3-EOS22) for jeffbenite, which provides the following values of unit-cell volume, V0, bulk modulus, K0T and first pressure derivative, K´(at T = 298.15 K):

(V0 can be expressed in J/bar providing a value of 11.532, which is already rescaled by a 0.9787 factor to take into account the typical overestimation from the DFT calculation. The correction factor was estimated starting from the same identical overestimation on pyrope). The reason for the overestimation of the cell volume in ab initio calculations (at the DFT or HF/DFT level of the theory) is well understood23,24 and it has long been proved to be not an issue in the estimation of the second derivatives of the energy versus volume function on which, in turn, bulk moduli and vibrational frequencies are computed.

The full elastic constant tensor of jeffbenite have been computed at the static level (i.e. T = 0 K, P = 0 GPa and no zero point effects included) by fitting the second derivatives of the energy with respect to strain components, then using stress–strain relations25. Jeffbenite (tetragonal, space group I\(\overline{4}\)2d) has six independent elastic stiffnesses, calculated as follows: C11 = C22 = 319.2 GPa; C12 = 140.7 GPa; C13 = C23 = 123.5 GPa; C33 = 257.0 GPa; C44 = C55 = 100.5 GPa; C66 = 129.1 GPa. The aggregate elastic moduli (bulk and shear moduli) inferred by the elastic tensor through a Voigt-Reuss-Hill averaging scheme are KVRH = 184.2 GPa and GVRH = 98.4 GPa, respectively. The former value is in excellent agreement with that obtained at T = 0 K and P = 0 GPa from the static BM3-EOS (i.e. K0 = 182.8 GPa), which supports the internal consistency of ab initio elastic data computed for jeffbenite in this work.

The evolution of the bulk modulus as a function of temperature is shown in Fig. 1a. The temperature dependency of the bulk modulus is expressed as:

Dependency of bulk modulus, K0 (in GPa) (a) and that of the volume thermal expansion coefficient, α (in K−1) (b) as a function of temperature, T (in K), for jeffbenite in this study.

The volume thermal expansion coefficient is given by:

The thermal expansion coefficient evolution as a function of temperature is shown in Fig. 1b.

The values of entropy and Gibbs energy of formation (starting from pyrope) are given as:

To show that the thermodynamic properties of jeffbenite calculated in this work are reliable, we have compared the Raman spectrum calculated from our data with an experimental one for the holotype jeffbenite approved by IMA (Fig. 2)1. The spectra are nearly identical, with the calculated one showing a higher resolution (this is quite typical as it is unlikely that an experimental spectrum can reach the resolution of the computed one); the experimental spectrum appears to be slightly shifted toward higher wavenumbers: this could be related to the differences in chemical composition between the pure Mg3Al2Si3O12 jeffbenite used for the calculation in this work and the natural sample, which shows an average of 4.6 wt% of FeO (of which about 20% as Fe3+), some Cr3+ and slightly less Al than the pure end member. However, although such limited chemical differences, the overlap is satisfactory.

Comparison between the calculated Raman spectrum of jeffbenite form this study and the experimental one1.

As the entropy (at least the vibrational contribution to it, which is the only one if order/disorder phenomena are excluded, as it is in the present case) is uniquely determined by the phonon spectrum, the excellent match between the calculated and the experimental Raman spectrum in Fig. 2 provides assurance that the value of entropy we determined for jeffbenite is reliable (see next section).

To analyse the possible phase transition between the two polymorphs jeffbenite and pyrope we need first to compute the Gibbs energy. The differences in Gibbs energy between jeffbenite and pyrope, as functions of pressure, at different (fixed) temperatures is described as follows:

The free energy of pyrope has been evaluated in several different ways:

Full quantum–mechanical evaluation;

Quantum–mechanical evaluation but with a correction for the entropy at standard conditions (\(S_{0}\)) taken from the H&P 2011 database

From the H&P 2011 database

From the H&P 2002 database

From the Stixrude database

The entropy of pyrope at standard conditions is a critical parameter affecting the pressure of transition from jeffbenite to pyrope as the temperature increases. The quantum–mechanical evaluation of S0 is 276.31 J/mol K [the modified-Kieffer model26], has been used for the evaluation of the acoustic mode contribution, with frequencies taken from the original Kieffer publication27; the value of the entropy without such acoustic contribution is significantly lower: 263.82 J/mol K, about 4.5% lower]. The value of S0 we determined in this work by quantum–mechanical evaluation is nearly identical to a previous one28 (276.85 J/mol K), at the ab initio level, through a super-cell approach for the evaluation of both the phonon dispersion effects and the acoustic contribution. Furthermore, the vibrational entropy values calculated in this work for pyrope are consistent with previous computational investigations performed at the hybrid HF/DFT level29. In Fig. 3, the entropy as functions of temperature, at zero pressure and from different works, is shown.

Entropy, S (in J/mol K), as a function of temperature, T (in K), for the pure end member pyrope (the experimental values are from ref.30 up to 350 K and from ref.31 for data at higher temperatures).

The difference between the entropies from Baima et al.28 and the current work are negligible, at least up to a temperature of 600 K. Experimental values are generally lower than the corresponding ab initio values. Indeed, the experimental \(S_{0}\) is 266.27 J/mol K. The values of \(S_{0}\) for pyrope adopted in thermodynamic databases are: 266.30, 269.50 and 242.36 J/mol K for HP0232, HP1133 and Stx34, respectively.

In this work, the Gibbs energy of jeffbenite was evaluated by using the same methods and computational parameters as those employed for pyrope. In Fig. 4, a comparison of the \(\Delta G\left( P \right)\) between the two polymorphs, at 300 (Fig. 4a), 500 (Fig. 4b) and 700 K (Fig. 4c) is shown. In Fig. 4, the straight line \(\Delta G = 0\) (zero line) marks the transition from jeffbenite to pyrope, that occurs at 6.02 GPa at 300 K (Fig. 4a), as seen from the intersection of the solid line with the zero line. In this case, the Gibbs energy of pyrope is evaluated at the ab initio level. Dashed lines, in colour, refer to the evaluation of the Gibbs energy of pyrope by means of the thermodynamics databases HP02, HP11 and Stx. The estimated transition pressures are 6.25 GPa (HP02), 6.11 GPa (HP11) and 6.08 (Stx). At this relatively low temperature (300 K), the impact of entropy on the computed \(\Delta G\) is almost negligible. At higher temperatures, the situation significantly changes: in particular, at a temperature of 500 K (Fig. 4b), the transition pressure decreases to 4.27 GPa (red dashed line).

ΔG (in KJ/mol) for jeffbenite in this study as a function of pressure, P (in GPa), at T = 300 K (a), at T = 500 K (b) and 700 K (c).

However, by recognising the fact that the entropy at standard conditions (S0) of pyrope, computed at the ab initio level, is overestimated with respect to the experimental value, a correction could be applied that results in black solid line of the Fig. 4b (S0 corr., which refers to such a correction of the entropy in the standard state). The correction here adopted corresponds to the HP11 value of S0. In this case, the transition pressure is 4.94 GPa. Indeed, the higher value of the transition pressure in the latter case is due to a relative decrease of the Gibbs energy of pyrope (in turns, due to the decrease of its entropy). The transition pressures computed by employing the Gibbs energies of pyrope from the databases are 5.04 (HP02) and 4.93 GPa (HP11). The curve resulting from the Stixrude database is not reported, as the corresponding value of S0 for pyrope is too far from the experimental one to be considered reliable; in addition, at variance with the entropy reported in the other databases, with the experimental measurements and with the quantum–mechanical estimations, the Stixrude database reports a value for S0 that is lower than the corresponding value for jeffbenite computed in the present work; this leads to an increase of the transition pressure as the temperature is increased (6.45 GPa, at 500 K; this P–T point is not represented in Fig. 4b).

At higher temperatures (see Fig. 4c for T = 700 K) the observed trends are confirmed. The estimation of the transition pressure, if pyrope is dealt at the ab initio level (with the correction for S0 described above), is 3.20 GPa; otherwise, the pressure is 3.53 GPa (pyrope from the HP11 database) or 3.82 GPa (pyrope from the HP02 database). Even with the Stixrude database, at 700 K, the transition pressure decreases (5.92 GPa; again, this P–T point this is not shown in Fig. 4c).

By entering the thermodynamic data of Mg3Al2Si3O12 jeffbenite determined in this work into the Stixrude database34 as implemented in Perple_X software35, the phase diagram section for the Mg3Al2Si3O12 system can be calculated and is shown in Fig. 5. Surprisingly, and in total disagreement with respect to the literature, the stability field indicates that jeffbenite is a low pressure polymorph of pyrope; in detail, by using the Stixrude database34, jeffbenite stability field expands toward higher temperatures and lower pressures. At the maximum temperature, e.g., 1400 K, jeffbenite is stable at about 0.90 GPa, while it reaches a maximum of 6 GPa just above 400 K.

Pressure–temperature stability field of the pure Mg end-members jeffbenite and pyrope calculated using Perple_X35.

The main target of this work was to compute pressure–temperature stability field of jeffbenite and to establish its polymorphic relationship with pyrope. However, thermodynamic and thermoelastic data of jeffbenite were lacking in literature and thus we had to calculate them.

With respect to pyrope, in term of bulk modulus, jeffbenite shows a value of 175.39 GPa which lies within the average value through several data published in literature for pyrope, that spans between about 164 and 182 GPa, with an average value around 170.2 GPa25,36,37,38,39,40,41,42,43,44,45,46,47. The first pressure derivative, K´, is 4.09 for jeffbenite and appears to be slightly lower than the average value of all published values for pyrope, which is 4.63 (ranging between 3.2 and 6.4). The computed bulk modulus and K´ for pyrope (same computational parameters as those employed for jeffbenite) are 162.8 GPa and 4.36, respectively, and are consistent with the ranges of the experimentally measured values.

In terms of bulk modulus dependency with temperature, we obtained for jeffbenite a value equal to -0.020 GPa/K against an average (experimental) value of − 0.022 GPa/K of pyrope [although this value has a quite significant data scatter in literature going from − 0.0194(30)43, to − 0.021(9)48, up to − 0.026(4)47]. However, the computed dK0/dT for pyrope is -0.033 GPa/K; that is, it is slightly higher from that of jeffbenite.

The volume thermal expansion coefficient for jeffbenite is here calculated as α0V = 1.717 × 10–5 K-1 (at 298.15 K) and, differently with respect to what we observed for the bulk modulus, we find a significant difference between jeffbenite and pyrope, with this last showing an average value of 2.19 × 10–5 K-1 (with values ranging between about 2 and 2.5 × 10–5 K-1, ref.44,49,50). The ab initio computed value of α0V for pyrope is 3.0 × 10–5 K-1 (at 298.15 K).

Thanks to the above calculated thermodynamic data, here we have reported, for the first time, the pressure–temperature stability field of jeffbenite, at least for its Mg end-member with composition Mg3Al2Al3O12, which represents the ideal formula of jeffbenite reported in the official mineral list updated to July 2022 by the International Mineralogical Association. The stability field definitively indicates that jeffbenite is not a high pressure mineral and is the lower pressure polymorph of pyrope (see Fig. 5).

Although, based on what is now well accepted in literature, jeffbenite is considered only as a super-deep inclusion in diamonds and thus the higher pressure polymorph of pyrope, our results contradict this statement and, at the same time, they are totally consistent with the analysis of the molar volumes of the two phases. Indeed, as we also mentioned in the first section of the Results, even without any of our calculation and completely neglecting our work, a clear contradiction was already evident at the experimental level by comparing the molar volume of jeffbenite, which has V0 = 11.532 (J/bar), and that of pyrope, which has a V0 = 11.316 (J/bar)51.

By taking into account the thermoelastic parameters of jeffbenite and pyrope as calculated in this work, jeffbenite has a larger molar volume with respect to pyrope in the whole P–T stability range investigated in this work (see Fig. 6). This definitively demonstrates that the former cannot be a higher pressure polymorph of the Mg3Al2Si3O12 end-member phase but instead it is the lower pressure polymorph. This is consistent with our pressure–temperature stability field for jeffbenite in Fig. 5.

Primitive unit cell volume difference (in Å3) between jeffbenite and pyrope, in the indicated pressure–temperature ranges.

As to the other thermodynamic properties, it is interesting to note that some attempts was made in the past to empirically estimate the thermodynamic properties of jeffbenite in such a way to reconcile the phase relations observed in type III inclusions in diamonds from Brazil with an hypothetical stability field of this garnet phase52,53. In particular, a minimum pressure and temperature of 25 GPa and 2273 K were suggested for the formation of jeffbenite on the basis of observations on the Ca content of type III inclusions in diamond where jeffbenite coexists with two silicate perovskites52. Nevertheless, the purely hypothetical HP-HT stability field of jeffbenite in the predicted phase diagram for the enstatite (MgSiO3)—pyrope (Mg3Al2Si3O12) join, besides not being supported by any experimental evidence, is clearly flawed by physical unsoundness of the thermodynamic data assumed for jeffbenite. In fact, the former thermodynamic assessment53 was forced to assume a much higher compressibility of jeffbenite with respect to pyrope as well as huge entropy values for jeffbenite to stabilize this phase at high pressures and temperatures, respectively. This clearly contradicts our first principles results, which define both internally- and physically-consistent thermoelastic parameters and reliable entropy values for jeffbenite and pyrope as well. As an example, the entropy value assessed for jeffbenite at T = 970 K53 is S0 970 = 827.35 J/mol·K, which is overestimated by roughly 10% as compared to our ab initio value (i.e. S0 970 = 754.5 J/mol·K). This discrepancy is well beyond the level of confidence by which DFT is able to predict vibrational entropy of silicate minerals54.

Jeffbenite is a rare silicate only found in super-deep diamonds. Although jeffbenite is a rare mineral, however, it covers a crucial role in super-deep diamond research as it is considered a very high pressure mineral marker2 stable at least at the transition zone depths between 410 and 660 km. However, before our study, no pressure–temperature stability fields for jeffbenite was published (with the exception of one very Ti–rich synthetic jeffbenite11) because no thermodynamic data were available for such rare mineral. Here we have calculated all thermoelastic and thermodynamic data and determined the first pressure–temperature stability field for the Mg jeffbenite end-member. Of the 16 natural analysed jeffbenite known in literature, well 8 jeffbenites show a Mg# between 0.90 and 0.92, 7 jeffbenites have a Mg# between 0.81 and 0.89 and only one jeffbenite shows a Mg# equal to 0.43. Thus, such data indicate that the Mg jeffbenite end-member is the most abundant and critical one to understand the behaviour of jeffbenite and this is why we focused on it.

Very surprisingly, our results definitively show that the Mg end-member of jeffbenite is not a high-pressure polymorph of pyrope and is stable at low pressure and temperature conditions (Fig. 5). However, although several Mg-rich jeffbenites were indicated by authors as super-deep inclusions, using our data on pure Mg3Al2Si3O12 jeffbenite, we cannot speculate about the thermodynamic stability of Fe-richer and Ti-richer jeffbenites (and in some jeffbenites even the Cr content could be significant); actually, for Fe-rich and Ti–rich jeffbenites authors reported synthesis experiments demonstrating that they can be obtained at high pressure (Ti–rich up to 13 GPa11, Fe-dominant jeffbenite at 15 GPa17).

Combining our results and those from experiment laboratories11,17, we would suggest to be cautious in using extremely Mg-rich and extremely Ti-poor jeffbenites to claim super-deep origin for their diamond hosts.

Structures (unit cell parameters and atomic fractional coordinates), static energies and vibrational frequencies at the Γ point of the Brillouin zone of jeffbenite were computed at different values of the primitive unit cell volume, in the [359, 406 Å3] range (10 points in the range). The full elastic tensor of jeffbenite, at the equilibrium static volume was also computed. The calculations were performed at the ab initio level by using the CRYSTAL17 code55. The hybrid Hartree–Fock/Density Functional WC1LYP56,57,58,59 was employed. The localized basis sets chosen for the atoms were of the type 85-11G(1d) for Mg, 85-11G(2d) for Al, 88-31G(2d) for Si, and 8-411G(2d) for O. The thresholds controlling the computation of the Coulombic and exchange integrals (ITOL1 to ITOL5 in the CRYSTAL17 input56) were set to 9, 9, 9, 9 and 22. The shrinking factor (IS) controlling the sampling of points in the BZ where the electronic Hamiltonian is diagonalized was set to 4, resulting in 13 independent k points in the BZ. An XXL grid56 for the numerical evaluation of the integrals of the DFT functionals of the electron density was chosen, which corresponded to 219,069 points in the unit cell; the very high accuracy of such numerical evaluation can be measured by the integration of the electron density over the unit cell, resulting in 399.999958 electrons in the cell, out of 400. Quantum–mechanical results are included in the manuscript related files.

By using the QM-thermodynamic software59, which implements a standard statistical thermodynamics formalism, in the limit of the Quasi-Harmonic Approximation (QHA), vibrational frequencies and static energies at each unit cell volume were employed to compute (i) the Equation of State (third-order Birch-Murnaghan) parameters (V0, K0, and K´) at 298.15 K; (ii) the dependence of K0 by the temperature (dK0/dT); (iii) the specific heat at constant pressure (CP) and its temperature dependence; (iv) the entropy at standard conditions (S0); (v) the Gibbs free energy at standard conditions (G0); (vi) the thermal expansion and its temperature dependence. A correction to CP, S0 and G0, in order to take into account the contribution of the acoustic phonons to those quantities, was made by employing the modified Kieffer-model as described in previous work26. This method allows to define shear and longitudinal seismic velocities along different propagation (and polarization) directions of the single crystal from the ab initio elastic constant tensor by solving the Christoffel determinant. A set of directionally-averaged seismic velocities are then used to calculate the acoustic contributions to thermodynamic properties according to a sine wave dispersion relation assumed for the three acoustic branches in the phonon spectrum27.

Thermodynamics quantities were estimated up to a temperature of 1700 K, and a maximum pressure of about 20 GPa at a temperature of 298.15 K (24 GPa at 1700 K). Results are provided with the deposited files (see the below Data availability statement).

In order to consistently compare the stability of jeffbenite with that of pyrope, in the appropriate P/T region, identical ab initio calculations at the same HF/DFT level, which employed the same parameters as those already set for jeffbenite, where also performed for the pyrope case. The Gibbs free energy of jeffbenite at standard conditions (G0), in the appropriate scale, was computed from the difference of the ab initio energies of jeffbenite and pyrope (energies estimated at the same ab initio level) and rescaled to the G0’s found in the thermodynamic databases used for subsequent computations33,60. The Perple_X program35 was then used to compute pseudosections for fixed global stoichiometries of the system corresponding to several significant mineral assemblages.

All data generated during this study are included in this published article. In detail, we deposited the following 9 files: (1) Frequencies.pdf = vibrational frequencies of jeffbenite as a function of the unit cell volume and the static pressure as a function of the primitive unit cell volume. (2) jeff_pvt_eosfit.dat = contains the volume-pressure–temperature data. (3) jeffbenite.eos = includes all thermoelastic parameters and has a format readable by EoSFIT-inc free software (http://www.rossangel.com/text_eosfit.htm). (4) R1_jeffbenite_WC1LYP_elastcon_P = 0.mb = zero-pressure static elastic constants tensor calculated for jeffbenite at WC1LYP level of theory (Cijkl, in Mbars). (5) R2_jeffbenite_WC1LYP_VRH_P = 0.txt = the aggregate elastic moduli (bulk modulus K, shear modulus G, Young’s modulus E, in Mbars; Poisson’s ratio) and longitudinal and shear seismic velocities (VP and VS, in km/s) of jeffbenite, computed according to the Voigt-Reuss-Hill scheme61. (6) R3_jeffbenite_WC1LYP_elastcon_P = 0_VpG_S1S2.txt = the longitudinal and shear seismic velocities (VP, VS1 and VS2, in km/s) along any direction of propagation (and polarization) in the single-crystal of jeffbenite, obtained by solving the Christoffel determinant26,62. Longitudinal and shear seismic velocities calculation of jeffbenite were calculated using CAREWARE package63. (7) Input_file_calc.pdf = input file including the complete processing of the ab-initio data to produce the thermodynamic parameters, by means of the Qm-Thermodynamic Python program (https://qm-thermodynamics.readthedocs.io/en/main/). 59 .This file allows any readers to reproduce all results obtained in this work. (8) jeffbenite_WC1LYP_elastcon_P = 0.d12 = input file of zero-pressure elastic constant tensor calculation of jeffbenite at WC1LYP level of theory. (9) jeffbenite_WC1LYP_elastcon_P = 0.txt = output file of zero-pressure elastic constant tensor calculation of jeffbenite at WC1LYP level of theory calculated with CRYSTAL.

A Correction to this paper has been published: https://doi.org/10.1038/s41598-023-30111-2

Nestola, F. et al. Tetragonal Almandine-Pyrope Phase, TAPP: Finally a name for it, the new mineral jeffbenite. Min. Mag. 80, 1219–1232 (2016).

Article Google Scholar

Harris, J. W., Hutchison, M. T., Hursthouse, M., Light, M. & Harte, B. A new tetragonal silicate mineral occurring as inclusions in lower mantle diamonds. Nature 387, 486–488 (1997).

Article ADS CAS Google Scholar

Hutchinson, M. T. Constitution of the deep transition zone and lower mantle shown by diamonds and their inclusions. Ph.D. Thesis (University of Edinburgh, 1997).

McCammon, C., Hutchison, M. & Harris, J. Ferric iron content of mineral inclusions in diamonds from São Luiz: A view into the lower mantle. Science 278, 434–436 (1997).

Article ADS CAS Google Scholar

Stachel, T., Harris, J. W., Brey, G. P. & Joswig, W. Kankan diamonds (Guinea) II: Lower mantle inclusion parageneses. Contrib. Min. Petrol. 140, 16–27 (2000).

Article ADS CAS Google Scholar

Hutchison, M. T., Hurtshouse, M. B. & Light, M. E. Mineral inclusions in diamonds: Associations and chemical distinctions around the 670-km discontinuity. Contrib. Min. Petrol. 142, 119–126 (2001).

Article ADS CAS Google Scholar

Harte, B., Harris, J. W., Hutchison, M. T., Watt, G. R. & Wilding, M. C. in Mantle petrology; field observations and high-pressure experimentation: A tribute to Francis R. (Joe) Boyd. Vol. 6, (eds. Y. Fei, C. M. Bertka & B. O. Mysen) 125–153 (Geochemical Society, 1999).

Kaminsky, F. V. et al. Superdeep diamonds from the Juina area, Mato Grosso State, Brazil. Contrib. Min. Petrol. 140, 734–753 (2001).

Article ADS CAS Google Scholar

Hayman, P. C., Kopylova, M. G. & Kaminsky, F. V. Lower mantle diamonds from Rio Soriso (Juina area, Mato Grosso, Brazil). Contrib. Min. Petrol. 149, 430–445 (2005).

Article ADS CAS Google Scholar

Bulanova, G. P. et al. Mineral inclusions in sublithospheric diamonds from Collier 4 kimberlite pipe, Juina, Brazil: Subducted protoliths, carbonated melts and primary kimberlite magmatism. Contrib. Min. Petrol. 160, 489–510 (2010).

Article ADS CAS Google Scholar

Armstrong, L. S. & Walter, M. J. Tetragonal almandine pyrope phase (TAPP): Retrograde Mg-perovskite from subducted oceanic crust?. Eur. J. Min. 24, 587–597 (2022).

Article Google Scholar

Harte, B. & Hudson, N. F. C. in Proc. 10th Int. Kimberlite Conf., Vol. 2, 235–253 (Journal of the Geological Society of India, 2013).

Zedgenizov, D. A., Kagi, H., Shatsky, V. S. & Ragozin, A. L. Local variations of carbon isotope composition in diamonds from Sao Luis (Brazil); evidence for heterogenous carbon reservoir in sublithospheric mantle. Chem. Geol. 363, 114–124 (2014).

Article ADS CAS Google Scholar

Nestola, F. Inclusions in super-deep diamonds: Windows on the very deep Earth. Rend. Lincei-Sci. Fis. 28, 595–604 (2017).

Article Google Scholar

Zedgenizov, D., Hirioyuki, K., Ohtani, E., Tsujimori, T. & Komatsu, K. Retrograde phases of former bridgmanite inclusions in superdeep diamonds. Lithos 370–371, 105659 (2020).

Article Google Scholar

Smith, E. M. & Nestola, F. in Mantle Convection and Surface Expressions, Geophysical Monograph 263 (eds. Marquardt, H., Ballmer, M., Cottaar, S. & Konter, J.) 179–192 (American Geophysical Union, Wiley, 2021).

Smyth, J. R. et al. Ferromagnesian jeffbenite synthesized at 15 GPa and 1200 °C. Am. Min. 107, 405–412 (2021).

Article ADS Google Scholar

Wang, F. et al. High-pressure crystal structure and equation of state of ferromagnesian jeffbenite: Implications for stability in the transition zone and uppermost lower mantle. Contrib. Min. Petrol. 176, 93 (2021).

Article ADS CAS Google Scholar

Thomson, A. R. et al. Origin of sub-lithospheric diamonds from the Juina-5 kimberlite (Brazil): Constraints from carbon isotopes inclusion compositions. Contrib. Min. Petrol. 168, 1081 (2014).

Article ADS Google Scholar

Smith et al. Blue boron-bearing diamonds from Earth’s lower mantle. Nature 560, 84–87.

Geiger, C. A. & Feenstra, A. Molar volumes of mixing of almandine–pyrope and almandine–spessartine garnets and the crystal chemistry of aluminosilicate garnets. Am. Min. 82, 571–581 (1997).

Article ADS CAS Google Scholar

Birch, F. Finite elastic strain of cubic crystals. Phys. Rev. 71, 809–824 (1947).

Article ADS CAS MATH Google Scholar

Cremer, D. Density functional theory: Coverage of dynamic and non-dynamic electron correlation effects. Mol. Phys. 99, 1899–1940 (2001).

Article ADS CAS Google Scholar

Prencipe, M. Quantum mechanics in Earth sciences: A one-century-old story. Rend. Lincei-Sci. Fis. 30, 239–259 (2019).

Article Google Scholar

Erba, A., Mahmoud, A., Belmonte, D. & Dovesi, R. High pressure elastic properties of minerals from ab initio simulations: The case of pyrope, grossular and andradite silicate garnets. J. Chem. Phys. 140, 124703 (2014).

Article ADS CAS PubMed Google Scholar

Belmonte, D., Gatti, C., Ottonello, G., Richet, P. & VetuschiZuccolini, M. Ab initio Thermodynamic and Thermophysical Properties of Sodium Metasilicate, Na2SiO3, and Their Electron-Density and Electron-Pair-Density Counterparts. J. Phys. Chem. A 120, 8881–8895 (2016).

Article CAS PubMed Google Scholar

Kieffer, S. W. Thermodynamics and lattice vibrations of minerals: 4: Application to chain and sheet silicates and orthosilicates. Rev. Geophys. Space Ge. 18, 862–886 (1980).

Article ADS CAS Google Scholar

Baima, J., Ferrabone, M., Orlando, R., Erba, A. & Dovesi, R. Thermodynamics and phonon dispersion of pyrope and grossular silicate garnets from ab initio simulations. Phys. Chem. Min. 43, 137–149 (2016).

Article ADS CAS Google Scholar

De La Pierre, M. & Belmonte, D. Ab initio investigation of majorite and pyrope garnets: Lattice dynamics and vibrational spectra. Am. Min. 101, 162–174 (2016).

Article ADS Google Scholar

Haselton, H. T. & Westrum, E. F. Low-temperature heat capacities of synthetic pyrope, grossular, and pyrope60 grossular40. Geochim. Cosmochim. Ac. 44, 701–709 (1980).

Article ADS CAS Google Scholar

Téqui, C., Robie, R. A., Hemingway, B. S., Neuville, D. R. & Richet, P. Melting and thermodynamic properties of pyrope (Mg3Al2Si3O12). Geochim. Cosmochim. Ac. 55, 1005–1010 (1991).

Article ADS Google Scholar

Holland, T. J. B. & Powell, R. An internally consistent thermodynamic data set for phases of petrological interest. J. Metamorph. Geol. 16, 309–343 (1998).

Article ADS CAS Google Scholar

Holland, T. J. B. & Powell, R. An improved and extended internally consistent thermodynamic dataset for phases of petrological interest, involving a new equation of state for solids. J. Metamorph. Geol. 29, 333–383 (2011).

Article ADS CAS Google Scholar

Stixrude, L. & Lithgow-Bertelloni, C. Thermodynamics of mantle minerals – II, Phase equilibria. Geophys. J. Int. 184, 1180–1213 (2011).

Article ADS CAS Google Scholar

Connolly, J. A. D. Computation of phase equilibria by linear programming: A tool for geodynamic modeling and its application to subduction zone decarbonation. Earth Planet. Sc. Lett. 236, 524–541 (2005).

Article ADS CAS Google Scholar

Babuška, V., Fiala, J., Kumazawa, M., Ohno, I. & Sumino, Y. Elastic properties of garnet solid-solution series. Phys. Earth Planet. In. 16, 157–176 (1978).

Article ADS Google Scholar

Leitner, B. J., Weidner, D. J. & Liebermann, R. C. Elasticity of single crystal pyrope and implications for garnet solid solution series. Phys. Earth Planet. In. 22, 111–121 (1980).

Article ADS CAS Google Scholar

O’Neill, B., Bass, J. D., Rossman, G. R., Geiger, C. A. & Langer, K. Elastic properties of pyrope. Phys. Chem. Min. 17, 617–621 (1991).

ADS Google Scholar

Ottonello, G., Bokreta, M. & Sciuto, P. F. Parametrization of energy and interaction in garnets: End-member properties. Am. Min. 81, 429–447 (1996).

Article ADS CAS Google Scholar

Conrad, P., Zha, C. S., Mao, H.-K. & Hemley, R. J. The high-pressure, single-crystal elasticity of pyrope, grossular, and andradite. Am. Min. 84, 384–388 (1999).

Article ADS Google Scholar

Chen, G., Cooke, J. A., Gwanmesia, G. D. & Liebermann, R. C. Elastic wave velocities of Mg3Al2Si3O12-pyrope garnet to 10 GPa. Am. Min. 84, 384–388 (1999).

Article ADS CAS Google Scholar

Mittal, R., Chaplot, S. L., Choudhury, N. & Loong, C. K. Inelastic neutron scattering and lattice-dynamics studies of almandine Fe3Al2Si3O12. Phys. Rev. B 61, 3983–3988 (2001).

Article ADS Google Scholar

Sinogeikin, S. V. & Bass, J. D. Elasticity of majorite and majorite-pyrope solid solution to high pressure: Implications for the transition zone. Geophys. Res. Lett. 9, 2453–2456 (2002).

Google Scholar

Milani, S. et al. Diamond–garnet geobarometry: The role of garnet compressibility and expansivity. Lithos 227, 140–147 (2015).

Article ADS CAS Google Scholar

Chantel, J. et al. Elastic wave velocities in polycrystalline Mg3Al2Si3O12-pyrope garnet to 24 GPa and 1300 K. Am. Min. 101, 991–997 (2016).

Article ADS Google Scholar

Hu, Y., Wu, Z., Dera, P. K. & Bina, C. R. Thermodynamic and elastic properties of pyrope at high pressure and high temperature by first-principles calculations. J. Geophys. Res. 121, 6462–6476 (2016).

Article ADS Google Scholar

Angel, R. J., Gillio, M., Mazzucchelli, M. L. & Alvaro, M. Garnet EoS: A critical review and synthesis. Contrib. Min. Petrol. 177, 54 (2022).

Article ADS CAS Google Scholar

Zou, Y. et al. Thermal equation of state of Mg3Al2Si3O12 pyrope garnet up to 19 GPa and 1,700 K. Phys. Chem. Min. 39, 589–598 (2012).

Article ADS CAS Google Scholar

Skinner, B. J. Physical properties of end-members of the garnet group. Am. Min. 41, 428–436 (1956).

CAS Google Scholar

Thieblot, L., Roux, J. & Richet, P. High-temperature thermal expansion and decomposition of garnets. Eur. J. Min. 10, 7–15 (1998).

Article CAS Google Scholar

Geiger, C. A. A tale of two garnets: The role of solid solution in the development toward a modern mineralogy. Am. Min. 101, 1735–1749 (2016).

Article ADS Google Scholar

Gasparik, T. & Hutchinson, M. T. Experimental evidence for the origin of two kinds of inclusions in diamonds from the deep mantle. Earth Planet. Sc. Lett. 181, 103–114 (2000).

Article ADS CAS Google Scholar

Gasparik, T. Phase Diagrams for Geoscientists: An Atlas of the Earth’s Interior (Springer, 2014).

Book Google Scholar

Benisek, A. & Dachs, E. The accuracy of standard enthalpies and entropies for phases of petrological interest derived from density-functional calculations. Contrib. Min. Pet. 173, 90 (2018).

Article ADS Google Scholar

Dovesi, R. et al. WIREs Computational Molecular Science 8, e1360 (2018).

Article Google Scholar

Dovesi, R. et al. CRYSTAL17 User’s Manual (University of Torino, 2017).

Google Scholar

Wu, Z. & Cohen, R. E. More accurate generalized gradient approximation for solids. Phys. Rev. B 73, 235116 (2006).

Article ADS Google Scholar

Lee, C., Yang, W. & Parr, R. G. Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density. Phys. Rev. B 37, 785–789 (1988).

Article ADS CAS Google Scholar

Prencipe, M. Mauro-Prencipe/QM-thermodynamics: QmTh-v2.4.1-alpha (Version 2.4.1). Zenodo. https://doi.org/10.5281/zenodo.5061097 (2021).

Stixrude, L. & Lithgow-Bertelloni, C. Thermodynamics of mantle minerals: I. Physical properties. Geophys. J. Int. 162, 610–632 (2005).

Article ADS Google Scholar

Hill, R. W. The elastic behaviour of a crystalline aggregate. Proc. Phys. Soc. London 65A, 349–354 (1952).

Article ADS CAS Google Scholar

Musgrave, M. J. P. Crystal Acoustics. Holden-Day, San Francisco, California, 288 pp. (1970).

Mainprice, D. A Fortran program to calculate seismic anisotropy from the lattice preferred orientation of minerals. Comput. Geosci. 16, 385–393 (1990).

Article ADS Google Scholar

Download references

F.N. thanks the ERC Starting Grant n. 307322. D.B. warmly acknowledges financial support by the Italian MIUR PRIN2017 (Project Number 2017KY5ZX8).

Dipartimento di Geoscienze, Università degli Studi di Padova, Via Gradenigo 6, 35131, Padua, Italy

Fabrizio Nestola

Dipartimento di Scienze della Terra, Università degli Studi di Torino, Via Valperga Caluso 35, 10125, Turin, Italy

Mauro Prencipe

Dipartimento di Scienze della Terra, dell’Ambiente e della Vita, Università degli Studi di Genova, Corso Europa 26, 16132, Genoa, Italy

Donato Belmonte

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

F.N. conceived the study and analysed jeffbenite literature data. M.P and D.B. performed the calculations and analysed jeffbenite and garnet literature data. F.N., M.P. and D.B. wrote and revised the manuscript.

Correspondence to Fabrizio Nestola.

The authors declare no competing interests.

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

The original online version of this Article was revised: The Acknowledgments section in the original version of this Article was incomplete. It now reads: “F.N. thanks the ERC Starting Grant n. 307322. D.B. warmly acknowledges financial support by the Italian MIUR PRIN2017 (Project Number 2017KY5ZX8).”

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.

Reprints and Permissions

Nestola, F., Prencipe, M. & Belmonte, D. Mg3Al2Si3O12 jeffbenite inclusion in super-deep diamonds is thermodynamically stable at very shallow Earth’s depths. Sci Rep 13, 83 (2023). https://doi.org/10.1038/s41598-022-27290-9

Download citation

Received: 19 October 2022

Accepted: 29 December 2022

Published: 03 January 2023

DOI: https://doi.org/10.1038/s41598-022-27290-9

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.

SHARE