Journal Pre-proof High-pressure, high-temperature phase stability of iron-poor dolomite and the structures of dolomite-IIIc and dolomite-V
Jannes Binck, Stella Chariton, Michal Stekiel, Lkhamsuren Bayarjargal, Wolfgang Morgenroth, Victor Milman, Leonid Dubrovinsky, Björn Winkler PII:
S0031-9201(19)30087-1
DOI:
https://doi.org/10.1016/j.pepi.2019.106403
Reference:
PEPI 106403
To appear in: Received date:
9 April 2019
Revised date:
1 November 2019
Accepted date:
18 November 2019
Please cite this article as: J. Binck, S. Chariton, M. Stekiel, et al., High-pressure, hightemperature phase stability of iron-poor dolomite and the structures of dolomite-IIIc and dolomite-V, (2019), https://doi.org/10.1016/j.pepi.2019.106403
This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of record. This version will undergo additional copyediting, typesetting and review before it is published in its final form, but we are providing this version to give early visibility of the article. Please note that, during the production process, errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
© 2019 Published by Elsevier.
Journal Pre-proof
High-pressure, high-temperature phase stability of iron-poor dolomite and the structures of dolomite-IIIc and dolomite-V Jannes Bincka,∗, Stella Charitonb , Michal Stekiela , Lkhamsuren Bayarjargala , Wolfgang Morgenrotha , Victor Milmanc , Leonid Dubrovinskyb , Bj¨ orn Winklera a Institut
ro
of
f¨ ur Geowissenschaften, Goethe-Universit¨ at Frankfurt, Altenh¨ oferallee 1, 60438 Frankfurt am Main, Germany b Bayerisches Geoinstitut, Universit¨ at Bayreuth, Bayreuth 95447, Germany c BIOVIA Dassault Syst` emes, 334 Science Park, Cambridge CB4 0WN, UK
-p
Abstract
re
The stability of Fe-poor dolomite, CaMg0.98 Fe0.02 (CO3 )2 , was studied by Raman spectroscopy and single-crystal X-ray diffraction at high pressures (P < 60 GPa)
lP
and high temperatures (T < 2300 K). Density functional theory calculations were employed to complement the experimental study. Between 40–60 GPa and 1800–2300 K, we observed the formation of a high P, T phase, “dolomite-V”,
na
which, after quenching to ambient temperature, remained stable down to 12 GPa. The dolomite-V phase crystallizes in the space group C2/c with Z = 4 formula
ur
units. The structure of the high pressure polymorph “dolomite-IIIc” was solved, which crystallizes in the space group P ¯1 with Z = 8 formula units. The combined
Jo
experimental and theoretical findings show that at high pressures and low to moderate temperatures Dol-IIIc is formed, while at high pressures and high temperatures Dol-V becomes stable. Assuming that thermodynamic equilibrium is obtained at high-pressure, high-temperature conditions, the current study extends our understanding of the phase stability of Fe-poor dolomite polymorphs at upper and lower mantle conditions. Keywords: dolomite, structure, polymorph, carbonate
∗ Corresponding
author. Email address:
[email protected] (Jannes Binck)
Preprint submitted to Journal of LATEX Templates
November 1, 2019
Journal Pre-proof
1. Introduction Carbonates are well-known host minerals for carbon and thus are important components of the global carbon cycle [1]. The Earth’s crust contains carbon mainly stored in carbonate sediments [1]. Large amounts of oceanic crust, 5
including carbonate deposits, are continuously subducted along convergent plate margins [2]. Estimates suggest that teragram of carbon are subducted into the
of
deep Earth every year [3]. The solubility of carbon in major mantle minerals was shown to be very low and led to the conclusion that carbon may be stored
environments, which are consistent with conditions in subducting slabs, are
-p
10
ro
in carbonates that survive decomposition during subduction [4]. Cold oxidized
believed to prevent carbonate break down [5]. However, carbonates may also
re
react with silicates or metals and decompose into diamond, when encountering reducing conditions in the Earth’s mantle [6, 7].
15
lP
Dolomite CaMg(CO3 )2 is thought to constitute up to 50 % of the Earth’s accessible carbonate reservoirs [8]. Since there is evidence for dolomite inclusions in ultra-deep diamonds [9], high-pressure, high-temperature polymorphs of
na
dolomite could also be important carbon-containing phases in the Earth’s mantle. Some experimental constraints concerning the stability of dolomite and Fe-
20
ur
bearing dolomite at mantle conditions are available [10, 11, 12, 13]. Iron is a major chemical component of the mantle and, depending on its redox state,
Jo
affects the thermodynamic stability of carbonates [14]. Recently it has been shown that at high pressures (>35 GPa) distinct polymorphs of dolomite are formed as a function of Fe-content [12, 15]. In order to understand the fate of subducted carbon it is essential to study both Fe-poor and Fe-rich solid solutions 25
of the largest known carbonate reservoir at conditions like they are assumed to prevail in the mantle. In the following paragraphs, results of several earlier studies are discussed which have investigated solid solutions of Fe-poor and Fe-rich dolomite at P, T conditions of the Earth’s mantle.
2
Journal Pre-proof
Dolomite (Dol-I): Dolomite has been used as a term for solid solutions 30
with an Fe content of up to 40 at % [10, 11, 12, 13]. Hence, Ca(Mg,Fe)(CO3 )2 polymorphs were labeled as “dolomite” and a consecutive index. At ambient conditions, dolomite (Dol-I) crystallizes in the space group R¯3 with Z = 3 formula units [16]. The ordered structure is build up by alternating layers of octahedral CaO6 , triangular planar CO3 and octahedral MgO6 groups along the c-axis. A partial substitution of Mg2+ by Fe2+ is observed in natural dolomites
of
35
[17], which form solid solutions with the isostructural ankerite. Nearly Fe-free
ro
dolomite (Fe < 2 at %) was shown to decompose into aragonite and magnesite at higher pressures and higher temperatures[18, 19, 20]. Only recently, however,
40
-p
pure dolomite was shown to undergo two high pressure phase transitions rather than decomposing [12]. An increasing Fe content may stabilize the dolomite
re
structure at higher pressures [21, 11]. For a nearly Fe-free (Fe < 2 at %) dolomite, Raman spectra have indicated the formation of a dolomite-Ib (Dol-Ib) phase at
lP
around 11 GPa [13]. However, the Dol-Ib phase was not found by a more recent study that has used a sample with a slightly higher Fe content (Fe = 8 at %) [22]. At temperatures > 1100 K, dolomite shows an order-disorder transition
na
45
into a phase isostructural with calcite (R¯3c, Z = 6) between ambient pressure
ur
and ∼3 GPa [23, 24].
Dolomite-II (Dol-II):
Pure and Fe-rich (Fe = 40 at %) Dol-II crystallizes
50
Jo
in space group P ¯ 1 with Z = 2 formula units. Its structure was solved by Merlini et al. [11, 12]. The structure has topological similarities with that of Dol-I, but has a lower symmetry due to the tilting of the triangular CO3 groups. It was further shown, that the structure of Dol-II is equivalent to that of calcite-II [11]. The phase transition of Dol-I into Dol-II is of second order [11], soft phonon driven [25] and was reported for Fe-poor dolomite (< 8 at %) at pressures between 55
14–18 GPa [26, 12, 13, 22] and Fe-rich dolomite (40 at %) at ∼17 GPa [12], respectively. At around 30 GPa, laser annealed Dol-II decomposes into aragonite + magnesite in a temperature range between 1600–1800 K [10].
3
Journal Pre-proof
Dolomite-III, -IIIb and -IIIc (Dol-III, IIIb, IIIc): Dolomite-III was first introduced by Mao et al. [10]. A structure model was elucidated one year 60
later by Merlini et al. [11]. Two different high-pressure stuctures have been proposed for a Fe-rich (Fe = 40 at %) dolomite at ambient temperature: Dol-III (P ¯ 1, Z = 8) at 56 GPa [11], and Dol-IIIb (R3, Z = 3) at 37 GPa [12]. Another high-pressure polymorph (Dol-IIIc) was reported for pure CaMg(CO3 )2
65
of
at about 41 GPa and 300 K [12], but due to a limited number of reflections the authors could not solve its structure. The pressure at which Dol-II transforms
ro
into Dol-III, Dol-IIIb, or Dol-IIIc depends on the Fe content, and ranges from 35 to 41 GPa [11, 12]. Raman spectra on Fe-poor (Fe < 8 at %) dolomite at
-p
high pressures have been reported by Efthimiopoulos et al. [13] and Vennari and
70
or Dol-IIIc.
Dol-IV (P nma, Z = 6) is a high-pressure, high-
lP
Dolomite IV (Dol-IV):
re
Williams [22]. It is unclear whether those spectra correspond to Dol-III, Dol-IIIb
temperature polymorph obtained at 115 GPa and 2500 K of an iron-rich (Fe =
na
40 at %) dolomite. Its structure was determined at a pressure of 115 GPa and ambient temperature by Merlini et al. [12]. The coordination of the carbon atoms 75
was found to be no longer triangular, but tetrahedral. Dol-IV was preserved
ur
upon decompression to at least 70 GPa, before the diamond anvils failed.
Jo
Further potential polymorphs of dolomite:
The report of the Dol-III
phase described by Mao et al. [10] was based on diffraction data of a dolomite powder sample (Fe = 8 at %) in a pressure range between 26–38 GPa and upon 80
heating to 1500 K. Since the diffraction patterns of Dol-III reported by Mao et al. [10] differ from those reported by Merlini et al. [11, 12], speculations arose that another high-pressure, high-temperature phase had been found [13]. Based on density functional theory calculations, equations of state have been reported for CaMg(CO3 )2 up to 80 GPa for hypothetical low-energy structures
85
with space groups P ¯ 1, P 2/c, C2/c (trigonally coordinated carbon) and C2/c (4-fold C) [27]. The structure with C2/c space group and Z = 4 formula units
4
Journal Pre-proof
(trigonally coordinated carbon) turned out to be the most stable phase [27]. Furthermore, the thermodynamic stability of the C2/c phase was shown to increase when significant proportions of iron are incorporated into the structure 90
[15]. Summing up the current state of knowledge, polymorphs of both Fe-free
of
and Fe-bearing dolomite can exist at pressures up to at least 112 GPa [12]. At pressures > 35 GPa, heating experiments on Fe-rich dolomite (40 at %) showed
95
ro
that polymorphs can exist at conditions of the Earth’s deep lower mantle [11, 12]. In contrast much less in known regarding the phase stability of Fe-poor dolomite
-p
at the P, T conditions of the mantle. Hence, the aim of our study was to extend the phase diagram of Fe-poor (≤ 2 at %) dolomite up to pressures of 60 GPa
re
and temperatures of 2300 K. Also, we report the crystal structures of Dol-IIIc
100
here.
na
2. Experimental
lP
[12] and the C2/c phase [27], which will be referred to as “dolomite-V (Dol-V)”
2.1. Characterization of samples
ur
A natural iron-bearing dolomite (CaMg0.98 Fe0.02 (CO3 )2 ) from Algeria has been used for high P, T experiments in the laser-heated diamond anvil cell (LHDAC). The stoichiometry and homogeneity of this sample has been previously
Jo
105
characterized by Heinrich et al. [28]. In addition, the sample was further characterized by single-crystal X-ray diffraction with a Bruker IµS Inside diffractometer at the Bayerisches Geoinstitut (Germany). The diffractometer was operated using AgKα radiation (λ = 0.55941 ˚ A) at 50 kV and 880 µA. Lattice param110
eters of the ordered dolomite at ambient conditions were a = b = 4.796(1) ˚ A, c = 15.974(4) ˚ A and V = 318.23(12) ˚ A3 , which are consistent with the parameters of the established ordered structure of iron-free dolomite [16].
5
Journal Pre-proof
2.2. Preparation of high-pressure, high-temperature experiments High-pressure, high-temperature experiments were carried out using Boehler115
Almax diamond anvil cells (DAC) [29]. Diamonds with culets of either 300 or 350 µm diameter were used for high P, T Raman experiments. For high pressure single-crystal X-ray diffraction studies, diamonds with 300 µm culets and Boehler-Almax cells with opening angles of 70◦ were used. Sample chambers
120
of
of 150–175 µm in diameter were laser drilled in Re gaskets pre-indented to 40–48 µm. Samples investigated by Raman spectroscopy were loaded together
ro
with KCl as thermal insulation. No additional thermal insulator was used for samples investigated by X-ray diffraction. All cells were loaded with a ruby
-p
crystal pressure reference and with Ne as pressure transmitting medium. Before and after each Raman or X-ray diffraction measurement the pressure was determined using ruby reference scales for non- and quasi-hydrostatic con-
re
125
ditions [30, 31]. The accuracy of the pressure determination by Mao et al. [31]
lP
is estimated to fall between 0.05 and 1 GPa for a pressure range up to ∼60 GPa [32]. Pressures were determined before and after the measurement with an
130
na
accuracy of better than ±1 GPa. In some cases, laser heating led to changes in the pressure by up to 8 %.
ur
2.3. Raman spectroscopy in the LH-DAC Raman spectroscopy was carried out at the Institute of Geosciences at the
Jo
Goethe Universit¨at Frankfurt (Germany). Single crystals of 40–90 µm diameter and 20 µm thickness were measured in three different runs. Raman spectra were 135
measured in 2–4 GPa steps upon compression or decompression covering a range between ambient pressure and ∼55 GPa. The measured frequency range was 100– 1300 cm−1 . A frequency doubled 532.14 nm Nd:YAG Oxxius laser (LCX-532S) was focused on the sample and spectra were collected in backscattering geometry, using a grating spectrometer (Acton, SP-2356) equipped with a CCD detector
140
(Pixis 256E) and a microscope objective (Mitutoyo) [33]. The laser power was set to 430 mW and spectra were collected for 100 s. The sample was heated from both sides with a pulsed CO2 laser (Diamond K-250 from Coherent, λ = 6
Journal Pre-proof
10.6 µm) [33]. For the spectroradiometric temperature determination we used a grating spectrometer in combination with a CCD camera (Acton with CCD 145
camera (Pixis 256E)). The temperatures during laser heating were determined by the two-colour pyrometer method, employing Planck and Wien fits [34]. The accuracy of temperature measurement is around ±150 K.
of
2.4. High-pressure X-ray diffraction The synthesis of Dol-V was carried out using the laser set-up [33] described above. A single crystal of dolomite (Dol-I) with dimensions of 25 × 20 × 15 µm3
ro
150
was compressed to 46.2(8) GPa in a DAC and laser-heated to 1800(150) K
-p
for 5 min. The formation of the Dol-V phase was initially identified using Raman spectroscopy. X-ray diffraction experiments were then performed at the
155
re
high-pressure X-ray diffraction beamline ID15b at the European Synchrotron Radiation Facility (ESRF, Grenoble, France), with a wavelength of 0.4113 ˚ A
lP
and a MAR555 flat-panel detector (experimental details are listed in Tab. 1). The spot-size of the X-ray beam was adjusted to 10 × 10 µm2 .
na
After the measurement at the ESRF the pressure of the laser-heated sample was increased to 59.3(8) GPa and again heated to 2300(150) K for 5 min in 160
Frankfurt. Powder diffraction measurements were eventually carried out upon
ur
decompression down to ambient conditions at the extreme conditions beamline P02.2 at PETRA III (Desy, Hamburg, Germany) using a wavelength of 0.2893 ˚ A
Jo
and a Perkin Elmer XRD 1621 flat-panel detector. The spot-size of the X-ray beam was set to 9 × 3 µm2 . 165
Further single crystal diffraction measurements at P02.2 were aimed for determining the structure of Dol-IIIc. Single crystals of dolomite (Dol-I) with dimensions of 10 × 15 × 10 µm3 were compressed and measured in intervals of 5-10 GPa up to a maximal pressure of ∼43 GPa. For Dol-IIIc, diffraction data with two different orientations of the DAC have been obtained in order to increase
170
the completeness of the data set. Integration of the reflection intensities and absorption correction were performed with the CrysAlisPro software (Agilent technology). The data collected 7
Journal Pre-proof
for Dol-IIIc in two different orientations were rescaled and combined using XPREP [35]. Structure solution and refinements of Dol-IIIc and Dol-V were 175
carried out using the JANA software package [36]. The solution of the Dol-IIIc and Dol-V structures was achieved with a charge-flipping algorithm [37], using superflip [38] and Fourier-difference analysis.
of
2.5. Density functional theory In order to obtain theoretical Raman spectra of the phases investigated density functional perturbation theory (DFPT) calculations were performed employing
ro
180
the CASTEP code [39]. The amount of iron in our samples is ≈ 2 at %. It is to
-p
be expected that this will not lead to a measureable change in the Raman spectra at extreme conditions. Hence we chose to compute structure-property relations
185
re
of the iron-free end-member. The code is an implementation of Kohn–Sham DFT based on a plane wave basis set in conjunction with pseudopotentials. The plane
lP
wave basis set makes converged results straightforward to obtain in practice, as the convergence is controlled by a single adjustable parameter, the plane wave
na
cut-off, which we set to 1020 eV. The norm-conserving pseudopotentials were generated “on the fly” from the CASTEP information provided in the CASTEP 190
data base. These pseudopotentials have extensively been tested for accuracy
ur
and transferability [40]. The number of valence electrons were 4, 6, 2, and 10 for C, O, Mg, and Ca, respectively. The corresponding core radii (in a.u.) were 1.2,
Jo
1.2, 1.8, and 2.0 for C, O, Mg, and Ca, respectively. All calculations employed the GGA-PBE exchange-correlation functional [41]. The Brillouin zone integrals 195
were performed using Monkhorst–Pack grids [42] with spacings between grid ˚−1 . Geometry optimizations were defined as being points of less than 0.037 A converged when the energy charge between iterations was <0.5×10−6 eV/atom, the maximal residual force was < 0.01 eV/˚ A, and the maximal residual stress was < 0.02 GPa.
8
Journal Pre-proof
200
3. Results and discussion 3.1. Crystal structure of Dol-IIIc Single crystal diffraction data could be indexed with the Dol-I phase up to a pressure of 11.8(2) GPa, while diffraction data collected between 20.5(4) and 37.0(8) GPa were indexed with the Dol-II phase [11]. Diffraction pat-
205
terns changed again at ∼43 GPa, which is indicative of the Dol-II to Dol-IIIc
of
transition reported by Merlini et al. [12]. The phase transition is of first or-
ro
der as a significant volume decrease of 3.1 % was observed (see appendix Fig. S1). A similar volume decrease was also shown for the transitions of iron-rich
210
-p
Dol-II into the iron-rich Dol-III and Dol-IIIb phases [11, 12]. We could index a Dol-IIIc data set measured at 43.4(8) GPa with the triclinic unit cell
re
reported by Merlini et al. [12] and solved its structure (Tab. 2 and appendix Tab. S1). Dol-IIIc crystallizes in space group P ¯1 with Z = 8 formula units,
lP
corresponding to 80 atoms per unit cell. At 43.4(8) GPa the lattice parameters are a = 4.4518(14) ˚ A, b = 11.1683(17) ˚ A, c = 13.6960(17) ˚ A, α = 69.044(13)◦ , β = 88.343(9)◦ , γ = 89.344(10)◦ and V = 635.64(15) ˚ A3 . There is an excellent
na
215
agreement between the experimental and theoretical structural data for the same pressure (Tab. 2). Dol-IIIc is a distorted polymorph of Dol-II, which is
ur
characterized by planar triangular CO3 groups that are no longer co-planar, alternating distorted MgO6 octahedra, and distorted polyhedra of CaOn with an oxygen coordination of Ca ranging from 7 to 9 (Fig. 1). Some of the Ca and
Jo
220
Mg cation polyhedra share faces and edges. A significant difference between the Fe-poor Dol-IIIc and its Fe-rich analogues (Dol-III and Dol-IIIb) is the lower degree of tilting of the CO3 groups. It was previously shown, that upon further compression (up to 102 GPa) Dol-IIIc undergoes no phase transition at ambient 225
temperature [12]. 3.2. Crystal structure of Dol-V The Dol-V phase was observed by Raman spectroscopy at 46.2(8) GPa and after heating the crystal to 1800(150) K. The structure of Dol-V was studied
9
Journal Pre-proof
using synchrotron X-ray diffraction. The reflections could be indexed with a 230
monoclinic unit cell and the structure was readily solved and refined (Tab. 2 and appendix Tab. S2). Dol-V crystallizes in space group C2/c with Z = 4 formula units, corresponding to 40 atoms per unit cell. Unit cell parameters are a = 8.051(14) ˚ A, b = 7.848(3) ˚ A, c = 5.105(15) ˚ A, β = 106.4(3)◦ and V = 309.4(11) ˚ A3 at 46.2(8) GPa. There is very good agreement between the experimental and theoretical structural data (Tab. 2). The structure is ordered
of
235
and consists of CO3 triangular planar units that are tilted against each other,
ro
distorted MgO6 octahedra and distorted CaO8 square antiprisms. The building blocks of the structure are arranged in alternating layers along the a-axis (Fig.
240
-p
2). Cation polyhedra and some of the planar CO3 groups share edges. At around 46 GPa, a volume decrease of about 2.6 % is observed after
re
transforming Dol-IIIc upon heating to Dol-V (see appendix Fig. S1). The experimentally determined structure of Dol-V is identical to a DFT-based struc-
lP
ture model of a pure dolomite reported by Solomatova and Asimow [27]. A comparison between diffraction data of Dol-V (at 46 GPa) and a monoclinic Dol-III phase reported by Mao et al. [10] at 49 GPa showed that these two
na
245
phases are distinct.
ur
3.3. Decomposition of Dol-V determined by X-ray diffraction Our data displayed some additional reflections at 46.2(8) GPa, which could
250
Jo
not be assigned to Dol-V, diamond, neon or the rhenium gasket. We conclude that new phases were formed due to the decomposition of Dol-V. When Dol-V was pressurized to ∼46 GPa and heated to ∼1800 K, powder rings started to appear in addition to the single crystal reflections. At ∼59 GPa and after heating to 2300 K the powder rings became dominant. Powder diffraction data were collected from 59 GPa down to ambient conditions in 10 GPa steps. Powder 255
patterns at 59 and 46.2 GPa could be indexed with Dol-V and its decomposition products post-aragonite (CaCO3 ) and ferromagnesite ((Mg,Fe)CO3 ), as reported by Ono et al. [43] and Lavina et al. [44], respectively (Fig. 3a & b). In order to calculate reflection positions of post-aragonite and ferromagnesite at the 10
Journal Pre-proof
pressures at which our experiments have been carried out, we used the equation 260
of state data from Ono et al. [43], Lavina et al. [44]. At ambient conditions, the powder patterns could be indexed with the Dol-I phase using the structure reported by Althoff [45]. Additional weak reflections were observed, which could be indexed with calcite (CaCO3 ) and ferromagnesite reported by Sass et al. [46] and Lavina et al. [44] (Fig. 3c). 3.4. Dolomite high-pressure polymorphs determined by Raman spectroscopy
of
265
ro
Raman spectra were measured up to pressures of ∼42 GPa (Fig. 4a) and at ambient temperatures. According to group theory, the following Raman and
-p
infrared modes are expected for Dol-I at ambient conditions: Γ = 4Ag (R) + 4Eg (R) + 5 Au (IR) + 5Eu (IR). Six Raman-modes of Dol-I could be traced when increasing the pressure up to ∼6 GPa (Fig. 5a & b). At pressures between 9.1(1)
re
270
and 14.0(1) GPa, the splitting of the characteristic 176 cm −1 (Eg ) mode of Dol-I
lP
was observed, which indicated that the Dol-Ib phase has formed. Although Dol-Ib was found at a pressure 1.9 GPa lower than reported earlier by Efthimiopoulos
275
na
et al. [13], our observation still shows an excellent reproducibility of the formation of Dol-Ib for the same sample. The second order phase transition of Dol-Ib into Dol-II was found at around 14.5 GPa (Fig. 5a & b). Within an uncertainty of
ur
2 GPa, our result agrees with that of Efthimiopoulos et al. [13]. The transition from Dol-Ib to Dol-II is characterized by the splitting of Raman modes below
280
Jo
600 cm−1 , while Raman bands of 720 and 880 cm−1 start to split into doublets at pressures between 18 and 20 GPa. A splitting of the 1140 cm−1 CO3 stretching mode remains up to 28 GPa, but disappears at higher pressures. For the Dol-II structure reported by Merlini et al. [12], the total number of IR and R phonon frequencies is expected to be Γ = 27Au (IR) + 30Ag (R). We observed at least 18 Raman active modes. Our theoretical Raman spectra are in good agreement 285
with experimental data of Dol-II (Fig. 7). The Dol-II to Dol-IIIc first order phase transition was found at around ∼36.2 GPa (Fig. 5a & b). The Dol-IIIc phase is characterized by a further splitting of the Raman-modes of Dol-II and the occurrence of at least 7 new modes that occur in a frequency interval between 11
Journal Pre-proof
700 and 900 cm−1 and between 1150 and 1250 cm−1 . The Dol-IIIc phase was 290
measured up to a pressure of ∼42 GPa. The factor group analysis of Dol-IIIc gives: Γ = 117Au (IR) + 120Ag (R). We observed at least 29 of the Raman-active modes. The spectra recorded here are similar to previously reported Raman spectra of CaMg0.92 Fe0.08 (CO3 )2 that have been measured in a pressure range between 42–86 GPa [22]. The spectra were discussed to correspond to the DolIII phase, which has been found by Merlini et al. [11]. However, the Raman
of
295
spectra have only been measured for iron-poor dolomite [13, 22], making it more
ro
likely that the formation of Dol-IIIc was observed rather than the presence of Dol-III, or Dol-IIIb. Upon decompression, the back-transformations of Dol-IIIc
respectively.
re
300
-p
into Dol-II and Dol-II into Dol-I were observed at around 30 GPa and 12 GPa,
3.5. High-pressure high-temperature polymorphs of dolomite determined by Ra-
lP
man spectroscopy
Raman spectra of quenched samples that had been heated to temperatures of
305
na
1200-2300 K were measured up to pressures of ∼55 GPa (Fig. 4b and appendix Fig. S2). Close to the Dol-I–Dol-II transition, Dol-I remained stable at 13(1) GPa and upon heating to 1600(150) K (appendix Fig. S2a). No indication for Dol-Ib
ur
was observed before and after heating in this run. At 17 GPa and 1700–1800 K, characteristic Raman-spectra of Dol-II started to change, most notably by the
310
Jo
appearance of a CO3 stretching mode at 1136 cm−1 (appendix Fig. S2b). The new peak at 1136 cm−1 is very likely the CO3 -stretching mode of aragonite as its Raman shift corresponds to that at ∼18 GPa reported by Bayarjargal et al. [33]. A decomposition of Dol-II into aragonite + magnesite has previously been found for pressures between 26–32 GPa and temperatures between 1600–2350 K [10, 11]. Hence, we infer that decomposition of Dol-II started at these conditions. 315
At ∼39.5(8) GPa, heating Dol-IIIc to 1880(150) K led to significant changes in the Raman-spectra (Fig. 4b). The new modes could not be assigned to Dol-IIIc and indicated the synthesis of a new phase assemblage, that we have identified to be Dol-V + ferromagnesite ((Mg,Fe)CO3 ) + post-aragonite (CaCO3 ) by X-ray 12
Journal Pre-proof
diffraction (Fig. 3). The comparison of experimental and theoretical Raman 320
spectra made it possible to distinguish between characteristic modes of Dol-V and its decomposition products (Fig. 7). In comparison to the theoretical data, the experimental data show additional peaks between 800 and 900 cm−1 and a splitting of the CO3 stretching mode at around 1200 cm−1 . Characteristic features that facilitate the detection of Dol-V are new modes in the low frequency range between 146 and 600 cm−1 , as well as two additional intense bands around
of
325
900 cm−1 . Factor group analyses yields the following Raman- and infrared-active
ro
modes for Dol-V: Γ = 14Ag (R) + 13Au (IR) + 16 Bg (R) + 14Bu (IR). We observe at least 14 of the Raman-active modes (Fig. 5). The Dol-V phase was further
330
-p
pressurized to 55.2(8) GPa and again heated to 2300(150) K. A strong splitting of the CO3 stretching mode is observed, which is indicative of the progressive
re
decomposition of Dol-V.
On pressure release, Dol-V was detectable down to 12.6(3) GPa, while char-
lP
acteristic modes shifted to lower frequencies (Fig. 5a & b). The appearance of new modes that might indicate phase transitions of high pressure CaCO3 polymorphs were also observed during pressure release. The splitting of the CO3
na
335
stretching mode between 1150 and 1250 cm−1 decreases continuously during pressure release. Below 25 GPa, only one broad peak can be observed. Below 44
ur
GPa, characteristic modes at around 850 cm−1 disappear completely. Although
340
Jo
it was not possible to properly assign all unknown Raman bands from the high pressure Raman measurements to a particular high pressure carbonate phases, measurements at ambient conditions indicated the reformation of Dol-I, with additional Raman modes that can be explained by a calcite-magnesite-siderite phase assemblage (Fig. 6). Our results from Raman spectroscopy support those from X-ray diffraction and confirm that decomposition occurred at pressures 345
between 39.5-60 GPa and temperatures between 1800-2300 K.
13
Journal Pre-proof
3.6. Implications of the observation of high-pressure, high-temperature polymorphs of dolomite Summing up the combined observations from Raman spectroscopy and Xray diffraction, we were able to improve the current knowledge regarding the 350
phase stability of Fe-poor dolomite (CaMg0.98 Fe0.02 (CO3 )2 ) at conditions of the Earth’s upper and lower mantle. Based on previously reported data, which have
of
investigated dolomite samples of similar compositions [19, 24, 10, 13, 12], and assuming that thermodynamic equilibrium was reached in our experiments, we
355
ro
suggest a phase diagram for Fe-poor dolomite (Fig. 8).
No decomposition of Dol-I or Dol-Ib was found in our experiments, since the
-p
temperatures were too low (Fig. 8). Our data on heated Dol-II complemented earlier findings indicating that Dol-II decomposes into aragonite and magnesite
re
between 17 and 30 GPa at temperatures > 1600 K. Dol-IIIc forms when Dol-II is pressurized to ∼39 GPa at ambient temperature. Upon heating, Dol-IIIc transforms into Dol-V, which after temperature quenching remains present on
lP
360
pressure release down to ∼13 GPa. It is likely that Dol-IIIc is meta-stable at
na
all P, T conditions, since Dol-V formed instead of Dol-IIIc after heating and quenching. Theoretical considerations on Fe-free dolomite [27] indicated that Dol-V is destabilized at higher pressures and 0 K with respect to post-aragonite + magnesite. Our observations indicate that post-aragonite and Fe-bearing
ur
365
Jo
magnesite (ferromagnesite) would form at P, T conditions of the Earth’s lower mantle, whereas Dol-V starts to decompose. However, Dol-V was present together with post-aragonite and ferromagnesite at 40-52 GPa and 1800 K and at 55-60 GPa and 2300 K. This observation either implies that these conditions 370
are close to the univariant reaction boundary of Dol-V * ) post-aragonite + ferromagnesite, or that the reaction kinetics are very slow, which are considered unlikely. Support for the stability of Fe-rich Dol-V at high pressures and 0 K is given by the thermodynamic calculations of Solomatova and Asimow [15], which show that small amounts of Fe can stabilize the C2/c structure of Dol-V as well
375
as other double carbonates at lower mantle conditions. Then, the stability field of Dol-V could intersect the Earth’s lower mantle geotherm. Carbonates are 14
Journal Pre-proof
likely decomposed in high temperature slabs at deep upper mantle depths [47]. However, cold oceanic crust that incorporates carbonate-silicate assemblages is believed to descent down to possibly the core-mantle boundary [48]. Hence, 380
it is not unlikely that substantial quantities of Fe-poor Dol-V assembled with silicates are hosted by deep subducted oceanic slabs, as temperatures in very
of
cold slabs are assumed to be lower than 1500 K at depths < 1500 km [49].
4. Conclusions
knowledge on the phase stabilities of Fe-poor high-pressure, high-temperature
-p
385
ro
Our study on a natural CaMg0.98 Fe0.02 (CO3 )2 dolomite sample extended our
dolomite polymorphs. Structures have been determined for Dol-IIIc at 43 GPa,
re
and for Dol-V after heating to 1800 K at 46 GPa. Experimental structural data were shown to be in excellent agreement with those obtained by DFT-based
390
lP
calculations. Dol-IIIc transforms into Dol-V at pressures between 40 and 60 GPa and temperatures exceeding 1800 K. Dol-V decomposes into post-aragonite (CaCO3 ) and ferromagnesite ((Mg,Fe)CO3 ) at P, T conditions of the Earth’s
na
lower mantle. The results of the present study emphasize the dependence of phase stabilities on composition. Hence, further combined Raman spectroscopy and
395
ur
X-ray diffraction studies along the dolomite-ankerite join at extreme conditions would be of great interest with respect to the phase stability of chemically
Jo
complex carbonates in the Earth’s mantle.
5. Acknowledgements The authors acknowledge funding by the Deutsche Forschungsgemeinschaft (DFG)-Germany (FOR2125/CarboPaT, BA4020, 400
WI1232) and BMBF
(05K16RFA, 05K16RFB). DESY (Hamburg, Germany), a member of the Helmholtz Association (HGF), and the ESRF (Grenoble, France) are acknowledged for the provision of experimental facilities. We would like to thank Hanns-Peter Liermann and his team for assistance in using beamline P02.2. Dr. Ilias
15
Journal Pre-proof
Efthimiopoulos is thanked for supplying sample materials within the CarboPaT 405
collaboration/DFG Research Group.
References [1] R. M. Hazen, R. T. Downs, A. P. Jones, L. Kah, Carbon mineralogy and crystal chemistry, Reviews in Mineralogy and Geochemistry 75 (2013) 7–46.
[2] N. H. Sleep, K. Zahnle, Carbon dioxide cycling and implications for climate
ro
410
of
doi:10.2138/rmg.2013.75.2.
on ancient Earth, Journal of Geophysical Research: Planets 106 (2001)
-p
1373–1399. doi:10.1029/2000JE001247.
[3] R. Dasgupta, M. M. Hirschmann, The deep carbon cycle and melting in
doi:10.1016/j.epsl.2010.06.039.
lP
415
re
Earth’s interior, Earth and Planetary Science Letters 298 (2010) 1–13.
[4] H. Keppler, M. Wiedenbeck, S. S. Shcheka, Carbon solubility in olivine and
na
the mode of carbon storage in the Earth’s mantle, Nature 424 (2003) 414. doi:10.1038/nature01828.
420
ur
[5] N. Martirosyan, T. Yoshino, A. Shatskiy, A. Chanyshev, K. Litasov, The CaCO3 –Fe interaction: Kinetic approach for carbonate subduction to the
Jo
deep Earth’s mantle, Physics of the Earth and Planetary Interiors 259 (2016) 1–9. doi:10.1016/j.pepi.2016.08.008. [6] V. Stagno, D. Frost, C. McCammon, H. Mohseni, Y. Fei, The oxygen fugacity at which graphite or diamond forms from carbonate-bearing melts 425
in eclogitic rocks, Contributions to Mineralogy and Petrology 169 (2015) 16. doi:10.1038/nature11679. [7] Y. N. Palyanov, Y. V. Bataleva, A. G. Sokol, Y. M. Borzdov, I. N. Kupriyanov, V. N. Reutsky, N. V. Sobolev, Mantle–slab interaction and redox mechanism of diamond formation, Proceedings of the National Academy
430
of Sciences 110 (2013) 20408–20413. doi:10.1073/pnas.1313340110. 16
Journal Pre-proof
[8] D. H. Zenger, J. B. Dunham, R. L. Ethington, Concepts and Models of Dolomitization, SEPM Society for Sedimentary Geology, 1980. doi:10.2110/pec.80.28. [9] F. E. Brenker, C. Vollmer, L. Vincze, B. Vekemans, A. Szymanski, 435
K. Janssens, I. Szaloki, L. Nasdala, W. Joswig, F. Kaminsky, Carbonates from the lower part of transition zone or even the lower mantle, Earth
of
and Planetary Science Letters 260 (2007) 1–9. doi:10.1016/j.epsl.2007.02.038.
ro
[10] Z. Mao, M. Armentrout, E. Rainey, C. E. Manning, P. Dera, V. B. Prakapenka, A. Kavner, Dolomite III: A new candidate lower mantle carbonate, Geophysical Research Letters 38 (2011). doi:10.1029/2011GL049519.
-p
440
re
[11] M. Merlini, W. A. Crichton, M. Hanfland, M. Gemmi, H. M¨ uller, I. Kupenko, L. Dubrovinsky, Structures of dolomite at ultrahigh pressure and their
lP
influence on the deep carbon cycle, Proceedings of the National Academy of Sciences 109 (2012) 13509–13514. doi:10.1073/pnas.1201336109. [12] M. Merlini, V. Cerantola, G. D. Gatta, M. Gemmi, M. Hanfland, I. Kupenko,
na
445
P. Lotti, H. M¨ uller, L. Zhang, Dolomite-IV: Candidate structure for a carbonate in the Earth’s lower mantle, American Mineralogist 102 (2017)
ur
1763–1766. doi:10.2138/am-2017-6161.
450
Jo
uller, Com[13] I. Efthimiopoulos, S. Jahn, A. Kuras, U. Schade, M. Koch-M¨ bined high-pressure and high-temperature vibrational studies of dolomite: Phase diagram and evidence of a new distorted modification, Physics and Chemistry of Minerals 44 (2017) 465–476. doi:10.1007/s00269-017-0874-5. [14] V. Cerantola, E. Bykova, I. Kupenko, M. Merlini, L. Ismailova, C. McCammon, M. Bykov, A. I. Chumakov, S. Petitgirard, I. Kantor, V. Svitlyk, J. Ja455
cobs, M. Hanfland, M. Mezouar, C. Prescher, R. R¨ uffer, V. B. Prakapenka, L. Dubrovinsky, Stability of iron-bearing carbonates in the deep Earth’s interior, Nature Communications 8 (2017) 15960. doi:10.1038/ncomms15960.
17
Journal Pre-proof
[15] N. V. Solomatova, P. D. Asimow, First-principles calculations of highpressure iron-bearing monoclinic dolomite and single-cation carbonates with 460
internally consistent Hubbard U, Physics and Chemistry of Minerals 45 (2018) 293–302. doi:10.1007/s00269-017-0918-x. [16] F. J. Sans, H. Steinfink, Refinement of the crystal structure of dolomite, American Mineralogist 44 (1959) 679–682.
solid-solution series; an X-ray, Moessbauer, and TEM study, American
ro
465
of
[17] R. J. Reeder, W. A. Dollase, Structural variation in the dolomite-ankerite
Mineralogist 74 (1989) 1159–1167.
-p
[18] C. Biellmann, P. Gillet, J. Peyronneau, B. Reynard, et al., Experimental evidence for carbonate stability in the Earth’s lower mantle, Earth and Plan-
[19] I. Martinez, J. Zhang, R. J. Reeder, In situ X-ray diffraction of aragonite
lP
470
re
etary Science Letters 118 (1993) 31–41. doi:10.1016/0012-821X(93)90157-5.
and dolomite at high pressure and high temperature: Evidence for dolomite breakdown to aragonite and magnesite, American Mineralogist 81 (1996)
na
611–624. doi:10.2138/am-1996-5-608. [20] A. Buob, R. W. Luth, M. W. Schmidt, P. Ulmer, Experiments on CaCO3 MgCO3 solid solutions at high pressure and temperature, American Miner-
ur
475
Jo
alogist 91 (2006) 435–440. doi:10.2138/am.2006.1910. [21] E. Franzolin, M. Schmidt, S. Poli, Ternary Ca–Fe–Mg carbonates: subsolidus phase relations at 3.5 GPa and a thermodynamic solid solution model including order/disorder, Contributions to Mineralogy and Petrology 480
161 (2011) 213–227. doi:10.1007/s00410-010-0527-x. [22] C. E. Vennari, Q. Williams, A novel carbon bonding environment in deep mantle high-pressure dolomite, American Mineralogist 103 (2018) 171–174. doi:10.2138/am-2018-6270. [23] R. Reeder, H.-R. Wenk, Structure refinements of some thermally disordered
485
dolomites, American Mineralogist 68 (1983) 769–776. 18
Journal Pre-proof
[24] S. M. Antao, W. H. Mulder, I. Hassan, W. A. Crichton, J. B. Parise, Cation disorder in dolomite, CaMg(CO3 )2 , and its influence on the aragonite + magnesite = dolomite reaction boundary, American Mineralogist 89 (2004) 1142–1147. doi:10.2138/am-2004-0728. 490
[25] M. Stekiel, A. Girard, T. Nguyen-Thanh, A. Bosak, V. Milman, B. Winkler, Phonon-driven phase transitions in calcite, dolomite, and magnesite,
of
Physical Review B 99 (2019) 054101. doi:10.1103/PhysRevB.99.054101.
ro
[26] A. Zucchini, P. Comodi, S. Nazzareni, M. Hanfland, The effect of cation ordering and temperature on the high-pressure behaviour of dolomite, Physics and Chemistry of Minerals 41 (2014) 783–793. doi:10.1007/s0026.
-p
495
re
[27] N. V. Solomatova, P. D. Asimow, Ab initio study of the structure and stability of CaMg(CO3 )2 at high pressure, American Mineralogist 102 (2017)
lP
210–215. doi:10.2138/am-2017-5830.
[28] W. Heinrich, P. Metz, W. Bayh, Experimental investigation of the mechanism of the reaction: 1 tremolite + 11 dolomite = 8 forsterite + 13 calcite
na
500
+ 9 CO2 + 1 H2 O, Contributions to Mineralogy and Petrology 93 (1986)
ur
215–221. doi:10.1007/BF00371323. [29] R. Boehler, New diamond cell for single-crystal X-ray diffraction, Review
505
Jo
of Scientific Instruments 77 (2006) 115103. doi:10.1063/1.2372734. [30] H. Mao, P. Bell, J. t. Shaner, D. Steinberg, Specific volume measurements of Cu, Mo, Pd, and Ag and calibration of the ruby R1 fluorescence pressure gauge from 0.06 to 1 Mbar, Journal of applied physics 49 (1978) 3276–3283. doi:10.1063/1.325277. [31] H. Mao, J.-A. Xu, P. Bell, Calibration of the ruby pressure gauge to 800 510
kbar under quasi-hydrostatic conditions, Journal of Geophysical Research: Solid Earth 91 (1986) 4673–4676. doi:10.1029/JB091iB05p04673.
19
Journal Pre-proof
[32] A. Dewaele, P. Loubeyre, M. Mezouar, six metals above 94 GPa,
Equations of state of
Physical Review B 70 (2004) 094112.
doi:10.1103/PhysRevB.70.094112. 515
[33] L. Bayarjargal, C.-J. Fruhner, N. Schrodt, B. Winkler, CaCO3 phase diagram studied with raman spectroscopy at pressures up to 50 GPa and high temperatures and DFT modeling, Physics of the Earth and Planetary
of
Interiors 281 (2018) 31–45. doi:10.1016/j.pepi.2018.05.002.
520
ro
[34] L. R. Benedetti, P. Loubeyre, Temperature gradients, wavelength-dependent emissivity, and accuracy of high and very-high temperatures measured in
-p
the laser-heated diamond cell, High Pressure Research 24 (2004) 423–445.
re
doi:10.1080/08957950412331331718.
[35] G. M. Sheldrick, A short history of SHELX, Acta Crystallographica Section
525
lP
A 64 (2008) 112–122. doi:10.1107/S0108767307043930. [36] V. Petˇr´ıˇcek, M. Duˇsek, L. Palatinus, Crystallographic computing system
na
JANA2006: General features, Zeitschrift f¨ ur Kristallographie-Crystalline Materials 229 (2014) 345–352. doi:10.1515/zkri-2014-1737.
ur
[37] G. Oszl´ anyi, A. S¨ ut˝o, Ab initio structure solution by charge flipping, Acta Crystallographica Section A: Foundations of Crystallography 60 (2004) 134–141. doi:10.1107/S0108767303027569.
Jo
530
[38] L. Palatinus, G. Chapuis,
SUPERFLIP–a computer program for
the solution of crystal structures by charge flipping in arbitrary dimensions,
Journal of Applied Crystallography 40 (2007) 786–790.
doi:10.1107/S0021889807029238. 535
[39] S. J. Clark, M. D. Segall, C. J. Pickard, P. J. Hasnip, M. I. Probert, K. Refson, M. C. Payne,
First principles methods using CASTEP,
Zeitschrift f¨ ur Kristallographie-Crystalline Materials 220 (2005) 567–570. doi:10.1524/zkri.220.5.567.65075.
20
Journal Pre-proof
[40] K. Lejaeghere, G. Bihlmayer, T. Bjoerkman, P. Blaha, S. Bluegel, V. Blum, 540
D. Caliste, I. E. Castelli, S. J. Clark, A. Dal Corso, S. de Gironcoli, T. Deutsch, J. K. Dewhurst, I. Di Marco, C. Draxl, M. Dulak, O. Eriksson, J. A. Flores-Livas, K. F. Garrity, L. Genovese, P. Giannozzi, M. Giantomassi, S. Goedecker, X. Gonze, O. Granaes, E. K. U. Gross, A. Gulans, F. Gygi, D. R. Hamann, P. J. Hasnip, N. A. W. Holzwarth, D. Iusan, D. B. Jochym, F. Jollet, D. Jones, G. Kresse, K. Koepernik, E. Kuecuekbenli,
of
545
Y. O. Kvashnin, I. L. M. Locht, S. Lubeck, M. Marsman, N. Marzari,
ro
U. Nitzsche, L. Nordstrom, T. Ozaki, L. Paulatto, C. J. Pickard, W. Poelmans, M. I. J. Probert, K. Refson, M. Richter, G.-M. Rignanese, S. Saha,
550
-p
M. Scheffler, M. Schlipf, K. Schwarz, S. Sharma, F. Tavazza, P. Thunstroem, A. Tkatchenko, M. Torrent, D. Vanderbilt, M. J. van Setten, V. Van Spey-
re
broeck, J. M. Wills, J. R. Yates, G.-X. Zhang, S. Cottenier, Reproducibility
lP
in density functional theory calculations of solids, Science 351 (2016). [41] J. P. Perdew, K. Burke, M. Ernzerhof, proximation made simple,
Physical review letters 77 (1996) 3865.
doi:10.1103/PhysRevLett.77.3865.
na
555
Generalized gradient ap-
[42] H. J. Monkhorst, J. D. Pack, Special points for Brillouin-zone integrations,
ur
Physical review B 13 (1976) 5188. doi:10.1103/PhysRevB.13.5188.
Jo
[43] S. Ono, T. Kikegawa, Y. Ohishi, J. Tsuchiya, Post-aragonite phase transformation in CaCO3 at 40 GPa, American Mineralogist 90 (2005) 667–671. 560
doi:doi.org/10.2138/am.2005.1610. [44] B. Lavina, P. Dera, R. T. Downs, O. Tschauner, W. Yang, O. Shebanova, G. Shen, Effect of dilution on the spin pairing transition in rhombohedral carbonates, High Pressure Research 30 (2010) 224–229. doi:10.1080/08957959.2010.485391.
565
[45] P. Althoff, Structural refinements of dolomite and a magnesian calcite and implications for dolomite formation in the marine environment, American Mineralogist 62 (1977) 772–783. 21
Journal Pre-proof
[46] R. L. Sass, R. Vidale, J. Donohue, Interatomic distances and thermal anisotropy in sodium nitrate and calcite, Acta Crystallographica 10 (1957) 570
567–570. doi:10.1107/S0365110X57002029. [47] A. Rosenthal, E. Hauri, M. Hirschmann, Experimental determination of C, F, and H partitioning between mantle minerals and carbonated basalt, CO2 /Ba and CO2 /Nb systematics of partial melting, and the CO2 contents
77–87. doi:10.1016/j.epsl.2014.11.044.
ro
575
of
of basaltic source regions, Earth and Planetary Science Letters 412 (2015)
[48] K. D. Litasov, A. Shatskiy, E. Ohtani, G. M. Yaxley, Solidus of alkaline car-
-p
bonatite in the deep mantle, Geology 41 (2013) 79–82. doi:10.1130/G33488.1.
re
[49] F. Maeda, E. Ohtani, S. Kamada, T. Sakamaki, N. Hirao, Y. Ohishi, Diamond formation in the deep lower mantle: A high-pressure reaction of MgCO3 and SiO2 , Scientific reports 7 (2017) 40602. doi:10.1038/srep40602.
lP
580
[50] P. Gillet, C. Biellmann, B. Reynard, P. McMillan, Raman spectroscopic
na
studies of carbonates part I: High-pressure and high-temperature behaviour of calcite, magnesite, dolomite and aragonite, Physics and Chemistry of
585
ur
Minerals 20 (1993) 1–18. doi:10.1007/BF00202245. [51] M. Stekiel, T. Nguyen-Thanh, S. Chariton, C. McCammon, A. Bosak,
Jo
W. Morgenroth, V. Milman, K. Refson, B. Winkler, High pressure elasticity of FeCO3 -MgCO3 carbonates, Physics of the Earth and Planetary Interiors 271 (2017) 57–63. doi:10.1016/j.pepi.2017.08.004. [52] T. Katsura, A. Yoneda, D. Yamazaki, T. Yoshino, E. Ito, Adiabatic 590
temperature profile in the mantle, Physics of the Earth and Planetary Interiors 183 (2010) 212–218. doi:10.1016/j.pepi.2010.07.001.
22
Journal Pre-proof
c* b*
c*
re
-p
ro
of
b*
Figure 1: Crystal structure of Dol-IIIc in projection along [100] at 43.4(8) GPa (left) and
Jo
ur
na
lP
reconstructed reciprocal lattice plane in the b∗ c∗ plane (right). (color online)
Figure 2: Crystal structure of Dol-V in projections along [001] and [100] at 46.2(8) GPa and after annealing at 1800(150) K. (color online)
23
Journal Pre-proof
10
Ne
12
lP 8
2
10
12
10
12
ur
na
Dol-V Ferromagnesite Post-aragonite c)4 1bar 6
re
-p
2
Ne
Ne
of
Dol-V Ferromagnesite Post-aragonite 8 K b)4 46.2 GPa, 6annealed at 1800
Intensity (a.u.)
Ne
59 GPa, annealed at 2300 K
ro
a)
Jo
Dol-I Ferromagnesite Calcite
4
6
8 2 ( )
Figure 3: Phase assemblages measured by X-ray diffraction (wavelength = 0.2893 ˚ A). a) After heating to 2300(150) K and quenching to ambient temperature at 59.3(8) GPa, powder patterns indicated Dol-V + ferromagnesite [44] + post-aragonite [43]. b) After heating to 1800(150) K and quenching to ambient temperature, powder patterns indicated Dol-V + Ferromagnesite [44] + post-aragonite [43]. c) After heating to 2300 K at 59 GPa and quenching to ambient conditions, powder patterns indicated a phase assemblage of Dol-I [45], Calcite [46], and Ferromagnesite [44]. (color online)
24
ro
55.2(8) GPa 52.4(8) GPa 47.7(8) GPa
na
Dol-V + Decomp.
43.8(8) GPa 36.4(7) GPa 33.9(7) GPa
25.1(5) GPa 21.3(4) GPa 17.9(4) GPa 12.6(3) GPa 1 bar
Dol-I + Decomp.
300 500 700 900 1100 Raman shift (cm 1)
Raman spectra of CaMg0.98 Fe0.02 (CO3 )2 polymorphs. a) Raman spectra of
ur
Figure 4:
of
b)
-p
Intensity (a.u.)
26 24 22 20 18 16 14 12 10 8 6 4 2 0
re
14 a) 12 41.8(8) GPa 39.5(4) GPa Dol-IIIc 10 34.7(8) GPa 27.8(4) GPa 8 21.3(4) GPa Dol-II 17.3(4) GPa 6 14.0(4) GPa 11.5(2) GPa Dol-Ib 4 9.1(1) GPa 6.0(2) GPa 2 1.8(1) GPa Dol-I 0 300 500 700 900 1100 Raman shift (cm 1)
lP
Intensity (a.u.)
Journal Pre-proof
Dol-I (black), Dol-II (red) and Dol-IIIc (blue) were measured upon compression at ambient
Jo
temperature. b) Raman spectra of quenched Dol-V + decomposition products (green) and Dol-I + decomposition products (black). Spectra between 43 and 52 GPa were measured upon compression and after annealing at 1800 K. The spectrum at 55 GPa was measured after annealing at 2300 K. Spectra between 36 GPa and ambient conditions were measured upon decompression and after heating to 2300 K. (color online)
25
Journal Pre-proof
Eg
100 0 Figure 5:
lP
Eg
20 40 Pressure (GPa)
na
200
1000
of
Ag
-p
300
1100
900
re
Ag
ro
Raman shift (cm 1)
500 400
1200
Dol-IIIc
Dol-I Dol-Ib
Dol-II
b)
800 700
Ag
Dol-II
Dol-IIIc Dol-V + decomp. product
Eg
600 0
20 40 Pressure (GPa)
Pressure dependence of the Raman shift for high-pressure, high-temperature
ur
Raman shift (cm 1)
600
1300
a) Dol-I Dol-Ib
700
dolomite polymorphs as well as its decomposition products in the low (a) and high (b)
Jo
frequency spectrum: Dol-I (black), Dol-II (red), Dol-IIIc (blue) and Dol-V + decomposition products (green). Data are presented for measurements during compression (full circles) and decompression (open circles). Diamond marker correspond to modes of Dol-II (dark-red) and Dol-V (light-green), which have been calculated by DFT. Irreducible representations of Raman modes are labeled for ambient conditions [50]. (color online)
26
Figure 6:
re
-p
ro
of
Quenched sample Dolomite Calcite Magnesite Siderite
lP
14 12 10 8 6 4 2 0
200
400
600 800 Raman shift (cm 1)
1000
1200
na
Intensity (a.u.)
Journal Pre-proof
Raman spectrum of quenched dolomite (CaMg0.98 Fe0.02 (CO3 )2 ) at ambient
ur
conditions (quenched after laser-heating at 2300 K and 55 GPa), which decomposed into a phase assemblage of dolomite, calcite, and magnesite with slight proportions of a siderite
Jo
component. Reference spectra of calcite, magnesite and siderite are plotted for comparison. All presented Raman spectra were measured at ambient conditions. The dolomite reference is the sample used in the current study before any pressure or temperature treatment was applied. Reference spectra of calcite and siderite were obtained from characterized reference materials [33, 51]. The reference spectrum of magnesite was obtained from the RRUFF open database (R050676). (color online)
27
of
Journal Pre-proof
4
DFT
re
3
ro
43.0(8) GPa, 1800(150) K
-p
5
34.7(8) GPa DFT
2 1 400
600
Dol-V + decomposition product Dol-V Dol-II Dol-II
800 1000 1200 1400 1600 Raman shift (cm 1)
na
200
lP
Intensity (a.u.)
6
ur
Figure 7: Experimental Raman spectra of Dol-II (red) and Dol-V + decomposition products (green) in comparison to DFT-calculated spectra of Dol-II (dark-red) and Dol-V (light-green).
Jo
For Dol-V, the frequency part of 800-1000 cm−1 shows a certain mismatch between theoretical and experimental data. The high-frequency part doesn’t show a splitting of the 1200 cm−1 mode in the calculations, however it is seen in the experiment. The mismatch is explained by additional modes coming from a carbonate decomposition assemblage. (color online)
28
na
lP
re
-p
ro
of
Journal Pre-proof
Figure 8: Synthesis conditions for polymorphs of Fe-poor dolomite (Fe ≤ 8 at %). Closed and open symbols correspond to data obtained upon compression and decompression, respectively.
ur
Dol-I (black), disordered Dol-I (orange), Dol-II (red), Dol-IIIc (blue), Dol-V + decomposition products (green), aragonite + magnesite (purple), decomposing Dol-II into aragonite + magnesite (red-purple), Fe-dolomite-III (grey). Data presented as circles were obtained in this
Jo
study on temperature quenched samples, squares by Mao et al. [10], triangles by Merlini et al. [12], left pointing triangles by Martinez et al. [19] and right pointing triangles by Antao et al. [24]. Dashdotted line (purple) corresponds to the dolomite-aragonite + magnesite transition as determined by Martinez et al. [19]. Dashed lines (black, red, blue) correspond to phase boundaries including the Dol-Ib phase described by Efthimiopoulos et al. [13]. Stability fields of the current study are delineated by solid red and green lines corresponding the Dol-I–Dol-II and Dol-II–Dol-V transitions, respectively. Solid grey line corresponds to an adiabatic geotherm, as described in [52]. (color online)
29
Journal Pre-proof
Table 1: Experimental set-up for single-crystal X-ray diffraction measurements of Dol-IIIc and Dol-V Dol-V
CaMg0.98 Fe0.02 (CO3 )2
CaMg0.98 Fe0.02 (CO3 )2
Size of crystal (µm3 )
10×15×10
25×20×15
Synchrotron source
PETRA III
ESRF
Beamline
P02.2
ID15b
Energy (keV)
41.7
30.1
Wavelength (˚ A)
0.2893
0.4113
9×3
10 × 10
ro
Spot-size
(µm2 )
of
Dol-IIIc Composition
Perkin Elmer XRD1621
MAR555
Omega range/step/exposure
±21◦ /0.5◦ /2s
±33◦ /0.5◦ /1s
Jo
ur
na
lP
re
-p
Detector
30
Journal Pre-proof
Table 2: Crystallographic information about experimental and theoretical data and details of crystal structure refinements of Dol-IIIc and Dol-V. Phase
Dol-IIIc DFT
Exp
DFT
Pressure (GPa) annealing Temp. (K)/duration (min)
43.4(8)
43.4
46.2(8)
43
-
-
1800(150)[a] /5
-
Temperature (K)
298
-
298
-
Crystal system
Triclinic
Triclinic
Monoclinic
Monoclinic
Space group
P¯ 1
P¯ 1
C2/c
C2/c
a (˚ A)
4.4518(4)
b (˚ A)
11.1683(17)
of
Exp
ro
analytical Method
Dol-V
4.4879
8.051(14)
8.1930
11.2572
7.848(3)
7.9595
13.6960(17)
13.7041
5.105(15)
5.1235
α (◦ )
69.044(13)
69.20
90.0
90.0
88.343(9)
88.73
106.4(3)
106.11
89.344(10)
89.77
90.0
90.0
(◦ )
re
β
-p
c (˚ A)
γ (◦ ) V (˚ A3 )
647.08
309.4(11)
320.00
3.85
3.78
3.95
3.81
8
8
4
4
lP
635.64(15)
ρ (g/cm3 ) Z
736
368
Theta range for data collection (◦ )
2.03–15.80
2.88–20.27
Completeness to d = 0.8 ˚ A(%)
30.85
40.52
Index ranges
-7 < h < 7
-13 < h < 12
-13 < k < 9
-13 < k < 13
0 < l < 21
-6 < l < 5
ur
Jo
Redundancy
na
F(000)
Reflections collected Rint
1.00
2.00
2063
410
0.05
0.026
Unique reflections (I>3σ(I))
1289
153
Unique reflections (all)
2062
205
161
24
R1 /wR2 (I>3σ(I))
0.060/0.064
0.057/0.060
R1 /wR2 (all)
0.086/0.065
0.081/0.064
No. of parameters
[a] [b]
[b]
Data collection was performed on temperature quenched samples. Due to the limited amount of available reflections, nearly all atoms were refined in the
isotropic approximation. For Dol-V it was possible to refine Ca anisotropically.
31
Journal Pre-proof
120
Dol-I Dol-II Dol-IIIc Dol-V
100 90 80
of
V/Z (Å 3)
110
600
20
P (GPa)
30
-p
10
ro
70
40
50
re
Figure S1: Pressure dependence of volume per formula unit of Dol-I (black), Dol-II (red),
Jo
ur
na
lP
Dol-IIIc (blue) and Dol-V (green). (color online)
S0
Dol-I
13(1) GPa, 1530(150) K 13.4(1) GPa
700
900
1100
900
1100
500 700 900 1 Raman shift (cm )
1100
ro
500
of
13(1) GPa, 1600(150) K
-p
Dol-II + decomposition Aragonite Dol-II
re
17(1) GPa, 1800(150) K 17(1) GPa, 1700(150) K 17.9(1) GPa 17.2(1) GPa
lP
500
700
Dol-II
ur
na
14 a) 12 10 8 6 4 2 0 300 14 b) 12 10 8 6 4 2 0 300 14 c) 12 10 8 6 4 2 0 300
Jo
Intensity (a.u.)
Intensity (a.u.)
Intensity (a.u.)
Journal Pre-proof
27(2) GPa, 1200(150) K 27(2) GPa
Figure S2: Raman spectra of temperature quenched dolomite samples at high pressures. a) Laser-annealed (up to 1600 K) Dol-I. No indication for Dol-Ib is present. b) Laser-annealed Dol-II (up to 1800 K). Splitting of the CO3 stretching mode is indicative for the decomposition of Dol-II. A reference spectrum of aragonite at ∼18 GPa is plotted for a phase identification [33]. c) No decomposition was observed for laser-annealed Dol-II at 1200 K (color online)
S1
Journal Pre-proof
Table S1: Refined atomic coordinates and isotropic displacement factors for Dol-IIIc. Dol-IIIc y
z
Uiso
0.2317(3)
0.4250(2)
0.67558(14)
0.0094(2)
Ca2
0.2485(3)
0.2098(2)
0.44820(14)
0.0090(2)
Ca3
0.7704(3)
0.2974(2)
1.03875(14)
0.0107(2)
Ca4
0.6891(2)
0.0512(2)
0.83600(14)
0.0088(2)
Mg1
0.7370(5)
0.2913(5)
0.5642(3)
0.0092(4)
Mg2
0.7477(4)
0.5835(4)
0.8124(2)
0.0065(4)
Mg3
0.2613(4)
0.2064(4)
0.9133(2)
Mg4
0.2851(4)
0.9212(4)
0.6920(2)
O1
0.9156(9)
0.9255(9)
0.7658(5)
0.0112(8)
O2
0.5139(9)
0.7027(9)
0.7064(5)
0.0110(8)
O3
0.5072(9)
0.2535(8)
0.7897(5)
0.0090(7)
O4
0.9603(9)
0.4691(9)
O5
0.5053(9)
0.4040(9)
O6
0.0151(9)
0.3961(9)
O7
0.9364(9)
O8 O9
of
x
Ca1
0.0081(4) 0.0076(4)
ro
-p
Atoms
0.0099(7)
0.4343(5)
0.0090(7)
0.4511(5)
0.0117(8)
0.1785(8)
0.6946(5)
0.0089(7)
0.6215(9)
0.9544(9)
0.5979(5)
0.0112(8)
0.9396(9)
0.8768(9)
0.9569(5)
0.0111(8)
0.7194(9)
0.8527(5)
0.0127(8)
0.0539(9)
0.8890(5)
0.0104(8)
0.3006(9)
0.8295(5)
0.0136(8)
0.4490(9)
0.7778(5)
0.0129(8)
0.8457(8)
0.8282(5)
0.0087(7)
lP
re
0.9252(5)
0.8697(10)
O11
0.1840(9)
O12
0.9543(10)
O13
0.6037(10)
O14
0.4806(9)
O15
0.3120(10)
0.2731(9)
0.6004(5)
0.0114(8)
O16
0.3799(9)
0.0811(9)
0.7089(5)
0.0112(8) 0.0115(8)
ur
O17
na
O10
0.3511(9)
0.3456(9)
0.9579(5)
0.1112(10)
0.5785(9)
0.7354(5)
0.0134(8)
0.2391(10)
0.5772(10)
0.3828(6)
0.0148(9)
O20
0.0784(9)
0.7847(9)
0.6693(5)
0.0105(8)
O21
0.3866(9)
0.8884(9)
1.0115(5)
0.0115(8)
O22
0.0761(9)
0.0370(8)
0.5750(5)
0.0086(7)
O23
0.4011(9)
0.5534(9)
0.9045(5)
0.0105(8)
O24
0.7027(9)
0.1587(9)
0.5116(5)
0.0108(8)
C1
0.2095(11)
0.1737(11)
0.6724(6)
0.0071(9)
C2
0.2536(12)
0.4575(12)
0.4273(7)
0.0091(9)
C3
0.1735(12)
0.9399(12)
0.9530(7)
0.0097(10)
C4
0.6914(12)
0.3351(11)
0.7997(6)
0.0080(9)
C5
0.2360(12)
0.6828(12)
0.6998(7)
0.0096(9)
C6
0.7937(12)
0.0451(11)
0.5627(7)
0.0092(9)
C7
0.2405(13)
0.4574(13)
0.9282(7)
0.0121(10)
C8
0.7574(13)
0.8313(12)
0.8154(7)
0.0107(10)
O18
Jo
O19
S2
ro
of
Journal Pre-proof
-p
Table S2: Refined atomic coordinates and isotropic/anisotropic displacement factors for
re
Dol-V.
Dol-V
Ca1
1
Mg1
0.5
y
z
Uani
0.6680(3)
0.75
0.018(3)
0.5888(4)
0.75
0.0128(8)
0.4406(6)
0.548(3)
0.0176(12)
lP
x
Uiso
0.8662(12)
O2
0.8389(12)
0.2423(6)
0.837(2)
0.0194(12)
O3
0.6224(12)
0.4034(5)
0.641(3)
0.0163(12)
C1
0.3665(8)
0.681(4)
0.0173(15)
na
O1
ur
Atoms
Jo
0.7827(17)
S3
Figure 1
Figure 2
Figure 3
Figure 4
Figure 5
Figure 6
Figure 7
Figure 8