HOSTED BY Geoscience Frontiers - HPSTAR

7
Research Paper Aluminum solubility in bridgmanite up to 3000 K at the top lower mantle Zhaodong Liu a, b, * , Ran Liu b , Yucheng Shang b , Fangren Shen b , Luyao Chen b , Xuyuan Hou b , Mingguang Yao b , Tian Cui b , Bingbing Liu b , Tomoo Katsura a, c a Bayerisches Geoinstitut, University of Bayreuth, Bayreuth, 95440, Germany b State Key Laboratory of Superhard Materials, Jilin University, Changchun 130012, China c Center for High Pressure Science and Technology Advanced Research, Beijing 100094, China ARTICLE INFO Keywords: Bridgmanite LiNbO 3 -type phase Corundum Temperature Entropy Lower mantle ABSTRACT The temperature dependence of the Al 2 O 3 solubility in bridgmanite has been determined in the system MgSiO 3 Al 2 O 3 at temperatures of 27503000 K under a constant pressure of 27 GPa using a multi-anvil appa- ratus. Bridgmanite becomes more aluminous with increasing temperatures. A LiNbO 3 -type phase with a pyrope composition (Mg 3 Al 2 Si 3 O 12 ) forms at 2850 K, which is regarded as to be transformed from bridgmanite upon decompression. This phase contains 30 mol% Al 2 O 3 at 3000 K. The MgSiO 3 solubility in corundum also increases with temperatures, reaching 52 mol% at 3000 K. Molar volumes of the hypothetical Al 2 O 3 bridgmanite and MgSiO 3 corundum are constrained to be 25.95 0.05 and 26.24 0.06 cm 3 /mol, respectively, and interaction parameters of non-ideality for these two phases are 5.6 0.5 and 2.2 0.5 KJ/mol, respectively. The increases in Al 2 O 3 and MgSiO 3 contents, respectively, in bridgmanite and corundum are caused by a larger entropy of Al 2 O 3 bridgmanite plus MgSiO 3 corundum than that of MgSiO 3 bridgmanite plus Al 2 O 3 corundum with temperature, in addition to the conguration entropy. Our study may help explain dynamics of the top lower mantle and constrain pressure and temperature conditions of shocked meteorites. 1. Introduction Geochemical and petrological studies suggest that Earths lower mantle (6602900 km depth), which occupies more than 50% of Earths volume, is the largest geochemical reservoir in the Earth (e.g., Ringwood, 1975; Anderson, 1983; McDonough and Sun, 1995). Bridgmanite is the most abundant phase in this region, and comprises about 80% of this region in volume (Irifune, 1994). Under lower-mantle conditions, bridgmanite can contain 25 mol% of alumina (Al 2 O 3 )(Liu et al., 2016, 2017a), which is far beyond its contents in the pyrolite and mid-oceanic ridge basalt (MORB) compositions (Green et al., 1979; Sun, 1982). Bridgmanite is thus the major host phase for the Al 2 O 3 in the lower mantle (Irifune et al., 1996; Liu et al., 2016, 2017a). Al 2 O 3 incorporation can greatly change chemical and physical properties of bridgmanite (e.g., McCammon, 1997; Xu et al., 1998; Zhang and Weidner, 1999; Brodholt, 2000). Therefore, the Al 2 O 3 solubility in bridgmanite is of great signi- cance for understanding the mineralogy of the lower mantle. The Al 2 O 3 solubility in bridgmanite has been studied in the system MgSiO 3 Al 2 O 3 by several workers (Irifune et al., 1996; Kubo and Akaogi, 2000; Hirose et al., 2001; Akaogi et al., 2002). Liu et al. (2016, 2017a) recently demonstrated that the Al 2 O 3 solubility in bridgmanite has both positive pressure and temperature dependences at pressures of 2752 GPa and temperatures of 17002500 K. They also predicted that a LiN- bO 3 -structured phase (LN) with a pyrope (Mg 3 Al 2 Si 3 O 12 ) composition should form at 27502800 K under 27 GPa upon decompression. LN has been regarded as a metastable phase formed from high-pressure stable bridgmanite by a diffusionless transformation upon decompression (e.g., Ross et al., 1989). The formation of LN can help understand the complicated crystal chemistry of bridgmanite and constrains shock con- ditions of meteorites (Ishii et al., 2017; Liu et al. 2019c). However, the temperature dependence of Al 2 O 3 solubility in bridgmanite is still poorly understood because the data obtained by our recent (Liu et al. 2017a) and earlier (Irifune et al., 1996; Kubo and Akaogi, 2000; Hirose et al., 2001; Akaogi et al., 2002) studies did not agree well. Furthermore, the chemistry of Al 2 O 3 component in bridgmanite remains unknown at temperatures higher than 2500 K. At higher temperatures, oxygen va- cancies in the form of an MgAlO 2.5 component should increase, and might thereby decrease the Al 2 O 3 component (Brodholt, 2000; Liu et al., * Corresponding author. Bayerisches Geoinstitut, University of Bayreuth, Bayreuth, 95440, Germany. E-mail address: [email protected] (Z. Liu). Peer-review under responsibility of China University of Geosciences (Beijing). HOSTED BY Contents lists available at ScienceDirect Geoscience Frontiers journal homepage: www.elsevier.com/locate/gsf https://doi.org/10.1016/j.gsf.2020.04.009 Received 21 October 2019; Received in revised form 1 February 2020; Accepted 12 April 2020 Available online 29 April 2020 1674-9871/© 2020 China University of Geosciences (Beijing) and Peking University. Production and hosting by Elsevier B.V. This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/). Geoscience Frontiers 12 (2021) 929935

Transcript of HOSTED BY Geoscience Frontiers - HPSTAR

Geoscience Frontiers 12 (2021) 929–935

HOSTED BY Contents lists available at ScienceDirect

Geoscience Frontiers

journal homepage: www.elsevier.com/locate/gsf

Research Paper

Aluminum solubility in bridgmanite up to 3000 K at the top lower mantle

Zhaodong Liu a,b,*, Ran Liu b, Yucheng Shang b, Fangren Shen b, Luyao Chen b, Xuyuan Hou b,Mingguang Yao b, Tian Cui b, Bingbing Liu b, Tomoo Katsura a,c

a Bayerisches Geoinstitut, University of Bayreuth, Bayreuth, 95440, Germanyb State Key Laboratory of Superhard Materials, Jilin University, Changchun 130012, Chinac Center for High Pressure Science and Technology Advanced Research, Beijing 100094, China

A R T I C L E I N F O

Keywords:BridgmaniteLiNbO3-type phaseCorundumTemperatureEntropyLower mantle

* Corresponding author. Bayerisches Geoinstitut,E-mail address: [email protected] (Z. LiuPeer-review under responsibility of China Univ

https://doi.org/10.1016/j.gsf.2020.04.009Received 21 October 2019; Received in revised forAvailable online 29 April 20201674-9871/© 2020 China University of GeoscienceCC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/).

A B S T R A C T

The temperature dependence of the Al2O3 solubility in bridgmanite has been determined in the systemMgSiO3–Al2O3 at temperatures of 2750–3000 K under a constant pressure of 27 GPa using a multi-anvil appa-ratus. Bridgmanite becomes more aluminous with increasing temperatures. A LiNbO3-type phase with a pyropecomposition (Mg3Al2Si3O12) forms at 2850 K, which is regarded as to be transformed from bridgmanite upondecompression. This phase contains 30 mol% Al2O3 at 3000 K. The MgSiO3 solubility in corundum also increaseswith temperatures, reaching 52 mol% at 3000 K. Molar volumes of the hypothetical Al2O3 bridgmanite andMgSiO3 corundum are constrained to be 25.95 � 0.05 and 26.24 � 0.06 cm3/mol, respectively, and interactionparameters of non-ideality for these two phases are 5.6 � 0.5 and 2.2 � 0.5 KJ/mol, respectively. The increases inAl2O3 and MgSiO3 contents, respectively, in bridgmanite and corundum are caused by a larger entropy of Al2O3

bridgmanite plus MgSiO3 corundum than that of MgSiO3 bridgmanite plus Al2O3 corundum with temperature, inaddition to the configuration entropy. Our study may help explain dynamics of the top lower mantle and constrainpressure and temperature conditions of shocked meteorites.

1. Introduction

Geochemical and petrological studies suggest that Earth’s lowermantle (660–2900 km depth), which occupies more than 50% of Earth’svolume, is the largest geochemical reservoir in the Earth (e.g., Ringwood,1975; Anderson, 1983; McDonough and Sun, 1995). Bridgmanite is themost abundant phase in this region, and comprises about 80% of thisregion in volume (Irifune, 1994). Under lower-mantle conditions,bridgmanite can contain 25 mol% of alumina (Al2O3) (Liu et al., 2016,2017a), which is far beyond its contents in the pyrolite and mid-oceanicridge basalt (MORB) compositions (Green et al., 1979; Sun, 1982).Bridgmanite is thus the major host phase for the Al2O3 in the lowermantle (Irifune et al., 1996; Liu et al., 2016, 2017a). Al2O3 incorporationcan greatly change chemical and physical properties of bridgmanite (e.g.,McCammon, 1997; Xu et al., 1998; Zhang and Weidner, 1999; Brodholt,2000). Therefore, the Al2O3 solubility in bridgmanite is of great signifi-cance for understanding the mineralogy of the lower mantle.

The Al2O3 solubility in bridgmanite has been studied in the systemMgSiO3–Al2O3 by several workers (Irifune et al., 1996; Kubo and Akaogi,

University of Bayreuth, Bayreuth).ersity of Geosciences (Beijing).

m 1 February 2020; Accepted 12

s (Beijing) and Peking University.

2000; Hirose et al., 2001; Akaogi et al., 2002). Liu et al. (2016, 2017a)recently demonstrated that the Al2O3 solubility in bridgmanite has bothpositive pressure and temperature dependences at pressures of 27–52GPa and temperatures of 1700–2500 K. They also predicted that a LiN-bO3-structured phase (LN) with a pyrope (Mg3Al2Si3O12) compositionshould form at 2750–2800 K under 27 GPa upon decompression. LN hasbeen regarded as a metastable phase formed from high-pressure stablebridgmanite by a diffusionless transformation upon decompression (e.g.,Ross et al., 1989). The formation of LN can help understand thecomplicated crystal chemistry of bridgmanite and constrains shock con-ditions of meteorites (Ishii et al., 2017; Liu et al. 2019c). However, thetemperature dependence of Al2O3 solubility in bridgmanite is still poorlyunderstood because the data obtained by our recent (Liu et al. 2017a)and earlier (Irifune et al., 1996; Kubo and Akaogi, 2000; Hirose et al.,2001; Akaogi et al., 2002) studies did not agree well. Furthermore, thechemistry of Al2O3 component in bridgmanite remains unknown attemperatures higher than 2500 K. At higher temperatures, oxygen va-cancies in the form of an MgAlO2.5 component should increase, andmight thereby decrease the Al2O3 component (Brodholt, 2000; Liu et al.,

, 95440, Germany.

April 2020

Production and hosting by Elsevier B.V. This is an open access article under the

SH-USER1
Text Box
HPSTAR 1146-2021

Fig. 1. (a) Cross section of the cell assembly. (b) Generated temperature as afunction of the heating powder. The inner picture is the BSE image of therecovered assembly of the run IRIS624.

Table 1Starting materials, experimental conditions, and runs products.

Run No. Start Comp. P (27 GPa)/T (K)/t(minutes)

Phases

IRIS624 En75Cor25 2750/10 LN þ Cor þ trace Brg þSti

IRIS635 En75Cor25 2850/3 LN þ Cor þ ReEn65Cor35 LN þ Cor þ Re

IRIS627 En65Cor35 3000/3 LN þ Cor þ Re

Abbreviations: Brg, bridgmanite; Cor, corundum; LN, LiNbO3-type phase; Sti,stishovite.

Z. Liu et al. Geoscience Frontiers 12 (2021) 929–935

2019a, b; Navrotsky et al., 2003), which should decrease ferropericlase inthe lower mantle. Temperatures in hot plumes from the lower mantleshould be significantly higher by 500–1000 K than the surroundingmantle (e.g., Farnetani, 1997). Consequently, temperature effects arevital to constrain dynamics of the lower mantle. Nevertheless, our limitedknowledge about the Al2O3 solubility in bridgmanite at very high tem-peratures prevents our understandings mantle dynamics.

Here, we investigated the Al2O3 solubility in bridgmanite at tem-peratures up to 3000 K and a constant pressure of 27 GPa in a multi-anvilpress. This pressure can prevent the formation of majoritic garnet. Basedon our new results, we have determined the temperature dependence ofthe Al2O3 solubility in bridgmanite together with the MgSiO3 solubilityin corundum, estimated thermoelastic parameters of bridgmanite andcorundum, and discussed their implications for dynamics of the lowermantle.

2. Experimental methods

En75Cor25 (Mg3Al2Si3O12, En: MgSiO3, Cor: Al2O3; the number rep-resents mol%) and En65Cor35 glasses were used as starting materials.Detailed chemical composition of the En75Cor25 glass was reported in Liuet al. (2016), and that of the En65Cor35 glass was confirmed in the presentstudy to have the intended composition using an electron probe micro-analyzer. Starting materials were put into rhenium (Re) capsules made of25 μm thick foils, and then heated at 500 K for half an hour to purgewater. The capsules with starting materials were finally put into cellassemblies. Quench experiments at a constant pressure of 27 GPa andtemperatures of 2700–3000 K were performed using a Cr2O3-doped MgO(OMCR, Mino Ceramic Co., LTD.) octahedra with a 7-mm edge lengthand LaCrO3 sleeves for heating in combination with tungsten carbidecubes with 3 mm truncated edge lengths in a multi-anvil apparatus

930

(IRIS-15) with a press load of 15 MN at the Bayerisches Geoinstitut,University of Bayreuth (Ishii et al., 2016). Temperature was monitoredwith a W97Re3–W75Re25 thermocouple placed adjacent to the Re capsule.The adapted cell assembly can be referred to Fig. 1. The pressure at hightemperatures was calibrated by the decomposition of pyrope intoaluminous bridgmanite and corundum and the Al2O3 solubility inbridgmanite at various temperatures (Liu et al., 2017b). Pressure un-certainties of these quench experiments are �1 GPa.

Phases in quench runs were identified using a micro-focused X-raydiffractometer (XRD, Bruker, D8 DISCOVER) equipped with a Co tubeoperated at 40 kV and 500 μA. X-ray beams were focused to 50 μm indiameter using an IFG polycapillary X-ray mini-lens. The XRD profile ofeach sample was collected for 3 h. Textural observation was performedusing a LEO1530 scanning electron microscope (SEM) operating at anacceleration voltage of 15 kV and a beam current of 10 nA. Chemicalcompositions of each phase present in the quench runs were determinedusing a JEOL JXA-8200 electron probe microanalyzer (EPMA) operatingat acceleration voltages of 15 kV and a beam current of 5 nA with stan-dards of enstatite for Mg and Si, and corundum for Al.

3. Results

Experimental conditions and phases present in the recovered samplesare listed in Table 1.

Fig. 1a shows cross section of the used cell assembly, and Fig. 1b plotsthe generated temperature as a function of the power supplied by theLaCrO3 heater at 27 GPa. It is clearly seen that the temperature cansuccessfully read 2399 �C. Above this temperature, we cannot readtemperature because it exceeded the limit of temperature display of thethermometer. We thereby estimated the temperature from the extrapo-lation by a cubic polynomial fitting of the temperatures below 2399 �Cand heating power. In the run 627, the heating power reached ~760 Wand the estimated temperature was ~3000 � 50 K.

Fig. 2 shows XRD patterns of run products. Fig. 3 shows their back-scattered electron (BSE) images. At 2750 K, both XRD and BSE obser-vations exhibit that run products from En75Cor25 consist of LN with traceamounts of corundum, bridgmanite, and stishovite (Figs. 2a and 3a). Weassumed that LN is a product of back-transformation of bridgmanite upondecompression based on previous results (Funamori et al., 1997; Liuet al., 2016; Miyajima et al., 1999; Ishii et al., 2017). Therefore, wereferred to the composition of LN as “bridgmanite composition” here-after. The formation of large amounts of LN in this bulk compositionindicates that bridgmanite incorporates close to 25 mol% of Al2O3 evenat the pressure much lower than previous studies, above 40 GPa (Liuet al., 2017a). Trace amounts of bridgmanite may be a metastableremnant from the back-transition. The BSE image shows no distin-guishable BSE signal intensities between LN and bridgmanite, indicatingidentical compositions of these two phases. One of these phases (bridg-manite) should therefore be metastable based on no binary phase loop ofthese two phases. From the run at 2850 K, the run product completelyconsists of a single LN with grain sizes larger than 10 μm. From the runsat 2850 and 3000 K from En65Cor35, the run product consists of a mixtureof LN and corundum based on XRD identifications and BSE observations(Figs. 2c, d and 3c, d).

Fig. 2. XRD profiles of the run products. The number in parenthesis represents the miller induces of the first appearing phase. Question masks represent the unknownpeaks. Abbreviations: Brg, bridgmanite; Cor, corundum; LN, LiNbO3-type phase; Sti, stishovite; Re, rhenium.

Z. Liu et al. Geoscience Frontiers 12 (2021) 929–935

One question is whether melting occurs at the highest experimentaltemperature of 3000 K (run 627). SEM observations, however, did notshow any features of melting in this run as follows. First, phase compo-sitions are identical within analytical uncertainties in the radial direc-tion. Second, although particles of Re or its oxide were found in theboundary of LN and corundum (Fig. 3b–d), the amounts of these particlessignificantly increase from the center to the ends of the sample capsule.These Re phases are thought to have been formed by a solid diffusion ofthe capsule at very high temperatures (Th�evenin et al., 1993). Further-more, Kudo and Ito (1996) studied the melting phase relations in thissystem at 25 GPa and reported that bridgmanite with 10 mol% Al2O3melted at higher temperature than the melting temperature of the pureMgSiO3 bridgmanite, which is 2900 K. Their report suggested that theincorporation of Al2O3 component must raise the melting temperature ofbridgmanite. These observations and report lead to the conclusion thatmelting did not occur at the temperature of the run 627. Therefore, themelting temperature of bridgmanite with the En75Cor25 compositionshould be higher than 3000 K at 27 GPa.

Compositional analysis by EPMA confirmed that LN/bridgmanitecontained 24 � 1 mol% Al2O3 at 2750 K from the En75Cor25 startingmaterial (Table 2). The LN in the run 635 from this starting materialcontained 25.0 � 0.2 mol% Al2O3 at 2850 K, which is identical to that ofthe starting material. From the En65Cor35 starting material in the samerun, the Al2O3 content in LN reached 26 � 1 mol%, while the MgSiO3content in corundum was 42 � 1 mol% (Table 2). In the run 627 (3000K), the Al2O3 content in LN and the MgSiO3 content in corundum werefound to be 29 � 1 and 52 � 1 mol%, respectively (Table 2). Latticeparameters of this LN (a ¼ b ¼ 4.849 (1) Å; c ¼ 12.712 (10) Å; the

931

number in the parenthesis represents the standard deviation for the lastdigit.) are slightly higher than obtained by Ishii et al. (2017) those (a¼ b¼ 4.783 (2) Å; c ¼ 12.680 (11) Å). These differences can be explained bythe higher Al2O3 content in the present study. As already mentioned, theformation of these Al-rich LN was recovered from bridgmanite synthe-sized at lower-mantle conditions through a diffusionless transition, inwhich the A (Mg) cations are substituted and BO6 (B, Si) octahedra aredistorted due to the incorporation of large amounts of Al (Liu et al., 2016,2019c; Ishii et al., 2017).

4. Discussion

4.1. Thermodynamics of Al partitioning in bridgmanite and corundum

To further understand the Al3þ exchange between bridgmanite andcorundum, we conducted the thermodynamics calculation using thefollowing reaction:

MgSiO3 (Brg) þ Al2O3 (Cor) ¼ Al2O3 (Brg) þ MgSiO3 (Cor) (1)

where Al2O3 and MgSiO3 are the hypothetical end-member of bridg-manite (Brg) and corundum (Cor), respectively. The Gibbs free energychange of reaction (1) can be expressed as:

μBrgMgSiO3þ μCorAl2O3

¼ μBrgAl2O3þ μCorMgSiO3

(2)

where μBrgMgSiO3and μBrgAl2O3

are the chemical potentials of MgSiO3 and Al2O3

in bridgmanite, respectively, and μCorAl2O3and μCorMgSiO3

are those of Al2O3

and MgSiO3 components in corundum, respectively. Chemical potentials

Fig. 3. BSE images of run products. Abbreviations: Brg, bridgmanite; Cor, corundum; LN, the LiNbO3-type phase; Sti, stishovite; Re, rhenium.

Table 2Chemical compositions of the bridgmanite and LN-type phases.

Run.No.

IRIS624 IRIS635 IRIS627

Comp. En75Cor25 En75Cor25 En65Cor35 En65Cor35

Phases Brg/LN (n¼ 12)

LN (n ¼11)

LN (n ¼10)

Cor (n¼ 8)

LN (n¼ 12)

Cor (n¼ 8)

MgO 30.58 (68) 30.26(22)

29.85(70)

16.47(49)

28.19(45)

20.62(83)

Al2O3 23.77 (54) 25.42(20)

26.50(55)

58.30(72)

28.96(81)

49.22(119)

SiO2 45.34 (91) 44.73(45)

43.87(43)

24.70(93)

42.60(42)

31.09(75)

Total 99.70(119)

97.82(58)

100.21(96)

99.48(76)

99.74(41)

100.93(68)

Mg 0.767 (11) 0.754 (5) 0.745(13)

0.416(11)

0.707(11)

0.244(14)

Al 0.471 (13) 0.501 (3) 0.523(12)

1.165(21)

0.574(16)

0.718(13)

Si 0.763 (9) 0.747 (4) 0.735(4)

0.419(13)

0.716(7)

0.277(12)

Σcation 2.001 (5) 2.002 (3) 2.003(4)

1.999(4)

1.997(2)

1.998(5)

Component (mol%)MgSiO3 76 (1) 75 (0) 74 (1) 42 (1) 71 (1) 52 (1)Al2O3 24 (1) 25 (0) 26 (1) 58 (1) 29 (1) 48 (1)Total 100 100 100 100 100 100

Oxide analyses are reported in wt.%. n: number of analysis points. The oxygennumber of Brg and LN is normalized to 3. Number in parentheses representsstandard deviation for the last digit (s). Abbreviations: Brg, bridgmanite; LN,LiNbO3-type.

Z. Liu et al. Geoscience Frontiers 12 (2021) 929–935

of these components can be expressed as the following equations in thenon-ideal solution model:

932

μBrgMgSiO3¼ μ�BrgMgSiO3

þ RT ln aBrgMgSiO3(3)

μCorAl2O3¼ μ�CorAl2O3

þ RT ln aCorAl2O3(4)

μBrgAl2O3¼ μ�BrgAl2O3

þ RT ln aBrgAl2O3(5)

μCorðMg0:5Si0:5Þ2O3¼ μ�CorMgSiO3

þ RT ln aCorMgSiO3(6)

where μ�BrgMgSiO3, μ�CorAl2O3

, μ�BrgAl2O3, and μ�CorðMg0:5Si0:5Þ2O3

are the standard

chemical potentials of the (hypothetical) endmembers of these fourcomponents, and aBrgMgSiO3

, aCorAl2O3, aBrgAl2O3

, and aCorMgSiO3are their activities. At

equilibrium, the standard state Gibbs free energy change of reaction (1)can be then expressed as:

ΔG0R1 ¼ μ�BrgAl2O3

þ μ�CorðMg0:5Si0:5Þ2O3� μ�BrgMgSiO3

� μ�CorAl2O3¼RT ln aBrgMgSiO3

þRT ln aCorAl2O3�RT ln aBrgAl2O3

� RT ln aCorMgSiO3(7)

The activity coefficients for these components in bridgmanite andcorundum can be expressed as:

aBrgAl2O3¼ �

XBrgAl2O3

� γBrgAl2O3

�2(8)

aBrgMgSiO3¼�XBrgMgSiO3

� γBrgMgSiO3

�2(9)

aCorAl2O3¼�XCorAl2O3

� γCorAl2O3

�2(10)

aCorMgSiO3¼�XCorMgSiO3

� γCorMgSiO3

�2(11)

Fig. 4. Molar volume of bridgmanite and corundum in the system MgSiO3-

–Al2O3 in previous studies.

Fig. 5. Gibbs free energy of reaction (1) as a function of temperature. The dataat temperatures of 1700–2500 K are from Liu et al. (2016, 2017a).

Z. Liu et al. Geoscience Frontiers 12 (2021) 929–935

By assuming the symmetric solutions for bridgmanite and corundum,their activity coefficients can be written as:

RT lnγBrgAl2O3¼WBrg

Al

�1� XBrg

Al2O3

�2(12)

RT lnγBrgMgSiO3¼WBrg

Al

�1� XBrg

MgSiO3

�2(13)

RT lnγCorAl2O3¼WCor

Al

�1� XCor

Al2O3

�2(14)

RT lnγCorMgSiO3¼WCor

Al

�1� XCor

MgSiO3

�2(15)

where XBrgAl2O3

and XBrgMgSiO3

are the mole fractions of Al2O3 and MgSiO3 in

bridgmanite, respectively, XCorAl2O3

and XCorMgSiO3

are the mole fractions of

Al2O3 and MgSiO3 components in the corundum, respectively, and WBrgAl

andWCorAl are the interaction parameters of bridgmanite and corundum of

the symmetric solutions, respectively. ΔG0R1 can be expressed as:

ΔG0R1 ¼ � 2RT ln

XCorMgSiO3

XBrgAl2O3

XCorAl2O3

XBrgMgSiO3

þ 2WCorAl

�1� 2XCor

Al2O3

�þ 2WBrg

Al

�2XBrg

Al2O3� 1

(16)

Molar volumes of bridgmanite and corundum as a function of theAl2O3 and MgSiO3 contents, respectively, are summarized in Fig. 4(D’Amour et al., 1978; Ito et al., 1978; Irifune et al., 1996; Kubo andAkaogi, 2000; Liu et al., 2016, 2017a). It is clearly seen that molar vol-umes of bridgmanite and corundum, respectively, increase almost line-arly with increasing Al2O3 and MgSiO3 contents within analyticaluncertainties. A linear relation leads to the following equation for thebridgmanite volume in cm3/mol:

VBrg ¼ 24.44 þ 0.0151(5) � χAl2O3 (17)

where χAl2O3 represents the mole fraction of the Al2O3 in bridgmanite.The derived molar volume of the hypothetical Al2O3 bridgmanite is to be25.95 � 0.05 cm3/mol. Since the molar volume of corundum is 25.56cm3/mol under ambient conditions (D’Amour et al., 1978), the volume ofAl2O3 bridgmanite is larger than that of pure corundum.

The same equation for corundum produces the following equation incm3/mol:

933

VCor ¼ 25.56 þ 0.0068 (6) � χMgSiO3(18)

where χMgSiO3represents mole fractions of the MgSiO3 content in

corundum. It is thus found that the effect of MgSiO3 contents on thevolume of corundum is significantly smaller than that of Al2O3 forbridgmanite. We have derived the molar volume of the hypotheticalMgSiO3 corundum of 26.24 � 0.06 cm3/mol. This value is very close tothe volume of MgSiO3 akimotoite, which is 26.35 cm3/mol (Horiuchiet al., 1982). Since akimotoite has an ilmenite-structure, which is amodification of the corundum structure by cation ordering, theMgSiO3-bearing corundum may form a continuous solid solution withakimotoite, and cation ordering may occur from a certain composition inthe Al2O3–MgSiO3 system (Panero et al., 2006).

We derived WCorAl and WBrg

Al by an empirical method for the non-ideality of solid solutions due to a mismatch of the component volumesusing the following formula (Davies and Navrotsky, 1983; Akaogi andIto, 1999):

WG ¼ 100:8ΔV – 0:4 ðkJ=molÞ (19)

ΔV ¼ VA � VB

ðVA þ VBÞ=2 (20)

where VA and VB are the molar volumes of the larger and smaller com-ponents, respectively. We then obtained the WBrg

Al and WCorAl values of 5.6

� 0.2 and 2.2� 0.2 kJ/mol, respectively, from the present molar volumeresults and equations of (17) and (18).

In Fig. 5, the derived ΔG0R1 from reaction (16) changes from 106� 13

kJ mol�1 at 1700 K to 75 � 7 kJ mol�1 at 2300 K and to 38 � 4 kJ mol�1

at 3000 K. The composition data of bridgmanite and corundum at tem-peratures of 1700–2500 K is from Liu et al. (2016, 2017a). By fitting thepresent experimental data to a linear function of ΔG0

R1 ¼ ΔH0R1 � TΔS0R1,

we obtained ΔS0R1 ¼ ��

∂ΔG0R1

∂T

�P¼ 46� 3 J/(mol⋅K). This fact suggests

that the configuration entropy of Al2O3 bridgmanite and MgSiO3corundum increase more than MgSiO3 bridgmanite and Al2O3 corundumwith increasing temperatures, which is consistent with ab initio calcula-tions (Panero et al., 2006; Jung et al., 2010). It further means that heatcapacities of two former components are larger than the lattercomponents.

Fig. 6 illustrates the solubility of Al2O3 in bridgmanite and that ofMgSiO3 in corundum, respectively, as function of temperature up to3000 K at 27 GPa in the present and previous studies (Irifune et al., 1996;

Fig. 6. The Al2O3 and MgSiO3 contents, respectively, in bridgmanite and corundum as a function of temperature. Dashed lines are thermodynamics calculation results.

Z. Liu et al. Geoscience Frontiers 12 (2021) 929–935

Kubo and Akaogi, 2000; Akaogi et al., 2002; Liu et al., 2016, 2017a). Inthe present study, the Al2O3 solubility in bridgmanite apparently in-creases from 24 � 1 to 29 � 1 mol% with increasing temperature from2750 to 3000 K. At 2850 K, bridgmanite contains 25 mol% Al2O3, andthereby transforms into LN upon decompression. At 3000 K, we foundthat corundum incorporates the MgSiO3 component as much as 52 � 1mol%, which is significantly higher than previous studies (Irifune et al.,1996; Kubo and Akaogi, 2000; Akaogi et al., 2002; Liu et al., 2016,2017a). Dashed lines are the thermodynamic calculations using thefollowing equation:

ΔGR1ðP; T ; XÞ¼ΔH0T �TΔS0T þ

Z P

1 atmΔVP; T þ 2RTln

XCorMgSiO3

XBrgAl2O3

XCorAl2O3

XBrgMgSiO3

� 2WCorAl

�1� 2XCor

Al2O3

�� 2WBrg

Al

�2XBrg

Al2O3� 1

�(21)

where ΔH0T is the enthalpy; ΔS0T is the configuration entropy; T is tem-

perature; P is the pressure; ΔVP; T is the molar volume at high pressureand high temperature. Thermoelastic parameters of bridgmanite andcorundum can be found in the present study for molar volume and en-tropy, and other parameters from Akaogi and Ito (1999) and Liu et al.(2019b). Our thermodynamics calculations show almost consistent re-sults for the experimental data of bridgmanite. However, there existssome difference between our calculation and experimental data forcorundum at temperature above 2400 K. This may be caused by poorlyconstrained thermoelastic parameters and large uncertainties of theinteraction parameter for the Al2O3 corundum–MgSiO3 akimotoite,which needs further studies.

4.2. Implications

It is clearly found that the Al2O3 solubility in bridgmanite signifi-cantly increases with increasing temperatures. Bridgmanite can containthe Al2O3 content up to 30 mol% at 3000 K even at a relatively lowpressure of 27 GPa. This value is far beyond it in MORB compositions.From subducted basaltic slabs to the surrounding ambient lower mantle,and then to upwelling of hot plumes at the top region of the lowermantle,bridgmanite would become more aluminous due to increasing tempera-tures in this order, if the hot plumes have a higher Al2O3 content than

934

slabs and ambient lower mantle. If the plume at the top region of thelower mantle is dominated by bridgmanite, bridgmanite with a higherAl2O3 content would become denser. Therefore, it may play the role of abarrier for the high temperature plumes upwelling from the lowermantle.

The LN can be used an indicator for constraining pressure and tem-perature conditions of shocked meteorites (e.g., Sharp et al., 1997; Ishiiet al., 2017). The present study shows a clear temperature dependence ofthe Al2O3 content in LN at the top lower mantle. Therefore, the presenceof LN with aluminous enstatite compositions could be used to constrainthe formation conditions of shocked meteorites. It is emphasized that,however, the presence of LN or bridgmanite with a garnet compositiondoes not directly mean that the shock pressure was above 40 GPa (Ishiiet al., 2017; Liu et al., 2019c). If the temperature had been higher than2000 K, the shock pressure should be lower than 40 GPa, and it can be 27GPa if the temperature were above 2800 K.

5. Conclusions

Phase relations in the system MgSiO3–Al2O3 have been studied attemperatures of 2750–3000 K under a constant pressure of 27 GPa in amulti-anvil press. It is found that both the Al2O3 and MgSiO3 contents,respectively, in bridgmanite and corundum have a positive temperaturedependence. Bridgmanite with the Al2O3 content higher than 25 mol%transforms into LN upon decompression due to the incorporation of largeamounts of Al. Bridgmanite and corundum, respectively, contain theAl2O3 and MgSiO3 contents up to 30 and 52 mol.% at a temperature of3000 K. We constrained the partial molar volumes and interaction pa-rameters of the hypothetical end-members of Al2O3 bridgmanite andMgSiO3 corundum. The increase in Al2O3 and MgSiO3 contents, respec-tively, in bridgmanite and corundum with temperature is enhanced bytheir positive entropy in addition to the configuration entropy. Tem-perature dependence of the Al2O3 content in bridgmanite would helpunderstand the dynamics of the lower mantle and constrain the pressureand temperature conditions for shocked meteorites.

Declaration of competing interest

The authors declare that they have no known competing financialinterests or personal relationships that could have appeared to influence

Z. Liu et al. Geoscience Frontiers 12 (2021) 929–935

the work reported in this paper.

Acknowledgments

We thank E. Posner, D. Krauβe, R. Njul, H. Fischer, and S. Übelhackfor their assistance with sample and high-pressure assembly preparation.The manuscript is greatly improved by the constructive comments ofShatskiy Anton and one anonymous reviewers and the Handing Editor. Z.L. was financially supported by the Bayerisches Geoinstitut Visitor’sProgram and the Fundamental Research Funds for the Central Univer-sities of Ministry of Education of China (Grant No. 45119031C037). Thisproject has received funding from the European Research Council (ERC)under the European Union’s Horizon 2020 research and innovationprogramme (Proposal No. 787 527) It is also supported by researchgrants to T. K. (BMBF: 05K13WC2, 05K16WC2; DFG: KA3434/3–1,KA3434/7–1, KA3434/8–1, KA3434/9–1) and Z. L. (the National ScienceFoundation of China Grant No. 41902034).

References

Akaogi, M., Ito, E., 1999. Calorimetric study on majorite–perovskite transition in thesystem Mg4Si4O12–Mg3Al2Si3O12: transition boundaries with positivepressure–temperature slope. Phys. Earth Planet. In. 114, 129–140.

Akaogi, M., Tanaka, A., Ito, E., 2002. Garnet-ilmenite-perovskite transitions in the systemMg4Si4O12–Mg3Al2Si3O12 at high pressures and high temperatures: phase equilibria,calorimetry and implications for mantle structure. Phys. Earth Planet. In. 132,303–324.

Anderson, D.L., 1983. Chemical composition of the mantle. J. Geophys. Res. 88, 41–B52.Brodholt, J.P., 2000. Pressure-induced changes in the compression mechanism of

aluminous perovskite in the Earth’s mantle. Nature 407, 620–622.Davies, P.K., Navrotsky, A., 1983. Quantitative correlations of deviations from ideality in

binary and pseudo-binary solid solutions. J. Solid State Chem. 46, 1–22.D0Amour, H., Schiferl, D., Denner, W., Schulz, H., Holzapfel, W.B., 1978. High-pressure

single-crystal structure determinations for ruby up to 90 kbar using an automaticdiffractometer P ¼ 0 kbar. J. Appl. Phys. 49, 4411–4416.

Farnetani, C.G., 1997. Excess temperature of chemical stratification mantle plumes: therole of chemical stratification across D". Geophys. Res. Lett. 24, 1583–1586.

Funamori, N., Yagi, T., Miyajima, N., Fujino, K., 1997. Transformation in garnet: fromorthorhombic perovskite to LiNbO3 phase on release of pressure. Science 275,513–515.

Green, D.H., Hibberson, W.O., Jaques, A.L., 1979. Petrogenesis of mid-ocean ridgebasalts. In: McElhinny, M.W. (Ed.), The Earth: its Origin, Structure and Evolution.Academic Press, London, pp. 269–299.

Hirose, K., Fei, Y., Yagi, T., Funakoshi, K., 2001. In situ measurements of the phasetransition boundary in Mg3Al2Si3O12: implications for the nature of the seismicdiscontinuities in the Earth’s mantle. Earth Planet Sci. Lett. 184, 567–573.

Horiuchi, H., Hirano, M., Ito, E., Matsui, Y., 1982. MgSiO3 (ilmenite-type): single crystalX-ray diffraction study. Am. Mineral. 67, 788–793.

Irifune, T., 1994. Absence of an aluminous phase in the upper part of the Earth’s lowermantle. Nature 370, 131–133.

Irifune, T., Koizumi, T., Ando, J., 1996. An experimental study of the garnet-provskitetransformation in the system MgSiO3–Mg3Al2Si3O12. Phys. Earth Planet. In. 96,147–157.

Ishii, T., Shi, L., Huang, R., Tsujino, N., Druzhbin, D., Myhill, R., Li, Y., Wang, L.,Yamamoto, T., Miyajima, N., Kawazoe, T., Nishiyama, N., Higo, Y., Tange, Y.,Katsura, T., 2016. Generation of pressure over 40 GPa using Kawai-type multi–anvil

935

press with tungsten carbide anvils. Rev. Sci. Instrum. 87 (2) https://doi.org/10.1063/1.4941716.

Ishii, T., Sinmyo, R., Komabayashi, T., Boffa-Ballaran, T., Kawazoe, T., Miyajima, N.,Hirose, K., Katsura, T., 2017. Synthesis and crystal structure of LiNbO3-typeMg3Al2Si3O12: a possible indicator of shock conditions of meteorites. Am. Mineral.102, 1947–1952.

Ito, E., Matsui, Y., 1978. Synthesis and crystal-chemical characterization of MgSiO3perovskite. Earth Planet Sci. Lett. 38, 443–450.

Jung, D.Y., Vinograd, V.L., Fabrichnaya, O.B., Oganov, A.R., Schmidt, M.W., Winkler, B.,2010. Thermodynamics of mixing in MgSiO3–Al2O3 perovskite and ilmenite from abinitio calculations. Earth Planet Sci. Lett. 295, 477–48.

Kubo, A., Akaogi, M., 2000. Post–garnet transitions in the systemMg4Si4O12–Mg3Al2Si3O12 up to 28 GPa: phase relations of garnet, ilmenite andperovskite. Phys. Earth Planet. In. 121, 85–102.

Kudo, R., Ito, E., 1996. Melting relations in the system Mg4Si4O12 (En)-Mg3Al2Si3O12 (Py)at high pressures. Phys. Earth Planet. In. 96, 159–169.

Liu, Z.D., Akaogi, M., Katsura, T., 2019b. Increase of the oxygen vacancy component inbridgmanite with temperature. Earth Planet Sci. Lett. 505, 141–151.

Liu, Z.D., Boffa Ballaran, T., Frost, D., Katsura, T., 2019a. Strong correlation of oxygenvacancy in bridgmanite with Mg/Si ratios. Earth Planet Sci. Lett. 523, 115697.

Liu, Z.D., Dubrovinsky, L., McCammon, C, Ovsyannikov, S.V., Koemets, I., Chen, L.Y,Cui, Q., Su, N., Cheng, J.G., Cui, T., Liu, B.B., Katsura, T., 2019c. A new (Mg0.5Fe3þ0.5)(Si0.5Al3þ0.5)O3 LiNbO3-type phase synthesized at lower mantle conditions. Am.

Miner. 104, 1213–1216.Liu, Z.D., Irifune, T., Nishi, M., Tange, Y., Arimoto, T., Shinmei, T., 2016. Phase relations

in the system MgSiO3– Al2O3 up to 52 GPa and 2000 K. Phys. Earth Planet. In. 257,18–27.

Liu, Z.D., Ishii, T., Katsura, T., 2017b. Rapid decrease of MgAlO2.5 component inbridgmanite with pressure. Geochem. Perspect. Lett. 5, 12–18.

Liu, Z.D., Nishi, M., Ishii, T., Fei, H.Z., Miyajima, N., Boffa Ballaran, T., Ohfuji, H.,Sakai, T., Wang, L., Shcheka, S., Arimoto, T., Tange, Y., Higo, Y., Irifune, T.,Katsura, T., 2017a. Phase relations in the system MgSiO3–Al2O3 up to 2300 K atlower-mantle pressures. J. Geophys. Res. 122 (10), 7775–7788.

McCammon, C.A., 1997. Perovskite as a possible sink for ferric iron in the lower mantle.Nature 387, 694–696.

McDonough, W.F., Sun, S.-s., 1995. The composition of the Earth. Chem. Geol. 120,223–253.

Miyajima, N., Fujino, K., Funamori, N., Kondo, T., Yagi, T., 1999. Garnet-perovskitetransformation under conditions of the Earth’s lower mantle: an analyticaltransmission electron microscopy study. Phys. Earth Planet. In. 116, 117–131.

Navrotsky, A., Schoenitz, M., Kojitani, H., Xu, H., Zhang, J., Weidener, D.J., Jeanloz, R.,2003. Aluminum in magnesium silicate perovskite: formation, structure, andenergetics of magnesium-rich defect solid solutions. J. Geophys. Res. 108, 2330.

Panero, P.W., Akber-Knutson, S., Akber-Knutson, L., 2006. Al2O3 incorporation in MgSiO3perovskite and ilmenite. Earth Planet Sci. Lett. 252, 152–161. https://doi.org/10.1016/j.epsl.2006.09.036.

Ringwood, A.E., 1975. Composition and Petrology of the Earth’s Mantle. McGraw–Hill,New York.

Ross, N.L., Ko, J., Prewvitt, C.T., 1989. A new phase transition in MnTiO3: LiNbO3-perovskite structure. Phys. Chem. Miner. 16, 621–629.

Sharp, T.G., Lingemann, C.M., Dupas, C., St€offler, D., 1997. Natural occurrence ofMgSiO3-ilmenite and evidence for MgSiO3-perovskite in a shocked L chondrite.Science 277, 352–355.

Sun, S., 1982. Chemical composition and the origin of the Earth’s primitive mantle.Geochem. Cosmochim. Acta 46, 179–192.

Th�evenin, Th, Arl�es, L., Boivineau, M., Vermeulen, J.M., 1993. Thermophysical propertiesof rhenium. Int. J. Thermophys. 14, 441–448.

Xu, Y., McCammon, C.A., Poe, B.T., 1998. The effect of alumina on the electricalconductivity of silicate perovskite. Science 282, 922–924.

Zhang, J., Weidner, D.J., 1999. Thermal equation of state of aluminium-enriched silicateperovskite. Science 284, 782–784.