paleoproterozoic high-pressure metamorphic history of the...

19
1 RESEARCH Paleoproterozoic high-pressure metamorphic history of the Salma eclogite on the Kola Peninsula, Russia Takeshi Imayama 1 , Chang-Whan Oh 2 , Shauket K. Baltybaev 3 , Chan-Soo Park 4 , Keewook Yi 4 , and Haemyeong Jung 5 1 RESEARCH INSTITUTE OF NATURAL SCIENCES, OKAYAMA UNIVERSITY OF SCIENCE, 1-1 RIDAI-CHO, KITA-KU, OKAYAMA 7000005, JAPAN 2 DEPARTMENT OF EARTH AND ENVIRONMENTAL SCIENCES, AND EARTH AND ENVIRONMENTAL SCIENCE SYSTEM RESEARCH CENTER, CHONBUK NATIONAL UNIVERSITY, 567 BACKJAEDARO, DUCKJIN-GU, JEONJU 54896, REPUBLIC OF KOREA 3 INSTITUTE OF PRECAMBRIAN GEOLOGY AND GEOCHRONOLOGY, RUSSIAN ACADEMY OF SCIENCES, 2, MAKAROVA, ST. PETERSBURG, 199034, RUSSIA 4 DIVISION OF EARTH AND ENVIRONMENTAL SCIENCE RESEARCH, KOREA BASIC SCIENCE INSTITUTE, 161, YEONGUDANJI, OCHANG-EUP, CHEONGWON-GU, CHEONGJU, 28119, OCHANG 34133, REPUBLIC OF KOREA 5 SCHOOL OF EARTH AND ENVIRONMENTAL SCIENCES, SEOUL NATIONAL UNIVERSITY, 1 GWANAK-RO, GWANAK-GU, SEOUL 08826, REPUBLIC OF KOREA ABSTRACT The Precambrian Salma eclogites on the Kola Peninsula, Russia, represent some of the oldest eclogites in the world; however, there has been much debate regarding whether the timing of their eclogite facies metamorphism is Archean (2.72–2.70 Ga) or Paleoproterozoic (1.92–1.88 Ga). New microstructural observations, pressure-temperature (P- T ) analyses, zircon inclusion analyses, and U-Pb zircon dating performed in this study suggest that eclogite facies metamorphism occurred at ca. 1.87 Ga under P-T conditions of 16–18 kbar and 750–770 °C. Metamorphic zircons with the age of 1.87 Ga have inclusions of garnet (Grt) + omphacite (Omp) + Ca-clinopyoxene (Cpx) + amphibole (Amp) + quartz (Qz) + rutile (Rt) ± biotite (Bt), as well as flat heavy rare earth element (HREE) patterns due to the presence of abundant amounts of garnet during peak eclogite facies metamorphism. The Paleoproterozoic ages (1.92–1.88 Ga) presented in previous studies are reinterpreted to represent prograde ages, rather than peak ages, because these ages have been inferred from U-Pb dating in zoisite-bearing zircon and Sm-Nd and Lu-Hf geochronologic analyses of garnet showing growth zoning. In contrast, the 2.73–2.72 Ga unzoned zircons with dark cathodolumines- cence contain inclusions of Grt + Amp + plagioclase (Pl) + Qz + rutile (Rt) ± Bt and are relatively enriched in HREEs, suggesting that an initial amphibolite facies metamorphic event occurred during the Archean. This study also proposes that the Salma eclogites underwent granu- lite facies retrograde metamorphism at 10–14 kbar and 770–820 °C, with rapid decompression occurring soon after peak metamorphism ca. 1.87 Ga. The final period of retrograde amphibolite facies metamorphism occurred at 8–10 kbar and 590–610 °C. Whole-rock chemical analyses indicate that the Salma eclogites were originally tholeiitic basalts formed at a mid-ocean ridge. The occurrence of eclogite facies metamorphism ca. 1.87 Ga suggests that the collision between the Kola and Karelian continents occurred during the Paleoproterozoic, rather than the Archean. These results, as well as those of previous studies, imply that the subduction required to form eclogites may have begun during or before the Paleoproterozoic. LITHOSPHERE GSA Data Repository Item 2017317 https://doi.org/10.1130/L657.1 INTRODUCTION One of the most important questions in Earth sciences involves the ini- tiation and evolution of subduction during the Precambrian (e.g., Cawood et al., 2006; van Hunen and Moyen, 2012, and references therein). Many researchers have inferred that subduction began during the Archean (e.g., Komiya et al., 1999; Brown, 2006, 2009; Cawood et al., 2006; Van Kranendonk et al., 2007), based on the presence of indicators of plate tectonics, such as accretionary prisms, orogens, and paired metamorphic belts. Brown (2006) suggested that the first appearance of Neoarchean high-pressure granulite reflects the initiation of subduction, which has a geothermal gradient higher than that observed in modern subduction characterized by blueschist. Blueschist facies metamorphism is believed to have begun during the Neoproterozoic (Maruyama et al., 1996; Stern, 2005; Tsujimori and Ernst, 2014). These occurrences indicate that an early style of subduction without blueschist may have dominated between the Neoarchean and the Neoproterozoic (Brown, 2006, 2009). However, the evolution of this early style of subduction prior to the Neoproterozoic remains unclear. Because eclogite is a typical high-pressure rock formed within subduc- tion zones, unraveling the metamorphic evolution of Precambrian eclogites with pre-Neoproterozoic ages is important in order to study the evolution of the early style of subduction occurring prior to the Neoproterozoic. Precambrian eclogites are rare worldwide. The Paleoproterozoic eclogites from Tanzania (2.0 Ga in the Usagaran belt; 1.89–1.86 Ga in the Ubendian belt) are well-known examples that are considered to represent remnants of the subducted Paleoproterozoic oceanic lithosphere (e.g., Möller et al., 1995; Collins et al., 2004; Boniface et al., 2012). However, the existence of Archean eclogites remains controversial. Mints et al. (2010, 2014) suggested that the eclogite in the Salma area of the Kola Peninsula is an Archean eclogite, based on the 2.87–2.82 Ga zircon age recorded in the eclogite. The 2.72–2.70, 2.4, and 1.9 Ga zircon ages obtained from the Salma eclogite were interpreted to be retrograde metamorphic ages (Mints et al., 2010, 2014). In contrast, Skublov et al. (2010a, 2011) suggested that © 2017 Geological Society of America. For permission to copy, contact [email protected]

Upload: others

Post on 27-May-2020

7 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Paleoproterozoic high-pressure metamorphic history of the …hosting03.snu.ac.kr/~hjung/pdf/Imayama_et_al-17.pdf · 2017-09-20 · Paleoproterozoic high-pressure metamorphic history

LITHOSPHERE | Volume 9 | Number 5 | www.gsapubs.org 1

RESEARCH

Paleoproterozoic high-pressure metamorphic history of the Salma eclogite on the Kola Peninsula, Russia

Takeshi Imayama1, Chang-Whan Oh2, Shauket K. Baltybaev3, Chan-Soo Park4, Keewook Yi4, and Haemyeong Jung5

1RESEARCH INSTITUTE OF NATURAL SCIENCES, OKAYAMA UNIVERSITY OF SCIENCE, 1-1 RIDAI-CHO, KITA-KU, OKAYAMA 7000005, JAPAN2DEPARTMENT OF EARTH AND ENVIRONMENTAL SCIENCES, AND EARTH AND ENVIRONMENTAL SCIENCE SYSTEM RESEARCH CENTER, CHONBUK NATIONAL UNIVERSITY, 567 BACKJAEDARO, DUCKJIN-GU, JEONJU 54896, REPUBLIC OF KOREA

3INSTITUTE OF PRECAMBRIAN GEOLOGY AND GEOCHRONOLOGY, RUSSIAN ACADEMY OF SCIENCES, 2, MAKAROVA, ST. PETERSBURG, 199034, RUSSIA4DIVISION OF EARTH AND ENVIRONMENTAL SCIENCE RESEARCH, KOREA BASIC SCIENCE INSTITUTE, 161, YEONGUDANJI, OCHANG-EUP, CHEONGWON-GU, CHEONGJU, 28119, OCHANG 34133, REPUBLIC OF KOREA

5SCHOOL OF EARTH AND ENVIRONMENTAL SCIENCES, SEOUL NATIONAL UNIVERSITY, 1 GWANAK-RO, GWANAK-GU, SEOUL 08826, REPUBLIC OF KOREA

ABSTRACT

The Precambrian Salma eclogites on the Kola Peninsula, Russia, represent some of the oldest eclogites in the world; however, there has been much debate regarding whether the timing of their eclogite facies metamorphism is Archean (2.72–2.70 Ga) or Paleoproterozoic (1.92–1.88 Ga). New microstructural observations, pressure-temperature (P-T ) analyses, zircon inclusion analyses, and U-Pb zircon dating performed in this study suggest that eclogite facies metamorphism occurred at ca. 1.87 Ga under P-T conditions of 16–18 kbar and 750–770 °C. Metamorphic zircons with the age of 1.87 Ga have inclusions of garnet (Grt) + omphacite (Omp) + Ca-clinopyoxene (Cpx) + amphibole (Amp) + quartz (Qz) + rutile (Rt) ± biotite (Bt), as well as flat heavy rare earth element (HREE) patterns due to the presence of abundant amounts of garnet during peak eclogite facies metamorphism. The Paleoproterozoic ages (1.92–1.88 Ga) presented in previous studies are reinterpreted to represent prograde ages, rather than peak ages, because these ages have been inferred from U-Pb dating in zoisite-bearing zircon and Sm-Nd and Lu-Hf geochronologic analyses of garnet showing growth zoning. In contrast, the 2.73–2.72 Ga unzoned zircons with dark cathodolumines-cence contain inclusions of Grt + Amp + plagioclase (Pl) + Qz + rutile (Rt) ± Bt and are relatively enriched in HREEs, suggesting that an initial amphibolite facies metamorphic event occurred during the Archean. This study also proposes that the Salma eclogites underwent granu-lite facies retrograde metamorphism at 10–14 kbar and 770–820 °C, with rapid decompression occurring soon after peak metamorphism ca. 1.87 Ga. The final period of retrograde amphibolite facies metamorphism occurred at 8–10 kbar and 590–610 °C. Whole-rock chemical analyses indicate that the Salma eclogites were originally tholeiitic basalts formed at a mid-ocean ridge. The occurrence of eclogite facies metamorphism ca. 1.87 Ga suggests that the collision between the Kola and Karelian continents occurred during the Paleoproterozoic, rather than the Archean. These results, as well as those of previous studies, imply that the subduction required to form eclogites may have begun during or before the Paleoproterozoic.

LITHOSPHERE GSA Data Repository Item 2017317 https://doi.org/10.1130/L657.1

INTRODUCTION

One of the most important questions in Earth sciences involves the ini-tiation and evolution of subduction during the Precambrian (e.g., Cawood et al., 2006; van Hunen and Moyen, 2012, and references therein). Many researchers have inferred that subduction began during the Archean (e.g., Komiya et al., 1999; Brown, 2006, 2009; Cawood et al., 2006; Van Kranendonk et al., 2007), based on the presence of indicators of plate tectonics, such as accretionary prisms, orogens, and paired metamorphic belts. Brown (2006) suggested that the first appearance of Neoarchean high-pressure granulite reflects the initiation of subduction, which has a geothermal gradient higher than that observed in modern subduction characterized by blueschist. Blueschist facies metamorphism is believed to have begun during the Neoproterozoic (Maruyama et al., 1996; Stern, 2005; Tsujimori and Ernst, 2014). These occurrences indicate that an early style of subduction without blueschist may have dominated between the Neoarchean and the Neoproterozoic (Brown, 2006, 2009). However, the

evolution of this early style of subduction prior to the Neoproterozoic remains unclear.

Because eclogite is a typical high-pressure rock formed within subduc-tion zones, unraveling the metamorphic evolution of Precambrian eclogites with pre-Neoproterozoic ages is important in order to study the evolution of the early style of subduction occurring prior to the Neoproterozoic. Precambrian eclogites are rare worldwide. The Paleoproterozoic eclogites from Tanzania (2.0 Ga in the Usagaran belt; 1.89–1.86 Ga in the Ubendian belt) are well-known examples that are considered to represent remnants of the subducted Paleoproterozoic oceanic lithosphere (e.g., Möller et al., 1995; Collins et al., 2004; Boniface et al., 2012). However, the existence of Archean eclogites remains controversial. Mints et al. (2010, 2014) suggested that the eclogite in the Salma area of the Kola Peninsula is an Archean eclogite, based on the 2.87–2.82 Ga zircon age recorded in the eclogite. The 2.72–2.70, 2.4, and 1.9 Ga zircon ages obtained from the Salma eclogite were interpreted to be retrograde metamorphic ages (Mints et al., 2010, 2014). In contrast, Skublov et al. (2010a, 2011) suggested that

© 2017 Geological Society of America. For permission to copy, contact [email protected]

Page 2: Paleoproterozoic high-pressure metamorphic history of the …hosting03.snu.ac.kr/~hjung/pdf/Imayama_et_al-17.pdf · 2017-09-20 · Paleoproterozoic high-pressure metamorphic history

IMAYAMA ET AL.

2 www.gsapubs.org | Volume 9 | Number 5 | LITHOSPHERE

the Salma eclogite is a Paleoproterozoic eclogite, based on the presence of metamorphic zircons recording ages of ca. 1.92–1.88 Ga, low Th/U ratios, and flat heavy rare earth element (HREE) patterns. Lu-Hf and Sm-Grt ages (1.90–1.88 Ga) obtained from eclogite and eclogitized ultrabasite were also interpreted to reflect eclogite facies metamorphic ages (Skublov et al., 2010b; Herwartz et al., 2012; Mel’nik et al., 2013). However, garnet from massive eclogite (sample 46, Table 1) exhibits prograde zoning (Skublov et al., 2011), implying that these ages may represent prograde metamorphic ages rather than ages of peak metamorphism. These previous studies pre-sented no direct evidence with which to determine which zircons formed during eclogite facies metamorphism, such as omphacite inclusions in zircon. Therefore, the timing of the eclogite facies metamorphism of the Salma eclogite is still uncertain, and it is necessary to confirm whether the Salma eclogite is Archean or Paleoproterozoic in age based on direct evidence. In addition, the study of the pressure-temperature (P-T) condi-tions of eclogites in the Salma area is necessary, because only minimum pressure conditions have been determined using geothermobarometry (Mints et al., 2010, 2014; Shchipansky et al., 2012). Determining the age and petrogenesis of the Salma eclogite in the Kola Peninsula is thus essential for understanding Precambrian geodynamics and the tectonic evolution of the Fennoscandian shield.

In this paper we provide direct evidence for Paleoproterozoic eclogite based on zircon U-Pb dating coupled with analyses of REEs and inclu-sions in zircon. The P-T conditions of the different metamorphic stages of the Salma eclogite were estimated using conventional geothermobarom-etry and pseudosection modeling. In addition, whole-rock chemistry was analyzed to characterize the tectonic setting in which these eclogites

originally formed. We suggest that Paleoproterozoic subduction zones were relatively warmer than Phanerozoic subduction zones but colder than Neoarchean subduction zones.

GEOLOGICAL BACKGROUND

The Fennoscandian shield records a general trend in which the age of geological activity decreases toward the southwest. The northern part of the shield is dominated by Archean rocks, whereas the major part of the shield comprises the Paleoproterozoic 2.0–1.8 Ga Svecofennian Province and the 1.8–1.65 Ga Transscandinavian Igneous Belt. The 1.2–0.9 Ga Sveconorwegian Province is farther to the southwest (Daly et al., 2006; Fig. 1A). In the northern part of the shield, the Kola-Karelian orogen is located between the Kola and Karelian cratons. The Kola-Karelian orogen mainly consists of three Paleoproterozoic tectonic belts (the Kola suture belt, the Tanaelv belt, and the Lapland and Umbra granulite belts), the Neo-archean Inari microcontinent, and the Belomorian mobile belt (Fig. 1B).

The Belomorian mobile belt is principally composed of 2.9–2.6 Ga tonalite-trondhjemite-granodiorite (TTG) gneisses (Hölttä et al., 2008; Mints et al., 2014) and includes a ca. 2.9 Ga paragneiss complex and 2.9–2.8 Ga greenstone belts (Slabunov et al., 2006). The available geologi-cal, isotopic, and geochemical data from the mafic-ultramafic rocks of the greenstone complex are compatible with their interpretation as the tectoni-cally disrupted and metamorphosed remnants of a Mesoarchaean ophiolitic association (Slabunov et al., 2006). This belt underwent multiple deforma-tion and metamorphic events during both the Archean and Paleoprotero-zoic (Daly et al., 2001, 2006; Mints et al., 2014). The Paleoproterozoic

TABLE 1. AGE CONSTRAINTS AND INTERPRETATIONS FROM THE SALMA ECLOGITES AND RELATED ROCKS

Lithology (sample identification) Method Age (Ma) Interpretation of source given References

Eclogite (S-198/107) U-Pb Zrn 2703 ± 9 Retrograde granulite facies metamorphism Mints et al. (2010)

Fe-Ti eclogite (S-204-2B) U-Pb Zrn 28201913

Magmatic protolith agePartially reset age

Mints et al. (2010)

Garnetite (S-204-23B) U-Pb Zrn 1891 ±17 Metamorphic event Mints et al. (2010)Plagiogranite vein (S-204-28) U-Pb Zrn 2866 ± 10

2781 ± 15Eclogite facies metamorphism to ca. 2.87 Ga or older Mints et al. (2010)

Eclogite (S-198/107) U-Pb Zrn 2724 ± 35 Granulite facies metamorphism Kaulina et al. (2010)Eclogite (Ex198) U-Pb Zrn 2917 ± 360

2939 ± 811820 ± 180

Magmatic protolith ageMagmatic protolith age

Metamorphic event

Kaulina et al. (2010)

Garnetite (S-204-23) U-Pb Zrn 1891 ± 17 Metamorphic event Kaulina et al. (2010)Plagiogranite (S-204-28) U-Pb Zrn 2866 ± 36

2778 ± 231874 ± 29

Magmatic protolith ageMagmatic protolith age

Metamorphic event

Kaulina et al. (2010)

Massive eclogite (sample 46) U-Pb Zrn 2865 ± 35(2879 ± 34)1923 ± 75

(1878 ± 36)

Magmatic protolith eventEclogite facies metamorphism

Skublov et al. (2010a) Skublov et al. (2011)

Eclogitized ultrabasic (sample 21) U-Pb Zrn 1907 ± 11 Eclogite facies metamorphism Skublov et al. (2010a)Pegmatite vein (sample 62) U-Pb Zrn 1841 ± 12 Retrograde amphibolite facies metamorphism Skublov et al. (2010a)Massive eclogite (sample 46) Lu-Hf Grt 1901 ± 5 Eclogite facies metamorphism Herwartz et al. (2012)Eclogitized ultrabasite (sample 21) Lu-Hf Grt 1894 ± 4 Eclogite facies metamorphism Herwartz et al. (2012)Massive eclogite (sample 46) Sm-Nd Grt 1897 ± 16 Eclogite facies metamorphism Mel’nik et al. (2013)Garnetites (sample 48) U-Pb Zrn

Sm-Nd Grt2864 ± 431927 ± 501887 ± 19

(1839 ± 11)

Magmatic protolith ageMetamorphic eventMetamorphic event

Mel’nik et al. (2013)

Massive eclogite (sample 46) Sm-Nd Grt 1789 ± 23 Retrograde metamorphic event Skublov et al. (2010b)Eclogitized ultrabasic (sample 21) Sm-Nd Grt 1878 ± 12 Metamorphic event Skublov et al. (2010b)Eclogite (RPB1B) U-Pb Zrn 2716 ± 10

1865 ± 15Amphibolite facies metamorphism

Eclogite facies metamorphismThis study

Eclogite (RPB3A) U-Pb Zrn 2727 ± 81868 ± 171720 ± 79

Amphibolite facies metamorphismEclogite facies metamorphism

Retrograde amphibolite facies metamorphism

This study

Note: Ages in parentheses are data from Skublov et al. (2011). Zrn—zircon; Grt—garnet.

Page 3: Paleoproterozoic high-pressure metamorphic history of the …hosting03.snu.ac.kr/~hjung/pdf/Imayama_et_al-17.pdf · 2017-09-20 · Paleoproterozoic high-pressure metamorphic history

LITHOSPHERE | Volume 9 | Number 5 | www.gsapubs.org 3

Paleoproterozoic high-pressure metamorphism of the Salma eclogite | RESEARCH

reworking was associated with collisional events, resulting in granulite facies metamorphism (Balagansky et al., 2014, and references therein).

In the Belomorian mobile belt, several eclogite exposures occur (Fig. 1) in the Salma area (Kaulina et al., 2010: Mints et al., 2010; Skublov et al., 2010a, 2010b, 2011), in the Kuru-Vaara quarry (Shchipansky et al., 2012; Balagansky et al., 2014), and in the Gridino area in the northeastern region of Karelia (Volodichev et al., 2004, 2012; Dokukina et al., 2014). The Salma and Kuru-Vaara eclogites are related to the high-pressure metamor-phism of oceanic lithosphere and exhibit peak P-T conditions of >13–14 kbar and 700–750 °C (e.g., Mints et al., 2010, 2014; Shchipansky et al., 2012). The Gridino eclogites were originally gabbroic dike swarms that intruded felsic gneiss and were metamorphosed under peak P-T condi-tions of 17–18 kbar (probably to 22 kbar) and 740–865 °C (Volodichev et al., 2004; Dokukina et al., 2014).

In this study we focus on eclogite and granulite samples collected from two outcrops in the Salma area of the Belomorian mobile belt, which is located near the Tanaelv belt (Fig. 1B). In the Salma area, mafic bodies (including eclogite) occur within TTG gneisses, which mainly consist of quartz diorite and trondhjemite. In the mafic bodies, eclogites occur as layers that are intercalated with garnet granulite, garnet amphibolite, amphibolite, and garnetite (Fig. 2A). Although some eclogite layers are fresh and coarse grained, most eclogite layers are medium grained and have retrograded into granulite and amphibolite (Figs. 2B–2D). In the coarse-grained eclogites, garnets occur within a bright green matrix, which mainly consists of clino-pyroxene (omphacite and calcic clinopyroxene) with minor amphibole (Fig. 2C). In the medium-grained eclogite, less garnet occurs within a matrix that is dark green, due to the presence of abundant amphiboles that formed during retrograde metamorphism (Figs. 2B, 2D). The interlayered granu-lite and amphibolite can be considered to represent strongly retrograded eclogite. During retrograde metamorphism, omphacite was first retrograded into calcic clinopyroxene, thus forming symplectite around garnet and in the matrix; then, clinopyroxene was retrograded to amphibole. Mints et al.

(2014) inferred that the protolith of the layered mafic body was originally a series of normal gabbro norite, olivine gabbro, and Fe-Ti oxide gabbro intercalated with local troctolite, which resembled the gabbroic suite from the modern oceanic crust of the slow-spreading Southwest Indian Ridge (Dick et al., 2000). The different degrees of retrograde metamorphism may be due to compositional differences or varying amounts of water infiltration. The retrograde eclogite includes a few leucosomes (Figs. 2B, 2D) consist-ing of quartz, plagioclase, K-feldspar, and relict garnet, thus indicating that partial melting occurred during metamorphism.

PETROGRAPHY AND MINERAL CHEMISTRY

The mineral compositions of the five samples were analyzed using a Shimadzu electron probe microanalyzer (EPMA-1600) at the Jeonju branch of the Korea Basic Science Institute (South Korea). Multiple eclogite sam-ples (RPB3A, RPB1A, and RPB1B) and garnet-clinopyroxene granulite samples (RPB3B and RPB3C) were collected. The operating conditions used for the analyses included an accelerating voltage of 15 kV and a beam current of 20 nA. The probe diameter for the mineral composition spot analyses was 3 µm. Natural and synthetic silicates and oxides were used as standards. The ZAF method was employed for matrix correction. Rep-resentative mineral compositions are listed in Data Repository Table DR11.

Retrograded Eclogite

Sample RPB3A is a retrograde eclogite mainly consisting of gar-net, clinopyroxene, amphibole, biotite, plagioclase, quartz, and rutile. Garnet is characterized by an inclusion-rich core and an inclusion-free

1 GSA Data Repository Item 2017317, which includes four tables and one figure, is available at http://www.geosociety.org /datarepository /2017, or on request from [email protected].

KolaCraton

Kola-KareliaOrogen

BalticSea

Sveconorwegian

Belt

WhiteSea

Exposed Archean basement

A

Paleoproterozoic accretional orogenMesoproterozoic orogenPaleozoic orogen

Reworked Archean complex

SvecofennianOrogen

Transscandinavian Igneous belt

Caled

onia

n Bel

t Karelia Craton

Paleoproterozoic collisional orogen

Paleoproterozoic igneous belt

▲▲

▲▲

▲▲

▲▲

▲ ▲▲▲

▲▲

▲ ▲▲

Paleoproterozoic granulite belts

0 100km

Barents Sea

36o

69o

36o30o

Belomorian Mobile belt

Murmansk terrane

Kola Province terrane

Lapland belt

Umba belt

Kola suture belt

White Sea

Tanaelv belt

Inari unit

Tectonic melange

Meso-Neoarchean TTG gneiss & greenstone belt

Paleoproterozoic volcanics & sediments

Neoarchan microcontinent

Reworked Archean TTG gneiss & greenstone belt

Paleoproterozoic post-orogenic granites

B

Eclogite samples

▲●

SalmaKuru-Vaara

●▲

GridinoReworked Archean TTG gneiss & greenstone belt

3A, 3B, 3C

1A, 1B

69o

67o

Figure 1. (A) Simplified geological map of the Fennoscandian shield showing the locations of the eclogites in the Belomorian mobile belt. (B) The tectonic division of the northeastern Fennoscandian shield region. Sample locations: 3A, 3B, and 3C: N67°28′32″/E32°22′39″; samples 1A and 1B: N67°31′07″/E32°21′03″. Modified from Berthelsen and Marker (1986), Zhao et al. (2002), and Daly et al. (2006). TTG—tonalite-trondhjemite-granodiorite.

Page 4: Paleoproterozoic high-pressure metamorphic history of the …hosting03.snu.ac.kr/~hjung/pdf/Imayama_et_al-17.pdf · 2017-09-20 · Paleoproterozoic high-pressure metamorphic history

IMAYAMA ET AL.

4 www.gsapubs.org | Volume 9 | Number 5 | LITHOSPHERE

rim (Fig. 3A). The main inclusions in the garnet core are amphibole, clinozoisite, quartz, and rutile (Figs. 3B, 3C), with minor biotite. Most clinopyroxene crystals in the matrix are symplectic with plagioclase (Fig. 3D) and are calcic clinopyroxene (commonly diopside and minor augite). Omphacite is found as relics (Fig. 3D). Omphacite records jadeite contents of as much as 21%, whereas the symplectic diopside records low jadeite contents (3%–10%; Fig. 4A). Garnet is characterized by homogeneous cores [almandine (Alm

41–44), pyrope (Prp

33–34), grossular (Grs

20–26), spes-

sartine (Sps1)] and Mg-rich inner rims (Alm

39–40Prp

36–39Grs

22–23Sps

1) and

outer rims (Alm41–42

Prp35

Grs22–23

Sps1; Fig. 5A). A decrease in X

Fe from 0.54

to 0.55 to 0.50–0.52 is observed toward the inner rim, and XFe

increases to 0.54 at the outer rim. All amphiboles within garnets and the matrix plot within the compositional field of pargasite (Fig. 4B) with an X

Fe ratio of

0.08–0.26. Plagioclase has an anorthite content (Xan

) of 0.21–0.35. Cli-nozoisite records low Fe3+/(Fe3+ + Al) ratios (0.02–0.03).

Samples RPB1A and RPB1B are also retrogressed eclogites that are relatively dark in color and consist of garnet, clinopyroxene, amphibole, plagioclase, quartz, and rutile with minor biotite. The retrogressed eclog-ites include many subhedral garnet porphyroblasts ranging in size from 3 to 10 mm. The main inclusions in the garnet are amphibole, clinopyroxene, quartz, and rutile. Subhedral clinopyroxene is dominant within the matrix, and plagioclase lamellae were found within the clinopyroxene in sample RPB1A (Fig. 3E). The symplectites of calcic clinopyroxene + plagioclase

and amphibole + plagioclase formed around garnet due to the breakdown of omphacite during granulite facies metamorphism (Fig. 3F). In contrast, the formation of coronitic plagioclase is related to the replacement of garnet rims during amphibolite facies metamorphism (Fig. 3F). In sample RPB1A, the jadeite contents of the clinopyroxene in the garnet and matrix range from 11% to 18% and 15% to 21%, respectively, whereas the jadeite content of symplectic clinopyroxene is 9% (Fig. 4C). In sample RPB1B, clinopyroxene in garnet records jadeite contents ranging between 14% and 27%, whereas the clinopyroxene in the symplectite records jadeite contents ranging between 5% and 10% (Fig. 4C). The garnets in both samples have homogeneous cores (Alm

41–43Prp

36–38Grs

18–21Sps

1–2 with X

Fe

values of 0.52–0.54 in sample RPB1A; Alm44–46

Prp29–32

Grs23–25

Sps1 with

XFe

values of 0.59–0.61 in sample RPB1B) and relatively Fe-rich outer-most rims (with X

Fe ratios of 0.57–0.58 in sample RPB1A and X

Fe ratios of

0.63–0.64 in sample RPB1B) (Fig. 5B). The XFe

ratio in garnet increases toward the rim, thus reflecting retrograde metamorphism (Fig. 5B). The amphibole inclusions in garnet are magnesiohornblende with X

Fe contents

of 0.11–0.12 in sample RPB1A and pargasite with XFe

values of 0.23–0.29 in sample RPB1B (Fig. 4D). In contrast, amphibole in contact with garnet records a wide compositional range, from magnesiohornblende to par-gasite, with X

Fe contents of 0.26–0.38 in sample RPB1A and 0.21–0.29

in sample RPB1B (Fig. 4D). The Xan

values of plagioclase are 0.18–0.31 in sample RPB1A and 0.23–0.30 in sample RPB1B.

amphiboliteRetrogradedeclogite

Granulizedeclogite

BA

C D

Figure 2. Outcrop photographs of the Salma eclogites from the Kola Peninsula, Russia. (A) Eclogite outcrop in which layered-type eclogite is interlayered with amphibolite and granulite. (B) Eclogite that has retrograded into granulite or amphibolite, with relict eclogite. (C) Rather fresh eclogite consisting of garnet and omphacite with minor amphibolite. (D) Thin leucosome in eclogite.

Page 5: Paleoproterozoic high-pressure metamorphic history of the …hosting03.snu.ac.kr/~hjung/pdf/Imayama_et_al-17.pdf · 2017-09-20 · Paleoproterozoic high-pressure metamorphic history

LITHOSPHERE | Volume 9 | Number 5 | www.gsapubs.org 5

Paleoproterozoic high-pressure metamorphism of the Salma eclogite | RESEARCH

Garnet-Clinopyroxene Granulite

Samples RPB3B and RPB3C are garnet-clinopyroxene granulites consisting of garnet, clinopyroxene, amphibole, plagioclase, and quartz. Clinopyroxene crystals in the 2 granulites are diopside with jadeite con-tents of <5% (Fig. 4A). In sample RPB3B, the composition of the garnet core is Alm

39–42Prp

31–33Grs

25–27Sps

1; the almandine component increases

at the rim (Alm44–48

Prp23

-29

Grs25–27

Sps2), thus exhibiting retrograde zoning.

The garnet in sample RPB3C also exhibits retrograde zoning. The garnet core has a composition of Alm

42–43Prp

27–30Grs

26–28Sps

1, whereas the gar-

net rim has an Fe-rich composition of Alm46–49

Prp23–24

Grs26–29

Sps1–2

. The values of X

Fe increase toward the rim from 0.59 to 0.62 and to 0.66–0.68

in RPB3B and RPB3C. The amphiboles in sample RPB3B plot on the boundary between the compositional fields of magnesiohornblende and pargasite (Fig. 4B). The X

Fe values of amphibole in garnet (0.11–0.22) are

slightly lower than those in amphiboles in contact with garnet (0.26–0.38). The amphibole in sample RPB3C is magnesiohornblende (Fig. 4B) with X

Fe values of 0.19–0.30. The X

an values of plagioclase in samples RPB3B

and RPB3C are 0.43–0.65 and 0.33–0.36, respectively.

P-T ESTIMATES

Metamorphic Stages Based on Petrography

On the basis of microstructural observations and mineral relationships, several metamorphic stages have been recognized from the Salma eclog-ites. (1) Evidence of prograde metamorphism is preserved within garnet in sample RPB3A (Fig. 3A). The prograde assemblage is clinozoisite (Czo) + amphibole (Amp) + garnet (Grt) + biotite (Bt) + quartz (Qz) + rutile (Rt) (Whitney and Evans, 2010). (2) The omphacite in garnet in

0.5 mm

RelictOmp

Ca-Cpx+ Pl

Amp

Grt

Grt

Omp

Ca-Cpx+ Pl

Amp+ Pl

Qz

400 μm

1.0 mm

-

BtAmp

Grt

Amp

CoroniticPl

Bt

Bt

Amp

Czo

Ca-Cpx

Pl lamellae

Grt

Czo

Czo

Qz

Grt

Czo

Amp

Fig.3b

Fig.3c

Qz

Qz

Grt

Amp+Pl

100 μm 150 μm

100 μm

Rt

B C

E F

A

D

Figure 3. Photomicrographs of the Salma eclogites. (A) Inclusion-rich core and inclusion-free rim in garnet in sample RPB3A. (B) Backscattered elec-tron (BSE) image of clinozoisite and quartz in garnet in sample RPB3A. (C) BSE image of amphibole, rutile, clinozoisite, and quartz in garnet in sample RPB3A. (D) Plane-polarized light photomicrograph of symplectite of Ca-clinopyroxene + plagioclase and relict omphacite in sample RPB3A. (E) BSE image of plagioclase lamellae in Ca-clinopyroxene in sample RPB1A. (F) BSE image of omphacite inclusion in garnet and Ca-clinopyroxene + plagio-clase and amphibole + plagioclase symplectites surrounding garnet porphyroblasts in sample RPB1B. Abbreviations: Amp—amphibole, Grt—garnet, Czo—clinozoisite, Cpx—clinopyroxene, Omp—omphacite, Pl—plagioclase, Rt—rutile, Qz—quartz.

Page 6: Paleoproterozoic high-pressure metamorphic history of the …hosting03.snu.ac.kr/~hjung/pdf/Imayama_et_al-17.pdf · 2017-09-20 · Paleoproterozoic high-pressure metamorphic history

IMAYAMA ET AL.

6 www.gsapubs.org | Volume 9 | Number 5 | LITHOSPHERE

sample RPB1B (Fig. 3F) and the relict omphacite in the matrix of sample RPB3A (Fig. 3D) formed during eclogite facies metamorphism. The Mg-rich garnet rim in sample RPB3A and the omphacite-bearing garnet core in sample RPB1B grew during this stage. The omphacite-bearing garnets from sample RPB1B also include amphibole, quartz, and rutile. The leu-cosome associated with eclogite within the outcrop (Fig. 2D) indicates that melt existed during metamorphism. The inferred peak assemblage consists of omphacite (Omp) + Amp + Grt + Qz + Rt + melt ± Bt. (3) The first overprint resulted in the growth of plagioclase and Ca-clinopyroxene,

producing the mineral assemblage Ca-Cpx + Amp + Grt + Qz + Rt +Pl + melt ± Bt. The symplectite of Ca-clinopyroxene + plagioclase observed in all samples (e.g., Figs. 3D, 3F), as well as the plagioclase lamellae observed in the Ca-clinopyroxene of sample RPB1A (Fig. 3E), developed during this stage. (4) The secondary amphiboles replacing garnet rims (Figs. 3A, 3F) and surrounding the symplectites (Fig. 3D) represent an amphibolite facies overprint that formed during cooling. The coronitic plagioclase surrounding the garnet rims also formed during this stage, along with secondary amphibole.

20 20

40 40

WoEnFs

omphacite aegirine-augite

Jd AegRPB1A

In grt

Symplectite

Matrix

RPB1B

CA

20 20

40omphacite aegirine-augite

WoEnFs

Jd Aeg

In grtSymplectite

Relict

RPB3A RPB3B RPB3C

40

D

00

0.5

1

(Na+

K +

2C

a)

in A

site

0.5 1 1.5 2(Al + Fe3+ + 2Ti) in C site

Edenite

Tremolite

Pargasite

Mgnesio- Hornblend

Tschermakite

Sadanagaite

B

0

0.5

1

0 0.5 1 1.5 2

(Na+

K +

2C

a)

in A

site

(Al + Fe3+ + 2Ti) in C site

Edenite

Tremolite

Pargasite

Mgnesio- Hornblend

Tschermakite

Sadanagaite

In grtcontactmatrix

RPB3A RPB3B RPB3C

RPB1AIn grtcontact

RPB1B

Figure 4. Clinopyroxene compositions plotted on the wollastonite + enstatite + ferrosilite (WoEnFs)–jadeite–aegirine (Jd-Aeg) diagram of Morimoto (1988) (grt—garnet) for (A) samples RPB3A–RPB3C and (B) samples RPB1A, RPB1B. Classification of Ca-amphiboles according to Hawthorne et al. (2012) based on a (Na + K + 2Ca) in A site (Al + Fe+3 + 2Ti) in C site diagram for (C) samples RPB3A–RPB3C and (D) samples RPB1A, RPB1B.

Page 7: Paleoproterozoic high-pressure metamorphic history of the …hosting03.snu.ac.kr/~hjung/pdf/Imayama_et_al-17.pdf · 2017-09-20 · Paleoproterozoic high-pressure metamorphic history

LITHOSPHERE | Volume 9 | Number 5 | www.gsapubs.org 7

Paleoproterozoic high-pressure metamorphism of the Salma eclogite | RESEARCH

Conventional Geothermobarometry

The P-T conditions of the different metamorphic stages of the metaba-sites were estimated using the garnet-clinopyroxene and garnet-amphibole geothermometers (Ellis and Green, 1979; Graham and Powell, 1984) and the garnet-clinopyroxene-plagioclase-quartz and garnet-amphibole-plagioclase-quartz geobarometers (Powell and Holland, 1988; Kohn and Spear, 1990). The P-T results obtained using conventional thermobarom-etry are listed in Table 2.

The homogeneous garnet cores from sample RPB3A include clinozo-isite; their compositions are relatively Mg poor compared to their inner rim compositions. These data imply that the homogeneous cores formed during the prograde stage. The metamorphic temperatures estimated using the compositions of the amphibole inclusions in garnet and the adjacent garnet cores in sample RPB3A are ~610–660 °C, which represent the P-T conditions during prograde metamorphism. However, the uncertainties of

the estimated temperatures may be high, due to the possibility of com-positional changes occurring during peak or retrograde metamorphism. Sample RPB3A records Mg-rich inner rims in garnet and relict ompha-cite, and the garnet core in sample RPB1B includes omphacite. Thus, we infer that these garnet and omphacite compositions are related to eclogite facies metamorphism. They yield peak eclogite-stage temperatures rang-ing from 730 to 810 °C, assuming a pressure of 17 kbar, as obtained using compositional isopleths in the pseudosection (see later section). The P-T conditions of the first overprint (11.5–12.5 kbar and 770 °C), which indi-cate granulite facies metamorphism, were inferred from the compositions of garnet rims and the Ca-clinopyroxene and plagioclase located near the garnet in sample RPB1A. The retrograde P-T conditions of amphibolite facies metamorphism (8.0–10.0 kbar and 590–610 °C) were calculated using the compositions of garnet rims and the plagioclase and amphibole surrounding the garnet in samples RPB3A, RPB1A, and RPB1B.

P-T Pseudosection

The P-T pseudosections for samples RPB3A and RPB1B were cal-culated in the modal chemical system NCKFMASHTO (Na

2O-CaO-

K2O-FeO-MgO-Al

2O

3-SiO

2-H

2O-TiO

2-O

2) using the Perplex X program

(Connolly, 1990) with an internally consistent thermodynamic data set (Holland and Powell, 1998; updated in 2002). The following solid-solution models were used in these calculations: garnet (White et al., 2000), biotite (Tajcmanová et al., 2009), plagioclase (Newton et al., 1980), K-feldspar (Waldbaum and Thompson, 1968), clinopyroxene and orthopyroxene (Holland and Powell, 1996), phengite (parameters from thermodynamic dataset of Powell and Holland, 1988), amphibole (Dale et al., 2005), and melt (White et al., 2001). All fluid was assumed to be H

2O; its content

was obtained from the values of weight loss on ignition. The ferrous/ferric (Fe2+/Fe3+) ratio was determined by using the titration of FeO to calculate the O

2 component. The haplogranitic melt of White et al. (2001) may not

always be appropriate for modeling partial melting in metabasites. As a result, it is possible that the assemblages containing melt in the calculated pseudosections may be metastable. However, the topology of the phase relationship between amphibolite facies and granulite facies metabasites does not significantly change when mineral assemblages coexist with fluids or melts (Pattison, 2003). The pseudosection approach is useful for inferring mineral assemblages including melt for the Salma eclogites. The bulk-rock compositions used in the pseudosection calculations were analyzed using inductively coupled plasma–mass spectrometry (ICP-MS; Perkin Elmer Optima 3000) at Activation Laboratories Ltd. (Canada).

The effective bulk composition was possibly modified by the growth of zoned garnet due to crystal fractionation (e.g., Evans, 2004). Because garnet in sample RPB3A displays prograde zoning with homogeneous cores and Mg-rich inner rims, its effective bulk composition was calcu-lated. First, the pseudosection used to estimate prograde P-T conditions was constructed using the bulk composition determined from the ICP-MS analyses (Fig. 6). Second, the modal percentage of garnet cores within the rock was estimated (~10 vol%), and the composition of the garnet cores was subtracted from the bulk chemical data (Konrad-Schmolke et al., 2008). The recalculated bulk composition was used for the pseu-dosection in order to estimate the peak and retrograde P-T conditions (Fig. 7). The molar proportion of the unfractionated and effective bulk rock composition used for the pseudosection modeling is shown in the captions for Figures 6 and 7.

The P-T pseudosection constructed using the actual measured bulk composition indicates that the prograde assemblage of zoisite (Czo) + Amp + Grt + Bt + Qz + Rt + H

2O occurs at P-T conditions of 13–18 kbar

and 640–720 °C (Fig. 6). This P-T range can be considered reasonable

0.30 

0.40 

0.50 

0.60 

0.70 

Xalm

Xsps

Xpyp

0.00 

0.10 

0.20 

0.0  0.5  1.0  1.5  2.0  2.5  3.0  3.5 

Xgrs

XFe

B

Distance (mm)

B’RPB1B

0.00

0.10

0.20

0.30

0.40

0.50

0.60A A’

Distance (mm)

RPB3A

Xalm

Xsps

Xpyp

Xgrs

XFe

B

A

0.0 0.5 1.0 1.5 2.0  2.5  3.0  3.5 4.0  4.5

Figure 5. Representative compositional zoning profiles of garnet in (A) sample RPB3A and (B) sample RPB1B. Abbreviations: alm—alman-dine; pyp—pyrope; grs—grossular; sps—spessartine.

Page 8: Paleoproterozoic high-pressure metamorphic history of the …hosting03.snu.ac.kr/~hjung/pdf/Imayama_et_al-17.pdf · 2017-09-20 · Paleoproterozoic high-pressure metamorphic history

IMAYAMA ET AL.

8 www.gsapubs.org | Volume 9 | Number 5 | LITHOSPHERE

8

11

14

17

20

23

Pre

ssur

e (k

bar)

Temperature (oC) 660 720 780 840600 900

Bt Cpx Melt Pl Grt

Bt Pl Grt Qz H2O

Bt Pl Grt Opx Qz H2O Bt Melt Pl Grt Qz H2O

Bt Melt Pl Grt Opx Qz

Bt Melt Pl Grt Opx

Bt Melt Pl Ilm Grt Opx (-Rt)

Bt Melt Pl

Ilm Grt Opx

Bt Cpx Melt Pl Ilm Grt Opx

Bt Cpx Melt Pl Ilm Grt Opx (-Rt)

Chl Ph Grt Zo

Bt Chl Ph Grt Zo

Ph Grt Zo H2O

Cpx Ph Grt Ky Zo H2O

Bt Cpx Grt Ky QzH2O

Bt Cpx Melt Grt H2O (-Amp)

Bt Cpx Melt Grt (-Amp)

Bt Chl Ph Grt Zo Qz

Bt Cpx Chl Ph Grt Zo

Cpx Chl Ph Grt Zo H2O

Chl Ph Grt Zo Q H2O

Cpx Ph Grt Zo Qz H2O

Bt Cpx Melt Grt Qz H2O

Bt Cpx Melt Grt H2O

Cpx Melt Grt H2O

Bt Ph Grt Zo Qz H2O

Bt Cpx Ph Grt Qz H2O

Bt Melt Pl Ilm Opx H2O (-Rt)

Cpx Ph Grt Ky Zo Qz H2O

Bt Cpx Melt Pl Ilm Opx(-Rt)

Bt Cpx Melt Grt Ky Qz H2O

Cpx Melt Grt H2O (-Amp)

Bt Cpx Ph Grt Zo Qz H2O

Ph Grt Ky Zo H2O

Bt Cpx Melt Grt Qz H2O (-Amp) Bt Cpx Melt Grt Ky Qz H2O (-Amp)

Bt Cpx Grt Qz H2O (-Amp)

Bt Cpx Ph Grt Ky Qz H2O

Chl Ph Grt Zo Qz

Bt Cpx Grt Zo Qz H2O

Bt Pl Ilm Opx H2O (-Rt)

Bt Melt Pl Ilm Grt Opx H2O (-Rt)

Bt Melt Pl Ilm Opx (-Rt)

Chl Ph Grt Zo H2O

Chl Ph Grt Ky Zo H2O

Ph Grt Zo Qz H2O

Bt Melt Pl Grt Opx H2O

Bt Pl Grt Opx H2O

Bt Chl Grt Qz

Bt Chl Grt Zo Qz

Bt Chl Grt Zo Qz H2O

Bt Grt Ky Zo H2O

Bt Chl Grt Qz H2O

Bt Chl Grt Ky Zo Qz H2O

Bt Grt Ky Zo Qz H2O

Bt Ph Grt Ky Zo H2O

Cpx Ph Grt Zo H2O

Bt Grt Qz H2O

Bt Grt Zo Qz H2O

Cpx Ph Grt Ky H2O

Bt Cpx Ph Grt Ky H2O

Ph Grt Ky Zo Qz H2O

Cpx Ph Grt Ky Qz H2O

Cpx Ph Grt Qz H2O Bt Melt Grt Qz H2O

Bt Pl Opx H2O

Bt Melt Pl Grt Qz

Bt Melt Grt Qz

Bt Cpx Grt Ky Qz H2O

Bt Cpx Ph Grt Ky Qz H2O (-Amp)

Bt Cpx Melt Grt

Bt Melt Pl Grt

Bt Cpx Melt Pl Grt Opx

Bt Cpx Melt Grt Opx

Bt Ph Grt Ky Zo Qz H2O

Bt Cpx Ph Grt Ky Qz H2O

Bt Chl Ilm Grt Qz

Bt Chl Ilm Grt Qz H2O

Bt Pl Ilm Grt Opx Qz H2O

12

3

456

7

8

9

10

11

12

13

14

15

16

17

18 19

20

21

22

23

24

2526

27

28

29 30

31

32

33

1

2

3

4

5

6

7

8

9

10

11

12

13

14

15

16

17

18

19

20

21

22

23

24

25

26

27

28

29

30

31

32

33

Bt Cpx Grt Qz H2O

Bt Melt Pl Ilm Opx H2O

Bt Pl Ilm Opx H2O RPB3A NCKFMASHTO (+ Amp, Rt)

Bt Cpx Melt Grt Qz

Bt Cpx Melt Pl Grt Qz

Figure 6. Pressure-temperature (P-T) pseudosection of sample RPB3A calculated in the NCKFMASHTO (Na2O-CaO-K2O-FeO-MgO-Al2O3-SiO2-H2O-TiO2-O2) system. The bulk compositions (mol%) used are SiO2 (45.15), TiO2 (0.33), Al2O3 (8.95), FeO (7.46), MgO (16.85), CaO (10.59), Na2O (2.01), K2O (0.13), O2 (1.03), and H2O (7.49). The prograde assemblage is shown in italics. Bt—biotite, Chl—chlorite, Ilm—ilmenite, Opx—orthopyroxene, Zo—zoisite, Ky—kyanite, Ph—phengite. Abbreviations as in Figure 3.

TABLE 2. PRESSURE-TEMPERATURE RESULTS OF THERMOBAROMETRIC CALCULATIONS

Metamorphic grade Sample Grt Loc. XFe Xgrs Amp Loc. XFe Cpx Loc. XFe XNa Pl Loc. Xan GHT

(ºC)

GCT

(ºC)

GHPQP

(kbar)

GCPQP

(kbar)

Epidote-Amphibolite RPB3A 42 core 0.564 0.205 31 inc 0.250 660RPB3A 12 core 0.524 0.206 10 inc 0.182 610

Eclogite RPB3A 15 near rim

0.509 0.214 51 relic 0.141 0.233 730

RPB1B 10 core 0.604 0.245 3 inc 0.232 0.278 810Granulite RPB1A 1 rim 0.543 0.211 3 cont 0.181 0.198 27 cont 0.297 770 11.5–12.5Amphibolite RPB3A 1 rim 0.542 0.219 19 cont 0.173 18 cont 0.314 590 8.5–10.0

RPB1B 33 rim 0.632 0.228 31 cont 0.243 32 cont 0.237 610 8.0–9.0RPB1A 60 rim 0.596 0.210 62 cont 0.210 61 cont 0.313 590 8.5–10.0

Note: Amp—amphibole; Grt—garnet; Cpx—clinopyroxene; Pl—plagioclase; Loc.—location; inc—inclusion in garnet; cont—contact to garnet; P—pressure; T—tem-perature. The numbers in the columns below the mineral abbreviations represent analytical spot numbers. XFe = Fe/(Fe + Mg), Xgrs = Ca/(Fe + Mg + Ca + Mn), Xan = Ca/(Ca + Na) (grs is grossular, an is anorthite). GH and GC—garnet-hornblende and garnet-clinopyroxene Fe-Mg exchange thermometers; GHPQ and GCPQ—garnet-hornblende-plagioclase-quartz and garnet-clinopyroxene-plagioclase-quartz barometers.

Page 9: Paleoproterozoic high-pressure metamorphic history of the …hosting03.snu.ac.kr/~hjung/pdf/Imayama_et_al-17.pdf · 2017-09-20 · Paleoproterozoic high-pressure metamorphic history

LITHOSPHERE | Volume 9 | Number 5 | www.gsapubs.org 9

Paleoproterozoic high-pressure metamorphism of the Salma eclogite | RESEARCH

because it matches with the temperature estimate (610–660 °C) in the conventional geothermobarometry section and overlaps that of the epidote-amphibolite facies on the metamorphic facies diagram delineated by Oh and Liou (1998). The eclogite facies assemblage of Omp + Amp + Grt + Bt + melt + Qz + Rt yields P-T conditions of 13–20 kbar and 730–820

°C (Fig. 7). The compositional isopleths of XNa

in clinopyroxene and XFe

in garnet are shown in Figure 7. The isopleths of the inner rims of garnet (X

Fe = 0.50–0.52) and relict omphacite (X

Na = 0.21–0.23) constrain the

peak P-T conditions to 17–18 kbar and 750–770 °C, which are within the range inferred from the mineral phase assemblages and within the tem-perature estimate (730–800 °C) in the conventional geothermobarometry section. The plagioclase-forming reaction occurred at conditions of 9–14 kbar and 700–850 °C, thus indicating that the growth of plagioclase with Ca-clinopyroxene occurred during decompression under granulite facies conditions. The granulite facies assemblage of Ca-Cpx + Amp + Grt + Qz + Rt + plagioclase (Pl) + melt + Bt constrains the P-T conditions to 12–14 kbar and 780–820 °C (Fig. 7). As a result, a clockwise P-T path including rapid uplift was estimated.

The garnet core in sample RPB1B is homogeneous, and the pseudo-section modeling was done without considering the effective bulk-rock

composition (Fig. 8). This bulk composition has relatively higher SiO2,

Na2O, and CaO contents than sample RPB3A. The calculated P-T pseu-

dosection shows an increase in the stability of plagioclase, which is stable at P-T conditions below 12–16 kbar at 700–850 °C (Fig. 8). The melt stability field occurs at temperatures above ~680–700 °C. The eclogite facies assemblage of Omp + Grt + Rt + Amp + Qz + melt is stable at the P-T conditions of 15–18.5 kbar and 740–790 °C (Fig. 8). The isopleths of X

Na = 0.27–0.28 for omphacite in garnet and X

Fe = 0.58–0.60 for garnet

cores yield peak P-T conditions of 16–17 kbar and 750–770 °C (Fig. 8), consistent with those of sample RPB3A. The granulite facies assemblage of Ca-Cpx + Amp + Grt + Qz + Rt + Pl + melt plots within the P-T condi-tions of 10–15 kbar and 720–800 °C. The isopleths of the X

Fe (0.63–0.64)

of garnet rims and the maximum XNa

(0.10) of clinopyroxene symplectite yield P-T conditions of 10–11 kbar and 790–820 °C, indicating that they reflect a rapid uplifting P-T path (Fig. 8).

WHOLE-ROCK CHEMISTRY

The major and trace element contents of the five studied metabasites were analyzed using an ICP-MS (Perkin Elmer Optima 3000, Activation

Temperature (oC) 660 720 780 840600 900

8

11

14

17

20

Pre

ssur

e (k

bar)

Bt Chl Qz H2O

Bt Chl Zo Qz

Bt Chl Grt Zo Qz H2O

Bt Grt Zo Qz H2O Bt Melt

Grt Zo Qz H2O

Bt Melt Grt Zo Qz

Bt Melt Pl Opx Qz

Bt Cpx Melt Grt Zo Qz

Bt Melt Pl Ilm Opx H2O

Bt Melt Pl

Grt Opx

Bt Melt Pl Ilm Opx Bt Melt Pl IIm Opx H2O (-Ru)

Bt Chl Ph Grt Zo H2O

Cpx Ph Grt Zo H2O

Bt Cpx Grt Zo Qz H2O

Ph Grt Zo H2O

Ph Grt Ky Zo H2O

Ph Grt Zo Qz H2O

Bt Cpx MeltGrt H2O

Bt Chl Ph Grt Zo Qz H2O

Cpx Ph Gt Ky Zo Qz H2O

Bt Chl Grt Zo H2O

Bt Pl Opx Qz H2O

Bt Pl Grt QzH2O

Bt Melt Pl Grt Qz H2O

Bt Melt Pl Opx Qz H2O Bt Melt Pl Grt Opx Qz

Bt Pl Opx H2O

Bt Melt Pl Opx H2O

Br Melt Pl Opx

Bt Melt Pl Ilm Opx (-Rt)

Bt Cpx Melt Pl Grt Opx

Bt Cpx MeltPl Ilm Opx (-Rt)

Bt Cpx Melt Pl Ilm

Grt Opx (-Rt)

Bt Cpx Melt Pl Ilm Grt Opx (-Rt)

Bt Qz H2O

Bt Chl Zo Qz H2O

Bt Zo Qz H2O Bt Grt

Qz H2O Bt Melt Grt Qz H2O

Bt Melt Pl Grt Qz

Bt Melt Grt Qz

Bt Melt Pl Grt

Bt Cpx Melt Pl Grt

Bt Cpx Melt Grt Opx

Bt Ph Grt Zo H2O

Cpx Phe Grt Ky H2O

Cpx Ph Grt Zo Qz H2O

Cpx Ph Grt Ky Qz H2O

Bt Cpx Grt Ky Qz H2O

Bt Cpx Grt Qz H2O

Bt Cpx Melt Grt Ky Qz

H2O

Bt Cpx Melt Grt

Bt Cpx Phe Grt Zo Qz H2O

Bt Chl Grt Zo Qz

Bt Ph Grt Ky Zo H2O

Bt Phe Grt Zo Qz H2O

Bt Cpx Melt Grt Zo Qz H2O

Bt Cpx Melt Grt Qz H2O

Bt Chl Ph Grt Zo Qz

Bt Grt Ky Zo Qz H2O

Chl Ph Grt Zo H2O

Chl Ph Grt Ky Zo H2O

RPB3A(effective bulk composition)

1

2

3

4

5

6

7

8

9

10

11

12

13

14

15

16

17

18

19

20

1

2 3

4

5

6

7

8

910

11

1213

14

15

16

17

18

1920

Cpx Ph Grt Ky Zo H2O

Bt Cpx Ph Grt Ky Qz H2O (-Amp)

Bt Pl Grt Opx Qz H2O

21

22

23

24

25

2122

23 25

26

26

27

27

28

29

30

31

28

29

30

31

Bt CpxMelt Pl Grt Qz

24

NCKFMASHTO (+ Amp, Rt)

50525456

6062

4858

24

20

16

12

Bt Cpx Melt Grt Qz

Grt100*XFe100*XNa

Cpx

16 58

Figure 7. Pressure-temperature (P-T) pseudosection of sample RPB3A calculated using the effective composition after 10% garnet fractionation. NCK-FMASHTO is Na2O-CaO-K2O-FeO-MgO-Al2O3-SiO2-H2O-TiO2-O2. The bulk compositions (mol%) used are SiO2 (45.39), TiO2 (0.36), Al2O3 (8.36), FeO (6.19), MgO (17.13), CaO (10.72), Na2O (2.23), K2O (0.14), O2 (1.14), and H2O (8.32). The peak and retrograde assemblages are shown in bold and white letters, respectively. Compositional isopleths of garnet for XFe and clinopyroxene for XNa from sample RPB3A are also shown. The bold circle represents the peak P-T conditions. Abbreviations are as in Figures 3 and 6.

Page 10: Paleoproterozoic high-pressure metamorphic history of the …hosting03.snu.ac.kr/~hjung/pdf/Imayama_et_al-17.pdf · 2017-09-20 · Paleoproterozoic high-pressure metamorphic history

IMAYAMA ET AL.

10 www.gsapubs.org | Volume 9 | Number 5 | LITHOSPHERE

Laboratories Ltd., Canada). The whole-rock data are listed in Data Reposi-tory Table DR2. The five metabasites have basaltic compositions, with SiO

2 contents of 47.0–52.4 wt% and low alkali concentrations in terms

of Na2O (1.49–2.16 wt%) and K

2O (0.02–0.33 wt%). These metabasites

were originally low-alkali tholeiitic basalts, as shown in Figure 9A. They plot in the island arc basalt (IAB) and mid-oceanic ridge basalt (MORB) fields on the Ti/100-Zr-Y and Nb-Zr/4-Y diagrams (Figs. 9B, 9C).

The REE patterns normalized using C1 chondritic values are basically depleted in light REEs, thus yielding flat patterns showing an affinity to normal (N) MORB (Figs. 10A, 10B). Sample RPB1B exhibits a negative Eu anomaly. The incompatible trace element abundances normalized using primitive mantle values are shown in Figures 10C and 10D. The elements ranging from Nd to Yb produce low and flat trends in most samples, except for sample RPB1B. In contrast, the elements from Rb to Sr, which are highly incompatible, record variable values, which were likely produced by disturbances during subduction. These whole-rock data indicate that

the origin of the Salma eclogites is likely subducted oceanic crust that originated at a spreading center. This interpretation agrees well with those of previous studies (Mints et al., 2010, 2014; Konilov et al., 2011).

ZIRCON U-Pb AGES AND GEOCHEMISTRY

Analytical Procedure

Zircon grains from the two eclogite samples (RPB3A and RPB1B) were separated using the standard heavy liquid technique and were then hand-picked under a binocular microscope. Cathodoluminescence (CL) and backscattered electron (BSE) images were obtained using the JEOL 6610LV scanning electron microscope at the Korea Basic Science Insti-tute (KBSI, Ochang, South Korea). The CL and BSE images from sam-ples RPB3A and RPB1B are shown in Figures 11 and 12, respectively. Microinclusions in zircons were identified using the scanning electron

Pre

ssur

e (k

bar)

8

11

14

17

20

23

Temperature (oC) 660 720 780 840600 900

Melt Pl Ilm Grt Opx Qz (-Rt)

Bt Pl Ilm Grt Qz

Bt Pl Grt Qz

Bt Grt Qz

Bt Grt Zo Qz

Bt Grt Zo Qz H2O

Ph Grt Zo Qz

Ph Grt Zo Qz H2O

Cpx Ph Grt Zo Qz H2O

Bt Pl Ilm Grt Qz

Cpx Ph Grt Ky Zo Qz H2O

Cpx Ph Grt Ky Qz H2O

Cpx Ph Grt Qz H2O

Bt CpxGrt Zo QzH2O

Bt Pl Grt Zo Qz

Cpx Ph Grt Qz H2O (-Amp)

Bt Cpx Ph Grt Qz H2O

Bt Melt Grt Zo Qz

Bt Melt Grt Qz H2O

Bt MeltGrt Zo Qz H2O

Bt Cpx Melt Grt Qz H2O

Bt Melt Pl Grt Qz

Bt Melt Grt Qz

Bt Cpx Melt Grt Qz

Bt Cpx Grt Qz H2O (-Amp)

Bt Pl Grt Qz H2O

Bt Pl Ilm GrtQz H2O

Cpx Melt Grt Qz

Bt Cpx Ph Grt Qz H2O (-Amp)

Bt Pl Ilm Grt Opx Qz H2O (-Rt)

Bt Melt Pl Ilm Grt

Cpx Melt Grt Qz H2O (-Amp)

Bt Cpx Melt Grt Qz H2O (-Amp)

Cpx Melt Grt Qz (-Amp) Bt Melt Pl Ilm Grt Opx Qz (-Rt)

Cpx Melt Pl Grt Qz

Melt Pl Ilm Opx Qz (-Rt)

Cpx Melt Pl Ilm Opx Qz (-Rt)

Cpx Melt Pl Ilm Grt Opx Qz (-Rt)

Bt Cpx Melt Pl Ilm Grt Qz (-Rt)

Cpx Melt Pl Grt Qz (-Amp)

Cpx Melt Pl Ilm Opx Qz(-Rt, -Amp)

Cpx Melt Pl Ilm Grt

Opx Qz (-Rt, -Amp)

Cpx Melt Pl Grt Opx Qz

Cpx Melt Pl Ilm Grt Opx Qz

Cpx Melt Pl Ilm Grt Opx Qz (-Amp)

Cpx Melt Pl Grt Opx Qz (-Amp)

Bt Phe Grt Zo Qz H2O

Bt Cpx Ph Grt Zo Qz H2O

Bt Melt Pl Ilm Grt Qz H2O (-Rt)

Bt Cpx Melt Grt Zo Qz H2OBt Melt Pl Grt Zo Qz Bt Melt Pl Grt Qz H2O

Bt Melt Pl Ilm Grt Qz

Bt Melt Pl Ilm Grt Opx Qz H2O (-Rt)

Cpx Melt Pl Ilm Grt Qz

Bt Cpx Melt Pl Ilm

Grt Qz

Bt Cpx Melt Pl Ilm Grt Opx Qz (-Rt)

Bt Cpx Grt Qz H2O

Bt Cpx Melt Pl Grt Qz

1

2

3

4

5

6

7

8

9

10

11

12

13

14

13

4

5

6

7

8

910

11 12

13

14

RPB1B

15

NCKFMASHTO (+ Amp, Rt)

60

6412

16

20

2456

28

2

Grt100*XFe100*XNa

Cpx

5816

Figure 8. Pressure-temperature (P-T) pseudosection of sample RPB1B calculated in the NCKFMASHTO (Na2O-CaO-K2O-FeO-MgO-Al2O3-SiO2-H2O-TiO2-O2) system. The bulk compositions (mol%) used are SiO2 (52.81), TiO2 (1.14), Al2O3 (8.46), FeO (9.87), MgO (11.14), CaO (10.45), Na2O (1.46), K2O (0.04), O2 (0.86), and H2O (3.77). The peak and retrograde assemblages are shown in bold and white letters, respectively. Compositional isopleths of garnet for XFe and clinopyroxene for XNa from sample RPB1B are also shown. The bold and dashed circles represent the peak and retrograde P-T conditions, respectively. Abbreviations are as in Figures 3 and 6.

Page 11: Paleoproterozoic high-pressure metamorphic history of the …hosting03.snu.ac.kr/~hjung/pdf/Imayama_et_al-17.pdf · 2017-09-20 · Paleoproterozoic high-pressure metamorphic history

LITHOSPHERE | Volume 9 | Number 5 | www.gsapubs.org 11

Paleoproterozoic high-pressure metamorphism of the Salma eclogite | RESEARCH

RPB1A RPB1B

RPB3A RPB3B RPB3C

A B C

Tholeiitic

Calc-Alkaline

Na2O+K2O MgO

FeOt

IAT

MORBCAB

WPB

Zr Y*3

Ti/100

AI

AII

B

CD

Zr/4 Y

Nb*2

Figure 9. The whole-rock compositions of retrograded eclogite and granulite in the Salma area. (A) Plotted on the FeOtotal-(Na2O + K2O)-MgO classifica-tion diagram. (B) Plotted on the Ti/100-Zr-Y*3 tectonic discrimination diagram. IAT—island-arc tholeiites; CAB—calc-alkaline basalts; WPB—within-plate basalts; MORB—mid-oceanic ridge basalt. (C) Plotted on the Nb*2-Zr/4-Y tectonic discrimination diagram. AI—within-plate alkali basalts; AII—within-plate alkali basalts and within-plate tholeiites; B—enriched-type MORB; C—within-plate tholeiites and volcanic-arc basalts; D—normal-type MORB and volcanic-arc basalts (Rollinson, 1993, and references therein).

.1

1

10

100

300

RbBa

ThU

NbTa

CeSr

NdSm

ZrHf

EuGd

TbDy

YEr

Yb

Sam

ple/

Prim

itive

Man

tle

.1

1

10

100

300

RbBa

ThU

NbTa

CeSr

NdSm

ZrHf

EuGd

TbDy

YEr

Yb

Sam

ple/

Prim

itive

Man

tle

N-MORB

E-MORB

OIB

RPB1A RPB1B

increasing incompatibility

N-MORB

E-MORB

OIB

RPB3A RPB3B RPB3C

increasing incompatibility

RPB1A RPB1B

.1

1

10

100

300

La CePr Nd SmEuGdTbDyHoErTmYbLu

Sam

ple/

C1

Cho

ndrit

e

.1

1

10

100

300

La CePr Nd SmEuGdTbDyHoErTmYbLu

Sam

ple/

C1

Cho

ndrit

e

RPB3A RPB3B RPB3C

N-MORB

E-MORB

OIB

N-MORB

E-MORB

OIB

DC

BA

Figure 10. Chondrite-normalized rare earth element patterns. MORB—mid-oceanic ridge basalt (N—normal; E—enriched); OIB—oceanic island basalts for (A) samples RPB3A–RPB3C and (B) samples RPB1A, RPB1B. Primitive mantle-normalized trace element patterns for (C) samples RPB3A–RPB3C and (D) samples RPB1A, RPB1B. Data are normalized using the values of chondrite and primitive mantle from Sun and McDonough (1989).

Page 12: Paleoproterozoic high-pressure metamorphic history of the …hosting03.snu.ac.kr/~hjung/pdf/Imayama_et_al-17.pdf · 2017-09-20 · Paleoproterozoic high-pressure metamorphic history

IMAYAMA ET AL.

12 www.gsapubs.org | Volume 9 | Number 5 | LITHOSPHERE

12.1

12.2

1845 ± 21

2733 ± 148.1

7.1

2811 ± 8

1776 ± 48

15.1

15.2

1783 ± 60

1855 ± 46

Grt

Qz

1892 ± 239.2

1904 ± 16

1855 ± 37

17.2

17.1

50μm 50μm

CL CL

CL CL CL

BSE

50μm

50μm50μm 50μm

A B C

E FD

CL BSECL

CL CL CL

CL CL CL

50μm 50μm 50μm

50μm50μm50μm

50μm 50μm 50μm

3.2

2703 ± 14

3.1 1799 ± 82

2.1

2757 ± 7

20.2

1833 ± 67

11.1

1848 ± 31Omp

Cpx

Qz13.1

13.2

1769 ± 101

2729 ± 10 18.1

1914 ± 28

17.1

1720 ± 79

17.2

1924 ± 34

2a.1

1863 ± 56

2a.2

1850 ± 36

20.1

1932 ± 44

A B C

E F

G H

D

I

Figure 11. (A–F) Representative cathodoluminescence (CL) and backscattered electron (BSE) images of dated zircon crystals in sample RPB3A. The analyzed spots are shown with their 207Pb/206Pb ages and spot numbers. Abbreviations as in Figure 3.

Figure 12. (A–I) Representative cathodoluminescence (CL) and backscattered electrons (BSE) images of dated zircon crystals in sample RPB1B. The analyzed spots are shown with their 207Pb/206Pb ages and spot numbers. Abbreviations as in Figure 3.

Page 13: Paleoproterozoic high-pressure metamorphic history of the …hosting03.snu.ac.kr/~hjung/pdf/Imayama_et_al-17.pdf · 2017-09-20 · Paleoproterozoic high-pressure metamorphic history

LITHOSPHERE | Volume 9 | Number 5 | www.gsapubs.org 13

Paleoproterozoic high-pressure metamorphism of the Salma eclogite | RESEARCH

microscope with energy-dispersive X-ray spectroscopy (SEM-EDX) detector at the KBSI and the Thermo Scientific DXR micro-Raman microscope equipped with a 532 nm laser at the Tectonophysics Labora-tory in the School of Earth and Environmental Sciences (Seoul National University). The EDX spectra obtained from the inclusions in the zircon were used to identify mineral inclusions by comparing them with those of minerals in thin sections within the same samples. The mineral inclusion assemblages in the zircons are listed in Table 3, and the EDX spectra of garnet, omphacite, and Ca-clinopyroxene inclusions in zircons are shown in Data Repository Figure 1.

The REE composition of zircon was analyzed using laser ablation (LA)-ICP-MS at the KBSI. The LA-ICP-MS system consists of a laser ablation system (213 nm Nd-YAG [neodymium-doped yttrium aluminum garnet laser] UP213, New Wave Research, a division of ESI), ICP, and a quadrupole mass spectrometer (X2 series, Thermo Scientific). The ana-lytical procedures for the REE analyses of the zircons followed those of Yuan et al. (2004) and Liu et al. (2007). Ablation signals were collected by ICP-MS using a time-resolved analysis of 45 s. The Nd-YAG laser was operated at a repetition rate of 10 Hz, a spot size of 55 mm, and an energy level of 80% (27 J/cm2). NIST 612 glass was used as an external calibration standard, and each analysis was normalized to the silicon content (29Si) as an internal standard. GLITTER software (http://www .glitter -gemoc.com/) was used for data reduction. Zircon REE data are given in Data Repository Table DR3.

The zircon U-Pb ages were analyzed using the SHRIMP (sensitive high-resolution ion microprobe) IIe ion microprobe at the KBSI. The analytical procedures for SHRIMP dating were mainly based on those proposed by Williams (1998). A spot size of 15–20 µm and a 1.5–2 nA negative ion oxygen beam (O

2−) were used for all analyses. The measured

206Pb/238U ratio was calibrated using the FC1 zircon standard (ca. 1099 Ma; Paces and Miller, 1993). The SL13 zircon standard was also used for the calibration of U concentrations (238 ppm; Hoskin, 1998). Data reduction, age calculations, and common Pb corrections were conducted using SQUID 2.50 (Ludwig, 2009) and Isoplot 3.6 software (Ludwig, 2008). The zircon U-Pb ages are listed in Data Repository Table DR4.

Results

Sample RPB3AThe zircon grains from sample RPB3A are subhedral and have sub-

rounded edges (Figs. 11A–11F). Most zircon grains have dark CL cores surrounded by pale gray CL mantles with sector or patchy zoning (Figs. 11A, 11B). In the BSE images of these grains, the cores are brighter than the mantles (Fig. 11C). However, some zircons have pale gray CL cores with sector or patchy zoning that are similar to the mantles surrounding the dark CL cores (Figs. 11D, 11E). A few grains have bright CL cores surrounded by pale gray CL mantles (Fig. 11F). Thin, bright CL rims are locally observed surrounding pale gray CL cores and mantles (Figs. 11E, 11F). The dark CL cores contain inclusions of garnet, amphibole, plagioclase, quartz, biotite, and rutile; the pale gray CL cores and mantles contain inclusions of garnet (Fig. 11F), amphibole, biotite, and quartz (Table 3; Data Repository Item). K-feldspar occurs along fractures in zircons and is likely associated with later alteration. On the chondrite-normalized diagram (Fig. 13A), the dark CL cores are enriched in HREEs, with Lu

N/Gd

N ratios of 8.6–33.6. Significant negative Eu anomalies can

be observed in these zircons (Eu/Eu* = 0.07–0.20). The 207Pb/206Pb ages of the dark CL cores are 2875–2387 Ma. The weighted mean 207Pb/206Pb age of 10 nearly concordant ages from the dark CL cores is 2716 ± 10 Ma (mean square of weighted deviates, MSWD = 2.7, n = 10, 2σ). In contrast, the REE patterns of the pale gray domains exhibit flat HREE

patterns (LuN/Gd

N ratio = 0.6–1.2) and small negative Eu anomalies (Eu/

Eu* = 0.40–0.61; Fig. 13A). They yield 207Pb/206Pb ages of 2378–1776 Ma, with a major cluster at 1900–1780 Ma. The weighted mean 207Pb/206Pb age of the concordant data is 1865 ± 15 Ma (MSWD = 1.7, n = 14, 2σ; Fig. 14A). The U-Pb zircon data from sample RPB3A produce a discordia

0.001

0.01

0.1

1

10

100

1000

10000

La Ce Pr* Nd SmEuGd Tb Dy Ho Er TmYb Lu

Pale gray domain

(A) RPB3A

Zirc

on/C

hond

rite

Dark CL core

0.001

0.01

0.1

1

10

100

1000

10000

La Ce Pr*Nd SmEuGd Tb Dy Ho Er TmYb Lu

Zirc

on/C

hond

rite

Pale gray domainDark CL core

(B) RPB1B

Figure 13. Chondrite-normalized rare earth element patterns for dif-ferent zircon domains. CL—cathodoluminescence. (A) Sample RPB3A. (B) Sample RPB1B. Data are normalized using the chondritic values of Sun and McDonough (1989).

TABLE 3. MINERAL INCLUSIONS IN ZIRCONS FROM THE SALMA ECLOGITES

Sample Domain Inclusion type

Qz Grt Amp Omp Ca-Cpx Bt Rt Pl K-Fsp Ap

RPB3A dark CL core • • • • • • rpale gray domain • • • •

bright CL rimRPB1B dark CL core •

pale gray domain • • • • • •bright CL rim

Note: CL is cathodoluminescence. Mineral abbreviations: Qz—quartz; Grt—gar-net; Amp—amphibole; Omp—omphacite; Cpx—clinopyroxene; Bt—bitote; Rt—rutile; Pl—plagioclase; K-Fsp—potassium feldspar; Ap—apatite. Triangle represents mineral formed by later alteration.

Page 14: Paleoproterozoic high-pressure metamorphic history of the …hosting03.snu.ac.kr/~hjung/pdf/Imayama_et_al-17.pdf · 2017-09-20 · Paleoproterozoic high-pressure metamorphic history

IMAYAMA ET AL.

14 www.gsapubs.org | Volume 9 | Number 5 | LITHOSPHERE

1.2 1.6 2.0 2.4 2.8 3.2 3.6 4.0 4.4

1600

2400

0.06

0.10

0.14

0.18

0.22P

b*20

7 Pb*

/206

238U/ 206Pb*

1820 ± 70 & 2905 ± 85 Ma

Dark CL core

Pale gray domain

Intercepts at

MSWD = 3.6

2000

2800

(A) RPB3A

(B) RPB1B

1680

1760

1840

1920

2000

1865 ± 15 MaMSWD = 1.7

1.4 1.8 2.2 2.6 3.0 3.4 3.8

1600

2000

2400

2800

0.06

0.08

0.10

0.12

0.14

0.16

0.18

0.20

0.22

207 P

b*/2

06P

b*

238U/ 206Pb*

1881 ± 38 & 2764 ± 33 MaIntercepts at

MSWD = 1.3

Dark CL core

Bright CL rim

Pale gray domain

1550

1650

1750

1850

1950

2050

1868 ± 17 MaMSWD = 1.4

Figure 14. Concordia diagrams for the sensitive high-resolution ion microprobe (SHRIMP) U-Pb analyses of zircon. The dashed line represents the discordia line. All error ellipses are quoted at the 1σ level. MSWD—mean square of weighted deviates. The mean and discordia ages are shown at the 2σ level. CL—cathodoluminescence. (A) Sample RPB3A. (B) Sample RPB1B.

Page 15: Paleoproterozoic high-pressure metamorphic history of the …hosting03.snu.ac.kr/~hjung/pdf/Imayama_et_al-17.pdf · 2017-09-20 · Paleoproterozoic high-pressure metamorphic history

LITHOSPHERE | Volume 9 | Number 5 | www.gsapubs.org 15

Paleoproterozoic high-pressure metamorphism of the Salma eclogite | RESEARCH

line, yielding an upper intercept age of 2905 ± 85 Ma and a lower intercept age of 1820 ± 70 Ma (Fig. 14A; MSWD = 3.6, 2σ), indicating that a Pb-loss event occurred during the period of Paleoproterozoic metamorphism. The brighter CL rims were not analyzed due to their insufficient thickness.

Sample RPB1BThe zircon grains from sample RPB1B are subhedral with rounded or

subrounded edges (Figs. 12A–12I). Three domains are observed on the basis of CL and BSE images. Most crystals have dark CL cores, which are surrounded by pale gray CL mantles (Figs. 12A, 12B, 12D). The BSE images show brighter cores and relatively darker mantles (Fig. 12C). Several zircons have pale gray CL cores with sector or patchy zoning that are similar in their CL brightness to the mantles surrounding the dark CL cores (Figs. 12E, 12F). In some zircons, the internal structure of the pale gray cores features a cloudy zoning pattern (Figs. 12G–12I). These pale gray cores are mostly surrounded by brighter CL outer rims, which vary in thickness from narrow (Figs. 12E–12H) to broad (Fig. 12I). Apatite occurs as inclusions in the dark CL cores, whereas Ca-clinopyroxene (Fig. 12D), omphacite (Fig. 12H), quartz, amphibole, rutile, and apatite occur in the pale gray CL domains (Table 3; Data Repository Fig. 1). The dark CL cores display HREE-enriched patterns with a steep slope from the middle REEs to the HREEs (Lu

N/Gd

N = 22.3–34.3). They also record negative

Eu anomalies (Eu/Eu* = 0.21–0.30; Fig. 13B). The 207Pb/206Pb ages of the dark CL cores are 2757–2342 Ma. The weighted mean 207Pb/206Pb age of the concordant data is 2727 ± 8 Ma (MSWD = 0.6, n = 8, 2σ). The pale gray domains have flat HREE patterns with moderate to shallow slopes from the middle REEs to the HREEs (Lu

N/Gd

N = 1.1–8.8) and slightly

smaller negative Eu anomalies (Eu/Eu* = 0.32–0.56) compared to the dark CL zircons (Fig. 13B). Their 207Pb/206Pb ages are Paleoproterozoic and range from 1967 to 1750 Ma, with the exception of 2 analytical spots that yield ages of 1687–1636 Ma. The weighted mean 207Pb/206Pb age of the pale gray domains is 1868 ± 17 Ma (MSWD = 1.4, n = 22, 2σ; Fig. 14B). The omphacite-bearing zircon domains yield ages of 1863 ± 56 Ma and 1850 ± 36 Ma (Fig. 12H). These U-Pb zircon data produce a discordia line yielding upper and lower intercept ages of 2764 ± 33 Ma and 1881 ± 38 Ma, respectively (MSWD = 1.3, 2σ; Fig. 14B), thus indicating that a Pb-loss event occurred during the period of Paleoproterozoic meta-morphism. These upper and lower intercept ages are consistent (within error) with the weighted mean ages of the dark CL cores and pale gray domains, respectively. One analytical spot from the brighter CL rim yields an apparent 207Pb/206Pb age of 1720 ± 79 Ma (Fig. 12I).

DISCUSSION

Meaning of Archean Zircon Ages

Petrographic, geochemical, and geochronological data from the Salma eclogites in the Kola Peninsula reveal the polymetamorphic history of this area. The ages of the magmatic protoliths of the Salma eclogites are known to be 2.94–2.92 Ga (Kaulina et al., 2010) and 2.88–2.82 Ga (Mints et al., 2010; Skublov et al., 2010a, 2011; Mel’nik et al., 2013). However, the 2.88–2.87 Ga zircon age obtained from Skublov et al. (2010a, 2011) should be interpreted as a metamorphic age, rather than as a magmatic age, because these zircons do not record zoning patterns that are typical for igneous zircons, such as concentric or banded zoning (Hoskin and Schaltegger, 2003); instead, they exhibit unzoned patterns, which are char-acteristic of metamorphic zircons. Mints et al. (2014) also interpreted 2.82 Ga to be the earliest age of metamorphism based on 176Hf/177Hf isotopic ratios, and suggested that subduction occurred in the Salma area between 2.87 and 2.82 Ga based on the intrusions of mafic dikes at 2.86–2.83 Ga

and felsic veins at 2.82–2.78 Ga. Mints et al. (2014) believed that high-pressure metamorphism, which formed the Salma eclogite, may have occurred during the subduction stage; however, the Archean age of the eclogite facies metamorphism in the Salma area is uncertain, because they did not provide direct evidence for it.

The 2.73–2.72 Ga ages obtained from the dark CL zircons in samples RPB3B and RPB1B are interpreted to represent an Archean amphibolite facies metamorphic event. The unzoned regions present in the dark CL zircons are generally produced by metamorphism. We also found a rep-resentative metamorphic mineral assemblage (i.e., garnet) in the dark CL zircons from sample RPB3A. Although high to moderate Th/U ratios in the dark CL zircons (0.2–5.1 for RPB3A, 0.2–0.3 for RPB1B) may indicate their magmatic origins (cf. Skublov et al., 2010a), metamorphic zircons with high Th/U ratios have also been reported in high-grade rocks (e.g., Harley et al., 2007).

Some researchers reported 2.72–2.70 Ga (retrograde) granulite facies metamorphism (Kaulina et al., 2010; Mints et al., 2010). Although there is no direct petrological evidence to link these ages to granulite facies metamorphism, the REE pattern with HREE enrichment in zircons was explained by zircon growth in equilibrium with melt during granulite facies metamorphism (Kaulina et al., 2010). However, melt can be pro-duced from temperatures of 680–700 °C, which correspond to conditions of upper amphibolite facies metamorphism in bulk-rock composition of the Salma eclogites (Figs. 7 and 8). Metamorphic zircons have been reported in amphibolites in the orogeny (e.g., Oh et al., 2014), thus imply-ing that (upper) amphibolite facies metamorphism can produce abundant zircons. In this study, the mineral inclusions within the dark CL zircons with 2.73–2.72 Ga ages are characterized by an amphibolite facies mineral assemblage (Grt + Amp + Pl + Qz + Rt ± Bt). The relatively enriched HREE patterns and remarkably negative Eu anomalies observed in the dark CL zircons, compared to the pale gray zircons, can be explained by the presence of less garnet and abundant plagioclase. These indicate that amphibolite facies metamorphism occurred at ca. 2.73–2.72 Ga.

Interpretation of Paleoproterozoic Zircon Ages

The pale gray CL zircons from samples RPB3A and RPB1B that contain garnet and omphacite yield 207Pb/206Pb age mean ages of 1865 ± 15 Ma and 1868 ± 17 Ma, respectively. Direct evidence for Paleoprotero-zoic eclogite is provided by 207Pb/206Pb ages of 1863 ± 56 and 1850 ± 36 Ma from omphacite-bearing zircon domains. These results indicate that eclogite facies metamorphism (16–18 kbar and 750–770 °C) occurred ca. 1.87 Ga. The flat HREE patterns indicate that these metamorphic zircons formed in equilibrium with garnet during eclogite facies metamorphism (e.g., Rubatto, 2002; Whitehouse and Platt, 2003; Imayama et al., 2012). Metamorphic zircons characterized by flat HREE patterns have been obtained from many eclogites around the world (e.g., northeast Green-land, Gilotti et al., 2004, McClelland et al., 2006; South Korea, Kim et al., 2006; Central Alps, Liati et al., 2009; Bohemian Massif, Bröcker et al., 2010). High HREE abundances in garnet produce flat HREE pat-terns in zircon. In addition, the very weak Eu anomalies in the pale gray CL zircons, compared to the dark CL zircons, indicate that plagioclase-free mineral assemblages exist in the eclogite. The presence of small Eu anomalies and Ca-clinopyroxene (instead of only omphacite) in the pale gray CL zircons may indicate that zircon growth continued to granulite facies during decompression. Nevertheless, the absence of plagioclase inclusions in the pale gray CL zircons means that the zircons yielding ages of ca. 1.87 Ga mainly grew during eclogite facies metamorphism.

Skublov et al. (2010a, 2011) reported that eclogite facies metamor-phism occurred ca. 1.92–1.88 Ga, based on the analysis of metamorphic

Page 16: Paleoproterozoic high-pressure metamorphic history of the …hosting03.snu.ac.kr/~hjung/pdf/Imayama_et_al-17.pdf · 2017-09-20 · Paleoproterozoic high-pressure metamorphic history

IMAYAMA ET AL.

16 www.gsapubs.org | Volume 9 | Number 5 | LITHOSPHERE

zircon rims surrounding Archean magmatic zircon cores (ca. 2.88–2.87 Ga) within the massive eclogite (sample 46, Table 1). However, the ca. 1.92–1.88 Ga metamorphic zircons include zoisite and quartz (Skublov et al., 2010a, 2011), which appear to represent prograde metamorphism, rather than peak eclogite facies metamorphism, which is represented by the presence of garnet and omphacite. Garnet in the analyzed eclogite records prograde zoning, with increasing pyrope contents from core to rim (Skublov et al., 2011). Studies of Lu-Hf and Sm-Nd garnet geochro-nology from the same eclogite yielded garnet–whole-rock isochron ages of 1901 ± 5 Ma and 1897 ± 16 Ma, respectively (Herwartz et al., 2012; Mel’nik et al., 2013). Although these ages were interpreted to reflect the timing of peak eclogite facies metamorphism, the Lu-Hf and Sm-Nd ages of garnets with growth zoning are generally interpreted to represent prograde metamorphic ages, due to the low diffusivities of REEs (Baxter and Scherer, 2013). The omphacite-bearing metamorphic zircons that formed ca. 1.87 Ga found in this study provide the first clear age of peak eclogite facies metamorphism in the Salma eclogite.

P-T Path During Paleoproterozoic Metamorphism

In the Salma eclogites, the identification of epidote-amphibolite facies prograde metamorphism was based on the presence of clinozoisite and amphibole inclusions in garnet cores and prograde zoned garnets with homogeneous cores and Mg-rich inner rims in sample RPB3A. Because the garnet in sample RPB1A only contains amphibole inclusions but lacks epidote (Fig. 4D), some Salma eclogites may have undergone amphibo-lite facies metamorphism prior to eclogite facies metamorphism. The P-T conditions of the prograde stage are estimated to be 13–18 kbar and 640–720 °C. These results closely match with the boundary between the amphibolite, epidote-amphibolite, and eclogite facies on the metamor-phic facies diagram developed by Oh and Liou (1998). This prograde metamorphism likely occurred ca. 1.92–1.88 Ga (Skublov et al., 2010a, 2011; Herwartz et al., 2012; Mel’nik et al., 2013), as mentioned herein.

The granulite facies overprinting (10–14 kbar and 770–820 °C) occurred during the subsequent exhumation stage from 17 to 18 kbar, lead-ing to the breakdown of omphacite to Ca-clinopyroxene and plagioclase. Because the zircon growth ca. 1.87 Ga could have continued to undergo granulite facies metamorphism, this probably occurred soon after the eclogite facies metamorphism, thus implying that rapid decompression occurred. Upon cooling, the amphibolite facies overprint occurred at conditions of 8–10 kbar and 590–610 °C.

Tectonic Implications

Some researchers have interpreted the eclogites in the Belomorian mobile belt to represent evidence of Archean subduction followed by col-lision, leading to the amalgamation of the Karelia craton, the Kola craton, and the microcontinent between them (Mints et al., 2010, 2014). However, in this study, an age of ca. 1.87 Ga for eclogite facies metamorphism was obtained from zircons with omphacite inclusions and flat HREE patterns within the Salma eclogite. The Paleoproterozoic zircons collected from the eclogites from the Kuru-Vaara quarries also exhibit flat HREE pat-terns (Skublov et al., 2011) and the P-T paths of the eclogites from the Grindino and Salma areas are similar (Mints et al., 2014), indicating that the eclogites in the Belomorian mobile belt formed during the Paleopro-terozoic; these findings do not support the Archean subduction-collision model. The Archean subduction-collision model is not able to explain the regional occurrence of Paleoproterozoic granulite facies metamorphism (i.e., the Lapland and Umba belts) in the Kola-Karelian collisional zone (cf. Daly et al., 2006). The 1.87 Ga metamorphic age of the Salma eclogite

in this study supports the model of Paleoproterozoic collision between the Kola and Karelian cratons suggested by Berthelsen and Marker (1986), Zhao et al. (2002), and Daly et al. (2006).

Eclogites that formed within a transitional eclogite-granulite facies P-T range could have formed in a deep continental crustal root zone (e.g., De Paoli et al., 2009). However, the whole-rock chemistry of the Salma eclogites in this study is characterized by depleted light REEs, which reflects their origins as N-MORB and is consistent with the results of pre-vious studies (Konilov et al., 2011; Mints et al., 2014). It is likely that the 2.94–2.93 Ga ages inferred from the magmatic zircon in the Salma eclog-ite represent the protolith age of the N-MORB (Kaulina et al., 2010). This study also indicates that the Salma eclogites underwent an amphibolite facies metamorphic event ca. 2.73–2.72 Ga. During the Paleoproterozoic subduction of the unit including the protolith of the Salma eclogites, these rocks underwent progressive metamorphism from epidote-amphibolite facies ca. 1.92–1.88 Ga to eclogite facies ca. 1.87 Ga. This subduction stage was followed by the collision of the Kola and Karelian cratons; the unit including the Salma eclogites was rapidly uplifted and recorded strong overprinting produced first by granulite facies metamorphism and then by amphibolite facies metamorphism during or after the collision.

Secular Changes in the Geothermal Gradients of Subduction Zones

Subduction in the Precambrian may have proceeded differently than modern subduction, due to the hotter conditions in the mantle of the early Earth (e.g., Davies, 1992); however, it is debatable when ancient subduction changed into modern subduction (e.g., Cawood et al., 2006; van Hunen and Moyen, 2012, and references therein). Determining the changes in the patterns of metamorphism at plate boundary zones allows us to understand the evolutions and geodynamics of subduction zones (e.g., Brown, 2006, 2009). Determining the timing of the first appearances of high-pressure granulite, eclogite, and blueschist is thus very important for understanding the changes in the patterns of metamorphism and the geothermal gradient in Precambrian subduction zones over time.

The oldest high-pressure granulite is Neoarchean (ca. 2.5 Ga) and is present in the Jianping complex; it was produced by a subduction-collision event in the eastern region of the North China Craton (Wei et al., 2001; O’Brien and Rötzler, 2003; Liu et al., 2011; Lu et al., 2017). The P-T conditions of this high-pressure granulite metamorphic event were 10–13 kbar and 780–850 °C (Wei et al., 2001; Wang and Cui, 1992; Lu et al., 2017; Fig. 15). Eclogite facies metamorphism occurred at ca. 2.0 Ga in the Usagaran orogen of Tanzania, with peak metamorphic conditions of

~750 °C and 18 kbar (Möller et al., 1995; Collins et al., 2004). The peak metamorphic conditions and P-T paths of the Tanzanian eclogites are similar to those of the Salma eclogites (Fig. 15). These rocks underwent rapid decompression after peak metamorphism and were then retrograded into first granulites and then amphibolites. Because the eclogite facies metamorphism in the Salma eclogite and the Usagaran eclogite occurred during the Paleoproterozoic, subduction accompanying the development of eclogite likely began during or prior to the Paleoproterozoic. The oldest reported blueschist (ca. 750–730 Ma) is from the Aksu Group of western China (Liou et al., 1989, 1996; Nakajima et al., 1990; Maruyama et al., 1996; Zhu et al., 2011; Yong et al., 2013); the P-T conditions of this blue-schist facies metamorphism were 4–10 kbar and 300–400 °C (Liou et al., 1989, Zhang et al., 1999; Fig. 15). The geothermal gradient required for a formation of high-temperature eclogite (>~750 °C) is higher than that of blueschist and lower than that of high-pressure granulite. Consequently, the first appearance of high-pressure granulite, eclogite, and blueschist in the Neoarchean, Paleoproterozoic, and Neoproterozoic, respectively,

Page 17: Paleoproterozoic high-pressure metamorphic history of the …hosting03.snu.ac.kr/~hjung/pdf/Imayama_et_al-17.pdf · 2017-09-20 · Paleoproterozoic high-pressure metamorphic history

LITHOSPHERE | Volume 9 | Number 5 | www.gsapubs.org 17

Paleoproterozoic high-pressure metamorphism of the Salma eclogite | RESEARCH

may reflect a decrease in the geothermal gradients of subduction zones from the Neoarchean to the Neoproterozoic due to the cooling of the Earth.

Changes in metamorphic facies during prograde metamorphism in the Phanerozoic to Paleoproterozoic eclogites and Neoarchean high-pressure granulites can provide clearer information about changes in the geothermal gradients at subduction zones from the Neoarchean to the Phanerozoic. This study provides evidence that there was a prograde metamorphic event that involved epidote-amphibolite facies and/or amphibolite facies during the formation of the Paleoproterozoic Salma eclogites. Eclogites that have undergone low-grade epidote-amphibolite facies conditions during prograde metamorphism are known from several Phanerozoic subduction zones, such as the high- and ultrahigh-pressure eclogites of Sambagawa (Takasu, 1984; Enami et al., 1994; Itaya et al., 2011; Fig. 15). However, blueschist facies metamorphism mainly occurs prior to high- and ultrahigh-pressure eclogite facies metamorphism in most Pha-nerozoic subduction zones, such as the western Alps and New Caledonia

(e.g., Oh and Liou, 1998; Rubatto and Hermann, 2001; Fig. 15). More-over, eclogites that formed via prograde metamorphism from amphibolite facies are almost absent in Phanerozoic subduction zones (Oh and Liou, 1998). However, the prograde mineral assemblage of the Neoarchean granulite in the Jianping complex is Amp + Pl ± Qz ± Bt, which repre-sents amphibolite facies metamorphism (Wang and Cui, 1992; Liu et al., 2011). These data imply that Paleoproterozoic subduction zones were relatively warmer than Phanerozoic subduction zones but colder than Neoarchean subduction zones.

CONCLUSIONS

Based on the petrologic, thermobarometric, geochemical, and geochro-nological data presented in this study and previous studies, the following conclusions are proposed for the tectonothermal evolution of the Salma eclogites in the Kola Peninsula.

1. The source rocks for the Salma eclogites formed in a mid-ocean ridge ca. 2.94–2.93 Ga and first underwent amphibolite facies metamor-phism ca. 2.73–2.72 Ga. This amphibolite facies metamorphism is con-firmed by inclusions of Grt + Amp + Pl + Qz + Rt ± Bt in 2.73–2.72 Ga unzoned dark CL zircons, which are characterized by enriched HREE patterns and remarkably negative Eu anomalies.

2. The Salma eclogites may have undergone epidote-amphibolite facies or amphibolite facies prograde metamorphism at ca. 1.92–1.88 Ga. The ca. 1.87 Ga peak eclogite facies metamorphism can be inferred from the U-Pb age dating of pale gray CL metamorphic zircons with inclusions of Grt + Omp + Ca-Cpx + Amp + Qz + Rt ± Bt. These zircons record flat HREE patterns and nearly lack Eu anomalies. The peak metamorphic P-T conditions were ~16–18 kbar and 750–770 °C. Soon after this peak meta-morphism, granulite facies metamorphism occurred after decompression at 10–14 kbar and 770–820 °C and was followed by amphibolite facies overprinting at 8–10 kbar and 590–610 °C during cooling.

3. The Paleoproterozoic subduction and subsequent continent-conti-nent collision between the Kola and Karelian continents is supported by the occurrence of eclogite facies metamorphism at ca. 1.87 Ga.

4. The oldest appearances of high-pressure granulite, eclogite, and blueschist occurred in the Neoarchean, Paleoproterozoic, and Neopro-terozoic, respectively, which may reflect a decrease in the geothermal gradients of Precambrian subduction zones due to the cooling of the Earth. The prograde metamorphism from epidote-amphibolite facies or amphibolite facies to eclogite facies in the subduction zone during the Paleoproterozoic also implies that the Paleoproterozoic subduction zone was relatively warmer than the Phanerozoic subduction zone but was colder than the eclogite-free Neoarchean subduction zones.

ACKNOWLEDGMENTSWe thank K. de Jong, Seoul National University, Korea, for helpful discussion; Y. Park, Seoul National University, Korea, for assistance with micro-Raman analyses; and Juhn G. Liou and two anonymous reviewers for constructive and critical reviews that significantly helped to improve the manuscript. We also thank R. Damian Nance for careful editorial handling. This work was supported by National Research Foundation of Korea (NRF) grants 657 NRF-2013R1A1A2058525, NRF-2014R1A2A2A01003052, and NRF-2017R1A2B2011224.

REFERENCES CITEDBalagansky, V., Shchipansky, A., Slabunov, A.I., Gorbunov, I., Mudruk, S., Sidorov, M., Azimov,

P., Egorova, S., Stepanova, A., and Voloshin, A., 2014, Archaean Kuru-Vaara eclogites in the northern Belomorian Province, Fennoscandian Shield: Crustal architecture, timing, and tectonic implications: International Geology Review, v. 57, p. 1543–1565, doi: 10 .1080 /00206814 .2014 .958578.

Baxter, E.F., and Scherer, E.E., 2013, Garnet geochronology: Timekeeper of tectonometamor-phic processes: Elements, v. 9, p. 433–438, doi: 10 .2113 /gselements .9 .6 .433.

Berthelsen, A., and Marker, M., 1986, Tectonics of the Kola collision suture and adjacent Ar-chean and early Proterozoic terrains in the northeastern region of the Baltic Shield: Tec-tonophysics, v. 126, p. 31–55, doi: 10 .1016 /0040 -1951 (86)90219 -2.

200 400 600 800 10000

10

20

EG

BS

Pre

ssur

e (k

bar)

EA

GS

LG

HG

AM

Temperature (oC)

previous studythis study

P-T path

Thisstudy

TZ

JC

AG

TZ

30

NC (HP)

WA(HP)

SA (UHP)

SA(HP)

WA(UHP)

Figure 15. Pressure-temperature (P-T) diagram showing schematic P-T-time path (red arrow) of the Salma eclogites determined in this study and that (blue arrow) of the Usagaran eclogite facies rocks in Tanzania (TZ; Collins et al., 2004). Solid, dashed, and dotted red rectangles indicate the P-T conditions of the eclogite, granulite, and amphibolite facies metamorphic stages, respectively, of the Salma eclogite. Orange and purple rectangles represent the P-T conditions of Neoarchean high-pressure granulites in the Jianping Complex (JC) of the North China Craton and Neoproterozoic blue-schist in the Aksu Group (AG) of western China. Dotted and dashed green curves represent schematic field P-T curves for high-pressure (HP) and ultrahigh-pressure eclogites (UHP) of the representative Phanerozoic sub-duction zones: western Alps (WA), New Caledonia (NC), and Sambagawa (SA). The P-T curves and petrogenetic grid are from Oh and Liou (1998), Rubatto and Hermann (2001), and Itaya et al. (2011). BS—blueschist facies, GS—greenschist facies, EG—eclogite facies, HG—high-pressure granulite facies, LG—low-pressure granulite facies, EA—epidote-amphibolite facies, AM—amphibolite facies.

Page 18: Paleoproterozoic high-pressure metamorphic history of the …hosting03.snu.ac.kr/~hjung/pdf/Imayama_et_al-17.pdf · 2017-09-20 · Paleoproterozoic high-pressure metamorphic history

IMAYAMA ET AL.

18 www.gsapubs.org | Volume 9 | Number 5 | LITHOSPHERE

Boniface, N., Schenk, V., and Appel, P., 2012, Paleoproterozoic eclogites of MORB-type chemis-try and three Proterozoic orogenic cycles in the Ubendian belt (Tanzania): Evidence from monazite and zircon geochronology, and geochemistry: Precambrian Research, v. 192–195, p. 16–33, doi: 10 .1016 /j .precamres .2011 .10 .007.

Bröcker, M., Klemd, R., Kooijman, E., Berndt, J., and Larionov, A., 2010, Zircon geo chronology and trace element characteristics of eclogites and granulites from the Orlica-Śnieźnik complex, Bohemian Massif: Geological Magazine, v. 147, p. 339–362, doi: 10.1017 /S0016756809990665.

Brown, M., 2006, Duality of thermal regimes is the distinctive characteristic of plate tectonics since the Neoarchean: Geology, v. 34, p. 961–964, doi: 10 .1130 /G22853A .1.

Brown, M., 2009, Metamorphic patterns in orogenic systems and the geological record, in Cawood, P.A., and Kröner, A., eds., Earth Accretionary Systems in Space and Time: Geo-logical Society of London Special Publication 318, p. 37–74, doi: 10 .1144 /SP318 .2.

Cawood, P.A., Kröner, A., and Pisarevsky, S., 2006, Precambrian plate tectonics: Criteria and evidence: GSA Today, v. 16, p. 4–11, doi: 10 .1130 /GSAT01607 .1.

Collins, A.S., Reddy, S.M., Buchan, C., and Mruma, A., 2004, Temporal constraints on Palaeo-proterozoic eclogite formation and exhumation (Usagaran Orogen, Tanzania): Earth and Planetary Science Letters, v. 224, p. 175–192, doi: 10 .1016 /j .epsl .2004 .04 .027.

Connolly, J.A.D., 1990, Multivariable phase diagrams: An algorithm based on generalized ther-modynamics: American Journal of Science, v. 290, p. 666–718, doi: 10 .2475 /ajs .290 .6 .666.

Dale, J., Powell, R., White, R.W., Elmer, F.L., and Holland, J.B., 2005, A thermodynamic model for Ca-Na clinoamphiboles in Na2O-CaO-FeO-MgO-Al2O3-SiO2-H2O-O for petrological calculations: Journal of Metamorphic Geology, v. 23, p. 771–791, doi: 10 .1111 /j .1525 -1314 .2005 .00609 .x.

Daly, J.S., Balagansky, V.V., Timmerman, M.J., Whitehouse, M.J., de Jong, K., Guise, P., Bogdanova, S., Gorbatschev, R., and Bridgwater, D., 2001, Ion microprobe U-Pb zircon geochronology and isotopic evidence for a trans-crustal suture in the Lapland–Kola Orogen, northern Fennoscan-dian Shield: Precambrian Research, v. 105, p. 289–314, doi: 10 .1016 /S0301 -9268 (00) 00116 -9.

Daly, J.S., Balagansky, V.V., Timmerman, M.J., and Whitehouse, M.J., 2006, The Lapland–Kola orogen: Palaeoproterozoic collision and accretion of the northern Fennoscandian litho-sphere, in Gee, D.G., and Stephenson, R.A., eds., European lithosphere dynamics: Geo-logical Society of London Memoir 32, p. 579–598, doi: 10 .1144 /GSL .MEM .2006 .032 .01 .35.

Davies, G.F., 1992, On the emergence of plate tectonics: Geology, v. 20, p. 963–966, doi: 10 .1130 /0091 -7613 (1992)020 <0963: OTEOPT>2 .3 .CO;2.

De Paoli, M.C., Clarke, G.L., Klepeis, K.A., Allibone, A.H., and Turnbull, I.M., 2009, The eclogite-granulite transition: Mafic and intermediate assemblages at Breaksea Sound, New Zea-land: Journal of Petrology, v. 50, p. 2307–2343, doi: 10 .1093 /petrology /egp078.

Dick, H.J.B., et al., 2000, A long in situ section of the lower ocean crust: Results of ODP Leg 176 drilling at the Southwest Indian Ridge: Earth and Planetary Science Letters, v. 179, p. 31–51, doi: 10 .1016 /S0012 -821X (00)00102 -3.

Dokukina, K.A., Kaulina, T.V., Konilov, A.N., Mints, M.V., Van, K.V., Natapov, L., Belousova, E., Simakin, S.G., and Lepekhina, E.N., 2014, Archaean to Palaeoproterozoic high-grade evolu-tion of the Belomorian eclogite province in the Gridino area, Fennoscandian shield: Geo-chronological evidence: Gondwana Research, v. 25, p. 585–613, doi: 10 .1016 /j .gr .2013 .02 .014.

Ellis, D.J., and Green, E.H., 1979, An experimental study of the effect of Ca upon garnet-clinopyroxene Fe-Mg exchange equilibria: Contributions to Mineralogy and Petrology, v. 71, p. 13–22, doi: 10 .1007 /BF00371878.

Enami, M., Wallis, S., and Banno, Y., 1994, Paragenesis of sodic pyroxene-bearing quartz schists: Implications for the P-T history of the Sanbagawa belt: Contributions to Mineral-ogy and Petrology, v. 116, p. 182–198, doi: 10 .1007 /BF00310699.

Evans, Y.P., 2004, A method for calculating effective bulk composition modification due to crystal fractionation in garnet-bearing schist: Implications for isopleth thermobarometry: Journal of Metamorphic Geology, v. 22, p. 547–557, doi: 10 .1111 /j .1525 -1314 .2004 .00532 .x.

Gilotti, J.A., Nutman, A.P., and Brueckner, H.K., 2004, Devonian to Carboniferous collision in the Greenland Caledonides: U-Pb zircon and Sm-Nd ages of high-pressure and ultrahigh-pressure metamorphism: Contributions to Mineralogy and Petrology, v. 148, p. 216–235, doi: 10 .1007 /s00410 -004 -0600 -4.

Graham, C.M., and Powell, R., 1984, A garnet-hornblende geothermometer: Calibration, test-ing, and application to the Pelona Schist, southern California: Journal of Metamorphic Geology, v. 2, p. 13–31, doi: 10 .1111 /j .1525 -1314 .1984 .tb00282 .x.

Harley, S.L., Kelly, N.M., and Möller, A., 2007, Zircon behaviour and the thermal histories of mountain chains: Elements, v. 3, p. 25–30, doi: 10 .2113 /gselements .3 .1 .25.

Hawthorne, F., Oberta, R., Harlow, G.E., Maresch, W.V., Martin, R.F., Schumacher, J.C., and Welch, M.D., 2012, Nomenclature of the amphibole supergroup: American Mineralogist, v. 97, p. 2031–2048, doi: 10 .2138 /am .2012 .4276.

Herwartz, D., Skublov, S.G., Berezin, A.V., and Mel’nik, A.E., 2012, First Lu-Hf garnet ages of eclogites from the Belomorian mobile belt (Baltic shield, Russia): Doklady Earth Sciences, v. 443, p. 377–380, doi: 10 .1134 /S1028334X12030130.

Holland, T., and Powell, R., 1996, Thermodynamics of order-disorder in minerals. 2. Symmet-ric formalism applied to solid solutions: American Mineralogist, v. 81, p. 1425–1437, doi: 10 .2138 /am -1996 -11 -1215.

Holland, T.J.B., and Powell, R., 1998, An internally consistent thermodynamic data set for phases of petrological interest: Journal of Metamorphic Geology, v. 16, p. 309–343, doi: 10 .1111 /j .1525 -1314 .1998 .00140 .x.

Hölttä, P., Balagansky, V., Garde, A.A., Mertanen, S., Peltonen, P., Slabunov, A., Sorjonen-Ward, P., and Whitehouse, M., 2008, Archean of Greenland and Fennoscandia: Episodes, v. 31, p. 13–19.

Hoskin, P.W.O., 1998, Minor and trace element analysis of natural zircon (ZrSiO4) by SIMS and laser ablation ICPMS: A consideration and comparison of two broadly competitive techniques: Journal of Trace and Microprobe Techniques, v. 16, p. 301–326.

Hoskin, P.W.O., and Schaltegger, U., 2003, The composition of zircon and igneous and metamor-phic petrogenesis, in Hanchar, J.M., and Hoskin, P.W.O., eds., Zircon: Mineralogical Society of America Reviews in Mineralogy and Geochemistry, v. 53, p. 27–62, doi: 10 .2113 /0530027.

Imayama, T., et al., 2012, Two-stage partial melting and contrasting cooling rates within the Higher Himalayan crystalline sequences in the far-eastern Nepal Himalaya: Lithos, v. 134–135, p. 1–22, doi: 10 .1016 /j .lithos .2011 .12 .004.

Itaya, T., Tsujimori, T., and Liou, J.G., 2011, Evolution of the Sanbagawa and Shinmato high-pressure belts in SW Japan Insights from K-Ar (Ar-Ar) geochronology: Journal of Asian Earth Sciences, v. 42, p. 1075–1090, doi: 10 .1016 /j .jseaes .2011 .06 .012.

Kaulina, T.V., Yapaskurt, V.O., Prresnyakov, S.L., Savchenko, E.E., and Simakin, S.G., 2010, Metamorphic evolution of the Archean eclogite-like rocks of the Shirokaya and Uzkaya Salma area (Kola Peninsula): Geochemical features of zircon, composition of inclusions, and age: Geochemistry International, v. 48, p. 871–890, doi: 10 .1134 /S001670291009003X.

Kim, S.W., Oh, C.W., Williams, I.S., Rubbato, D., Ryu, I.-C., Rajesh, V.J., Kim, C.-B., Guo, J., and Zhai, M., 2006, Phanerozoic high-pressure eclogite and intermediate-pressure granulite facies metamorphism in the Gyeonggi Block, South Korea: Implications for the eastward extension of the Dabie-Sulu continental collision zone: Lithos, v. 92, p. 357–377, doi: 10 .1016 /j .lithos .2006 .03 .050.

Kohn, M.J., and Spear, F.S., 1990, Two new barometers for garnet amphibolites with applica-tions to eastern Vermont: American Mineralogist, v. 75, p. 89–96.

Komiya, T., Maruyama, S., Masuda, T., Nohda, S., Hayashi, M., and Okamoto, K., 1999, Plate tectonics at 3.8–3.7 Ga: Field evidence from the Isua accretionary complex, southern West Greenland: Journal of Geology, v. 107, p. 515–554, doi: 10 .1086 /314371.

Konilov, A.N., Shchipansky, A.A., Mints, M.V., Dokukina, K.A., Kaulina, T.V., Bayanova, T.B., Natapov, L.M., Belousova, E.A., Griffin, W.L., and O’Reilly, S.Y., 2011, The Salma eclog-ites of the Belomorian Province, Russia: HP/UHP metamorphism through the subduction of Mesoarchean oceanic crust, in Dobrzhinetskaya, L.F., et al., eds., Ultrahigh-pressure metamorphism: 25 years after the discovery of coesite and diamond: London, Elsevier, p. 623–670, doi: 10 .1016 /B978 -0 -12 -385144 -4 .00018 -7.

Konrad-Schmolke, M., O’Brien, P.J., de Capitani, C., and Carswell, D.A., 2008, Garnet growth at high- and ultra-high pressure conditions and the effect of element fractionation on mineral modes and composition: Lithos, v. 103, p. 309–332, doi: 10 .1016 /j .lithos .2007

.10 .007.Liati, A., Gebauer, D., and Fanning, C.M., 2009, Geochronological evolution of HP metamorphic

rocks of the Adula Nappe, Central Alps, in pre-Alpine and Alpine subduction cycles: Jour-nal of the Geological Society [London], v. 166, p. 797–810, doi: 10 .1144 /0016 -76492008 -033.

Liou, J.G., et al., 1989, Proterozoic blueschist belt in western China: Best documented Pre-cambrian blueschists in the world: Geology, v. 17, p. 1127–1131, doi: 10 .1130 /0091 -7613 (1989) 017 <1127: PBBIWC>2 .3 .CO;2.

Liou, J.G., Graham, S.A., Maruyama, S., and Zhang, R.Y., 1996, Characteristics and tectonic sig-nificance of the Late Proterozoic Aksu blueschists and diabasic dikes, northwest Xinjiang, China: International Geology Review, v. 38, p. 228–244, doi: 10 .1080 /00206819709465332.

Liu, S., Santosh, M., Wang, W., Bai, X., and Yang, P., 2011, Zircon U-Pb chronology of the Jianping Complex: Implications for the Precambrian crustal evolution history of the northern margin of North China Craton: Gondwana Research, v. 20, p. 48–63, doi: 10 .1016 /j .gr .2011 .01 .003.

Liu, X., Gao, S., Diwu, C.R., Yuan, H., and Hu, Z.C., 2007, Simultaneous in-situ determination of U-Pb age and trace elements in zircon by LA-ICP-MS in 20 µm spot size: Chinese Sci-ence Bulletin, v. 52, p. 1257–1264, doi: 10 .1007 /s11434 -007 -0160 -x.

Lu, J.-S., Zhai, M.-G., Lu, L.-S., and Zhao, L., 2017, P-T-t evolution of Neoarchaean to Paleo-proterozoic pelitic granulites from the Jidong terrane, eastern North China Craton: Pre-cambrian Research, v. 290, p. 1–15, doi: 10 .1016 /j .precamres .2016 .12 .012.

Ludwig, K.R., 2008, User’s manual for Isoplot 3.6: A geochronological toolkit for Microsoft Excel: Berkeley, California, Berkeley Geochronology Center Special Publication 4, 77 p.

Ludwig, K.R., 2009, SQUID 2 Rev. 2.50: A user’s manual: Berkeley, California, Berkeley Geo-chronology Center Special Publication 5, 110 p.

Maruyama, S., Liou, J.G., and Terabayashi, M., 1996, Blueschists and eclogites of the world -and their exhumation: International Geology Review, v. 38, p. 485–594, doi: 10 .1080 /00206819709465347.

McClelland, W.C., Power, S.E., Gilotti, J.A., Mazdab, F.K., and Wopenka, B., 2006, U-Pb SHRIMP geochronology and trace element geochemistry of coesite-bearing zircons, north-east Greenland Caledonides, in Hacker, B.R., et al., eds., Ultrahigh-pressure metamorphism: Deep continental subduction: Geological Society of America Special Paper 403, p. 23–43, doi: 10 .1130 /2006 .2403 (02).

Mel’nik, A.E., Skublov, S.G., Marin, Y.B., Berezin, A.V., and Bogomolov, E.S., 2013, New data on the age (U-Pb, Sm-Nd) of garnetites from Salma eclogites of the Belomorian mobile belt: Doklady Earth Sciences, v. 448, p. 78–85, doi: 10 .1134 /S1028334X13010133.

Mints, M.V., Belousova, E.A., Konilov, A.N., Natapov, L.M., Shchipansky, A.A., Griffin, W.L., O’Reilly, S.Y., Dokukina, K.A., and Kaulina, T.V., 2010, Mesoarchean subduction processes: 2.87 Ga eclogites from the Kola Peninsula, Russia: Geology, v. 38, p. 739–742, doi: 10

.1130 /G31219 .1.Mints, M.V., Dokukina, K.A., and Konilov, A.N., 2014, The Meso-Neoarchaean Belomorian

eclogite province: Tectonic position and geodynamic evolution: Gondwana Research, v. 25, p. 561–584, doi: 10 .1016 /j .gr .2012 .11 .010.

Möller, A., Appel, P., Mezger, K., and Schenk, V., 1995, Evidence for a 2 Ga subduction zone: Eclogites in the Usagaran belt of Tanzania: Geology, v. 23, p. 1067–1070, doi: 10 .1130 /0091

-7613 (1995)023 <1067: EFAGSZ>2 .3 .CO;2.Morimoto, N., 1988, Nomenclature of pyroxenes: Mineralogical Magazine, v. 52, p. 535–550,

doi: 10 .1180 /minmag .1988 .052 .367 .15.Nakajima, T., Maruyama, S., Uchiumi, S., Liou, J.G., Wang, X., Xiao, X., and Graham, S.A.,

1990, Evidence for late Proterozoic subduction from 700-Myr-old blueschists in China: Nature, v. 346, p. 263–265, doi: 10 .1038 /346263a0.

Newton, R.C., Charlu, T.V., and Kleppa, O.J., 1980, Thermochemistry of the high structural state plagioclases: Geochimica et Cosmochimica Acta, v. 44, p. 933–941, doi: 10 .1016 /0016 -7037 (80)90283 -5.

Page 19: Paleoproterozoic high-pressure metamorphic history of the …hosting03.snu.ac.kr/~hjung/pdf/Imayama_et_al-17.pdf · 2017-09-20 · Paleoproterozoic high-pressure metamorphic history

LITHOSPHERE | Volume 9 | Number 5 | www.gsapubs.org 19

Paleoproterozoic high-pressure metamorphism of the Salma eclogite | RESEARCH

O’Brien, P.J., and Rötzler, J., 2003, High-pressure granulites: Formation, recovery of peak con-ditions and implications for tectonics: Journal of Metamorphic Geology, v. 21, p. 3–20, doi: 10 .1046 /j .1525 -1314 .2003 .00420 .x.

Oh, C.W., and Liou, J.G., 1998, A petrogenetic grid for eclogite and related facies under high-pressure metamorphism: The Island Arc, v. 7, p. 36–51, doi: 10 .1046 /j .1440 -1738 .1998 .00180.x.

Oh, C.W., Imayama, T., Yi, S., Kim, T., Ryu, I., Jeon, J., and Yi, K., 2014, Middle Paleozoic meta-morphism in the Hongseong area, South Korea and its tectonic meaning to Paleozoic orogeny in Northeast Asia: Journal of Asian Earth Sciences, v. 95, p. 203–216, doi: 10 .1016 /j .jseaes .2014 .08 .011.

Paces, J.B., and Miller, J.D., 1993, Precise U-Pb ages of Duluth Complex and related mafic inclu-sions, northeastern Minnesota: Geochronological insights into physical, petrogenetic, pa-leomagnetic, and tectonomagmatic processes associated with the 1.1 Ga midcontinent rift system: Journal of Geophysical Research, v. 98, p. 13,997–14,013, doi: 10 .1029 /93JB01159.

Pattison, D.R.M., 2003, Petrogenetic significance of ortho pyroxene-free garnet + clinopy-roxene + plagioclase ± quartz-bearing metabasites with respect to the amphibolite and granulite facies: Journal of Metamorphic Geology, v. 21, p. 21–34, doi: 10 .1046 /j .1525

-1314 .2003 .00415 .x.Powell, R., and Holland, T.J.B., 1988, An internally consistent thermodynamic dataset with

uncertainties and correlations. 3. Applications to geobarometry, worked examples and a computer program: Journal of Metamorphic Geology, v. 6, p. 173–204, doi: 10 .1111 /j .1525 -1314 .1988 .tb00415 .x.

Rollinson, H.R., 1993, Using geochemical data: Evaluation, presentation and interpretation: Longman Geochemistry Series: London, Pearson Education Ltd., 352 p.

Rubatto, D., 2002, Zircon trace element geochemistry: Partitioning with garnet and the link between U-Pb ages and metamorphism: Chemical Geology, v. 184, p. 123–138, doi: 10

.1016 /S0009 -2541 (01)00355 -2.Rubatto, D., and Hermann, J., 2001, Exhumation as fast as subduction: Geology, v. 29, p. 3–6,

doi: 10 .1130 /0091 -7613 (2001)029 <0003: EAFAS>2 .0 .CO;2.Shchipansky, A.A., Khodorevskaya, L.I., and Slabunov, A.I., 2012, The geochemistry and iso-

topic age of eclogites from the Belomorian Belt (Kola Peninsula): Evidence for subducted Archean oceanic crust: Russian Geology and Geophysics, v. 53, p. 262–280, doi: 10 .1016 /j .rgg .2012 .02 .004.

Skublov, S.G., Balashov, Y.A., Marin, Y.B., Berezin, A.V., Mel’nik, A.E., and Paderin, I.P., 2010a, U-Pb age and geochemistry of zircons from Salma eclogites (Kuru-Vaara deposit, Belo-morian Belt): Doklady Earth Sciences, v. 432, p. 791–798, doi: 10 .1134 /S1028334X10060188.

Skublov, S.G., Berezin, A.V., Marin, Y.B., Rizvanoca, N.G., Bogomollov, E.S., Sergeeva, N.A., Vasil’eva, I.M., and Guseva, V.F., 2010b, Complex isotopic-geochemical (Sm-Nd, U-Pb) study of Salma eclogites: Doklady Earth Sciences, v. 434, p. 1396–1400 (in Russian origi-nal pages 802–806), doi: 10 .1134 /S1028334X10100247.

Skublov, S.G., Berezin, A.V., and Mel’nik, A.E., 2011, Paleoproterozoic eclogites in the Salma area, northwestern Belomorian mobile belt: Composition and isotopic geochronologic characteristics of minerals and metamorphic age: Petrology, v. 19, p. 470–495, doi: 10 .1134 /S0869591111050055.

Slabunov, A.I., et al., 2006, The Archean of the Baltic shield: Geology, geochronology, and geodynamic settings: Geotectonics, v. 40, p. 409–433, doi: 10 .1134 /S001685210606001X.

Stern, R.J., 2005, Evidence from ophiolites, blueschists, and ultrahigh-pressure metamorphic terranes that the modern episode of subduction tectonics began in Neoproterozoic time: Geology, v. 33, p. 557–560, doi: 10 .1130 /G21365 .1.

Sun, S.S., and McDonough, W.F., 1989, Chemical and isotopic systematics of oceanic basalts; Implications for mantle composition and processes, in Saunders, A.D., and Norrey, M.J., eds., Magmatism in ocean basins: Geological Society of London Special Publication 42, p. 313–345, doi: 10 .1144 /GSL .SP .1989 .042 .01 .19.

Tajcmanová, L., Connolly, J.A.D., and Cesare, B., 2009, A thermodynamic model for titanium and ferric iron solution in biotite: Journal of Metamorphic Geology, v. 27, p. 153–165, doi: 10 .1111 /j .1525 -1314 .2009 .00812 .x.

Takasu, A., 1984, Prograde and retrograde eclogites in the Sambagawa Metamorphic Belt, Besshi District, Japan: Journal of Petrology, v. 25, p. 619–643, doi: 10 .1093 /petrology /25 .3 .619.

Tsujimori, T., and Ernst, W.G., 2014, Lawsonite blueschists and lawsonite eclogites as proxies for paleo-subduction zone processes: A review: Journal of Metamorphic Geology, v. 32, p. 437–454, doi: 10 .1111 /jmg .12057.

van Hunen, J., and Moyen, J.-F., 2012, Archean subduction: Fact or fiction?: Annual Review of Earth and Planetary Sciences, v. 40, p. 195–219, doi: 10 .1146 /annurev -earth -042711 -105255.

Van Kranendonk, M.J., Smithies, R.H., Hickman, A.H., and Champion, D., 2007, Review: Secular tectonic evolution of Archean continental crust: Interplay between horizontal and verti-cal processes in the formation of the Pilbara Craton, Australia: Terra Nova, v. 19, p. 1–38, doi: 10 .1111 /j .1365 -3121 .2006 .00723 .x.

Volodichev, O.I., Slabunov, A.I., Bibikova, E.V., Konilov, A.N., and Kuzenko, T.I., 2004, Archean eclogites in the Belomorian mobile belt, Baltic shield: Petrology, v. 12, p. 540–560.

Volodichev, O.I., Slabunov, A.I., Sibelev, O.S., Skublov, S.G., and Kuzenko, T.I., 2012, Geochro-nology, mineral inclusions, and geochemistry of zircons in eclogitized gabbronorites in the Gridino area, Belomorian Province: Geochemistry International, v. 50, p. 657–670, doi: 10 .1134 /S0016702912060080.

Waldbaum, D.R., and Thompson, J.B., 1968, Mixing properties of sanidine crystalline solu-tions: II. Calculations based on volume data: American Mineralogist, v. 53, p. 2000–2017.

Wang, C.Q., and Cui, W.Y., 1992, Studies of garnets from the Archaean metamorphic complex in Liaoxi (West Liaoning)-Chifeng area: Acta Petrological Mineralogica, v. 11, p. 157–168.

Wei, C.J., Zhang, A.L., Wu, T.H., and Li, J.H., 2001, Metamorphic P-T conditions and geologi-cal significance of high-pressure granulite from the Jianping complex, western Liaoning province: Acta Petrological Sinica, v. 17, p. 269–282.

White, R.W., Powell, R., Holland, T.J.B., and Worley, B.A., 2000, The effect of TiO2 and Fe2O3 on metapelitic assemblages at greenschist and amphibolites facies conditions: Mineral equilibria calculations in the system K2O-FeO-MgO-Al2O3-SiO2-H2O-TiO2-Fe2O3: Journal of Metamorphic Geology, v. 18, p. 497–511, doi: 10 .1046 /j .1525 -1314 .2000 .00269 .x.

White, R.W., Powell, R., and Holland, T.J.B., 2001, Calculation of partial melting equilibria in the system Na2O-CaO-K2O-FeO-MgO-Al2O3-SiO2-H2O (NCKFMASH): Journal of Metamor-phic Geology, v. 19, p. 139–153, doi: 10 .1046 /j .0263 -4929 .2000 .00303 .x.

Whitehouse, M.J., and Platt, J.P., 2003, Dating high-grade metamorphism—Constraints from rare-earth elements in zircon and garnet: Contributions to Mineralogy and Petrology, v. 145, p. 61–74, doi: 10 .1007 /s00410 -002 -0432 -z.

Whitney, D.L., and Evans, B.W., 2010, Abbreviations for names of rock-forming minerals: American Mineralogist, v. 95, p. 185–187, doi: 10 .2138 /am2010 .3371.

Williams, I.S., 1998, U-Th-Pb geochronology by ion microprobe, in McKibben, M.A., et al., eds., Applications of microanalytical techniques to understanding mineralizing processes: Society of Economic Geologists Reviews in Economic Geology Volume 7, p. 1–35, doi: 10 .5382 /Rev .07 .01.

Yong, W., Zhang, L., Hall, C.M., Mukasa, S.B., and Essen, E.J., 2013, The 40Ar/39Ar and Rb-Sr chronology of the Precambrian Aksu blueschists in western China: Journal of Asian Earth Sciences, v. 63, p. 197–205, doi: 10 .1016 /j .jseaes .2012 .05 .024.

Yuan, H., Gao, S., Liu, X., Li, H., Günter, D., and Wu, F., 2004, Accurate U-Pb age and trace element determinations of zircon by laser ablation-inductively coupled plasma-mass spectrometry: Geostandards and Geoanalytical Research, v. 28, p. 353–370, doi: 10 .1111 /j .1751 -908X .2004 .tb00755 .x.

Zhang, L., Jiang, W., Wei, C., and Dong, S., 1999, Discovery of deerite from the Aksu Precam-brian blueschist terrane and its geological significance: Science in China, ser. D, v. 42, p. 233–239, doi: 10 .1007 /BF02878960.

Zhao, G., Cawood, P.A., Wilde, S.A., and Sun, M., 2002, Review of global 2.1–1.8 Ga orogens: Implications for a pre-Rodinia supercontinent: Earth-Science Reviews, v. 59, p. 125–162, doi: 10 .1016 /S0012 -8252 (02)00073 -9.

Zhu, W., Zheng, B., Shu, L., Ma, D., Wu, H., Li, Y., Huang, W., and Yu, J., 2011, Neoproterozoic tectonic evolution of the Precambrian Aksu blueschist terrane, northwestern Tarim, China: Insights from LA-ICP-MS zircon U-Pb ages and geochemical data: Precambrian Research, v. 185, p. 215–230, doi: 10 .1016 /j .precamres .2011 .01 .012.

MANUSCRIPT RECEIVED 16 FEBRUARY 2017 REVISED MANUSCRIPT RECEIVED 12 JUNE 2017 MANUSCRIPT ACCEPTED 24 JULY 2017

Printed in the USA