Metamorphic zirconology of continental subduction zones

Metamorphic zirconology of continental subduction zones

Accepted Manuscript Metamorphic zirconology of continental subduction zones Ren-Xu Chen, Yong-Fei Zheng PII: DOI: Reference: S1367-9120(17)30208-0 ht...

5MB Sizes 1 Downloads 100 Views

Accepted Manuscript Metamorphic zirconology of continental subduction zones Ren-Xu Chen, Yong-Fei Zheng PII: DOI: Reference:

S1367-9120(17)30208-0 http://dx.doi.org/10.1016/j.jseaes.2017.04.029 JAES 3065

To appear in:

Journal of Asian Earth Sciences

Received Date: Revised Date: Accepted Date:

21 January 2017 25 April 2017 27 April 2017

Please cite this article as: Chen, R-X., Zheng, Y-F., Metamorphic zirconology of continental subduction zones, Journal of Asian Earth Sciences (2017), doi: http://dx.doi.org/10.1016/j.jseaes.2017.04.029

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

Review article

Metamorphic zirconology of continental subduction zones

Ren-Xu Chen*, Yong-Fei Zheng

CAS Key Laboratory of Crust-Mantle Materials and Environments, School of Earth and Space Sciences, University of Science and Technology of China, Hefei 230026, China

*

Corresponding author. Email: [email protected]

Abstract Zircon is widely used to date geological events and trace geochemical sources in high-pressure (HP) to ultrahigh-pressure (UHP) metamorphic rocks of continental subduction zones. However, protolith zircons may be modified by three different types of metamorphic recrystallization via mechanisms of solid-state transformation, metasomatic alteration and dissolution reprecipitation; new zircon growth may be induced by dehydration reactions below the wet solidus of crustal rocks (metamorphic zircon) or peritectic reactions above the wet solidus (peritectic zircon). As a consequence, there are different origins of zircon domains in high-grade metamorphic rocks from collisional orogens. Thus, determining the nature of individual zircon domains is substantial to correct interpretation of their origin in studies of isotopic geochronology and geochemical tracing. We advocate an integrated study of zircon mineragraphy (internal structure and external morphology), U-Pb ages, mineral inclusions, trace elements, and Lu-Hf and O isotope compositions. Only in this way we are in a position to advance the simple zircon applications to metamorphic zirconology, enabling discrimination between the different origins of zircon and providing constraints on the property of fluid activity at subduction-zone conditions. The metamorphic recrystallization of protolith zircons and the new growth of metamorphic and peritectic zircons are prominent in HP to UHP metamorphic rocks of collisional orogens. These different types of recrystallized and grown zircons can be distinguished by their differences in element and isotope compositions. While the protolith nature of metamorphosed rocks dictates water availability, the P-T conditions of subduction zones dictate the property of subduction-zone fluids. The fluids of different properties may be produced at different positions of subducting and exhuming crustal slices, and they may physically and chemically mix with each other in continental subduction channels. Such fluids can act as an important agent not only for the physical transport of protolith zircons but also for the chemical transport of element Zr and other fluid-mobile incompatible trace elements from the subducted crust to the mantle wedge. Therefore, the discrimination between the different types of zircons provides a powerful means to decipher the role of fluids in subduction zone processes.

Keywords: Zircon; regional metamorphism; subduction zone; dehydration reaction; peritectic reaction; fluid action

1. Introduction Zircon is a common accessory mineral in crustal rocks. Because of its high physiochemical stability and refractory nature, zircon is widely used to date geological events and trace geochemical sources (e.g., Hermann et al., 2001, 2013; Valley, 2003; Zheng et al., 2004, 2006a; Watson and Harrison, 2005; Kemp et al., 2006; Harley and Kelly, 2007a; McClelland and Lapen, 2013; Kohn et al., 2015; Taylor et al., 2016). This is because zircon is rich in U and Th but poor in Pb and has the high closure temperature of Pb diffusion (Zheng and Fu, 1998; Cherniak and Watson, 2003). Thus, zircon U-Pb dating has been one of the most commonly used and effective methods in geochronological studies (e.g., Wu and Zheng, 2004; Harley et al., 2007; Rubatto and Hermann, 2007a). As a phase enriched in Hf relative to radioactive Lu, zircon retains a strong fingerprint of the isotopic feature of crustal sources from which it crystallized. This provides robust evidence for growth and reworking of crustal rocks in the Earth’s history (e.g., Kemp et al., 2006; Zheng et al., 2006a, 2007a; Scherer et al., 2007). Zircon may contain significant amounts of temperature- or process-sensitive trace elements such as rare earth elements (REE) and Y as well as high field strength elements (HFSE), which can provide compelling evidence for conditions of zircon growth. This is important to reconstruction of of magmatic and metamorphic processes and to tracing the origin of host rocks (e.g., Harley et al., 2007; Rubatto and Hermann, 2007a). Although mineral O isotopes have a high sensitivity to low-temperature surface processes, zircon has the high stability in preserving its O isotope composition. Thus, zircon O isotope studies have been an effective means to discriminate the role of low-temperature versus high-temperature processes in its host rocks (e.g., Valley, 2003; Chen et al., 2011). Because of these unique properties, zircon is widely used in geological dating and geochemical tracing in various processes such as magma crystallization, partial melting, metamorphism, fluid action and hydrothermal mineralization, leading to the emergence of a new research branch named as zirconology. Zircon can grow or recrystallize under various conditions during hydrothermal, metamorphic, anatectic to magmatic processes (e.g., Rubatto, 2002; Wu and Zheng, 2004; Hoskin, 2005; Harley and Kelly, 2007b; Schaltegger, 2007). Geochemical information recorded in zircon can be correctly interpreted only when zircon growth conditions can be clearly linked to the evolutionary history of host rocks in subduction zones. High-pressure (HP) to ultrahigh-pressure (UHP) metamorphic rocks in collisional orogens generally experience multistages of evolution (Rumble et al., 2003; Zheng et al., 2003a), and metamorphic dehydration and partial melting of crustal rocks at subduction-zone conditions may produce different types of fluids such as aqueous solutions, hydrous melts and supercritical fluids (Zheng et al., 2011a; Zheng and Hermann, 2014). For these reasons, there are different origins of zircon in HP to UHP metamorphic rocks. Relict zircons (protolith zircons of either magmatic or detrital origins) may suffer variable extents of metamorphic recrystallization

through structural modification and chemical alteration in response to changes in P-T conditions and fluid accessibility; new zircon may grow during metamorphic dehydration and partial melting at subduction-zone conditions. As a consequence, the orogenic metamorphic rocks may contain not only recrystallized zircons of different types but also newly grown zircons of different types. The recrystallization of protolith zircons and the growth of new zircons in crustal rocks may take place under a broad range of P-T conditions during prograde, peak or post-peak greenschist-, amphibolite-, eclogite- and granulite-facies conditions (e.g., Fraser et al., 1997; Wu et al., 2006; Baldwin and Brown, 2008; Liu and Liou, 2011). With the advanced application of zircon studies to dating of geological events and tracing of geochemical sources and processes, the term zirconology has being used increasingly in the literature (e.g., Zheng, 2009; Xia et al., 2009, 2010; Chen et al., 2010, 2011; Nemchin, et al., 2012; Tichomirowa et al., 2012; Li et al., 2013). This involves an integrated study of zircon mineragraphy (internal structure and external morphology), U-Pb ages, mineral inclusions, trace elements, and Lu-Hf and O isotopes. Such a zirconological study is necessary in order to correctly interpret observations from zircons in crustal and mantle rocks. This is a big step in applying a single mineral to studies of metamorphic geology and geochemistry. For this reason, this paper provides a review on the studies of metamorphic zirconology in UHP metamorphic rocks from continental subduction zones. Although many of examples are taken from the Dabie-Sulu orogenic belt in China, available data from the other typical UHP terranes on Earth are also taken into account. The results indicate that different types of zircons can be discriminated by studying their properties during recrystallization and growth in continental subduction zones. This also provides constraints on the fluid action during subduction-zone processes.

2. Fundamentals for metamorphic zirconology In the present review of metamorphic zirconology, magmatic zircon is referred to as that crystallized from magmatic melts, and relict zircon is referred to as those inherited from crustal protoliths (some are of magmatic origin whereas the other is of detrital origin). On the other hand, metamorphic zircon is referred to as that formed through dehydration reactions below the wet solidus of crustal rocks, and peritectic zircon is referred to as those crystallized through peritectic reactions above the wet solidus of crustal rocks. While magmatic melts have separated from their parental rocks and transported upwards with large extent of evolution by fractional crystallization, anatectic melts are not separated from their parental rocks and thus only experienced the smallest extent of fractional crystallization. Nevertheless, anatectic zircon may grow from anatectic melts in which the local oversaturation of Zr is achieved by fractional crystallization of Zr-poor minerals. Therefore, the physical and chemical properties of subduction-zone fluids are a key to mineralogical

processes that form or rework zircon at continental subduction-zone conditions. Understanding these properties and processes is substantial not only to petrogenetic interpretation of zircons from high-grade metamorphic rocks in collisional orogens but also to tectonic interpretation of their U-Pb ages and geochemical signatures (e.g., Wu and Zheng, 2004; Harley and Kelly, 2007; Rubatto and Hermann. 2007; Zheng, 2009, 2012; Hermann et al., 2013). As defined in petrology of magmatic rocks, a magma is a mixture of crystal and melt, whereas the melt is short of crystalline minerals. Thus, a melt is part of a magma rather than whole magma. The difference between melt and magma is the occurrence and amount of crystalline minerals in the melt. A melt becomes a magma as soon as rock-forming minerals have significantly crystallized from the melt. A felsic melt is produced by partial melting of crustal rocks. Anatexis is referred to as partial melting of lower degrees to result in migmatization, generating anatectic melts that have not left their parental rocks. The anatectic melts were produced through peritectic reactions at temperatures above the wet solidus of crustal rocks. Typically, the anatectic melt is produced by migmatization, and its crystallized product is veinlets in metamorphic rocks, leucosomes in migmatites (metatexite and diatexite) and pegmatite veins in felsic gneisses. Geochemically, the anatectic melt has achieved thermodynamic equilibrium with the peritectic mineral in partitioning of water and incompatible trace elements, but it is not with the relict mineral. On the other hand, magmatism requires partial melting of higher degrees with significant transport and accumulation of anatectic melts. Thus, magmatic melts have escaped from their parental rocks (migmatites) with significant evolution in petrology. The anatectic melt becomes the magmatic melt after its significant evolution with fractional crystallization of rock-forming minerals. Therefore, the magmatic melt has achieved thermodynamic equilibrium with the crystallized minerals in partitioning of water and incompatible trace elements. Zircon grown under different environments can entrap the concurrently grown minerals, fluids and melts as its inclusions, which provide important records of its formation conditions and mechanism (e.g., Hermann et al., 2001; Liu and Liou, 2011; Li et al., 2013). The occurrence of inclusions (mineral, fluid or melt) in zircon provides an opportunity to correlate the growth zones of zircon with metamorphic/anatectic conditions. This is usually realized by identifying the distribution of mineral inclusion assemblage, CL images and trace element composition of zircon or mineral inclusions (e.g., Hermann et al., 2001; Liu and Liou, 2011). However, the relationship between inclusion species and host zircon domains is commonly complicated, limiting its application to zircon genesis. Three mechanisms have been proposed to account for the occurrence of inclusions in zircon (e.g., Gebauer et al., 1997; Liu et al., 2001; Zheng et al., 2011b): (1) entrapping of concurrently grown minerals, fluids or melts during zircon growth; (2) squeezing/crystallizing of inclusions along fractures into the preexisting zircon; (3) transforming of

precursor mineral inclusions into new mineral inclusions during metamorphism. Once inclusions were entrapped into zircon, their composition can hardly change because of the refractoriness of zircon. Elastic models suggest that transformation of a precursor quartz to coesite appears unlikely to happen (e.g., Gillet et al., 1984; Van Dermolen and Van Roermund, 1986). In this regard, the third mechanism seems unlikely. Healed structures of earlier deformed minerals and disturbed structures such as patch zoning and resorption can be found around secondary inclusions by detailed microstructure imaging (e.g., Gebauer et al., 1997; Dubińskaa et al., 2004). In contrast, cracks or healing traces generally cannot be observed from primary mineral inclusions. As such, detailed structure analyses can distinguish between the secondary and primary origins of mineral inclusions, and only the primary inclusions are of petrological meaning in the zircon genesis. Coesite and other eclogite-facies minerals as well as low-P (e.g., feldspar, biotite and quartz) minerals were found in zircon cores from schist and gneiss from the main hole of Chinese Continental Scientific Drilling (CCSD-MH) in the Sulu orogen (Zhang et al., 2006) and jadeite quartzite from Shuanghe in the Dabie orogen (Gao et al., 2015). However, U-Pb ages, CL images and trace element compositions indicate that these zircon cores are of magmatic rather than metamorphic origin. Detailed structure observations reveal that there are cracks and disturbed structure, indicating these inclusions are formed through metasomatic alteration along fractures due to fluid infiltration during UHP metamorphism (Gao et al., 2015). Thus, it is critical to distinguish between the primary and secondary mineral inclusions when interpreting the formation conditions of host zircon. Detailed studies of mineral inclusions have been carried out on zircons from various UHP metamorphic rocks from the Dabie-Sulu orogenic belt (Liu and Liou, 2011, and references therein). The results show that, at the prograde HP eclogite-facies stage during subduction, mineral inclusions in metamorphic zircons are predominated by quartz, garnet, omphacite, phengite, rutile, K-feldspar, dolomite and apatite. At the peak UHP eclogite-facies metamorphic stage, mineral inclusions are coesite, garnet, omphacite, phengite, rutile, jadeite, kyanite, titanite, K-feldspar, aragonite, magnesite and apatite. At the amphibolite-facies retrogression stage during exhumation, mineral inclusions are only composed of low-pressure minerals such as quartz, plagioclase, albite, amphibole, calcite and apatite. The metamorphic assemblage of mineral inclusions is not only related to metamorphic P-T conditions, but also controlled by the composition of host rocks (Liu and Liou, 2011). For HP to UHP eclogite-facies metamorpohic zircons, inclusion mineral assemblages are Coe/Qtz + Grt + Omp + Phe ± Mgs ± Ap for eclogite/amphibolite in granitic orthogneiss, Coe/Qtz + Grt + Omp + Phe + (Mgs+Arg)/Dol + Ap for eclogite in marble, Coe/Qtz + Grt ± Omp/Jd + Phe + Ttn + Ap for paragneiss, Coe/Qtz + Phe + Ap ± Grt ± Jd ± Rt ± Ttn ± Ky ± Kfs for granitic orthogneiss, Coe/Qtz + Grt + Phe + Rt +Ap + Omp/Jd + Ky for quartzite, and

Coe/Qtz + Grt + Omp + Ap + (Mgs + Arg)/Dol for marble. For the late amphibolite-facies metamorphic zircons, inclusion mineral assemblages are Amp+Pl+Ap for eclogite in granitic orthogneiss, Qtz+Cal+Amp for eclogite in marble, Qtz+Ab for paragneiss, orthogneiss and quartzite and Cal for marble. Analyses of these mineral inclusions, in combination with zircon U-Pb dates, can provide important constraints on P-T-t paths of zircon domains and further their host rocks (Hermann et al., 2001; Liu and Liou, 2011). Magmatic zircon generally exhibits high REE abundances and steep REE patterns, positive Ce anomalies and negative Eu anomalies (e.g., Rubatto, 2002; Whitehouse and Platt, 2003). It has been documented in a number of studies that distribution of trace elements in zircon reflects the metamorphic conditions of crustal rocks and that the individual stages of zircon growth are often associated with specific metamorphic conditions (e.g., Schaltegger et al., 1999; Rubatto, 2002; Rubatto and Hermann, 2003; Whitehouse and Platt, 2003; Kelly and Harley, 2005). Available studies suggest that the trace element composition of newly growth zircon is mainly dictated by the following issues. (1) The ability of trace elements to enter its crystal lattice during subduction-zone metamorphism. For example, during eclogite-facies metamorphism, contraction of the zircon lattice favors the substitution of Zr4+ by Hf4+ due to its smaller ionic radius, but hinders the substitution of Zr4+ by elements other than Hf due to their larger ionic radii. This may induce the fractionation of Th to U, Lu to Hf and LREE to HREE to some extent (e.g., Wu and Zheng, 2004; Chen et al., 2010). Wang and Griffin (2004) argued that enrichment of Hf, and depletion of Y, U and Th in metamorphic zircon may be ascribed to different partition coefficients during metamorphism. (2) The composition of coexisting fluid/melt phases (e.g., Rowley et al., 1997; Keay et al., 2001; Rubatto, 2002; Wu et al., 2007; Li et al., 2013), which records the nature of dehydration/peritectic reactions. (3) The concurrent growth or recrystallization of specific minerals in hosting specific trace elements, such as garnet for HREE and Y, rutile for Nb, Ta and Ti, plagioclase for Eu and epidote for LREE, Th and U (e.g., Rubatto, 2002; Whitehouse and Platt, 2003; Wu and Zheng, 2004). For example, the activity of garnet during zircon growth in metamorphic systems, which dictates the garnet effect on zircon REE partition (Rubatto, 2002; Whitehouse and Platt, 2003). If both garnet and zircon occur as the products of metamorphic and peritectic reactions, the zircon is characterized by the flat HREE pattern due to the preferential partition of HREE into the garnet. Similarly, if garnet is a residual phase during these reactions, due to its high HREE, the product zircon would also exhibit depletion in HREE. In contrast, if garnet is involved in these reaction as reactant, the product zircon would be enriched in HREE. (4) The feature of formation enviroment (Rubatto, 2002; Wu and Zheng, 2004). The openness and closure of metamorphic and peritectic systems would affect zircon composition. Growth velocity of grown zircon was also suggested as a controlling factor (Vavra et al., 1999).

Many laboratory experiments and field-based studies have indicated that aqueous solutions can only dissolve fluid-mobile incompatible trace elements such as LILE, U, Sr and Pb and thus they cannot transport considerable amounts of melt-mobile incompatible trace elements such as LREE and Th (e.g., Hermann et al., 2006a; Zheng et al., 2011a). In contrast, hydrous melts can dissolve higher contents of solutes such as SiO2, Al2O3, CaO, NaO, Th, U, LILE and LREE (Zheng and Hermann, 2014, and references therein). As a result, aqueous solutions and hydrous melts can be distinguished by characteristic trace element contents and specially their ratios. For example, because U is more mobile than Th at temperatures below the wet solidus of crustal rocks, aqueous solutions are characterized by low Th contents and thus low Th/U ratios (e.g., Rollinson and Windley, 1980; Rowley et al., 1997). On the other hand, Th becomes mobile at temperatures above the wet solidus due to the breakdown of Th-rich minerals such as allanite and monazite (Hermann, 2002). Thus, the hydrous melts generally exhibit both higher Th and U contents and higher Th/U ratios than the aqueous solutions. It is expected that the newly grown zircon from aqueous solutions and hydrous melts can be distinguished by its contents and ratios of some trace elements, especially Th and U contents and their ratios. The further evolution of hydrous melts into magmatic melts leads to the common observation that the magmatic zircon shows uniformly higher Th/U ratios than both metamorphic and peritectic zircons. The Ti-in-zircon thermometer is widely used to link temperature to time in metamorphic rocks (e.g., Watson et al., 2006; Ferry and Watson, 2007; Tomkins et al., 2007). Due to the limited diffusivity of the 4+ cations within their lattices as constrained by the experimental diffusion study of Cherniak and Watson (2007), the Ti-in-zircon thermometer is particularly relevant to high-T and ultrahigh-T geological processes (e.g., Kelsey and Hand, 2015). The Ti-in-zircon thermometer not only depends on the Ti contents of zircon, but also depend on the activities of SiO2, and TiO2 as well as pressure (e.g., Watson et al., 2006; Ferry and Watson, 2007; Ferris et al., 2008). Although zircon Ti contents can be well measured, the other parameters especially for pressure cannot be well constrained. As shown by studies from the Dabie-Sulu orogenic belt, the Ti contents of zircon domains formed during a metamorphic event vary significantly (quite common in all types of metamorphic rocks), resulting in a very large variation in Ti-in-zircon temperatures (e.g., Zong et al., 2010; Chen et al., 2013b; Xu et al., 2013; Liu et al., 2015). In addition, nearly all metamorphic zircon domains have similar Ti-in-zircon temperatures despite their formation under different metamorphic P-T conditions. In this regard, it is critical to determine the saturation of Ti, Zr and Si in the target rock, pressure and the possible later resetting of Ti in zircon when applying the Ti-in-zircon thermometer (e.g., Ferry and Watson, 2007; Tomkins et al., 2007; Kelsey and Hand, 2015; Kohn, et al., 2015; Taylor et al., 2016). While the saturation can be demonstrated petrologically by the presence of rutile and quartz, the pressure can only be constrained by index

mineral and barometry. Studies that report Ti-in-zircon temperatures without having established the saturating criteria should be viewed with considerable caution. The Lu–Hf isotope system of a rock can be divided into two subsystems: zircon with low Lu/Hf ratios but matrix (mainly other REE-rich minerals such as garnet, epidote, monazite and apatite) with high Lu/ Hf ratios (Amelin et al., 2000; Kinny and Maas, 2003; Zheng et al., 2005b). In a closed system, zircon and matrix evolve in different ways. After a long time, the zircon has a low

176

Hf/177Hf ratio, whereas the matrix has a high

176

Hf/177Hf ratio. Due to the breakdown or

recrystallization of REE-rich minerals, newly grown zircon would show variably elevated 176

Hf/177Hf ratios (Amelin et al., 2000; Zheng et al., 2005b). For zircon formed via the dissolution

reprecipitation of pre-existing zircon in a closed system, on the other hand, it can inherit the rock Hf isotope composition by weighted meaning of Hf element and isotopes between protolith zircon and matrix (Zheng et al., 2005b, 2006a; Flowerdew et al., 2006). In this regard, the more matrices were involved, the higher

176

Hf/177Hf ratios the newly grown zircons have. If the more protolith zircons

were involved, the newly grown zircons have 176Hf/177Hf ratios that are still higher than, but close to, those for the protolith zircon. In either case, the newly grown zircon in a closed system has similar or elevated

176

Hf/177Hf ratios relative to the protolith zircon. In an open system, on the other hand,

the newly grown zircon would have significantly variable and different Lu–Hf isotope compositions from the inherited zircon (Zheng et al., 2005b; Wu et al., 2006a). Therefore, the Hf isotope composition of newly grown zircon can provide insights into its formation mechanism and system nature (open or closed).

3. Geological setting Because the present review mainly focuses on zircons from UHP metamorphic rocks in the Dabie-Sulu orogenic belt, east-central China, it is necessary to outline its geological setting in order to better understand metamorphic zirconology. There are two UHP terranes in this orogenic belt, which were synchronously generated by northward subduction of the South China Block beneath the North China Block in the Triassic (e.g., Li et al., 1999; Zheng et al., 2003a). They are separated into Dabie and Sulu orogens by 500 km of left-lateral strike-slip displacement along the Tan-Lu fault (Fig. 1). The two orogens are composed of several fault-bounded HP and UHP metamorphic units in association with Mesozoic magmatic rocks and sedimentary covers (e.g., Zheng et al., 2005a; Xu et al., 2006). The UHP metamorphism is demonstrated by the occurrence of coesite and microdiamond in metamorphic minerals from eclogite and gneiss in the Dabie-Sulu orogenic belt (e.g., Okay et al., 1989; Xu et al., 1992; Liu and Liou, 2011). Available observations indicate that this collisional orogen contains one of the largest and best-exposed UHP metamorphic terranes on Earth (e.g., Liou et al., 2009; Zheng, 2012).

The Dabie-Sulu UHP metamorphic rocks are mainly composed of granitic orthogneiss, with minor proportions of other rock types such as paragneiss, eclogite, peridotite, marble and quartzite. In terms of the differences in metamorphic P-T conditions, eclogite-bearing slices are subdivided into three UHP zones: low-T/UHP, mid-T/UHP and high-T/UHP zones (e.g., Zheng et al., 2005a; Xu et al., 2006; Liu and Li, 2008). The maximum pressure estimates lie in the diamond stability field of >120 km, and the maximum temperatures vary from 730 to 850°C depending on P-T paths of different UHP slices during continental collision. All the three UHP zones underwent amphibolite-facies retrogression during exhumation, and the high-T/UHP zone also underwent widespread migmatization and magmatism at the postcollisional stage (e.g., Wu et al., 2007; Zhao and Zheng, 2009; Chen et al., 2015). The high-T/UHP zone is characterized by overprinting of HP granulite-facies metamorphism over the UHP fabrics during the early exhumation. The maximum temperatures are obtained at the stage of early exhumation, named as "hot" exhumation (Zhao et al., 2007a; Zheng et al., 2011a). Quartz veins and leucosomes within UHP rocks are widely found in the Dabie-Sulu orogenic belt, indicating the presence of fluid flow and partial melting during continental collision (e.g., Zheng et al., 2007b, 2011a; Sheng et al., 2012; Chen et al., 2013a, 2013b; Li et al., 2014, 2016a). The crustal rocks in the Dabie-Sulu orogenic belt underwent UHP eclogite-facies metamorphism at 225-240 Ma, quartz eclogite-facies metamorphism at 215-225 Ma and amphibolite-facies metamorphism at 215-205 Ma (e.g., Zheng et al., 2009). The protoliths of most UHP metaigneous rocks are of middle Neoproterozoic age (mainly at 740 to 780 Ma), whereas some of UHP metasedimentary rocks have Archean and Paleoproterozoic protoliths (Zheng, 2008). Rock-forming minerals in the UHP metaigneous rocks exhibit negative to low δ18O values (e.g., Zheng et al., 2003a, 2009; Chen et al., 2007; Tang et al., 2008a, 2008b). The U-Pb dating of low to negative δ18O zircons demonstrates that the protoliths of metaigneous rocks were strongly altered by unusually

18

O-depleted surface water at high temperatures in the middle Neoproterozoic (e.g.,

Rumble et al., 2002; Zheng et al., 2004, 2009; Chen et al., 2011; He et al., 2016).

4. Newly grown zircons during subduction-zone metamorphism 4.1 Metamorphic zircon The metamorphic zircon is generally produced through subsolidus dehydration reactions between silicate and accessory minerals with the presence of aqueous solutions in the products. Previous studies often ascribe the origin of metamorphic zircon to direct precipitation from metamorphic fluids (e.g., Rubatto and Hermann, 2003; Dubinska et al., 2004; Zheng et al., 2007b). However, the metamorphic fluids produced at subsolidus conditions are highly undersaturated with

HFSE and thus not able to precipitate zircon (e.g., Zheng and Hermann, 2014, and references therein). For this reason, we prefer to interpret the metamorphic zircon as the product of mineral reactions during metamorphic dehydration below the wet solidus of crustal rocks. Metamorphic zircons in UHP metamorphic rocks commonly occur either as small anhedral grains or as overgrowths around relict cores of protolith zircon (Fig. 2). These domains exhibit no zoning, cloudy zoning and weak zoning in CL images (e.g., Zheng et al., 2004, 2005b, 2006a; Wu et al., 2006a; Xia et al., 2009; Chen et al., 2010; Li et al., 2013) and concordant U-Pb metamorphic ages. Metamorphic zircons in UHP metamorphic rocks from the Dabie-Sulu orogenic belt generally exhibit lower REE, Th and HFSE contents but higher Hf contents than the protolith zircons of magmatic origin (Chen et al., 2010; Liu and Liou, 2011). However, REE patterns for the metamorphic zircons are variable (Fig. 3). Eclogite-facies metamorphic zircons exhibit relative lack of negative Eu anomalies due to the breakdown of plagioclase under eclogite-facies conditions. Such zircons in eclogites are generally characterized by flat HREE patterns with low (Lu/Gd)N ratios because of the concurrent growth of metamorphic garnet during eclogitization. In contrast, the majority of metamorphic zircons in eclogite-facies felsic gneisses show steep HREE patterns with high (Lu/Gd)N ratios (Fig. 3d-f), leaving only the minority of metamorphic zircons in these gneisses to exhibit the flat REE patterns similar to those in mafic eclogites (Fig. 3b). This may reflect the garnet effect in the metamorphic system of felsic rocks. While the flat HREE patterns indicate the concurrent growth of metamorphic garnet during dehydration reaction in the felsic rocks, the steep HREE patterns suggest the relative lack of garnet growth during the eclogite-facies metamorphism of felsic rocks (Fig. 3b, d-f). Therefore, the geochemical composition of metamorphic zircons is related not only to the metamorphic P-T condition but also to the composition of host rocks In principle, amphibolite-facies metamorphic zircons would exhibit steep REE patterns and marked negative Eu anomalies. This is indeed observed in some UHP metamorphic rocks from several UHP terranes (e.g., Hermann et al., 2001). However, metamorphic zircons in amphibolite-facies retrograde eclogites from the Dabie-Sulu orogenic belt often exhibit flat REE patterns similar to those in eclogites (Fig. 3d and Liu and Liou, 2011). Mineral inclusions in these zircon domains do indicate their growth under amphibolite-facies conditions. In this regard, the trace element composition of metamorphic zircons may suggest their growth in association with recrystallization of garnet but the limited crystallization of plagioclase. On the other hand, amphiboles formed under amphibolite-facies conditions would inherit the trace element composition from their precursor garnet/omphacite (Sassi et al., 2000), implying the limited release of HREE from these precursors.

Metamorphic zircons also occur as large euhedral crystals in quartz veins within UHP eclogites (Fig. 2), recording the local focus of aqueous solutions during continental collision (Zheng et al., 2007b; Wu et al., 2009; Chen et al., 2012; Sheng et al., 2012, 2013). Although aqueous solutions have the great capacity to dissolve and transport silica and fluid-mobile incompatible trace elements such as LILE, they have very limited capacity to dissolve and transport REE and HFSE (Zheng and Hermann, 2014). In this regard, the occurrence of metamorphic zircons in the quartz veins does not mean the chemical transport of element Zr by the aqueous solutions. Instead, it suggests the growth of metamorphic zircons via locally focused dehydration reactions at decompressional conditions. The occurrence of fluid inclusions in these zircons (Fig. 2f) confirms their origin from dehydration reaction under subsolidus conditions. As such, the metamorphic zircons were carried by the aqueous solutions from their growth sites into the veining sites. This requires the generation of abundant aqueous solutions through the subsolidus dehydration reactions, allowing for sufficient growth of metamorphic zircons as the large euhedral crystals. If only very minor amounts of the aqueous solutions were produced by metamorphic dehydration, the newly grown zircon would occur as small anhedral grains as commonly observed in UHP eclogites. Despite the big difference in zircon morphology and thus fluid abundances, there are no significant differences in trace element and Lu-Hf isotope compositions between the two kinds of the metamorphic zircons in the UHP metamorphic rocks (e.g., Chen et al., 2012a; Sheng et al., 2012, 2013). This indicates the origin of aqueous solutions from the metamorphic dehydration of UHP rocks themselves in both cases. Most zircons exhibit flat REE patterns, indicating their growth in association with the concurrent growth of metamorphic garnet under eclogite-facies conditions. This is confirmed by the mineral inclusions of Grt and Omp in these zircon domains (Fig. 2f). Rare zircons exhibit steep REE patterns with negative Eu anomalies, indicating their growth away from the garnet growth (Fig. 3c). Relict cores of the protolith zircons are common in the quartz veins (Fig. 2g; Chen et al., 2012a; Sheng et al., 2012, 2013), suggesting the physical transport of protolith zircons by the metamorphic fluids. This explains the occurrence of both old relict and newly grown zircons in metamorphic veins and orogenic peridotites. In view of the difference in element abundances between metamorphic zircons and protolith magmatic zircons, the metamorphic zircons are expected to show the difference in some element ratios from the protolith magmatic zircons. The metamorphic zircons generally have lower Th/U ratios (generally <0.1), higher Eu/Eu* ratios, higher Hf/Y ratios and lower (Lu/Gd) N ratios than the protolith magmatic zircons (Figs. 3 and 4). The metamorphic zircons exhibit low and elevated

176

176

Lu/177Hf ratios

Hf/177Hf ratios (Fig. 5a). They may have O isotope compositions similar to, or

different from, the protolith magmatic zircons depending on the O isotope composition of metamorphic minerals and thus the protolith nature of host rocks. For example, metamorphic

zircons in UHP eclogites from Qinglongshan in the Sulu orogen exhibit low or negative δ18O values (Fig. 5b), because the protoliths of UHP metabasites there underwent meteoric hydrothermal alteration in the middle Neoproterozoic (Zheng et al., 2003; Chen et al., 2011). Metamorphic zircons may exhibit higher 18O values than normal mantle values of 5.3±0.3‰ (Valley et al., 1998) if they are produced by metamorphism from supracrustal rocks, otherwise they show lower 18O values than normal values if they are generated by metamorphism from intracrustal rocks that underwent low 18O fluid alteration. Two major episodes of metamorphic zircons have been identified during continental subduction-zone metamorphism in the Dabie-Sulu orogenic belt: one at the prograde HP-UHP transition stage and the other at the retrograde UHP-HP transition stage (e.g., Zheng et al., 2005b; Wu et al., 2006; Zheng, 2009; Xia et al., 2013). This is indicated by the two episodes of zircon growth at c. 240 Ma and c. 225 Ma, respectively, in the UHP eclogites (Zheng et al., 2009). During the prograde HP-UHP metamorphism, aqueous solutions were produced by the breakdown of hydrous HP minerals such as paragonite, amphibole, lawsonite and zoisite. The breakdown of paragonite may be common during the transformation of blueschist to eclogite (Li et al., 2004), and the breakdown of amphibole may be common during the transformation of amphibolite to eclogite (Liu and Ye, 2004). During the breakdown of amphibole, ilmenite was also involved. Ilmenite has been considered as a significant reservoir of Zr in mafic lithology, and its breakdown is especially important to zircon growth (Bingen et al., 2004). Thus, these dehydration reactions may release SiO2 and probably ZrO2 and thus lead to the growth of metamorphic zircons. In addition, zircon is a major reservoir of Zr in metamorphic rocks (e.g., Bea et al., 2006). The Lu-Hf isotope composition of metamorphic zircons suggests that the protolith zircon was also involved in the zircon growth. During exhumation, the breakdown of hydrous UHP minerals such as lawsonite, phengite and zoisite can provide considerable amounts of aqueous solutions for veining, metasomatism and even anatexis (Li et al., 2004; Zheng et al., 2007b). Furthermore, decompression exsolution of structural hydroxyl and molecular water from nominally anhydrous UHP minerals such as garnet, omphacite and rutile is also significant in UHP metamorphic rocks (Zheng et al., 1999, 2003a; Chen et al., 2007b; Zheng, 2009), providing an important source of retrograde fluids for hydration of the UHP rocks during exhumation. The Zr solubility in OH-rich fluids an be enhan e even at o er te

eratures Dubińska et a ., 2004), so that the aqueous solutions generated by local sinking of the

liberated hydroxyls are highly alkaline and thus oxidized (Zheng et al., 2007b). It is an efficient agent to dissolve and transport Zr and Si from relevant minerals for the growth of metamorphic zircons (Zheng, 2009). In addition, rutile is an important reservoir of Zr in eclogites (Sassi et al., 2000; Zack et al., 2002). Rutile may release a lot of Zr due to retrograde recrystallization of

eclogites under HP eclogite-facies conditions.

4.2 Peritectic zircon Partial melting of UHP metamorphic rocks is common not only in the Dabie-Sulu orogenic belt but also elsewhere in the other collisional orogens (e.g., Zheng et al., 2011a; Gao et al., 2012; Liu et al., 2012; Chen et al., 2013a, 2013b; Gordon et al., 2013; Li et al., 2013, 2014, 2016a; Gilotti et al., 2014). It occurs at varying scales from thin sections to outcrops, forming different types of migmatites. Anatectic melts were produced through peritectic reactions above the wet solidus of crustal rocks. Alkalic to granitic dykes and intrusions of synexhumation age are highly evolved products of the anatectic melts (Yang et al., 2005; Liu et al., 2009c; Zhao et al., 2012). Peritectic zircons primarily occur in migmatitic rocks and pegmatites (Wallis et al., 2005; Liu et al., 2009a, 2010, 2012; Zong et al., 2010; Zeng et al., 2011; Chen et al., 2013a, 2013b; Xu et al., 2013), whereas magmatic zircons typically occur in synexhumation magmatic intrusions (Zhao et al., 2012). Because of the low Zr solubility in anatectic melts, zircon is rarely crystallized from the anatectic melts unless the anatectic melts have experienced given degrees of evolution with fractional crystallization of Zr-poor minerals from the melts. Like magmatic zircons, anatectic zircons can only be crystallized from the anatectic melts in which the local Zr oversaturation was achieved during their evolution. While peritectic zircons would generally grow above the wet solidus with increasing temperature, the anatectic zircons would only grow along a temperature decrease path with temperartures on the wet solidus (Fig. 6). In this regard, peritectic zircons would generally occur in migmatitic rocks and pegmatites, whereas anatectic zircons can only occur in the leucosome and pegmatite which have underwent large extent of fractional crystallization along the temperature decrease path.
Magmatic zircons generally show oscillatory zoning, high REE contents and steep HREE patterns, high Th/U (>0.1) and Lu/Hf ratios, positive Ce and negative Eu anomalies. In contrast, peritectic zircons usually exhibit weak zonation (no zoning to cloudy zoning, occasionally oscillatory zoning) (Fig. 7) and low Th/U ratios (generally <0.1), containing multiphase solid inclusions of Qtz±Ab±Kfs±Ap. Compared to the protolith zircons of magmatic origin, the peritectic zircons often exhibits similar or lower REE contents (Fig. 4), variable REE patterns (Fig. 8), lower or similar Th and U contents, higher Hf contents, similar or higher Nb and Ta contents, similar Ce and Eu anomalies, similar or lower Nb/Ta ratios, similar or lower Lu/Hf ratios but higher 176

Hf/177Hf ratios (Figs. 4 and 5). The U-Pb dating indicates that the anatexis of UHP metamorphic

rocks primarily takes place at high-T/HP granulite-facies conditions during the early exhumation.

The partial melting of UHP metamorphic rocks are generally ascribed to the HP granulite-facies overprinting during “hot” exhumation of the deeply subducted continental crust (Zheng et al., 2011a). It is expected that zircon grown under granulite-facies conditions would exhibit negative Eu anomalies and flat REE patterns due to the stability of both plagioclase and garnet (Whitehouse and Platt, 2003). However, peritectic zircons from the North Qaidam orogen in northern Tibet show large variations in mineragraphy and trace element composition (e.g., Chen et al., 2012b; Yu et al., 2014, 2015; Zhang et al., 2015). Peritectic zircons in HP granulites from Dulan in North Qaidam exhibit no zoning, irregular zoning, well-developed sector zoning or oscillatory zoning, variably high Th/U ratios (generally >0.1), weakly negative Eu anomalies and relatively flat REE patterns (Fig. 9b) and contain inclusions of Grt + Qtz + Cpx + Pl + Rt + Ap (Yu et al., 2014). In contrast, peritectic zircons from leocosomes within UHP rocks from Xitieshan and Lüliangshan in North Qaidam exhibit weak zoning or oscillatory zoning, low Th/U ratios <0.1 and flat to steep HREE patterns (Fig. 9b), and contain mineral inclusions of quartz + feldspar + apatite (Chen et al., 2012b; Yu et al., 2015; Zhang et al., 2015). Despite the large difference of trace elements, these peritectic zircons have no large difference in their U-Pb ages, suggesting their growth under similar P-T conditions during continental collision (Zhang et al., 2015). The part of peritectic zircons shows flat HREE patterns and is generally associated with a significant amount of garnet in the leucosome. This indicates the concurrent growth of peritectic garnet and thus the active presence of garnet, which is confirmed by identification of mineral inclusions in these peritectic zircons. The other part of peritectic zircons exhibits steep HREE patterns and occurs in migmatite or leucosome, in which HP/UHP metamorphic zircons also exhibit steep HREE patterns, consistent with the relative lack of garnet in the host rocks. The significant depletion of LREE in peritectic zircons is usually ascribed to the presence of LREE-rich minerals such as allanite, monazite and epidote during zircon growth. However, peritectic zircons from the leucosomes without these minerals also exhibit marked depletion in LREE. In this regard, the composition of peritectic zircons is primarily controlled not only by the composition of reactants but also by that of other peritectic minerals during crustal anatexis. In the Western Gneiss Region of Norway, peritectic zircons with concordant U-Pb ages of 410-406 Ma, coeval with peak or near-peak UHP metamorphism, exhibit low Th/U ratios and flat HREE patterns with lack of negative Eu anomalies. This indicates their growth in association with the concurrent growth of peritectic garnet. The other zircons, including those from crosscutting pegmatite and showing younger U-Pb ages consistent with dates for amphibolite-facies retrogression, exhibit steep HREE patterns and negative Eu anomalies (Gordon et al., 2013). This suggests that their growth is associated with the plagioclase crystallization but with the garnet inertness. The latter observation indicates that peritectic zircons record the transition from UHP

eclogite-facies (garnet-active) to lower P (plagioclase-active) conditions. In this regard, the composition of peritectic zircons is related not only to the metamorphic conditions but also to the activity of coexisting minerals. The composition of peritectic zircons produced at all stages of exhumation changes with the P-T conditions (Fig. 9a), which has also been found in the North-East Greenland (Gilotti et al., 2014).
Usually, the low abundances of trace elements in the metamorphic zircon relative to the magmatic zircon are ascribed to the effect of concurrently grown or recrystallized, specific minerals such as garnet, epidote and rutile. However, the experimental study of Rubatto and Hermann (2007b) for REE partition between zircon, melt and garnet indicates that REE (especially HREE) were preferentially incorporated into zircon rather than garnet when both minerals crystallized from melts at temperatures below 850 °C. Such a temperature effect is of critical importance to petrogenetic interpretation of peritectic, anatectic and magmatic zircons. Thermodynamic equilibrium partition of trace elements between coexisting minerals is also independent of relative proportions between different minerals, implying that even if a large amount of garnet is present during zircon growth or recrystallization in a closed system of anatectic/magmatic melts at the temperatures below 850 °C, the resultant zircon does not become HREE-depleted relative to the coexisting garnet. In this regard, it is necessary to distinguish the magmatic minerals from peritectic minerals in their compositions. Because different minerals may be involved in partial melting via peritectic reactions, the variation of trace element abundances in peritectic zircons may be caused by variable involvement of different reactants in the crustal anatexis. Therefore, it is critical to determine the trace element partition between peritectic zircon and other peritectic minerals at different P-T conditions. Most peritectic, anatectic and magmatic zircons of synexhumation growth exhibit higher 176

Hf/177Hf ratios than the relict cores of protolith zircon (Liu et al., 2009c, 2009d, 2010a; Zhao et

al., 2012; Chen et al., 2013b), suggesting that the anatexis is associated with decomposition or recrystallization of high Lu/Hf minerals in the UHP metamorphic rocks. Although the peritectic, anatectic and magmatic zircons from migmatites, pegmatites and plutonic rocks show significant differences in element contents and ratios, most of them show similar Hf isotopic compositions and thus suggest the similar origins. Therefore, they record the partial melting of UHP metamorphic rock with petrological evolution from the anatectic melt to the magmatic melt during the exhumation of deeply subducted continental crust. Sometimes different generations of peritectic, anatectic and magmatic zircons occur in the same samples (Liu et al., 2009c, 2009d, 2010a, 2012; Chen et al., 2013b). Nevertheless, they often exhibit different Hf and O isotope compositions (Liu et al., 2009d; Chen et al., 2013b), implying that their growth is associated with different origins of

anatectic and magmatic melts. Anatexis of the UHP metamorphic rocks mainly occurs during the exhumation of deeply subducted continental crust (Zheng et al., 2011a), corresponding to the stage of HP granulite-facies overprinting. In fact, the anatexis would have started at the late UHP eclogite-facies stage during the initial exhumation. This has been demonstrated by the following observations: (1) coesite inclusions occur in some domains of the peritectic zircon from the Dabie-Sulu orogenic belt (Chen et al., 2013a, 2013b); (2) leucosomes from the Western Gneiss Region show similar compositions to the experimental melts produced at UHP conditions (Labbrouse et al., 2011); (3) peritectic zircons with concordant U-Pb ages of 410-406 Ma, coeval with peak or near-peak UHP metamorphism, have been found in some leucosomes within UHP rocks in the Western Gneiss Region (Gordon et al., 2013). Furthermore, crustal anataxis in the Kokchetav and Bohemian UHP terranes has temperatures as high as 1000C in the diamond stability field (e.g., Shatsky et al., 1999; Kotkova et al., 2016; Stepanov et al., 2016). The resulted peritectic zircons exhibit sector, fir-tree or oscillatory zoning, high Th/U ratios of >0.1, flat to steep HREE patterns with weakly negative or no Eu anomalies (Fig. 9c). This suggests that the breakdown of Th-rich hydrous minerals is significant whereas the metamorphic garnet is partially activated during the anatexis. For the anataxis during exhumation, with decreasing temperatures, anatectic melts may undergo large extent of fractional crystallization to achieve the local Zr oversaturation. In this regard, anatectic zircons would also grow besides peritectic zircons in felsic veins, especially in the pegmatites which generally have the largest extent of fractional crystallization. This can explain the observed large variation in zircon element compositions and U-Pb ages. However, these anatectic zicons do not show large differences in trace element composition from the peritectic zircons (Figs. 4 and 7). The composition of anatectic zircons are controlled by the composition of evolved melts and concurrent growth minerals. On the other hand, UHP metamorphic rocks may suffer partial melting during their final subduction into the UHP regime, which has been documented in the Dabie-Sulu orogenic belt (Xia et al., 2013; Li et al., 2014, 2016a). Peritectic zircons produced by this stage of crustal anatexis generally exhibit dark luminescent and no zoning in CL images, variable Th/U ratios, flat to steep HREE patterns with variable Eu anomalies (Fig. 9d). Coesite and multiple phase solid inclusions were found in some of these peritectic zircons (Li et al., 2014, 2016a), suggesting at least part of them growth at UHP conditions. Nevertheless, the occurrence of flat to steep HREE patterns in these peritectic zircons suggests discontinuous growth of the peritectic garnet in this period. The UHP metamorphic rocks, especially for those in the high-T/UHP zone from the Dabie-Sulu orogenic belt, also underwent partial melting in the postcollisional stage at pressures below 1.0 GPa, resulting in extensive migmatization and bimodal magmatism (e.g., Zhao and Zheng, 2009; Chen et

al., 2015a). Peritectic zircons of this stage exhibit unzoning, weak zoning and oscillatory zoning, variable Th/U ratios (from <0.1 to >1) and steep HREE patterns (Chen et al., 2015). The steep HREE patterns are consistent with the low-P anatexis, whereas the variable Th/U ratios may be related to the differential involvement of Th- and U- rich minerals such as allanite, epidote and monazite. In summary, peritectic zircons in UHP metamorphic rocks can grow at all stages of continental collision from the final subduction via the peak UHP metamorphism to the initial exhumation, with additional growth at the postcollisional stage. As a consequence, peritectic zircons may exhibit large variations in CL structure from unzoning to oscillatory zoning, in Th/U ratios from <0.1 to >1.0, in REE contents from low to high, in HREE patterns from flat to steep, and in Eu anomalies from no to negative. The variations in the composition of peritectic zircons are primarily related not only to anatectic P-T conditions but also to product mineral species in peritectic reactions. In addition, it is also influenced by the activity of preexisting minerals, either reactant or product. As a consequence, the peritectic zircons may exhibit variable compositions between metamorphic and magmatic zircons. Despite the large variations in element contents and ratios, they generally exhibit higher

176

Hf/177Hf ratios than the inherited magmatic zircons (Fig. 5a), indicating the involvement

of HFSE- and HREE-rich minerals other than zircon in the crustal anatexis.

4.3 Distinction between metamorphic and peritectic zircon Both metamorphic and peritectic zircons show large variations in CL imaging, Th/U ratios, and trace element contents and ratios, depending on P-T conditions and minerals involved in breakdown, recrystallization and growth. This often results in similar geochemical compositions between some metamorphic and peritectic zircons (Fig. 4). Nevertheless, the two types of zircons can be distinguished by a study of zirconology, including CL images for mineragraphy (structure and morphology), U-Pb ages, trace element contents and ratios, stable and radiogenic isotope compositions, and geothermobarometries (Table 1). Lots of important information can be obtained by such a study for zircons from the same samples. Both metamorphic and peritectic zircons exhibit variable CL images. Although most peritectic zircons generally show no zoning in CL images similar to common metamorphic zircons, oscillatory zoning, which is typical for magmatic zircons, often occurs in the peritectic zircons but is rarely observed in metamorphic zircons (e.g., Rubatto and Hermann, 2003). This difference is evident in the peritectic zircons in leucosome and pegamatite relative to the metamorphic zircons in quartz veins (Figs. 2 and 6). Although metamorphic minerals are common in both types of zircons, fluid inclusions only occur in metamorphic zircons (e.g., Wu et al., 2009) whereas melt inclusions

(e.g., multiphase solid inclusions) only occur in peritectic zircons (e.g., Li et al., 2014, 2016a; Chen et al., 2015). For example, multiphase solid inclusions in peritectic zircons are common in migmatites from the high-T/UHP metamorphic zones in North Dabie and Weihai of the Dabie-Sulu orogenic belt, consistent with their growth via peritectic reactions during partial melting of UHP metamorphic rocks (Chen et al., 2015; Li et al., 2016a). In contrast, fluid inclusions in metamorphic zircons are common in quartz veins inside the UHP slices of the Dabie orogen (Wu et al., 2009). Because U is more mobile than Th in aqueous solutions, metamorphic fluids produced at temperatures below the wet solidus of crustal rocks are characterized by low Th contents and thus low Th/U ratios. In contrast, Th becomes mobile during crustal anatexis due to the breakdown of Th-rich accessory minerals such as allanite and monazite (Hermann, 2002). Thus, anatectic melts commonly exhibit both higher Th and U contents and higher Th/U ratios than aqueous solutions. It is expected that the metamorphic zircons grown through subsolidus dehydration reactions generally have low Th contents but high U contents and thus low Th/U ratios of <0.1, whereas the peritectic zircons grown through supersolidus peritectic reactions may have high Th and U contents and thus high Th/U ratios of >0.1. Indeed, metamorphic zircons from the Dabie-Sulu orogenic belt generally exhibit low Th but high U contents and thus low Th/U ratios of <0.1. Nevertheless, most peritectic zircons there also exhibit low Th/U ratios of <0.1, whereas high Th/U ratios generally only occur in peritectic zircons but are absent in metamorphic zircons. The peritectic zircons with high Th/U ratios were usually observed in allanite-bearing gneisses, where they exhibit similar REE patterns to magmatic zircons and high Nb and Ta contents (Liu et al., 2009b). In this regard, the low Th/U ratios of <0.1 cannot be used to distinguish metamorphic zircons from periectic zircons, but high Th/U ratios certainly occur in the peritectic zircons rather than in the metamorphic zircons. On the other hand, the peritectic zircons with low Th/U ratios are generally ascribed to the inactive presence of Th- and U-rich minerals such as allanite and monazite during crustal anatexis. However, peritectic zircons without these minerals in host rocks also exhibit low Th/U ratios. As such, the presence of Th-rich minerals and their involvement in crustal anataxis are substantial to the high Th/U ratios for peritectic zircons. Because of the low Zr solubility in anatectic melts, zircon is rarely crystallized from anatectic melts unless the anatectic melts have experienced given degrees of evolution with fractional crystallization of Zr-poor minerals from the melts. Therefore, highly evolved leucosomes in migmatites are expected to contain anatectic zircons with geochemical compositions between peritectic and magmatic zircons. Temperature is also a critical factor to dictate the occurrence of dehydration reactions below or above the wet solidus of crustal rocks. Dehydration melting of crustal rocks commonly takes place at temperature of 750–800 °C or higher (Brown, 2010; Clemens, 2006), but this process usually proceeds in the postcollisional stage at low to medium pressures. The wet solidus of granitic rocks

is generally above 650–700 °C (e.g., Hermann et al., 2006a; Zheng et al., 2011a). When the temperature is below the wet solidus, aqueous solutions are produced with possible dissolution of fluid-mobile incompatible trace elements such as LILE, U, Pb and Sr. At the temperature above the wet solidus of granitic rocks, hydrous melts are generated with possible dissolution of melt-mobile incompatible trace elements such as LILE, LREE, U, Th, Pb and Sr (e.g., Reed et al., 2000; Zajacz et al., 2008; Zheng and Hermann, 2014). Although metamorphic zircons are produced via the subsolidus dehydration reactions rather than direct growth from the metamorphic fluids, they may approach geochemical equilibrium with each other in the fluid-mobile incompatible trace element composition. Likewise, peritectic zircons are generated via the supersolidus peritectic reactions rather than direct crystallization from the anatectic melts, but they also approach geochemical equilibrium with each other in the melt-mobile incompatible trace element composition. Whole-rock Zr contents can be used to estimate the zircon saturation temperature during partial melting of felsic rocks (Watson and Harrison, 2003; Boehnke et al., 2011), but the presence of relict protolith zircons makes this temperature estimate maximum relative to the real anatectic temperature (e.g., Hanchar and Watson, 2003; Miller et al., 2003). On the other hand, the Ti-in-zircon thermometry provides a temperature estimate of zircon crystallization from hydrous melts that have achieved Zr saturation if the activities of both SiO2 and TiO2 as well as pressure can be well constrained (Ferry and Watson, 2007). In this regard, the Ti-in-zircon temperatures generally give the minimum estimate of zircon crystallization from anatectic and magmatic melts. Comparing such temperatures with the wet solidus of crustal rocks, it is possible to recognize the presence of hydrous melts and even anatectic behaviors and thus to distinguish the peritectic zircons from metamorphic zircons (Li et al., 2013). Studies of experimental and field-based geochemistry indicate that the stability of specific accessory minerals during metamorphic dehydration and partial melting is substantial to the abundances of fluid/melt-mobile incompatible trace elements in metamorphic and peritectic minerals (e.g., Hermann et al., 2006; Zheng et al., 2011a; Zheng and Hermann, 2014). For example, garnet is a major host for HREE and Y, rutile for HFSE, and monazite, epidote and allanite for LREE and Th. As a consequence, metamorphic zircons generally exhibit lower Y contents and thus higher Hf/Y ratios than peritectic zircons. Similar to Th/U ratios, (Lu/Gd)N ratios can also be used to distinguish between metamorphic and peritectic zircons. With increasing temperature, the solubility of LREE in hydrous melts increases more rapidly than the solubility of HREE (e.g., Kessel et al., 2005). This may be ascribed to the breakdown of LREE-rich minerals such as allanite and epidote. As a result, peritectic zircons may have lower (Lu/Gd)N ratios than metamorphic zircons at elevated temperatures. Nb/Ta ratios may be significantly fractionated between hydrous melt and aqueous solution under certain condition (Pettke et al., 2005). Experimental studies also

confirm this fractionation by finding higher Nb/Ta ratios for hydrous melts (e.g., Kessel et al., 2005). Therefore, metamorphic zircon may also be distinguished from anatectic zircons by lower Nb/Ta ratios (Li et al., 2013). However, the trace element compositions of newly grown zircons depend not only on the property of subduction-zone fluids, but also on the P-T conditions of dehydration reactions. Sometimes this results in overlapped trace element compositions between metamorphic and peritectic zircons (Fig. 4). As shown by the results from UHP metamorphic rocks in the Dabie-Sulu orogenic belt, the metamorphic zircons generally exhibit flat HREE patterns with relative lack of negative Eu anomalies, and some of the peritectic zircons show similar REE patterns. In this regard, the generally used REE patterns and contents cannot be used to distinguish the metamorphic zircons from the peritectic zircons. This is also true for the other elements such as Nb and Ta. Nevertheless, when the specific mineral effect can be evaluated, the trace element contents can be used to distinguish between different types of zircon, especially for zircons from the same samples. For example, two distinct zircon domains in UHP metagranites from Taohang in the Sulu orogen exhibit different trace element compositions (Fig. 10), indicating their growth during a transition from subsolidus dehydration reactions to supersolidus peritectic reactions along a temperature-increasing path. In contrast, zircon domains in pegmatite veins within the UHP metagranites record their growth during a transition from hydrous melts to aqueous solutions along a temperature-decreasing path (Fig. 10).


5. Metamorphic recrystallization Protolith zircons may be metamorphosed through structural modification and chemical alteration during subduction-zone metamorphism, resulting in varying degrees of recrystallization. The mechanisms of metamorphic recrystallization can be categorized into solid-state transformation, metasomatic alteration and dissolution reprecipitation, respectively (Xia et al., 2009, 2010, 2013; Chen et al., 2010, 2011). The three mechanisms are primarily dictated by the extent of fluid action during metamorphism. In association with modification in crystal structure, the metamorphosed zircons may show partial decouple not only in some trace elements but also in O, U-Th-Pb and Lu-Hf isotope systems. The differential behaviors of different element and isotope systems during metamorphism are mainly caused by the difference in their diffusion behaviors in zircon and their mobilities upon the action of metamorphic fluids or anatectic melts. As a consequence, different types of metamorphosed zircons show different characteristics in mineragraphy, U-Pb ages, inclusions, trace elements, and Lu-Hf and O isotopes.

5.1 Solid-state transformation The solid-state transformation of protolith zircons commonly occurs during fluid-absent metamorphism. The resulted zircons show partial reworking in internal structure and U-Th-Pb isotopes, resulting in blurred oscillatory zoning (Fig. 11a, b) and discordant U-Pb ages with apparent

206

Pb/238U ages close to the protolith age or decreasing towards the metamorphic age (Xia

et al., 2009; Chen et al., 2010). The younger ages are generally ascribed to preferential loss of radiogenic Pb by diffusion from metamorphosing zircons. This is because the radiogenic Pb occurs in the cation of either Pb4+ or Pb6+ which is much smaller than Pb2+ in size, and thus it is more susceptible to diffusion loss than the common Pb. Previous studies also suggested purging of non-essential structural constituent elements (e.g., LREE in addition to the radiogenic Pb) from the crystal structure of protolith zircons by metamorphic recrystallization. Cations with ionic radii significantly different from Zr (and Hf) and Si are preferentially purged from the zircons (e.g., Vavra et al., 1996; Hoskin and Black, 2000; Hoskin and Schaltegger, 2003). The concentrations of trace elements in solid-state recrystallized zircons sometimes show large variations. However, their apparent

206

Pb/238U ages are not positively

correlated with HREE, MREE, Th+U, Nb+Ta contents (Figs. 12 and 13), indicating that these variations are not caused by the simple loss of incompatible trace elements from the protolith zircons. Instead, the variations are largely primary, inherited from the protolith zircons. This is consistent with the compositional variation in magmatic zircons that has long been observed by a lot of studies (e.g., Hoskin and Ireland, 2000). Although trace elements are lost from zircon at different degrees, the REE patterns are still typical of magmatic zircon (Fig. 12). Although Th/U ratios decrease to different extents, they are still greater than 0.1. In this regard, it is likely that the solid-state recrystallized zircons have largely preserved the structural and geochemical characteristics of protolith zircons. Except for the radiogenic Pb, the loss of other incompatible trace elements during the recrystallization was very limited at best. There is no significant correlation between 206Pb/238U ages and Hf contents for metamorphosed zircons in the CCSD-MH samples of the Sulu orogen (Fig. 13d), indicating that the large variation in the Hf contents of these zircons may be largely primary, inherited from the protolith zircons. In view of the immobility of Hf, the Lu-Hf isotope composition of solid-state recrystallized zircons would also be retained. This is confirmed by the observation that no obvious correlations are observed not only between 176

206

Pb/238U ages and Hf(t) values but also between

176

Lu/177Hf and

Hf/177Hf ratios (Fig. 13e, f). All these observations indicate that the Lu-Hf isotope system was not

disturbed considerably.

5.2 Metasomatic alteration The metasomatic alteration of protolith zircons proceeds through fluid infiltration along grain fractures and boundaries, leading to heterogeneous element and isotope exchange between zircon and fluid. While porphyritic texture may develop due to possible dissolution reprecipitation along the grain fractures and boundaries, matrix domains generally show blurred oscillatory zoning, planar zoning, weak zoning or even unzoning in CL images (Fig. 11c, d). As a whole, the metasomatically recrystallized zircons exhibit lowered Th/U ratios and discordant U-Pb ages with apparent

206

Pb/238U ages clearly moving towards the metamorphic age (Fig. 13b). Their

ratios roughly show a negative correlation with apparent 176

206

176

Lu/177Hf

Pb/238U ages, indicating that the

Lu/177Hf ratios were also slightly lowered by the metasomatic alteration (Fig. 13e). In contrast,

there is no correlation between 176Hf/177Hf ratios and apparent that their

176

206

Pb/238U ages (Fig. 13f), indicating

Hf/177Hf ratios generally remain unchanged (Chen et al., 2010; Xia et al., 2010). Their

O isotope composition may show a large variation (Fig. 5b), with δ18O values varying between metamorphic and protolith zircons (Chen et al., 2011; Sheng et al., 2012). The decoupling in zircon U-Pb isotopes, trace elements and Lu-Hf isotopes was ascribed to the differential diffusion loss of incompatible trace elements (e.g., Martin et al., 2008), but it had better to be explained by the metasomatic recrystallization. Trace elements in protolith zircons were variably modified by metasomatic alteration, depending on the nature and composition of infiltrating fluids. If the metasomatic fluid is an aqueous solution, fluid-mobile incompatible trace elements such as LILE, Pb and U would be modified to large extent, whereas fluid-immobile elements such as LREE and Th were less modified and even unchanged. Nevertheless, the aqueous solution may act as a catalyzer to enhance the diffusivity of some melt-mobile trace elements and thus result in lowering of Th and LREE in protolith zircons (Chen et al., 2010). This was observed in zircon from the CCSD-MH samples (Fig. 12a, b). If the metasomatic fluid is a hydrous melt, melt-mobile incompatible trace elements such as LILE, LREE, Th, U and Pb would all be significantly modified (Xia et al., 2009, 2010). This was seen in zircon from the UHP gneiss in the South Dabie (Fig. 12c). Mineral inclusions in this type of metamorphosed zircons are mainly the protolith minerals of magmatic origin. Because of the fluid metasomatism, however, a few metamorphic/metasomatic minerals can also be found in these zircons. In addition, there is a difference in the mechanism of fluid action during the metasomatic recrystallization of protolith zircons. While the fluid action along grain fractures and boundaries is mainly the dissolution reprecipitation, that on matrix zircons is primarily the diffusion-driven reactions (Geisler et al., 2007; Xia et al., 2009; Chen et al., 2010). Thus, the metasomatic recrystallization is realized by the composite processes, which would induce the occurrence of nanopores and the possible occurrence of nanocrystalline ZrO2 in the recrystallized

zircon. The recrystallized zircon domains would also have high contents of non-formula elements such as Ca, Ba, Al, Fe, Mn and possibly common Pb contents.

5.3 Dissolution recrystallization The dissolution reprecipitation of protolith zircons causes the largest extent of reworking by metasomatic fluids (Xia et al., 2009, 2010; Chen et al., 2010, 2011). The protolith zircons underwent dissolution at first and then reprecipitation at the same time or immediately afterwards. They were recrystallized to generate the metamorphosed zircons that are rich in all incompatible trace elements. The process of dissolution reprecipitation was also observed by laboratory experiments (Geisler et al., 2007). The supercritical fluid may have acted as a solvent and catalyst. The nature of mineralogical reactions is also a key to the composition of product minerals. For example, the reaction of (Zr,Hf,Y,REE)(Si,P)O4 → Zr,Hf)SiO4 + (Y,REE)PO4 was suggested as a possible mechanism to generate trace elements-depleted but Hf-enriched zircon and xenotime (Pan, 1997). This reaction mechanism, with the addition of U-Th-silicate as a product, was also advocated by Tomaschek et al. (2003). During the reaction, porosity is created and occupied by fluid inclusions in the porous zircon shell; trace elements including Y, HREE, Th and P are expelled from the protolith zircons and reprecipitated as discrete phases such as xenotime and (Y, HREE, and Th)-silicates. In this regard, the dissolution recrystallized zircons would contain micrometric to nanometric sized pores and metasomatic mineral inclusions such as thorite and xenotime. Although the pores would be healed by later tectonothermal events, the resulted inclusions would be preserved (Vonlanthen et al., 2012). In this regard, the dissolution recrystallized zircon generally exhibits spongy or porous structure in CL images (Fig. 11e, f). Due to the largest extent of reworking during the dissolution recrystallization, protolith zircon U-Pb ages would be generally lowered to that close to the metamorphic age (Xia et al., 2010; Chen et al., 2011). Concordant U-Pb ages can be achieved if the dissolution recrystallization proceeds completely. The Lu-Hf isotope composition may change depending on the openness or closure of the system (Chen et al., 2010, 2011). In a relatively close system, the Lu-Hf isotope composition was primarily controlled by the protolith zircon, so that the

176

Lu/177Hf and

176

Hf/177Hf ratios are

similar to those of the protolith zircon. In an open system, the 176Lu/177Hf ratios would decrease and the

176

Hf/177Hf ratios would slightly increase due to the dissolution of Lu-rich minerals such as

garnet. The O isotope composition of protolith zircons was completely reworked (Fig. 5b), exhibiting relative y ho ogeneous δ18O values similar to that of newly grown zircons (Chen et al., 2011; Sheng et al., 2012). The dissolution recrystallized zircon has trace element composition different from the protolith zircon. There are two processes including dissolution and reprecipitation. In terms of the

relationship between dissolution and reprecipitation, there are two subtypes of dissolution recrystallization: coupled dissolution-reprecipitation, in which dissolution and reprecipitation processes

are

temporally

and

spatially

connected

with

each

other,

and

decoupled

dissolution-reprecipitation, in which the dissolution and reprecipitation processes are temporally and spatially disconnected. In the former case, precursor zircons would be redistributed into two parts: one with unzoned texture is rich in Hf, and the other with spongy texture is rich in the other trace elements. They exhibit steep MREE-HREE patterns with high (Lu/Gd)N ratios. In the latter case, zircon exhibit no zoning or weak zoning in mineragraphy, shallow MREE-HREE patterns with low (Lu/Gd)N ratios (Chen et al., 2010). This was seen in zircons from the CCSD-MH samples (Fig. 12a, b). Compared to the metasomatic recrystallization, there is the maximum extent of geochemical exchange between dissolution recrystallized zircon and fluid, and the composition of the resultant zircon is more related to that of the fluid. In this regard, the zircon composition is also related to the physicochemical property and geochemical composition of metasomatic fluids. Supercritical fluids may be involved in the dissolution recrystallization of protolith zircons (Xia et al., 2010), resulting in the dissolution reprecipitation of high-field-strength incompatible trace elements. This would cause the elevation of REE and HFSE in the resulted zircons, which was seen in the UHP gneiss zircon from South Dabie (Xia et al., 2010). The similar case was also observed in hydrothermally altered zircons that are relatively enriched in LREE (e.g., Hoskin, 2005).

5.4 Distinction between different types of metamorphosed zircons The different types of metamorphosed zircons show different geochemical compositions and thus provide different constraints on the metamorphic evolution of host rocks. The U-Pb dating of solid-state recrystallized zircons has a capacity to give a lower limit to the protolith age, whereas the U-Pb dating of dissolution recrystallized zircons tend to give an upper limit to the metamorphic age. The metasomatically recrystallized zircon would give mixed U-Pb ages, of which the youngest ones are obtained from newly grown veinlets along fractures and boundaries and they are close to the metamorphic age whereas the oldest ones occur in the matrix domains and are close to the protolith age. In this regard, identifying the different types of recrystallized zircons is critical to genetic interpretation of the zircon geochemical data, especially for the protolith tracing. The metamorphic recrystallization of protolith zircons is primarily controlled by the property of the zircons themselves and the property of acting fluids. Crystalline zircons may become metamict ones due to radiation damage, which is commonly caused by the presence of high concentrations of radioactive elements U and Th (Booth et al., 2005). This can result in amorphous domains and a porous structure, which are much more susceptible to fluid infiltration and geochemical exchange (e.g., Geisler et al., 2007). Thus, the dissolution reprecipitation often occurs in such U- and Th-rich

zircons, whereas crystalline zircons with no fracture are resistant to metasomatic alteration. In either case, once zircons experienced geochemical exchange with fluids, their structure and composition were disturbed to leave possible records for geochemical identification. Consequently, the metamorphosed zircons commonly show discordant U-Pb ages, the variable disappearance of the primary zoning in CL and BSE images, the occurrence of metasomatic mineral inclusions, element compositions deviating from typical magmatic zircons as well as the disturbance of O isotopic systematics. In addition, amorphous zircon domains would exhibit difference in Raman spectra from well crystalline zircons, depending on the metamict extents (e.g., Nasdala et al., 2001). Since the recrystallization of zircon is related to the property of zircon, different types of recrystallization is possible to exhibit difference in Raman spectra (Davies et al., 2015). Therefore, the different types of metamorphic recrystallization can be discriminated by a combined study of CL and BSE images, Raman analyses, U-Pb ages, trace elements, mineral inclusions, and Lu-Hf and O isotopes (Table 1). For the solid-state recrystallization, the action of metamorphic fluids is insignificant on the protolith zircons. As a result, the primary zoning of protolith zircons is only partially disturbed by diffusion at elevated temperatures, and some non-formula trace elements (particularly the radiogenic Pb) could be expelled due to the lattice deformation (Chen et al., 2010; Geisler et al., 2007; Hoskin and Black, 2000; Xia et al., 2009). Although the resulted zircons often have discordant U-Pb ages, they still show high Th/U ratios of >0.1 and steep HREE patterns typical of magmatic zircon. For the metasomatic recrystallization, the action of metamorphic fluids is only significant along grain fractures and boundaries, yielding a kind of porphyritic structures in which the metasomatic domains surround the solid-state recrystallized domains (relict cores). Sometimes the metasomatic domains contain nanoscale pores with variable concentration of fluid-mobile elements such as Ca, Al, and Fe, locally up to several wt.% (Geisler et al., 2007; Rayner et al., 2005). This type of recrystallized zircons would exhibit REE patterns different from typical magmatic zircons. For the dissolution recrystallization, protolith zircons would be dissolved in and reprecipitated from the supercritical fluids in the almost in-situ fashion. As a consequence, the reprecipitated domains generally exhibit skeletal textures and contain micron-sized pores and mineral inclusions such as xenotime and (Y, HREE, and Th)-silicates (Geisler et al., 2007; Rubatto et al., 2008; Tomaschek et al., 2003; Xia et al., 2010). Micron-sized pores may disappear due to later annealing at elevated temperatures (Vonlanthen et al., 2012), but the mineral inclusions cannot be completely erased by this later process. Therefore, a comprehensive study of zirconology has a great potential to provide insights not only into the origin of metamorphosed zircons themselves but also into the petrogenesis of host rocks.

While the diffusion is a principal mechanism for the loss of radiogenic Pb in the solid-state transformed zircon, this process is not effective for the mobility of incompatible trace elements and their pertinent isotopes in the absence of fluid phases (metamorphic fluids or anatectic melts). As soon as the fluid phases were accessible to protolith zircons in the metamorphosing rocks, they were variably active to the protolith zircons for metasomatic and dissolution recrystallizations. This is the reason why the protolith zircons in the same rocks often show different extents of metamorphic recrystallization. If the fluid phases are not active, magmatic minerals rather than metamorphic/peritectic minerals would occur as inclusions in the newly grown zircons. Nevertheless, the extent of variations in U-Pb ages, trace elements and Hf-O isotopes is offent dictated by the difference in the diffusion behaviors of different elements in zircon. The radiogenic Pb generally has the largest diffusion rate, whereas the other elements such as REE, Th, U, Hf and O exhibit lower diffusion rate (Cherniak et al., 2003). As a consequence, although the crystal structure and U-Pb isotope composition of a metamorphosed zircon were variably reworked, its trace elements, Lu-Hf and O isotopes have almost inherited from the protolith zircon. This leads to a kind of the decoupling between different geochemical variables in the metamorphosed zircons.

6. Zircon in orogenic peridotites Orogenic peridotites are one of the common components in UHP terranes of collisional orogens. They are commonly categorized into mantle type (M-type) and crustal type (C-type) (e.g., Carswell et al., 1983; Brueckner and Medaris, 2000; Zhang et al., 2000). M-type peridotites were originally located in the subcontinental lithospheric mantle (SCLM) wedge overlying subducting slabs and were tectonically entrained by the subducting/exhuming crustal slices into subduction channels. They would have undergone varying degrees of crustal metasomatism by fluid phases released from subducting crustal rocks in view of the stability of hydrous minerals in them (Zheng et al., 2016). Thus, M-type peridotites in continental subduction zones provide a direct lithological record of the mass transfer from the subducted slab to the SCLM wedge (Zheng, 2012). Traditionally, zircon in peridotites was ascribed to crustal contamination during their emplacement into continental crust. This is because zircon is theoretically unable to crystallize from the primary peridotite due to the low Zr content and Si activity of the bulk rock (e.g., Palme and O'Neill, 2003; Hermann et al., 2006b; Zheng, 2012). However, zircon has been found in garnet peridotites from several UHP terranes, including Dabie-Sulu (Rumble et al., 2002; Zhang et al., 2005, 2011; Zheng et al., 2006b, 2008, 2014; Li et al., 2016b), North Qaidam (Song et al., 2005; Xiong et al., 2011, 2014; Chen et al., 2017); Alps (Hermann et al., 2006b) and Erzgebirge (Liati and Gebauer, 2009). It has also been found in ophiolitic complexes (e.g., Yamamoto et al., 2013; Robinson et al., 2014; Belousova et al., 2015) and mantle xenoliths trapped by basaltic volcanics

(e.g., Liati et al., 2004; Liu et al., 2010). As illustrated in Fig. 14a, zircon grains were also recognized in the thin sections of peridotites (e.g., Zhang et al., 2005, 2011; Xiong et al., 2014; Chen et al., 2017), and the paragenesis and composition of mineral inclusions in the peridotite zircon are generally consistent with those of peridotite constituent minerals. Such observations clearly demonstrate that zircons in the peridotites were emplaced together with their host peridotites rather than originated from the crustal contamination. The presence of fluid/melt inclusions in the peridotite zircon indicates its growth during crustal metasomatism of the host peridotite (Zhang et al., 2005; Liati and Gebaur, 2009). Therefore, zircon only occurs as an extraordinarily rare phase in refertilized peridotites, providing a mineralogical evidence for the crustal metasomatism (e.g., Gebauer, 1996; Bea et al., 2001; Katayama et al., 2003; Zheng et al., 2003b, 2012).
As illustrated in Fig. 14b and c, both newly grown zircon and old relict zircon have been found in M-type orogenic peridotites (e.g., Liati and Gebauer, 2009; Yang et al., 2009; Zheng et al., 2014; Li et al., 2016b). However, their origin has been controversial. The origin of newly grown zircons is generally ascribed to one of the following two mechanisms: (1) crystallization from metasomatic fluids of mantle origin (e.g., Grieco et al., 2001; Zheng et al., 2006b); and (2) crystallization from metasomatic fluids derived from dehydration of the deeply subducted crust (e.g., Hermann et al., 2006b; Liati and Gebauer, 2009; Li et al., 2016b; Chen et al., 2017). In either case, these two mechanisms emphasize the role of metasomatic agents in dictating the growth of zircon in orogenic peridotites (Zheng et al., 2011a; Zheng and Hermann, 2014). Because either the asthenospheric or lithospheric mantle has very low Zr contents, zircon cannot directly crystallize from mantle-derived fluids produced by devolatilization of normal mantle rocks. Thus, the first mechanism does not hold under close scrutiny. In contrast, crustal rocks have variably high Zr contents and zircon is susceptible to growth through dehydration reactions of the crustal rocks. As a consequence, zircon growth is common due to the crustal metasomatism at the slab-mantle interface in continental subduction channels (Zheng, 2012). Zircons were generally found in orogenic peridotite that underwent large extent of crustal metasomatism and thus contain secondary hydrous minerals such as phlogopite, amphibole or Ti-clinohumite (Katayama et al., 2003; Zhang et al., 2005, 2011; Hermann et al., 2006b). So far no or rare zircons were found in phlogopite-free peridotites. As illustrated in Fig. 14d and 14e, mineral inclusions such as apatite, phlogopite, amphibole and U-oxide as well as felsic melt inclusions were found in newly grown zircons of orogenic peridotites from Dabie-Sulu, Alps and Bohemia (Zhang et al., 2005, 2011; Hermann et al., 2006b; Liati and Gebauer, 2009). The element analysis of these mineral inclusions indicates that they contain crustal signatures (Hermann et al., 2006b). Newly grown zircons in the orogenic peridotite from Tengjia in the Sulu orogen exhibit negative to low

δ18O values of -11.3 to 0.9‰ (Li et al., 2016b), consistent with the O isotope composition of UHP metaigneous rocks in this region. All these observations indicate that the newly grown zircons in orogenic peridotites were generated by the crustal metasomatism, and thus they are named as metasomatic zircon. These zircons generally exhibit no zoning, and some of them exhibit oscillatory zoning or relict cores (Liati and Gebauer, 2009; Yang et al., 2009; Zheng et al., 2014; Li et al., 2016b). There are generally different episodes of zircon growth in orogenic peridotites during subduction zone metamorphism (e.g., Hermann et al., 2006b; Liati and Gebauer, 2009; Li et al., 2016b; Chen et al., 2017). As depicted in Fig. 15a, metasomatic zircons in M-type peridotites from the Dabie-Sulu orogenic belt show variable U-Pb ages from 205 to 244 Ma, indicating different episodes of zircon growth during continental collision. Most of these U-Pb ages are in the range of 212 to 227 Ma (Fig. 15a), later than the UHP metamorphic ages of 225-240 Ma, indicating their growth by crustal metasomatism at the early exhumation stage of the deeply subducted continental crust. These metasomatic zircons show large variations in Th and U contents and Th/U ratios from <0.01 to 1 (Fig. 15b), suggesting their growth through dehydration reactions at temperatures below and above the wet solidus of crustal rocks. Such processes also give rise to different compositions of fluids (e.g., aqueous solutions versus hydrous melts) in metamorphic/anatectic systems. As illustrated in Fig. 14f, melt inclusions occur in peridotite zircon from the Bohemia massif (Liati and Gebauer, 2009) whereas fluid inclusions occur in peridotite zircon from the Sulu orogen (Zhang et al., 2005), confirming that both fluid phases were involved in the crustal metasomatism.
For old zircon domains in orogenic peridotites, two mechanisms have been proposed for their origin: (1) injection of granite-related melts into previously emplaced peridotites, carrying crustal zircon and possibly crystallizing new zircon (Belousova et al., 2015); (2) capture of xenocrysts from crustal rocks recycled by deep subduction (e.g., Zheng, 2012; Yamamoto et al., 2013; Robinson et al., 2015; Li et al., 2016b). The first mechanism links the old and new zircons to granitic magmatism in the continental crust, emphasizing zircon introduction by secondary alteration of the peridotites after their emplacement into the continental crust. Thus it ascribes the occurrence of zircons in the peridotites to the emplacement of granitic magmas into the crustal level. Granitic magmatism generally requires partial melting of crustal rocks at temperatures of at least 650C, which is the upper limit for the temperature at the crust-mantle transition zone of continental lithosphere (Zheng and Chen, 2016). At such low temperatures, only leucocratic granite can be produced by eutectic melting (Johannes and Holtz, 1996), with the preservation of detrital zircons but the very limited growth of magmatic zircon. Thus, this mechanism is only applicable to the association of orogenic peridotites with granites, and zircons in peridotites and granites would have

consistent U-Pb ages and Hf-O isotopes. However, orogenic peridotites in continental subduction zones are generally associated with UHP gneiss/eclogite rather than granites. Thus the first mechanism cannot hold for the occurrence of old zircons in orogenic peridotites. The second mechanism links the old zircon in the peridotites to the recycling of deeply subducted crustal zircons, with their survival during metamorphic dehydration and partial melting at the slab-mantle interface in subduction channels. This corresponds to the physical transport of detrital zircons from deeply subducted crustal rocks to the SCLM wedge (Zheng and Hermann, 2014). Such an interpretation is reinforced by the case study of peridotite zircons in the Dabie-Sulu orogenic belt (Li et al., 2016b). In addition to the metasomatic zircon of Triassic U-Pb ages, there are also relict zircons with older ages in the M-type orogenic peridotites from the Dabie-Sulu orogenic belt (Yang et al., 2009; Zheng et al., 2014; Li et al., 2016b). These relict zircons generally occur as cores, and exhibit no zoning, blurred oscillatory zoning or oscillatory zoning. They show a large variation in U-Pb ages (Fig. 15a, c). Some of them have concordant Neoproterozoic U-Pb ages of 740-780 Ma, a lot of them have U-Pb ages between Neoproterozoic and Triassic, and few of them have Paleoproterozoic and even Archean U-Pb ages (Fig. 15). They contain mineral inclusions of apatite, biotite, calcite, amphibole, xenotime and sulfide, but no peridotite constituent minerals such as olivine, and pyroxene occur in these zircons (Zheng et al., 2014; Li et al., 2016b). This indicates that they were not originated from the peridotites themselves, but physically transported by metasomatic agents into the peridotites. As presented below, several lines of evidence indicate that these relict zircons were originated from the deeply subducted continental crust of the South China Block, part of it being exhumed as the UHP metamorphic rocks along the Dabie-Sulu orogenic belt. (1) The Neoproterozoic U-Pb ages in the M-type peridotites are common for relict zircons in the Dabie-Sulu UHP metaigneous rocks, and U-Pb ages older than Neoproterozoic also occur in relict zircons from the UHP metasedimentary rocks (e.g., Zheng et al., 2009; Liu and Liou, 2011; Zhang and Zheng, 2013; Zhang et al., 2014). 2) The negative to o

δ18O relict zircons not only occur in the Tengjia

peridotite (Li et al., 2016b) but are also common in the Dabie-Sulu UHP metaigneous rocks (Zheng et al., 2004; Tang et al., 2008a, b; Chen et al., 2011; Sheng et al., 2012). (3) The relict zircons in the M-type peridotites have comparable Hf isotope compositions to those of the Dabie-Sulu UHP metaigneous rocks (e.g., Li et al., 2016b). (4) Relict zircons with similar U-Pb ages, Hf-O isotope compositions have been found in quartz veins and leucosomes inside the UHP rocks (e.g., Sheng et al., 2012; Li et al., 2013; Xu et al., 2013). Therefore, the deeply subducted continental crust of the South China Block would have delivered the metasomatic agents not only for the new growth of metasomatic zircons but also for the physical transport of relict zircons in the M-type peridotites in

the Dabie-Sulu orogenict belt.

7. Zircon records fluid action in continental subduction channel Generally speaking, the new growth of metamorphic, peritectic and metasomatic zircons requires supply of Zr to their host rocks. This element can be derived either from the breakdown of Zr-bearing minerals such as biotite, hornblende and ilmenite (Zheng, 2012), or from the dissolution of protolith zircons (Vavra et al., 1999; Ayers et al., 2003). Because Zr is a water-insoluble element and generally immobile in most geological environments, it is usually expected that the newly grown zircon would spatially occur close to Zr-releasing reactions in the rock matrix. This process has been documented by Degeling et al. (2001) from garnet-bearing migmatites in SW Norway, where numerously small grains of zircon grew directly in the cordierite corona that formed as the result of garnet breakdown. The in-situ growth of zircon from the breakdown of garnet and ilmenite was also reported by Fraser et al. (1997) and Bingen et al. (2001), respectively, during granulite-facies metamorphism. It is expected that zircons in this case would exhibit a large variation in trace element composition depending on the composition of their parental minerals. However, an integrated study of textural and in-situ microanalyses by Sláma et al. (2007) for zircon in high-grade metamorphic rocks from the Bohemian Massif shows that the newly grown zircon is not spatially associated with the reaction corona despite the reaction resulting in significant release of Zr. A lot of other studies also indicate that metamorphic zircons generally have relatively consistent rather than scattered trace element compositions (e.g., Chen et al., 2010; Liu and Liou, 2011). In this regard, these zircons would be produced through metamorphic reactions with possible homogenization in trace element composition upon on the action of fluid phases. Because of the Zr immobility in aqueous solutions, it is unlikely for Zr to be efficiently transferred via the fluid phase. Nevertheless, Zr is a melt-mobile incompatible trace element and it is susceptible to dissolution into hydrous melts. In particular, aqueous solutions and hydrous felsic melts have much lower Zr solubility than mafic melts (Harrison and Watson, 1983; Watson, 1996; Harrison et al., 2007). The important role of fluid phases in the growth and recrystallization of zircon during subduction-zone metamorphism has been well recognized (e.g., Rubatto et al., 1999; Rubatto and Hermann, 2003; Wu et al., 2006; Zheng, 2009). This is exemplified by zircons from UHP eclogites at Yangkou and Qinglongshan in the Sulu orogen (Zheng, 2009). The Yangkou eclogite was produced in the absence of metamorphic fluids as indicated by the occurrence of both intergranular coesite and its pseudomorph in the matrix of eclogite and the preservation of igneous textures in the eclogitized metagabbro (Liou and Zhang, 1996; Zhang and Liou, 1997), and evidence for disequilibrium in the Sm-Nd isochron system between eclogite minerals (Zheng et al., 2002). Thus, there was no new growth of metamorphic zircons in the Yangkou eclogite (Zheng et al., 2004). In

contrast, the Qinglongshan eclogite was generated in the presence of metamorphic fluids as indicated by the occurrence of negative δ18O values and equilibrium Sm-Nd and Rb-Sr isochron systems between coexisting minerals (Zheng et al., 2002, 2003c), and thus contain a great amount of metamorphic zircons (Zheng et al., 2004; Chen et al., 2011). As discussed before, the presence of fluid phases is necessary in order to transform the protolith mineral assemblage to the metamorphic mineral assemblage via dehydration/hydration reactions, otherwise the protolith texture may be preserved despite the eclogite-facies metamorphism. A similar case was also observed in eclogitized granulites in the Western Gneiss Region of Norway, where eclogitization only took place along granulite fractures through which metamorphic fluids were able to flow (Austrheim, 1987). In addition, the presence of fluids would also significantly facilitate the recrystallization of protolith zircons. Therefore, the fluid availability dictates not only the growth of metamorphic/peritectic zircon but also the extent of metamorphic recrystallization during subduction-zone metamorphism. In this regard, the zirconological study of HP to UHP rocks from collisional orogens can provide advanced constraints on fluid action in continental subduction zones. Although the present study focuses on zircon, it merits to point out that important information about the evolutionay history and fluid action of HP to UHP rocks can also be obtained by dating other chronometric minerals such as monazite, apatite, rutile, titanite and allanite and investigating their internal textures and position within individual samples (e.g., Taylor et al., 2016; Kohn, 2016). The U-Pb dating of metamorphic zircons in the Dabie-Sulu orogenic belt indicates that metamorphic zircon would have grown in the whole period of continental collision with two peak at the prograde HP-UHP transition stage and the retrograde UHP-HP transition stage (e.g., Zheng et al., 2005b; Wu et al., 2006; Zheng, 2009; Xia et al., 2013). Thus, these metamorphic zircons record the two major episodes of fluid action in the continental subduction zone during subduction and exhumation, respectively. Compared to the fluid action during subduction, the fluid action during exhumation is more extensive and intensive. This is consistent with the observation that the majority of metamorphic zircons would have grown during the exhumation of UHP metamorphic rocks in continental subduction zones. The growth of peritectic zircons during subduction is also recognized from UHP migmatites with metasedimentary protoliths at Weihai and Taohang in the Sulu orogen (Li et al., 2014, 2016a), indicating that the anataxis of UHP metamorphic rocks also took place during the subduction. Numerical models and experimental studies have indicated that partial melting of metasediments atop the subducting slab would be triggered by infiltration of aqueous solutions that were derived from dehydration of the underlying rocks in a deeper slab interior (Gao et al., 2015; Li et al., 2016a). The released fluids from the subducting slab interior would be transported upwards into the

metasedimentary rock atop the subducted slab (Gao et al., 2015). With heating of the overlying mantle wedge, hydration melting of the metasedimentary rocks would take place atop the subducting slab. During the initial exhumation of UHP crustal slices, the subducted slab would undergo either near-isothermal decompression (e.g., Nakamura and Hirajma, 2000) or an increase in temperature by adiabatic decompression (Zhao et al., 2007a; Li et al., 2014). The latter has been named as the “hot” exhu ation (Zheng et al., 2011a), which leads to the dehydration melting that has been reported in several UHP terranes such as Dabie-Sulu (e.g., Zheng et al., 2011a; Wang et al., 2014), Bohemia (e.g., Kotková et al., 2016), North Greenland (e.g., Gilotti et al., 2014), Kokchetav (e.g, Ragozin et al., 2009) and North Qaidam (e.g., Zhang et al., 2015). Peritectic zircons recognized from these UHP terranes show a large variation in trace element composition (Fig. 9), suggesting that their growth is associated with variable activities of garnet and plagioclase during the exhumation from UHP eclogite-facies to HP amphibolite-facies conditions. U-Pb dating of these zircons also suggests that anatexis of the UHP metamorphic rocks generally occurred at the HP granulite-facies stage, but started at UHP conditions. A numerical modelling for melt-bearing UHP crustal rocks suggests that the fluid- and melt-induced weakening of crustal rocks at mantle depths would have significant geodynamic effects on exhumation mechanisms (Jamieson et al., 2011; Labrousse et al., 2011, 2015). The results indicate that when the partial melting of UHP metamorphic rocks occurs, the rocks at different positions in the subducting slices and the different lithologies of subducting rocks may have different P-T-t paths during continental collision. As illustrated in Fig. 16 for the UHP metamorphic rocks in the Sulu orogen, different P-T-t paths are obtained for three UHP migmatite samples from the same outcrop at Weihai and two UHP metagranite samples from the same outcrop at Taohang (Li et al., 2014, 2016a). This difference suggests differential behaviors of dehydration and anataxis during subduction and exhumation of continental crust. The different portions of a subducting slice and its different lithologies would undergo different degrees of dehydration and anataxis during continental collision. During subduction, fluids released from the subducting slab interior would be transported upwards into the UHP rock atop the subducted slab (Gao et al., 2015). With heating of the overlying mantle wedge, hydration melting of the UHP rocks (especially for metasedimentary rock) atop the subducting slab could occur, whereas those in the interior of the subducting crust would remain below the wet solidus. During the exhumation of UHP crustal slices, on the other hand, the temperature atop the subducted slab would decrease, but the crustal interior would either retain the high temperature by near-isothermal decompression or undergo an increase by adiabatic decompression. Therefore, the anataxis of UHP rocks during the exhumation was able to take place in continental subduction zones.


Different lithologies (metagranite, metasedimentary and metabasite) all underwent different extents of anatexis during exhumation, whereas metasedimentary rocks atop the subducted slab would also undergo anataxis during subduction (e.g., Li et al., 2016a). This difference is controlled by the wet solidus of crustal rocks and thus the availability of water inside the UHP rocks. The nature of protoliths is the major controlling factor in dictating the water availability in the UHP rocks during subduction-zone metamorphism (e.g., Zheng, 2009; Xia et al., 2013). This primarily depends on the species and abundance of water in the protoliths. Volcanic and sedimentary rocks contain abundant water, mainly in the form of molecular water (pore fluid), whereas intrusive rocks contain less water, which is primarily in the form of structural hydroxyl. Because the molecular water has higher mobility than the structural hydroxyl, both metavolcanic and metasedimentary rocks have the capacity to release more water than the metaintrusive rocks during continental collision. This is illustrated by UHP metamorphic rocks from the Dabie-Sulu orogenic belt that exhibit different episodes of fluid activity and zircon growth in the same lithotectonic units (Fig. 16). The occurrence of quartz veins and leucosomes in UHP metamorphic rocks indicates that both metamorphic fluid and anatectic melt were produced with focused flow during continental collision (e.g., Wu et al., 2009; Chen et al., 2012a, b; Sheng et al., 2012, 2013; Xu et al., 2013). Besides newly grown zircons (metamorphic and peritectic zircons), relict protolith zircons were found in the quartz veins and leucosomes. Both metasomatic and inherited zircons also occur in orogenic peridotites. These observations indicate that fluids released from the sudbucting continental crust would not only chemically transport incompatible elements such as Zr and Si, but also physically transport tiny grains of refractory minerals. These released fluids resulted in the formation of quartz veins and leucosomes within the UHP rocks in the subducting/exhuming slices on the one hand, transfer of the incompatible elements and tiny minerals into the overlying mantle wedge on the other hand. The former case is illustrated by the study of three UHP migmatite samples from the same outcrop at Weihai in the Sulu orogen (Li et al., 2016a), which exhibit different U-Pb ages and 18O values for peritectic zircons (Fig. 17). The first episode of anataxis occurred in UHP metasedimentary rock at 230-227 Ma during subduction into the coesite stability field and the second episode of anataxis occurred in both metasedimentary and metaigneous rocks at ca. 220 Ma during exhumation into the lower crust (Fig. 16b). This observation further indicates that different types and origins of UHP rocks were mixed in the local place to undergo dehydration-driven peritectic reactions at different times and thus produce the different compositions of anatectic melts. The leucosomes at Weihai contain peritectic zircons grown at different episodes and are associated with the melts of different compositions (Fig. 17). Similar cases were found in quartz veins and leucosomes within UHP metamorphic rocks in the Sulu orogen. This indicates that the veins and

leucosomes are generally mixed results from different episodes of fluid/melt action. Because of the possible evolution and differentiation during fluid/melt transport, it should be cautious to use their whole-rock compositions to trace the composition of fluid/melt (Li et al., 2016a).
Numerical modelling was carried out to study zircon behaviors during metamorphic dehydration and partial melting of hydrous MORB and felsic rocks (Kelsey et al., 2008; Kelsey and Powell, 2011; Kohn et al., 2015). The results suggest that the solubility of Zr would increase in metamorphic fluids during prograde and peak metamorphism and thus zircon would dissolve into slightly rather than grow from the fluids. Partial melting may drive massive zircon dissolution, but crystallization of Zr-poor minerals from the melt leads to regrowth of zircon. In view of the mass balance, they proposed that zircon growth from the fluid phases cannot be attributed to the prograde amphibolite-eclogite transition to UHP facies, or to partial melting. Instead, zircon could grow mainly during late-stage exhumation and melt cooling, particularly during oxide transitions from rutile to ilmenite during the melt crystallization. However, this would make the anatectic zircon rather than the metamorphic and peritectic zircons. Thus, it is important to distinguish the anatectic minerals from the metamorphic and peritectic minerals in high-grade metamorphic rocks. The study of natural samples from UHP terranes indicates that most zircons grew at different stages during exhumation, and they are the products of metamorphic and peritectic reactions rather than direct crystallization from subsolidus metamorphic fluids or supersolidus anatectic melts. In addition, there was considerable growth of both metamorphic and peritectic zircons during subduction. In either case, the newly grown zircons would not directly crystallize from the metamorphic fluids or the anatectic melts, otherwise the local oversaturation of Zr in the fluids/melts was achieved by significant crystallization of Zr-poor minerals from these fluids/melts with changes in metamorphic P-T conditions. This again makes the anatectic zircon rather than the metamorphic and peritectic zircons. Instead, both metamorphic and peritectic zircons would form by mineralogical reactions in association with dehydration or hydration. The metamorphic and peritectic reactions have the capacity to transform reactants to solid products with the local saturation of Zr for the growth of both zircon and garnet, leaving dilute fluids/melts as the liquid product. On the other hand, the solubility of Zr in aqueous solutions and hydrous melts is generally too low (e.g., Ayers et al., 2012) to reach oversaturation for direct zircon growth from metamorphic fluids and anatectic melts. In this regard, the occurrence of both metamorphic and peritectic zircons in the same UHP metamorphic rocks indicates that the host rocks underwent multistage reworking at temperatures below and above the wet solidus of crustal rocks during collisional orogeny. In addition, anatectic zircons could grow from anatectic melts if they were considerably evolved with significant crystallization of Zr-poor minerals. In this case, their U-Pb ages can give the timing of

melt crystallization and thus the lower limit of anatectic event. In contrast, the peritectic zircons formed through peritectic reactions during crustal anataxis, and their U-Pb ages record the timing of anatectic events. In summary, Zr is a HFSE element, so that Zr in minerals is not susceptible to dissolution and transportion by common fluid phases such as aqueous solutions and hydrous melts. However, it can be efficiently dissolved and transported by alkaline and supercritical fluids. As an incompatible trace element, Zr tends to concentrate in hydrous melts during crustal anatexis, but it does not mean that zircon can directly precipitate from anatectic melts without considerable evolution into magmatic melts in which Zr-poor minerals have fractionally crystallized significantly. New zircon growth starts as soon as the Zr saturation is achieved in local environments, which is common in the products of metamorphic and peritectic reactions (Fig. 18). The protolith zircon of magmatic and detrital origins can be modified by metamorphic recrystallization, approaching different extents of thermodynamic reequilibration (Fig. 18). This primarily depends on its physiochemical property (zircon crystallinity, and the presence or absence of cracks in the zircon crystal) and the accessibility to metamorphic fluids or anatectic melts. Thus, the property of zircon in premetamorphic protoliths and the behaviors of Zr-bearing minerals during subduction-zone metamorphism dictate the growth and recrystallization of zircon during collisional orogeny.


8. Concluding remarks The protolith zircons of magmatic and detrital origins can be modified by metamorphic recrystallization under HP to UHP conditions, approaching different extents of thermodynamic reequilibration during continental subduction-zone metamorphism. This mainly depends on its physiochemical property (zircon crystallinity, and the presence or absence of cracks in the zircon crystal) and the accessibility to metamorphic fluids or anatectic melts. The metamorphic recrystallization of protolith zircons in UHP metamorphic rocks can proceed via the mechanisms of solid-state transformation, metasomatic alteration and dissolution reprecipitation. Solid-state transformed zircons exhibit the lowest degrees of reworking and thus almost inherit all geochemical features from the protolith zircons. Thus, they are often characterized by discordant U-Pb ages close to or below the protolith age, steep MREE-HREE patterns typical of magmatic origin, high trace element contents, and protolith Hf and O isotope compositions. Dissolution reprecipitated zircons show the highest degrees of reworking and thus concordant or nearly concordant U-Pb ages for the metamorphic event, different REE patterns from the protolith zircon, changed trace element abundances relative to the protolith zircons, almost unchanged Hf isotope ratios and similar O isotope compositions to the metamorphic zircon. Metasomatically altered zircons display

intermediate degrees of reworking and thus have their many geochemical features in between. They often show considerable decoupling in many geochemical variables due to differential metasomatism of different domains. The distinction between the three types of metamorphic recrystallization is critical to tracing of the protolith nature of UHP metamorphic rocks. New growth of zircons in UHP metamorphic rocks is primarily realized by metamorphic and peritectic reactions, respectively, below and above the wet solidus of crustal rocks. Although aqueous solutions were also produced by metamorphic dehydration and hydrous melts were also generated by dehydration-driving partial melting, the solubility of Zr in these fluid phases are too low to reach the Zr oversaturation for zircon growth directly from them. In this regard, the composition of reactant minerals is the key to the species of metamorphic and peritectic minerals. Because protolith minerals may be incompletely consumed during these mineralogical reactions, it is possible that relict minerals coexist with the newly grown minerals in UHP metamorphic rocks. Nevertheless, the newly grown zircons exhibit concordant U-Pb ages, variable CL images, variable trace element compositions, elevated

176

Hf/177Hf ratios, and variable 18O values. The variable trace

element compositions are mainly controlled by the metamorphic P-T conditions and the dissolution and preservation of specific minerals during dehydration and anataxis. The elevated

176

Hf/177Hf

ratios indicate that Zr-rich minerals other than zircon were involved in the dehydration and anataxis to provide the elements of Zr and Hf for zircon growth. The O isotope composition of newly grown zircons is primarily controlled by the composition of host rocks. As a consequence, the metamorphic and peritectic zircons may exhibit similar geochemical features on the one hand and different geochemical features on the other hand. Nevertheless, some characteristic element contents and ratios, CL images and inclusions can be used to distinguish between metamorphic and peritectic zircons. An integrated study of mineralogical and geochemical compositions holds a great potential to make the distinction between the two types of newly grown zircons. The protolith nature of UHP metamorphic rocks dictates fluid availability and thus the recrystallization of protolith zircons and the new growth of zircons during continental subduction-zone metamorphism. While the dehydration and anataxis of UHP metamorphic rocks are intensive during exhumation, the anataxis of UHP metamorphic rocks also occurred during subduction. The differences in the positions and lithology of the subducting crust also have influences on the behavior of dehydration and anataxis during continental collision. The UHP metasedimentary rocks atop the deeply subducted continental slab may undergo hydration melting during the final subduction, whereas their underlying metagranite and metabasite would undergo dehydration or hydration melting during exhumation of the deeply subducted continental crust. Different episodes of fluid phases may have different origins and they could physically and chemically mix with each other at the slab-mantle interface in continental subduction channels.

New growth of zircon in orogenic peridotites is primarily associated with metasomatism of crustally derived fluid phases in continental subduction channels. This process is accompanied by both physical transport of tiny protolith zircon and chemical transport of element Zr. The dehydration and anataxis of UHP metamorphic rocks during the exhumation would result in not only the formation of veining and amphibolite-facies retrogression within the UHP slices but also the metasomatism of the overlying mantle wedge. Therefore, both growth and recrystallization of zircon in HP to UHP metamorphic rocks and orogenic peridotites are primarily dictated by the fluid availability from the breakdown of hydrous minerals and the exsolution of structural hydroxyl and molecular water in nominally anhydrous minerals.

Acknowledgments This study was supported by funds from the Chinese Ministry of Science and Technology (2015CB856106), the Natural Science Foundation of China (41422301, 41590624, 41673030 and 41372073), Youth Innovation Promotion Association CAS (2013283) and the Fundamental Research Funds for the Central Universities (WK3410000009). Thanks are due to Yi-Xiang Chen, Hai-Yong Li, Wan-Cai Li, Ying-Ming Sheng, Yuan-Bao Wu, Xiong-Xia Xia and Long Zhang for their assistance during different phases of this work. We are grateful to two anonymous reviewers for their comments that greatly helped the improvement of the presentation.

References Amelin, Y., Lee, D.C., Halliday, A.N., 2000. Early-middle Archean crustal evolution deduced from Lu–Hf and U–Pb isotopic studies of single zircon grains. Geochim. Cosmochim. Acta 64, 4205–4225. Austrhei , H., 1987. E ogitization of o er rusta granu ites by flui

igration through shear

zones. Earth Planet. Sci. Lett. 81, 221–232. Auzanneau, E., Vielzeuf, D., Schmidt, M.W., 2006, Experimental evidence of decompression melting during exhumation of subducted continental crust. Contrib. Mineral. Petrol. 152, 125–148. Ayers, J.C., Delacruz, K., Miller, C., Switzer, O., 2003. Experimental study of zircon coarsening in quartzite ± H2O at 1.0 GPa and 1000°C, with implications for geochronological studies of high-grade metamorphism. Am. Mineral. 88, 365–376. Ayers, J.C., Zhang, L., Luo, Y., 2012. Zircon solubility in alkaline aqueous fluids at upper crustal conditions. Geochim. Cosmochim. Acta 96, 18-28. Baldwin, J.A., Brown, M., 2008. Age and duration of ultrahigh-temperature metamorphism in the Anápolis–Itauçu Complex, Southern Brasília Belt, central Brazil — constraints from U–Pb geochronology, mineral rare earth element chemistry and trace-element thermometry. J. Metamorph. Geol. 26, 213–233. Bea, F., Fershtater, G.B., Montero, P., Whitehouse, M., Levin, V.Y., Scarrow, J.H., Austrheim, H., Pushkariev, E.V., 2001. Recycling of continental crust into the mantle as revealed by Kytlym dunite zircons, Ural Mts, Russia. Terra Nova 13, 407-412. Bea, F., Montero, P., Ortega, M., 2006. A LA-ICP-MS evaluation of Zr reservoirs in common crustal rocks: implication for Zr and Hf geochemistry, and zircon-forming processes. Can. Mineral. 44, 693-714. Be ousova, E.A., Ji énez, J.M.G., Graha , I., Griffin, W.L., O’Rei y, S.Y., Pearson, N., Martin, L., Craven1, S., Talavera, C., 2015. The enigma of crustal zircons in upper-mantle rocks: clues from the Tumut ophiolite, southeast Australia. Geology 43, 119-122. Bingen, B., Austrheim, H., Whitehouse, M.J., 2001. Ilmenite as a source for zirconium during high-grade metamorphism? Textural evidence from the Caledonides of Western Norway and implications for zircon geochronology. J. Petrol. 42, 355 –375. Bingen, B., Austrheim, H., Whitehouse, M.J., Davis, W.J., 2004. Trace element signature and U–Pb geochronology of eclogite-facies zircon, Bergen Arcs, Caledonides of Western Norway. Contrib. Mineral. Petrol. 147, 671-683. Boehnke, P., Watson, E.B., Trail, D., Harrison, T.M., Schmitt, A.K., 2013. Zircon saturation re-revisited. Chemical Geology 351, 324-334.

Bohlen, S.R., Boettcher, A.L., 1982. The quartz-coesite transformation: A precise determination and the effects of other components. J. Geophys. Res. 87, 7073–7078 Booth, A.L., Kolodny, Y., Chamberlain, C.P., McWilliams, M., Schmitt, A.K., Wooden, J., 2005. Oxygen isotopic composition and U-Pb discordance in zircon. Geochim. Cosmochim. Acta 69, 4895–4905. Brown, M., 2010. Melting of the continental crust during orogenesis: the thermal, rhe- ological, and compositional consequences of melt transport from lower to upper continental crust. Can. J. Earth Sci. 47, 655–694. Brueckner, H.K., Medaris, L.G., 2000. A general model for the intrusion and evolution of 'mantle' garnet peridotites in high-pressure and ultra-high-pressure metamorphic terranes. J. Metamorph. Geol. 18, 123-133. Bundy, F.P., 1980. The P, T phase and reaction diagram for elemental carbon, 1979. J. Geophys. Res. 85, 6930–6936. Carswell, D.A., Harvey, M.A., Al-Samman, A., 1983. The petrogenesis of contrasting Fe-Ti and Mg-Cr garnet peridotite types in the high grade gneiss complex of Western Norway. Bull. Mineral. 106, 727-750. Chen, R.-X., Zheng, Y.-F., Gong, B., Zhao, Z.-F., Gao, T.-S., Chen, B., Wu, Y.-B., 2007a. Oxygen isotope geochemistry of ultrahigh-pressure metamorphic rocks from 200-4000 m core samples of the Chinese Continental Scientific Drilling. Chem. Geol. 242, 51-75. Chen, R.-X., Zheng, Y.-F., Gong, B., Zhao, Z.-F., Gao, T.-S., Chen, B., Wu, Y.-B., 2007b. Origin of retrogra e fl ui in u trahigh-pressure metamorphic rocks: constraints from mineral hydrogen isotope and water content changes in eclogite-gneiss transitions in the Sulu orogen. Geochim. Cosmochim. Acta 71, 2299 –2325. Chen, R.-X., Zheng, Y.-F., Xie, L.-W., 2010. Metamorphic growth and recrystallization of zircon: distinction by simultaneous in-situ analyses of trace elements, U-Th-Pb and Lu-Hf isotopes in zircon from eclogite-facies rocks in the Sulu orogen. Lithos 114, 132-154. Chen, Y.-X., Zheng, Y.-F., Chen, R.-X., Zhang, S.-B., Li, Q.L., Dai, M.L., Chen, L., 2011. Metamorphic growth and recrystallization of zircons in extremely

18

O-depleted rocks during

eclogite-facies metamorphism: evidence from U-Pb ages, trace elements, and O-Hf isotopes. Geochim. Cosmochim. Acta 75, 4877-4898. Chen, R.-X., Zheng, Y.-F., Hu, Z.-C., 2012a. Episodic fluid action during exhumation of deeply subducted continental crust :Geochemical constraints from zoisite-quartz vein and host metabasite in the Dabie orogen. Lithos 155, 146-166. Chen, D.-L., Liu, L., Sun, Y., Sun, W.-D., Zhu, X.-H., Liu, X.-M., Guo, C.-L., 2012b. Felsic veins within UHP eclogite at Xitieshan in North Qaidam, NW China: Partial melting during

exhumation. Lithos 136-139, 187-200. Chen, Y.-X., Zheng, Y.-F., Hu, Z., 2013a. Synexhumation anatexis of ultrahigh-pressure metamorphic rocks: Petrological evidence from granitic gneiss in the Sulu orogen. Lithos 156-159, 69-96. Chen, Y.-X., Zheng, Y.-F., Hu, Z.C., 2013b. Petrological and zircon evidence for anatexis of UHP quartzite during continental collision in the Sulu orogen. J. Metamorph. Geol. 31, 389-413. Chen, R.-X., Ding, B.H., Zheng, Y.-F., Hu, Z., 2015. Multiple episodes of anataxis in a collisional orogen: zircon evidence from migmatite in the Dabie orogen. Lithos 212-215, 247-265. Chen, R.-X., Li, H.-Y., Zheng, Y.-F., Zhang, L., Gong, B., Hu, Z., Yang, Y., 2017. Crust-mantle interaction in a continental subduction channel: evidence from orogenic peridotites in North Qaidam, Northern Tibet. J. Petrol. doi:10.1093/petrology/egx011 Cherniak, D.J., Watson, E.B., 2003. Diffusion in zircon. Rev. Mineral. Geochem. 53, 113-143. Cherniak, D.J., Watson, E.B., 2007. Ti diffusion in zircon. Chem. Geol. 242, 473–48 6. Clemens, J.D., 2006. Melting of the continental crust: fluid regimes, melting reactions, and source-rock fertility. In: Brown, M., Rushmer, T. (Eds.), Evolution and Differentiation of the Continental Crust. Cambridge University Press, Cambridge, pp. 297–331. Davies, J.H.F.L., Stern, R.A., Heaman, M.L., Rojas, X., Walton, E.L., 2015. Resolving oxygen isoto i

isturban e in zir on: A ase stu y fro

the o

δ 18O Scourie dikes, NW Scotland.

Am. Mineral. 100, 1952-1966. Degeling, H., Eggins, S., Ellis, D.J., 2001. Zr budgets for metamorphic reactions, and the formation of zircon from garnet breakdown. Mineral. Mag. 65, 749-758. Dubińskaa, E., By inab, P., Koz o skia, A. Dorr, W., Nejbert, k., S hastok, J., Ku i ki, C., 2004. U-Pb dating of serpentinization: hydrothermal zircon from a metasomatic rodingite shell (Sudetic ophiolite, SW Poland). Chem. Geol. 203, 183-203. Ferry, J.M., Watson, E.B., 2007. New thermodynamic models and revised calibrations for the Ti-in-zircon and Zr-in-rutile thermometers. Contrib. Mineral. Petrol. 154, 429-437. Ferriss, E.D.A., Essene, E.J., Becker, U., 2008. Computational study of the effect of pressure on the Ti-in-zircon geothermometer. Eur. J. Mineral. 20, 745-755. Flowerdew, M.J., Millar, I.L., Vaughan, A.P.M., Horstwood, M.S.A., Fanning, C.M., 2006. The sour ce of granitic gneisses and migmatites in the Antarctic peninsular: a combined U–Pb SHRIMP and laser ablation HF isotope study of complex zircons. Contrib. Mineral. Petrol. 151, 751–768. Fraser, G., Ellis, D., Eggins, S., 1997. Zirconium abundance in granulite-facies minerals, with implications for zircon geochronology in high-grade rocks. Geology 25, 607-610. Gao, X.-Y., Zheng, Y.-F., Chen, Y.-X., 2012. Dehydration melting of ultrahigh-pressure eclogite in

the Dabie orogen: evidence from multiphase solid inclusions in garnet. J. Metamorph. Geol. 30, 193-212. Gao, X.-Y., Zheng, Y.-F., Chen, Y.-X., Tang, H., Li, W.-C., 2015, Zircon geochemistry records the action of metamorphic fluid on the formation of ultrahigh-pressure jadeite quartzite in the Dabie orogen. Chem. Geol. 419, 158–175. Gebauer, D., 1996. A P-T-t path for a (ultra?-) high-pressure ultrarmafic/mafic rock associations and their felsic country-rocks based on SHRIMP-dating of magmatic and metamorphic zircon domains. Example: Alpe Arami (Central Swiss Alps). in A. Basu, & S. Hart, eds. Sp ecial A GU-monog raph d edicated to Profs. Tilton and Tatsumoto: Earth Processes: Reading the Isotopic Code , Geophysical Monograph Series, 95, 307–329. Gebauer, D., Schertl, H.P., Brix, M., Schreyer, W., 1997. 35 Ma old ultrahigh-pressure metamorphism and evidence for very rapid exhumation in the Dora Maira Massif, Western Alps. Lithos 41, 5–24 Geisler, T., Schaltegger, U., Tomaschek, F., 2007. Re-equilibration of zircon in aqueous fluids and melts. Elements 3, 43-50. Gillet, P., Ingrin, J., Chopin, C., 1984. Coesite in subducted continental crust: P–T history deduced from an elastic model. Earth Planet. Sci. Lett. 70, 426–436. Gilotti, J.A., McClelland, W.C., Wooden, J.L., 2014. Zircon captures exhumation of an ultrahigh-pressure terrane, North-East Greenland Caledonides. Gond. Res. 25, 235-256. Gordon, S.M., Whitney, D.L., Teyssier, C., Fossen, H., 2013. U-Pb dates and trace-element geochemistry of zircon from migmatite, Western Gneiss Region, Nor-way: Significance for history of partial melting in continental subduction. Lithos 170-171, 35-53. Green, T.H., Hellman, P.L., 1982. Fe-Mg partitioning between coexisting garnet and phengite at high pressure, and comments on a garnet-phengite geothermometer. Lithos 15, 253–266. Grieco, G., Ferrario, A., Quadt, A.V., Koeppel, V., Mathez, E.A., 2001. The zircon-bearing chromitites of the phlogopite peridotite of Finero (Ivrea zone, Southern Alps): evidence and geochronology of a metasomatized mantle slab. J. Petrol. 42, 89-101. Hanchar, J.M., Watson, E.B., 2003. Zircon saturation thermometry. Rev. Mineral. Geochem. 53, 89-112. Harley, S.L., Kelly, N.M., Moller, A., 2007. Zircon behavior and the thermal histories of mountain chains. Elements 3, 25-30. Harley, S.L., Kelly, N.M., 2007a. Zircon tiny but timely. Elements 3, 13-18. Harley, S.L., Kelly, N.M., 2007b. The impact of zircon-garnet REE distribution data on the interpretation of zircon U-Pb ages in complex high-grade terrains: an example from the Rauer Islands, East Antarctica. Chem. Geol. 241, 62-87.

Harrison, T.M., Watson, E.B., 1983. Kinetics of zircon dissolution and zirconium diffusion in granitic melts of variable water content. Contributions to Mineralogy and Petrology 84, 66-72. Harrison, T.M., Watson, E.B., Aikman, A.B., 2007. Temperature spectra of zircon crystallization in plutonic rocks. Geology 35, 635-638. Hayden, L.A., Watson, E.B., Wark, D.A., 2008. A thermobarometer for sphene (titanite). Contrib. Mineral. Petrol. 155, 529-540. He, Q., Zhang, S.B., Zheng, Y.-F., 2016. High temperature glacial meltwater-rock reaction in the Neoproterozoic: evidence from zircon in-situ oxygen isotopes in granitic gneiss from the Sulu orogen. Precamb. Res. 284, 1-13. Hermann, J., Rubatto, D., Korsakov, A., Shatsky, V., 2001. Multiple zircon growth during fast exhumation of diamondiferous, deeply subducted continental crust (Kokchetav Massif, Kazakhstan). Contrib. Mineral. Petrol. 141, 66-82. Hermann, J., 2002. Allanite: thorium and light rare earth element carrier in subducted crust. Chem. Geol. 192, 289-306. Hermann, J., Spandler, C., Hack, A., Korsakov, A.V., 2006a. Aqueous fluids and hydrous melts in high-pressure and ultra-high pressure rocks: implications for element transfer in subduction zones. Lithos 92, 399-417. Hermann, J., Rubatto, D., Trommsdorff, V., 2006b, Sub-solidus Oligocene zircon formation in garnet peridotite during fast decompression and fluid infiltration (Duria, Central Alps), Mineral. Petrol., 88, 181-206. Hermann, J., Spandler, C., 2008. Sediment melts at sub-arc depths: An experimental study. J. Petrol. 49, 717–740. Hermann, J., Zheng, Y.-F., Rubatto, D., 2013. Deep fluids in subducted continental crust. Elements 9, 281-287. Holdaway, M., 2000. Application of new experimental and garnet Margules data to the garnet-biotite geothermometer. Am. Mineral. 85, 881–892. Holland, T.J.B., 1980. The reaction albite = jadeite + quartz determined experimentally in the range 600–1200 degrees C. Am. Mineral. 65, 129–134. Hoskin, P.W.O., Black, L.P., 2000. Metamorphic zircon formation by solid-state recrystallization of protolith igneous zircon. J. Metamorph. Geol. 18, 423-439. Hoskin, P.W.O., Ireland, T.R., 2000. Rare earth element chemistry of zircon and its use as a provenance indicator. Geology 7, 627-630. Hoskin, P.W.O., Schaltegger, U., 2003. The composition of zircon and igneous and metamorphic petrogenesis. Rev. Mineral. Geochem. 53, 27-55. Hoskin, P.W.O., 2005. Trace-element composition of hydrothermal zircon and the alteration of

Hadean zircon from the Jack Hills, Australia. Geochim. Cosmochim. Acta 69, 637-648. Huang, W.L., Wyllie, P.J., 1981, Phase relationships of S-type granite with H2O to 35 kbar: Muscovite granite from Harney Peak, South Dakota. J. Geophys. Res. 86, 10515–10529. Jamieson, R.A., Unsworth, M.J., Harris, N.B.W., Rosenberg, C.L., Schulmann, K., 2011. Crustal melting and the flow of mountains. Elements 7, 253-260. Johannes, W., Holtz, F., 1996. Petrogenesis and Experimental Petrology of Granitic Rocks. Springer, Berlin Heidelberg. 335 pp. Katayama, I., Muko, A., Iizuka, T., Maruyama, S., Terada, K., Tsutsumi, Y., Sano, Y., Zhang, R.Y., Liou, J.G., 2003. Dating of zircon from Ti-clinohumite-bearing garnet peridotite: Implication for timing of mantle metasomatism. Geology 31, 713-716. Keay, S., Lister, G., Buick, I., 2001. The timing of partial melting, Barrovian metamorphism and granite intrusion in the Naxos metamorphic core complex, Cyclades, Aegean Sea, Greece. Tectonophysics 342, 275-312. Kemp, A.I.S., Hawkesworth, C.J., Paterson, B.A., Kinny, P.D., 2006. Episodic growth of the Gondwana supercontiental from hafnium and oxygen isotopes in zircon. Nature 439, 580-583. Kelly, N.M., Harley, S.L., 2005. An integrated microtextural and chemical approach to zircon geochronology: refining the Archean history of the Napier Complex, east Antarctica. Contrib. Mineral. Petrol. 149, 57-84. Kelsey, D.E., Clark, C., Hand, M. 2008. Thermobarometric modelling of zircon and monazite growth in melt-bearing systems: examples using model metapelitic and metapsammitic granulites. J. Metamorph. Geol. 26, 199–212. Kelsey, D.E., Powell, R., 2011. Progress in linking accessory mineral growth and breakdown to major mineral evolution in metamorphic rocks: a thermody-namic approach in the Na2O-CaO-K2O-FeO-MgO-Al2O3-SiO2-H2O-TiO2 -ZrO2 system. J. Metamorph. Geol. 29, 151–166. Kelsey, D.E., Hand, M., 2015. On ultrahigh temperature crustal metamorphism: phase equilibria, trace element thermometry, bulk composition, heat sources, timescales and tectonic settings. Geo. Front. 6, 311–356. Kinny, P.D., Maas, R., 2003. Lu-Hf and Sm-Nd isotope systems in zircon. Rev. Mineral. Geochem. 53, 327-341. Kohn, M.J., Corrie, S.L., Markley, C., 2015. The fall and rise of metamorphic zircon. Am. Mineral. 100, 897-908. Kohn. M.J., 2016. Metamorphic chronology–a tool for all ages: Past achievements and future prospects. Am. Mineral. 101, 25-42. Kotková, J., Whitehoise, M., Schaltegger, U., D'Abzac, F.-X., 2016. The fate of zircon during

UHT-UHP

eta or his : isoto i

U/Pb, δ18O, Hf) and trace element constraints. J.

Metamorph. Geol. 34, 719-739. Labrousse, L., Prouteau, G., Ganzhorn, A.C., 2011. Continental exhumation triggered by partial melting at ultrahigh pressure. Geology 39, 1171-1174. Labrousse, L., Duretz, T., Gerya, T., 2015. H2O-fluid-saturated melting of subducted continental crust facilitates exhumation of ultrahigh-pressure rocks in continental subduction zones. Earth and Planetary Science Letters 428, 151-161. Li, S.G., Jagoutz, E., Lo, C.-H., Chen, Y.Z., Li, Q.L., Xiao, Y.L., 1999. Sm/Nd, Rb/Sr, and 40 Ar/39Ar isotopic systematics of the ultrahigh-pressure metamorphic rocks in the Dabie-Sulu belt, Central China: A retrospective view. Intern. Geol. Rev. 41, 1114-1124. Li, X.-P., Zheng, Y.-F., Wu, Y.-B., Chen, F.K., Gong, B., Li, Y.-L., 2004. Low-T eclogite in the Dabie terrane of China: etro ogi a an isoto i

onstrain ts on flui activity and radiometric

dating. Contrib. Mineral. Petrol. 148, 443–470. Li, T.F., Yang, J.S., Zhang, R.Y., 2008. Geochemical characteristics, UHP metamorphic age, and genesis of the Hujialin garnet clinopyroxenite, Sulu Terrane, China. Intern. Geol. Rev. 50, 48-60. Li, W.-C., Chen, R.-X., Zheng, Y.-F., Li, Q.L., Hu, Z.C., 2013. Zirconological tracing of transition between aqueous fluid and hydrous melt in the crust: Constraints from pegmatite vein and host gneiss in the Sulu orogen. Lithos 162-163, 157-174. Li, W.-C., Chen, R.-X., Zheng, Y.-F., Hu, Z., 2014. Dehydration and anatexis of UHP metagranite during continental collision in the Sulu orogen. J. Metamorph. Geol. 32, 915-936. Li, W.-C., Chen, R.-X., Zheng, Y.-F., Tang, H., Hu, Z., 2016a. Two episodes of partial melting in ultrahigh-pressure migmatittes from deeply subducted continental crust in the Sulu orogen, China. Geol. Soc. Am. Bull. 128, 1521-1542. Li, H.-Y., Chen, R.-X, Zheng, Y.-F., Hu, Z., 2016b. The crust-mantle interaction in continental subduction channels: Zircon evidence from orogenic peridotite in the Sulu orogen. J. Geophys. Res. 121, 1-26. Liati, A., Franz, L., Gebauer, D., Fanning, C.K., 2004. The timing of mantle and crustal events in South Namibia, as defined by SHRIMP-dating of zircon domains from a Grt peridotite xenolith of the Gibeon Kimberlite Province. J. African Earth Sci. 39, 147-157. Liati, A., Gebauer, D., 2009. Crustal origin of zircon in a garnet peridotite: a study of U-Pb SHRIMP dating, mineral inclusions and REE geochemistry (Erzgebirge, Bohemian Massif). Eur. J. Mineral. 21, 737-750 Liou, J.G., Zhang, R.Y., 1996. Occurrences of intergranular coesite in ultrahigh-P rocks from the Sulu region, eastern China: Implications for lack of fluid during exhumation. Am. Mineral. 81,

1217-1221. Liou, J.G., Ernst, W.G., Zhang, R.Y., Tsujimori, T., Jahn, B.M., 2009. Ultrahigh-pressure minerals and metamorphic terranes-The view from China. J. Asian Earth Sci. 35, 199-231. Liu, J.B., Ye, K., Maruyama, S., Cong, B.L., Fan, H.R., 2001. Mineral inclusions in zircon from gneisses in the ultrahigh-pressure zone of the Dabie Mountains, China. J. Geol. 109, 523–535. Liu, J.B., Ye, K., 2004. Transformation of garnet epidote amphibolite to eclogite, western Dabie Mountains, China. J. Metamorph. Geol. 22, 383– 394. Liu, D., Jian, P., Kroner, A., Xu, S., 2006. Dating of prograde metamorphic events deciphered from episodic zircon growth in rocks of the Dabie-Sulu UHP complex, China. Earth Planet. Sci. Lett. 250, 650-666. Liu, Y.C., Li, S.G, 2008. Detachment within subducted continental crust and multi-slice successive exhumation of ultrahigh-pressure metamorphic rocks: Evidence from the Dabie-Sulu orogenic belt. Chin. Sci. Bull. 53, 3105-3119. Liu, F.L., Gerdes, A., Zeng, L.S., Xue, H.M., 2008. SHRIMP U–Pb dating, trace elements and the Lu–Hf isotope system of coesite-bearing zircon from amphibolite in the SW Sulu UHP terrane, eastern China. Geochim. Cosmochim. Acta 72, 2973-3000. Liu, F.L., Gerdes, A., Xue, H.M., 2009a. Differential subduction and exhumation of crustal slices in the Sulu HP-UHP metamorphic terrane: insights from mineral inclusions, trace elements, U-Pb and Lu-Hf isotope analyses of zircon in orthogneiss. J. Metamorph. Geol. 27, 805-825. Liu, F.L., Gerdes, A., Liou, J.G., Liu, P.F., 2009b. Unique coesite-bearing zircon from allanite-bearing gneisses: U-Pb, REE and Lu-Hf properties and implications for the evolution of the Sulu UHP terrane, China. Eur. J. Mineral. 21, 1225-1250. Liu, F.L., Xue, H., Liu, P., 2009c. Partial melting time of ultrahigh-pressure metamorphic rocks in the Sulu UHP terrane: constrained by zircon U-Pb ages, trace elements and Lu/Hf isotope compositions of biotite-bearing granite. Acta Petrol. Sin. 25, 1039-1055 (in Chinese with English abstract). Liu, F.L., Wang, F., Liu, P., 2009d. Genetic relationship between pegmatite formation and anatexis of ultrahigh-pressure (UHP) metamorphic rocks in the Weihai area, North Sulu UHP terrane. Acta Geol. Sin. 83, 1687-1702 (in Chinese with English abstract). Liu, F.L., Robinson, P.T., Gerdes, A., Xue, H.M., Liu, P.H., Liou, J.G., 2010a. Zircon U-Pb ages, REE concentrations and Hf isotope compositions of granitic leucosome and pegmatite from the north Sulu UHP terrane in China: Constraints on the timing and nature of partial melting. Lithos 117, 247-268. Liu, Y.S., Gao, S., Hu, Z.C., Gao, C.G., Zong, K.Q., Wang, D. 2010b. Continental and oceanic crust recycling-induced melt-peridotite interactions in the Trans-North China Orogen: U-Pb dating,

Hf isotopes and trace elements in zircons from mantle xenoliths. J. Petrol. 51, 537-571. Liu, F.L., Liou, J.G., 2011. Zircon as the best mineral for P-T-time history of UHP metamorphism: A review on mineral inclusions and U-Pb SHRIMP ages of zircons from the Dabie-Sulu UHP rocks. J. Asian Earth Sci. 40, 1-39. Liu, F.L., Robinson, P.L., Liu, P.H., 2012. Multistage partial melting events in the Sulu UHP terrane: zircon U–Pb dating of granitic leucosomes within amphibolite and gneiss. J. Metamorph. Geol. 30, 887-906. Liu, Y.-C., Deng, L.-P., Gu, X.-F., Groppo, X., Rolfo, F., 2015. Application of Ti-in-zircon and Zr-in-rutile thermometers to constrain high-temperature metamorphism in eclogites from the Dabie orogen, central China. Gond. Res. 27, 410-423. Martin, L.A.J., Duchene, S., Deloule, E., Vanderhaeghe, O., 2008. Mobility of trace elements and oxygen in zircon during metamorphism: consequences for geochemical tracing. Earth Planet. Sci. Lett. 267, 161-174. McClelland, W.C., Lapen, T.J., 2013. Linking time to the pressure-temperature path for ultrahigh-pressure rocks. Elements 9, 273-279. Miller, C.F., McDowell, S.M., Mapes, R.W., 2003. Hot and cold granites? Implications of zircon saturation temperatures and preservation of inheritance. Geology 31, 529–532. Nakamura, D., Hirajima, T., 2000, Granulite-facies overprinting of ultrahigh-pressure metamorphic rocks, northeastern Su-Lu region, eastern China. J. Petrol. 41, 563–582. Nasdala, L., Wenzel, M., Vavra, G., Irmer, G., Wenzel, T., Kober, B., 2001. Metamictization of natural zircon: accumulation versus thermal annealing of radioactivity-induced damage. Contrib. Mineral. Petrol. 141, 125-144. Nemchin, A.A, Grange, M.L., Pidgeon, R.T., Meyer, C., 2012. Lunar zirconology. Aust. J. Earth Sci. 59, 277-290. Okay, A.I., Xu, S.-T., Sengor, A.M.C., 1989. Coesite from the Dabie Shan eclogites, central China. Eur. J. Mineral. 1, 595–598. Palme, H., O'Neill, H. St. O., 2003. Cosmochemical constraints of mantle composition, Treatise on Geochemistry 2, 1-38. Pan, Y., 1997. Zircon- and monazite-forming metamorphic reactions at Manitouwadge, Ontario. Can. Mineral. 35, 105-118. Pettke, T., Audétat, A., Schaltegger, U., Heinrich, C.A., 2005. Magmatic-to-hydrothermal crystallization in the W-Sn mineralized Mole Granite (NSW, Australia): part II: evolving zircon and thorite trace element chemistry. Chem. Geol. 220, 191-213. Ragozin, A.L., Liou, J.G., Shatsky, V.S., Sobolev, N.V., 2009. The timing of the retrograde partial melting in the Kumdy-Kol region (Kokchetav Massif, Northern Kazakhstan). Lithos 109,

274-284. Ravna, E.K., 2000. Distribution of Fe2+ and Mg between coexisting garnet and hornblende in synthetic and natural systems: An empirical calibration of the garnet-hornblende Fe-Mg geothermometer. Lithos 53, 265–277. Rayner, N., Stern, R.A., Carr, S.D., 2005. Grain-scale variations in trace element composition of fluid-altered zircon, Acasta Gneiss Complex, northwestern Canada. Contrib. Mineral. Petrol. 148, 721-734. Reed, M.J., Candela, P.A., Piccoli, P.A., 2000. The distribution of rare earth elements between monzogranitic melt and the aqueous volatile phase in experimental investigations at 800°C and 200 MPa. Contrib. Mineral. Petrol. 140, 251–262. Robinson, P.T., Trumbull, R.B., Schmitt, A., Yang, J.S., Li, J.W., Zhou, M.F., Erzinger, J., Dare, S., Xiong, F., 2015. The origin and significance of crustal minerals in ophiolitic chromitites and peridotites. Gond. Res. 27, 485-506. Rollinson, H.R., Windley, B.F., 1980. Selective elemental depletion during metamorphism of Archean granulites. Contrib. Mineral. Petrol. 72, 257-263. Rowley, D.B., Xue, F., Tucker, R.D., Peng, Z.X., Baker, J., Davis, A., 1997. Ages of ultrahigh pressure metamorphism and protolith orthogneisses from the eastern Dabie Shan: U/Pb zircon geochronology. Earth Planet. Sci. Lett. 151, 191-203. Rubatto, D., Gebauer, D., Compagnoni, R., 1999. Dating of eclogite-facies zircons: the age of Alpine metamorp hism in the Sesia-Lanzo Zone (Western Alps). Earth Planet. Sci. Lett. 167, 141–158. Rubatto, D., 2002. Zircon trace element geochemistry: partitioning with garnet and the link between U-Pb ages and metamorphism. Chem. Geol. 184, 123-138. Rubatto, D., Hermann, J., 2003. Zircon formation during fluid circulation in eclogites (Monviso, Western Alps): implications for Zr and Hf budget in subduction zones. Geochim. Cosmochim. Acta 67, 2173-2187. Rubatto, D., Hermann, J., 2007a. Zircon behavior in deeply subducted rocks. Elements 3, 31-35. Rubatto, D., Hermann, J., 2007b. Experimental zircon/melt and zircon/garnet trace element partitioning and implications for the geochronology of crustal rocks. Chem. Geol. 241, 38-61. Rubatto, D., Müntener, O., Barnhoorn, A., Gregory, C., 2008. Dissolution–reprecipitation of zircon at low-temperature, high-pressure conditions (Lanzo Massif, Italy). Am. Mineral. 93, 1519-1529. Rumble, D., Giorgis, D., Ireland, T., Zhang, Z., Xu, H., Yui, T.F., Yang, J., Xu, Z., Liou, J.G., 2002. Lo

δ18O zircons, U-Pb dating, and the age of the Qinglongshan oxygen and hydrogen isotope

anomaly near Donghai in Jiangsu Province, China. Geochim. Cosmochim. Acta 66, 2299-2306.

Rumble, D., Liou, J.G., Jahn, B.-m., 2003. Continental crust subduction and ultrahigh pressure metamorphism. Treatise on Geochemistry 3, 293-319. Sassi, R., Harte, B., Carswell, D.A., Han, Y.J., 2000. Trace element distribution in central Dabie eclogites. Contrib. Mineral. Petrol. 139, 298–315. Schaltegger, U., Fanning, C.M., Günther, D., Maurin, J.C., Schulmann, K., Gebauer, D., 1999. Growth, annealing and recrystallization of zircon and preservation of monazite in high-grade metamorphism:

conventional

and

in-situ

U-Pb

isotope,

cathodoluminescence

and

microchemical evidence. Contrib. Mineral. Petrol. 134, 186-201. Schaltegger, U., 2007. Hydrothermal zircon. Elements 3, 51. Scherer, E.E., Whitehouse, M.J., Munker, C., 2007. Zircon as a monitor of crustal growth. Elements 3, 19-24. Shatsky, V.S., Jagoutz, E., Sobo ev, N.V., Koz’ enko, O.A., Parkho enko, V.S., Troes h, M., 1999. Geochemistry and age of ultrahigh pressure metamorphic rocks from the Kokchetav massif (Northern Kazakhstan). Contrib. Mineral. Petrol. 137, 185–205. Sheng, Y.-M., Zheng, Y.-F., Chen, R.-X., Li, Q.L., Dai, M.N., 2012. Fluid action on zircon growth and recrystallization during quartz veining within UHP eclogite: Insights from U-Pb ages, O–Hf isotopes and trace elements. Lithos 136-139, 126-144. Sheng, Y.-M., Zheng, Y.-F., Li, S.N., Hu, Z.C., 2013. Element mobility during continental collision: insights from polymineralic metamorphic vein within UHP eclogite in the Dabie orogen. J. Metamorph. Geol. 31, 221-241. S á a, J., Koš er, J., Pe ersen, R.B., 2007. Behaviour of zir on in high-grade metamorphic rocks: evidence from Hf isotopes, trace elements and textural studies. Contrib. Mineral. Petrol. 154, 335-356. Song, S.G., Zhang, L.F., Niu, Y.L., Su, L., Jian, P., Liu, D.Y., 2005. Geochronology of diamond-bearing zircons from garnet peridotite in the North Qaidam UHPM belt, Northern Tibetan Plateau: a record of complex histories from oceanic lithosphere subduction to continental collision. Earth Planet. Sci. Lett. 234, 99-118. Song, Y., Xu, H., Zhang, J., Wang, D., Liu, E., 2014. Syn-exhumation partial melting and melt segregation in the Sulu UHP terrane: Evidences from leucosome and pegmatitic vein of migmatite. Lithos 202–203, 55–75. Stepanov, A.S., Rubatto, D., Hermann, J., Korsakov, A.V., 2016. Constrasting P-T paths within the Barchi-Kol UHP terrain (Kokchetav complex): implications for subduction and exhumation of continental crust. Am. Mineral. 101, 788-807.Tang, J., Zheng, Y.-F., Wu, Y.-B., Gong, B., Zha, X.P., Liu, X.-M., 2008a. Zircon U-Pb age and geochemical constraints on the tectonic affinity of the Jiaodong terrane in the Sulu orogen, China. Precambr. Res. 161, 389-418.

Tang, J., Zheng, Y.-F., Gong, B., Wu, Y.-B., Gao, T.-S., Yuan, H., Wu, F.-Y., 2008b. Extreme oxygen isotope signature of meteoric water in magmatic zircon from metagranite in the Sulu orogen, China: Implications for Neoproterozoic rift magmatism. Geochim. Cosmochim. Acta 72, 3139–3169. Taylor, R.J.M., Kirkland, C.L., Clark, C., 2016. Accessories after the facts: constraining the timing, duration and conditions of high-temperature metamorphic processes. Lithos 264, 239-257. Tichomirowa, M., Sergeev, S., Berger, H-J., Leonhardt, D., 2012. Inferring protoliths of high-grade metamorphic gneisses of the Erzgebirge using zirconology, geochemistry and comparison with lower-grade rocks from Lusatia (Saxothuringia, Germany). Contrib. Mineral. Petrol. 164, 375-396. Tomaschek, F., Kennedy, A.K., Villa, I.M., Markus, L., Chris, B., 2003. Zircons from Syros, Cyclades, Greece-recrystallization and mobilization of zircon during high-pressure metamorphism. J. Petrol. 44, 1977-2002. Tomkins, H., Powell, R., Ellis, D., 2007. The pressure dependence of the zirconium-in-rutile thermometer. J. Metamorph. Geol. 25, 703–71 Valley, J.W., Kinny, P.D., Schulze, D.J., Spicuzza, M.J., 1998. Zircon megacrysts from kimberlite: oxygen isotope variability among mantle melts. Contrib. Mineral. Petrol. 133, 1-11. Valley, J.W., 2003. Oxygen isotopes in zircon. Rev. Mineral. Geochem. 53, 343-385. Van DerMolen, I., Van Roermund, H.L.M., 1986. The pressure path of solid inclusions in minerals: the retention of coesite inclusions during uplift. Lithos 19, 317–324. Vavra, G., Gebauer, D., Schmid, R., 1996. Multiple zircon growth and recrystallization during polyphase Late Carboniferous to Triassic metamorphism in granulites of the Ivrea Zone (Southern Alps): an ion microprobe (SHRIMP) study. Contrib. Mineral. Petrol. 122, 337-358. Vavra, G., Schmid, R., Gebauer, D., 1999. Internal morphology, habit and U–Th–Pb microanalysis of amphibole to granulite facies zircon: geochronology of the Ivrea Zone (Southern Alps). Contrib. Mineral. Petrol. 134, 380-404. Vielzeuf, D., Holloway, J.R., 1988. Experimental de-termination of the fluid-absent melting relations in the pelitic system. Contrib. Mineral. Petrol. 98, 257–276. Vonlanthen, P., Gerald, J.D.F., Rubatto, D., Hermann, J., 2012. Recrystallization rims in zircon (Valle d'Arbedo, Switzerland): an integrated cathodoluminescence, LAICP-MS, SHRIMP, and TEM study. Am. Mineral. 97, 369-377. Wallis, S., Tsuboi, M., Suzuki, K., Fanning, M., Jiang, L.L., Tanaka, T., 2005. Role of partial melting in the evolution of the Sulu (eastern China) ultrahigh-pressure terrane. Geology 33, 129-132. Wang, X., Griffin, W.L., 2004. Unusual Hf contents in metamorphic zircon from coesite-bearing

eclogites of the Dabie Mountains, east-central China: implications for the dating of ultrahigh–pressure metamorphism. J. Metamorph. Geol. 22, 629-637. Wang, L., Kusky, T.M., Polat, A., Wang, S., Jiang, X., Zong, K., Wang, J., Deng, H., and Fu, J., 2014. Partial melting of deeply subducted eclogite from the Sulu orogen in China. Nat. Commun. 5, 5604; doi:10.1038/ncomms6604. Watson, E.B., Harrison, T.M., 1983. Zircon saturation revisited: Temperature and composition effects in a variety of crustal magma types. Earth and Planetary Science Letters 64, 295–304. Watson, E.B., 1996. Dissolution, growth and survival of zircons during crustal fusion: kinetic principles, geological models and implications for isotopic inheritance. Trans. Roy. Soc. Edinburgh 87, 43–56. Watson, E.B., Harrison, T.M., 2005. Zircon thermometer reveals minimum melting conditions on earliest Earth. Science 308, 841–844. Watson, E.B., Wark, D.A., Thomas, J.B., 2006. Crystallization thermometers for zircon and rutile. Contrib. Mineral. Petrol. 151, 413-433. Whitehouse, M.J., Platt, J.P., 2003. Dating high-grade metamorphism-constraints from the rare-earth elements in zircons and garnet. Contrib. Mineral. Petrol. 145, 61-74. Wu, Y.-B., Zheng, Y.F., 2004. Genesis of zircon and its constraints on interpretation of U-Pb age. Chin. Sci. Bull. 49, 1554-1569. Wu, Y.-B., Zheng, Y.-F., Zhao, Z.-F., Gong, B., Liu, X., Wu, F.-Y., 2006. U/Pb, Hf and O isotope evidence for two episodes of fluid-assisted zircon growth in marble-hosted eclogites from the Dabie orogen. Geochim. Cosmochim. Acta 70, 3743-3761. Wu, Y.-B., Zheng, Y.-F., Zhang, S.-B., Zhao, Z.-F., Wu, F.-Y., Liu, X.-M., 2007. Zircon U-Pb ages and Hf isotope compositions of migmatite from the North Dabie terrane in China: constraints on partial melting. J. Metamorph. Geol. 25, 991-1009. Wu, Y.-B., Gao, S., Zhang, H.F., Yang, S.H., Liu, X.C., Jiao, W.F., Liu, Y.S., Yuan, H.L., Gong, H.J., He, M.C., 2009. U-Pb age, trace-element, and Hf-isotope compositions of zircon in a quartz vein from eclogite in the western Dabie Mountains: constraints on fluid flow during early exhumation of ultrahigh-pressure rocks. Am. Mineral. 94, 303-312. Xia, Q.-X., Zheng, Y.-F., Yuan, H.L., Wu, F.Y., 2009. Lu-Hf and U-Th-Pb isotope distinction between metamorphic growth and recrystallization of zircon from eclogite-facies metagranite in the Dabie orogen. Lithos 112, 477-496. Xia, Q.-X., Zheng, Y.-F., Hu, Z.C., 2010. Trace elements in zircon and coexisting minerals from low-T/UHP metagranite in the Dabie orogen: implications for action of supercritical fluid during continental subduction-zone metamorphism. Lithos 114, 385-412. Xia, Q.-X., Zheng, Y.-F., Chen, Y.-X., 2013. Protolith control on fluid availability for zircon growth

during continental subduction-zone metamorphism in the Dabie orogen. J. Asian Earth Sci. 67-68, 93-113. Xiong, Q., Zheng, J.P., Griffin, W.L., O'Reilly, S.Y., Zhao, J.H., 2011. Zircons in the Shenglikou ultrahigh-pressure garnet peridotite massif and its country rocks from the North Qaidam terrane (western China): Meso-Neoproterozoic crust–mantle coupling and early Paleozoic convergent plate-margin processes. Precamb. Res. 187, 33-57. Xiong, Q., Zheng, J.P., Griffin, W.L., O'Reilly, S.Y., Pearson, N.J., 2014. Pyroxenite dykes in orogenic peridotite from North Qaidam (NE Tibet, China) Track metasomatism and segregation in the mantle wedge. J. Petrol. 35, 2347-2376. Xu, S.-T., Okay, A.I., Ji, S.Y., Sengor, A.M.C., Wen, S., Liu, Y.C., Jiang, L.L., 1992. Diamond from the Dabie–Shan metamorp hic rocks and its implication for tectonic setting. Science 256, 80–82. Xu, Z.-Q., Zeng, L.-S., Liu, F.-L., Yang, J.-S., Zhang, Z.-M., McWilliams, M., Liou, J.G., 2006. Polyphase subduction and exhumation of the Sulu high-pressure-ultrahigh pressure metamorphic terrane. Geol. Soc. of Am. Special Paper 403, 93-113. Xu, H.J., Ye, K., Song Y, Chen, Y., Zhang, J., Liu, Q., Guo, S., 2013. Prograde metamorphism, decompressional partial melting and subsequent melt fractional crystallization in the Weihai migmatitic gneisses, Sulu UHP terrane, eastern China. Chem. Geol. 341, 16-37. Yamamoto, S., Komiya, T., Yamamoto, H., Kaneko, Y., Terabayashi, M., Katayama, I., Iizuka, T., Maruyama, S., Yang, J.S., Kon, Y., Hirata, T., 2013. Recycled crustal zircons from podiform chromitites in the Luobusa ophiolite, southern Tibet. Island Arc 22, 89-103. Yang, J.H., Chung, S.L., Wilde, S.A., Wu, F.Y., Chu, M.F., Lo, C.H., Fan, H.R., 2005. Petrogenesis of post-orogenic syenites in the Sulu Orogenic Belt, East China: geochronological, geochemical and Nd–Sr isotopic evidence. Chem. Geol. 214, 99-125. Yang, J.S., Li, T.F., Chen, S.Z., Wu, C.L., Robinson, P.T., Liu, D.Y., Wooden, J.L., 2009. Genesis of garnet peridotites in the Sulu UHP belt: examples from the Chinese continental scientific drilling project - main hole, PP1 and PP3 drillholes. Tectonophysics, 475, 359-382. Yu, S.Y., Zhang, J.X., Mattinson, C.G., Del Real, P.G., Li, Y.S., Gong, J.H., 2014. Paleozoic HP granulite-facies metamorphism and anatexis in the Dulan area of the North Qaidam UHP terrane, western China: constraints from petrology, zircon U–Pb and amphibole Ar–Ar geochronology. Lithos 198, 58–76. Yu, S.Y., Zhang, J.X., Sun, D. Padblo Garcia del R., Li, Y., Zhao, X., Hou, K., 2015. Petrology, geochemistry, zircon U-Pb dating and Lu-Hf isotope of granitic leucosomes within felsic gneiss from the North Qaidam UHP terrane: constraints on the timing and nature of partial melting. Lithos 218-219, 1-21.

Zack, T., Kronz, A., Foley, S.F., Rivers, T., 2002. Trace element abundances in rutiles from eclogites and associated garnet mica schists. Chem. Geol. 184, 97– 122. Zajacz, Z., Halter, W.E., Pettke, T., Guillong, M., 2008. Determination of fluid/melt partition coefficients by LA-ICPMS analysis of co-existing fluid- and silicate melt inclusions: controls on element partitioning. Geochim. Cosmochim. Acta 72, 2169–2197. Zeng, L.S., Gao, L.E., Yu, J.J., Hu, G.Y., 2011. SHRIMP zircon U/Pb dating on ultrahigh pressure rocks Yangkou: implications for the timing of partial melting in the Sulu UHP metamorphic belt. Acta Petrol. Sinica 2011, 27, 1085-1094 (in Chinese with English abstract). Zhang, R.Y., Liou, J.G., 1997. Partial transformation of gabbro to coesite-bearing eclogite from Yangkou, the Sulu terrane, eastern China. J. Metamorph. Geol. 15, 183-202. Zhang, R.Y., Liou, J.G., Yang, J.S., Yui, T.F., 2000. Petrochemical constraints for dual origin of garnet peridotites from the Dabie-Sulu UHP terrane, eastern-central China. J. Metamorph. Geol. 18, 149–166. Zhang, R.Y., Zhang, J.S., Wooden, J.L., Liou, J.G., 2005. U-Pb SHRIMP geochronology of zircon in garnet peridotite from the Sulu UHP terrane, China: implications for mantle metasomatism and subduction-zone UHP metamorphism. Earth Planet. Sci. Lett. 237, 729-743. Zhang, Z.M., Shen, K., Xiao, Y.L., Hoefs, J., Liou, J.G., 2006. Mineral and fluid inclusions in zircon of UHP metamorphic rocks from the CCSD-main drill hole: A record of metamorphism and fluid activity. Lithos 92, 378-398. Zhang, Z.M., Dong, X., Liou, J.G., Liu, F., Wang, W., Yui, F., 2011. Metasomatism of garnet peridotite from Jiangzhuang, southern Sulu UHP belt: constraints on the interactions between crust and mantle rocks during subduction of continental lithosphere. J. Metamorph. Geol. 29, 917-937. Zhang, S.-B., Zheng, Y.-F., 2013. Formation and evolution of Precambrain continental lithosphere in South China. Gond. Res. 23, 1241-1260. Zhang, S.-B., Tang, J., Zheng, Y.-F., 2014. Contrasting Lu-Hf isotopes in zircon from Precambrian metamorphic rocks in the Jiaodong Peninsula: constraints on the tectonic suture between North China and South China. Precamb. Res. 245, 29-50. Zhang, L., Chen, R.-X., Zheng, Y.-F., Hu, Z., 2015. Partial melting of deeply subducted continental crust during exhumation: insights from felsic veins and host UHP metamorphic rocks in North Qaidam, northern Tibet. J. Metamorph. Geol. 33, 671-694. Zhao, R., Liou, J. G., Zhang, R. Y. & Li, T., 2006. SHRIMP U-Pb zircon dating of the Rongcheng eclogite and associated peridotite: New constraints for ultrahigh-pressure metamorphism of mantle-derived

afi -u tra afi

bo ies fro

the Sulu terrane. In: Ultrahigh-pressure

Metamorphism: Deep Continental Subduction, Special Paper 403 (eds Hacker, B. R.,

McClelland, W. C. & Liou, J. G.), pp. 115–125. Geological Society of America, Boulder, CO. Zhao, Z.-F., Zheng, Y.-F., Chen, R.-X., Xia, Q.-X., Wu, Y.-B., 2007a. E e ent

obi ity in

felsic ultrahigh-pressure metamorphic rocks during continental collision.

afi and Geochim.

Cosmochim. Acta 71, 5244–5266. Zhao, R.X., Zhang, R.Y., Liou, J.G., Booth, A.L., Pope, E.C., Chamberlain, C.P., 2007b. Petrochemistry, oxygen isotopes and U-Pb SHRIMP geochronology of mafic-ultramafic bodies from the Sulu UHP terrane, China. J. Metamorph. Geol. 25, 207-224. Zhao, Z.-F., Zheng, Y.-F., 2009. Remelting of subducted continental lithosphere: petrogenesis of Mesozoic magmatic rocks in the Dabie–Sulu orogenic belt. Sci. China (D) 52, 1295-1318. Zhao, Z.F., Zheng, Y.-F., Zhang, J., Dai, L., Liu, Q., Liu, X.M., 2012. Syn-exhumation magmatism during continental collision: Evidence from alkaline intrusives of Triassic age in the Sulu orogen. Chem. Geol. 328, 70–88. Zheng, Y.-F., Fu, B., 1998. Estimation of oxygen diffusivity from anion porosity in minerals. Geochem. J. 32, 71–89. Zheng, Y.-F., Wang, Z.-R., Li, S.-G., Zhao, Z.-F. 2002. Oxygen isotope equilibrium between eclogite minerals and its constraints on mineral Sm–Nd chronometer. Geochim. Cosmochim. Acta 66, 625–634. Zheng, Y.-F., Fu, B., Gong, B., Li, L., 2003a. Stable isotope geochemistry of ultrahigh pressure metamorphic rocks from the Dabie-Sulu orogen in China: implications for geodynamics and fluid regime. Earth Sci. Rev. 62, 105-161. Zheng, Y.-F., Yang, J.-J., Gong, B., Jahn, B.-m., 2003b. Partial equilibrium of radiogenic and stable isotope systems in garnet peridotite during ultrahigh-pressure metamorphism. Am. Mineral. 88, 1633-1643. Zheng, Y.-F., Zhao, Z.-F., Li, S.G., Gong, B. 2003c. Oxygen isotope equilibrium ultrahigh-pressure metamorphic minerals and its constraints on Sm–Nd and Rb –Sr chronometers. Geol. Soc. Spec. Publ. 220, 93–117. Zheng, Y.-F., Chen, F.-K., Gong, B., Li, L., 2004. Zircon U-Pb and oxygen isotope evidence for a large-scale 18O depletion event in igneous rocks during the Neoproterozoic. Geochim. Cosmochim. Acta 68, 4145-4165. Zheng, Y.-F., Zhou, J.-B., Wu, Y.-B., Xie, Z., 2005a. Low-grade metamorphic rocks in the Dabie –Sulu Orogenic Belt; a passive-margin accreti onary wedge deformed during continent subduction. Intern. Geol. Rev. 47, 851– 871. Zheng, Y.-F., Wu, Y.-B., Zhao, Z.-F., Zhang, S.-B., Xu, P., Wu, F.-Y., 2005b. Metamorphic effect on zircon Lu-Hf and U-Pb isotope systems in ultrahigh-pressure metagranite and metabasite. Earth Planet. Sci. Lett. 240, 378-400.

Zheng, Y.-F., Zhao, Z.-F., Wu, Y.-B., Zhang, S.-B., Liu, X.M., Wu, F.-Y., 2006a. Zircon U-Pb age, Hf and O isotope constraints on protolith origin of ultrahigh-pressure eclogite and gneiss in the Dabie orogen. Chem. Geol. 231, 135-158. Zheng, J.P., Griffin, W.L., O'Reilly, S.Y., Yang, J.S., Zhang, R.Y., 2006b. A refractory mantle protolith in younger continental crust, east-central China: Age and composition of zircon in the Sulu ultrahigh-pressure peridotite. Geology 34, 705-708. Zheng, Y.-F., Zhang, S.-B., Zhao, Z.-F., Wu, Y.-B., Li, X., Li, Z., Wu, F.-Y., 2007a. Contrasting zircon Hf and O isotopes in the two episodes of Neoproterozoic granitoids in South China: Implications for growth and reworking of continental crust. Lithos 96, 127-150. Zheng, Y.-F., Gao, T.-S., Wu, Y.-B., Gong, B., Liu, X., 2007b. Fluid flow during exhumation of deeply subducted continental crust: Zircon U-Pb age and O isotope studies of quartz vein in eclogite. J. Metamorph. Geol. 25, 267-283. Zheng, Y.-F., 2008. A perspective view on ultrahigh-pressure metamorphism and continental collision in the Dabie–Sulu or ogenic belt. Chin. Sci. Bull. 53, 3081–3104. Zheng, J.P., Sun, M., Griffin, W.L., Zhou, M.F., Zhao, G.C., Robinson, P., Tang, H.Y., Zhang, Z.H., 2008. Age and geochemistry of contrasting peridotite types in the Dabie UHP belt, eastern China: petrogenetic and geodynamic implications. Chem. Geol. 247, 282-304. Zheng, Y.-F., 2009. Fluid regime in continental subduction zones: petrological insights from ultrahigh-pressuremetamorphic rocks. J. Geol. Soc. 166, 763-782. Zheng, Y.-F., Chen, R.-X., Zhao, Z.-F., 2009. Chemical geodynamics of continental subduction-zone

eta or his : insights fro

stu ies of the Chinese Continenta S ientifi

Drilling (CCSD) core samples. Tectonophysics 475, 327–358. Zheng, Y.-F., Xia, Q.-X., Chen, R.-X., Gao, X.-Y., 2011a. Partial melting, fluid supercriticality and element mobility in ultrahigh-pressure metamorphic rocks during continental collision. Earth Sci. Rev. 107, 342-374. Zheng, Y.-F., Gao, X.-Y., Chen, R.-X., Gao, T.-S., 2011b. Zr-in-rutile thermometry of eclogite in the Dabie orogen: Constraints on rutile growth during continental subduction-zone metamorphism. J. Asian Earth Sci. 40, 427-451. Zheng, Y.-F., 2012. Metamorphic chemical geodynamics in continental subduction zones. Chem. Geol. 328, 5-48. Zheng, Y.-F., Hermann, J., 2014. Geochemistry of continental subduction-zone fluids. Earth, Planets and Space 66:93; doi:10.1186/1880-5981-66-93. Zheng, J.P., Tang, H.Y., Xiong, Q., Griffin, W.L., O’Rei y, S.Y., Pearson, N., Zhao, J.H., Wu, Y.B., Zhang, J.F., Liu, Y.S., 2014. Linking continental deep subduction with destruction of a cratonic margin: strongly reworked North China SCLM intruded in the Triassic Sulu UHP belt. Contrib.

Mineral. Petrol. 167: 1028. Zheng, Y.-F., Chen, Y.-X., 2016. Continental versus oceanic subduction zones. National Science Review 3, 495-519. Zheng, Y.F., Chen, R.X., Xu, Z., Zhang, S.B. 2016. The transport of water in subduction zones. Science China: Earth Sciences 59, 651-682. Zong, K.Q., Liu, Y.S., Hu, Z.C., Kusky, T., Wang, D.B., Gao, C.G., Gao, S., Wang, J.Q., 2010. Melting-induced fluid flow during exhumation of gneisses of the Sulu ultrahigh-pressure terrane. Lithos 120, 490-510.

Figure Captions Figure 1. Sketch map of geology in the Dabie-Sulu orogenic belt, east-central China. Abbreviations: UHP = ultrahigh pressure, HP = high pressure, LP = low pressure, HT = high temperature, MT = medium temperature, LT = low temperature, NCB = North China Block, SCB = South China Block.

Figure 2. Plane-polarized light (PL) images of mineral inclusions in zircon and CL images of zircon. (a) Metamorphic zircon in UHP eclogite (after Chen et al., 2010); (b) metamorphic zircon in UHP gneiss (after Chen et al., 2010); (c) and (d) metamorphic zircon in phengite eclogite at 1074.3 m depth of the CCSD-MH, which contains a thick low-luminescence core with Coe and Omp inclusions, and a thin bright-luminescence rim without inclusion. Note that an isolated H 2O+CO2 fluid inclusion with negative crystal shape occurs in the core of zircon and an Omp inclusion occurs nearby (after Zhang et al., 2006); (e) and (f) CL images and corresponding photomicrograph for metamorphic zircon in quartz vein (after Wu et al., 2009); metamorphic zircon as rim (g) and multi-growth new grain (h) in a quartz vein (after Chen et al., 2012a).

Figure 3. Chondrite-normalized REE patterns for zircon from eclogite, gneiss and quartz vein from the Dabie-Sulu orogenic belt. (a) Metamorphic zircon in UHP eclogite from the CCSD-MH (after Chen et al., 2010); (b) metamorphic zircon in UHP gneiss from the CCSD-MH (after Chen et al., 2010); (c) metamorphic zircon in quartz vein; (d) zircon in amphibolite from the Sulu orogen (Liu et al., 2008); (e) zircon in UHP orthogneiss from the Sulu orogen (Liu et al., 2009a); (f) zircon in UHP paragneiss from the Sulu orogen (Liu et al., 2009b).

Figure 4. Diagrams of representative trace element contents and ratios for different types of newly grown zircons from the Dabie-Sulu orogenic belt. Data source: metamorphic zircon in UHP rocks (Liu et al., 2008, 2009a; Chen et al., 2010; Li et al., 2014); metamorphic zircon in quartz vein (Wu et al., 2009; Zheng, 2009; Zheng et al., 2011b; Chen et al., 2012a; Sheng et al., 2012); relict magmatic zircon (Liu et al., 2009c, 2009d, 2010a; Chen et al., 2010, 2013a, 2013b; Sheng et al., 2012); peritectic zircon in UHP rocks, including gneiss, quartzite and migmatite (Liu et al., 2010a, 2012; Zong et al., 2010; Chen et al., 2013a, 2013b; Xu et al., 2013; Li et al., 2014, 2016); peritectic/anatectic zircon in pegmatite (Liu et al., 2009d, 2010a; Xu et al., 2013); magmatic zircon in synexhumation granite and syenite (Liu et al., 2009c; Zhao et al., 2012 and unpublished data).

Figure 5. Diagrams of U-Pb ages, Lu-Hf and O isotope compositions for zircons from UHP metamorphic rocks in the Dabie-Sulu orogenic belt. (a) Zircon Lu-Hf isotope relationships between

metamorphic zircon and peritectic zircon, and inherited magmatic zircon. Abbreviations: MZ, metamorphic zircon; RMZ, relict magmatic zircon; PZ, peritectic zircon. Data sources: quartz vein (Sheng et al., 2012); eclogite and gneiss (Chen et al., 2010); peritectic zircon and relict magmatic zircon in leucosome and pegmatite (Liu et al., 2010a). (b) Correlations of SIMS data between apparent 206Pb/238U ages an δ18O values for eclogite and gneiss (after Chen et al., 2011).

Figure 6. Cartoon showing the generation of anatectic and magmatic melts as well as peritectic, anatectic and magmatic zircons in collisional orogens. The anatectic melt is produced by partial melting of crustal rocks through dehydration/hydration reactions, without complete separation from its parental rocks. The magmatic melt is highly evolved from anatectic melts and thus has separated from their parental rocks, with significant fractional crystallization. The anatectic melts may be lowly evolved to crystallize minerals in leocosomes, veinlets and pegmatites at the final stage of their evolution. Magmatic zircon refers to zircon crystallized from magmas, anatectic zircon refers to zircon crystallized from evolved anatectic melts, and peritectic zircon refers to zircon produced by peritectic reaction during anatexis. Because of the low Zr solubility in anatectic melts, zircon is rarely crystallized from the anatectic melts unless the anatectic melts have experienced given degrees of evolution with fractional crystallization of Zr-poor minerals from the melts.

Figure 7. Plane-polarized light (PL) images of mineral inclusions in zircon and CL images of zircon in Dabie-Sulu UHP metamorphic rocks. (a) Peritectic zircon in gneiss (after Chen et al., 2013a); (b) peritectic zircon in quartzite (after Chen 2013b); (c) peritectic zircon as rim of relict Neoproterozoic magmatic zircon in leucosome (after Zeng et al., 2011); (d) and (e) peritectic zircon as rim of metamorphic zircon in leucosome (after Liu et al., 2010a); (f) pertectic zircon in leucosome (after Liu et al., 2012); (g) peritectic/anatectic zircon in pegmatite (after Liu et al., 2010a); (h) peritectic/anatectic zircon in pegmatite (after Liu et al., 2009d); (i) peritectic/anatectic zircon in Kfs-rich pegmatite (after Xu et al., 2013); (j) and (k) magmatic zircon in biotite-bearing granite (Liu et al., 2009c); (l) magmatic zircon in granite and syenite (Zhao et al., 2012, unpublished data ).

Figure 8. Chondrite-normalized REE patterns for inherited magmatic zircon and peritectic zircon in UHP metamorphic rocks from the Dabie-Sulu orogenic belt. (a) Inherited magmatic zircon in gneiss, quartzite, migmatite and pegmatite; (b) peritectic zircon in UHP rocks and leucosome; (c) peritectic/anatectic zircon in pegmatite; (d) magmatic zircon in synexhumation granite and syenite. The data for granite and syenite are from Zhao et al. (2012).

Figure 9. Chondrite-normalized REE patterns for peritectic zircons in the typical UHP terranes. (a)

North-East Greenland (after Gilotti et al., 2014); (b) North Qaidam (peritectic zircon in HP-granulite from Yu et al., 2014; that in Grt-free leucosome from Zhang et al., 2015); (c) Bohemia (Kotoková et al., 2016); (d) Sulu (Taohang from Li et al., 2014; Weihai from Li et al., 2016a).

Figure 10. Correlation of Hf/Y and Ta/Nb ratios for anatectic and peritectic zircons from pegmatite vein and gneiss at Taohang in the Sulu orogen (after Li et al., 2013). (a) Pegmatite vein PV1; (b) gneiss HG1. The data of Yankee Lode are from Pettke et al. (2005). Solid red pentagon, empty red pentagon and solid blue pentagon denote the early magmatic, late magmatic and hydrothermal zircon domains, respectively, from Yankee Lode. The zircon domains in the host gneisses record their growth during a transition from metamorphic dehydration to partial melting along a temperature-increasing path, whereas the zircon domains in the pegmatite veins record their growth during a transition from an anatectic melt to an aqueous solution along a temperature-decreasing path. Despite the difference in the direction of fluid/melt evolution, the all zircon domains in these gneisses and pegmatite veins record the same event of crustal anatexis in the UHP gneisses.

Figure 11. CL images for different types of metamorphosed zircon. Solid transformation (a) (after Chen et al., 2010) and (b) (after Xia et al., 2009); metasomatic alteration (c) (after Chen et al., 2010) and (d) (after Xia et al., 2009); dissolution reprecipitation (e) (after Xia et al., 2010) and (f) (after Chen et al., 2010). The s a e bar is 50 μ .

Figure 12. Chondrite-normalized REE patterns for different types of metamorphosed zircons from the Dabie-Sulu orogenic belt. Panels (a) and (b) after Chen et al. (2010), and (c) after Xia et al. (2010). Abbreviations SR, MR, DR and MG denote the solid-state recrystalization, metasomatic recrystallization, dissolution recrystallization and metamorphic growth, respectively. The grey zone denote the protolith zircon of magmatic origin.

Figure 13. The relationship between

206

Pb/238U ages and selected trace element contents/ratios and

Lu-Hf isotope ratios for zircons from UHP metamorphic rocks in the Dabie-Sulu orogenic belt. (a) 206

Pb/238U age vs. HREE, (b) 206Pb/238U age vs. Th/U, (c) 206Pb/238U age vs. (Lu/Gd)N, (d) 206Pb/238U

age vs. Hf, (e) 206Pb/238U age vs. 176Lu/177Hf, (f) 206Pb/238U age vs. 176Hf/177Hf. Data sources: C2010, Chen et al., 2010; X2010, Xia et al., 2010. Abbreviations SR, MR, DR and MG denote the solid-state recrystalization, metasomatic recrystallization, dissolution recrystallization and metamorphic growth, respectively.

Figure 14. Photograph showing zircons in thin sections. (a) phlogopite-bearing peridotite from the Sulu orogen (after Zhang et al., 2011); (b) CL images of zircons in peridotie from the Sulu orogen (after Li et al., 2016b); (c) core-rim structure from the Bohemian massif (after Liati and Gebauer, 2009); (d) photo showing inclusions of phlogopite and olivine in zircon from the Zhimafang peridotite in the Sulu orogen (after Zhang et al., 2011); (e) CL image of zircon in Alpine peridotite showing core-rim structure and inclusions of clinopyroxene, amphibole and U-oxide (after Hermann et al., 2006b); (f) BSE image of zircon in the Bohemian peridotite, showing quartz inclusion and melt inclusion composed of albite, K-feldspar, quartz and melt (after Liati and Gebauer, 2009).

Figure 15. Histrogram of U-Pb ages and Th vs. U for metasomatic zircon and relict zircon in M-type orogenic peridotites from the Dabie-Sulu orogenic belt. Data sources: Rumble et al. (2002), Zhang et al. (2005, 2011), Liu et al. (2006), Zhao et al. (2006, 2007b), Zheng et al. (2006b, 2014), Li et al. (2008, 2016), Yang et al. (2009).

Figure 16. Pressure-temperature-time (P-T-t) paths for UHP metamorphic rocks in the Sulu orogen (after Li et al., 2014, 2016a). (a) Two metagranites at Taohang, and (b) three migmatite at Weihai. The dashed part of P-T path in (b) is conjectural. Gray and red boxes show metamorphic zircon produced by subsolidus dehydration reaction and peritectic zircon produced by supersolidus dehydration melting. The range of these box show temperatures calculated from the Ti-in-zircon thermometry (Ferry and Watson, 2007). Anatectic ages were also shown in the diagram. Thin lines denote the calculated P-T relations from the garnet-phengite thermometer of Green and Hellman (1982), the garnet-amphibole thermometer of Ravna (2000), the garnet-biotite thermometer of Holdaway (2000), and the Zr-in-titanite thermometer of Hayden et al. (2008). Capital letters D, M, and L denote the diatexite, metatexite, and leucosome, respectively. Reaction curves: diamond/graphite is after Bundy (1980), coesite/quartz is after Bohlen and Boettcher (1982), jadeite + quartz/albite is after Holland (1980), amphibole out is after Hermann and Spandler (2008), biotite and plagioclase out are after Auzanneau et al. (2006). The curve for the wet solidus of granite is after Huang and Wyllie (1981). The curves for phengite-dominated partial melting are based on experimental petrology: A2006—Auzanneau et al. (2006) for the metagraywacke system; H2002—Hermann (2002) for the KCMASH system; VH1988—Vielzeuf and Holloway (1988) for the pelitic system. The P-T paths are consistent with hydration melting at 237–227 Ma due to the tectonic transition from HP to UHP eclogite-facies metamorphism during the final subduction, and phengite-dominated dehydration melting at 223–217 Ma during the exhumation of UHP metamorphic rocks, with subsequent growth at 218–214 Ma during amphibolite-facies

retrogression.

Figure 17. Zircon U-Pb ages and O isotopes for UHP migmatites in the Sulu orogen (after Li et al., 2016a). (a) Zircon U-Pb ages and O isotopes for UHP diatexite, metatexite, and leucosome at Weihai; (b) the distributions of metamorphic and anatectic ages in the Sulu orogen. The U-Pb age distributions of UHPM (ultrahigh-pressure metamorphism) and RM (retrograde metamorphism) are from Liu and Liou (2011). The anatectic ages were collected from the literature (Wallis et al., 2005; Zhao et al., 2007; Zeng et al., 2009; Zong et al., 2010; Liu et al., 2012; Chen et al., 2013a, 2013b; Li et al., 2013, 2014; Song et al., 2014; Wang et al., 2014). Abbreviations: Coe-Ec—coesite-eclogite facies; Qz-Ec—quartz eclogite facies; Gr—granulite facies; Am—amphibolite facies.

Figure 18. Cartoon showing the zircon behavior in continental subduction zones. Growth of metmamorphic and peritctic zircons are controlled by dehydration reactions at tempetaurers below and above the wet solidus of crustal rocks, respectively, corresponding to different properties of fluid action during continental collision. Metmamorphic zircons have grown during the whole continental collision in the Dabie-Sulu orogenic belt, showing with two age peaks at the prograde HP-UHP transition stage during subduction and the retrograde UHP-HP transition stage during exhumation, respectively. Peritectic and anatectic zircons have grown mainly during exhumation of the UHP metamorphic rocks, the growth of peritectic zircons rarely occurs in metsasedimentary rocks atop the subducted continental slab during subduction. While peritectic zircon is produced by peritectic reaction during anataxis, anatectic zircon crystallizes from the anatectic melt during siognificant crystallization of Zr-poor minerals. Zircon recrystallization can occur in the whole collisional orogeny with episodic activities of fluid action at different stages. While the soid-state recrystallization of protolith zircons occurs in the lack of fluid action, metasomatic and dissolution recrystallizations occur in the presence of fluid action and intensively during exhumation. Abbreviations: SR = solid-state recrystallization, MR = metsasomatic recrystallization, DR = dissolution recrystallization.

Figure 1

Figure 2

Figure 3

Figure 4

Figure 5

Figure 6

Figure 7

Figure 8

Figure 9

Figure 10

Figure 11

Figure 12

Figure 13

Figure 14

Figure 15

Figure 16

Figure 17

1 2 3 4 5 6 7

8 9 10 11

Figure 18

79

Table 1 A summary of mineragraphy, inclusions, trace elements, U-Pb ages, Lu-Hf and O isotopes in different types of zircons from UHP metamorphic rocks Type Metamorphosed zircons Newly grown zircons Subtype

Solid-state recrystallization

Metasomatic recrystallization

Dissolution recrystallization

Metamorphic zircon

Peritectic zircon

Agents

Absent of limited fluid

Aqueous solution or hydrous melt

Aqueous solution or hydrous melt or supercritical fluid

Aqueous solution

Hydrous melt

Mineragraphy

euhedral morphology; oscillatory zoning to weakly zoning

euhedral morphology; patched or planar zoning

anhedral morphology; spongy texture

euhedral to anhedral morphylogy; unzoned ,weakly or cloudy zoned texture

unzoning, weakly zoning

Inclusions

protolith magmatic minerals

protolih magmatic minerals, rare metasomatic minerals

protolith magmatic minerals, abundant metamorphic and metasomatic minerals

metamorphic minerals, fluid inclusions

peritectic minerals, melt inclusions

Th and U contents

similar to or slightly lower than protolith zircon

variably lower than protolith zircon

lower than protolith zircon

very low Th, similar or lower U relative to protolith magmatic zircon

lower Th but similar U contents to protolith magmatic zircon

Th/U ratios

slightly lowered but >0.1

lowered but >0.1

significantly lower than <0.1

generally lower than 0.1

U-Pb ages

discordant between protolith and metamorphic ages

discordant between protolith and metamorphic ages

concordant for metamorphism

concordant for anatexis

Lu-Hf isotopes

unchanged 176Lu/177Hf and 176 Hf/177Hf ratios

O isotopes

variable values between protolith and metamorphic zircons

Trace elements

Mechanism

lowered 176Lu/177Hf ratios but almost unchanged 176Hf/177Hf ratios variable values between protolith and metamorphic zircons

Variable from <0.1 to >1 (generally <0.1) concordant or nearly concordant for prograde metamorphism lowered 176Lu/177Hf ratios but almost unchanged 176 Hf/177Hf ratios similar to metamorphic zircons

almost unchanges in contents; steep HREE patterns; high (Lu/Gd)N and Lu/Hf ratios

lowered or elevated REE depending on the property of fluids; steep-shallow HREE patterns; almost unchanged Hf; high (Lu/Gd)N and Lu/Hf ratios

similar or elevated Hf; lowered or elevated REE, Y, Th, U, Nb, Ta and Ti; variable HREE patterns; variable (Lu/Gd)N and Lu/Hf ratios depending on the property of fluids

ion diffusion

fluid alteration

dissolution reprecipitation

80

lowered 176Lu/177Hf but elevated 176Hf/177Hf ratios variable values depending on the composition of host rock lower REE+Y, Th+U and Nb+Ta than protolith magmatic zircon; flat to steep HREE patterns with variable Eu anomalies depending on metamorphic conditions; high Hf and low Lu/Hf ratios dehydration reaction below the wet solidus

lowered 176Lu/177Hf ratios but elevated 176Hf/177Hf ratios variable values depending on the composition of host rock lower than protolith zircon, high Hf, lowered Lu/Hf ratios, flat to steep HREE patterns with variable Eu anomalies depending on anatectic conditions peritectic reaction above the wet solidus

Graphical abstract

81

Highlights  Protolith zircons may be modified by solid-state transformation, metasomatic alteration and dissolution reprecipitation.  New zircon growth occurs via either metamorphic or peritectic reactions depending on temperature.  Different types of zircon can be distinguished by their differences in geochemical compositions.  Property of protolith zircon and behaviors of Zr-bearing minerals dictate the new growth of zircon.  A zirconological study has the advantage to date geological events and trace geochemical sources.

82