WSRC-MS-2000-00187

 

 

Crustal Fluid Evolution and Changes in Deformation Conditions during
Regional Syn- to Post-Orogenic Exhumation: Southeastern
Piedmont, Southern Appalachians

M. A. Evans
University of Pittsburgh
Pittsburgh, PA 15260

M. J. Bartholomew
University of South Carolina
Columbia, SC 29208

 

This document was prepared in conjunction with work accomplished under Contract No. DE-AC09-96SR18500 with the U.S. Department of Energy.

DISCLAIMER

This report was prepared as an account of work sponsored by an agency of the United States Government. Neither the United States Government nor any agency thereof, nor any of their employees, makes any warranty, express or implied, or assumes any legal liability or responsibility for the accuracy, completeness, or usefulness of any information, apparatus, product or process disclosed, or represents that its use would not infringe privately owned rights. Reference herein to any specific commercial product, process or service by trade name, trademark, manufacturer, or otherwise does not necessarily constitute or imply its endorsement, recommendation, or favoring by the United States Government or any agency thereof. The views and opinions of authors expressed herein do not necessarily state or reflect those of the United States Government or any agency thereof.

This report has been reproduced directly from the best available copy.

Available for sale to the public, in paper, from:  U.S. Department of Commerce, National Technical Information Service, 5285 Port Royal Road, Springfield, VA 22161,  phone: (800) 553-6847,  fax: (703) 605-6900,  email:  orders@ntis.fedworld.gov   online ordering:  http://www.ntis.gov/support/ordering.htm

Available electronically at  http://www.osti.gov/bridge/

Available for a processing fee to U.S. Department of Energy and its contractors, in paper, from: U.S. Department of Energy, Office of Scientific and Technical Information, P.O. Box 62, Oak Ridge, TN 37831-0062,  phone: (865 ) 576-8401,  fax: (865) 576-5728,  email:  reports@adonis.osti.gov

 

Abstract

Fluid inclusion microthermometric data from veins in the southeastern Piedmont province record the changes in fluid composition and deformation conditions during regional exhumation and cooling related to Late Paleozoic syn- to post-orogenic processes and early Mesozoic rifting. In general, the composition of post-metamorphic fluids that were trapped late during the Alleghanian orogeny and during post-orogenic exhumation are remarkably consistent across the southeastern Piedmont, indicating regional fracture connectivity.

The first fluids were trapped in veins that formed during during the last phases of the Alleghanian. These syn-deformational fluids are CO2-saturated low salinity brines (salinities of 2.6 to 5.7 wt.% NaCl equivalent) with homogenization temperatures in the range of 200° to 365°C. They were trapped under lithostatic pressures between 240 and 280 MPa, indicating burial depths of 11.2 to 12.7 km. These depths are similar to emplacement depths of post-kinematic plutons, suggesting a period of rapid isobaric cooling.

Low-salinity H2O inclusions and rare CO2-rich inclusions are evidence for Early Mesozoic regional decompression as fracturing above the brittle-to-ductile transition allowed regional pore-fluid pressure to drop to hydrostatic levels. Convective circulation of meteoric water resulted in the dilution of 'in-situ' fluids, and ultimately to a system saturated with meteoric water. These fluids continued to be trapped in vein minerals through much of the Mesozoic as rift basins formed during the opening of the Atlantic Ocean.

Late Paleozoic through Mesozoic exhumation rates for the eastern Piedmont province average 0.063 km m.y.-1 and cooling rates average ~1.9°C m.y.-1. These low rates may be directly related to thinned crust and lithosphere resulting from delamination processes during the late Alleghanian orogeny.

Introduction

Metamorphic terranes of orogenic belts may undergo a period of exhumation (as defined by England and Molnar, 1990, p.1173) as a result of syn-orogenic crustal delamination (Sacks and Secor, 1990; Nelson, 1992). However, our understanding of post-metamorphic rock processes during this period of exhumation is limited by mineral instability and sluggishness of retrograde reactions (Craw, 1990). As with all other stages of orogenesis, fluids play a critical role during uplift. Studies of these fluids and their interactions with the rock can yield much information on processes that occur in exhumed terranes (Hollister and others, 1979; Frey and others, 1980; Boullier and others, 1989; Craw and Koons, 1989). The presence of fluids during unroofing and cooling of the metamorphic belt is evidenced by veins that cross-cut the metamorphic rocks. Fluid inclusion studies of fluids trapped in these post-metamorphic veins provide information on fluid evolution and changes in environmental conditions during uplift (Sauniac and Touret, 1983; Craw, 1987; Holm and others 1989; Craw, 1990; Jenkin and others, 1994).

This study examines late syn- to post-orogenic veins in the southeastern Piedmont province of the southern Appalachian orogen. During the late Paleozoic and early Mesozoic, the area experienced a period of regional exhumation with concurrent unroofing and cooling (Dallmeyer and others, 1986). During this time, rocks metamorphosed at amphibolite grade (~600°C) during the Alleghanian orogeny were cooled to <200°C, resulting in changes in environmental conditions. This was followed by continued exhumation during Mesozoic rifting and opening of the Atlantic Ocean. By Late Cretaceous time, the metamorphic rocks were at surface conditions (~20°C).

The goals of this paper are twofold. First, we will characterize the chemistry and evolution of fluids present in Piedmont rocks during late Alleghanian orogenesis and into early Mesozoic rifting. This will be achieved by integrating field observations, microthermometry of fluid inclusions in vein minerals, and vein petrography and mineral paragenesis. Second, we determine environmental conditions during fluid trapping and relate those to the exhumation and cooling history.

Regional Geology

Following the usage by Maher and others (1994), the eastern Piedmont of South Carolina and Georgia is divided into multiple terranes, each with a distinct lithotectonic assemblage and geologic history. The Carolina terrane (Fig. 1) is divided into two parts. The western portion of the Carolina terrane consists of Precambrian to Cambrian mafic to intermediate metavolcanic rocks and metasedimentary rocks penetrated by a series of mafic to ultramafic intrusive complexes (Secor, 1987; Dennis and Shervais, 1991, 1996). These rocks were metamophosed to upper greenschist to lower amphibolite facies between 500 and 430 ma (Butler, 1991). The eastern portion of the Carolina terrane consists of Cambrian felsic to intermediate metavolcanic rocks and metasedimentary rocks broadly encompassed by the Persimmon Fork, Richtex, and Asbill Pond Formations (Fig. 1) (Dennis and Shervais, 1991, 1996). These rocks were regionally metamorphosed to chlorite and biotite grade between 480 to 450 Ma (Butler, 1991). The Savannah River terrane (Mahar and others, 1991) consists of migmatitic sillimanite-bearing schists, paragneisses, and granite orthogneisses, and subordinate metamorphosed mafic to ultramafic rocks (Mahar and others, 1994). The Modoc fault zone separates the Carolina and Savannah River terranes. It consists of originally lower grade Carolina terrane rocks that underwent significant granitic sheet intrusion, prograde metamorphism and penetrative strain during the late Paleozoic Alleghanian orogeny (Secor and others, 1986).

In this area, the Alleghanian orogeny is divided into three deformation events that reflect the continuum of orogenesis (Secor and others, 1986). D2 is associated with amphibolite facies regional metamorphism and mid-crustal felsic plutonism that occurred between 315 to 295 Ma (Dallmeyer and others 1986; McSween and others, 1991). D2 reflects dextral strike-slip displacement along the Modoc fault zone (Fig. 1) (West and others, 1995). The D3 event occurred between 290 to 310 Ma (Dallmeyer and others, 1986; Pray and others, 1997) and folded the Modoc fault zone and high grade gneisses and schists of the Savannah River terrane and juxtaposed them against Carolina terrane rocks. This resulted in retrograde metamorphism from amphibolite to lower greenschist conditions (Mahar, 1987). Much of the study area was overprinted by ductile dextral shearing (Irmo deformation event, D4 of Secor and others, 1986) that terminated by 275 Ma (Pray and others, 1997).

Syn-orogenic granitoid and gabbroic plutons were emplaced between 327± 21 Ma and 286 Ma, and some as young as 266 Ma (McSween and others, 1991; Speer and others, 1994, D. Secor, personal communication, 2000). Many plutons are syn-kinematic and contain pervasive foliations and lineations. However, plutons discussed in this paper are considered to be post-kinematic because they lack ductile deformation structures, but may contain later brittle features including faults and joints. These plutons are considered to have been emplaced late in the orogeny (between 309± 2 Ma and 286± 3 Ma) (Speer and others, 1991). These late plutons were emplaced as water saturated (Pfluid = 0.5 Ptotal), relatively hot magmas (650° to 750°C ) at 290 to 330 MPa pressure (11.2 to 12.7 km depth) (Speer and others, 1994).

After the Alleghanian orogeny, during the early Mesozoic, this region was affected by deformation related to rifting that eventually resulted in the formation of the Atlantic Ocean. The initial phase of this deformation is the transition from the D4 dextral strike-slip event to north-south extension (D5) (Bartholomew and others, 1997). The second phase of Mesozoic rifting is characterized by northwest-southeast extension (D6). During this event, circulating hydrothermal fluids produced reactions in the wall-rock of D6 veins and earlier-formed reopened veins, altering calcium feldspar to K-feldspar (Mauldin and others, 1997). This 'pinking' event is dated at approximately 220 Ma (Kish, 1992). The last deformation event (D7) associated with rifting is the emplacement of syn-rifting diabase dikes throughout the Piedmont (Burt and others, 1978; Ragland and others, 1983; Bell, 1988; Goldsmith and others, 1988; Withjack and others, 1998). This event is dated at approximately 200 Ma (McHone and Butler, 1984; Sutter, 1988).

Veins

Data collection

The goal of sampling process was to sample only those veins that are post-metamorphic in origin. In order to accomplish this, a set of criteria was established to determine which veins were metamorphic. Veins that were parallel to foliation, boudinaged, folded, or exhibited obvious recrystallization textures in the field were assumed to be pre- to syn-metamorphic and were not sampled. In contrast, veins that cut across foliation (generally at a high angle) or are associated with normal faulting were assumed to be late syn- to post-metamorphic. These veins are generally planar to subplanar, 0.5 cm to 10’s of cm wide, and are vertical to steeply dipping. In general, veins in the post-kinematic granites and in the country rock are 1) the same orientation, 2) contain the same paragenetic sequence, and 3) contains the same fluid inclusion history. Therefore, the veins examined are late Alleghanian or later.

At each structural station (Fig. 1), vein orientations, along with other structural data (joints, faults, bedding, cleavage, fold axes, etc.) were measured. Samples of vein material from each vein set represented at the outcrop were collected for petrologic, and microthermometric analysis.

Orientation and Mineral Paragenesis

Five major vein sets are defined based on mineral paragenesis and vein orientation (Fig. 2, Table 1). V4 veins are found only in the Modoc fault zone and in the northeastern Carolina terrane. They are interpreted to have formed during D4 deformation (Secor and others, 1986; Pray and others, 1997) as the rock temperatures cooled below the brittle to ductile transition for quartz-feldspar aggregates (~300° to 400°C). The V4 set is divided into two sub-orthogonal subsets. V4a veins strike 325° ± 20° , contain the assemblage quartz ± pyrite and are orthogonal to the Modoc fault zone. V4b veins strike 058° ± 12° , contain the assemblage quartz ± muscovite, and are parallel to the Modoc fault zone (Fig. 2, Table 1). V5 veins strike 279° ± 21° and are parallel to east-west striking normal faults. They contain the mineral assemblage quartz ± muscovite ± epidote plus late calicte ± zeolite ± siderite (Fig. 2, Table 1). V5 is the most common vein set, occurring throughout the study area. V6 veins strike 000° ± 9° and contain the assemblage quartz ± chlorite plus late calicte ± zeolite ± prehnite. V7 veins (Fig. 2, Table 1) strike 014° ± 24° and are found only in the northeast part of the Carolina terrane. In general, the V7 set veins are subparallel to the strike of the Jurassic diabase dikes in the Carolinas (Ragland, 1991). V7 veins contain a distinctive calcite± prehnite± zeolite± quartz assemblage. V7 veins parallel V6 veins and in some cases are reopened V6 veins.

Fluid Inclusion Analysis

Methods

Fluid inclusion compositions and densities were determined by measuring the temperatures of phase transitions within fluid inclusions using a modified U.S. Geological Survey-type heating-freezing stage manufactured by FLUID Inc. The stage was calibrated at 0°C (ice bath), 374.1°C (critical point of water), and -56.6°C (CO2 triple point) (the latter two standards supplied by SYNFLINC, Inc.). Individual temperature measurements were found to be reproducible to ± 0.2°C at low temperatures (CO2 melting, ice, clathrate; CO2 homogenization), and to ± 2°C at high temperatures (aqueous liquid-vapor or aqueous carbonic homogenization).

Quartz and calcite hosted fluid inclusions were first warmed to their homogenization temperatures in order to reduce the risk of inclusion stretching due to ice formation (Lawler and Crawford, 1983; Meunier, 1989). The inclusions were then cooled below -100°C and subsequently warmed rapidly to determine the approximate temperature of ice melting (TmIce) and homogenization (ThTot) of constituent phases. Subsequent runs were then performed at slower rates to more accurately determine these temperatures. Depending on the composition of each inclusion, the following temperatures and phase ratios were determined: TmCO2, percent CO2 vapor relative to CO2 liquid at TmCO2, TmIce (water ice), Tmcl (clathrate), ThCO2(CO2 L-V), percent CO2 at 35°C, ThTot (H20 L-V), and ThTot (H2O-carbonic). Because of the difficulty of viewing phase changes in small inclusions, clathrate and ice melting measurements could not be performed on all inclusions for which ThTot was determined. In the CO2-bearing aqueous inclusions, salinity was derived from either 1) the dissociation temperature of the clathrate (Tmcl) in the presence of CO2-liquid and vapor (Chen, 1972; Bozzo and others, 1973; Diamond, 1992) or 2) from the melting temperature of ice in the presence of CO2 clathrate and CO2 vapor, using the experimental data of Bozzo and others (1973).

Fluid Inclusion Observations

Fluid Inclusion Populations

A fluid inclusion population comprises 'a group of inclusions that formed at the same time from the same parent fluid under identical physicochemical conditions and which have since undergone the same post-entrapment history' (Guscott and Burley, 1993). In this study, the populations are defined by inclusion fluid composition and inclusion homogenization temperature. Four populations of fluid inclusions are found in the veins studied.

H2O. These inclusions range in size from <2 m m to ~ 25 m m and contain a vapor bubble that occupies from 1 to 2 volume % of the inclusion. Homogenization temperatures (ThTot) range from 65° to 224°C (Figs. 3 and 4) and always occurs by homogenization to a liquid. Water ice melting temperatures (TmIce) range from –0.2° to 0.0°C, but are concentrated at 0.0°C and are interpreted to represent meteoric water (Fig. 3, Table 1) Several positive TmIce values (1.0 to 2.3°C) represent metastable 'superheated' ice (Roedder, 1984). Primary H2O inclusions occur in calcite as irregular clusters without a planar or other clear geometric distribution. Pseudosecondary H2O inclusions occur in calcite as planar arrays along discordant, annealed fractures that terminate within the crystal margins. These inclusions are compositionally the same as the primary inclusions, suggesting that they are related through a continuous temporal sequence. In quartz, the H2O inclusions always occur as planar clusters and appear to be secondary. In cross-cutting planar arrays, the array with the lower average ThTot is always the latest (Fig. 3a).

Low-Salinity H2O. These inclusions range in size from <2 m m to ~ 50 m m and contain a vapor bubble that occupies from 1 to 2 volume % of the inclusion. Homogenization temperatures (ThTot) range from 86° to 181°C (Figs. 3 and 4, Table 1) and always occurs by homogenization to a liquid. Water ice melting temperatures (TmIce) range from –3.2° to -0.3°C, representing salinities in the range of 0.5 to 5.3 wt. % NaCl equivalent. Low-salinity H2O inclusions occur only in quartz as isolated inclusions and irregular clusters interpreted as primary and as planar clusters that are interpreted as pseudosecondary or secondary (Fig. 3b). This inclusion population is interpreted as fluids representing formed as a result of mixing of meteoric water with more saline in-situ fluids.

Low XCO2. These inclusions range in size from <2 m m to over 100 m m and contain two or three phases at room temperature depending upon the ThCO2 of the CO2 vapor. When three phases are present, they include a carbonic vapor, a carbonic liquid, and an immiscible aqueous liquid (Figs. 3c and 3d). Upon homogenization of the carbonic vapor between 17.4° to 30.9°C, the two remaining phases are a carbonic bubble that occupies from 20 to 30 volume % of the inclusion and an aqueous liquid. More rarely, inclusions contain a vapor bubble that occupies 80 to 90 volume % of the inclusion. These homogenize to a vapor. Many inclusions were too small to observe ThCO2, so for these inclusions, only ThTot is reported (Figs. 3e and 3f). CO2 melting temperatures range from -56.8° to -56.4°C. Values lower than –56.6°C may be accounted for by small amounts of CH4. However, melting temperatures above –56.6°C cannot be accounted for by any known phase equilibria, although Burruss (1981) suggested that such high values may be the result of a mixture of CO2 and SO2. Total homogenization temperatures (ThTot) range from 198° to 365°C, and may be to a vapor or liquid (Figs. 3 and 4, Tables 1 and 2). Several inclusions decrepitated before reaching homogenization. These are used to provide a lower limit to the trapping pressure. Water ice melting temperatures (TmIce) range from -4.2° - 0.9°C and represent salinities in the range of 1.5 – 6.7 wt.% NaCl equivalent (Tables 1 and 2). In larger inclusions, CO2 clathrate melting may be observed between 7.8 - 9.2°C indicating salinities in the range of (1.7 - 4.3 wt. % NaCl equivalent). Only two eutectic temperature values could be determined, -34°C and -28°C, suggesting the fluid contains a mixture of NaCl-CaCl2 (Te = -52°C [Schafer and Lax, 1962]), NaCl-CaCl2-MgCl2 (Te = -57°C [Luzhnaya and Vereshtechetina, 1946]), or NaCl-MgCl2 (Te = -35°C [Schafer and Lax, 1962]). CO2 density was derived from the ThCO2 measured in the metastable absence of CO2 clathrate, assuming CO2 to be pure (Angus and others, 1973). Densities range from 0.37 to 0.97. These inclusions occur as both primary and pseudosecondary inclusions in quartz. The pseudosecondary inclusions have ThTot values similar to, or lower than the ThTot values of the primary inclusions, suggesting a coeval or sequential trapping sequence. In intersecting planar arrays, the younger array has lower ThTot that the older array (Fig. 3f)

CO2-rich. These inclusions range in size from < 5 m m to over 75 m m. They are two-phase at room temperature and include a carbonic vapor bubble that occupies 10 to 20 volume % of the inclusion and an immiscible carbonic liquid. The presence of up to 4 mole % CH4 (Table 3) is indicated by the homogenization of a very small vapor bubble at approximately –82.2°C. Homogenization temperatures range from 17.4° to 30.9°C (Table 3). Although no visible aqueous liquid is present, the presence of water (most likely along the inclusion walls) is indicated by CO2 clathrate melting temperatures of 7.7° to 8.2°C which represents salinities in the range of 3.9 to 4.4 wt. % NaCl equivalent (Table 3). Low-salinity H2O inclusions are found in the same irregular cluster as the CO2-rich inclusions and are assumed to be coeval (Fig. 3g). The CO2-rich inclusions only occur as primary and pseudosecondary inclusions in a single quartz sample from the Savannah River terrane.

Fluid Trapping Conditions

Isochores were constructed for the H2O and Low-salinity H2O inclusions using the MacFlincor software (Brown and Hagemann, 1994, 1995) and the equations of state for the H2O-CO2-NaCl system (Brown and Lamb, 1989). A geothermal gradient range of 20° to 30°C km-1 and hydrostatic (10 MPa km-1) pressure gradient was used to provide upper and lower bounds to the trapping pressure (Fig. 5).

For Low XCO2 inclusions, the salinity, CO2 density, and homogenization temperatures of the inclusions were used to estimate CO2 content and the minimum pressure of entrapment at ThTot using the MacFlincor software (Brown and Hagemann, 1994, 1995) and the procedure outlined in Brown and Lamb (1989). Homogenous fluids could have been trapped at higher temperatures along appropriate isochores for each fluid composition, therefore, the values determined represent a minimum trapping pressure and a minimum trapping temperature (ThTot). The true entrapment pressure-temperature conditions lie along isochores in the single-phase region for each fluid inclusion composition.

Isochores for CO2-rich inclusions were constructed using data from Swanenberg (1979). ThTot values from associated Low-salinity H2O inclusions were then used to constrain the trapping temperatures.

Fluid Compositions and Trapping Conditions

Modoc fault zone. Two distinct fluids are trapped in V4 veins in rocks of the northeastern Modoc fault zone. The earlier-trapped fluids are Low XCO2 (Table 2) with a CO2 content ranging from 6.5 to 30.1 mole %, a density ranging from 0.36 to 0.97 g cm-3, and salinities ranging from 3.1 to 4.4 weight % NaCl equivalent. ThTot values range from 201° to 363°C (Fig. 4a). The calculated minimum pressures of entrapment have a bimodal distribution (Fig. 5a) which is interpreted to be the result of pore-fluid pressure transients accompanying dilatant fracturing in the fault zone (Parry and Bruhn, 1990; Parry and others, 1991). The highest pressure values fall near the lithostatic thermobaric gradient, while the lowest pressures are near the hydrostatic thermobaric gradient. Similar bimodal distributions have been documented along the Dixie Valley fault in Nevada (Parry and Bruhn, 1990; Parry and others, 1991) and in syntectonic veins from the Irish Variscides (Meere, 1992). The pressure transients most likely occurred during periodic reactivation of the Modoc fault during D3 deformation. The scatter in values may be the result of errors in the ThTot measurements and pressure estimates, from post-entrapment changes in fluid inclusions, from entrapment of a non-homogenous fluid, or from real temperature and/or fluid pressure variations.

Maximum pressures are in the range of 240 to 280 MPa, indicating burial depths of 11.2 to 12.7 km at lithostatic conditions. The data indicate geothermal gradients of 35°C km-1 to as low as 20°C km-1. Minimum pressures are in the range of 98 to 145 MPa under hydrostatic conditions.

The second fluid population consists of Low-salinity H2O inclusions that represent an influx of meteoric fluids. Maximum pressures are assumed to be hydrostatic, although they could have periodically been higher. Using a thermobaric gradient of 10°C km-1, trapping of this population began at depths of nearly 10 km.

Northeastern Carolina Terrane. Veins in the Harbison granite, the Columbia granite, and the Asbill Pond Formation also contain two distinct fluid populations. The early Low XCO2 inclusions have ThTot values ranging from 202° to 338°C but with a bimodal salinity distribution, 0.9 to 2.8 wt. % NaCl equivalent and 3.2 to 4.0 wt. % NaCl equivalent (Fig. 4b). The higher salinity fluids are found only in the granites. Only one V4b vein from the Harbison granite contained Low XCO2 inclusions large enough to determine ThCO2. In this vein, the CO2 content is 0.11 to 0.14 mole %, the salinity 1.8 to 3.5 wt. % NaCl equivalent, and the CO2 density 0.90 to 0.92 g cm-3 (Table 2). Trapping conditions for the fluids in this vein are 255° to 275°C and 250 to 275 MPa (Fig. 5b). Assuming lithostatic conditions, this represents a burial depth of 9.6 to 10.6 km. There is no control on trapping pressure for the remaining Low XCO2 inclusions (Fig. 5b).

Vein mineral assemblages and mineral phase equilibria provide additional support for the pressure-temperature conditions determined by fluid inclusions. V5 veins in the Harbison granite contain the assemblage quartz ± epidote ± chlorite which is stable in the temperature interval of approximately 250° to 400°C (Fig. 5b).

Nearly pure H2O inclusions constitute the second fluid population (Fig. 4b, Table 1). These inclusions are found in all vein sets, but are the only type of fluid inclusions in the late calcite ± zeolite bearing V7 veins (Table 1). Pressure corrected trapping temperatures range from 70° to 225°C and pressures range from 20 to 95 MPa, corresponding to burial depths of up to 9.5 km (assuming a hydrostatic thermobaric gradient). The upper part of the temperature range is within the laumantite ± prehnite stability field (approximately 140° to 240°C) (Fig. 5b) and the presence of prehnite restricts pressures to less than 150 MPa. The presence of zeolite is related to regional heating and fluid flow associated with Mesozoic rifting and emplacement of Jurassic age dikes (Butler, 1977)

Southwest Carolina Terrane. In the southwestern portion of the Carolina terrane, both V5 and V6 quartz veins contain two distinct fluid inclusion populations. The earlier trapped fluids are Low XCO2 and have homogenization temperatures ranging from 200° to 318°C and clathrate melting temperatures in the range of 7.8° to 8.2°C, corresponding to salinities of 3.6 to 4.3 Weight % NaCl equivalent (Fig. 4c, Table 2). For most inclusions, a separate carbonic liquid bubble could not be discerned. Either the inclusions were too small to make out the small bubble, or the carbonic vapor bubble was nearly the same size as the carbonic liquid ‘bubble’. However, the presence of clathrate and the rare carbonic vapor homogenization attest to the presence of CO2 in the inclusions. The two inclusions used for estimating environmental conditions (Table 2) give minimum trapping temperature of approximately 250° C and pressures of 283 MPa corresponding to a depth of 10.9 km.

Low-salinity H2O inclusions are uncommon, occurring as pseudosecondary inclusions in the Persimmon Fork Formation, while veins in the Cuffytown Creek granite and the Persimmon Fork Formation contain common secondary H2O inclusions. Pressure corrected trapping temperatures range from 80° to 320°C and pressures range from 25 to 148 MPa, corresponding to maximum burial depths of 8.8 to 14.8 km (assuming a hydrostatic thermobaric gradient).

Savannah River Terrane. In this area, V4a and V4b veins contain Low XCO2 inclusions with salinities ranging from 2.2 to 6.2 weight % NaCl equivalent, and the Low-salinity H2O inclusions range from 0.6 to 5.7 weight % NaCl equivalent (Fig. 4d). This overlap suggests that there is a continuum in mixing of 'in-situ' fluids with meteoric fluids, but unlike the other areas there is not apparently a wholesale meteoric influx. In inclusions large enough to see definitive phase transitions, the CO2 content ranges from 0.10 to 0.16 mole %, the salinity 2.2 to 3.3 weight % NaCl equivalent, and the CO2 density 0.78 to 0.88 g cm-3 (Table 2). Trapping conditions for the fluids in this vein are 237° to 308°C and 210 to 249 MPa (Fig. 5d). Assuming lithostatic conditions, this represents a burial depth of 8.1 to 9.6 km. Again, the vein mineral assemblage quartz± epidote± chlorite indicates a temperature interval of approximately 250° to 400°C (Fig. 5b).

Rare CO2-rich inclusions contain up to 4% CH4 and have densities ranging from 0.53 to 0.76 (Table 3). Isochore calculated for CO2-rich and coeval Low-salinity H2O inclusions give pressure-corrected trapping conditions of 170° to 275°C and 48 to 108 MPa indicating maximum burial depths of up to 10.8 km under hydrostatic conditions.

For the Low-salinity H2O and H2O inclusions, pressure corrected trapping temperatures range from 100° to 250°C and pressures range from 25 to 115 MPa, corresponding to maximum burial depths of 7.5 to 11.5 km (assuming a hydrostatic thermobaric gradient) (Fig. 5d). The upper portion of the temperature range is supported by the presence of laumantite in V4a veins. The presence of late stilbite reflects temperatures lower than 140°C.

Discussion

Areal and Temporal Changes in Fluid Composition

The composition of post-metamorphic fluids trapped late during the Alleghanian orogeny and during post-orogenic uplift are remarkably consistent across the southeastern Piedmont. The earliest-trapped fluids are represented by Low-XCO2 inclusions and these have the same general trapping temperatures and composition in all tectonic and structural settings examined, including post-deformation plutons and host country rock. This suggests that a pervasive interconnected brittle fracture network developed after the rocks cooled to temperatures below the brittle-to-ductile transition for quartzo-feldspathic rocks. Similar regionally homogeneous fluid compositions are found in the Lachlan fold belt of southeastern Australia (Gray and others, 1991), the Ominica belt in western Canada (Nesbitt and Muehlenbachs, 1995), and the central Appalachian Valley and Ridge province (Evans and Battles, 1999).

The chemical composition of the fluids is similar to that found in other metamorphic settings (e.g. Hollister and Burruss, 1976; Sisson and Hollister, 1990). The origin of the fluids is most likely re-equilibration reactions and retrograde breakdown of metamorphic minerals. There may also be a component of the fluids associated with the intrusion of the late syn-post kinematic plutons. The limited scope of this study precludes a more detailed investigation into the origin of the fluids.

The systematic nature of the fracture/vein networks attest to their origin in relation to a tectonic stress. Although the V4 vein sets are limited to rocks within and near the Modoc fault zone, the V5 and V6 sets are regionally pervasive. These fractures would have aided in the regional circulation of fluid by advective cycling of fluid resulting from episodic dilatancy ‘pumping’ in the fracture networks associated with localized faulting (Sibson and others, 1975; Sibson 1989; Gray and others, 1991). The fluids were trapped in V4 veins at depth of 11.2 to 12.6 km in the Modoc fault zone, between 9.6 to 10.9 km in V5 and V6 veins in the Carolina terrane, and between 8.1 to 9.6 km in V5 and V6 veins in the Savannah River terrane. These depths are close to the 11.0 to 14.3 km depth expected for the brittle-to-ductile transition for quartzo-feldspathic rocks (300° to 400°C) using an average geothermal gradients of 30°C/km and a 20°C surface temperature.

With regional uplift, denudation, and cooling, connecting fracture networks reached the paleosurface and allowed meteoric derived surface fluids to penetrate deep into the section where thermal convective circulation aided in distribution and mixing (Fig. 6). Convective circulation can only involve surface fluids with pressures near hydrostatic, since the fluid cannot flow up the pressure gradient toward lithostatically pressured intervals.

The regional pore-fluid pressure drop is evidenced by the presence of Low-salinity H2O inclusions and CO2-rich inclusions. These fluids formed as a result of unmixing of the original CO2-H2O-NaCl fluid present in the rocks upon decompression into a CO2-rich phase and a H2O-NaCl component (Conevey, 1981; Bowers, 1991; Boullier and Robert, 1992; Perry, 1998). Additionally, the fluid pressure reduction would have resulted in the precipitation of quartz and the phase separation of CO2 (Sibson and others, 1988). As large quantities of meteoric water circulated through this portion of the orogenic belt, any metamorphic fluid that was present was diluted as has been observed in other metamorphic belts (Nesbitt and Muehlenbachs, 1989, 1991, 1995; Botrell and others, 1990; Jenkin and others, 1992, 1994). Similar records of deep meteoric fluid penetration have been reported from the Canadian Cordillera (Nesbitt and others, 1986; Nesbitt and Muehlenbachs, 1991), the Pyrenees (Wickham and Taylor, 1987), and the Alpine Schist metamorphic belt (Jenkin and others, 1994).

Finally, the H2O inclusions indicate the regional influx and circulation of meteoric waters nearly to the brittle-to-ductile transition zone. Veins in the Savannah River Terrane, however, do not contain pure H2O inclusions. The lack of pure H2O inclusions in this area may be attributed to minimal meteoric influx resulting from 1) the presence of foliated rocks in the Modoc fault zone that were structurally above the Savannah River terrane (Fig. 6) thereby creating a permeability barrier to surface derived fluids; or 2) the low inherent permeability of the migmatitic rocks of the Savannah River terrane restricting fluid circulation.

Exhumation History

The eastern Carolina terrane and Modoc fault zone have been studied extensively, and a large amount of geologic and radiometric age data is available (Secor and others, 1986, Dallmeyer and others, 1986; Pray and others, 1997). Therefore, the following discussion of the regional exhumation history will center on this area, although a similar history may be applied to the remainder of the study area.

Based on aluminum-in-hornblende barometry (Vyhnal and McSween, 1990), emplacement pressures for local late-synorogenic plutons range from 290± 50 MPa (11.2± 1.9 km) for the Harbison granite, 315± 15 MPa (12.1± 0.6 km)for the Winnsboro granite (Speer and others, 1994), and 320± 50 MPa (12.3± 1.9 km) for the Graniteville pluton. These values, along with the average emplacement temperature of 650° to 750°C, set a starting point for examining the late Paleozoic to early Mesozoic cooling and exhumation history (Figs.7 and 8). The rather shallow emplacement depth for these plutons may be attributed to their emplacement as 'collisional' granites during Alleghanian crustal thinning resulting from rapid delamination (Sacks and Secor, 1990; Nelson, 1992) or crustal arching related to dextral transpression (Gates and others, 1988).

The reminder of the cooling curve is constructed from radiogenic age data, fission track data, and stratigraphic data. Hornblendes from the northeastern Modoc fault zone yield 40Ar/39Ar age dates of 292 and 298 Ma (Dallmeyer and others, 1986). Muscovite from the Modoc fault zone and the Savannah River terrane give 40Ar/39Ar age dates ranging from 272 to 278 Ma with an average of 275 Ma (Maher and others, 1994; Pray and others, 1997). The closer temperature of argon retention in amphiboles and muscovite is dependent on rate of cooling and pressure (Baldwin and others, 1993). However, for the relatively slow cooling rates of less than 10°C/My encountered in this study, and the relatively low pressures of ~300 MPa, the closure temperatures or these minerals is close to that determined by the initial diffusion equations. Therefore, we will use 500± 25°C for the closure temperature for argon retention in hornblende (Harrison, 1981), and 375± 25°C for the closure temperature for argon retention in muscovite (Brown and Dallmeyer, 1996, based on the preliminary data of Robbins [1972] and the diffusion equations of Dodson [1973]) (Fig. 7). Apatite fission track ages of 140± 10 Ma have been reported by Zimmerman (1979) for the eastern Carolina terrane. These indicate cooling through ~125°C (Naeser and Cebula, 1978). Finally, by at least Santonian time (~87.5 Ma) the rocks of the inner Piedmont are exposed at the Paleosurface (Nystrom and others, 1992) (Fig. 7).

The fluid inclusion data can be used to construct a pressure-temperature curve that corresponds to the late Paleozoic to Mesozoic portion of the cooling curve (Fig. 8). The first fluids trapped in the rocks record temperatures up to 365°C and depths on the order of 11.0 to 12.7 km. These depths are nearly the same as the granite emplacement depths and suggest a period of rapid, near isobaric cooling (Fig. 8). Such rapid cooling could result from the crustal delamination exposing warm midcrustal rocks to near surface temperatures and the influx of cool meteoric waters into the recently uplifted rock (Fig. 6). As pointed out by Nesbitt and Meuhlenbachs (1989), the convection of meteoric water in the crust should have a significant impact of heat flow and cooling in the crust. For example, regional geothermal gradients must have rapidly been modified, from 60° to 70°C km-1 during plutonism to 20° to 40°C km-1 by the time Low XCO2 inclusions were trapped (Fig. 8).

The evidence for fluid pressure cycling in the Modoc zone suggests that these rocks were exhumed to above the brittle-to-ductile transition zone prior to, or during D4 deformation (Fig. 7). A regional drop in pore-fluid pressure to hydrostatic levels most likely occurred by late in D6 deformation as regional temperatures dropped to the 200° to 250°C range (Figs. 7 and 8).This event is recorded by the regional influx of Low salinity H2O and H2O fluids.

Fluid trapping continued after the Alleghanian orogeny and into the early Mesozoic. By this time, the upper crust was saturated with meteoric fluids. Veins containing these meteoric derived fluids formed during the later phase of Mesozoic rifting when the rift-to-shift transition set up a new stress field in the crust (Withjack and others, 1998).

By extrapolating fluid temperatures down the exhumation curve, fluid inclusions with trapping temperatures of 80° to 125°C may indicate trapping as late as 120 to 140 Ma (Fig. 7). These late trapped fluids might be related to the formation of microfractures resulting from crustal stresses associated with inception of the present-day compressive stress field (Zoback and Zoback, 1989).

In summary, the above data yield a two-stage cooling curve for the eastern portion of the study area. Between the time of emplacement of the Harbison pluton (the earliest of the local plutons), and the muscovite argon retention age, the regional cooling rate is ~ 9.5°C m.y.-1. Then, the regional cooling rate falls to ~1.9°C m.y.-1 throughout the Mesozoic (Fig. 7). Because there is a rapid decrease in the geothermal gradient from the time of pluton emplacement to the muscovite cooling age, we cannot determine an exhumation rate for this time interval. But, we can say that exhumation was minimal, as pore-fluid pressure was relatively constant. However, we can calculate an approximate exhumation rate for the late Paleozoic through Mesozoic. Here, we assume 1) an initial starting depth of 11.9 km at 275 Ma, 2) surface conditions at 87 Ma, and 3) a relatively constant geothermal gradient. The value determined is 0.063 km m.y.-1. This relatively slow exhumation rate is nearly 2 orders of magnitude less than those recorded in the Alps and Himalayas (e.g. Zeitler, 1985; and Copeland and others, 1987; Vance and Onions, 1992; Marshall and others, 1997). This may be due to the thickened crust in the active orogens being isostatically compensated for by rapid erosion, whereas in the Piedmont, the thinned crust and lithosphere resulting from the delamination process is already near isostatic equilibrium. Therefore, regional exhumation is slow as crust and lithosphere thickness only increases by cooling during the Mesozoic.

Conclusions

Fluid inclusion microthermometric data from veins in the southeastern Piedmont province record the changes in fluid composition and deformation conditions during regional exhumation and cooling related to Late Paleozoic syn- to post-orogenic processes and Triassic rifting. In general, the composition of post-metamorphic fluids trapped late during the Alleghanian orogeny and during post-orogenic uplift are remarkably consistent across the southeastern Piedmont, indicating regional fracture connectivity.

Syn-deformational fluids were trapped in veins that formed during the last phases of the Alleghanian orogeny. These fluids were CO2-saturated and had a temperature range of 200° to 365°C and salinities of 2.6 to 5.7 wt.% NaCl equivalent. They were trapped under lithostatic pressures between 240 and 280 MPa indicating burial depths of 11.2 to 12.7 km. These depths are similar to emplacement depths of post-kinematic plutons, suggesting a period of rapid isobaric cooling.

Low-salinity H2O inclusions and rare CO2-rich inclusions are evidence for regional decompression as fracturing above the brittle-to-ductile transition allowed the regional pore-fluid pressure to drop to hydrostatic levels. Convective circulation of meteoric water resulted in the dilution of 'in-situ' fluids, and ultimately to a pure-H2O saturated system. These fluids continued to be trapped in vein minerals through the early Cretaceous.

Late Paleozoic through Mesozoic exhumation rates for the eastern Piedmont province average 0.063 km m.y.-1 and cooling rates average ~1.9°C m.y.-1. These low rates may be directly related to thinned crust and lithosphere resulting from delamination processes during the late Alleghanian orogeny.

Acknowledgments

This work was partially supported through South Carolina University Research and Educational Task 269, Westinghouse Savannah River Company. We thank D. Secor, Jr. for providing many helpful suggestions for improvement of the manuscript.

References

  1. Baldwin, S. L., Lister, G. S., Hill, E. J., Foster, D. A., and McDougall, I., 1993, Thermochronologic constraints on the tectonic evolution of active metamorphic core complexes, D'Entrecasteaux Islands, Papua, New Guinea: Tectonics, v. 12, p. 611-628.
  2. Bartholomew, M. J., Heath, R. D., Brodie, B. M., and Lewis, S. E., 1997, Post-Alleghanian deformation of Alleghanian granites (Appalachian Piedmont) and the Atlantic Coastal Plain: Geological Society of America Abstracts with Programs, v. 29, p. 4.
  3. Bartholomew, M. J., Whitaker, A. E., and Barker, C. A., 1998, Preliminary Mesozoic-Cenozoic brittle-deformation history of Eocambrian rocks (Ridgeway gold mine, South Carolina), Carolina terrane: South Carolina Geology, v. 40, p. 19-27.
  4. Bell, H., III, 1988, Map showing Mesozoic diabase dikes in South Carolina described since 1960: United States Geological Survey Miscellaneous Field Studies Map MF-2030, scale, 1:250,000.
  5. Bird, D. K., and Helgeson, H. C., 1981, Chemical interaction f aqueous solutions with epidote-feldspar mineral assemblages in geologic systems II, Equilibrium constraints in metamorphic/geothermal processes: American Journal of Science, v. 281, p. 576-614.
  6. Bottrell, S. H., Greenwood, P. B., Yardley, B. W. D., Shepherd, T. J., and Spiro, B., 1990, Metamorphic and post-metamorphic fluid flow in the low-grade rocks of the Harlech Dome, north Wales: Journal of Metamorphic Geology, v. 8, p. 131-143.
  7. Boullier, A-M., and Pecher, A., 1981, Fluid inclusion in the MCT pile in Central Nepal: Conference on Deformation Mechanisms, Rheology, and Tectonics, University of Leeds, Leeds, March 1989, Abstract.
  8. Boullier, A-M., and Robert, F., 1992, Paleoseismic events recorded in Archean gold-quartz vein networks, Val d'Or, Abitibi, Quebec, Canada: Journal of Structural Geology, v. 14, p. 161-179.
  9. Bowers, T. S., 1991, The deposition of gold and other metals: Pressure-induced immiscibility and associated stable isotope signatures: Geochimica et Cosmochimica Acta, v. 55, p. 2417-2434.
  10. Bozzo, A. T., Chen, H-S, Kass, J. R., and Barduhn, A. J., 1975, The properties of the hydrates of chlorine and carbon dioxide: Desalination, v. 16, p. 303-320.
  11. Bramlett, K. W., Secor, D. T., Jr., and Prowell, D. C., 1982, The Belair fault: A Cenozoic reactivation structure in the eastern Piedmont: Geological Society of America Bulletin, v. 93, p. 1109-1117.
  12. Brown, M., and Dallmeyer, R. D., 1996, Rapid Variscan exhumation and the role of magma in core complex formation: southern Brittany metamorphic belt, France: Journal of Metamorphic Geology, v. 14, p. 361-379.
  13. Brown, P. E., and Hagemann, S. G., 1984, MacFlincor: A computer program for fluid inclusion data reduction and manipulation, in De Vivo, B., and Frezzoti, M. L., eds., IMA Short Course on Fluid Inclusions, Virginia Polytechnic and State University, Blacksburg, Virginia, p. 231-250.
  14. Brown, P. E., and Hagemann, S. G., 1985, MacFlincor and its application to fluids in Archean lode-gold deposits: Geochimica et Cosmochimica Acta, v. 59, p.3943-3952.
  15. Brown, P.E., and W.M. Lamb, 1989, P-V-T properties of the fluids in the system H2O - CO2 - NaCl: New graphical presentations and implications for fluid inclusion studies: Geochimica et Cosmochimica Acta, v. 53, p. 1209-1221.
  16. Burruss, R. C., 1981, Analysis of phase equilibria in C-O-H-S fluid inclusions In Hollister L. S. and Crawford, M. L., eds., Mineralogical Association of Canada Short Course in Fluid Inclusions, Mineralogical Association of Canada, Volume 6, p. 75-100.
  17. Burt, E. R., Carpenter, P. A., III, McDanial, R. D., and Wilson, W. F., 1978, Diabase dikes of the eastern Piedmont of North Carolina: North Carolina Department of Natural Resourced and Community Development, Geological Survey Section, Information Circular 23, 12 p.
  18. Butler, J. R. 1977, Comments on zeolite formation in the Carolina Piedmont: Geological Society of America Abstracts with Programs, v. 9, p. 124-125.
  19. Butler, J. R., 1991, Metamorphism in Horton, J. W., Jr., and Zullo, V. A., eds., The Geology of the Carolinas, The University of Tennessee Press, Knoxville, p. 127-141.
  20. Chen, H. S., 1972, The thermodynamics and composition of carbon dioxide hydrate: MS Thesis, Syracuse University, 78p.
  21. Cho, M., Maruyama, S., and Liou, J. G., 1987, An experimental investigation of heulandite-laumontite equilibrium at 1000 to 2000 bar P fluid: Contributions to Mineralogy and Petrology, v. 97, p. 43-50.
  22. Conevey, R. M., Jr., 1981, Gold quartz veins and auriferous granite at the Oriental mine, Alleghany District, California: Economic Geology, v. 76, p. 2176-2199.
  23. Copeland, P. Harrison, T. M., Kidd, W. S. F., Xu, R., and Zhang, Y., 1987, Rapid early Miocene acceleration of uplift in the Gangdese Belt, Xizang (southern Tibet), and its bearing on accommodation mechanisms of the India-Asia collision: Earth and Planetary Science Letters, v. 86, p. 240-252.
  24. Craw, D., 1990, Fluid evolution during uplift of the Annapurna Himal, central Nepal: Lithos, v. 24, p. 137-150.
  25. Craw, D. and Koons, P. O., 1989, Tectonically induced hydrothermal activity and gold mineralization adjacent to major fault zones: in Keays, R., Ramsay, R., and Groves, D., eds., The Geology of old Deposits. Economic Geology Monograph 6, p. 463-470.
  26. Dallmeyer, R. D., Wright, J. E., Secor, D. T., Jr., and Snoke, A. W., 1986, Character of the Alleghanian orogeny in the southern Appalachians: Part II. Geochronological constraints on the teconothermal evolution of the easten Piedmont in South Carolina: Geological Society of America Bulletin, v. 97, p. 1329-1344.
  27. Dennis, A. J., Sacks, P. E., and Maher, H. D., 1987, Nature of the late Alleghanian strike-slip eformation in the eastern South Carolina Piedmont: The Irmo shear zone: in Secor, D. T., Jr., ed., Anatomy of the Alleghanian orogeny as seen from the Piedmont of South Carolina and Georgia, South Carolina Geological Survey, Columbia, South Carolina, p. 49-66.
  28. Dennis, A. J., and Shervais, J. w., 1991, Evidence for arc rifting along the Carolina terrane boundary in northwestern South Carolina: Geology, v. 19, p. 226-229.
  29. Dennis, A. J., and Shervais, J. W., 1996, The Carolina terrane in northwestern South Carolina: Insights into the development of an island arc in Nance, R. D., and Thompson, M. D., eds. Avalonian and Related Peri-Gondwanan Terranes of the Circum-North Atlantic, Geological Society of America Special Paper 304, p. 237-256.
  30. Diamond, L. W., 1992, Stability of CO2 clathrate hydrate + CO2 liquid + CO2-vapor + aqueous KCl-NaCl solutions: Experimental determination and application to salinity estimates of fluid inclusions: Geochimica et Cosmochimica Acta, v. 56, p. 273-280.
  31. Dodson, M. H., 1973, Closure temperature and cooling geochronological and petrological systems: Contributions to Mineralogy and Petrology, v. 40, p. 259-274.
  32. Englund, P., and Molnar, P. 1990, Surface uplift, uplift of rocks, and exhumation of rocks: Geology, v. 18, p. 1173-1177.
  33. Evans, M. A. and Battles, D. A., 1999, Fluid inclusion and stable isotope analysis of veins from the central Appalachian Valley and Ridge province: Implications or regional syn-orogenic hydrologic structure and fluid migration: Geological Society of America Bulletin, v. 111, p. 1841-1860.
  34. Frey, M., Bucher, K., Fank, E., and Mullis, J., 1980, Alpine metamorphism along the geotraverse Basel-Chiasso: a review: Eclogea Geologica Helvetica Acta, v. 73, p. 527-546.
  35. Fullagar, P. D., 1981, Summary of Rb-Sr whole-rock ages for South Carolina, Geology, v. 25, p. 29-32.
  36. Fullagar, P. D., and Butler, J. R., 1979, 325-265 m.y.-old granitic plutons in the Piedmont of the southeastern Appalachians: American Journal of Science, v. 279, p. 161-185.
  37. Fullagar, P. D., and Kish, S. A., 1981, Mineral age traverses across the Piedmont of South Carolina and North Carolina, in Horton, J. W., Jr., Butler, J. R., and Milton, D. J., eds., Geological Investigations of the Kings Mountain belt and adjacent areas in the Carolinas: Carolina Geological Society Field Trip Guidebook 1981, Columbia, South Carolina Geologica Survey, p. 155-165.
  38. Gates, A. E., Speer, J. A., and Pratt, T. L., 1988, The Alleghanian southern Appalachian Piedmont: A transpressional model: Tectonics, v. 7, p. 1307-1324.
  39. Goldsmith, R., Milton, D. J., and Horton, J. W., Jr., 1988, Geologic Map of the Charlotte 1° x2° quadrangle, North Carolina and South Carolina: United States Geological Survey Miscellaneous Investigations Series Map I-1251-E, scale 1:250,000.
  40. Gray, D. R., Gregory, R. T., and Durney, D. W., 1991, Rock-buffered fluid-rock interaction in deformed quartz-rich turbidite sequences, eastern Australia: Journal of Geophysical Research, v. 96, p. 19681-19704.
  41. Guscott, S. C., and Burley, S. D., 1993, A systematic approach to reconstructing paleofluid evolution from fluid inclusions in authigenic quartz overgrowths in Parnell, J. ed., Proceedings of Geofluids 93, p. 323-328.
  42. Haas, J. L., Jr., 1976, Physical properties of the coexisting phases and thermochemical properties of the H2O component in boiling NaCl solutions: U. S. Geological Survey Bulletin 1421-A, 73 p.
  43. Harrison, T. M., 1981, Diffusion of 40Ar in hornblende: Contributions to Mineralogy and Petrology, 78, p. 324-331.
  44. Heath, R. D., and Bartholomew, M. J., 1997, Mesozoic phases of post-Alleghanian deformation in the Appalachian Piedmont: Geological Society of America Abstracts with Programs, v. 29, p. 23.
  45. Hollister , L. S., and Burruss, R. C., 1976, Phase equilibria in fluid inclusions from the Khtada Lake metamorphic complex: Geochimica et Cosmochimica Acta, v. 40, p. 163-175.
  46. Hollister, L. S., Burruss, R. C., Hendry, D. L., and Hendel, E. M., 1979, Physical conditions during uplift of metamorphic terrane as recorded by fluid inclusions: Bulletin Mineralogique, v. 102, p. 555-561.
  47. Holm, D., Norris, R. J., and Craw, D., 1989, Brittle and ductile deformation in a zone of rapid uplift, Central Southern Alps, New Zealand, Tectonics, v. 8, p. 153-168.
  48. Jenkin, G. R. T., Craw, D., and Fallick, A. E., 1994, Stable isotopic and fluid inclusion evidence for meteoric fluid penetration into an active mountain belt; Alpine Schist, New Zealand: Journal of Metamorphic Geology, v. 12, p. 429-444.
  49. Jenkin, G. R. T., Fallick, A. E., and Leake, B. E., 1992, A stable isotope study of retrograde alteraion in S. W. Connemara, Ireland: Contributions to Mineralogy and Petrology, v. 110, 269-288.
  50. Kish, S. A., 1992, An initial geochemical and isotopic study of granite from core C-10, Savannah River Site, S. C. in Fallow, W., and Price, V., eds., Geological Investigations of the Savannah River Area, South Carolina and Georgia: Carolina Geological Society Fieldtrip Guidebook, p. B-IV, 1-4.
  51. Lawler, J. P., and Crawford, M. L., 1983, Stretching of fluid inclusions resulting from a low temperature microthermometric technique: Economic Geology, v. 78, p. 527-529.
  52. Liou, J. G., 1971, Stilbite-laumontite equilibrium: Contributions to Mineralogy and Petrology, v. 31, p. 171-177.
  53. Liou, J. G., 1987, Very low-grade metamorphism of volcanic and volcanoclastic rocks: Mineral assemblages and mineral facies in Frey, M., ed., Low Temperature Metamorphism, Blackie and Son, Glasgow, p. 59-113.
  54. Liou, J. G,, Seki, Y., Guillemette, R., and Sakai, H., 1985, Compositions and paragenesis of secondary minerals in the Onikobe geothermal system, Japan, Chemical Geology, v. 49, p. 1-20.
  55. Luzhnaya, N. P., and Vereshtchetina, I. P., 1946, Sodium, calcium, and magnesium chlorides in aqueous solutions at -57° to +25° C (polythermic solubility): Zh. Prikl. Khimii, v. 19, p. 723-733.
  56. Maher, H. D., 1987a, D3 folding in the eastern Piedmont associated with Alleghanian thrusting: in Secor, D. T., Jr., ed., Anatomy of the Alleghanian orogeny as seen from the Piedmont of South Carolina and Georgia, South Carolina Geological Survey, Columbia, South Carolina, p. 35-48.
  57. Maher, H. D., 1987b, Kinematic history of mylonitic rocks from the Augusta fault zone, South Carolina and Georgia: American Journal of Science, v. 287, p. 795-816.
  58. Maher, H. D., Dallmeyer, R. D., Secor, D. T., Jr., and Sacks, P. E., 1994, 40Ar/39Ar constraints on chronology of Augusta fault zone movement and late Alleghanian extension, southern Appalachian Piedmont, South Carolina and Georgia: American Journal of Science, v. 294, p. 428-448.
  59. Mahar, H. D., Sacks, P. E., and Secor, D. T. Jr., 1991, The eastern Piedmont of South Carolina, in Horton, R. W., and Zullo, V., eds., The Geology of the Carolinas, Carolina Geological Society 50th Anniversary Volume, University of Tennessee Press, Knoxville, p. 93-108.
  60. Marshall , D., Kirschner, D., and Bussy, F., 1997, A Variscan pressure-temperature-time path for the N-E Mont Blanc massif: Contributions to Mineralogy and Petrology, v. 126, p. 416-428.
  61. Mauldin, J., Shervais, J., and Dennis, A. J., 1997, Basement core from the Savannah River Site, South Carolina: Geological Society of America Abstracts with Programs, v. 29, p. 57.
  62. McHone, J. G., and Butler, J. R., 1984, Mesozoic igneous provinces of New England and the opening of the North Atlantic Ocean: Geological Society of America Bulletin, v. 85, p. 757-765.
  63. McSween, H. Y., Jr., Speer, J. A., and Fullagar, P. D., 1991, Plutonic Rocks in Horton, J. W., Jr., and Zullo, V. A., eds., The Geology of the Carolinas, The University of Tennessee Press, Knoxville, p. 109-126.
  64. Meere, P. A., 1995, High density fluids in a quartz vein from the Irish Variscides: Journal of Structural Geology, v. 17, p. 435-446.
  65. Meunier, J. D., 1989, Assessment of Low-temperature fluid inclusions in calcite using microthermometry: Economic Geology, v. 84, p. 167-170.
  66. Naeser, C. W., and Cebula, G. T., 1978, Fission track dating of apatite and zircon; an interlaboratory comparison: United States Geological Survey Open-File report 78-108, 14 p.
  67. Nelson, K. D., 1992, Are crustal thickness variations in old mountain belts like the Appalachians a consequence of lithospheric delamination?: Geology, v. 20, p. 498-502.
  68. Nesbitt, B. E., and Muehlenbachs, K., 1989, Origins and movement of fluids during deformaion and metamorphism in the Canadian Cordillera: Science, v. 245, p. 733-736.
  69. Nesbitt, B. E., and Muehlenbachs, K., 1991, Stable Isotopic constrints on the nature of the syntectonic fluid regime of the Canadian Cordillera: Geophysical Research Letters, v. 18, p. 963-966.
  70. Nesbitt, B. E., and Muehlenbachs, K., 1995, Geochemical studies of the origins and effects of synorogenic crustal fluids in the southern Omineca Belt of British Columbia, Canada: Geological Society of America Bulletin, v. 107, p. 1033-1050.
  71. Nesbitt, B. E., Murowchick, J. B., and Muehlenbachs, K., 1986, Dual origins of lode gold deposits in the Canadian Cordillera: Geology, v. 14, p. 506-509.
  72. Nystrom, P. G., Jr., Willoughby, R. H., and Price, L. K., 1991, Cretaceous and Tertiary stratigraphy of the upper Coastal Plain, South Carolina, in Horton, J. W., Jr., and Zullo, V. A., eds., The Geology of the Carolinas, The University of Tennessee Press, Knoxville, p. 221-240.
  73. Parry, W. T., 1998, Fault-fluid compositions from fluid-inclusion observations and solubilities of fracture sealing minerals: Tectonophysics, v. 290, p. 1-26.
  74. Parry , W. T., and Bruhn, R. L., 1986, Pore fluid and seismogenic characteristics of fault rock at depth on the Wasatch fault, Utah: Journal of Geophysical Research, v. 91, p. 730-744.
  75. Parry, W. T., and Bruhn, R. L., 1990, Fluid pressure transients on seismogenic normal faults: Tectonophysics, v. 179, p. 335-344.
  76. Parry, W. T., Hedderly-Smith, D., and Bruhn, R. L., 1991, Fluid Inclusions and Hydrothermal alteration on the Dixie Valley fault, Nevada: Journal of Geophysical Research, v. 96, p. 19,733-19,748.
  77. Ragland, P. C., 1991, Mesozoic Igneous Rocks in Horton, J. W., Jr., and Zullo, V. A., eds., The Geology of the Carolinas, The University of Tennessee Press, Knoxville, p. 171-190.
  78. Ragland, P. C., Hatcher, R. D. Jr., and Whittington, D., 1983, Juxtaposed Mesozoic diabase dike sets from the Carolinas; A preliminary assessment: Geology, v. 11, p. 394-397.
  79. Robbins, G. A., 1972, Radiogenic argon diffusion in muscovite under hydrothermal conditions. MS Thesis, Brown University, Providence.
  80. Roedder, E. , 1984, Fluid Inclusions, Reviews in Mineralogy Volume 12, Mineralogical Society of America, 646p.
  81. Sacks, P. E., and Dennis, A. J., 1987, The Modoc zone – D2 (Early Alleghanian) in the eastern Appalachian Piedmont, South Carolina and Georgia: in Secor, D. T., Jr., ed., Anatomy of the Alleghanian orogeny as seen from the Piedmont of South Carolina and Georgia, South Carolina Geological Survey, Columbia, South Carolina, p. 18-34.
  82. Sacks, P. E., and Secor, D. T. Jr., 1990, Delamination in collisional orogens: Geology, v. 18, p. 999-1002.
  83. Sauniac, S., and Touret, J., 1983, Petrology and fluid inclusions of a quartz-kyanite segregation in the main thrust zone of the Himalayas: Lithos, v. 16, 35-45.
  84. Schafer, K. and Lax, E., 1962, Landolt-Bornstein Zahlenwerte und Funktionen aus Physik, Chemie, Astronomie, Geophysik und Technik, 6th ed. II, 2b, Springer-Verlag, Berlin, 760 p.
  85. Secor, D. T., Jr., 1987, Regional Overview: in Secor, D. T., Jr., ed., Anatomy of the Alleghanian orogeny as seen from the Piedmont of South Carolina and Georgia, South Carolina Geological Survey, Columbia, South Carolina, p. 1-18.
  86. Secor, D. T., Jr., Snoke, A. W., Bramlett, K. W., Costello, O. P., and Kimbrell, O. P., 1986, Character of the Alleghenian orogeny in the southern Appalachians: Part I. Alleghanian deformation in the eastern Piedmont of South Carolina: Geological Society of America Bulletin, v. 97, p. 1319-1328.
  87. Sibson, R. H., 1989, Earthquake faulting as a structural process: Journal of Structural Geology, v. 11, p. 1-14.
  88. Sibson, R. H., Moore, J. M., and Rankin, A. H., 1975, Seismic pumping – a hydrothermal fluid transport mechanism: Journal of the Geological Society of London, v. 131, p. 653-659.
  89. Sibson, R. H., Robert, F., and Poulsen, K. H., 1988, High angle reverse faults, fluid pressure cycling, and mesothermal gold-quartz deposits: Geology, v. 16, p. 551-555.
  90. Sisson, V. B., and Hollister, L. S., 1990, A fluid inclusion study of metamorphosed pelitic and carbonate rocks, south-central Maine: American Mineralogist, v. 75, p. 59-70.
  91. Speer, J. A., McSween, H. Y., Jr., and Gates, A., 1994, Generation segregation, ascent, and emplacemetn of Alleghanian plutons in the southern Appalachians, Journal of Geology, v. 102, p. 249-267.
  92. Stò we, K., and Barr, T. D., 1998, On uplift and exhumation during convergence: Tectonics, v. 17, p. 80-88.
  93. Sutter, J. F., 1988, Innovative approaches to the dating of igneous events in the early Mesozoic basins of the eastern United States:, in Froelich, A. J., and Robinson, G. R., Jr., eds., Proceedings, 2nd United States Geological Survey on the Early Mesozoic Basins of the astern United States: Unites States Geological Survey Circular 946, p. 110-114.
  94. Swanenberg , H. E. C., 1979, Phase equilibria in carbonic systems, and their application to freezing studies of fluid inclusions: Contributions to Mineralogy and Petrology, v. 68, p. 303-306.
  95. Vance, D., and Onions, R. K, 1992, Prograde and retrograde thermal histories from the central Swiss Alps: Earth and Planetary Science Letters, v. 114, p. 113-129.
  96. Vyhnal, C. R., and McSween, H. Y., Jr., 1990, Constraints on Alleghanian vertical displacements in the southern Appalachian Piedmont, based on aluminum-in-hornblende barometry: Geology, v. 18, p. 938-941.
  97. West, T. E., Secor, D. T., Pray, J. R., Boland, I. B., and Maher, H. D., 1995, New field evidence for the exposure of the Appalachian Decollement at the east end of the Pine Mountain terrane: Geology, v. 23, p. 621-624.
  98. Wickham, S. M., and Taylor, H. P., Jr., Stable isotope constraints on the origin and depth of penetration of hydrothermal fluids associated with Hercynian regional metamorphism and crustal anatexis in the Pyrenees: Contributions to Mineralogy and Petrology, v. 95, p. 255-268.
  99. Withjack, M. O., Schlische, R. W., and Olsen, P. E., 1998, Diachronous rifting, drifting, and inversion on the passive margin of central eastern North America: An analog for other passive margins: American Association of Petroleum Geologists Bulletin v. 82, p. 817-835.
  100. Zimmerman, R. A., 1979, Apatite fission track age evidence of post-Triassic uplift in the central and southern Appalachians: Geological Society of America Abstracts with Programs, v. 11, p. 219.
  101. Zoback, M. L., and Zoback, M. D., 1989, Tectonic stress field of the continental United States, in Pakiser, L. C., and Mooney, W. D., Geophysical framework of the continental United States: Boulder, Colorado, Geological Society of America Memoir 172, p. 523-539.

 

 

TABLE 1. Data Summary by Area

Area

Vein Set

Mineral Paragenesis

Inclusion Types

Th Range (°C)

Salinity
(Wt.% NaCl equiv.)

Modoc Fault Zone

Modoc Fault Zone

V4a

qtz± pyrite

Low XCO2

211-363

3.1-4.3

Low Salinity H2O

148-167

0.0-1.2

V4b

qtz± mus

Low XCO2

282-322

3.1-4.3

Low Salinity H2O

86-198

0.5-1.2

Northeast Carolina Terrane

Harbison Granite

V5

qtz± cc± ep± kfeld± chl± zeo± pre

H2O

66-108

0.0

V6

qtz± pyr± chl

Low XCO2

255-275

1.8-3.5

V7

cc± zeo

H2O

67-107

0.0

Columbia Granite

V5

qtz± cc

No data

No data

No data

V6

qtz± chl± cc± zeo± pre

Low XCO2

169-252

2.1-4.1

V7

cc± zeo

H2O

72-130

0.0-0.8

Asbill Pond Fmormation

V4a

qtz

Low XCO2

271-308

0.8-1.8

V4b

qtz± chl

Low XCO2

238-326

0.8-1.8

V5

qtz± chl

Low XCO2

241-297

0.8-1.8

Southwest Carolina Terrane

Persimmon Fork

V6

qtz

Low XCO2

200-259

3.6-4.3

Formation

Low Salinity H2O

138-162

1.0-1.1

H2O

82-224

0.0-0.9

Cuffytown Creek

V5

qtz

Low XCO2

287-303

4.9-4.5

Granite

H2O

77-191

0.0

High Grade

V5

qtz

Low XCO2

208-318

3.9-4.8

Carolina Terrane

Low Salinity H2O

152-162

No Data

Savannah River Terrane

Metavolcanics

V5

qtz

Low XCO2

198-322

No data

Low Salinity H2O

138-162

1.3-5.4

V6

qtz

CO2-rich

29-30

4.2-5.2

Low XCO2

217-307

2.2-6.2

Low Salinity H2O

86-221

0.6-5.4

Appling Granite

V5

qtz± cc± ep± kfeld± chl± zeo± sid

Low XCO2

215-341

2.6-5.7

Low Salinity H2O

152-181

1.5-1.8

cc = calcite, chl = chlorite, ep = epidote, kfeld = K-feldspar, pre = prehnite, pyr = pyrite, qtz = quartz, sid = siderite, zeo = zeolite.

 

Table 2. Summary of Data on H2O-NaCl-CO2 Fluid Inclusions

Sample

Mineral

Vein

Orientation

No.

TmCO2

C

Vapor

vol %

Tmcl

°C

ThCO2

°C

CO2

vol%

ThTot

°C

XH2O

XCO2

XNaCl

NaCl

wt %

 

Bulk r,
g cm-3

Southwest Carolina Terrane - Persimmon Fork Formation

CHL-A30-538

Quartz

328/73NE

1

-56.7

10

8.2

23.4L

0.29

251D

0.880

0.110

0.010

3.5

0.94

2

n.d.

10

8.2

23.3L

0.29

252L

0.880

0.110

0.010

3.5

0.94

Modoc Fault Zone -gneiss

NWB-C1-274

Quartz

330/70SW

1

-56.6

25

8.0

24.1L

0.59

362V

0.691

0.301

0.009

3.9

0.85

2

-56.6

65

7.9

25.2L

0.54

342V

0.735

0.255

0.010

4.1

0.85

3

-56.6

65

7.9

28.3L

0.50

324L

0.778

0.211

0.010

4.1

0.84

4

-56.6

25

7.8

28.3L

0.49

316D

0.785

0.205

0.011

4.3

0.84

5

-56.6

65

7.9

25.9L

0.54

344V

0.738

0.252

0.010

4.1

0.85

6

-56.7

50

8.0

19.2V

0.75

361V

0.802

0.188

0.010

3.9

0.40

7

-56.7

50

8.0

18.5V

0.52

243D

0.913

0.075

0.011

3.9

0.59

8

-56.7

50

8.0

19.5V

0.57

253D

0.895

0.094

0.011

3.9

0.55

9

-56.7

50

8.0

17.6V

0.51

241D

0.918

0.070

0.012

3.9

0.59

10

-56.6

40

7.8

18.2V

0.77

362V

0.790

0.199

0.011

4.3

0.37

11

-56.6

40

7.8

19.4V

0.47

222D

0.923

0.065

0.013

4.3

0.63

12

-56.6

40

7.9

17.4V

0.78

363V

0.786

0.204

0.010

4.1

0.36

13

-56.6

40

7.9

18.2V

0.75

349V

0.807

0.183

0.011

4.1

0.39

14

-56.7

40

7.8

18.2V

0.73

341V

0.821

0.168

0.011

4.3

0.41

15

-56.6

40

8.0

18.0V

0.73

341V

0.823

0.166

0.010

3.9

0.41

NWB-C1-276

Quartz

333/59SW

1

-56.6

50

7.7

24.0L

0.20

213V

0.917

0.070

0.013

4.4

0.97

2

-56.6

50

7.8

25.5L

0.23

225V

0.908

0.080

0.012

4.3

0.95

3

-56.6

50

7.8

25.2L

0.29

249L

0.881

0.107

0.012

4.3

0.93

4

n.d.

55

8.4

26.1L

0.40

296V

0.900

0.091

0.009

3.1

0.93

5

n.d.

60

8.4

25.7L

0.40

299V

0.899

0.092

0.010

3.1

0.93

6

-56.7

60

8.4

26.1L

0.41

309V

0.900

0.910

0.009

3.1

0.93

NWB-C1-267

Quartz

044/63SE

1

-56.6

35

7.7

30.5L

0.50

323V

0.798

0.190

0.011

4.4

0.80

2

-56.6

35

7.7

25.6L

0.41

297V

0.821

0.168

0.012

4.4

0.89

3

-56.6

35

7.8

25.6L

0.41

297V

0.821

0.168

0.011

4.3

0.89

Savannah River Terrane - Gneiss

CHL-A18-470

Quartz

348/60SW

1

-56.6

15

8.9

30.9L

0.47

308L

0.831

0.163

0.006

2.2

0.78

2

-56.6

15

8.8

30.9L

0.47

300L

0.831

0.163

0.006

2.4

0.78

CHL-A14-456

Quartz

271/72SW

1

n.d.

25

8.7

30.9V

0.47

305D

0.830

0.163

0.007

2.6

0.78

2

-56.7

25

8.8

29.2L

0.39

279D

0.852

0.142

0.006

2.4

0.86

3

-56.6

25

8.8

30.2L

0.37

271L

0.869

0.125

0.007

2.4

0.85

CHL-A54-739

Quartz

326/83SW

1

-56.6

25

7.9

30.9L

0.47

300V

0.826

0.163

0.011

4.1

0.79

2

-56.6

25

7.9

30.9L

0.37

270L

0.874

0.114

0.011

4.1

0.84

3

-56.6

25

8.8

30.2L

0.37

266V

0.869

0.125

0.007

2.4

0.85

4

-56.6

25

8.8

30.1L

0.30

237D

0.899

0.095

0.007

2.4

0.88

5

-56.7

25

7.8

30.9L

0.39

272L

0.882

0.106

0.010

4.3

0.85

CHL-A14-455

Quartz

001/81NW

1

-56.5

25

7.8

30.9L

0.36

264L

0.878

0.110

0.012

4.3

0.85

2

-56.6

15

7.8

30.9L

0.33

241L

0.890

0.098

0.012

4.3

0.86

3

-56.5

15

7.8

30.9L

0.36

256L

0.878

0.110

0.012

4.3

0.85

4

-56.6

15

7.8

30.9L

0.43

286L

0.846

0.142

0.012

4.3

0.81

Northeast Carolina Terrane - Harbison Granite

NWB-A1-1334

Quartz

345/88NE

1

-56.6

25

8.2

26.3

0.37

266L

0.849

0.141

0.010

3.5

0.90

2

-56.6

25

8.2

26.3

0.37

275L

0.848

0.143

0.010

3.5

0.90

3

-56.6

25

8.2

26.3

0.37

266L

0.849

0.141

0.010

3.5

0.90

4

-56.6

25

8.2

26.3

0.36

263L

0.852

0.138

0.010

3.5

0.90

5

-56.6

25

9.1

25.2

0.30

255L

0.884

0.112

0.005

1.8

0.92

6

-56.6

25

9.1

25.2

0.31

259L

0.879

0.116

0.005

1.8

0.92

7

-56.6

25

8.8

26.7

0.33

261L

0.871

0.122

0.008

2.4

0.90

8

-56.6

25

8.8

26.7

0.33

265L

0.871

0.122

0.007

2.4

0.91

9

-56.6

25

8.8

26.7

0.34

269L

0.867

0.127

0.008

2.4

0.91

TmCO2, melting temperature of CO2 solid; TmCl, melting temperature of clathrate; Vapor Vol. %, volume percent CO2 vapor relative to CO2 liquid at TmCO2; ThCO2, homogenization of CO2 vapor; CO2 vol. %, volume percent CO2 in inclusion at 35° C; ThTot, final homogenization temperature; r , density; L, homogenizes to a liquid phase; V, homogenizes to a vapor phase; D decrepitated.

 

 

TABLE 3. Summary of Data on High XCO2 Fluid Inclusions in Savannah River Terrane

Vein

TmCO2

Vapor

Tmcl

ThCO2

XCH4

CO2

Sample

Mineral

Orientation

No.

°C

vol %

°C

°C

Equiv. Den.*



g cm-3

CHL-A18-470

Quartz

348/60SW

1

-56.6

0.50

7.9

25.8L

0.04

0.550

Set V4a

2

-56.6

0.50

7.9

25.9L

0.04

0.550

3

-56.6

0.50

7.9

25.9L

0.04

0.550

4

-56.6

0.50

7.9

25.9L

0.04

0.550

5

-56.6

0.50

7.9

25.9L

0.04

0.550

6

-56.6

0.50

7.3

24.7L

0.04

0.600

7

-56.6

0.50

7.3

24.7L

0.04

0.600

8

-56.6

0.50

7.3

24.7L

0.04

0.600

9

-56.6

0.50

n.d.

25.4L

0.04

0.530

10

-56.6

0.50

n.d.

25.4L

0.04

0.530

11

-56.6

0.50

n.d.

25.4L

0.04

0.530

12

-56.6

0.50

n.d.

19.5L

0.04

0.760

13

-56.6

0.50

n.d.

19.5L

0.04

0.760

14

-56.6

0.50

n.d.

19.5L

0.04

0.760

15

-56.6

0.50

n.d.

20.9L

0.04

0.700

16

-56.6

0.50

n.d.

16.0L

0.04

0.770

17

-56.6

0.50

n.d.

18.5L

0.04

0.740

18

-56.7

0.50

n.d.

19.5L

0.04

0.760

19

-56.6

0.50

n.d.

18.9L

0.04

0.740

20

-56.6

0.50

n.d.

19.5L

0.04

0.760

Abbreviations as in Table 2. *Density is assuming inclusion is 100% CO2., n.d. = no data.

 


Figure 1. Geologic map of the study area (modified from Secor and others [1986] and Bramlett and others [1982]) showing sample localities (large filled circles). Also shown are post-kinematic granites: Appling (AP), Columbia (CO), Cuffytown Creek (CF), Graniteville (GV), Harbison (HB), and Winnsboro (WN).

 


Figure 2. Lower hemisphere equal-area stereonets of poles to veins. a) Modoc fault zone, b) northeast Carolina terrane, c) southwest Carolina terane, d) Savanah River terrane. Symbol code indicates the minerals present in each vein. V3a, V3b, V4a, V4b, and V5 are vein sets that correspond to deformation events discussed in text.

 


Figure 3. Fluid inclusion assemblage maps showing relationship between inclusion populations. a) Two intersecting planar clusters of pseudosecondary H2O inclusions. Note that the cluster with the lower average Th is clearly later. b) primary low-XCO2 inclusions and pseudosecondary planar clusters of low-salinity H2O inclusions. c) primary (right) and pseudosecondary (left) low-XCO2 inclusions. d) Primary low-XCO2 inclusions. e) primary low-salinity H2O inclusions. f) two intersections planar clusters of low-salinity H2O inclusions. Note that the cluster with the lower average Th is later. g) primary and pseudosecondary CO2-rich inclusions. Numbers are total homogenization temperatures for the aqueous phase and carbonic phase (values < 30.9°C). Maps are traced from microscope projections of polished thick sections. Maps a, b, c, and g are traced at 25°C, maps d, e, and f are traced at 15°C.

 


Figure 4. Summary of fluid inclusion data for each area. a) Modoc fault zone, b)northeast Carolina terrane, c) southwest carolina terrane, d) Savannah River terrane. TmIce is ice melting temperature, Tmcl is CO2 clathrate melting temperature. Qtz is quartz vein material, cc is calcite vein material. Some inclusions accounted for in the histograms do not have corresponding TmIce values.

 


Figure 5. P-T diagram showing trapping conditions for fluid inclusions in each area studied. Symbols are values derived from individual fluid inclusions (Table 2). Light shaded regions are range of trapping conditions for all Low XCO2 inclusions in each area. Dark shaded region is trapping conditions determined from CO2-rich inclusions (Table 3). Hatchured regions are trapping conditions determined for Low Salinity H2O and H2O inclusions. Also shown are lithostatic (L) and hydrostatic (H) thermobaric gradients for various geothermal gradients. Light shaded lines labeled 1 through 8 are mineral equilibria: 1. Prehnite = laumantite + kaolinite + quartz + water (Bird and Helgeson, 1981), 2. Stilbite = heulandite + water (Cho and others (1987), 3. Heulandite = laumantite + quartz + water (Cho and others, 1987), 4. Laumantite + prehnite + chlorite = pumpellyite (Liou and others, 1987), 5. Stilbite = laumantite + quartz + water (Liou, 1971), 6. Laumantite + prehnite = epidote + quartz + water (Liou and others, 1987), 7. Pumpellyite + laumantite = epidote + chlorite (Liou and others , 1987), 8. Epidote + mica + quartz = fledspar + water (Parray and Bruhn, 1986). Light shaded lines 9 and 10 are boiling curves for 0 wt. % and 5 wt. % NaCl equivalent salinity respectively (based on data from Haas, 1976).

 


Figure 6. Schematic cross-section across the study area (based on cross-section in Mahar, 1987). Section is restored to the late Alleghanian orogeny using pressures determined from fluid inclusion data and geobarometry of granites (Vyhnal and McSween, 1990). Arrows represent inferred fluid circulation paths.

 


Figure 7. Post-metamorphic thermal evolution of the northeastern Carolina terrane and Modoc fault zone. Shows a two-stage cooling history from the late Paleozoic through the late Mesozoic. Also shown is the range of trapping temperatures for inclusion types examined in this study. Emplacement temperatures of granites from Vyhnal and McSween (1990) and granite emplacement ages are: Appling (A) dated at 301± 3 Ma (D. Secor, personal communication, 2000), Columbia (C) dated at 286± 3 Ma (Fullagar and Butler, 1979), Harbison (H) dated at 309± 3 Ma (Fullagar and Kish, 1981), and Winnsboro (W) dated at 295± 2Ma (Fullagar, 1981). Retention of argon in Hornblende from Dallmeyer and others (1986). Retention of argon in muscovite from Maher and others (1994) and Pray and others (1997). Age of the potassium alteration (the 'Pinking event') from Kish (1992). Mesozoic dike emplacement age from Sutter (1988). Maintenance of apatite fission tracks from Zimmerman (1979), Earliest late Cretaceous surface from oldest Cretaceous sediments in the eastern Piedmont (Nystrom and others, 1992). Age range for Mesozoic rifting from Withjack and others (1999).

 

Figure 8. Evolution of pressure–temperature conditions for the northeastern Carolina terrane and Modoc fault zone. Note the similarity of granite emplacement pressures and the trapping of Low XCO2 inclusions at significantly lower temperatures reflecting near isobaric regional cooling. Dark shaded region is average conditions of granite emplacement from Vyhnal and McSween (1990) and Speer and others (1994). Light shaded area is range of trapping conditions determined from Low XCO2 fluid inclusions. Hatchured region is range of trapping conditions for Low Salinity H2O and H2O fluid inclusions.