ArticlePDF Available

HokieFlincs_H2O-NaCl: A Microsoft Excel spreadsheet for interpreting microthermometric data from fluid inclusions based on the PVTX properties of H2O–NaCl

Authors:
Short note
H
OKIE
F
LINCS
_H2O-N
A
C
L
: A Microsoft Excel spreadsheet for interpreting
microthermometric data from fluid inclusions based on the PVTX properties
of H
2
O–NaCl
Matthew Steele-MacInnis
n
, Pilar Lecumberri-Sanchez, Robert J. Bodnar
Department of Geosciences, Virginia Tech, Blacksburg, VA 24061, USA
article info
Article history:
Received 3 December 2011
Received in revised form
29 January 2012
Accepted 30 January 2012
Available online 6 February 2012
Keywords:
H
2
O–NaCl
Fluid inclusions
Microthermometry
PVTX
Isochores
1. Introduction
Sodium chloride (NaCl) is often the most common salt in
natural fluids (e.g., Fyfe et al., 1978). Consequently, phase equili-
brium (PTX) and volumetric data (PVTX) for H
2
O–NaCl are often
used to interpret microthermometric data from aqueous fluid
inclusions (FI), and for modeling natural aqueous electrolyte
fluids. The pressure–volume–temperature–composition (PVTX)
properties of H
2
O–NaCl have been well characterized by numer-
ous experimental and theoretical studies, as summarized in
Bodnar et al. (1985) and Bodnar (2003).
Several numerical models have been developed to represent
the PVTX properties of H
2
O–NaCl, facilitating interpretation of
data from FI. Those models can be categorized as (1) comprehen-
sive models (equations of state) that are not specifically designed
to interpret FI data (e.g., Anderko and Pitzer, 1993;Driesner,
2007;Driesner and Heinrich, 2007); or (2) correlation equations
designed specifically for FI research, with measurable quantities
(dissolution temperature, T
m
, and homogenization temperature,
T
h
) as independent variables (e.g., Bodnar et al., 1989;Bodnar,
1994;Bodnar and Vityk, 1994). Both types of models have
advantages and disadvantages in FI research. The comprehensive
models can be used as stand-alone internally consistent tools to
analyze fluid PVTX properties, but they require iterative calcula-
tions to interpret T
m
T
h
data, which can make data reduction time
consuming. The FI-specific models compute FI properties (com-
position, density, etc.) directly from T
m
and/or T
h
(without
iteration). However, no single model covers the complete range
of PVTX conditions of interest in most FI studies.
Owing to the piecemeal nature of the FI-specific models, inter-
preting microthermometric data can be tedious, requiring contin-
uous switching between the various models to obtain all the desired
information. To circumvent this issue, several easy-to-use computer
programs have been published that implement various models to
interpret H
2
O–NaCl (and other) microthermometric data (Bakker,
2003;Bodnar et al., 1989;Brown, 1989;Brown and Hagemann,
1995;Driesner, 2007;Driesner and Heinrich, 2007). These programs
have made the numerical models more accessible and applicable to
FI research. The programs provide powerful and adaptable tools for
interpreting FI data and for theoretical modeling (Bakker and Brown,
2003).Adrawbackofsomeoftheseprograms(e.g.,Bakker, 2003;
Driesner, 2007;Driesner and Heinrich, 2007)isthatafterresultsfor
one FI are computed, the program must be re-set before data for
another FI are entered. As a result, entry of large FI datasets is time
consuming, and it is not possible to have all data for an analytical
session saved to a single output file.
The ideal computer model allows the user to input tempera-
tures of phase changes commonly measured in the laboratory
during microthermometrynamely, dissolution temperatures
Contents lists available at SciVerse ScienceDirect
journal homepage: www.elsevier.com/locate/cageo
Computers & Geosciences
0098-3004/$ - see front matter &2012 Elsevier Ltd. All rights reserved.
doi:10.1016/j.cageo.2012.01.022
n
Corresponding author: Tel.: þ1 540 231 8575; fax: þ1 540 231 3386.
E-mail addresses: mjmaci@vt.edu (M. Steele-MacInnis),
pilar@vt.edu (P. Lecumberri-Sanchez), rjb@vt.edu (R.J. Bodnar).
Computers & Geosciences 49 (2012) 334–337
(T
m
) of various phases and homogenization temperature (T
h
)to
determine the composition, density, and pressure in the inclusion
at homogenization. From this information one can also calculate
the slope of the isochore and estimate the PT conditions of FI
formation, provided that either trapping Tor Pare known
independently and the inclusions are trapped in the single-phase
fluid field.
Here, we assemble the numerical formulae necessary to fully
characterize H
2
O–NaCl FI properties, based only on the tempera-
ture of dissolution of the last solid phase, T
m
, and the liquid-vapor
homogenization temperature, T
h
, obtained during microthermo-
metry. Note that an H
2
O–NaCl FI can exhibit more than two phase
changes during heating from temperatures less than the eutectic
temperature (21.2 1C; Hall et al., 1988) to the total homogeni-
zation temperature, but only the liquid–vapor homogenization
temperature and the temperature of disappearance of the last
solid phase are required to uniquely define the fluid composition
and density. Using these two input parameters, the program
calculates fluid density and salinity, and also estimates the
isochore slope, which allows the user to calculate a pressure
correction. Importantly, the program is designed such that the
user can quickly and easily reduce large datasets.
2. Program HokieFlincs_H2O-NaCl
H
OKIE
F
LINCS
_H2O-N
A
C
L
includes a set of FI-specific equations
(Fig. 1) that describe the PVTX properties of H
2
O–NaCl based only
on the measured T
m
, and T
h
. Depending on the mode of homo-
genization, different models may be required. For H
2
O–NaCl FI in
which the last solid phase to disappear is H
2
O–ice, salinity is
determined using the equation of Bodnar (1993). If the last solid
phase to disappear (in equilibrium with liquid and vapor) is either
hydrohalite or halite, equations from Sterner et al. (1988) are used
to determine the liquid composition. Correlation equations
derived in the present study are used to compute the densities
of liquid and vapor, the vapor composition, and the mass propor-
tions of liquid and vapor. The liquid and vapor compositions and
mass fractions are used to precisely determine the bulk composi-
tion. For homogenization in the absence of solids, pressure at T
h
and liquid density are represented by equations of Atkinson
(2002) and Bodnar (1983), respectively. The dP/dTslopes of
isochores in the liquid field are modeled using equations of
Bodnar and Vityk (1994), allowing direct determination of trap-
ping PT conditions if an independent estimate of either Por Tis
supplied. Note that although the slopes of fluid isochores gen-
erally have slight concave-down curvature, at crustal PT condi-
tions the isochores of H
2
O–NaCl fluids are essentially linear
(Bodnar and Vityk, 1994) and we assumed linearity in the
program. Critical properties are based on the model of Knight
and Bodnar (1989).
The numerical models described above cannot model the PVTX
properties of fluid inclusions for which T
h
oT
m
(i.e., inclusions
homogenize by halite disappearance). Therefore, in cases where
T
h
oT
m
, the salinity, pressure at homogenization, density and dP/
dTslope of the isochore in the liquid field are determined using
the model of Lecumberri-Sanchez et al. (in review), which is a
revision of an earlier model of Becker et al. (2008).
The Excel spreadsheet ‘‘H
OKIE
F
LINCS
_H2O-N
A
C
L
’’ (available in the
online Supporting Information, Appendix A) does not require
iterative procedures to interpret microthermometric data. The
user inputs a value of T
m
(specifying which solid phase, either
H
2
O–ice, hydrohalite or halite, is the last to dissolve) and T
h
,in1C
(Fig. 2). The fluid salinity, density and the isochore slope are
output (Fig. 2). If a pressure correction is required, the user inputs
an estimate of either Por Tof trapping, and the other quantity is
computed using the isochore slope and the slope-intercept
method. We avoided any formulae that require indirect solutions
(iteration) in order to avoid using macros in the program. There-
fore, H
OKIE
F
LINCS
_H2O-N
A
C
L
uses only Excel formulae and does not
employ Visual Basic (VBA) code. The program was designed in
this way because VBA macros are not supported in all versions of
Excel. Thus, because H
OKIE
F
LINCS
_H2O-N
A
C
L
does not use macros,
the program can be implemented using any version of Excel, on
both Mac and PC platforms.
As an example of the use of H
OKIE
F
LINCS
_H2O-N
A
C
L
, consider a FI
that contains liquid and vapor at room T, has final ice melting (in
the presence of vapor) at 10 1C, and homogenizes by vapor
bubble disappearance at 300 1C. To interpret these data, the value
‘‘10’’ is entered into the cell for T
m
(column C), and ‘‘ice’’ is
entered into the adjacent cell in column D(phase) (row 30 in the
screenshot; Fig. 2). Using the equation of Bodnar (1993) gives a
salinity of 13.9 wt% NaCl (column H;Fig. 2). To determine the
density of the FI and Pat homogenization, the homogenization
temperature (‘‘300’’) is entered into the same row in column E
(T
h
), and the program applies the equations of Bodnar (1983) and
Atkinson (2002) to estimate a bulk density of 0.86 g/cm
3
(column
J;Fig. 2) and Pat homogenization of 78 bar (column I;Fig. 2).
Using the equation of Bodnar and Vityk (1994), the isochore dP/dT
slope for this inclusion is 12.1 bar/1C (column K;Fig. 2). The
isochore slope can be used along with an independent estimate of
the trapping Tto determine the trapping P. For example, if
mineral equilibria indicate that the host phase for the FI formed
at 600 1C, the Pof formation would have been 3700 bar.
As another example, consider a FI that contains liquid, vapor
and halite at room temperature, and in which halite dissolves and
the bubble disappears at the same Tof 400 1C (i.e., the FI
homogenizes on the LVH curve). To interpret these microthermo-
metric data, ‘‘400’’ is entered into columns Cand E, and ‘‘halite’’ is
entered into column D(row 30; Fig. 2). Based on the mode of
homogenization, the models of Sterner et al. (1988) and Bodnar
(1983) are used to determine the salinity (47.4 wt% NaCl) and
density (1.09 g/cm
3
) of the FI (Fig. 2). The program uses the
models of Atkinson (2002) and Bodnar and Vityk (1994) to
estimate the P at homogenization (170 bar) and the slope of the
isochore (14.4 bar/1C) (Fig. 2). Note that most natural FI homo-
genize either by dissolution of the halite before the vapor bubble,
l
i
q
u
i
d
-
v
a
p
o
r
-
h
y
d
r
o
h
a
l
i
t
e
/
h
a
l
i
t
e
:
S
t
e
r
n
e
r
e
t
a
l
.
,
1
9
8
8
liquid-vapor-ice:
Bodnar, 1993
locus of critical points:Knight and Bodnar,1989
Atkinson, 2002
halite liquidus: Lecumberri-Sanchez et al., in review
Not shown:
+ liquid density: Bodnar, 1983;
+ isochore slope: Bodnar and Vityk, 1994
Pressure
Tem
p
erature
T = -21.2 to 700 °C
P = LVH to 6000 bar
X = 0 to 70 wt% NaCl
H
2
O liquid-vapor:
Atkinson, 2002
liquid-vapor curves:
Fig. 1. Schematic illustration of H
2
O–NaCl phase boundaries, showing the sources
of the various equations used in the present study. The halite liquidus line and
liquid–vapor curve both represent projections of constant liquid composition,
whereas the other curves represent a range of compositions. For complete
description of these features in PTX space, the reader is referred to Bodnar et al.
(1985) and Driesner and Heinrich (2007).
M. Steele-MacInnis et al. / Computers & Geosciences 49 (2012) 334–337 335
or by disappearance of the vapor bubble before the halite. The
significance of each mode of homogenization has been discussed
by Roedder and Bodnar (1980) and Bodnar (1994), and
Lecumberri-Sanchez et al. (in review) provide a model to interpret
data from FI in which T
h
oT
m
. All three modes of homogenization
can be modeled using (H
OKIE
F
LINCS
_H2O-N
A
C
L
).
For some FI it is not possible to measure the melting tem-
perature. For example, in some liquidþvapor FI the vapor bubble
occupies less than 10% of the inclusion volume and the ice
melts metastably during freezing experiments (Roedder, 1967).
Also, for very small FI the salinity is sometimes determined from
Raman analysis because ice melting cannot be observed during
microthermometry. In these cases, it is useful to have the option
to model the FI properties without entering a measured T
m
and
instead entering an estimate of salinity. H
OKIE
F
LINCS
_H2O-N
A
C
L
allows the user to input salinity directly for those FI for which
T
m
is not known (Column F,Fig. 2).
H
OKIE
F
LINCS
_H2O-N
A
C
L
can be used to determine properties of FI
that homogenize to the liquid phase. The program is generally
valid from 21.2 to 700 1C, the LV curve to 6000 bar and 0 to
70 wt% NaCl for FI that homogenize by vapor bubble disappear-
ance; and from T
h
of 100 to 600 1C, the LVH curve to 3000 bar and
28–75 wt% NaCl for FI that homogenize by halite disappearance.
For additional details, see Bodnar and Vityk (1994) and
Lecumberri-Sanchez et al. (in review). The program includes a
series of ‘‘if/then’’ statements to identify and alert the user to
potentially invalid input data. For example, if the user inputs a
value of 25 1C in the column for ice melting temperature, the
program recognizes that the input T
m
is less than the H
2
O–NaCl
eutectic temperature (21.2 1C; Hall et al., 1988) and a warning
appears in the same row in column ‘‘S’’. Similarly, the program
verifies that T
m
,T
h
and all estimated fluid properties (density,
pressure at homogenization, etc.) are within the ranges of validity
of the various numerical models.
A major advantage of H
OKIE
F
LINCS
_H2O-N
A
C
L
over other avail-
able programs for interpreting FI microthermometric data is the
speed and ease with which the user can interpret large datasets.
The spreadsheet is arranged such that T
m
and T
h
each occupy one
column (Fig. 2), and the maximum number of input data is only
limited by the number of rows available in Excel. An inclusionist
who regularly enters and stores microthermometric data in a
spreadsheet format needs only copy and paste T
m
and T
h
into the
proper columns to reduce the data. Because all computations are
direct, the results are immediately displayed when T
m
and T
h
are
entered. This is a significant advantage compared to previous
computer tools for interpreting FI data, especially in applications
where large numbers of FI are analyzed (e.g, thermal history
reconstructions, or studies of fluid evolution in hydrothermal ore
deposits). H
OKIE
F
LINCS
_H2O-N
A
C
L
is available for download as an
electronic annex to this paper (see Appendix A).
Acknowledgments
We thank Carlos Marques de Sa
´for discussions about inter-
pretation of microthermometric data, which motivated us to
create H
OKIE
F
LINCS
_H2O-N
A
C
L
. John Mavrogenes provided sugges-
tions for the program layout and experimented with an early
version. Two anonymous reviewers provided critical comments
that clarified and improved the manuscript and the program.
Financial support for MS-M was provided by the Institute for
Critical Technology and Applied Science (ICTAS) at Virginia Tech.
This material is based upon work supported by the US National
Science Foundation under grants nos. OCE-0928472 and EAR-
1019770 to RJB.
Appendix A. Supporting materials
Supplementary data associated with this article can be found
in the online version at doi:10.1016/j.cageo.2012.01.022.
References
Anderko, A., Pitzer, K.S., 1993. Equation-of-state representation of phase equilibria
and volumetric properties of the system NaCl–H
2
O above 573 K. Geochimica
et Cosmochimica Acta 57, 1657–1680.
Atkinson Jr., A.B., 2002. A model for the PTX properties of H
2
O–NaCl. Unpublished
M.Sc. Thesis, Virginia Tech, Blacksburg, pp. 133.
Bakker, R.J., 2003. Package FLUIDS 1. Computer programs for analysis of fluid
inclusion data and for modelling bulk fluid properties. Chemical Geology 194,
3–23.
Bakker, R.J., Brown, P.E., 2003. Computer modelling in fluid inclusion research. In:
Samson, I., Anderson, A., Marshall, D. (Eds.), Fluid Inclusions: Analysis and
Interpretation, Short Course 32, 32. Mineralogical Association of Canada short
course, pp. 175–212.
Becker, S.P., Fall, A., Bodnar, R.J., 2008. Synthetic fluid inclusions. XVII. PVTX
properties of high salinity H
2
O–NaCl solutions (430 wt% NaCl): Application to
fluid inclusions that homogenize by halite disappearance from porphyry
copper and other hydrothermal ore deposits. Economic Geology 103, 539–554.
Bodnar, R.J., 1983. A method of calculating fluid inclusion volumes based on vapor
bubble diameters and P-V-T-Xproperties of inclusion fluids. Economic Geology
78, 538–542.
Bodnar, R.J., 1993. Revised equation and table for determining the freezing point
depression of H
2
O–NaCl solutions. Geochimica et Cosmochimica Acta 57,
683–684.
Fig. 2. Screenshot of the Excel spreadsheet. Microthermometric data are placed in the leftmost columns (under ‘‘Inputs’’), and calculated PVTX results appear in the
adjacent columns to the right (under ‘‘Outputs’’). Further to the right, other columns allow the user to input an estimate of formation Tor Pto construct an isochore to
estimate trapping conditions.
M. Steele-MacInnis et al. / Computers & Geosciences 49 (2012) 334–337336
Bodnar, R.J., 1994. Synthetic fluid inclusions. XII. Experimental determination of
the liquidus and isochores for a 40 wt% H
2
O–NaCl solution. Geochimica et
Cosmochimica Acta 58, 1053–1063.
Bodnar, R.J., 2003. Introduction to aqueous-electrolyte fluid inclusions. In: Samson, I.,
Anderson, A., Marshall, D. (Eds.), Fluid Inclusions: Analysis and Interpretation, Short
Course 32, 32. Mineralogical Association of Canada short course, pp. 81–100.
Bodnar, R.J., Burnham, C.W., Sterner, S.M., 1985. Synthetic fluid inclusions in
natural quartz. III. Determination of phase equilibrium properties in the
system H
2
O–NaCl to 1000 1C and 1500 bars. Geochimica et Cosmochimica
Acta 49, 1861–1873.
Bodnar, R.J., Sterner, S.M., Hall, D.L., 1989. SALTY: a FORTRAN program to calculate
compositions of fluid inclusions in the system NaCl–KCl–H
2
O. Computers and
Geosciences 15, 19–41.
Bodnar, R.J., Vityk, M.O., 1994. Interpretation of microthermometric data for H
2
O–
NaCl fluid inclusions. In: de Vivo, B., Frezzotti, M.L. (Eds.), Fluid Inclusions in
Minerals, Methods and Applications. Blacksburg, pp. 117–130.
Brown, P.E., 1989. FLINCOR: A microcomputer program for the reduction and
investigation of fluid inclusion data. American Mineralogist 74, 1390–1393.
Brown, P.E., Hagemann, S.G., 1995. Macflincor and its application to fluids in Archean
lode-gold deposits. Geochimica et Cosmochimica Acta 59, 3943–3952.
Driesner, T., 2007. The system H
2
O–NaCl. Part II: Correlations for molar volume,
enthalpy, and isobaric heat capacity from 0 to 1000 1C, 1 to 5000 bar, and 0 to
1X
NaCl
. Geochimica et Cosmochimica Acta 71, 4902–4919.
Driesner, T., Heinrich, C.A., 2007. The system H
2
O–NaCl. Part I: Correlation
formulae for phase relations in temperature–pressure–composition space
from 0 to 1000 1C, 0 to 5000 bar, and 0 to 1 X
NaCl
. Geochimica et Cosmochimica
Acta 71, 4880–4901.
Fyfe, W.S., Price, N.J., Thompson, A.B., 1978. Fluids in the Earth’s Crust: Their
Significance in Metamorphic, Tectonic, and Chemical Transport Processes.
Elsevier Science Ltd., Amsterdam. 383pp.
Hall, D.L., Sterner, S.M., Bodnar, R.J., 1988. Freezing point depression of NaCl–KCl–
H
2
O solutions. Economic Geology 83, 197–202.
Knight, C.L., Bodnar, R.J., 1989. Synthetic fluid inclusions. IX. Critical PVTX proper-
ties of NaCl–H
2
O solutions. Geochimica et Cosmochimica Acta 53, 3–8.
Lecumberri-Sanchez, P., Steele-MacInnis, M., Bodnar, R.J., in review. A numerical
model to estimate trapping conditions of fluid inclusions that homogenize by
halite disappearance. Geochimica et Cosmochimca Acta.
Roedder, E., 1967. Metastable superheated ice in liquid water inclusions under
high negative pressure. Science 155, 1412–1417.
Roedder, E., Bodnar, R.J., 1980. Geologic pressure determinations from fluid
inclusion studies. Annual Review of Earth and Planetary Sciences 8, 263–301.
Sterner, S.M., Hall, D.L., Bodnar, R.J., 1988. Synthetic fluid inclusions. V. Solubility
relations in the system NaCl–KCl–H
2
O under vapor-saturated conditions.
Geochimica et Cosmochimica Acta 52, 989–1006.
M. Steele-MacInnis et al. / Computers & Geosciences 49 (2012) 334–337 337
... To ensure reproducibility, the eutectic (Te; i.e., the temperature Most of the analyzed FIs show a range of Te between -23.0° and -16.8°C, thus their compositions are best described by the H2O-NaCl-(±KCl) system. Salinities were calculated using Tm(ice) and the Excel spreadsheet of Steele-MacInnis et al. (2012). The salinity of halite-bearing inclusions was calculated from their Tm(halite) and total homogenization temperature (Th tot) using the equations of Sterner et al. (1988), whereas the equations of Bodnar (1993) were used for two-phase FI Tm(ice). ...
... Therefore, temperatures that are presented and discussed in this contribution are uncorrected. For reference purposes, pressure-corrected values for lithostatic and hydrostatic regimes at the maximum depth of emplacement (Bodnar and Vityk, 1994;Steele-MacInnis et al., 2012) are collated for all FIs in Appendix Table A4. ...
Article
Full-text available
The Sadisdorf Li-Sn-(W-Cu) prospect in eastern Germany is characterized by vein-and greisen-style mineral-ization hosted in and around a small granite stock that intruded into a shallow crustal environment. The nature and origin of this mineral system are evaluated in this contribution by a combination of petrography and fluid inclusion studies, complemented by Raman spectroscopy and whole-rock geochemical analyses. The early magmatic-hydrothermal evolution is characterized by a single-phase low-salinity (7.0 ± 4 wt % NaCl equiv), high-temperature (>340°C), CO2-CH4-bearing aqueous fluid, which caused greisen alteration and mineraliza-tion within the apical portions of the microgranite porphyry. The bimodal distribution of brine and vapor fluid inclusions, and the formation of a magmatic-hydrothermal breccia associated with the proximal vein mineral-ization are interpreted to mark the transition from lithostatic to hydrostatic pressure. The vein-and stockwork-style mineralization (main stage) displays lateral zonation, with quartz-cassiterite-wolframite-molybdenite mineral assemblages grading outward into base-metal sulfide-dominated assemblages with increasing distance from the intrusion. Late fluorite-bearing veinlets represent the waning stage in the evolution of the mineral system. The similarity in the homogenization temperature (250°-418°C) of fluid inclusions in quartz, cassiter-ite, and sphalerite across the Sadisdorf deposit suggests that cooling was not a significant factor in the mineral zonation. Instead, fluid-rock interaction along the fluid path is considered to have controlled this zonation. In contrast to quartz-, cassiterite-and sphalerite-hosted fluid inclusions, which have a salinity of 0.0 to 10.0 wt % NaCl equiv, the fluid inclusions in late fluorite veins that overprint all previous assemblages have a salin-ity of 0.0 to 3.0 wt % NaCl equiv and homogenize at temperatures of 120° to 270°C, thus indicating cooling with or without admixture of meteoric fluids during the waning stage of the mineral system. The Sadisdorf deposit shares similar characteristics with other deposits in the Erzgebirge region, including a shallow level of emplacement, similar mineralization/alteration styles, and a hydrothermal evolution that includes early-boiling, fluid-rock interaction, and late cooling. In contrast to most systems in the region, both proximal and distal mineralization are well preserved at Sadisdorf. The recognition of such spatial zoning may be a useful criterion for targeting greisen-related Li and Sn resources.
... Salinity (wt.% NaCl) and homogenization temperature were calculated by HokieFlincs (Steele-MacInnis et al., 2012). ...
Thesis
Full-text available
At the Toodoggone District of northern British Columbia, the Ranch area and the Silver Pond zone within the Lawyers area represent two distinct high-sulfidation epithermal occurrences that show some broad similarities but many significant differences. Similarities include some overlaps in terms of gangue mineralogy and alteration style, but pronounced differences include ore mineralogy (relatively simple, comprising few common sulfides at Silver Pond, versus more complex, comprising an array of sulfides and sulfosalts at Ranch) and especially textural styles of mineralization (with common open-space filling textures of veins and breccias at Silver Pond, versus mostly disseminated/replacement style at Ranch with the only open space being sparse vugs). Here, we explore the mineralogy, paragenesis, and fluid evolution of these two nearby yet disparate occurrences to deduce factors that influenced their different styles. Fluid inclusion results clearly indicate persistent boiling throughout the whole paragenesis of both systems. However, microthermometric analyses reveal distinct trends, with evidence of progressive cooling and dilution at Silver Pond, versus apparently isothermal mixing between saline and dilute fluids at Ranch. Specifically, the well-constrained paragenetic sequence of open-space vein minerals at Silver Pond shows a clear and monotonic progression from an early, ~325 °C fluid of ~15 wt% NaCl eq., towards a later fluid of ~100 °C and nearly zero salinity, consistent with progressive incursion of cold meteoric water. In contrast, the fluids at Ranch show little cooling but span an array of salinities from ~25 wt% NaCl eq. down to nearly zero, suggesting mixing between a saline magmatic fluid and hot, deeply-circulated groundwater. Hence, these two occurrences seem to represent interesting case studies that document different hydrothermal processes and fluid evolutions, leading to distinct mineralization styles in the epithermal environment. ii
Article
Full-text available
Phanerozoic orogenic gold deposits worldwide are commonly considered to be formed from metamorphic devolatilization of marine carbonaceous sedimentary rocks. Here we show that the Yindongpo gold deposit from the Qinling orogen (central China) is genetically associated with the metamorphism of volcanic rocks during the late Paleozoic orogeny, which involved the closure of the Shangdan ocean. Gold mineralization at Yindongpo is hosted in lower Paleozoic metavolcanic-sedimentary sequences and occurs mainly as lenticular to stratiform ore bodies that formed in three paragenetic stages represented by quartz-ankerite-pyrite (stage I), quartz-carbonate-sulfide (stage II) and quartz-calcite assemblages (stage III), respectively. Rutile grains coexisting with auriferous pyrite from stage II yield U–Pb ages of 395 ± 9 to 400 ± 13 Ma (2σ). Fluid inclusions in quartz of stages I and II are dominated by CO2-rich (~ 10 mol%) aqueous fluids with low salinities (< 4.9 wt% NaCl equivalent) and total homogenization temperatures ranging from 241 to 352 ºC, whereas the values for H2O-rich inclusions of stage III are 0.2 to 2.6 wt% NaCl equivalent and 151 to 164 °C. Based on secondary ion mass spectrometry analysis of oxygen isotopes of quartz (Qz-1 to Qz-4), the calculated δ¹⁸Ofluid values for the quartz-forming fluids are 1.3 to 7.0‰ in stage I, –3.1 to 6.6‰ in stage II, and –9.6 to –3.7‰ in stage III. These data indicate a metamorphic origin of ore fluids that underwent Rayleigh fractionation and incursion of meteoric water. The large variation in ⁴⁰Ar*/⁴He ratios (1.7–30.0), caused by accumulation of radiogenic Ar* and He loss within some pyrite samples, can be ascribed to regional metamorphism and deformation. Ore sulfides have sulfur (δ³⁴SV-CDT = –2.1 to 3.3‰) and lead (²⁰⁶Pb/²⁰⁴Pb = 17.008–17.152, ²⁰⁷Pb/²⁰⁴Pb = 15.402–15.493, and ²⁰⁸Pb/²⁰⁴Pb = 38.254–38.564) isotopic compositions that are consistent with those of pyrite in the metavolcanic host rocks. Results presented here suggest that the ore fluids and, by inference, gold of the Yindongpo deposit were derived primarily from the volcanic sequences during regional metamorphism and deformation in response to the Early Devonian Qinling collisional orogeny. The Yindongpo deposit represents the first recognized Paleozoic orogenic gold deposit in the Qinling orogen, and thus has important implications for regional metallogeny and gold exploration.
Article
Full-text available
The Xianghualing large tin-polymetallic skarn deposit is located in the Nanling W-Sn metallogenic belt, South China, showing distinct spatial zoning of mineralization. From the contact between granite and carbonate rocks, the mineralization transitions from proximal skarn Sn ore to cassiterite-sulfide ore and more distal Pb–Zn-sulfide ore. This study reveals the fluid evolution and genetic links among these different ore types. The physical and chemical characteristics of fluid inclusions from each ore types indicate that the skarn Sn ore, cassiterite-sulfide ore, and Pb–Zn-sulfide ore all originated from the identical magmatic fluid exsolved from the Laiziling granite. Their formation, however, is controlled by diverse fluid evolutionary processes and host rock characteristics. The Sn–Pb-Zn-rich fluids were primarily derived from cooled and diluted magmatic brine, which is generated by boiling of initial single phase magmatic fluid. Mixing of magmatic brine with meteoric water is crucial to form skarn Sn ore. Redox reactions of aqueous Sn (II) complexes with As (III) species and/or minor CO2 during short cooling period of ore-forming fluid is likely an effective mechanism to form high-grade cassiterite-sulfide ores, accompanied by favorable pH conditions maintained through interaction with carbonate host rocks. The later stage addition of meteoric water prompts the formation of Pb–Zn-sulfide ore. Comparing these findings with the characteristics of initial or pre-ore magmatic fluids in both mineralized and barren granitic systems indicates that high Sn content in the pre-ore fluids and the suitable fractional crystallization degree of the parent magma may determine high Sn mineralization potential in granitic magmatic-hydrothermal systems.
Article
Full-text available
The Lindero deposit is located in the Puna plateau, northwest Argentina, at the southern end of the Central Volcanic Zone of the Central Andes. The high-K calc-alkaline dioritic composition of the subvolcanic intrusions, the shallow emplacement depth (< 1.5 km), and the gold-rich and copper-depleted mineralization style suggest that the Lindero deposit is a porphyry gold deposit. Porphyry gold deposits are scarce worldwide and the factors controlling their formation are still poorly known. Here we present a detailed study of fluid inclusions in order to characterize the mineralizing fluids that precipitated the Au mineralization at Lindero. Different types of fluid inclusions in quartz veins (A-type and banded quartz), which are associated with the K-silicate alteration, were analyzed using Raman spectroscopy, microthermometry, and LA-ICP-MS (laser ablation inductively coupled plasma mass spectrometry). Four inclusion types can be recognized in quartz veins: (i) Salt melt inclusions, which are characterized by a dense packing of daughter minerals (mainly Fe-chloride, sylvite, halite, anhydrite, and hematite), by a distorted vapor bubble, and by the lack of liquid phase; (ii) Halite-bearing inclusions which contain liquid, vapor, and halite; (iii) Two-phase aqueous inclusions that contain liquid and vapor; (iv) Vapor-rich inclusions containing only vapor. The inclusion types are related to different stages of hydrothermal evolution. Stage 1 is the main mineralization stage, characterized by vapor-rich inclusions coexisting with salt melt inclusions. Salt melt inclusions commonly show total homogenization temperature (ThL) > 1000 °C. This Na-K-Fe-Cl-rich highly saline brine (~ 90 wt% NaCl eq.) was of magmatic origin and responsible for the Au mineralization. Two later stages involving cooler fluids (ThL < 300 °C) and gradually lower salinities (from 36.1 to 0.2 wt% NaCl eq.) trapped by halite-bearing and two-phase aqueous inclusions during stages 2 and 3, respectively, correspond to a late magmatic-hydrothermal system, that is probably related to a deep supercritical fluid exsolution. Salt melt inclusions represent the most likely parental fluid of K-silicate alteration and associated Au mineralization at Lindero. This uncommon type of fluid must have played an important role in Au transport and precipitation in shallow porphyry gold deposits.
Article
Full-text available
Porphyry-epithermal veins hosting Re-rich molybdenite and rheniite (ReS2) from the Maronia Cu-Mo ± Re ± Au porphyry in Thrace, NE Greece, provide new insights into the hydrothermal processes causing extreme Re enrichment. Quartz trace element chemistry (Al/Ti, Ge/Ti), Ti-in-quartz thermometry, and cathodoluminescence imaging reveal multiple quartz generations in consecutive hydrothermal quartz-sulfide veins associated with potassic, sericitic, and argillic alteration. Fluid inclusions in different quartz generations indicate that phase separation and fluid cooling are the main ore-forming processes in the porphyry stage (~ 500 – 350 °C), whereas mixing of a vapor-rich fluid with metalliferous (e.g., Pb, Zn, Au) meteoric water forms the epithermal veins (~ 280 °C). These processes are recorded by trace element ratios in pyrite that are sensitive to changes in fluid temperature (Se/Te), fluid salinity (As/Sb, Co/As), and mixing between fluids of magmatic and meteoric origin (Se/Ge). Highly variable intra-grain δ³⁴S values in pyrite record S isotope fractionation during SO2 disproportionation and phase separation, emphasizing the importance of in situ δ³⁴S analysis to unravel ore-forming processes. High δ³⁴S (~ 4.5‰) values of sulfides are indicative of low SO4²⁻/H2S fluid ratios buffered by the local host rocks and mixing of the magma-derived fluid with meteoric water. The formation of Re-rich molybdenite (~ 6600 ppm) is favored by cooling and reduction of a magma-derived, high-temperature (~400 °C), oxidized, and Re-rich fluid triggering efficient Re precipitation in early veins in the potassic alteration zone. The systematic temporal fluid evolution therefore reveals that coeval cooling and reduction of oxidized Re-rich fluids cause extreme Re enrichment at the Maronia porphyry system.
Article
Full-text available
FLINCOR is designed to reduce laboratory data gathered on fluid inclusions and to calculate P-T isochores for geologically important fluids composed of H 2O, CO 2, CH 4, NaCl, CO, and N 2. Requirements for input data have been oriented toward the types of values derived from fluid-inclusion observations. Alternatively, fluid mixtures expressed in a variety of units (for instance, mole fraction), can serve as input data. The user can choose from among multiple published equations of state describing the fluid behaviour and can readily compare the results obtained by using different equations. Calculated output can be either printed in tabular form or saved as an ASCII text file for exportation to other applications. FLINCOR is a Microsoft Windows application. -Author
Article
Full-text available
Fluid inclusions (FI) that homogenize by halite disappearance are common in some geological environments, and interpretation of microthermometric data from these inclusions has been limited by the lack of a model describing the PVTX relationships over the complete range of PTX conditions found in nature. In this study, a system of equations has been developed that can be used to estimate salinity, pressure and density of FI that homogenize by halite disappearance. The salinity, pressure, density and dP/dT slope of the FI isochore are calculated as functions of liquid–vapor homogenization temperature (Th) and halite dissolution temperature (Tm). The equations are based on a numerical model describing the isochoric pressure–temperature trajectory followed by halite-saturated fluids during heating. The model is valid for Th and Tm from 100 to 600 °C, and for pressures along the liquid–vapor–halite curve to 300 MPa.
Article
Realistic simulations of fluid flow in geologic systems have severely been hampered by the lack of a consistent formulation for fluid properties for binary salt-water fluids over the temperature-pressure-composition ranges encountered in the Earth's crust. As the first of two companion studies, a set of correlations describing the phase stability relations in the system H2O-NaCl is developed. Pure water is described by the IAPS-84 equation of state. New correlations comprise the vapor pressure of halite and molten NaCl, the NaCl melting curve, the composition of halite-saturated liquid and vapor, the pressure of vapor + liquid + halite coexistence, the temperature-pressure and temperature-composition relations for the critical curve, and the compositions of liquid and vapor on the vapor + liquid coexistence surface. The correlations yield accurate values for temperatures from 0 to 1000 degrees C, pressures from 0 to 5000 bar, and compositions from 0 to 1 X-NaCl (mole fraction of NaCl). To facilitate their use in fluid flow simulations, the correlations are entirely formulated as functions of temperature, pressure and composition.
Article
A set of correlations for the volumetric properties and enthalpies of phases in the system H2O–NaCl as a function of temperature, pressure, and composition has been developed that yields accurate values from 0 to 1000 °C, 1 to 5000 bar, and 0 to 1 XNaCl. The volumetric properties of all fluid phases from low-density vapor to hydrous salt melts and single-phase binary fluids at high pressures and temperatures, can be described by a simple equationVsolution(T,P,XNaCl)=VH2O(n1+n2T,P)i.e., for a given pressure and composition, the molar volume of the solution is related to the molar volume of pure water at the same pressure by a linear scaling of temperature. The parameters n1 and n2 are simple functions of pressure and composition. This linear relation could be demonstrated for all (P, XNaCl) pairs where accurate volumetric data are available over a sufficiently large temperature interval. Extrapolations over as much as 300 °C predict high temperature data within their experimental uncertainty of 1–2%. With a simple fit of parameters n1 and n2, the vast majority of several thousand experimental data points from the literature can be reproduced within experimental error, including dilute solutions in the compressible region. Although a strict theoretical foundation is lacking, this behavior can phenomenologically be rationalized as reflecting the relation between the linear to near-linear isochores of binary fluids to those of pure water. Accordingly, small deviations from linearity that amount to 0.1–0.2% error in molar volume occur only at low temperatures where the pure water isochores behave non-linearly. This effect is accounted for by introducing a deviation function. Given its high accuracy, the formulation for the molar volumes could be integrated along isotherms–isobars to generate data for the specific enthalpy of binary solutions. In order to allow direct computation rather than numerical integration, a correlation scheme was developed that is mathematically similar to that for molar volumes. Specific enthalpies computed from the correlation typically agree within 1–3% with those obtained from other studies. Similarly, isobaric heat capacities show good agreement with published data except for high salinities at moderate pressures and temperatures.
Article
Realistic simulations of fluid flow in geologic systems have severely been hampered by the lack of a consistent formulation for fluid properties for binary salt–water fluids over the temperature–pressure–composition ranges encountered in the Earth’s crust. As the first of two companion studies, a set of correlations describing the phase stability relations in the system H2O–NaCl is developed. Pure water is described by the IAPS-84 equation of state. New correlations comprise the vapor pressure of halite and molten NaCl, the NaCl melting curve, the composition of halite-saturated liquid and vapor, the pressure of vapor + liquid + halite coexistence, the temperature–pressure and temperature–composition relations for the critical curve, and the compositions of liquid and vapor on the vapor + liquid coexistence surface. The correlations yield accurate values for temperatures from 0 to 1000 °C, pressures from 0 to 5000 bar, and compositions from 0 to 1 XNaCl (mole fraction of NaCl). To facilitate their use in fluid flow simulations, the correlations are entirely formulated as functions of temperature, pressure and composition.
Article
An algorithm for estimating compositions of fluid inclusions approximated by the NaClKClH2O system has been developed using recently published data for phase equilibrium properties of vaporsaturated NaClKClH2O solutions. The algorithm satisfactorily reproduces experimental results for all compositions in the ternary system and requires only the temperatures of two phase changes to calculate the total salinity and the NaCl/(NaCl + KCl weight fraction of the inclusion fluid. The required input data are the temperatures at which the final two solid phases dissolve during heating. These data are obtainable easily during standard microthermometric analysis of the inclusions.