You are on page 1of 20

Bodnar RJ (2003) Reequilibration of fluid inclusions. In I. Samson, A. Anderson, & D. Marshall, eds.

Fluid Inclusions: Analysis and Interpretation. Mineral. Assoc. Canada, Short Course 32, 213-230.

CHAPTER 8. REEQUILIBRATION OF FLUID INCLUSIONS

Robert J. Bodnar
The Bubble Factory
Virginia Tech
Blacksburg, VA 24061 USA
rjb@vt.edu

INTRODUCTION
A fluid inclusion provides microthermo - (Fig. 8-1). The term was first introduced by
metric data that can be used to infer trapping Larson et al. (1973) to describe permanent,
conditions only if the inclusion satisfies several nonelastic deformation of the walls of inclusions
criteria (Roedder 1984). First, the inclusion must in fluorite and sphalerite as a result of
trap a single, homogeneous phase. Secondly, the overheating. Stretching usually results when an
inclusion must remain a constant volume system inclusion reequilibrates in a low strain rate
after trapping (excluding reversible elastic environment in which the host phase deforms
volume changes) and, thirdly, nothing can be plastically, and is more common in softer
added to, or removed from the inclusion follow- minerals than in harder minerals. Any material
ing trapping. These criteria are here referred to as that is lost from the inclusion occurs as a result of
"Roedder's Rules", and must be met if one wishes diffusion or leakage along dislocations (Wilkins
to use inclusions to reconstruct the thermobaric et al. 1981, Bodnar & Bethke 1984, Wilkins
history of the host rocks. If the inclusion volume 1986, Bakker & Jansen 1991, 1994, Hall &
changes, or if anything is added to or lost from Sterner 1993, Vityk et al. 2000).
the inclusion following trapping, the inclusion is With increasing strain rate, fractures may
said to have reequilibrated. Reequilibration can develop in the host adjacent to inclusions, with
occur in nature during continued burial or uplift partial loss of fluid by advection along the
of the rock to the surface, or reequilibration may microfractures. This type of reequilibra tion is
occur during sample preparation and/or data referred to as leakage or partial decrepitation.
collection in the laboratory. Less often considered Sometimes, the extent of fracturing is so intense
by inclusionists is the possibility that fluid that all fluid is lost, leaving an empty, dark cavity
inclusions in samples collected from surface in the host phase. This is referred to as
outcrops may have experienced elevated temper- decrepitation. Note that the resulting inclusions
atures during exposure to the sun or to forest fires, are probably not sensu stricto empty but, rather,
or may have undergone freezing in colder likely contain a low-density version of the
climates. Especially for low temperature original fluid (water vapor or CO2 gas, for
inclusions, these temperature fluctuations may example). The inclusions are referred to as being
have been sufficient to cause reequilibration. empty because no fluid phase is observed during
As a result of numerous studies of both petrographic or microthermometric examination.
natural and synthetic fluid inclusions over the past The term decrepitation was introduced
few decades, we now have a fairly comprehensive by researchers in the former Soviet Union to
understanding of the reequilibration process. describe the explosive release of inclusion
Before I discuss the physical and chemical factors contents during heating. In the early days of fluid
that affect the reequilibration process, it is first inclusion research, before the development of
necessary to introduce the various terms that are microscope heating/cooling stages and high-
used to describe fluid inclusion reequilibration. In quality long-working distance microscope
the fluid inclusion literature, reequilibration is a objectives, workers would heat mineral samples
general term that has been used to describe fluid in a furnace with a sensitive microphone attached.
inclusions that have cha nged volume, or have lost The microphone records acoustic emissions (i.e.,
or gained components, or both. Stretching is a noise) associated with decrepitating fluid
term that is generally used to describe fluid inclusions, and the decrepitation temperatures
inclusions whose volume has changed with little would be related to inclusion trapping
or no discernible loss of fluid from the inclusion temperatures (incorrectly, in many cases!).

213
Type of Deformation Deformation Fluid Loss Observational
Reequilibration Environment Mechanism Mechanism Result

STRETCHING LOW PLASTIC DIFFUSION UNDETECTABLE


STRAIN FLUID LOSS (±H2)
RATE

LEAKAGE

DETECTABLE
FLUID LOSS
PARTIAL
DECREPITATION

HIGH
STRAIN COMPLETE
DECREPITATION RATE BRITTLE ADVECTION FLUID LOSS

FIG. 8-1 Schematic representation of the relationship between different types of fluid inclusion
reequilibration (stretching, leakage, decrepitation) and the environment of deformation (strain rate), the
mechanisms of deformation (plastic vs. brittle) and fluid loss (diffusion vs. advection), and the amount of
fluid lost from the inclusion during deformation (undetectable to complete). According to these
relationships, stretching is associated with plastic deformation in a low strain environment, where
undetectable amounts of fluid may be lost from the inclusion by diffusion. Conversely, decrepitation is
associated with brittle deformation in a high strain environment, with partial to complete loss of fluid by
advection along fractures. (from Vityk et al. 2000).
Related to this, Hall & Bodnar (1989) showed during cooling, and were able to reverse this
that energy detected during acoustic emission process during laboratory heating. In this chapter,
studies of natural rocks and minerals is directly these somewhat specialized cases of
related to decrepitation of fluid inclusions in the reequilibration will not be considered further.
sample. The type of reequilibration that occurs is
An additional type of reequilibration a function of many factors, and in a given fluid
sometimes occurs when the fluid in the inclusion inclusion assemblage some inclusions may show
reacts with the host phase after trapping to evidence of significant reequilibration, while
generate a new mineral assemblage. For example, others may show no evidence of reequilibration.
Kleinefeld & Bakker (2002) report the presence Below I outline some of the many factors that
of calcite, paragonite and pyrophyllite crystals in affect the reequilibration process. Of particular
fluid inclusions in plagioclase, and relate their importance is the effect of time on the
origin to reaction of the CO2 -rich fluid with the reequilibration process, because this may affect
calcite host. Similarly, Heinrich & Gottschalk the applicability of laboratory-based studies to
(1995) observed inclusions containing calcite and reequilibration of fluid inclusions in nature. The
quartz crystals in wollastonite. These phases were intent of this chapter is to provide information
generated by the reaction of the wollastonite host that inclusionists might gain from petrographic
with the CO 2 -bearing fluid during retrogression. and microthermometric analyses of inclusions,
Kodera et al. (2003) reported the occurrence of and excludes detailed theoretical discussions of
ferropyrosmalite [(Fe,Mn)8Si6 O15 (OH,Cl)10 ] in the reequilibration process. Note, however, that
fluid inclusions in magmatic quartz. These the empirical observations presented below are
workers attribute the origin of this phase to generally consistent with theoretical models for
reaction of an original clinopyroxene daughter mineral deformation. For a more detailed
mineral with the saline solution in the inclusion discussion of the theoretical basis for mineral

214
deformation, the reader is referred to Poirier Among the first to recognize the rel-
(1985) and to Küster & Stöckhert (1997) for a ationship between mineral hardness and the ease
discussion of rock deformation theory applied to with which fluid inclusions reequilibrate were
fluid inclusion reequilibration. Tugarinov & Naumov (1970), who noted a linear
relationship between Mohs hardness and the
WHAT DETERMINES HOW EASILY AN pressure required to decrepitate fluid inclusions in
INCLUSION WILL REEQUILIBRATE? several minerals (Fig. 8-2). Thus, fluid inclusions
Mineralogy in soft minerals such as orpiment and stibnite
The most important factor that (Mohs hardness ˜ 2) will reequilibrate
determines the ease with which fluid inclusions (decrepitate) when the internal pressure reaches
reequilibrate is the structure and composition of only a few hundred bars, whereas about 800 bars
the host phase. As fluid inclusion reequilibration of internal pressure is required to initiate
involves the breaking or modification of bonds reequilibration of inclusions in quartz (Mohs
within the host mineral, the ease with which a hardness ˜ 7). Faiziev & Alidovov (1976) also
mineral reequilibrates can be correlated with the noted that inclusions in soft minerals with perfect
weakest bond in that mineral. While there are cleavage (e.g., fluorite, calcite, sphalerite, and
many ways of characterizing bond strength, one scheelite) show an increase in measured
of the most basic is related to the hardness of the homogenization temperature during repeated
mineral. Minerals with small atoms that are measurements. These workers attributed reequil-
tightly packed and involve covalent bonds are ibration to the presence of weak bonds between
harder than those containing larger atoms that are cleavage layers, consistent with observations by
held together by van der Waals or metallic bonds. Tugarinov & Maumov (1970). The consistent

FIG. 8-2. Relationship between the internal pressure necessary to initiate reequilibration of fluid inclusions
and the Moh s hardness of the host mineral (modified from Tugarinov & Naumov, 1970). Ranges in
pressure for calcite, barite, sphalerite, fluorite and quartz represent the variation in pressure as a function of
inclusion size. Size data for quartz are from Bodnar et al. (1989); data for fluorite and sphalerite are from
Bodnar & Bethke (1984); data for barite are from Ulrich & Bodnar (1988); data for calcite are estimated
from Hall et al. (1993) and Prezbindowski & Larese (1987).

215
relationship between Mohs hardness and the (1987, 1988) found that fluid inclusions in fluorite
pressure required to cause reequilibration of fluid decrepitated with complete loss of fluid when
inclusions makes it possible to estimate the overheated in the laboratory at one atmosphere
suitability of a given mineral for fluid inclusion confining pressure, whereas inclusions from the
studies. Thus, one would expect that gypsum, same sample that were overheated under 200 or
with a Mohs hardness of 2, might not be the best 400 bars of confining pressure reequilibrated by
host for fluid inclusion studies, whereas stretching.
corundum, with a Mohs hardness of 9, would be
an ideal mineral for fluid inclusion studies. Inclusion Size
It should be recognized that Mohs Whereas the mineralogy of the inclusion
hardness is not a quantitative scale but, rather, a host appears to be the most important factor
qualitative scale that makes it convenient to controlling the pressure required to reequilibrate
compare the relative hardness or tensile strength fluid inclusions, the "average" pressure can vary
of two minerals. The absolute hardness (or tensile significantly depending on inclusion size.
strength) of a mineral is not a linear function of Naumov et al. (1966) and Tugarinov & Naumov
Mohs hardness. Thus, the difference in absolute (1970) determined that fluid inclusions larger than
hardness between quartz (Mohs hardness = 7) and about 35 µm in quartz decrepitate when the
topaz (Mohs hardness = 8) is about the same as internal pressure reaches about 850 bars. Leroy
the difference in absolute hardness between talc (1979) reported that this pressure increases to 1.2
(Mohs hardness = 1) and quartz. As such, Figure kbar for 12 -13 µm fluid inclusions, and that
8-2 shows the relationship between pressure inclusions smaller than 12 µm did not decrepitate
required to initiate reequilibration and Mohs even when the internal pressure reached 2.7
hardness over the range of experimental kbars. Swanenburg (1980) reported that
observations (Mohs hardness ˜ 1-7), but should approximately 1 µm inclusions were able to
not be applied to minerals with a Mohs hardness maintain internal pressures of 5–7 kbars before
greater than that of quartz (=7). It is expected that decrepitating.
in the future additional empirical data will Bodnar et al. (1989) conducted an
become available for minerals with Mohs experimental study of the decrepitation behavior
hardness above that of quartz, allowing the trend of fluid inclusions in quartz and combined their
shown on Figure 8-2 to be extended to higher results with those of Leroy (1979) and
values of Mohs hardness. Swanenburg (1980) to produce the following
The type of reequilibration that occurs is relationship between inclusion size and the
also somewhat related to mineral hardness. As internal pressure required to initiate
noted above, soft minerals tend to reequilibrate by reequilibration (Fig. 8-3):
stretching to produce a gradual increase in
homogenization temperature as reequilibration Pressure (kbar) = 4.26 D-0.423 (1)
proceeds, whereas harder minerals such as quartz where D is the inclusion diameter in micrometers.
tend to maintain their original homogenization According to equation (1), the internal pressure
temperature until the inclusion decrepitates with required to reequilibrate inclusions in quartz
complete loss of fluid. It should be emphasized, varies from 0.8 kbar for a 50 µm inclusion, to 1.6
however, that many exceptions to this general kbars for a 10 µm inclusion, to 4.3 kbars for a 1
trend have been reported. For example, Ulrich µm inclusion (Fig.8- 3).
and Bodnar (1988) observed decrepitation of Bodnar and Bethke (1984) found a
inclusions in the soft mineral barite (Mohs similar relationship between inclusion size and
hardness ˜ 3), and Bodnar et al. (1989) noted that the minimum pressure required to initiate
some fluid inclusions in quartz reequilibrated by stretching of fluid inclusions in fluorite.
stretching (followed by later decrepitation). According to these workers, the internal pressure
Additionally, minerals that reequilibrate by brittle required to initiate stretching decreases with
failure (decrepitation) in the laboratory may inclusion size according to:
reequilibrate by stretching or leakage under the
higher confining pressures and much lower strain Pressure (bars) = 900 - 147 log10V (2)
rates common in nature. Thus, Guilhaumou et al.

216
FIG. 8-3. Relationship between inclusion size and the pressure required to reequilibrate fluid inclusions in
quartz under one atmosphere confining pressure. Inclusion volumes shown on the lower abscissa were
converted to inclusion diameter assuming a spherical geometry and are shown on the top abscissa (modified
from Bodnar et al. 1989).
where V is the inclusion volume in µm3 . concentration surrounding flaws (fluid inclusions)
Assuming a spherical geometry, this translates to in minerals.
about 635 bars for a 5 µm inclusion, and 200 bars Prezbindowski & Larese (1987)
for a 50 µm inclusion. Bodnar & Bethke (1984) conducted an experimental study of stretching of
also observed a clear relationship between fluid inclusions in calcite and found that larger
inclusion size and the internal pressure required to inclusions stretch more (i.e., the homogenization
stretch fluid inclusions in sphalerite, but were temperature increases by a larger amount) than
unable to accurately calculate the inclusion smaller inclusions subjected to the same P-T
volumes because the vapor bubbles were not history. Based on the results of Bodnar & Bethke
spherical at room temperature, precluding (1984), which indicate that the amount of
application of the technique described by Bodnar stretching (change in the homogenization
(1983). According to their data, the pressure temperature) for a group of inclusions that
necessary to stretch an inclusion smaller than initially had the same homogenization
about 5-10 µm in sphalerite is about 800 bars, temperature can be related to the temperature at
whereas an inclusion 50 µm in diameter requires which stretching begins, the results of
about 300-400 bars of internal pressure (Fig. 8-2). Prezbindowski & Larese (1987) would suggest
The relationship between inclusion size and the that the larger inclusions began to stretch before
pressure required to initiate stretching is less well the smaller inclusions. Goldstein (1986) observed
defined for inclusions in barite (Ulrich & Bodnar a similar size-pressure relationship for naturally
1988), although the trend indicates higher reequilibrated inclusions in low temperature
pressures for smaller inclusions, consistent with calcium carbonate cements. However, Hall et al.
other minerals studied. Prezbindowski & Tapp (1993) found no correlation between inclusion
(1991) showed that the size-pressure relationship size and ease of stretching during an experimental
observed in fluid inclusions is consistent with study of calcite, although they did notice that
theoretical models of crack propagation and stress inclusions closer to the mineral surface were more

217
prone to stretching compared to those deeper in HOW DO WE RECOGNIZE INCLUSIONS
the sample. THAT HAVE REEQUILIBRATED?
Often, one may be able to infer that
Inclusion shape reequilibration has occurred based on
A secondary factor that determines how petrographic observations of fluid inclusions and
easily a fluid inclusion will reequilibrate is the surrounding host mineral. If the fluid
inclusion shape. Bodnar et al. (1989) noted that inclusions within a fluid inclusion assemblage
the average pressure required to reequilibrate (FIA) satisfy "Roedder's Rules" described above
fluid inclusions in quartz increased by about 400 and have not reequilibrated, all of the inclusions
bars as the inclusion shape varied from very in the FIA will contain the same phases, and in
irregular, with many reentrants or "tentacles", to the same volume proportions, when observed at
one that was spherical or negative-crystal shaped. room temperature. Thus, an FIA composed of
Goldstein (1986) noted a similar dependence on two-phase (liquid + vapor) inclusions in which
inclusion shape for calcite, as did Bodnar & the liquid-to-vapor ratio is identical in all
Bethke (1984) for inclusions in fluorite and inclusions might indicate that reequilibration has
sphalerite. The relationship between inclusion not occurred. But, one must use care in
shape and the amount of pressure required to determining whether the inclusions have the same
initiate reequilibration is generally consistent with liquid-to-vapor ratio. For example, consider an
theoretical deformation models which indicate FIA containing inclusions with a salinity of 5 wt.
that stress is concentrated at sharp corners, and % NaCl and which homogenize at 150°C. If some
that the stress approaches infinity at crack tips of the inclusions reequilibrate by stretching such
(ends of tentacles in irregularly-shaped that the new homogenization temperature is
inclusions) or at the corners of angular inclusions 175°C, the volumes of the reequilibrated
(Prezbindowski & Tapp, 1991). However, the fact inclusions would have increased by 2.6 volume
that most angular inclusions do not spontaneously percent (see Bodnar & Bethke 1984, p. 153). If a
stretch or decrepitate during heating suggests that spherical vapor bubble in one such inclusion had
while theoretical models of deformation provide a diameter of 6 µm before stretching, the vapor
valuable insights into the reequilibration process, bubble diameter after stretching would be about
they may be less useful than empirical models for 6.6 µm! Inclusions with smaller bubbles would
predicting the reequilibration behavior of fluid show proportionately smaller increases in bubble
inclusions. diameter after stretching. It is unlikely that one
would be able to distinguish between two
Fluid Composition inclusions in the same FIA with vapor bubble
Little information is available diameters that differ by only 0.6 µm or less. Thus,
concerning the role of fluid composition in the an FIA could contain fluid inclusions, some of
reequilibration process. However, one would which had stretched by 25°C and others of which
expect that fluid inclusions containing water had not stretched (or stretched by a smaller
might reequilibrate more easily than those that are amount), and still appear to have consistent
water-free, owing to the presumed higher liquid-to-vapor ratios. However, this difference
solubilities of host minerals in aqueous fluids would become obvious during microthermometric
compared to, for instance, carbon dioxide. Hall & analysis of the inclusions, as described below.
Wheeler (1992) conducted a preliminary study of Finally, it should be emphasized that, whereas
the relationship between fluid composition and most reequilibration results in a "smearing" of
inclusion decrepitation and found that 3 µm data to produce a wide range in homogenization
diameter inclusions containing H2 O-NaCl temperatures, in some environments the reequilib-
required a lower pressure (3.0 kbars), and ration process may progress to the extent that the
inclusions containing H2 O-CO2 a higher pressure inclusions completely equilibrate at some new P-
(4.2 kbar), compared to pure H2 O inclusions (3.4 T condition, resulting in fluid inclusions with
kbars) of the same size. The role that fluid consistent phase ratios and microthermo metric
composition plays in reequilibration is an area of results, even though these are not representative
research that should be given high priority in of the original formation conditions (Vityk &
future studies. Bodnar 1995a, 1998).

218
Just as an FIA with real or apparently
consistent phase ratios is possible even if
inclusions have reequilibrated, a range in liquid-
to-vapor ratios is not always indicative of
reequilibration. If the inclusions were trapped in a
boiling (or immiscible fluid) system, a bimodal
distribution of phase ratios will result, with a
complete range between the end-member ratios
from trapping mixtures of the two phases present.
Fortunately, the range of possible phase ratios is
not entirely random, and is controlled by the
PVTX properties of the two immiscible fluids,
and the possible phase ratios that might result
from trapping mixtures may be calculated using
PVTX data (see Bodnar et al. 1985, their Fig.
5.5).
Textural features in the host mineral
surrounding fluid inclusions perhaps provide the
most diagnostic evidence for reequilibration.
When inclusions reequilibrate, the host phase FIG. 8-4. Relationship between the cooling/
surrounding the inclusions is deformed to produce unroofing path and the confining pressure and
a variety of features that are visible using internal pressure in a fluid inclusion trapped at
conventional optical microscopy or more sophist- temperature and pressure Tf, Pf.
icated techniques such as cathodoluminescence,
transmission electron microscopy (TEM), or X- special case that occurs only rarely in nature. In
ray tomography. Moreover, the textures that result most natural systems the rocks do not follow a P-
may be diagnostic of the P-T path followed dur- T path that approximates the inclusion isochore.
ing reequilibration. Most experimental studies, as Rather, the confining (geologic) pressure is either
well as studies of naturally reequilibrated inclus- greater than, or less than, the internal pressure in
ions have focused on the mineral quartz, and the the inclusions. In these cases the inclusion is not
discussion of fluid inclusion textures below is in equilibrium with the surrounding rock and
limited to this mineral. It is expected that petro- reequilibration may occur (see Diamond 2003,
graphic features observed for quartz may not Fig. 3-5).
always apply to other minerals that have signific- If the inclusion follows a P-T path that
antly different crystallographic and rheological results in internal overpressure, the first
properties or solubility in the inclusion fluid. modification that inclusions experience is a
Before discussing the various textural change in shape, often without a measurable
features associated with reequilibrated fluid change in volume. Specifically, the inclusions
inclusions in quartz, it is first necessary to become more regularly shaped, and elongated
consider the vario u s P-T histories under which inclusions become more rounded or equant (Fig.
fluid inclusions may be reequilibrated, because 8-5) (Gratier & Jenatton 1984, Pêcher & Boullier
each produces slightly different and, in some 1984, Bodnar et al. 1989, Bakker & Jansen 1991).
cases, diagnostic textural features. Following The evolution from less to more regular or
trapping, fluid inclusions may follow a P-T path negative-crystal shape represents the normal
that maintains an internal pressure in the inclusion maturation process that all fluid inclusions
equal to the confining (or lithostatic) pressure, or undergo as they try to minimize their surface area,
may follow paths such that the internal pressure is as discussed by Bodnar et al. (1985). Given
greater than or less than the confining pressure sufficient time, all inclusions should approach an
(Fig. 8-4). If the inclusion follows an isochoric equant or negative-crystal shape. Thus, the
path such that the confining pressure and the evolution from irregular to more-regular inclusion
internal pressure are always equal, reequilibration shape that occurs under conditions of internal
is unlikely to occur. This type of cooling path is a overpressure is not diagnostic of this type of

219
reequilibration, and the presence of equant or
negative-crystal shaped inclusions in a sample
does not necessarily indicate that the inclusions
have reequilibrated. With increasing departure of
the confining pressure from the inclusion
isochore, the largest inclusions begin to
reequilibrate as fractures are generated at sharp
corners of the inclusion and extend into the
surrounding host (Kotel'nikova 1994, Vityk et al.
1996) (Fig. 8-6). Some of the fluid that escapes
along these fractures may be trapped in the
surrounding quartz to produce a halo of very
small fluid inclusions adjacent to the fracture. As
the difference between the inclusion pressure and
the confining pressure increases further,
progressively smaller inclusions begin to
reequilibrate. Barker (1995) used the evolution in
shape of inclusions reequilibrated under
conditions of internal overpressure to identify
those inclusions that had experienced post-
entrapment modifications in the Caledonian Øse
Thrust, Norway. Detailed TEM studies by
Boullier et al. (1989), Bakker & Jansen (1994)
and Vityk et al. (2000) show an increased
dislocation density surrounding reequilibrated
inclusions in quartz (Fig. 8-7). This is associated
with diffusive loss of water into the quartz host,
which results in hydrolytic weakening of the
quartz and facilitates later fracturing and
additional water loss.
Fluid inclusions in quartz that
reequilibrate under conditions of internal
underpressure (confining pressure greater than the
internal pressure) show some distinct differences
compared to those reequilibrated under conditions
of internal overpressure. The first change
observed in these inclusions is that the inclusion
walls, which are initially smooth, become
corrugated and rough (Pêcher 1981, Sterner &
Bodnar 1989, Bakker & Jansen 1991) (Fig. 8-8).
Associated with this change is the development of
a halo of secondary inclusions surrounding the
FIG. 8-5. Evolution in the shape of fluid original inclusion. Bakker & Jansen (1991, 1994)
inclusions in quartz during reequilibration at and Vityk et al. (2000) suggested that fluid from
conditions of Pinternal > Pconfining. The the original inclusion (before reequilibration)
length/width ratio of the inclusion indicated by migrated into the surrounding quartz along
the arrow gradually decreases with time to dislocations and was later trapped to form the
produce an equant negative-crystal-shaped fluid secondary inclusion halo. Compared to conditions
inclusion after being held at a temperature of of internal overpressure, in which the largest
370°C and an internal pressure of 400 bars for inclusions begin to reequilibrate first (at the
5621 hours. The scale bar equals 50 µm. lowest amount of overpressure), the smallest
(modified from Bodnar et al., 1989). inclusions reequilibrate most easily under

220
FIG. 8-6. Relationship between the intensity of
reequilibration features and the size and pressure
differential (P internal - Pconfining) for methane
inclusions in quartz from the Ukrainian Carpathian
Mountains. At low pressure differential (top row)
none of the inclusions show reequilibration
textures. With increasing departure of the
confining pressure and internal pressure in the
inclusions, the largest inclusions begin to
reequilibrate, while the smaller inclusions show no
reequilibration textures. Progressively smaller
inclusions begin to reequilibrate as the pressure
differential increases. All inclusions in samples
with the largest pressure differential (bottom row)
show evidence of fracturing and development of
haloes of secondary inclusions surrounding the
original inclusion. (from Vityket al., 1996)

FIG. 8-7. TEM photomicrographs showing microstructures produced during reequilibration at 625°C and 2
kbar. The inclusions were originally formed at 700°C and 5 kbar, resulting in an internal overpressure
(Pinternal - Pconfining) of about 2.1 kbar at the reequilibration conditions. Fluid inclusions in "A" and "B"
are surrounded by poorly healed fractures (arrow 1), a few dislocations connected to the inclusion walls
(arrows labeled 2), and a Dauphine twin (arrow #4 in "B"). Note the high density of water bubbles in the
healed fracture in "A". Also note that some dislocations are pinned with bubbles (arrows #3 in "A"). (from
Vityk et al., 2000).

221
FIG. 8-8. Evolution of fluid
inclusions in quartz after
reequilibration at con-
ditions of Pinternal <
Pconfining. The inclusions
were originally formed at
800°C and 1.35 kbars, and
then reequilibrated at
700°C and 4 kbars,
producing a pressure
differential (Pconfining -
Pinternal) of about 2.75
kbars. Note the develop-
ment of corroded walls and
the formation of a halo of
secondary inclusions in the
surrounding quartz. The
scale bar in A, C and E
represents 25 µm. (from
Sterner et al. 1989).

conditions of internal underpressure (Fig. 8-9) range from 282-285°C. Another sample was
(Hurai & Horn 1992, Vityk et al. 1994). Careful formed at the same conditions and then cooled to
petrographic observations to determine whether 625°C and 2 kbar and held at these conditions for
the smallest or the largest inclusions in a given 30 days, resulting in an initial internal
sample show the most intense reequilibration overpressure in the inclusions of about 2.1 kbars.
features may provide a means of determining After 30 days the sample was quenched and the
whether the sample experienced an initially homogenization temperatures of the inclusions
isobaric cooling path, or an initially isothermal were measured. Not only had the temperatures
unroofing path, following formation. been shifted to higher values compared to the
original Th, but also the original three degree
Microthermometric Evidence Celsius range in Th (282-285°C) had expanded to
As the inclusion volume or composition about 65 degrees Celsius (˜ 310-375°C).
changes, the microthermometric behavior of The relationship between the range in Th
inclusions will change in a somewhat predictable and extent of stretching observed in laboratory
manner. An increase in volume without fluid loss studies may not be applicable to reequilibration in
results in a lower density and concomitant nature, as noted in the introduction to this chapter.
increase in the homogenization temperature. While it is true that stretching results in a wide
Based on numerous laboratory studies, when a range in Th during laboratory reequilibration
group of inclusions with initially uniform studies, if these same samples were held at the
homogenization temperatures is overheated by a reequilibration conditions for durations
certain amount, the amount of stretching (percent approaching geologic time scales (102 - 106
volume increase) of the inclusions in the FIA is years), it is likely that the homogenization
highly variable, resulting in a wide range of temperatures of all the inclusions would approach
homogenization temperatures. For example, Vityk some new "equilibrium" homogenization
& Bodnar (1998) reported that 160 fluid temperature. Results presented by Vityk &
inclusions in quartz that were trapped at 700°C
and 5 kbars all homogenized within a three degree

222
similar cements that had formed at the surface but
were later buried to significant depths and
exposed to temperatures well above 100°C. The
recent samples that had never been buried
contained mostly all-liquid inclusions, in addition
to liquid + vapor inclusions that had trapped water
plus air. Deeply buried samples showed a wide
range in homogenization temperatures and
salinities, reflecting continued reequilibration
during burial and exposure to fluids of different
salinities as burial progressed. Goldstein (1986)
concluded that the deeply buried cements were
unlikely to provide any useful information
concerning original formation conditions, but
could prove useful for interpreting burial pore
fluid history. Barker & Goldstein (1990) used the
fact that inclusions in calcite reequilibrate
relatively easily to suggest that inclusions in
calcite from sedimentary basins and geothermal
fields might be used to estimate the maximum
burial conditions.
One misconception that many workers
have concerns the reproducibility of
microthermometric data as a means of testing
whether or not reequilibration has occurred.
Workers often report that homogenization
temperatures of inclusions are reproducible when
measured multiple times. These workers then
interpret this to indicate that the inclusions have
not reequilibrated. However, a reproducible Th is
necessary but not sufficient to prove that an
inclusion has not reequilibrated. As noted by
FIG. 8-9. Relationship between the intensity of Bodnar & Bethke (1984), stretched fluid
reequilibration features and the size and pressure inclusions in fluorite gave reproducible (±0.2°C)
differential (Pconfining - Pinternal) for aqueous homogenization temperatures, even though these
inclusions in quartz from the Bayanovoo granite temperatures were ˜40°C higher than the Th
in central Mongolia. The inclusions were before the inclusions were stretched by laboratory
reequilibrated at 500°C and pressures of 1 (A-C), heating. Larson et al. (1973) also noted that some
2 (D-F), 3 (G-I), 4 (J-L) and 5 (M-O) kbar for 7 inclusions in fluorite gave reproducible Th after
days. At run conditions, all of the inclusions had four heating runs, even though some of these
an internal pressure of about 500 bars. With temperatures were as much as 70 degrees Celsius
increasing pressure differential the smaller too high (i.e., the inclusions had been stretched by
inclusions begin to reequilibrate first, followed 70 degrees Celsius). Thus, reproducibility of
by increasingly larger inclusions as the pressure homogenization temperatures is a necessary
differential increased from about 500 bars (A-C) criterion to confirm that inclusions have not
to 4.5 kbar (M -O). (from Vityk et al. 1994). reequilibrated, but is not in itself sufficient to
Bodnar (1995a, 1998) are consistent with this prove that reequilibration has not occurred.
suggestion.
Goldstein (1986) examined fluid Compositional Reequilibration
inclusions from recent calcium carbonate cements Compared to volumetric reequilibration,
that had formed at the surface and were never compositional reequilibration is less well
deeply buried. He compared these to inclusions in understood and more difficult to recognize. Most

223
compositional reequilibration involves the loss (or into the inclusions during uplift and cooling. The
gain) of hydrogen or water (OH-) from the extremely negative hydrogen isotopic (D/H)
inclusion (Hall & Sterner, 1995). In melt composition of the fluid inclusions was consistent
inclusion studies, other components, including with enrichment in the hydrogen isotope, as the
sodium and some trace elements, can be lost from larger deuterium diffuses through quartz at a
inclusions during heating. Melt inclusion much slower rate. Hall & Bodnar (1990)
reequilibration is not considered here. suggested that the small amount of methane often
Pure carbon dioxide (or carbon dioxide- reported in fluid inclusions from granulites is the
rich) inclusions are common in medium- to high- result of hydrogen diffusion into the inclusions
grade metamorphic environments. Based on following trapping, consistent with
thermodynamic considerations, water-free (or thermodynamic properties in the C-O-H system.
water-poor) fluids often are not in equilibrium Hall & Sterner (1993) conducted
with the mineral assemblage in the rocks, and this experimental studies of water loss from aqueous
has led many workers to invoke post-trapping fluid inclusions. As a result of water loss, the
reequilibration of the inclusions to explain the salinity of the inclusions increased by 22 %.
high CO2 contents. Hollister (1990) has Vityk et al. (2000) conducted similar experiments
suggested that CO2 inclusions in metamorphic and were able to increase the salinity by a few
rocks are produced when water leaks out of tenths of a weight percent NaCl as a result of
originally water-rich inclusions during plastic water loss. Audétat & Günter (1999) noted
deformation, to produce pure CO2 , or CO2 -rich, characteristic reequilibration textures associated
inclusions. Bakker & Jansen (1991) were able to with fluid inclusions in quartz from the Mole
reproduce this process experimentally. Water was Granite, Australia, and attributed variable and
lost from H2 O-CO2 inclusions during high salinities in the inclusions to loss of water to
reequilibration, resulting in inclusions more CO 2 - produce "sweat haloes" around deformed
rich than the original inclusions. Alternatively, inclusions. It is likely that ranges in salinities of
Lamb et al. (1987) and Lamb (1990) suggested apparently coeval inclusions in some magmatic
that the "primary appearing" CO2 inclusions in hydrothermal ore deposits, such as the porphyry
granulites were actually trapped during copper deposits, are due at least in part to water
retrogression and are not associated with peak loss from the inclusions following entrapment.
granulite grade conditions. One could interpret the discussion above
Owing to the small size of the hydrogen to indicate that inclusions reequilibrate by an
ion, hydrogen is able to diffuse through most increase in volume with no fluid loss (stretching)
mineral structures very easily. Mavrogenes & or by loss of fluid through leakage or
Bodnar (1994) showed experimentally that fluid decrepitation. In practice, these various
inclusions several tens of micrometers beneath the mechanisms of reequilibration likely operate
surface of a quartz crystal will reequilibrate to the together, with perhaps one or the other
external hydrogen fugacity in only a few hundred dominating, depending on the environment in
hours at 600-800°C, or in only a few tens of which reequilibration is occurring. Thus, Vityk et
thousands of hours (several years) at temperatures al. (2000) note that stretching of fluid inclusions
of 400°C. As part of this work, these workers in quartz requires the diffusive loss of fluid from
were able to confirm that the common failure of the inclusion into the surrounding quartz host to
sulfide daughter minerals to dissolve during initiate the stretching process. However, normal
heating is the result of post-entrapment hydrogen petrographic observations would not reveal the
loss from the inclusions. These workers were able presence of water-bearing dislocations, and one
to reverse this process experimentally and might assume that the change in homogenization
dissolve chalcopyrite daughter minerals - these temperature was due solely to volume change.
phases did not dissolve before hydrogen was Based on our experience and observations, we
experimentally diffused into the inclusion at high believe that in all, or certainly most, examples of
temperature and pressure. Hall et al. (1991) reequilibration, both volume change and fluid loss
attributed the presence of methane in fluid (or gain) occur. Possible exceptions might
inclusions from the Ducktown, Tennessee, USA, include, for example, the reequilibration of "pure"
massive sulfide deposits to hydrogen diffusion CO2 inclusions. Wanamaker & Evans (1989)

224
noted that following heating to generate high
internal pressures, the density of CO2 inclusions
in olivine had decreased. Assuming that CO2
cannot diffuse through olivine at these conditions,
these workers attributed the density decrease to
stretching (volume increase), in agreement with
their measurements of inclusion diameters before
and after heating.

Necking Down
Most fluid inclusions that we observe
today do not have the same shape as when the
fluid was originally trapped in the mineral.
Following entrapment in the host phase, most
inclusions change shape through a process
referred to as "necking down" that involves
dissolution and re-precipitation of the host phase
to reduce the surface area and produce many
smaller inclusions from the original larger
inclusion. For example, when a mineral fractures
during or after growth, some fluid may enter the
fracture as shown in Figure 8-10a. As necking
down proceeds, the fluid in the fracture becomes
isolated into progressively more and smaller fluid
inclusions. During this process, the inclusion
shape tends to become more regular. This process
is often referred to as maturation in the fluid
inclusion literature (cf. Bodnar et al., 1985). If
this process goes to completion, the former
fracture plane will be decorated by numerous
spherical or negative-crystal shaped fluid
inclusions, often showing a decrease in inclusion
size towards the original fracture tip (Fig. 8-10d).
If necking down occurs at the
temperature and pressure of trapping of the Fig. 8-10. Schematic representation of the
original fluid, the resulting inclusions will all formation of secondary fluid inclusions in a
show similar phase relations as well as fracture by necking down. Fluid enters the open
microthermometric behavior that is consistent fracture (a) and begins to dissolve material from
with the PTX trapping conditions. Indeed, the the walls and re -precipitate new mineral matter
process described above and illustrated in Figure to heal the fracture, isolating small pockets of
8-10 is identical to that used to trap synthetic fluid fluid in the process (b). With continued healing
inclusions according to the technique of Sterner & the originally irregularly shaped fluid inclusions
Bodnar (1984). Accordingly, fractured quartz (c) become more regular in shape and evolve
cores are placed into high temperature autoclaves through maturation to produce a plane decorated
along with a fluid of known composition. During by numerous negative-crystal shaped fluid
the constant-temperature experiment, fluid enters inclusions (d). (Modified from Roedder, 1962).
the fracture and the fracture heals as shown in down is simply the process by which inclusions
Figure 8-10 to produce numerous synthetic fluid mature and does not involve reequilibration of the
inclusions trapped at constant and known PTX type discussed in this chapter.
conditions. The microthermometric behavior of Inclusions that undergo necking down at
the resulting inclusions is consistent with the constant temperature and pressure, and in the one-
known formation conditions. In this case, necking phase fluid field, are of no concern in discussions

225
of reequilibration processes that lead to incorrect ature and/or pressure system, or in the presence of
microthermometric results. However, if the two or more phases, the fluid inclusions will show
temperature and/or pressure change during the variable phase relations and a range in
reequilibration process, or if two or more fluids ± homogenization temperatures. Such inclusions
solids are present in the fracture during necking should be avoided during microthermometric
down, the resulting inclusions likely will provide studies if the goal is to determine the original P-T
incorrect compositional and microthermometric formation conditions.
data. Figure 8-11 is a schematic representation of
necking down in a decreasing temp erature system REEQUILIBRATION OF MELT
in which phase separation to produce a vapor INCLUSIONS
bubble occurs during cooling. The original Compared to fluid inclusions,
inclusion at time T0 contains only liquid. During compositional reequilibration of melt inclusions is
cooling to temperature T1 a vapor bubble much more common. Anderson (1974) was one of
nucleates in the inclusion. During cooling from the first to consider the reliability of melt
temperature T1 to T2 , the original inclusion necks inclusions and the various post-entrapment
down to produce two inclusions. One inclusion changes that might lead to reequilibration.
(inclusion a) contains the larger vapor bubble that Nielsen et al. (1998) address the relative
formed during initial cooling, and the other advantages of microscope heating versus batch
inclusion contains only a small vapor bubble that heating of mineral grains to homogenize melt
formed after the two inclusions were separated. inclusions, and also discuss volatile loss and
With continued cooling, the larger inclusion contamination of melt inclusions from alteration
undergoes necking down again to produce two products. Manley (1996) discussed the use of melt
smaller inclusions, one containing the bubble that inclusion morphology as an indicator of the
was present at the moment necking occurred timing of inclusion formation during the eruption
(inclusion b) and the other (inclusion c) event. Qin et al. (1992) developed a general
containing a smaller bubble that nucleated after theoretical model for diffusive reequilibration of
necking during cooling from T2 t o T3 . At room melt inclusions, and applied this model to
temperature (T4 ) the three inclusions show very evaluate water loss from melt inclusions
different liquid to vapor ratios, and would following entrapment, as well as reequilibration
homogenize at different temperatures. Thus, if an of trace element concentrations in melt inclusions.
FIA is comprised of inclusions that have Gaetani & Watson (2000) evaluated Fe-Mg
undergone necking down in a changing temper- exchange between melt inclusions and host
olivine, and Danyushevsky et al. (2000) provide
data from natural melt inclusions in olivine that
have reequilibrated by Fe-loss.
A complete consideration of the melt
inclusion reequilibration process is beyond the
scope of this chapter. The reader is referred to the
sources listed above, as well as references therein,
for a more complete discussion of this problem.

HOW TO AVOID LABORATORY RE-


EQUILIBRATION OF FLUID INCLUSIONS
For most fluid inclusions, compositional
reequilibration during laboratory studies is
unlikely if care is taken during data collection.
Fig 8-11. Necking down of an originally large, Exceptions include hydrogen loss (or gain) by the
liquid -filled inclusion (T0 ) to produce three inclusions during microthermometry and
smaller inclusions. None of these inclusions compositional reequilibration of melt inclusions
would record the correct formation temperature. during heating. The main type of reequilibration
See text for details. (Modified from Roedder, that one has to be concerned with during
1962). laboratory studies of most fluid inclusions is

226
volumetric reequilibration. Fluid inclusions tend 400°C. For higher homogenization temperatures,
to reequilibrate volumetrically when the internal the internal pressure can become sufficiently high
pressure exceeds the strength of the host mineral to reequilibrate inclusions in even some of the
and, as noted above, this value depends on the more resistant minerals, such as quartz. For
mineral hardness as well as the inclusion size and example, a fluid inclusion containing a 25 wt.%
shape and fluid composition. In most cases, NaCl solution will have an internal pressure of
heating to some temperature above room about 900 bars at a homogenization temperature
temperature (and above the homogenization of 600°C (Bodnar & Vityk, 1994). According to
temperature) is required to generate pressures equation (1) above, this pressure is sufficient to
sufficiently high to cause most inclusions to cause decrepitation of inclusions in quartz larger
reequilibrate. Low temperature (Th =150°C) than about 25 micrometers (Bodnar et al., 1989).
aqueous inclusions may also reequilibrate during Many who study fluid inclusions from
cooling studies to measure ice-melting medium- to high-grade metamorphic environ-
temperatures. In this case, formation of ice during ments have experienced inclusion decrepitation
cooling, which occupies approximately 9 % more before homogenization occurs. This happens
volume than the liquid water it formed from, can because most inclusions from this environment
generate high pressures and cause the inclusion to contain significant amounts of carbon dioxide,
stretch. Lawler & Crawford (1983) coined the resulting in elevated pressure in the inclusions,
term "freeze stretching" to describe this even at room temperature. As an example, an
phenomenon. H2 O-CO2 inclusion containing 37 mole percent
In the early days of fluid inclusion CO2 and homogenizing at 263°C would have an
studies, most polished sections were prepared internal pressure of 2880 bars at homogenization
using Canada balsam or Lakeside cement or some (Sterner & Bodnar, 1991). This pressure would
other adhesive that required the sample to be cause decrepitation of all inclusions larger than
heated to temperatures approaching 100°C. In about 3 µm in quartz (equation (1) above), and
many cases, little care was taken to control this any inclusions larger than this would decrepitate
temperature and samples may have been heated to during heating, before the homogenization
several hundred degrees to mount the sample on a temperature is reached. Schmidt et al. (1999)
glass slide. Such high temperature would cause described application of the Hydrothermal
fluid inclusions to stretch before they were placed Diamond Anvil Cell as a pressurized fluid
into the fluid inclusion stage, resulting in inclusion stage to eliminate (or minimize)
erroneous microthermometric data. Today, most decrepitation of high pressure inclusions during
workers take care to minimize heating of the heating to homogenization.
sample during polishing, and often use adhesives In some cases it may simply not be
such as Superglue that sets at room temperature possible to heat (or cool) a fluid inclusion without
and requires no heating. In some extreme cases causing reequilibration to occur. Not every fluid
involving very low temperature inclusions, the inclusion is able to provide useful
samples must be stored in a temperature- microthermometric data. What is most important
controlled room and transported in ice to avoid to the fluid inclusionist is to be able to recognize
stretching of the inclusions (c f. Wilson et al., inclusions that reequilibrate during laboratory
2003). studies. This can best be accomplished by
Most workers believe that reequilibration following one simple rule during microthermo -
in the laboratory is not a problem as long as the metric analysis: Thou shalt collect no data from
inclusions are not heated above the any fluid inclusion that has not been observed
homogenization temperature. This is because the continuously during the very first heating (or
rate of pressure increase in the inclusion during cooling) run. If one obeys this rule, the
heating is much lower before the inclusion has probability of collecting data from inclusions that
homogenized than it is after homogenization have reequilibrated during laboratory analysis is
occurs. And, for gas-free aqueous inclusions in significantly reduced.
most minerals, the internal pressure is unlikely to
exceed a few hundred bars during heating, as long
as the homogenization temperature is below about

227
ACKNOWLEDGEMENTS BODNAR, R.J. (1983): A method of calculating
Much of the information presented in this fluid inclusion volumes based on vapor bubble
chapter represents studies by former students and diameters and P-V-T-X properties of inclusion
post-doctoral researchers and visitors to the Fluids fluids. Econ Geol, 78, 535- 542.
Research Laboratory. I wish to acknowledge Don
BODNAR, R.J. & BETHKE , P.M. (1984):
Hall, Mike Sterner and Max Vityk for their many
Systematics of stretching of fluid inclusions. I.
contributions that helped me to better understand
Fluorite and sphalerite at 1 atmosphere
the reequilibration process. I also give special
thanks to Phil Bethke and Ed Roedder for confining pressure. Econ. Geol, 79, 141-161.
encouraging me as a student to pursue this BODNAR, R.J., BINNS, P.R. & HALL , D.L. (1989):
important area of fluid inclusion research. Alan Synthetic fluid inclusions. VI. Quantitative
Anderson, Will Lamb , Iain Samson and one evaluation of the decrepitation behavior of fluid
anonymous reviewer are thanked for their inclusion in quartz at one atmosphere confining
comments and suggestions for improving the pressure. J. Meta. Geol. 7, 229-242.
manuscript. The National Science Foundation,
BODNAR, R.J., REYNOLDS, T.J. & KUEHN, C.A.
Department of Energy, NASA, and the American
(1985): Fluid inclusion systematics in
Chemical Society have supported work in the
Fluids Research Laboratory over the years. NSF epithermal systems. in Society of Economic
Geologists, Reviews in Economic Geology, 2,
Grants EAR-0001168 and EAR-0125918
Geology and Ge ochemistry of Epithermal
provided support during preparation of this
manuscript. Systems, B.R. Berger and P.M. Bethke, eds.,
73-98.
REFERENCES BODNAR, R.J. & VITYK , M.O. (1994)
A NDERSON, A.T. (1974): Evidence for a picritic, Interpretation of microthermometric data for
volatile-rich magma beneath Mt. Shasta, NaCl-H2O fluid inclusions. In De Vivo B,
Frezzotti ML (eds) Fluid inclusions in minerals:
California. J. Petrol. 15, 243 -267.
methods and applications. Published by Virginia
A UDETAT , A. & GUNTER, D. (1999): Mobility and Polytechnic Institute & State University,
H2O loss from fluid inclusions in natural quartz Blacksburg, VA, 117-131.
crystals. Contrib Mineral Petrol, 137, 1-14.
BOULLIER, A.M., M ICHOT , G., PÊCHER, A. &
BAKKER, R.J. & JANSEN , J.B.H. (1991): BARRES, O. (1989): Diffusion and plastic
Experimental post-entrapment water loss from deformation around fluid inclusions in synthetic
synthetic CO2 - H2O inclusions in natural quartz. In Bridgwater D (ed) Fluid movements -
quartz. Geochim. Cosmochim. Acta, 55, 2215- element transport and the composition of the
2230. deep crust. NATO ASI Series 281, Kluwer,
Dordrecht, 345-360 pp.
BAKKER, R.J. & JANSEN J.B.H. (1994): A
mechanism for preferential H2O leakage from DANYUSHEVSKY, L.V., DELLA-PASQUA, F.N. &
fluid inclusions in quartz, based on TEM SOKOLOV, S. (2000): Reequilibration of melt
observations. Contrib Mineral Petrol, 116, 7-20 inclusions trapped by magnesian olivine
phenocrysts from subduction-related magmas:
BARKER, A.J. (1995): Post-entrapment
petrological implications. Contrib. Mineral.
modification of fluid inclusions due to
Petrol. 138, 68-83.
overpressure: evidence from natural samples. J
Meta Geol, 13, 737-750. DIAMOND, L.W. (2003): Systematics of H2 O
inclusions. In I. Samson, A. Anderson, & D.
BARKER C.E. & GOLDSTEIN, R.H. (1990): Fluid
Marshall, eds. Fluid Inclusions: Analysis and
inclusion technique for determining the
Interpretation. Mineral. Assoc. Can., Short
maximum temperature in calcite and its
Course Ser. 32, 55-79.
comparison to the vitrinite reflectance
geothermometer. Geology, 18, 1003-1006. FAIZIEV, A.R. & ALIDOVOV, B.A. (1976):
Thermometry of minerals with perfect cleavage.
Soviet Geology & Geophysics, 17, 118-121.

228
GAETANI , G.A. & WATSON, E.B. (2000) Open HALL, D.L. & W HEELER, J.R. (1992): Fluid
system behavior of olivine-hoste d melt composition and the decrepitation behavior of
inclusions. Earth & Planetary Sci. Letters 183, synthetic fluid inclusions in quartz. Pan
27-41. American Conference on Research on Fluid
Inclusions, Program and Abstracts, 39.
GOLDSTEIN, R.H. (1986): Reequilibration of fluid
inclusions in low-temperature calcium HEINRICH, W. & GOTTSHALK, M. (1995):
carbonate cement. Geology, 14, 792-795. Metamorphic reactions between fluid inclusions
and mineral hosts. I. Progress of the reaction
GRATIER, J.P. & JENATTON, L. (1984):
calcite + quartz = wollastonite + CO2 in natural
Deformation by solution-deposition and
wollastonite-hosted fluid inclusions: Contrib
reequilibration of fluid inclusions in crystals
Mineral Petrol. 122, 51-61.
depending on temperature, internal pressure,
and stress. J Structural Geol, 6, 189-200. HOLLISTER, L.S. (1990): Enrichment of CO2 in
fluid inclusions in quartz by removal of H2O
GUILHAUMOU , N., COUTY, R. & DAHAN, N.
during crystal-plastic deformation. J Structural
(1987): Deformation of fluid inclusions in
Geol. 12, 895-901.
fluorite under confining pressure. Chem Geol,
61, 47-53. HURAI , V & HORN, E.E. (1992): A boundary
layer-induced immiscibility in naturally
GUILHAUMOU , N., TOURAY, J.-C. & BOUHLEL, S. reequilibrated H2O-CO2-NaCl inclusions from
(1988): Stretching of hydrocarbon fluid
metamorphic quartz (Western Carpathians,
inclusions in fluorite at 200 and 400 bars Czechoslovakia). Contrib Mineral Petrol. 112,
confining pressure. Application to low-pressure
414-427.
barometry. Bull. Miner. 111, 421-426.
KLEINEFELD , B. & BAKKER, R.J. (2002): Fluid
HALL , D.L. & BODNAR , R.J. (1989): Comparison
inclusions as microchemical systems: evidence
of fluid inclusion decrepitation and acoustic
and modelling of fluid-host interactions in
emission profiles of Westerley granite and
plagioclase. J. Meta. Geol. 20, 845-858.
Sioux quartzite. Tectonophysics 168, 283-296.
KODERA, P., MURPHY, P.J. & RANKIN, A.H.
HALL , D.L. & BODNAR , R.J. (1990): Methane in
(2003): retrograde mineral reactions in saline
fluid inclusions from granulites: a product of
fluid inclusions: The transformation
hydrogen diffusion? Geochim Cosmochim Acta
ferropyrosmalite clinopyroxene. Am. Mineral.
54, 641-652.
88, 151-158.
HALL , D.L., BODNAR, R.J. & CRAIG, J.R. (1991):
KOTEL 'NIKOVA , Z.A. (1994): The response of
Evidence for postentrapment diffusion of hydro-
fluid inclusions to changes in physicochemical
gen into peak metamorphic fluid inclusions parameters in the external medium. Geokhimia
from the massive sulfide deposits at Ducktown,
4: 476-485 (in Russian).
Tennessee. Am. Mineral. 76, 1344-1355.
KÜSTER, M. & STÖCKHERT , B. (1997): Density
HALL , D.L. & STERNER, S.M. (1993): Preferential
changes of fluid inclusions in high-pressure low-
water loss from synthetic fluid inclusions:
temperature metamorphic rocks from Crete: A
Contrib Mineral Petrol. 114 , 489-500.
thermobarometric approach based on the creep
HALL , D.L. & STERNER, S.M. (1995): strength of the host minerals. Lithos. 41, 151-
Experimental diffusion of hydrogen into 167.
synthetic fluid inclusions in quartz. J. Meta.
LAMB, W.M. (1990): Fluid inclusions in
Geol. 13, 345-355.
granulites: Peak vs. retrograde formation. in D.
HALL , D.L., STERNER, S.M. & W HEELER, J.R. Vielzeuf and Ph. Vidal (eds.), Granulites and
(1993): One-atmosphere decrepitation behavior Crustal Evolution, Kluwer Academic Press,
of synthetic fluid inclusions in natural calcite: Netherlands, 419-433.
Implications for the preservation of calcite-
hosted inclusions during burial. Archiwum
Mineralogiczne, 49, 91-92.

229
LAMB, W.M., VALLEY , J.W. & BROWN, P.E. PREZBINDOWSKI, D.R. & TAPP, J.B. (1991)
(1987): Post-metamorphic CO 2 -rich inclusions in Dynamics of fluid inclusion alteration in
granulites. Contrib Mineral Petrol. 96, 485-495. sedimentary rocks: a review and discussion.
Org, Geochem. 17, 131 -142.
LARSON, L.T., M ILLER, J.D., NADEAU, J.E. &
ROEDDER, E. (1973): Two sources of error in QIN , Z., LU, F. & A NDERSON A.T., JR. (1992)
low temperature inclusion homogenization Diffusive reequilibration of melt and fluid
determination and corrections on published inclusions. Am. Mineral. 77, 565-576.
temperatures for the East Tennessee and
ROEDDER, E. (1962): Ancient fluids in crystals.
Laisvall deposits. Econ. Geol., 68, 113-116.
Scientific American. 207, no. 4, 38-47
LAWLER, J.P. & CRAWFORD, M.L. (1983):
ROEDDER, E. (1984): Fluid Inclusions, Min. Soc.
Stretching of fluid inclusions resulting from a
low temperature microthermometric technique. Am. Rev. in Min. v. 12, 646 pp.
Econ. Geol., 78, 527-529. SCHMIDT , C., CHOU, I-M., BODNAR, R.J. &
BASSETT, W.A. (1998): Microthermometric
LEROY , J. (1979): Contribution à l'étalonnage de
analysis of synthetic fluid inclusions in the
la pression interne des inclusions fluides lors de
hydrothermal diamond-anvil cell. Am. Min., 83,
leur décrépitation. Bull. Mineral. 102, 584-593.
995-1007.
M ANLEY, C.R. (1996): Morphology and
STERNER, S.M. & BODNAR, R.J. (1984): Synthetic
maturation of melt inclusions in quartz
fluid inclusions in natural quartz. I.
phenocrysts from the Badlands rhyolite lava
Compositional types synthesized and
flow, southwestern Idaho. Am. Miner. 81, 158-
applications to experimental geochemistry.
168.
Geochim Cosmochim Acta 48, 2659-2668.
M AVROGENES, J.A. & BODNAR, R.J. (1994):
Hydrogen movement into and out of fluid STERNER, S.M. & BODNAR, R.J. (1989): Synthetic
fluid inclusions. VII. Reequilibration of fluid
inclusions in quartz: experimental evidence and
inclusions in quartz during laboratory simulated
geologic implications. Geochim Cosmochim
metamorphic burial and uplift: J Meta Geol 7,
Acta 58, 141-148.
243-260.
NAUMOV, V.B., BALITSKIY, V.S. & KHETCHIKOV,
STERNER, S.M. & BODNAR, R.J. (1991): Synthetic
L.N. (1966) Correlation of the temperatures of
formation, homogenization and decrepitation of fluid inclusions. X. Experimental determination
of P-V-T-X properties in the CO2-H2O system
gas-fluid inclusions. Doklady Acad. Sci. USSR
to 6 kb and 700°C. Am. J. Sci. 291, 1-54.
171, 146-148.
STERNER. S.M., HALL , D.L. & KEPPLER, H.
NIELSEN, R.L., MICHAEL , P.J. & SOURS-PAGE, R.
(1995): Compositional re-equilibration of fluid
(1998): Chemical and physical indicators of
inclusions in quartz. Contrib Mineral Petrol.
compromised melt inclusions. Geochim
119, 1-15.
Cosmochim Acta 62, 831-839.
PÊCHER, A. (1981): Experimental decrepitation SWANENBURG, H.E.C. (1980): Fluid inclusions in
high-grade metamorphic rocks from S.W.
and re-equilibration of fluid inclusions in
Norway. Unpub. Ph.D. Thesis, Geologica
synthetic quartz. Tectonophysics 78, 567-583.
Ultraiectina, University of Utrecht, The
PÊCHER, A. & BOULLIER, A.-M. (1984): Netherlands.
Évolution à pression et température élevées
TUGARINOV, A.I. & NAUMOV, V.B. (1970):
d'inclusions fluides dans un quartz synthétique.
Bull. Mineral. 107, 139-153. Dependence of the decrepitation temperature of
minerals on the composition of their gas-liquid
PREZBINDOWSKI, D.R. & LARESE, R.E. (1987): inclusions and hardness. Aksd. Nauk SSSR
Experimental stretching of fluid inclusions in Doklady, 195, 112-114.
calcite - Implications for diagenetic studies.
ULRICH, M.R. & BODNAR, R.J. (1988):
Geology 15, 333-336.
Systematics of stretching fluid inclusions. II.

230
Barite at one atmosphere confining pressure. VITYK , M.O., BODNAR , R.J. & DUDOK, I.V.
Econ. Geol. 83, 1037-1046. (1996): Fluid inclusions in "Marmarosh
Diamonds": Evidence for tectonic history of the
VITYK , M.O. & BODNAR, R.J. (1995a) Do fluid
Folded Carpathian Mountains, Ukraince.
inclusions in high grade metamorphic terranes Tectonophysics 255, 163-174.
preserve peak metamorphic density during
retrograde decompression? Am. Mineral. 80, VITYK , M.O., BODNAR , R.J. & SCHMIDT , C.S.
641-644. (1994): Fluid inclusions as tectonothermobaro-
meters: relation between pressure-temperature
VITYK , M.O. & BODNAR, R.J. (1995b): Textural
history and reequilibration morphology during
evolution of synthetic fluid inclusions in quartz
crustal thickening. Geology 22, 731-734.
during re -equilibration, with applications to
tectonic reconstruction. Contrib Mineral Petrol. WANAMAKER, B.J. & EVANS, B. (1989):
121, 309-323. Mechanical re-equilibration of fluid inclusions in
San Carlos olivine by power law creep. Contrib
VITYK , M.O. & BODNAR, R.J. (1998): Statistical
Mineral Petrol . 102, 102-111.
microthermometry of synthetic fluid inclusions
in quartz during decompression. Contrib W ILKINS, R.W.T. (1986): The mechanisms of
Mineral Petrol 132, 149-162. stretching and leaking of fluid inclusions in
fluorite. Econ. Geol., 81, 1003-1008.
VITYK , M.O., BODNAR , R.J. & DOUKHAN, J-C.
(2000): Synthetic fluid inclusions: XV. TEM W ILKINS, R.W.T., BIRD , J.R. & EWALD , A.H.
investigation of plastic flow associated with re- (1981): Observations on deformation
equilibration of synthetic fluid inclusions in microstructures and fluid inclusions in proton-
natural quartz. Contrib Mineral Petrol 139, 285- irradiated halite. N. Jb. Miner. Abh. 141, 240-
297. 257.
VITYK , M.O., BODNAR, R.J. & DUDOK, I.V. W ILSON, N.S.F., CLINE , J.S. & A MELIN , Y.V.
(1995): Natural and synthetic reequilibration (2003) Origin, timing and temperature of
textures of fluid inclusions in "Marmarosh secondary calcite-silica mineral formation at
Diamonds": Evidence of refilling under yucca mountain, Nevada. Geochim Cosmochim
conditions of compressive loading. Europ J Acta 67, 1145-1176.
Mineral. 7, 1071-1087.

231

You might also like