You are on page 1of 9

Physics and Chemistry of the Earth 30 (2005) 1038–1046

www.elsevier.com/locate/pce

The use and usefulness of boron isotopes in natural


silicate–water systems
a,* b
Annette Deyhle , Achim J. Kopf
a
Scripps Institution of Oceanography, UCSD, 9500 Gilman Drive, La Jolla, CA 92093-0212, USA
b
Research Centre Ocean Margins, University Bremen, Klagenfurter Strasse, 28359 Bremen, Germany

Received 25 March 2004; received in revised form 13 April 2005; accepted 19 April 2005
Available online 8 August 2005

Abstract

Boron and its two stable isotopes 10B and 11B are considered powerful tracers in low temperature regimes like ocean waters and
subduction zones, while very little is known about the isotopic fractionation in high pressure–temperature (PT) regimes. However,
recent studies indicate that boron fractionation in silicate–water systems may follow a systematic relationship with temperature,
regardless of the geological regime. In this paper we review previous B isotope studies and test the empirical relationship proposed
by Williams et al. [Boron isotope geochemistry during diagenesis. Part I. Experimental determination of fractionation during illiti-
zation of smectite, Geochim. Cosmochim. Acta 65 (2001a) 1769–1782; Boron isotope geochemistry during diagenesis. Part II. Appli-
cations to organic sediments, Geochim. Cosmochim. Acta 65 (2001b) 1783–1794; Application of boron isotopes to the
understanding of fluid–rock interactions in a hydrothermally stimulated oil reservoir in the Alberta Basin, Canada, Geofluids 1
(2001c) 229–240] by comparing it with the wealth of earlier data from studies of different geological scenarios. Our main conclusion
is that, given the large variations of B isotope fractionation patterns in natural silicate–water systems, the majority of these systems
are not represented by the proposed relationship. Factors more complex than temperature alone appear to control B isotope geo-
chemistry, which include species of the silicate mineral, starting fluid, pH, geological setting, fluid/rock ratio (i.e. porosity), or time
for equilibration between the solid and fluid phase.
Ó 2005 Elsevier Ltd. All rights reserved.

Keywords: Boron; Stable isotopes; Clay minerals; B isotope fractionation; Water–silicate

1. Introduction mostly from the coordination change in B between the



trigonal B(OH)3 of dissolved to tetrahedral BðOHÞ4 ,
Boron and B isotopes have been shown to represent a which substitutes for Al and Si in silicate minerals,
powerful proxy in fluid–rock interaction processes in the whereas 10B has a preference for tetrahedral bonds
geochemically very reactive area of low temperature (Kakihana and Kotaka, 1977; Palmer and Swihart,
alteration of geomaterials. This is mostly because of 1996; Palmer et al., 1987). In the field of interactive pro-
the geochemical characteristics of B, i.e., its high solubil- cesses between siliceous sediment/rock and aqueous
ity in aqueous fluids, its generally high volatility, and its fluid in diagenetic regimes, B is one of the key tracers
large isotopic variation in natural environments. The al- in subduction zone processes (Spivack and Edmond,
most 100&-span of B fractionation in nature results 1987; Bebout et al., 1993; You et al., 1993; Kopf et al.,
2000; Deyhle et al., 2001; Deyhle and Kopf, 2001,
* 2002; Kopf and Deyhle, 2002), alteration of igneous
Corresponding author. Tel.: +1 858 822 4345; fax: +1 858 822
4945. rock (Smith et al., 1995; Benton et al., 2001), arc volca-
E-mail address: adeyhle@ucsd.edu (A. Deyhle). nism (e.g., Rosner et al., 2003; Ishikawa and Nakamura,

1474-7065/$ - see front matter Ó 2005 Elsevier Ltd. All rights reserved.
doi:10.1016/j.pce.2005.04.003
A. Deyhle, A.J. Kopf / Physics and Chemistry of the Earth 30 (2005) 1038–1046 1039

1994; Ishikawa and Tera, 1997; Smith et al., 1995; Ishik- 30


awa and Tera, 1999; Ishikawa et al., 2001), or diagenesis errors are size of symbols
(Williams et al., 2001c; Kopf and Deyhle, 2002; Kopf 20 2
1.6 I/S
x -2 n for
et al., 2003). Isotope ratios are mostly a function of tem- .09 atio

∆ mineral-water (‰)
3
10 y=1 t ion
perature, pH, water chemistry, and the silicate minerals f rac

1100 °C
yg en
involved, as well as of pressure (Palmer et al., 1992; Her- 0 Ox

350 °C
vig et al., 2002).

300 °C
Many studies have examined the nature of B ex- -10
Bor
change and fractionation in silicate–water systems, both on f
ract
-20 i ona
in the laboratory (e.g., Couch and Grim, 1968; You

25 °C
y=1 tion
0.12
x+
et al., 1995; You and Gieskes, 2001; Williams et al., -30
2.44
2001a) and in natural settings (e.g., Ishikawa and
Nakamura, 1993; Kopf and Deyhle, 2002; Deyhle and -40
0 0.5 1 1.5 2 2.5 3 3.5 4
Kopf, 2002; Kopf et al., 2003; Deyhle et al., 2003,
2004). Based on their own studies and work by Palmer 1000/T (K)
et al. (1987), Williams et al. (2001a,b,c) have proposed Fig. 1. Results from hydrothermal experiments as well as boron
a linear B isotope fractionation relationship, designed isotope fractionation model by Williams et al. (2001a, their Fig. 13).
to describe a temperature-dependent B fractionation in For comparison, the authors also showed oxygen isotope fractionation
from Yeh and Savin (1977). Data are shown as Dd11Bmudd11Bfluid
natural silicate water systems, which can be applied
versus temperature (as 1000/TKelvin).
from low to high temperatures and to different silicate
minerals. The purpose of this manuscript is to review lit-
erature data from both the latter publications as well as (1100 °C) melt experiments of mid-oceanic ridge basaltic
all other work available on fluids and silica-bearing sol- glasses at 100 MPa confining pressure. The experiments
ids. This will allow testing the suggested B fractionation were doped with B standard (NBS SRM 951) and ran
trend for silicate systems, and it will additionally demon- for 24 h. The fractionation between aqueous fluid and
strate the complexity of B-isotope fractionation in differ- basaltic melt, which is related to the coordination
ent geological settings, and point to parameters other change from trigonal in the fluid to tetrahedral sites in
than temperature that might control B fractionation. the melt, resulted in a Dd11Bmud  d11Bfluid-value of
5& (see Fig. 1). In a later manuscript, Hervig et al.
(2002) have summarized experiments by other workers,
2. Data compilation which cover a wider T-range from 550 °C to 1100 °C
(their Fig. 3). These data show different fractionation
2.1. Recent experimental data and postulated trends, possibly because different mineral/rock systems
relationship are presented in the same diagram (e.g., some silicates
like tourmaline do not plot anywhere near the linear
In recent years, several papers (Williams et al., 2001a; relationship). Only the 1100 °C melt-fluid fractionation
Hervig et al., 2002; Williams and Hervig, 2002) have at- data were used for constructing the graph in Fig. 1,
tempted to establish an universal relationship for the whereas later data of melt experiments on rhyolites
fractionation behavior of B in silicate–water systems as and basalts by the same authors show considerable scat-
a function of temperature. From laboratory experiments ter, (Hervig et al., 2002) which suggests that melt-fluid
these authors established a systematic trend in B ex- systematics are also more complex and dependent on
change between siliceous rocks and interstitial waters. other factors than temperature alone.
Williams et al. (2001a) carried out a series of long In addition to the data from autoclave clay testing
term (120–150 days) autoclave reaction experiments of and melt experiments (see above), Williams et al.
a clay mineral–water slurry in an attempt to reach equi- (2001a,b,c) used a fourth data point from the literature
librium conditions. The temperature range was 300– (at 25 °C from Palmer et al., 1987) and fitted a linear
350 °C, and fluids were extracted under vacuum on an graph through these data. This graph follows a linear
extraction line. A total of 3 experiments gave two data relationship: y = 10.12x + 2.44, where x = 1000/T
points for clay (at 300 °C and 350 °C, respectively; see [K], as shown in Fig. 1. A different slope for the fraction-
Fig. 1, central dots). The B isotopic composition of ation curve (y = 12.2x + 5.66) was chosen by Hervig
the fluid was determined by a simple mass balance calcu- et al. (2002), who did not extrapolate their data to the
lation using the B content and d 11Bclay-value of the solid low temperature endmember by Palmer et al. (1987).
fraction (see Williams et al., 2001a, p. 1775). For details Despite the fact that the two data points from the
we refer to Williams et al. (2001a,b,c) or Williams and three experiments by Williams et al. (2001a) lead to a
Hervig (2002). Hervig and Moore (2000) as well as Her- steeper gradient than the one published (see grey line in
vig et al. (2002) have reported results from high-T Fig. 2; y = 21.42x + 21.37), and regardless of the fact
1040 A. Deyhle, A.J. Kopf / Physics and Chemistry of the Earth 30 (2005) 1038–1046

30 legend of Fig. 3 to which we refer for more information.


beginning
best fit through data by of tests In any case, the data compiled from the literature are
20 Williams et al. (2001a) I/S
y=21.41x + 21.37 nf
or similar to those from Williams et al. (2001a, Fig. 1) with
n atio 77)
tio 19
rac vin (
respect to the temperature range (i.e., approximately 10–
10 nf
x y ge d Sa 1000 °C), but they cover a larger variety of geological
n
∆ mineral-water (‰)

end of o ha
tests Ye settings. Furthermore, natural systems are open systems
0
prop
B5
and not closed as in the experimental setup. As a conse-
Willi osed D23
-10 ams fract
et a ionat D11
quence, more minerals and mineral assemblages are in-
l. (2 io
001 n tren volved, which are possibly the main reason for the
a,b, d
-20 c)
large scatter. It should also be noted that the natural
systems are assumed to allow the respective system more
-30
sand- time for equilibration between the fluid and solid phase
stone
-40 sand-
stone
(maybe excluding tectonically very active scenarios like
the Costa Rica forearc). Also, the corresponding fluid
mud-
-50 stone mud-
stone
and solid B signatures are not affected by possible arti-
facts in the fractionation pattern as a result of the
-60
0 0.5 1 1.5 2 2.5 3 3.5 4 artificial addition of B solutions (like ‘‘doping’’ the
1000/T (K) aliquot with boric acid standard NIST SRM 951 to
2000 ppm B, d11B = 0& solution, as in Hervig et al.
B isotope data Williams et al. (2001a) (2002), or a 10B spiked synthetic solution with a starting
B isotope data used by Williams et al. B ratio of 11B/10B = 0.038[d11B  990&], as in You
(2001a) to define fractionation trend
et al., 1995; You and Gieskes, 2001). Such B signatures,
O isotope data Williams et al. (2001a)
which are very dissimilar from any natural setting,
B isotope data Williams et al. (2001b)
might not be representative for a study that aims at sim-
PRE steam B isotope data Williams et al. (2001c) ulating natural processes (see discussion below). Also,
POST steam B isotope data Williams et al. (2001c) the spiked tests by You et al. (1995) and You and Gies-
B isotope data Williams et al. (2001d) * kes (2001) showed an opposite trend for d11B signatures
(fluids becoming more positive with increasing tempera-
Fig. 2. Results from hydrothermal experiments (Williams et al., ture) when compared with an unspiked test (fluids and
2001a), the proposed B fractionation relationship, and the oxygen
isotope fractionation graph proposed by Yeh and Savin (1977), as
solids becoming more negative with time and increase
shown in Fig. 1. We found it very informative to include the field data of T).
from the other studies by (some of) these authors (Williams et al., Furthermore, the temperature dependence of B isoto-
2001b,c,d) where reference is made to their fractionation model pic fractionation may be estimated only for a simplified
(Williams et al., 2001b, Fig. 1; 2001c, Fig. 2), as well as a best fit graph case where all monomeric units coexist without interac-
through the hydrothermal data (Williams et al., 2001a). All data shown
are from tables out of Williams et al. (2001a,b,c,d); some values were
tion (see Oi et al., 1989, where a simple approximation
averaged. has been used for polyborate ions).

that other field data gathered by the same authors do not


match the suggested fractionation trend (Williams et al., 3. Discussion
2001b,c,d), the authors postulate that the proposed B
fractionation trend represents a good geothermometer Our discussion will focus on the compatibility of the
for all natural silicate–water systems and that there is postulated B-fractionation trend by Williams et al.
no mineral specific fractionation of B isotopes. (2001a,b,c) of natural silicate–water systems with results
from available literature data from natural systems, as
2.2. Other global data on silicate–water systems given in Fig. 3. In the first part, we will try to illuminate
the problem with the proposed relationship by focusing
In order to test the geothermometer hypothesis as put on one group of silicate minerals, namely clay minerals
forward by Williams et al. (2001a,b,c) more completely, (and their transformation with PT). In the second part
we have compiled data from natural silicate water sys- of the discussion, we will discuss hypotheses that might
tems spanning the vast majority of geological scenarios. explain the mismatch of data from different natural sys-
Those data are summarized in Fig. 3. The different tems compiled from the literature and the proposed trend.
examples cover a range from low-T (<20 °C) marine sce-
narios such as the Mississippi delta or the Costa Rican 3.1. B fractionation in mineral-specific natural systems
Trench to volcanic (Mt. Etna, Sicily), hydrothermal
(Escanaba Trough) and magmatic systems (Mt. Shasta, From Fig. 3 it can be deduced that the linear
Cascades). A detailed list of references is given in the trend proposed by Williams et al. (2001a) is not readily
A. Deyhle, A.J. Kopf / Physics and Chemistry of the Earth 30 (2005) 1038–1046 1041

Fig. 3. Data from B isotope studies of silicate rocks and accompanying fluids published previously; some data are averaged out the entire set of
results from the publication(s). Graph from B fractionation model by Williams et al. (2001a,b,c) is given for reference. Underlined references in the
legend have been cited in the manuscripts in question (Williams et al., 2001a,b,c). Note that all data presented were taken from field studies of
geosystems (silicates of subrecent to 3.8 Ga age, active as well as passive tectonic and hydrogeologic settings, various temperature regimes, etc.; see
references) with variable degrees of fluid–rock equilibration. Even if some of those systems are not fully equilibrated, it is surprising that the proposed
fractionation model fails to agree with any of them. See above references for more details.
1042 A. Deyhle, A.J. Kopf / Physics and Chemistry of the Earth 30 (2005) 1038–1046

applicable to all natural systems; instead it may only In a recent study, Kopf and Deyhle (2002) have
represent a limited number of natural systems. However, examined clay mineral-rich muds from mud volcanoes
the distribution of the various silicate–water data pairs all over the world. In mud domes, the fluid and mud
in Fig. 3 is not random, but follows some systematic may be mobilized as deep as 10 km below the EarthÕs
trends. For instance, clay mineral-dominated systems surface, and is transported upwards due to its buoyancy
plot in the lower right corner of the diagram whereas (Kopf, 2002). The muds and firm mudstones studied
hydrothermally altered rocks plot in the central part, consist predominantly of clay (>70 vol.%) and were
and igneous rocks tend towards the left part of the dia- taken from the abyssal plain, an accretionary wedge,
gram (Fig. 3). An examination of each of the various its backstop, and from an onshore orogenic wedge of
groups of silicate minerals or siliceous rocks would be collisional tectonic settings (Kopf and Deyhle, 2002).
far beyond the scope of this short manuscript. We will As a result, the extruded material is expected to show
however illuminate our point of systematic trends by a prograde illitization of smectite with depth (i.e., tem-
focusing on some phyllosilicates. perature). A plot of the difference between d11Bmud/mud-
11
Early work on shales established that reversible stone and d Bfluid versus temperature has previously been
adsorption of B at low temperatures occurs (e.g., Har- shown to be useful as a first-order estimate of a system-
der, 1961; Couch and Grim, 1968; Keren and Mezuman, atic diagenetic process (assuming equilibrium between
1981; Palmer et al., 1987). With increasing temperature the solid and fluid phase; Williams et al., 2001a, see
and stress (e.g., during burial or tectonic processes), bor- Fig. 1). For the mud volcano samples, we find a good
on desorbs to enrich the fluid phase and is consequently correlation of the different areas with prograde illitiza-
depleted in the clay (e.g., You et al., 1996). The ad- tion due to inferred burial depth (Fig. 4) even though
sorbed species is predominantly the light isotope 10B, it is not clear whether equilibrium between mud and
so that fluids originating from clay-rich pore fluids ob- fluid has been reached since most of these are open sys-
tain lighter d11B values than seawater (e.g., Spivack tems. Interestingly, however, the comprehensive mud
et al., 1987). In general, B gets enriched in fluids by pro- volcano data resemble a trend established earlier from
cesses such as desorption from clay particles, mineral d18O isotopes of smectite–illite clays (Yeh and Savin,
dehydration reactions, increasing temperature (T), and 1977). This led Kopf and Deyhle (2002) to propose that
alteration of volcanic or igneous rocks (see Palmer and despite the fact that active flow may occur at some of the
Swihart, 1996, and references therein). Among others, mud domes sampled (e.g., on the Mediterranean Ridge;
You et al. (1993, 1995), Deyhle et al. (2001), Deyhle see Corselli and Basso, 1996), re-equilibration between
and Kopf, 2001 and Kopf et al. (2000) found that fluids
released in collision zones generally have elevated B con-
tents, and often show a bimodal d11B variation with
depth. Isotopes increase to compositions heavier than 30
seawater (up to >50&) at moderate depth (seawa-
I/S
ter = 39.5&; Spivack and Edmond, 1987), before 20 for
tion )
decreasing again to lower d11B values (25–30&) at great- tio
na 19
77
mineral-water (‰)

f rac v in,
er depth. Alteration of igneous rocks usually leads to 10
yg
en
nd
Sa
Ox ha
d11B lighter than seawater, with a less profound enrich- mY
e
ment in B content of the fluid (Palmer, 1991). 0 (fro oe
s
an
volc
One aspect which complicates the behavior of B in ud
-10 nm Barbados
clays is the transformation of smectite to illite when tem- I/ Si
n for
perature and stress increase (e.g., Colten-Bradley, 1987). tio Pakistan
-20 na Malaysia
ctio Mediterranean
Hydrothermal experiments have demonstrated that fra Makran
ron Ridge
some of the B in the pore fluid can be incorporated into -30 Bo
Caucasus
the illite mineral lattice (substituting Si and Al; e.g., Wil- Western Alps
-40
liams et al., 2001a), but may also be re-adsorbed if con- 0 0.5 1 1.5 2 2.5 3 3.5 4
fining pressures are sufficiently low. These laboratory 1000/T (K)
results suggest that in a natural scenarios, both muds
Fig. 4. Data from selected clay mineral–water systems from mud
and fluids may have high B concentrations. As a result volcanoes all over the world. Boron isotope ratios of different mud
of illitization (i.e., greater mobilization depth of the volcano provinces (as Dd11Bmudd11Bfluid) are plotted versus temper-
mud), the B content in the mud is expected to increase, ature (as 1000/TKelvin), where the temperatures are estimated from the
while d11B of the mud decreases (Williams et al., 2001a). average depth of the mud and fluid in each area using the regional
geothermal gradient and average d11B data from Kopf and Deyhle
This is in general agreement with earlier work by You
(2002). For reference, we show d18O isotope data of smectite-illite clays
et al. (1996, where an illite-rich sediment from the Nan- from Yeh and Savin (1977). Note that the dashed line through the mud
kai accretionary prism was used) and Deyhle et al. volcano data is not a calculated best fit, but the projected relationship
(2004, where mature claystones were considered). from the oxygen isotopes during illitization.
A. Deyhle, A.J. Kopf / Physics and Chemistry of the Earth 30 (2005) 1038–1046 1043

mud and fluid may have occurred in the majority of the There are other parameters which hamper hydrother-
features. The overall agreement of the data supports mal experiment results such as those by You et al.
the hypothesis that B geochemistry can be used to trace (1995), You and Gieskes (2001), and Williams et al.
the depth of mud volcano roots (i.e., faults) as a func- (2001), and limit their suitability as analogues to natural
tion of temperature (i.e., geothermal gradient), at least silicate–water systems. Hydrothermal experiments are
when a single group of minerals (or a process, like the carried out under closed conditions, and, therefore, the
progressive illitization of smectite) is considered. It fur- data from such experiments should only be extrapolated
ther suggests that some mineral-specific fractionation to natural settings which are close to a closed system. As
reactions may follow a trend (Fig. 4) which is opposite only very few hydrothermal studies on B-isotope frac-
to the one proposed by Williams et al. (2001a). tionation have been conducted to date (You et al.,
1995; You and Gieskes, 2001; Williams et al., 2001a;
3.2. Problems related to B fractionation trends Hervig et al., 2002; Kopf et al., 2002), more studies need
to be carried out in order to determine whether a simple
Unfortunately, none of the Williams et al. (2001b,c,d) linear relationship as established by Williams et al.,
natural silicate–water data plot near or on the proposed 2001a, can be applied to all silicate–water systems. Data
graph (Fig. 2). This raises doubts about their conclusion from the same authors (Hervig et al., 2002) hint towards
that ‘‘there is a good agreement between the field results B-isotope fractionation systematics which apply only to
and the experimentally derived B-isotope fractionation certain silicate minerals, or groups of minerals. For in-
curve indicating that it is a fair assessment of tempera- stance, experiments on basalt and rhyolite melts show
ture dependence of B-isotope fractionations between sil- different fractionation trends (Hervig et al., 2002) and
icates and waters.’’ (Williams et al., 2001b). When suggest that Si-content and possibly coordination of B
plotting data compiled from various publications in may be crucial players in B incorporation and fraction-
the literature into the same diagram (Fig. 3), it appears ation. Hervig et al. (2002) assume B in basaltic and rhy-
that these silicate–water systems are randomly scattered olitic glasses to be tetrahedrally coordinated, but this
over the diagram rather than linearly aligned with re- assumption has not been proven yet. In addition, B spik-
spect to Dd11Bsilicate  d11Bfluid and temperature. ing (SRM 951, 2000 ppm, 0&) of basalt (as for the
The deviation of natural silicate–water data (see hydrothermal test of Hervig et al., 2002) might also
Fig. 3) from the proposed trend by Williams et al. cause an unnatural B-isotope fractionation behavior.
(2001a,b,c,d) and Hervig et al. (2002) is probably be- It can be observed that the basaltic and ryhyolitic frac-
cause B-isotope fractionation is very complex and min- tionation trends are different (Hervig et al., 2002, their
eral-specific and not solely a function of temperature, Fig. 3). Given the different mineralogical composition
as claimed by those authors. Below we discuss several of rhyolite and basalt, it is impossible to decide whether
parameters that are likely to further influence the frac- mineralogy, Si-content, or the artificial spiking in the ba-
tionation behavior of B–water–silicate systems. salt experiment are responsible for the different B frac-
First, numerous workers have shown that B isotope tionation behavior observed.
fractionation is a function of mineralogy, temperature, Also experimental conditions, such as duration of the
and pH (e.g., Palmer et al., 1987; Hemming et al., hydrothermal test, fluid/rock ratio (and its variation
1995; You et al., 1996). Specifically the pH is critical during the test; see Kopf et al. (2002)), and specifically
for the B-fractionation, as at low pH all dissolved B is starting solution and silicate type or mineral composi-
essentially boric acid (B[OH]3) and at high pH all dis- tion might yield different results. Certainly, strong spik-
solved B is essentially borate ion (B[OH]4). A compre- ing of the fluid in a hydrothermal test with 10B causes a
hensive study on the pH-dependence of B fractionation fractionation behavior of B which is not only non-repre-
in silicate–water systems is beyond the scope of this pa- sentative for natural settings, but equally non-applicable
per for several reasons. First, the pH is not given for in any interpretation. For instance, both You et al.
the majority of the earlier studies. In addition, pH cannot (1995) and You and Gieskes (2001) used fluids with a
easily be obtained (or, in case of natural systems, be d11B ratio of 990& at the beginning and 100&
reconstructed). Generally it is accepted that the lower at the end of the experiment. Such d11B signatures do
the pH, the lower the expected resulting d11B signature. not exist in nature, and we suspect that oversaturation
This effect could possibly be the reason for some of the of the fluid with 10B may cause artifacts regarding incor-
shift to more negative d11B signatures in the fluid seen poration of the light vs. heavy species, and results in a
in hydrothermal tests with increasing temperature (e.g., shift to positive d11B values which is not necessarily
You and Gieskes, 2001, pH starting value 7.3, pH end meaningful. In contrast, d11B of unspiked testing solu-
6.4). We anticipate the Williams et al. (2001a) fluids to tions becomes more negative with increasing tempera-
show a similar trend, however, the d11B of those fluids ture (You et al., 1995; Kopf et al., 2002). The latter
was not measured, instead d11B signatures were calcu- trend nicely matches data from natural settings where
lated by these workers (see above). deep-seated fluids in subduction zone decollement zones
1044 A. Deyhle, A.J. Kopf / Physics and Chemistry of the Earth 30 (2005) 1038–1046

show similar shifts to more negative d11B (You et al., our compilatory work. We also acknowledge the edito-
1993; Kopf et al., 2000). rial support by Andrea Dini and Cedric Corteel.
Not only temperature, pH (or, more generally speak-
ing, the chemical composition of the fluid), the presence
of organic matter, and mineralogy are factors responsi-
ble for B fractionation, but also pressure has been References
named as a possible influence (Palmer et al., 1992; Her-
Bebout, G.E., Ryan, J.G., Leeman, W.P., 1993. B-Be systematics in
vig et al., 2002). Finally, the fluid/rock ratio should also
subduction-related metamorphic rocks; characterization of the
be taken into consideration and could possibly influence subducted component. Geochim. Cosmochim. Acta 57 (10), 2227–
the extent of reaction/fractionation between fluid and 2237.
rock. The effects are hard to anticipate, since natural Benton, L.D., Tomascak, P.B., Helz, R.T., 1999. A study of boron
sediments and rocks have high initial porosities and isotopes in Kilauea Iki lava lake, Hawaii. Eos Trans., AGU Abstr.
80, F1197.
are hence reactive; on the other hand, burial causes a
Benton, L.D., Ryan, J.G., Tera, F., 2001. Boron isotope systematics of
T-increase (in the then compressed rock with lower slab fluids as inferred from a serpentine seamount, Mariana
porosity), which also enhances fluid–rock interaction. forearc. Earth Planet. Sci. Lett. 187, 237–282.
Hydrothermal compression tests, where P is continu- Brumsack, H.J., Zuleger, E., 1992. Boron and boron isotopes in pore
ously increased in parallel experiments at various tem- waters from ODP Leg 127, Sea of Japan. Earth Planet. Sci. Lett.
113, 427–433.
peratures, have been carried out to illuminate the
Chaussidon, M., Uitterdijk Appel, P.W., 1997. Boron isotopic
counteracting processes of PT-increase and porosity compositions of tourmalines from the 3.8-Ga-old Isua subcrustals,
reduction and their effect on B geochemistry. At the West Greenland: implications on the d11B value of early Archean
range of temperatures (20–150 °C) and pressures (0– seawater. Chem. Geol. 136, 171–180.
80 MPa) regarded, B fractionation during deformation Colten-Bradley, V.A., 1987. Role of pressure in smectite dehydra-
tion—Effects on geopressure and smectite-to-illite transformation.
of fluid-saturated rock can be separated (Kopf et al.,
AAPG Bull. 71, 1414–1427.
2002). However, additional work on geomaterials other Corselli, C., Basso, D., 1996. First evidence of benthic communities
than clay-rich sediments has to be done in the future to based on chemosynthesis on the Napoli mud volcano (Eastern
more fully understand boron in silicate–water systems. Mediterranean). Mar. Geology 132, 227–240.
Couch, E.L., Grim, R.E., 1968. Boron fixation by illites. Clays and
Clay Minerals 16, 249–256.
Deyhle, A., 2000. Boron and boron isotope geochemistry: Evidence for
4. Conclusion fluid processes at convergent margins, PhD thesis, Christian-
Albrechts-Universität zu Kiel, Germany, 149 pp.
Deyhle, A., Kopf, A., 2001. Deep fluids and ancient pore waters at the
Results from two decades of intensive studies of the backstop: Stable isotope systematics (C, B, O) of mud volcano
geochemistry of boron and its isotopes have led to the deposits on the Mediterranean Ridge accretionary wedge. Geology
conclusion that the isotope fractionation of boron is 29/11, 1031–1034.
probably mineral specific and is additionally influenced Deyhle, A., Kopf, A., Eisenhauer, A., 2001. Boron systematics of
authigenic carbonates: A new approach to identify fluid processes
by other parameters such as pH, fluid/rock ratio, and
in accretionary prisms. Earth Planet. Sci. Lett. 187, 191–205.
pressure (c.f., Figs. 3 and 4). This suggests that the use Deyhle, A., Kopf, A., 2002. Strong B-enrichment and anomalous
of the boron isotopes for geothermometry is probably boron isotope geochemistry in the Japan Forearc. Mar. Geol. 183,
not possible without narrowing down the conditions 1–15.
that have affected the isotope fractionation processes Deyhle, A., Kopf, A., Aloisi, G., 2003. Boron and boron isotopes as a
tracer for diagenetic reactions and depth of mobilization, using
(Williams et al., 2001a,b). This first comprehensive com-
muds and authigenic carbonates from eastern Mediterranean mud
pilation presented in Fig. 4 points to potential problems, volcanoes. In: Maltman, A.J., Van Rensbergen, P., Hillis, R. (Eds.)
and does not allow any definite conclusions regarding Subsurface sediment mobilisation. Geol. Soc. London, Spec.
the influence of the most prominent players in B isotope Publications, vol. 216, pp. 493–505.
fractionation in natural silicate–water systems. There- Deyhle, A., Kopf, A., Pawlig, S., 2004. A cross section through the
frontal Japan Trench subduction zone: Geochemical evidence for
fore, we suggest that care has to be taken when using
fluid flow and fluid–rock interaction from DSDP and ODP pore
fractionation trends as geothermometers. To better an- waters and sediments, 13, 1, Island Arc.
swer these questions, further careful studies would have Harder, H., 1961. Beitrag zur Geochemie des Bors: III. Bor in
to be conducted which include different pressures, pHs, metamorphen Gesteinen und im geochemischen Kreislauf. Nachr.
sample compositions/mineralogy, and temperatures. Akad. Wiss. Göttingen II. Math.-Physikal. Kl. 1, 1–26.
Hemming, N.G., Reeder, R.J., Hanson, G.N., 1995. Mineral–fluid
partitioning and isotopic fractionation of boron in synthetic
calcium carbonate, Geochim. Cosmochim. Acta 59, 371–379.
Acknowledgements Hervig, R.L., Moore, G., 2000. Experimental determination of the
partitioning and isotopic fractionation of boron between mafic
melts and vapor. State of the Arc 2000 workshop, Auckland, NZ.
The authors thank Sonia Tonarini and Joris M. Gies- Hervig, R.L., Moore, G.M., Williams, L.B., Peacock, S.M., Holloway,
kes for their helpful criticism and expert comments on J.R., Roggensack, K., 2002. Isotopic and elemental partitioning of
A. Deyhle, A.J. Kopf / Physics and Chemistry of the Earth 30 (2005) 1038–1046 1045

boron between hydrous fluid and silicate melt. Am. Min. 87, tourmaline and aqueous vapor; a temperature- and pressure-
769–774. dependent isotopic system. Chem. Geol. 101 (1–2), 123–129.
Ishikawa, T., Nakamura, E., 1993. Boron isotope systematics of Palmer, M.R., Swihart, G.H., 1996. Boron isotope geochemistry: An
marine sediments. Earth Planet. Sci. Lett. 117, 567–580. overview. In: Grew, E.S., Anovitz, L.M. (Eds.), Boron. Mineral-
Ishikawa, T., Nakamura, E., 1994. Origin of the slab component in arc ogy, petrology, and geochemistry, Reviews in Mineralogy, vol. 33.
lavas from across-arc variation of B and Pb isotopes. Nature Mineralogical Society of America, pp. 709–744.
(London) 370 (6486), 205–208. Pennisi, M., Leeman, W.P., Tonarini, S., Pennisi, A., Nabelek, P.,
Ishikawa, T., Tera, F., 1997. Source, composition and distribution of 2000. Boron, Sr, O, and H isotope geochemistry of groundwaters
the fluid in the Kurile mantle wedge; constraints from across-arc from Mt. Etna (Sicily) – hydrologic implications. Geochim.
variations of B/Nb and B isotopes. Earth Planet. Sci. Lett. 152 Cosmochim. Acta 64, 961–974.
(1–4), 123–138. Rose, E.F., Shimizu, N., Layne, G.D., Grove, T.L., 2001. Melt
Ishikawa, T., Tera, F., 1999. Two isotopically distinct fluid components Production Beneath Mt. Shasta from Boron Data in Primitive Melt
involved in the Mariana Arc; evidence from Nb/B ratios and B, Sr, Inclusions. Science 293, 281–283.
Nd, and Pb isotope systematics. Geology (Boulder) 27 (1), 83–86. Rosner, M., Erzinger, J., Franz, G., Trumbull, R.B., 2003. Slab-
Ishikawa, T., Tera, F., Nakazawa, T., 2001. Boron isotope and trace derived boron isotope signatures in arc volcanic rocks from the
element systematics of the three volcanic zones in the Kamchatka Central Andes and evidence for boron isotope fractionation during
Arc. Geochim. Cosmochim. Acta 65 (24), 4523–4537. progressive slab dehydration. Geochem. Geophys. Geosyst. 4 (8),
James, R., Rudnicki, M.D., Palmer, M.R., 1999. The alkali element 9005. doi:10.1029/2002GC000438.
and boron geochemistry of the Escanaba Trough sediment-hosted Smith, H.J., Spivack, A.J., Staudigel, H., Hart, S.R., 1995. The boron
hydrothermal system. Earth Planet. Sci. Lett. 171, 157–169. isotopic composition of altered oceanic crust. Chem. Geol. 126 (2),
Kakihana, H., Kotaka, M., 1977. Equilibrium constants for boron 119–135.
isotope exchange reactions. Bull. Res. Lab. Nucl. Reactors 2, 1–12. Spivack, A.J., Edmond, J.M., 1987. Boron isotope exchange between
Keren, R., Mezuman, U., 1981. Boron adsorption to clay minerals seawater and oceanic crust. Geochim. Cosmochim. Acta 51, 1033–
using a phenomenological equation. Clays Clay Mins. 29, 198–204. 1043.
Kopf, A., Deyhle, A., Zuleger, E., 2000. Evidence for deep fluid Spivack, A.J., Palmer, M.R., Edmond, J.M., 1987. The sedimentary
circulation and gas hydrate dissociation using boron and boron cycle of the boron isotopes. Geochim. Cosmochim. Acta 51, 1939–
isotopes in forearc sediments from Costa Rica (ODP Leg 170). 1949.
Mar. Geol. 167, 1–28. Vengosh, A., Chivas, A.R., McCulloch, M.T., Starinsky, A., Kolodny,
Kopf, A.J., 2002. Significance of mud volcanism. Rev. Geophys. 40 Y., 1991. Boron isotope geochemistry of Australian salt lakes.
(2), 1005. doi:10.1029/2000RG000093. Geochim. Cosmochim. Acta 55, 2591–2606.
Kopf, A.J., Deyhle, A., 2002. Back to the roots: Source depths of mud Williams, L.B., Hervig, R.L., Holloway, J.R., Hutcheon, I., 2001a.
volcanoes and diapirs using boron and B isotopes. Chem. Geology Boron isotope geochemistry during diagenesis. Part I. Experimen-
192, 195–210. tal determination of fractionation during illitization of smectite.
Kopf, A.J., Castillo, P.R., Deyhle, A., 2002. Water–rock interaction in Geochim. Cosmochim. Acta 65, 1769–1782.
the upper seismogenic zone in the Nankai Trough subduction Williams, L.B., Hervig, R.L., Holloway, J.R., Hutcheon, I., 2001b.
factory. EOS, Trans. AGU (Supplement) 83/47, F1294. Boron isotope geochemistry during diagenesis. Part II. Applica-
Kopf, A., Deyhle, A., Lavrushin, V.Yu., Polyak, B.G., Buachidze, tions to organic sediments. Geochim. Cosmochim. Acta 65,
G.I., Eisenhauer, A., 2003. Isotopic evidence for deep gas and fluid 1783–1794.
migration from mud volcanoes in a zone of incipient continental Williams, L.B., Wieser, M.E., Fennell, J., Hutcheon, I., Hervig, R.L.,
collision (Caucasus, Russia). Int. J. Earth Sci. 92, 407–426. 2001c. Application of boron isotopes to the understanding of fluid–
Leeman, W.P., Tonarini, S., Pennisi, M., Ferrara, G., 1995. Boron rock interactions in a hydrothermally stimulated oil reservoir in the
isotopic variations in volcanis exhalatives and thermal waters from Alberta Basin, Canada. Geofluids 1, 229–240.
Vulcano Island, Italy: Implications for the evolution of volcanic Williams, L.B., Hervig, R.L., Wieser, M.E., Hutcheon, I., 2001d.
fluids. Eos, Transactions, American Geophysical Union, 76, 46 The influence of organic matter on the boron isotope geochemistry
Suppl., 655. of the Gulf Coast Sedimentary Basin. Chem. Geol. 174, 445–
London, D., Hervig, R.L., Morgan, G.B., 1988. Melt vapor solubilities 461.
and elemental partitioning in prealuminous granite-pegmatite Williams, L.B., Hervig, R.L., 2002. Exploring intra-crystalline B-
systems: experimental results with Macusani glass at 200 MPa. isotope variations in mixed-layer illite-smectite. Am. Min. 87,
Contrib. Mineral. Petrol. 99, 360–373. 1564–1570.
Matthews, A., Pulitz, B., Hamiel, Y., Hervig, R.L., 2003. Volatile Yeh, H.-W., Savin, S.M., 1977. Mechanism of burial metamorphism of
transport during crystallization of anatectic melts: Oxygen, argillaceous sediments: 3. O-isotope evidence. Geol. Soc. Am. Bull.
boron and hydrogen stable isotope study on the metamorphic 88, 1321–1330.
complex of Naxos, Greece. Geochim. Cosmochim. Acta 67 (17), You, C.-F., Spivack, A.J., Smith, J.H., Gieskes, J.M., 1993. Mobili-
3145–3163. zation of boron at convergent margins: Implications for boron
Oi, T., Nomura, M., Musashi, M., Ossaka, T., Okamoto, M., geochemical cycle. Geology 21, 207–210.
Kakihana, H., 1989. Boron isotopic compositions of some boron You, C.F., Spivack, A.J., Gieskes, J.M., Rosenbauer, R., Bischoff,
minerals. Geochim. Cosmochim. Acta 53 (12), 3189–3195. J.L., 1995. Experimental study of boron geochemistry; implications
Palmer, M.R., Spivack, A.J., Edmond, J.M., 1987. Temperature and for fluid processes in subduction zones. Geochim. Cosmochim.
pH controls over isotopic fractionation during adsorption of boron Acta 59 (12), 2435–2442.
on marine clay. Geochim. Cosmochim. Acta 51, 2319–2323. You, C.-F., Castillo, P.R., Gieskes, J.M., Chan, L.H., Spivack, A.J.,
Palmer, M.R., Sturchio, N.C., 1990. The boron isotope systematics of 1996. Trace element behavior in hydrothermal experiments: Impli-
the Yellowstone National Park (Wyoming) hydrothermal system: cations for fluid processes at shallow depths in subduction zones.
A reconnaissance. Geochim. Cosmochim. Acta 54, 2811–2815. Earth Planet. Sci. Lett. 140, 41–52.
Palmer, M.R., 1991. Boron isotope systematics of hydrothermal fluids You, C.-F., Gieskes, J.M., 2001. Hydrothermal alteration of hemi-
and tourmalines: A synthesis. Chem. Geol. 94, 111–121. pelagic sediments: experimental evaluation of geochemical
Palmer, M.R., London, D., Morgan VI, G.B., Babb, H.A., 1992. processes in shallow subduction zones. Appl. Geochem 16, 1055–
Experimental determination of fractionation of 11B/10B between 1066.
1046 A. Deyhle, A.J. Kopf / Physics and Chemistry of the Earth 30 (2005) 1038–1046

Xiao, Y.K., Sun, D., Wang, Y., Qi, H., Jin, L., 1992. Boron isotopic Zuleger, E., You, C.-F., Gieskes, J.M., Spivack, A.J., 1994. Boron and
compositions of brine, sediments, and source water in Da Qaidam boron isotopes in sediments and pore waters from ODP Leg 127
Lake, Qinghai, China. Geochim. Cosmochim. Acta 56, 1561– (Japan Sea). Eos Transactions, American Geophysical Union, 75,
1568. 44 Suppl., p. 740.

You might also like