You are on page 1of 15

JOURNAL OF PETROLOGY VOLUME 41 NUMBER 9 PAGES 1439–1453 2000

Geochemical Constraints from Zoned


Hydrothermal Tourmalines on Fluid
Evolution and Sn Mineralization:
an Example from Fault Breccias at Roche,
SW England

B. J. WILLIAMSON1∗, J. SPRATT2, J. T. ADAMS2, A. G. TINDLE3 AND


C. J. STANLEY2
1
DEPARTMENT OF EARTH SCIENCES, UNIVERSITY OF BRISTOL, WILLS MEMORIAL BUILDING, QUEEN’S ROAD,
BRISTOL BS8 1RJ, UK
2
DEPARTMENT OF MINERALOGY, THE NATURAL HISTORY MUSEUM, CROMWELL ROAD, LONDON SW7 5BD,
UK
3
DEPARTMENT OF EARTH SCIENCES, THE OPEN UNIVERSITY, MILTON KEYNES MK7 6AA, UK

RECEIVED JANUARY 26, 1999; REVISED TYPESCRIPT ACCEPTED FEBRUARY 7, 2000

Hydrothermal fluid evolution north of the St Austell granite, southwest exploration, particularly for tourmaline-breccia-hosted ores in Cu-
England, has been studied through geochemical analysis of tourmaline porphyry deposits.
from a fault breccia of <2 cm width within massive quartz–
tourmaline rocks at Roche. Brecciated tourmaline grains have
overgrowths of <400 m width [Fe/(Fe + Mg) = 0·31–0·99]
with four chemically distinct zones (1–4, towards the margins). KEY WORDS: breccia; Cornwall; hydrothermal; tin; tourmaline
Variations in overgrowth composition were caused by episodic mixing
between Mg-, Al-rich magmatic hydrothermal fluids (dominant in
zone 1), with an increasing component of more oxidizing, Fe-rich
formation waters (zones 2 and 4). More oxidizing conditions are INTRODUCTION
supported by high Sn contents in zone 2 (<0·35 wt %), with Sn The nature and origin of hydrothermal fluids responsible
probably present as Sn4+ rather than Sn2+, the usual form in for tin–tungsten mineralization in SW England have
hydrothermal fluids. From X-ray maps, zones 1 and 3 occur been the subject of numerous mineralogical (e.g. Farmer
exclusively as overgrowths on pre-existing grains, indicating that et al., 1991; London & Manning, 1995; Smith &
overgrowth formation was kinetically favoured over tourmaline Yardley, 1996; Williamson et al., 1997) and fluid
nucleation. In zones 2 and 4, nucleation and growth occurred, inclusion studies ( Jackson et al., 1982; Alderton &
possibly as a result of supersaturation with respect to tourmaline Rankin, 1983; Shepherd et al., 1985; Bottrell & Yardley,
during increased mixing with formation waters. Tourmaline is 1988; Wilkinson et al., 1994, 1995; Smith et al., 1996;
associated with the main episode of mineralization in many important Smith & Yardley, 1996). However, there are many
mineral deposits, often unaffected by alteration. This method of aspects of their evolution that remain poorly understood.
studying hydrothermal fluid evolution may therefore have uses in One of the most crucial of these is the source

∗Corresponding author. Telephone: +44-117-954-5235. Fax: +44-


117-925-3385. e-mail: Ben.Williamson@bristol.ac.uk  Oxford University Press 2000
JOURNAL OF PETROLOGY VOLUME 41 NUMBER 9 SEPTEMBER 2000

and behaviour of elements involved in mineralization, have greatly limited the extent of wall-rock reaction,
particularly whether they had a magmatic source or which may otherwise have altered the chemical and
were scavenged from crustal rocks by convective physical nature of the hydrothermal fluids. The com-
circulation of magmatic hydrothermal fluids and/or positions of hydrothermal vein minerals in the breccia
formation waters. What is also poorly understood is are therefore likely to yield valuable information on the
the strong association between tourmalinization and nature of the fluids from which they formed.
cassiterite mineralization. No direct role has been
identified for B and F in the formation of the Sn
deposits (Pollard et al., 1987; Taylor & Wall, 1993)
although it has been suggested that their presence in GEOLOGICAL SETTING
granite magmas lowers the solidus temperature to Roche Rock forms an isolated stock of MQT rock within
>650°C (Pichavant & Manning, 1984), allowing time Lower Devonian, mainly calcareous slate, Meadfoot Bed
for more extensive fractionation and partitioning of country rocks (Fig. 1). It lies >500 m north of the
ore-forming elements into hydrothermal fluids (Pollard northern margin of the St Austell granite, the smallest
et al., 1987). of the five main apophyses of the Hercynian batholith
Tourmaline is a particularly useful mineral in the study of SW England. The St Austell granite is a highly complex
of ore-forming fluids as it occurs in a wide variety of ore body consisting of six main granite types, many of which
deposits (e.g. Slack, 1982, 1996; Slack et al., 1993; London show extensive late magmatic and hydrothermal al-
& Manning, 1995; Griffin et al., 1996). It is also stable teration (Manning et al., 1996). Granite magmatism in
during most types of weathering processes and potentially the area occurred over an extended time period. The
up to upper amphibolite facies [reviewed by Henry & main biotite granites have yielded an age, by Rb–Sr
Dutrow (1996)], limiting problems with later alteration, whole-rock isochron, of 285 ± 4 Ma and the Brannel
which can be extremely extensive in mineral deposits. In elvan, a nearby late granite porphyry dyke, an age of
addition, it has a complex chemistry, with a number of 270 ± 9 Ma (Darbyshire & Shepherd, 1985). Present
different site occupancies (e.g. Burt, 1989; London & outcrop is considered to have been close to the roof of
Manning, 1995; Hawthorne, 1996; Jiang et al., 1998), the intrusion as indicated by the presence of numerous
which can accommodate a chemically varied suite of pegmatites occurring as sheets and containing abundant
ore-forming elements. Its trace element composition may miarolitic cavities carrying quartz, tourmaline, zinnwal-
therefore provide a tool in exploration [see review by dite, topaz and a wide range of other phases (Manning,
Slack (1996)]. This was recently demonstrated by a world- 1983; Weiss, 1989). The St Austell granite contains
wide study of massive sulphide deposits, which identified abundant cross-cutting tourmaline fissure veins and stock-
a good correlation between base metal proportions in works, which are often mineralized, containing amongst
tourmaline and the bulk composition of ore being mined others cassiterite, stannite and wolframite. The veins are
(Griffin et al., 1996). commonly bordered by greisens of <10 cm width, which
There have been several previous studies of tourmaline have been dated in the Goonbarrow pit, >2 km SE of
in SW England. These have mainly focused on the Roche, at 273·6 ± 1·1 Ma (K–Ar, Bray & Spooner,
discrimination of different genetic environments using 1983) and 276·8 ± 1·2 Ma ( 40Ar/39Ar, Chesley et al.,
‘major element’ variations (Manning, 1991; London & 1993).
Manning, 1995) and on the source of B using B isotopes MQT rocks mainly crop out at the margins of the
(Smith & Yardley, 1996). There have also been useful western part of the St Austell granite. In the past, probably
trace element studies on tourmaline from different settings from observations made in long since disused un-
in SW England. Li concentrations in tourmaline (and derground workings, these bodies have been described
other minerals) from granites and a wide variety of other as forming elongate, tin-bearing stripes proceeding from
rock types [including Roche massive quartz–tourmaline the main mass of St Austell granite (De La Beche, 1839;
(MQT) rock] have been determined by ion microprobe Collins, 1878). Today, their contacts with the granites
(Wilson & Long, 1983). Power (1968) has provided trace and surrounding country rocks cannot be seen. The
element data (including Sn) for a similar variety of rocks. origin of MQT rocks, and particularly whether they
There have, however, been no detailed studies on major formed from highly evolved silicate liquids or magmatic
and trace element variations within zoned hydrothermal hydrothermal fluids, remains problematic despite numer-
tourmalines and the bearing of this on the evolution of ous studies (see London & Manning, 1995). Badham
ore-forming fluids. Hydrothermal breccias at Roche are (1980) pointed out the lack of replacement textures in the
ideal for such a study as they formed within MQT rocks, MQT rocks and therefore suggested that they crystallized
which consist almost entirely of quartz and tourmaline, from a quartz–tourmaline magma that rose above the
the latter showing strong mechanical and chemical sta- roof of the main mass of granite. Charoy (1982) con-
bility (Slack, 1982; Henry & Dutrow, 1996). This will sidered that MQT rock is undoubtedly of magmatic

1440
WILLIAMSON et al. FLUID EVOLUTION FROM TOURMALINE COMPOSITIONS

Fig. 1. Map showing the location of Roche Rock and the main granite varieties of the St Austell granite [after Exley & Stone (1982) and
Manning et al. (1996)].

origin and formed by unmixing of a phase rich in B, diameter, with tourmaline extending into the voids. The
alkalis and silica. Smith & Yardley (1996) came to a similar tourmaline shows no embayments or other textures that
conclusion from B isotope studies, that the formation of may indicate partial resorption, and often has con-
MQT rocks was due to crystallization of volatiles pro- centrically zoned hydrothermal overgrowths. The over-
duced by the degassing of granite silicate liquids. There growths, which are developed parallel to the c-axis and
is therefore little argument that MQT rocks formed almost exclusively at one end of each tourmaline grain,
from magmatic ‘fluids’ of some kind, rather than from are dark blue–purple () to light blue or pale tan ()
formation waters. This is supported by the MQT rock (Fig. 2a and f ). The outermost part of the overgrowths
tourmaline being more similar in composition to mag- commonly consists of fine (<10 m width) feathery fibres
matic tourmaline from SW England rather than breccia (Fig. 2). Tourmaline also occurs as fine-grained needles
and replacement deposit tourmaline in the country rocks and sprays of feathery crystallites throughout the quartz
(London & Manning, 1995). matrix. The angular tourmaline fragments commonly
The hydrothermal breccia being studied forms a nar- contain rounded inclusions of quartz and more rarely
row (<2 cm width) vein in a large boulder of MQT rock inclusions of zircon, apatite and monazite.
at the southwestern margin of Roche Rock. The MQT Quartz in the breccia is granoblastic and considerably
rock immediately around the breccia is generally tour- more coarse grained (2–3 mm) than the tourmaline.
maline rich (70%) and contains a number of narrow Large, optically continuous crystals commonly enclose a
‘quartz’ veins with <5% acicular tourmaline. From the number of broken tourmaline crystals. In tourmaline-rich
elongate nature of the breccia, it is fault related rather areas, quartz occasionally forms narrow veins (<100 m
than representing an intrusive hydrothermal body, such
width), which cross-cut tourmaline aggregates and com-
as those at Wheal Remfry, in the western part of the St
monly contain acicular tourmalines. The quartz contains
Austell granite (Allman-Ward et al., 1982, fig. 1).
abundant fluid and mineral inclusions, the latter being
mainly haematite, and more rarely zircon, monazite and
apatite. Many of the fluid inclusions contain halite crystals
PETROGRAPHY and fluids associated with the formation of the breccias
The breccias contain 50–90% angular, fractured tour- are, therefore, likely to have been relatively saline. Fluid
maline grains and abundant cavities, up to 0·5 cm in inclusion studies on Sn–W veins from SW England by

1441
JOURNAL OF PETROLOGY VOLUME 41 NUMBER 9 SEPTEMBER 2000

Fig. 2. (a) Photomicrograph in plane-polarized light of brecciated tourmaline (light brown cores with pale blue rims) with light blue to black
zoned overgrowths. T, tourmaline; Qz, quartz. The matrix is quartz containing fine tourmaline needles and abundant fluid inclusions. (b–e) X-
ray element maps for Mg, Fe, Sn and F for the same field of view as in (a). The map for Sn is overlaid (in white) with the outline of four discrete
zones in the overgrowths (1–4) and the brecciated MQT tourmaline (0) determined from the maps for Fe, Mg and F. (f ) Photomicrograph in
plane-polarized light for a second area of the breccia showing a brecciated tourmaline grain with secondary hydrothermal overgrowth. (g, h) X-
ray element maps for Fe and Sn for the field of view in (f ) and an overlay for the different zones on the map for Sn.

1442
WILLIAMSON et al. FLUID EVOLUTION FROM TOURMALINE COMPOSITIONS

Alderton & Harmon (1991) indicated moderate salinities of 6 a.p.f.u., mainly in zones 2 and 4 (especially for
of between 5 and 15 wt % NaCl equivalent, with ho- traverse 2). Extensive checks were carried out to dem-
mogenization temperatures of between 230 and 400°C. onstrate that this was not an analytical problem, as is
More detailed fluid inclusion studies on the breccias also evident from its effects being strongest in certain
could not be carried out because when the quartz was zones. The high Si values are therefore considered to be
viewed in cathodoluminescence, it was found to be made an artefact of the recalculation procedure used to express
up of a complex ‘mosaic’ of angular (brecciated) fragments the data in formula units, particularly the assumptions
cemented by several generations of quartz. It would of total Fe as Fe2+, (OH, F) fixed at 4 a.p.f.u. and B
therefore be extremely difficult to identify fluid inclusions fixed at 3 a.p.f.u., although the latter seems likely from
in quartz cement, formed during the brecciation process, recent structural refinement studies (Bloodaxe et al.,
from those in the breccia tourmaline inherited from the 1999).
MQT rocks. X-ray element maps were generated for small
The MQT rocks at Roche are for the most part (>0·5 mm2) areas of the sample surface using a moving
granoblastic and homogeneous. They are medium stage rastering program on an automated Cameca
grained (<1 to 3 mm) and consist almost entirely of SX50 microprobe. The maps were obtained using an
tourmaline (30–70%) and quartz. They show very little accelerating voltage of 15 kV, a beam current of 100 nA,
in the way of internal structure apart from the presence of a spot size of >1 m, a spot interval of 2 m and a
rare, dark, commonly elongate, tourmaline-rich patches, count time of 200 ms per spot. Counts per second gain
‘quartz veins’ (containing <10% tourmaline) and numer- for each element were measured and plotted on separate
ous tourmaline- and quartz-lined vugs between <500 m maps using a colour scale, shown in the right-hand
and >50 cm in diameter. Tourmaline from the MQT margin, from black to red representing low to high
rocks commonly has cores showing relatively strong pleo- counts, respectively. X-ray element maps for Mg, Fe, Sn
chroism from dark brown () to pale tan–yellow () with and F are shown in Fig. 2.
rims that are mid-blue () to light blue–colourless ().
Some finer-grained tourmaline is entirely pale blue ()
to colourless ().

TOURMALINE CHEMISTRY
ANALYTICAL TECHNIQUES Tourmaline chemistry is complex with the possibility for
a wide range of substitutions (e.g. Burt, 1989; London &
Quantitative microprobe analysis was carried out using Manning, 1995; Hawthorne, 1996). A great deal of
an automated WDS Cameca SX50 microprobe with an
research has been undertaken on tourmaline structure
accelerating voltage of 20 kV, a beam current of 25 nA
and chemistry, which has recently culminated in a re-
and a spot size of >3 m, identical conditions to those
defined chemical classification scheme (Hawthorne &
used by London & Manning (1995). The following stand-
Henry, 1999). The basic formula for tourmaline is:
ards were used: jadeite for Na, wollastonite for Si and
XY3Z6(T6O18)(BO3)3V3W with common site occupancies
Ca, olivine for Mg and Fe, pure manganese for Mn,
being:
rutile for Ti, corundum for Al, synthetic KBr for K,
synthetic BaF for F, synthetic cassiterite for Sn. Detection
limits (3 SD) were <0·03 wt % (except for F, <0·1 wt %). X = Ca, Na, K or vacant;
Two traverses, parallel to the c-axis, were carried out Y = Li, Fe2+, Mg, Mn, Al, Cr3+, Fe3+, V3+, Ti4+;
across the contact of MQT fragments and secondary Z = Mg, Al, Fe3+, V3+, Cr3+;
overgrowths. The positions of the traverses are shown in T = Si, Al, B;
Fig. 2. Traverse 1 contains 25 points (length 190 m) B = B, (vacant);
compared with 49 points (length 284 m) in traverse 2, V = OH, O;
the latter being considerably more detailed. The data W = OH, F, O.
were recalculated to atoms per formula unit (a.p.f.u.),
based on 24·5 O, which assumes fixed B at 3 a.p.f.u.
and fixed (OH, F) at 4 a.p.f.u. The data have been The chemistry of tourmaline from the breccia is, how-
plotted on tourmaline ‘discrimination diagrams’ in Fig. ever, ‘relatively simple’ because of its restricted contents
3, and with distance along each traverse in Fig. 4. Data (in a.p.f.u.) of Ca <0·06, K <0·01, Mn <0·04 and Ti
for individual points from traverse 1, with an average <0·06 (Table 1). The main end-member compositions
for the MQT tourmaline, are presented in Table 1. between which the tourmalines in this study may fall
The recalculation procedure used produced slightly high have the general formula XY3Z6(Si6O18)(BO3)3V3W with
values (<6·1) for Si, above the maximum T site occupancy [after Hawthorne & Henry (1999)]:

1443
JOURNAL OF PETROLOGY VOLUME 41 NUMBER 9 SEPTEMBER 2000

Fig. 3. Variation diagrams for points in traverses 1 and 2 (except MQT rocks, which are shown as averages for each traverse) of (a) Mg/(Fe
+ Mg) vs Xvac/(Na + Xvac), (b) Fe vs Mg, (c) Al vs Fe and (d) Fe vs Xvac (in a.p.f.u.). Xvac (X vacancy) = 1 − (Na + K + Ca). Exchange
vector arrows and reference lines important in this study are indicated (Φ, vacancy). Also shown are data from SW England for tourmalines
from granites and pegmatites, MQT rocks, hydrothermal veins, breccias within granites, and replacements and breccias in country rocks [from
London & Manning (1995)].

broad geochemical trends, identified from the X-ray


(X) (Y3) (Z6) V3 W element maps (Fig. 2), and their outlines have been
overlaid on the X-ray element maps for Sn (Fig. 2d and
Elbaite Na+ Li+ 3+
1·5 Al1·3 Al63+ (OH)3 (OH) h). They have also been given different symbols in the
Schorl Na +
Fe 2+
3 Al63+ (OH)3 (OH) variation diagrams in Fig. 3 and are marked on the
Dravite Na+ Mg32+ Al63+ (OH)3 (OH) traverses in Fig. 4.
Olenite Na+ Al33+ Al63+ (O)3 (OH)
In the Xvac vs Mg/(Fe + Mg) discrimination diagram
Buergerite Na+ Fe33+ Al63+ (O)3 (F)
in Fig. 3a, the MQT tourmalines lie just within the
Povondraite Na+ Fe33+ Fe43+Mg22+ (OH)3 (O)
Schorl field, close to the boundary with Foitite. Because
of their high F contents (>0·5 a.p.f.u.), they may be
Rossmanite [] Li+ Al23+ Al63+ (OH)3 (OH)
described as F-Schorl (D. J. Henry, personal com-
Foitite [] Fe22+ Al63+ Al63+ (OH)3 (OH)
munication, 1999). The MQT tourmalines show a limited
Magnesio-foitite [] Mg22+ Al3+ Al63+ (OH)3 (OH)
range in compositions [Fe/(Fe + Mg) = 0·92–0·99],
similar to MQT tourmaline from Roche Rock, which
has a mean value of 0·937 ± 0·043 (3 SD; London &
For ease of description of chemical variations in the Manning, 1995). Their high Fe/(Fe + Mg), F and
tourmalines studied, the overgrowths have been divided Al are more similar to magmatic than hydrothermal
into four zones and the MQT tourmaline has been tourmaline in the Cornubian batholith and are very
described separately. The overgrowths can be clearly different from tourmaline formed in replacement deposits
distinguished from the MQT tourmaline on the X-ray and breccias in the country rock (Fig. 3).
element maps from their much lower F contents and Zone 1 contains considerably higher Mg and lower F
from their strong chemical zonation [Fe/(Fe + Mg) = compared with the MQT tourmaline. From a comparison
0·31–0·99]. The zones were delimited on the basis of of Figs 2, 3 and 4, the inner part of zone 1 (two points

1444
WILLIAMSON et al. FLUID EVOLUTION FROM TOURMALINE COMPOSITIONS

thousands of ppm Li found during ion probe studies of


tourmaline from Cornish hydrothermal breccias (Wilson
& Long, 1983).
Zones 2 and 4 contain high Fe (>3 a.p.f.u.) and low
Mg (<0·5 a.p.f.u.), and lie above the Schorl–Dravite
vector in Fig. 3b. The Y site has a maximum occupancy of
3 a.p.f.u. and the ‘excess’ Fe must therefore be contained
within a different site. It is highly unlikely to substitute
into the X site, in nine-fold co-ordination (Henry &
Dutrow, 1996), as its ionic radius is too small and because
Fe increases with Na towards the maximum Na site
occupancy in X of 1 a.p.f.u. (Table 1). The most likely
site for the ‘excess’ Fe is as Fe3+ in Z, because of the
substitution FeAl−1 (London & Manning, 1995; Dyar
et al., 1998). This is supported by the most Fe-rich
compositions showing proportionately large negative val-
ues for Al in the Y site (Fig. 4), which indicates a ‘deficit’
of Al in Z allowing the substitution of Fe3+ (London &
Manning, 1995). It is also possible that a component of
Fe in Y is contained as Fe3+ towards the composition of
Buergerite.
Tourmaline from the breccia and MQT tourmaline
were also analysed for a variety of ore-forming elements,
Sn, W and Cu. The only element to be found above
detection limits (dl) was Sn (<0·35 wt %; dl = 0·03 wt %),
and this was only in the overgrowths. Both the X-ray
Fig. 4. (a, b) Microprobe traverses from within the MQT tourmaline
and across the overgrowths. The locations of traverses 1 and 2 are element maps and traverses show significant con-
shown in Fig. 2. The units on the left-hand axis are in a.p.f.u. centrations of Sn in zone 2 and to a lesser extent in
zones 1 and 4. The homogeneous distribution of Sn
in traverse 1 and one point in traverse 2) is particularly throughout zone 2 is likely to indicate that the Sn is
Mg rich, and lies within the Dravite field of Fig. 3a. The ‘structurally bound’ rather than being contained as dis-
remainder of zone 1 straddles the Mg-Foitite, Foitite and crete mineral or amorphous inclusions.
Schorl fields, showing a poorly defined trend of decreasing A comparison of the photomicrograph in plane-po-
Mg/(Fe + Mg) and Xvac in the direction of the exchange larized light (Fig. 2a and f ) with the X-ray element maps
vectors (Fe2+Fe3+)(MgAl)−1, (NaFe2+)(ΦAl)−1 or shows that there is some correspondence between the
(Fe)(Mg)−1, or more probably a combination of these chemical composition of the tourmaline and its colour,
vectors (Fig. 3c and d). Zones 2, 3 and 4 fall within the although this is not always the case in tourmaline from
Schorl field with generally lower Xvac and Mg/(Fe + SW England (London & Manning, 1995). The more Fe-
Mg). From zone 2 to 3 there is a marked increase in rich zones are dark blue and the more Mg-rich zones
Xvac and Al and a slight increase in Mg/(Fe + Mg), are pale blue. The brecciated MQT tourmalines have
indicating the substitution (ΦAl)(NaFe2+)−1. The trend high F and are generally colourless–pale blue with light
within zone 3 is parallel to the exchange vector brown cores.
(ΦAl)(NaMg)−1 (Fig. 3c and d). Zone 4 shows higher Fe
and lower Al compared with zone 3 and similar Al, Fe
and Mg values to zone 2. The main difference between
DISCUSSION
zone 2 and 4 is the considerably higher Xvac in zone 4.
Because of the similarities in Al, Mg and Fe contents, Constraints on fluid evolution from
this cannot be explained by the substitution tourmaline chemistry
(ΦAl)(NaMg)−1 or (ΦAl)(NaFe2+)−1 and is more likely The relative contributions from magmatic waters, defined
to be due to (ΦFe3+)(NaFe2+)−1. as having equilibrated with a silicate melt (Bottrell &
Li contents in the tourmaline have been estimated Yardley, 1988), and formation waters in hydrothermal
using the technique of Burns et al. (1994) (see Table fluids are notoriously difficult to establish (Manning,
1). Calculated Li values are relatively low, with Li2O 1985). A dominantly magmatic source for early min-
<0·53 wt % (<1231 ppm) in the hydrothermal over- eralizing fluids in SW England is indicated by the close
growths. This is consistent with the hundred to several spatial and temporal association of mineralized veins to

1445
JOURNAL OF PETROLOGY VOLUME 41 NUMBER 9 SEPTEMBER 2000

Table 1: Tourmaline compositions for each point in traverse 1 (1–18 towards the margin) and an
average for the MQT tourmaline (point 0)

Point no.: Av. MQT 1 2 3 4 5 6


Zone: 0 1 1 1 1 1 1

Na2O 1·64 1·95 1·99 1·42 1·49 1·47 1·52


CaO 0·09 0·05 0·16 0·07 0·05 0·06 0·03
K2O 0·03 0·01 0·00 0·00 0·01 0·00 0·00
MgO 0·28 6·64 6·10 4·71 4·74 4·20 3·55
FeO 14·41 5·34 5·79 6·78 6·96 7·67 9·09
MnO 0·10 0·01 0·00 0·00 0·00 0·02 0·01
TiO2 0·08 0·09 0·17 0·21 0·20 0·19 0·19
Al2O3 35·43 34·73 35·28 36·00 35·87 35·95 35·50
SiO2 35·69 38·66 38·58 37·70 37·89 37·75 37·78
F 1·28 0·34 0·38 0·19 0·18 0·20 0·23
Li2O∗ 0·07 0·52 0·53 0·39 0·42 0·41 0·45
Total 89·09 88·34 88·98 87·47 87·81 87·92 88·35

Na 0·525 0·593 0·603 0·437 0·457 0·452 0·469


Ca 0·015 0·008 0·027 0·012 0·008 0·010 0·005
K 0·007 0·002 0·000 0·000 0·002 0·000 0·000
Mg 0·068 1·553 1·420 1·115 1·119 0·994 0·842
Fe 1·988 0·700 0·756 0·900 0·922 1·018 1·210
Mn 0·014 0·001 0·000 0·000 0·000 0·003 0·001
Ti 0·010 0·011 0·020 0·025 0·024 0·023 0·023
Al 6·887 6·418 6·493 6·737 6·693 6·725 6·659
Si 5·888 6·063 6·025 5·987 6·000 5·993 6·014
F 0·665 0·169 0·188 0·095 0·090 0·100 0·116
Li∗ 0·044 0·327 0·331 0·248 0·266 0·260 0·287
X 0·540 0·601 0·629 0·449 0·466 0·463 0·474
Y 2·069 2·254 2·176 2·015 2·041 2·015 2·054
Z 6·900 6·433 6·519 6·770 6·725 6·755 6·689
X+Y 2·610 2·855 2·806 2·464 2·506 2·477 2·528
Fe/(Fe + Mg) 0·967 0·311 0·347 0·447 0·452 0·506 0·590
Fe + Mg 2·056 2·253 2·176 2·015 2·041 2·012 2·052
XAl 0·788 0·496 0·544 0·757 0·725 0·747 0·703
Xvac 0·452 0·397 0·371 0·551 0·532 0·537 0·526

Sn wt % 0·01 0·04 0·06 0·14 0·19 0·17 0·19

granite plutons ( Jackson, 1979; Bray & Spooner, 1983; at Cligga Head, >30 km west of Roche (Smith & Yardley,
Chen et al., 1993) and from the high concentrations 1996). A contribution from formation waters is, however,
of transition metal chlorides in fluid inclusions from difficult to disprove, as they can commonly have B
tourmaline–topaz–quartz rocks at St Mewan Beacon contents up to 300 ppm (Leeman & Sisson, 1996), or
(Fig. 1; Bottrell & Yardley, 1988). Fluid inclusion and even higher if associated with evaporites, although these
isotopic studies suggest that Sn–W deposits at Cligga have not been found in the local country rocks (Smith
Head, Cornwall, were largely derived from magmatic & Yardley, 1996).
fluids (Smith et al., 1996). Boron in mineralizing fluids is Evidence for a magmatic component during the initial
also likely to have had a dominantly magmatic source, formation of zone 1 is the relatively high levels of Al
as indicated from B isotope studies of tourmalinized veins (Table 1, Fig. 4). Al is particularly soluble in low-pH

1446
WILLIAMSON et al. FLUID EVOLUTION FROM TOURMALINE COMPOSITIONS

Point no.: 7 8 9 10 11 12 13
Zone: 1 1 1 1 1 1 2

Na2O 1·69 1·85 1·95 1·89 1·82 2·12 2·54


CaO 0·03 0·04 0·03 0·02 0·07 0·12 0·22
K2O 0·01 0·00 0·00 0·01 0·02 0·02 0·03
MgO 3·26 3·55 3·82 2·37 1·93 0·78 0·16
FeO 10·21 10·03 10·05 12·37 13·70 17·04 21·00
MnO 0·02 0·02 0·00 0·03 0·03 0·02 0·03
TiO2 0·22 0·20 0·22 0·23 0·22 0·10 0·02
Al2O3 34·84 34·68 34·27 33·94 33·45 31·19 29·07
SiO2 37·34 37·27 37·57 37·11 36·92 35·95 35·07
F 0·24 0·24 0·19 0·25 0·28 0·38 0·44
Li2O∗ 0·38 0·35 0·42 0·40 0·35 0·31 0·04
Total 88·24 88·23 88·52 88·62 88·79 88·03 88·62

Na 0·527 0·577 0·606 0·594 0·575 0·692 0·849


Ca 0·005 0·007 0·005 0·003 0·012 0·022 0·041
K 0·002 0·000 0·000 0·002 0·004 0·004 0·007
Mg 0·781 0·851 0·912 0·573 0·469 0·196 0·041
Fe 1·372 1·348 1·346 1·677 1·868 2·400 3·026
Mn 0·003 0·003 0·000 0·004 0·004 0·003 0·004
Ti 0·027 0·024 0·027 0·028 0·027 0·013 0·003
Al 6·597 6·570 6·470 6·483 6·429 6·192 5·903
Si 6·001 5·992 6·019 6·016 6·022 6·056 6·043
F 0·122 0·122 0·096 0·128 0·144 0·202 0·240
Li∗ 0·247 0·228 0·271 0·263 0·230 0·209 0·026
X 0·532 0·584 0·611 0·597 0·588 0·714 0·889
Y 2·156 2·202 2·259 2·254 2·342 2·599 3·071
Z 6·633 6·602 6·505 6·521 6·465 6·208 5·907
X+Y 2·687 2·786 2·870 2·851 2·930 3·313 3·961
Fe/(Fe + Mg) 0·637 0·613 0·596 0·745 0·799 0·925 0·987
Fe + Mg 2·153 2·199 2·258 2·250 2·337 2·596 3·067
XAl 0·633 0·595 0·525 0·536 0·486 0·265 −0·050
Xvac 0·466 0·416 0·389 0·400 0·408 0·282 0·104

Sn wt % 0·16 0·12 0·08 0·08 0·07 0·08 0·11

(May et al., 1979), B- (Li + F)-rich aqueous fluids (Morgan or other resorption textures. Under normal cir-
& London, 1989). This type of fluid is likely to be of cumstances, for tourmaline formed within the granites
magmatic origin in the Roche area as formation waters or country rocks, tourmaline will have derived its Al
will have originated in, or at least passed through, local from the breakdown of local minerals such as feldspar
country rock calcareous slates. Al is considered to be and micas.
largely immobile in ‘typical’ aqueous crustal fluids Further evidence for a magmatic fluid component is
[Al(aq) = 5 × 10−5 to 10−4 m, at 400–700°C and from the relatively reduced nature of the hydrothermal
1 kbar; Ragnarsdottir & Walther, 1985]. There is no fluids during the initial formation of zone 1. The change
source of Al in the MQT rocks as MQT tourmaline, the in composition from the inner to outer part of zone 1 is
only major local Al-bearing phase, shows no embayments along the exchange vector (ΦAl)(NaMg)−1, combined

1447
JOURNAL OF PETROLOGY VOLUME 41 NUMBER 9 SEPTEMBER 2000

Table 1: continued

Point no.: 14 15 16 17 18
Zone: 2 3 3 3 4

Na2O 2·59 2·29 1·79 1·52 2·01


CaO 0·16 0·18 0·16 0·09 0·16
K2O 0·04 0·03 0·01 0·02 0·02
MgO 0·16 0·96 0·72 0·22 0·35
FeO 21·84 18·76 17·76 17·76 20·39
MnO 0·04 0·03 0·10 0·06 0·13
TiO2 0·02 0·02 0·02 0·01 0·01
Al2O3 28·31 30·71 31·62 32·10 29·31
SiO2 35·37 36·03 36·26 36·25 35·70
F 0·46 0·58 0·41 0·44 0·42
Li2O∗ 0·09 0·09 0·15 0·18 0·10
Total 89·08 89·68 89·00 88·65 88·60

Na 0·865 0·744 0·580 0·493 0·667


Ca 0·030 0·032 0·029 0·016 0·029
K 0·009 0·006 0·002 0·004 0·004
Mg 0·041 0·240 0·179 0·055 0·089
Fe 3·146 2·630 2·481 2·485 2·916
Mn 0·006 0·004 0·014 0·009 0·019
Ti 0·003 0·003 0·003 0·001 0·001
Al 5·746 6·067 6·224 6·328 5·907
Si 6·093 6·041 6·057 6·065 6·106
F 0·251 0·307 0·217 0·233 0·227
Li∗ 0·061 0·058 0·102 0·124 0·068
X 0·894 0·777 0·608 0·509 0·696
Y 3·193 2·874 2·674 2·548 3·024
Z 5·750 6·071 6·227 6·330 5·909
X+Y 4·087 3·651 3·282 3·057 3·720
Fe/(Fe + Mg) 0·987 0·916 0·933 0·978 0·970
Fe + Mg 3·187 2·870 2·660 2·540 3·005
XAl –0·158 0·112 0·284 0·395 0·015
Xvac 0·097 0·217 0·390 0·487 0·300

Sn wt % 0·35 0·10 0·07 0·10 0·11

Values are expressed as wt % oxides and a.p.f.u. (assuming 24·5 O), except Sn, which is in wt %. X = Na + Ca, Y = Fe +
Mg + Mn, Z = Al + 1·33Ti, Al in Y = Al + 1·33Ti + Si − 12 [following the method of London & Manning (1995)].
∗Calculated assuming 24·5 O (total Fe as Fe2+), assigning a value of three minus the Y-site total to Li, adding back the
equivalent wt % Li2O to the oxides, renormalizing again, then iterating this procedure until convergence is reached.

with a more minor component of (Fe)(Mg)−1, with no concentrations of Sn in the inner part of zone 1
evidence for substitutions involving Fe3+. Magmatic wa- (0·06 wt %, Fig. 4). Sn transport in aqueous fluids requires
ters will have been relatively reducing as the granites low pH and reducing conditions (Taylor & Wall, 1993),
from which they were derived show low magnetite abund- far more likely for magmatic than formation fluids, the
ance, low magnetic susceptibility (>1 × 10−5 e.m.u./g; latter having originated in or passed through calcareous
Willis-Richards & Jackson, 1989) and have been classified slate country rocks.
as belonging to the ilmenite series of Ishihara (1977). Substantial chemical changes in the hydrothermal fluid
This argument is supported by the presence of detectable are reflected by variations in tourmaline composition

1448
WILLIAMSON et al. FLUID EVOLUTION FROM TOURMALINE COMPOSITIONS

through zones 1–4. The main trend in zone 1 is parallel (above) it is likely that formation waters were relatively
to one or a combination of the exchange vectors depleted in Al and that magmatic fluids were the main
(Fe2+Fe3+)(MgAl)−1, (NaFe)(ΦAl)−1 or (Fe)(Mg)−1 (Fig. source for this element. Another limiting factor may have
3c and d). The field for zone 2 lies roughly along the been an increase in pH as a result of progressively higher
trend of zone 1, but is offset to higher Fe and lower degrees of mixing with formation waters from local
Xvac. Within zone 2, there is a poorly defined trend, calcareous slates. The amount of B required to stabilize
roughly in the direction of the exchange vector tourmaline increases with increasing pH, with tourmaline
(NaFe2+)(ΦAl)−1. Fe contents in zone 2 exceed the becoming unstable at pH >6·5–7 (Morgan & London,
maximum that can be accommodated in the tourmaline 1989). Other ‘ingredients’ for tourmaline formation,
Y site (3 a.p.f.u.). From the proportionate decrease in Al mainly Fe, Na, OH and Si, were abundant in the
in Z (Fig. 4), it is probable that the ‘excess’ Fe is hydrothermal fluids at this time, as indicated from the
accommodated in the Z site as Fe3+, indicating relatively presence of inclusions of haematite in matrix quartz and
oxidizing conditions during the formation of zone 2. high-salinity fluid inclusions. High hydrothermal fluid Fe
Slightly higher Mg contents in zone 3 are probably due contents are evident in the St Austell area from the
to the substitution (MgAl)(Fe2+Fe3+)−1 whereas the trend occurrence of bleached, biotite-depleted aureoles around
within zone 3 is parallel with the exchange vector hydrothermal quartz–tourmaline veins in biotite granites
(ΦAl)(NaMg)−1. Zone 4 shows almost identical com- (London & Manning, 1995). Similar bleached zones are
positions to zone 2, in terms of Fe, Mg and Al, but with also commonly seen in country rock metasediments.
a considerably higher Xvac. This is likely to be due Open-system formation of hydrothermal tourmaline,
to the substitution (ΦFe3+)(NaFe2+)−1, in addition to involving mixing between magmatic (providing B, Al,
(Fe2+Fe3+)(MgAl)−1, which caused the main com- Mg and Fe) and formation waters (probably providing
positional shift from zone 3 to 4. This may be indicating further Fe), is in agreement with the findings of London
even more oxidizing conditions in zone 4 than in zone & Manning (1995) from geochemical studies of many
2. Relatively oxidizing conditions are indicated from the different types of tourmaline from across SW England.
presence of haematite in matrix quartz, which is likely
to have formed contemporaneously with, or soon after,
zone 4.
The more oxidizing conditions could have been caused Ore-carrying potential of the hydrothermal
by either mixing with more oxidizing fluids or boiling. fluids
The latter was suggested for the formation of tourmalines In a comprehensive microprobe study of tourmaline
under relatively oxidizing conditions in the Larderello compositions from many different occurrences in Corn-
Geothermal Field, Italy (Cavarretta & Puxeddu, 1990). wall, none were found to contain Sn (London & Manning,
Boiling causes the loss of H into the vapour phase 1995). Why Sn was found in the overgrowths is not clear.
and the resulting oxidation of Fe in the aqueous phase It may simply be due to the refractory nature of the
(Drummond & Ohmoto, 1985). However, boiling is MQT rocks in which the breccia formed, with there
considered unlikely to have caused the apparent increase having been little cassiterite crystallization because of
in f O2 because F concentrations in the overgrowths limited wall rock reaction. Alternatively, Sn con-
increase in zones 2 and 4 (Table 1). This is the reverse centrations in the hydrothermal fluid may have been
of what would be expected during boiling, as, above below cassiterite saturation.
250°C, HF is preferentially partitioned into the vapour Sn concentrations show marked variations through the
phase (Reed & Spycher, 1985; Smith, 1995). overgrowths. In zone 1, Sn shows oscillatory zoning from
The increase in Fe and apparent increase in Fe3+/ >0·09 to 0·18 wt %, at a scale of >20 m, independent
Fe2+ through zone 1, which peak in zones 2 and 4, of any other element (Fig. 4, traverse 2). Zone 2 shows
suggest open-system behaviour during breccia formation, a peak for Sn of 0·35 wt %, which correlates with the
with an episodic but generally increasing contribution highest peak for Fe. Zones 3 and 4 show a decrease in
from relatively oxidizing formation waters. A significant Sn to similar levels as in zone 1. The origin of the
component of formation waters in the hydrothermal oscillatory zoning in zone 1 is uncertain, as this process
fluids of SW England has been suggested from numerous can be caused by a number of mechanisms, which
stable isotope determinations and fluid inclusion studies are often difficult to identify (e.g. Yardley et al., 1991;
(e.g. Jackson et al., 1982; Wilkinson, 1990; Alderton & Lanzirotti, 1995), particularly in minerals formed in open
Harmon, 1991; Wilkinson et al., 1995). systems [reviewed by Shore & Fowler (1996)]. These
One of the limiting factors for the formation of tour- mechanisms are either extrinsic, caused by external
maline in the breccia may have been a decrease in Al changes such as episodic fluid flow, variations in pressure,
in the hydrothermal fluid, as indicated from its general temperature, pH or f O2, or intrinsic, being caused by
decline through zones 1–4. From previous discussion ‘feed-back between crystal growth and solute diffusion

1449
JOURNAL OF PETROLOGY VOLUME 41 NUMBER 9 SEPTEMBER 2000

or surface effects as a source of non-linearity in the Taylor (1979)], being more easily accommodated within
crystal-growth kinetics’ (Shore & Fowler, 1996). The the tourmaline structure than Sn2+. This provides further
effects of extrinsic mechanisms can be recognized between evidence in support of more oxidizing conditions during
zones in the overgrowths as it is likely that each was the formation of zone 2. Which site it occupies is unclear,
formed from different mixtures of fluids. The oscillatory as again there are a variety of possible coupled sub-
zoning for Sn in zone 1 is better explained by intrinsic stitutions that could occur. Assuming that the high Sn4+
mechanisms, as this is the only element to show such content of the tourmaline reflects high concentrations in
regular oscillatory trends (Fig. 4). the hydrothermal fluid, its presence in the fluid is likely
How Sn was incorporated within the tourmaline struc- to be a transient phenomenon, as it would usually be
ture in zone 1 is problematic. From previous discussion, expected to precipitate out, forming cassiterite. It is
zone 1 (particularly the inner part) is likely to have been possible that Sn partitioning within tourmaline was kin-
formed under relatively reducing conditions. The Sn was etically favoured over cassiterite nucleation. Alternatively,
therefore probably contained in the hydrothermal fluid Sn concentrations may have been below cassiterite sat-
as Sn2+. Sn is dominantly thought to be transported as uration, although this seems unlikely because of the
Sn2+ in low redox hydrothermal fluids responsible for rare occurrence of such high tourmaline Sn contents.
the formation of Sn deposits (commonly in the form of Unfortunately, as the amount of Sn in the tourmaline
chloride and hydroxy chloride complexes; Jackson & probably relates to its valence state during tourmaline
Helgeson, 1985; Wilson & Eugster, 1990; Taylor & Wall, crystallization rather than to its concentration, no firm
1993). Sn2+ has an extremely large ionic radius of 0·122 conclusions can be reached on its magmatic or formation
nm in eight co-ordination (Shannon & Prewitt, 1969), water source. An interesting observation, however, is that
0·093 nm in six co-ordination (see Taylor, 1979). It is the Sn content of zone 4 is very much lower than that
therefore highly unlikely to substitute into an octahedral of zone 2, whereas (with the exception of Xvac) these
tourmaline site for Fe2+ (0·078 nm), Mg2+ (0·072 nm) or zones are almost identical in all other aspects of their
Al3+ (0·0535 nm) (six co-ordination; Shannon & Prewitt, geochemical compositions. Both zones were apparently
1969). The most likely substitution is within the X site,
formed under relatively oxidizing conditions with the
which is a nine co-ordinated trigonal antiprism located
hydrothermal fluid being dominated by formation waters.
along the three-fold axis of symmetry (Henry & Dutrow,
The most likely explanation for this is that the fluids
1996). This site is usually occupied by Na (0·102 nm),
were depleted in Sn by the time zone 4 was formed.
Ca (0·10 nm), with minor K (0·138 nm), or can be
Whether this was because of a decrease in cassiterite
vacant.
solubility in the magmatic fluid over time or was due to
The increasing concentrations of Sn in zone 2 can be
the diluting effect of mixing with formation waters is
explained either by an increase in Sn2+ concentrations
unclear.
in the hydrothermal fluid or by more oxidizing conditions,
with Sn4+ being more easily accommodated within the
tourmaline structure than Sn2+. The first possibility is
unlikely, as Sn in zone 2 is much higher than normally
found in tourmaline from SW England or mineral de-
Breccia formation
posits world-wide. Southwest England tourmaline was
found to have maximum Sn contents of <500 ppm from The X-ray element maps in Fig. 2 give an indication of
emission spectrography studies on 48 samples by Power the timing of tourmaline nucleation and growth. Zones
(1968). Proton microprobe studies of 32 samples of tour- 1 and 3, which show low Fe/Mg, are present only as
malines from massive sulphide deposits and tourmalinites overgrowths on, or fracture fills within, brecciated MQT
world-wide indicated Sn concentrations of <745 ppm tourmaline. Nowhere can they be seen to have formed
(Griffin et al., 1996). Nemec (1973) found a maximum of separate nucleation sites. Tourmaline nucleation appears
500 ppm in 80 tourmaline samples from different settings, to have occurred only during the formation of zones 2
Sn-mineralized and non-mineralized, world-wide. This and 4, with evidence for the latter being the common
suggests maximum Sn contents in tourmalines much less occurrence of feathery filaments of high-Fe tourmaline
than 1000 ppm, which is likely to reflect the relatively in the quartz matrix. The lack of nucleation during
reducing conditions that usually prevail during tour- the formation of zones 1 and 3 was probably due to
maline crystallization and the dominance, and in- crystallization on pre-existing tourmaline being kinetically
compatible nature, of Sn2+ in hydrothermal fluids, favoured. Nucleation during the formation of zones 2
especially where associated with Sn deposits. and 4 coincides with the proposed peak in mixing with
The increase in Sn concentration in zone 2 to an- formation waters and may therefore have been due to
omalously high levels is far more likely to be due to Sn4+, this process having led to supersaturation of the fluid
with an ionic radius of 0·069 nm [six co-ordination; see with respect to tourmaline.

1450
WILLIAMSON et al. FLUID EVOLUTION FROM TOURMALINE COMPOSITIONS

CONCLUSIONS AND WIDER popular approach of comparing the compositions of


single or a few tourmalines from many deposit types to
IMPLICATIONS OF THE STUDY identify genetic criteria to be used in exploration and
Compositional variations across secondary tourmaline genetic studies (e.g. Clarke et al., 1989; London & Man-
overgrowths in breccias at Roche have provided detailed ning, 1995; Griffin et al., 1996; reviewed by Slack, 1996).
information on the evolution of mineralizing fluids. Four The type of study presented here is likely to become
distinct zones have been recognized in the overgrowths, increasingly worth while with the development and rou-
zones 1–4 towards the margins. The overgrowths are tine use of more sophisticated methods of ‘point analysis’,
interpreted as having initially crystallized from relatively with progressively lower detection limits, such as laser
Mg-, Al-rich hydrothermal fluids of magmatic origin, ablation ICP-MS and ion probe, particularly for B and
which progressively mixed with more oxidized, Fe-rich ore-forming elements. Methods for studying the valency
formation waters, dominant during the formation of of Sn and Fe in situ (and non-destructively) in polished
zones 2 and 4. All four zones contain detectable con- sections of mineral samples are also eagerly awaited.
centrations of Sn, which reach a peak in zone 2. This
peak is considered to be due to more oxidizing conditions
during the formation of zone 2, with Sn being more
easily accommodated within the tourmaline structure as ACKNOWLEDGEMENTS
Sn4+ than as Sn2+. The presence of Sn4+ in hydrothermal All analyses were carried out at the Natural History
fluids is likely to be a transient phenomenon, as it would Museum, London, through internal funding. The authors
normally be expected to be precipitated out as cassiterite. wish to thank Terry Williams and Tony Wighton for
The identification of chemically distinct zones in the technical help, and Martin Smith, Darrell Henry, David
overgrowths has also yielded valuable information on the Manning and an anonymous referee for comments on
timing of tourmaline nucleation and growth. During the the manuscript.
formation of zones 1 and 3, tourmaline crystallization
was restricted to overgrowths on pre-existing brecciated
MQT tourmaline. During the formation of zones 2 and REFERENCES
4 there was also widespread tourmaline nucleation away Alderton, D. H. M. & Harmon, R. S. (1991). Fluid inclusion and stable
from the MQT tourmaline. These nucleation events are isotope evidence for the origin of mineralizing fluids in south-west
likely to relate to periods of supersaturation with respect England. Mineralogical Magazine 55, 605–611.
to tourmaline, probably as a result of a sudden increase Alderton, D. H. M. & Rankin, A. H. (1983). The character and
in mixing with formation waters. evolution of hydrothermal fluids associated with the kaolinised St
The study of just a few grains of tourmaline has Austell granite, SW England. Journal of the Geological Society, London
140, 297–310.
revealed details of the evolution of mineralizing fluids in
Allman-Ward, P., Halls, C., Rankin, A. & Bristow, C. M. (1982). An
the area. It is hoped that this will encourage more intrusive hydrothermal breccia body at Wheal Remfry in the western
extensive studies of chemical variations within single part of the St Austell granite pluton, Cornwall, England. In: Evans,
tourmalines as a tool in mineral exploration and in A. M. (ed.) Metallization Associated with Acid Magmatism. Chichester:
petrogenetic studies. Tourmalinization is commonly as- John Wiley, pp. 1–28.
sociated with mineralization in many of the world’s most Badham, J. P. N. (1980). Late magmatic phenomena in the Cornish
batholith—useful field guides for tin mineralisation. Proceedings of the
economically important mineral deposit types, par-
Ussher Society 5, 44–53.
ticularly in breccias related to Cu porphyry deposits. In Bloodaxe, E. S., Hughes, J. M., Dyar, M. D., Grew, E. S. & Guidotti,
the Los Bronces Cu porphyry, for example, which is one C.V. (1999). Linking structure and chemistry in the Schorl–Dravite
of the largest known of its type in the world, mineralization series. American Mineralogist 84, 922–928.
is concentrated within a tourmaline breccia, with tour- Bottrell, S. H. & Yardley, B. W. D. (1988). The composition of a
malinization having occurred before, during and after primary granite derived ore fluid from S.W. England, determined
the main episode of mineralization (Warnaars et al., 1985). by fluid inclusion analysis. Geochimica et Cosmochimica Acta 52, 585–588.
Bray, C. J. & Spooner, E. T. C. (1983). Sheeted vein Sn–W min-
Many of these types of deposit are highly altered from eralization and greisenization associated with economic ka-
the ingress of later fluids, which commonly produce olinization, Goonbarrow China Clay Pit, St Austell, Cornwall,
multiple generations of fluid inclusions and complex England: geologic relationships and geochronology. Economic Geology
mineral assemblages, which hamper the study of fluid 78, 1064–1089.
chemistry related to primary mineralization. Tourmaline Burns, P. C., MacDonald, D. J. & Hawthorne, F. C. (1994). The crystal
is commonly unaffected by later alteration and so can chemistry of manganese-bearing Elbaite. Canadian Mineralogist 32,
31–41.
preserve original information on the conditions of min- Burt, D. M. (1989). Vector representation of tourmaline compositions.
eralization. Its chemistry also often directly reflects that American Mineralogist 74, 826–839.
of related ores (Griffin et al., 1996). The study of individual Cavarretta, G. & Puxeddu, M. (1990). Schorl–dravite–ferridravite
tourmaline grains will complement the recently more tourmalines deposited by hydrothermal magmatic fluids during early

1451
JOURNAL OF PETROLOGY VOLUME 41 NUMBER 9 SEPTEMBER 2000

evolution of the Larderello Geothermal Field, Italy. Economic Geology Jiang, S.-Y., Palmer, M. R., Slack, J. F. & Shaw, D. R. (1998).
85, 1236–1251. Paragenesis and chemistry of multistage tourmaline formation in
Charoy, B. (1982). Tourmalinisation in Cornwall, England. In: Evans, the Sullivan Pb–Zn–Ag Deposit, British Columbia. Economic Geology
A. M. (ed.) Metallization Associated with Acid Magmatism. Chichester: 93, 47–67.
John Wiley, pp. 63–70. Lanzirotti, A. (1995). Yttrium zoning in metamorphic garnets. Geochimica
Chen, Y., Clark, A. H., Farrar, E., Wasteneys, H. A. H. P., Hodgson, et Cosmochimica Acta 59, 4105–4110.
M. J. & Bromley, A. V. (1993). Diachronous and independent Leeman, W. P. & Sisson, V. B. (1996). Geochemistry of boron and its
histories of plutonism and mineralization in the Cornubian batholith, implications for crustal and mantle processes. Mineralogical Society of
southwest England. Journal of the Geological Society, London 150, 1183– America, Reviews in Mineralogy 33, 645–695.
1191. London, D. & Manning, D. A. C. (1995). Chemical variation and
Chesley, J. T., Halliday, A. N., Snee, L. W., Mezger, K., Shepherd, significance of tourmaline from Southwest England. Economic Geology
T. J. & Scrivener, R. C. (1993). Thermochronology of the Cornubian 90, 495–519.
batholith in south-west England: implications for pluton em- Manning, D. A. C. (1983). Disseminated tin sulphides in the St Austell
placement and protracted hydrothermal mineralization. Geochimica granite. Proceedings of the Ussher Society 5, 411–416.
et Cosmochimica Acta 57, 1817–1835. Manning, D. A. C. (1985). Comparison of influence of magmatic
Clarke, D. B., Reardon, N. C., Chatterjee, A. K. & Gregoire, D. C. water on form of granite-hosted Sn–W deposits and associated
(1989). Tourmaline composition as a guide to mineral exploration: tourmalinisation from Thailand and southwest England. In: Halls,
a reconnaissance study from Nova Scotia using discriminant function C. (ed.) High Heat Production Granites, Hydrothermal Circulation and Ore
analysis. Economic Geology 84, 1921–1935. Genesis. London: Institute of Mining and Metallurgy, pp. 203–212.
Collins, J. H. (1878). The Hensbarrow Granite District: a Geological Description Manning, D. A. C. (1991). Chemical variation in tourmalines from
and a Trade History. Truro: Lake and Lake. south-west England. Proceedings of the Ussher Society 7, 327–332.
Darbyshire, D. P. F. & Shepherd, T. J. (1985). Chronology of granite Manning, D. A. C., Hill, P. I. & Howe, J. H. (1996). Primary lithological
magmatism and associated mineralisation, SW England. Journal of variation in the kaolinized St Austell Granite, Cornwall, England.
the Geological Society, London 142, 1159–1177.
Journal of the Geological Society, London 153, 827–838.
De La Beche, H. T. (1839). Report on the Geology of Cornwall, Devon
May, H. M., Helmke, P. A. & Jackson, M. L. (1979). Gibbsite solubility
and West Somerset. London: Longman, Orme, Brown, Green and
and thermodynamic properties of hydroxyaluminum ions in aqueous
Longmans.
solution at 25°C. Geochimica et Cosmochimica Acta 43, 861–868.
Drummond, S. E. & Ohmoto, H. (1985). Chemical evolution and
Morgan, G. B. & London, D. (1989). Experimental reactions of am-
mineral deposition in boiling hydrothermal systems. Economic Geology
phibolite with boron-bearing aqueous fluids at 200 MPa: implication
80, 126–147.
for tourmaline stability and partial melting in mafic rocks. Contributions
Dyar, M. D., Taylor, M. E., Lutz, T. M., Francis, C. A., Guidotti,
to Mineralogy and Petrology 102, 218–297.
C. V. & Wise, M. (1998). Inclusive chemical characterization of
Nemec, D. (1973). Tin in tourmalines. Neues Jahrbuch für Mineralogie,
tourmaline: Mössbauer study of Fe valence and site occupancy.
Monatshefte 2, 58–63.
American Mineralogist 83, 848–864.
Pichavant, M. & Manning, D. A. C. (1984). Petrogenesis of tourmaline
Exley, C. S. & Stone, M. (1982). Hercynian intrusive rocks. In:
and topaz granites; the contribution of experimental data. Physics of
Sutherland, D. S. (ed.) The Variscan Fold Belt in the British Isles.
Chichester: John Wiley, pp. 287–311. the Earth and Planetary Interiors 35, 31–50.
Farmer, C. B., Searl, A. & Halls, C. (1991). Cathodoluminescence and Pollard, P. J., Pichavant, M. & Charoy, B. (1987). Contrasting evolution
growth of cassiterite in the composite lodes at South Crofty Mine, of fluorine- and boron-rich tin systems. Mineralium Deposita 22,
Cornwall, England. Mineralogical Magazine 55, 447–458. 315–321.
Griffin, W. L., Slack, J. F., Ramsden, A. R., Tin Win, T. & Ryan, C. Power, G. M. (1968). Chemical variation in tourmalines from south-
G. (1996). Trace elements in tourmalines from massive sulfide west England. Mineralogical Magazine 36, 1078–1089.
deposits and tourmalinites: geochemical controls and exploration Ragnarsdottir, K. V. & Walther, J. V. (1985). Experimental de-
applications. Economic Geology 91, 657–675. termination of corundum solubilities in pure water between
Hawthorne, F. C. (1996). Structural mechanisms for light-element 400–700°C and 1–3 kbar. Geochimica et Cosmochimica Acta 49, 2109–
variations in tourmaline. Canadian Mineralogist 34, 123–132. 2115.
Hawthorne, F. C. & Henry, D. J. (1999). Classification of the minerals Reed, M. H. & Spycher, N. (1985). Boiling, cooling, and oxidation in
of the tourmaline group. European Journal of Mineralogy 11, 201–215. epithermal systems: a numerical modelling approach. In: Berger, B.
Henry, D. J. & Dutrow, B. L. (1996). Metamorphic tourmaline and R. & Bethke, P. M. (eds) Geology and Chemistry of Epithermal Systems.
its petrologic applications. Mineralogical Society of America, Reviews in Littleton, CO: Society of Economic Geologists, pp. 249–272.
Mineralogy 33, 503–557. Shannon, R. D. & Prewitt, C. T. (1969). Effective ionic radii in oxides
Ishihara, S. (1977). The magnetite-series and ilmenite-series granitic and fluorides. Acta Crystallographica B25, 925–946.
rocks. Mining Geology 27, 293–305. Shepherd, T. J., Miller, M. F., Scrivener, R. C. & Darbyshire, D. P.
Jackson, K. J. & Helgeson, H. C. (1985). Chemical and thermodynamic F. (1985). Hydrothermal fluid evolution in the relation to min-
constraints on the hydrothermal transport and deposition of tin: I. eralization in southwest England with special reference to the Dart-
Calculation of the solubility of cassiterite at high pressures and moor–Bodmin area. In: Halls, C. (ed.) High Heat Production Granites,
temperatures. Geochimica et Cosmochimica Acta 49, 1–22. Hydrothermal Circulation and Ore Genesis. London: Institute of Mining
Jackson, N. J. (1979). Geology of the Cornubian tin field: ‘a review’. and Metallurgy, pp. 345–364.
Bulletin of the Geological Society of Malaysia 11, 209–237. Shore, M. & Fowler, A. D. (1996). Oscillatory zoning in minerals: a
Jackson, N. J., Halliday, A. N., Sheppard, S. M. F. & Mitchell, J. G. common phenomenon. Canadian Mineralogist 34, 1111–1126.
(1982). Hydrothermal activity in the St Just mining district, Cornwall. Slack, J. F. (1982). Tourmaline in Appalachian–Caledonian massive
In: Evans, A. M. (ed.) Metallisation Associated with Acid Magmatism. sulphide deposits and its exploration significance. Transactions of the
Chichester: John Wiley, pp. 137–179. Institute of Mining and Metallurgy 91, B81–B89.

1452
WILLIAMSON et al. FLUID EVOLUTION FROM TOURMALINE COMPOSITIONS

Slack, J. F. (1996). Tourmaline associations with hydrothermal ore Wilkinson, J. J. (1990). The role of metamorphic fluids in the evolution
deposits. Mineralogical Society of America, Reviews in Mineralogy 33, of the Cornubian orefield: fluid inclusion evidence from south
559–643. Cornwall. Mineralogical Magazine 54, 219–230.
Slack, J. F., Palmer, M. R., Stevens, B. P. J. & Barnes, R. G. (1993). Wilkinson, J. J., Rankin, A. H., Mulshaw, S. C., Nolan, J. & Ramsey,
Origin and significance of tourmaline-rich rocks in the Broken Hill M. H. (1994). Laser ablation–ICP–AES for the determination of
district, Australia. Economic Geology 88, 505–541. metals in fluid inclusions: an application to the study of magmatic
Smith, M., Banks, D. A., Yardley, B. W. D. & Boyce, A. (1996). Fluid ore fluids. Geochimica et Cosmochimica Acta 58, 1133–1146.
inclusion and stable isotope constraints on the genesis of the Cligga Wilkinson, J. J., Jenkin, G. R. T., Fallick, A. E. & Foster, R. P. (1995).
Head Sn–W deposit, S. W. England. European Journal of Mineralogy Oxygen and hydrogen isotopic evolution of Variscan crustal fluids,
8, 961–974. south Cornwall, U.K. Chemical Geology 123, 239–254.
Smith, M. P. (1995). The geochemistry of water–rock interaction in Williamson, B. J., Stanley, C. J. & Wilkinson, J. J. (1997). Implications
progressively focused hydrothermal flow. Ph.D. Thesis, University from inclusions in topaz for greisenisation and mineralisation in
of Leeds. the Hensbarrow topaz granite, Cornwall, England. Contributions to
Smith, M. P. & Yardley, B. W. D. (1996). The boron isotopic com- Mineralogy and Petrology 127, 119–128.
position of tourmaline as a guide to fluid processes in the southwestern Willis-Richards, J. & Jackson, N. J. (1989). Evolution of the Cornubian
England orefield: an ion microprobe study. Geochimica et Cosmochimica ore field, southwest England: Part I. Batholith modelling and ore
Acta 60, 1415–1427. distribution. Economic Geology 84, 1078–1100.
Taylor, J. R. & Wall, V. J. (1993). Cassiterite solubility, tin speciation Wilson, G. A. & Eugster, H. P. (1990). Cassiterite solubility and tin
and transport in a magmatic aqueous phase. Economic Geology 88, speciation in supercritical chloride solutions. Geochemical Society Special
437–460. Publication 2, 179–195.
Taylor, R. G. (1979). Geology of Tin Deposits. New York: Elsevier. Wilson, G. C. & Long, J. V. P. (1983). The distribution of lithium in
Warnaars, F. W., Holmgren, D. C. & Barassi, F. S. (1985). Porphyry some Cornish minerals: ion microprobe measurements. Mineralogical
copper and tourmaline breccias at Los Bronces–Rio Blanco, Chile. Magazine 47, 191–199.
Economic Geology 80, 1544–1565. Yardley, B. W. D., Rochelle, C. A., Barnicoat, A. C. & Lloyd, G. E.
Weiss, S. (1989). The minerals of the china clay pits in Cornwall. Lapis (1991). Oscillatory zoning in metamorphic minerals: an indicator of
14, 11–24. infiltration metasomatism. Mineralogical Magazine 55, 357–365.

1453

You might also like