You are on page 1of 30

Chemical Geology, 33 (1981 ) 265--294 265

Elsevier Scientific Publishing Company, Amsterdam -- Printed in The Netherlands

THE GOLD-"CARBONATE ASSOCIATION: SOURCE OF CO2, AND COs


FIXATION REACTIONS IN ARCHAEAN LODE DEPOSITS

R. KERRICH and W.S. F Y F E


Department o f Geology, University of Western Ontario, London, Ont. N6A 5B7 (Canada)
(Accepted for publication May 18, 1981)

ABSTRACT

Kerrich, R. and Fyfe, W.S., 1981. The gold--carbonate association: source of CO2, and
CO 2 fixation reactions in Archaean lode deposits. Chem. Geol., 33: 265--294.

Abundant carbonate is a ubiquitous feature of Archaean lode gold deposits -- both in


domains of alteration enveloping veins, and in rocks enclosing auriferous sediments.
Detailed studies of chemical mass balance in metabasalts, progressively altered towards
gold-bearing quartz carbonate veins disposed within shear zones at Yellowknife, reveal
massive additions of CO 2, K, Si and Fe accompanying mineralisation, with concomitant
depletions of Na. Coherent behaviour of A1, Sc, Zr (also V and Nb) provides a reference for
constraining the volume relations during hydrothermal alteration. In the peripheral regions
of alteration, depletions of Ca and Mg result in overall volume reduction, but these ele-
ments are added to veins and their immediate alteration envelopes where there is a large
positive volume change. Precious metals, together with Cr, Ni, Cu, Zn, Pb, Rb, Cd and Ba,
have been added to the veins and altered wall rocks.
Whereas quartz, noble metals and other trace elements have been precipitated from
hydrothermal fluids, solution of Ca, Mg (and some Fe), indigenous to peripheral altera-
tion regions, combining with CO~ from the mineralising reservoir, appears to be the pro-
cess for forming the abundant ferri-dolomite gangue. Observations of many Archaean Au
deposits reveals that carbonate chemistry reflects the nature of wall rocks, with wall rocks
donating the bivalent metal cations, and hydrothermal fluids the CO 2. Alteration reactions
at Yellowknife involved hydrolysis of albite accompanied by fixation of Kaqueous to pro-
duce muscovite, with Na loss; and hydrolysis of chlorite + epidote with CO 2 fixation to
form ferri-dolomite.
Studies of oxygen isotopes, Fe2+/(Fe 2+ + Fe3÷), and structure reveal that Au, quartz
and carbonate were precipitated in the presence of fluids of probable metamorphic origin
(51~O + 8--9"/00 ), at low redox potential, and at ambient temperatures of 400--450°C,
during episodes of hydrofracturing.
The abundant CO 2 and K, requixed for extensive carbonate--nnuscovite replacement
alteration, could be supplied by fluids released during prograde metamorphism under
greenschist or greenschist---amphibolite facies conditions, where the relative proportion of
CO2/H~O is in the order of 0.2 to 0.5, and K/Na ~- 1. Given high CO~--CO in the hydro-
thermal reservoir, these molecules may act as complexing agents for transport of Au and
other rare elements (e.g., W, Pd, Ni, Cr).

0009-2541/81/0000--0000/$02.50 © 1981 Elsevier Scientific Publishing Company


266

INTRODUCTION

A ubiquitous feature of gold-bearing vein deposits in Archaean greenstone


belts is the presence of abundant carbonate minerals. Quartz and carbonates
are the dominant gangue in hydrothermal veins, and are typically present in
the extensive domains of alteration which envelop veins.
The nature of wall-rock alteration in proximity to Au-bearing veins has
been comprehensively documented by Bain (1933) from studies of deposits
in Ontario. Bain identified carbonate, chlorite and sericite as the predominant
+alteration minerals. Extensive development of carbonate minerals in the
Bendigo and Mother Lode systems was reported by Lindgren (1905) and
Knopf (1929}, respectively. The subject of carbonate alteration occurring in
conjunction with gold vein deposits has been discussed by several workers,
including Boyle (1961, 1976, 1979), Meyer and Hemley (1967), and Rose
and Burt (1979). In general, there is a concensus that reactions between CO2
and Ca-, Fe- and Mg-bearing silicates in the wall rocks are implicated in
formation of the carbonate minerals.
Au-bearing chemical sediments, like their vein counterparts, typically con-
tain a high proportion of carbonate minerals, and are bounded by flows or
volcaniclastic rocks possessing intense carbonate replacement alteration
(Fryer et al., 1979; Fyon et al., 1980; Kerrich, 1980; Pirie, 1980; Whitehead
et al., 1980; Kerrich et al., 1981). Low-temperature carbonate alteration of
serpentine by COs-rich metamorphic fluids has been reported by Knopf
(1906) and Barnes et al. (1973).
Whereas quartz precipitates in response to cooling of hydrothermal fluids,
in accord with field evidence and direct observation of geothermal systems
{Ellis and Mahan, 1977), all carbonate minerals have a negative temperature
coefficient of solubility, requiring specialised conditions such as boiling for
deposition from cooling solutions (Holland and Malinin, 1979). Hence the
abundance of carbonate and its coprecipitation with quartz from a common
cooling solvent in Au-bearing veins requires explanation.
This paper addresses the question of COs in Au-bearing vein deposits. The
source of carbonate, together with other essential gangue and alteration
mineral constituents, has been evaluated by means of chemical mass-balance
calculations on altered wall rock enveloping auriferous veins within the
Campbell, Con and Giant shear zones, at Yellowknife, North West Territories.
This approach enables discrimination of those chemical components donated
by the wall rocks to veins, from those introduced by hydrothermal fluids. In
addition, oxygen-isotope analyses were performed on whole rocks and mineral
separates in order to identify the ambient thermal conditions and source of
solutions implicated in mineralisation.
In the light of the chemical data, the process of CO2 production in hydro-
thermal reservoirs is discussed, together with reactions involving COs fixation
in carbonates during their coprecipitation with quartz. Factors governing the
K/Na ratio of metamorphic and other hydrothermal reservoirs are evaluated
267

in relation to observed patterns of massive K addition in auriferous veins.


Finally, some possible controls on Au transport and deposition are considered,
with particular reference to ultramafic host-rocks.

GEOLOGICAL SYNTHESIS

Au-bearing vein deposits are located within major ductile shear zones at
Yellowknife. The largest shear zones, the Con, Campbell and Giant, are
10--150 m wide, extending to depths in excess of 1--2 km. Shear zones
transect metabasic volcanic rocks and interflow metasediments at angles of
20~---70 ° in the vertical plane. In detail, the major structures consist of an
anastomosing network of second-order shear zones which envelop lens-shaped
domains of isotropic rock. The tectonic schistosity describes a sigmoidal form
between shear-zone boundaries, and contains a subvertically dipping lineation.
These geometrical relations indicate that the shear zones conform to the
simple shear model of Ramsay and Graham (1970).
After initial development of the shear zones, these structures subsequently
acted as permeable conduits for discharge of hydrothermal fluids which pre-
cipitated the Au-bearing quartz--carbonate veins. Such veins are typically
located within the confluence of second-order shear zones, at the apices of
the intervening domains of undeformed rock. The structures and ore bodies
contained in them have been described by Boyle (1961), Henderson and
Brown {1966) and Breakey (1975). From considerations of the ambient stress
regime for fluid flow along shear zones, Kerrich et al. (1977a) and Kerrich
and Allison (1978) interpreted the Au deposits to have been formed from
discharge of high-pressure hydrothermal fluids during successive episodes of
hydraulic fracturing.

Mineralogy

Basalt is the predominant rock type which the shear zones traverse. In the
vicinity of the Con and Giant Mines, the basalts are within the amphibolite
and epidote--amphibolite facies of metamorphism; actinolite + oligoclase +
epidote are the principal mineral constituents, with subordinate carbonate +
chlorite + quartz present. The texture is hornfelsic and the mineral fabric
isotropic. Within shear zones the basalts are transformed by deformation
into chlorite schists having an intense LS tectonic fabric. Chlorite + epidote
+ albite are accompanied by accessory quartz + carbonate + leucoxene +
sulphides (Boyle, 1961; Henderson and Brown, 1966).
Along discrete hydrothermal conduits in the shear zones, chlorite schists
have reacted with mineralising solutions to form sericite--ferroan dolomite--
chlorite schists. Altered rocks contain variable proportions of native Au,
electrum, pyrite, pyrrhotite, arsenopyrite, stibnite, chalcocite, sphalerite,
sulphosalts and galena (Boyle, 1961). A detailed account of the ore mineral-
ogy in the Campbell shear zone is given by Breakey (1975).
268

e~

¢o

o~ .=
5
8~
,~.~

a o . _~
o, .

~ oa ~

O~
269

ANALYTICAL METHODS

Twenty-eight samples covering the range from parent metabasalts and


chlorite schists, through partially- to highly-altered products at vein margins,
were collected over a 1600-m vertical section of the Campbell, Con and Giant
shear zones, Yellowknife. Major-element oxides, together with Ba, Rb, Sr, Y,
Zr and Nb, were determined by X-ray fluorescence spectrometry (XRF).
Cr, Ni, Cu, Zn, Pb and Cd were analysed by conventional flame atomic-
absorption spectrophotometry. Sc was determined by neutron activation
analysis (NAA). The chemical data are reported in Tables I--III. Isotopic
results are reported as 5~80-values in permille relative to SMOW {Standard
Mean Ocean Water). The overall reproducibility of 61SO-values has averaged
0.18%0 (2o).
Fractionations among minerals are quoted as A-values, defined as:

AA--B = 1000 In aA._ B ~ 6 A - 5 B

where ~A--B is the fractionation factor for the coexisting minerals A and B.

MASS-BALANCE CALCULATIONS

Simple inspection of tables of analyses is not satisfactory for deducing


chemical changes. Since the c o m p o n e n t oxides are constrained to a constant
sum (100%), no single c o m p o n e n t is independent of the others, and an in-
finite number of solutions exists. Gresens (1967) has pointed out that in many
instances comparisons of bulk composition are made on a quantitative basis
with the tacit assumption of constant volume, but this assumption is untenable
if alteration is accompanied by deformation and/or changes of volatile con-
tent. For instance, apparent gains or losses of immobile elements during meta-
somatism, arising from volume decrease and increase, respectively, could be
incorrectly deduced if the volume relations were not known.
Gresens (1967) suggested incorporating specific-gravity data into two-way
mass-balance calculations, such that fixing either volume change or the behav-
iour of one c o m p o n e n t during a reaction provides a unique solution. For any
transformation of parent-rock to altered product, volume factors ( f v ) may
be c o m p u t e d which correspond to the isochemical behaviour of individual
components. Clustering of the f y ' s of several components, which are empir-
ically known to be relatively immobile (e.g., A1203, Sc, Zr), then provides a
rational basis for estimating the volume change of the reaction as a whole.
In the notation used by Gresens (1966), f v = 1 signifies constant-volume
metasomatic alteration of reactants to products; f v = 0.8 and f v = 2.0, for
example, correspond to 20% volume reduction and 100% volume increase,
respectively.
The chemical data reported in Tables I--III have been used to estimate
volume factors for each c o m p o n e n t in the transformation of individual un-
altered wall rock to neighbouring metasomatised counterparts adjacent to a
270

..=

o.

E&

~J

E. =
_

=
o .

cOcO ~

¢x; r - 0 ~ ~
"~ ~ ° O~ ~0
~
~ C~
¢'q' uO ~-
~ ¢~~ u~. c~

< ~$~o-°o'.~o~ ~ ~ 0~
271

._=

E
o,~r-o0.~ or-co o~ ~
o

o
6
8

.-co

~E
.0 ~,

ii Ii

J =

i~<~m~oz~'~J ~ ~ > ~
272

vein. Inspection of the volume factors reveals that, in general, the relatively
immobile elements Al, $c and Zr behave coherently, maintaining approxi-
mately constant ratios to one another over an order of magnitude decrease
in abundance that corresponds to increasing intensity of alteration. Volume
factors for hydrothermal metasomatism have been estimated from the un-
weighted mean of the fv's of these three components.
In order further to test if A1, Sc and Zr behave coherently during hydro-
thermal alteration, and therefore if these elements may justifiably be used to
estimate volume factors as above, a multivariate profile analysis was conducted
{Morrison, 1967). This test may be used to evaluate whether or not the profile
shapes generated by the mean abundances of A1, Sc and Zr in each of several
groups of samples are different. The samples were subjectively arranged into
four groups based on the intensity of development of the alteration minerals,
namely parent-metabasalts (1) and three groups representing increasing esti-
mated states of alteration (2--4). (The group to which a given sample was
assigned is indicated by a superior suffix in Tables I--III.)
In this analysis the text statistic 0 (theta) is given by Cs/(1 + Cs), where
Cs is the largest eigenvalue of the matrix [HE] -~, and H and E are the be-
tween-groups and within-group components of the covariance matrix, respec-
tively. For the example under consideration the number of variables (chem-
ical elements) (p) is 3, the number of groups (k) is 4, and the total number of
samples (N) is 23.
The parameters of the 0-statistic are designated x, m and n, where s is the
minimum of (k - 1, p - 1), and:

m = ~' ( I k - p [ - 1 ) and n : - ~ ( N - k - p)

For the abundance data on A1, Sc and Zr, given in Tables I--IIl, and grouped
as above, the O-statistic is 4.45-10 -2 compared to 4.5-10 -2 for appropriate
values of s, m and n in the Heck distribution of largest eigenvalues, at a
probability of 0.05.
On the basis of this analysis, the profile shapes generated by the mean
abundances of AI, Sc and Zr in the four groups considered are not different
at a confidence level of 95%. The profile levels, however, are significantly low-
er between groups in the sequence of increasing states of alteration, indicating
a similar proportional dilution of these elements by the introduction of hydro-
thermal products. It is concluded that there is an equal retention of these
relatively immobile elements from wall rocks through increasing intensities
of hydrothermal alteration.
Hynes (1980) reports extensive mobility of Ti, Y and Zr during carbonate-
replacement alteration of metabasalts in the Ascot Formation, Quebec. At
Yellowknife, the covariance of AI, Sc and Zr indicates generally immobile be-
haviour of Zr in the presence of CO2-rich aqueous fluids, as discussed above.
Ti is enriched overall in rocks containing abundant carbonate, a pattern cor-
responding to the results of Hynes (1980).
273

Chemical composition of unmineralised parent-rocks

In order to evaluate chemical fluxes accompanying alteration reactions


between hydrothermal fluids and wall rocks, it is necessary to establish the
parent-rock composition. Inasmuch as hydrothermal fluids were focussed
along conduits within pre-existing shear zones, the problem is first to deter-
mine if deformation of the epidote--amphibolite metabasalt country rocks to
chlorite schists within shear zones took place under isochemical and isovolu-
metric conditions. This question has been addressed by Boyle (1961), and
most recently re-evaluated by Kerrich et al. (1977b), who concluded that
deformation involved depletion of SiO2 (maximum ~25%, average ~ 5%),
and Na~ O, together with hydration and decrease of specific gravity from
2.97 to 2.80, at approximately constant volume.
Based on the above considerations, compositions of three parent-rock
types were used for comparison with hydrothermally altered products:
(1) unaltered metabasalts located in the shear-zone wall, adjacent to stopes
from which hydrothermally metasomatised products were collected; (2) from
unmineralised domains of the shear zones, an average of analyses of nine
deformed schists (sample 1C, 12% depletion of SiO2 relative to parent-rock)
reported by Kerrich et al. (1977b); and (3) an analysis of the most deformed
schist obtained from within the shear zones (sample B51, ~25% depletion of
SiO2 ) (Table I). Substantial variations exist between the chemical composi-
tion of " u n a l t e r e d " metabasalts from the shear-zone walls, especially with
respect to the alkali, alkaline-earth and related trace elements. This chemical
heterogeneity may be due to seawater--hydrothermal alteration, during or
shortly after extrusion of the basalts, as discussed in the section on oxygen
isotopes.
The results of mass-balance calculations, based on the three assumed pre-
cursors and the estimated volume factors for metasomatism, are presented in
Tables IV and V. Chemical changes have been ranked according to increasing
estimated volume factors corresponding to increasing proximity to veins. All
reactions for the transformation of metabasalts outside shear zones to hydro-
thermally altered products at volume factors of 1.4 or less involve depletion
of SiO2. This is not considered a likely result inasmuch as the wall rocks inter-
acted with quartz-saturated solutions moving down temperature, under which
conditions SiO2 is precipitated. However, use of the analysis of schists from
within the shear zone (IC or B51) as the precursor for comparisons to altered
products, which is geologically more realistic, leads to gains of SiOz in almost
all transformations considered.
274

.0

+ + + + ~ 1 ++

~. ~ + + + + +~ +

+ +

~',--, 0

+ • ++ +
0'-.-

.~ .~ ._
~ + + + + ++
+

~.~_
+ + ' ! + + +

iSl
.-~ ! + + + + + + + +

c~ + ' + + + ' ~ + + +

~ E

2~E ~ + + + 1 ! I + + +

~..~ ~ + + + ~ + + l

N 7
+ ! + + I ! I + +

> ".,~
u~
o.

<
[.-,
275

T
tO
+ + 1 + + 1 + 1 + + +

~ + + ~ + 1 1 ÷ 1 +
+

T ÷ ++ ++
÷

~ 1 1 ~ + ~ . + + ~
+ + ++ +

~ 1 1 1 + 1 1 ÷÷
+

& & & & l l l l & &

~5
¢-~
+ ÷ ~ ! 1 + 1 ÷ ÷ ÷
c-

~ ÷ l + ÷ ÷ ÷ l & ÷ ÷
÷
O

÷ ÷ + I 1 ÷ + ÷

c-

& & & ÷ l l l l l & ' O


e~

+ + 1 ÷ + + 1 1 + + 1

II
Oo ~_~ ~
276

~ ._~ o

"~
o.o.~.~.m.~.~.,2.~
~r~- 0
i ,-..~
0 0
+
0 0
+
~0 ,-.-~ oO u~ u'~
,---, +
' ~.~
-~ ~,~ C~
o.
~.~..

L~
*"1 ~, t.~ t'--- Ol O0 0"~ ~ t~ 0 Le'a e ~
.~ .~_ ._~
C"~ + + + + ,.-~ + + ~
~1 + + +
" ~ . ~ .~
~ t'-,

~'~,

g 0

e- 0 g;
+ ++
.~ ~ +

~ O ~ ~ t ~ 0 0 ~

~o
+ + ' + ÷ +
+

LO

Ol
e...~

~ + + : ; ! I I + + ~
s° s +

t~
~ t ~ O ~ ~ ~ ~
~ 0 ~ 0 ~ ~ ~ 0 0
~0
c~ + ~ + + +

- ~ o ~ ÷ ~ I I + + +

>

o~o~OOOooo o ~ ~
8~g
277

G~

+ + + +

~ ~ ~ ~ 0
c~
T
~ l l i + + + l + + ~
+ +

0
¢'?

G + + 1 + + 1 + 1

t ~ 0 0 0 0 0 ~ 0 0 0 . ~ ~
+ : I + I + + ! +

.o
c~ 0 ~ ~ 0 ~ 0 0 ~

T
+ + + 1 + 1 1 + +

u~
t~
~ ~ o ~ ~ ~
.?
+ + + l + + + +

c~
t~

0
..~

t~
?
O 0 0 0 0 0 .~-~ ~ ,-d .~4 .-.~ , , - ~ 0 0 .r-q 0
+ ! I + +

II
< _~°oo
° o~ =~ ~o~'~
o. .~. .d ' d d . . . . .< .~
E~
278

,nter~s, l y of ~I- ,ntenslt~ of ~II .',tens,t~, ~t


-'
o'.e,o,
, oo q .•' e ~
0 , ~0~ .-
qi*erof,,,~

JLi_ _L •

-E

i -

ar-~rw~J-i
.c jr,.e fcc~or ,c,orne foe",:' vOIbme tO:*b"
i t ¢ i 1 i i i , , t e ~ t ~ : , , ~ ¢" ¢ i ¢ t ,
279

r'tensdyof .~ ,r'tensdy of ), ,ntensdy .of


® aiterot,on clterotton a:~era ' on

,!~ Fe2O, ,~ K20 ,~o] S,02

42

22

- 20

~r Al~tO3

g
6 6C :
~-. 4 3 4.5 • 4C

i.; 2 ' 2 " 20

-29"
"gO
• - u ~ "

-3=: -SS:

,×%-
Ni Sc

- I00 " - 30 -,OC


~clume ~OClOr ~91,~me~oc'ot lolu~ 'octor

c~r2t / 8 " !e : ' ~ : c~-2 [ ' ' c o l % ' 7961 b ,


072 i ~~ , i ~18 , , ~

Fig. 1. Gains and losses o f major-element o x i d e s and selected trace e l e m e n t s in: ( A ) Campbell
and Giant shear z o n e s ; and (B) the Con shear z o n e , as a f u n c t i o n o f increasing i n t e n s i t y
o f h y d r o t h e r m a l alteration ( i n t e n s i t y measured by f v -- the v o l u m e factor). Gains and
losses e x p r e s s e d as a percentage o f a b u n d a n c e in the assumed parent-rock.
280

RESULTS

Gains and losses of major-element oxides involved in hydrothermal a l t e r a


tion, expressed in grams per 100 grams of parent-rock, are compiled in
Tables IV and V; percentage gains and losses are illustrated in Fig. 1.
The behaviour of all major-element oxides and trace elements during hydr~
thermal alteration is broadly similar in direction of transport and absolute
magnitude for the three assumed precursors (excepting SiO2 at volume
factors of 1.4 or less, as discussed above). In addition, inferred chemical chang,,
are essentially the same for the Con, Campbell and Giant shear zom~,,~,at
equivalent estimated volume factors (Tables IV and V; Fig. 1).
Volume factors range from 0.8 at the periphery of alteration domains to . ;
at vein margins where there is a significant admixture of gangue minerals t,,,
the metabasic schists.
The salient chemical changes accompanying hydrothermal alteration are a.~
follows. Additions of SiO2 and volatiles occur in the majority of transforma
tions considered, which is consistent with precipitation of quartz from cooii,~
solutions, accompanied by hydration and fixation of CO2 in carbonates.
According to Boyle (1961), CO: increases from 1~/~ in umnineralizcd basalts
in the epidote amphibolite facies to 8~,;~in sericite--carbonate schists adjacem
to veins. CaO and MgO are depleted at volume factors of 1.3 or less, I)ut arv
added at f v -> 1.3. Na20 is depleted and K20 fixed over the entire range o1'
volume factors. Small positive and negative variations of total Fe occur at
fg < 1.7, with additions a t / v ~" 1.7. Al,~O.~, Sc and Zr are conserved at all
slates of hydrothermat alteratiom
Precious metals, together with Cr, Ni, Cu, Zn, Pb, Rb, Cd and l~a, have
been added from hydrothermal solutions to the veins and altered wall rocks:
V and Nb exhibit no large departures from background abundances. Rb and
Sr behave sympathetically with K and Ca, respectively (Fig. 1).
'['he simplest interpretation of these results is that the hydrothermal fluids
contribute SiO2, H:O * CO2, metals and K20 to the wall rocks, wiaich is
reflected in the presem:e of vein quartz, elevated metal abundances, hydra-
tion and carbonation of wall rocks, together with pervasive development or
sericite. Such additions to the chlorite schist precursor require volume factor.,
>1.
At increasing distances from veins, in the order of 2--15 m, additions becom~
progressively smaller as fluid penetration into wall rocks decreases, and volume.
factors correspondingly approach unity. In the peripheral regions of alteratio~
envelopes, depletions of CaO and MgO exceed additions of SiO2, K2 O,
volatiles, etc., such that volume factors are less than unity -- i.e. leaching pr~.
dominates over precipitation.
It is suggested that some Ca and Mg (also possibly some Fe) are taken int()
solution in the peripheral alteration regions where f v < 1, and combine with
CO2 contributed by the hydrothermal fluids to yield carbonate minerals in
regions of alteration where f v > 1, m~d in veins (Fig. 1). Cr, Ni, Sr and
281

sulphur may behave in the same manner. Similar patterns of alteration have
been noted for hydrothermally metasomatised mafic rocks adjacent to gold
veins at the Dome Mine, Timmins (Fryer et al., 1979); and at Red Lake by
Crocket et al. {1980), Kerrich (1980), and Pirie (1980).

DISCUSSION OF CHEMICAL MASS BALANCE

Important early studies of the geochemistry of altered rocks adjacent to


Au-bearing veins at Yellowknife were conducted by Boyle (1961). Boyle sug-
gested that Sb, As, Au, silica, transition metals and other elements were
liberated during carbonate replacement alteration of rocks, and that this
process is an important factor in the contribution of components to certain
Au-bearing lodes. This premise is based on simple comparison of element
abundances in carbonated rocks and their precursors, taking into account
changes of specific gravity (Boyle, 1961; 1976, table 7 and p. 23) but not pos-
sible large changes of volume.
Under the conditions of large volume increase required for carbonation
and veining, involving additions of H2 O, CO2 and SiO2, together with decrease
of specific gravity, all relatively immobile elements of the parent-rocks will
appear to undergo a similar proportional decrease in abundance by dilution
--e.g., A1203, Sc, Zr (Tables I--III). The analytically measured abundances
of Cr, Ni and other elements may decrease from country rocks to carbonated
alteration zones (cf. Boyle, 1961, pp. 61, 66; and Tables I--Ill), but if, as
demonstrated above, volume factors accompanying carbonate alteration may
exceed 7 then i n t r o d u c t i o n of these elements into the system is required
{Fig. 1). However, the apparent decrease of A1203, Cr, Ni, etc., in carbonated
alteration zones to veins is treated as an absolute reduction in abundance by
Boyle (1961; 1976, table 7); use of these inferred chemical gradients to de-
duce the direction or magnitude of chemical fluxes from wall rocks into veins
during mineralisation is not valid.
Observations of Au-bearing vein deposits in Archean greenstone belts, to-
gether with examination of the literature on this subject, reveals an empirical
relationship between carbonate composition and nature of the wall rock
which the vein transects. In.general, the predominant carbonate minerals
are ankerite plus ferroan dolomite in mafic rocks, magnesite + Fe-dolomite
in ultramafic rocks, calcite in felsic rocks and siderite plus rhodochrosite in
iron formation. Transitions in wall rock over the length of individual veins
are reflected by corresponding changes in carbonate gangue composition
(Bain, 1933). These observations support the deductions from chemical mass
balance that the wall rock dominates supply of the bivalent metal cations,
whereas the hydrothermal fluids donate CO2, quartz, plus precious metals to
the veins and their carbonate altered envelopes.
282

O x y g e n - i s o t o p e relations and r e d o x c o n d i t i o n s

Unmineralised shear zones. By evaluating geochemical parameters sensitive t(J


w a t e r - r o c k interaction such as oxygen-isotope ratios and the oxidation stat~
of Fe, it is possible to place constraints on the ambient physical conditions
under which the Au-bearing veins were precipitated, the source and redox
potential of the hydrothermal fluids, and an indication as to whether tl~e
chemical system was rock dominated or fluid dominated. 5'SO-values for
basaltic country rocks, their altered products, and veins are reported m
Table VI, together with data on the oxidation state of Fe synthesized from a
previous study (Kerrich et al., 1977a).
In traverses across unmineralised sections of the Campbell and Con shear
zones, from isotropic epidote--amphibolite facies metabasalts through the
deformed chlorite schists, the whole-rock 51SO-values remain approximately
constant at +7 to +7.5% 0, and Fe2÷/,VFe is constant at ~0.7 (Table Vl). These
data imply that deformation took place under conditions of low water/rock
ratio.
Taylor (1968) has shown that fresh, unaltered submarine basalts have
whole-rock 81SO-values of +5.5 to +6.5%0. The slight enrichment in 180
recorded for these volcanic rocks relative to primary submarine basalts is
attributed to oxygen-isotope exchange with seawater at low temperatures,
during thermally driven convective cooling of the ocean crust.

Mineralised shear zones. In proximity to veins, the whole-rock 61SO and Fe :~


~:Fe ratios exhibit a major perturbation from background values of -+7°/00
and 0.7, respectively, (Au <~ 4 ppb), to +9 to + 1 0 ° 0 and an average Fe~÷/2:Fe
of 0.92 in carbonate--sericite schists enveloping Au-bearing quartz veins
(Au 1--40 ppm). The 5180 of vein quartz is equal to that of quartz extracted
from the carbonated mafic schists, to within the limits of precision (Table VI}.
implying oxygen-isotope equilibrium between hydrothermal fluids and altered
wall rocks.
In general, studies of metamorphic processes indicate that ferrous/ferric
ratios undergo little change during regional metamorphism, but this is not the
case in environments where high water/rock ratios were involved. Oxidation-
reduction reactions under high-temperature metamorphic conditions occur
d o m i n a n t l y in response to the dissociation of water, according to the reaction
2H~O -+ 2H2 + O2. The hydrogen removed corresponds to the difference in
equilibrium hydrogen concentration at two different temperature and pres-
sure states. The predominance of Fe ~+ in mineralised wall rocks is believed to
result from the reduction of Fe 3÷ in silicates and metal oxides by hydrogen
during interaction with a large flux of H~-bearing aqueous solutions ascending
along a Tp gradient through conduits within the shear zones. Note that for an
aqueous fluid at high temperatures, the equilibrium PH, >> 2Po~ on account
of reactions of the type Fe 2. silicate + H:O -+ Fe304 + silicate + H2. Minimum
integrated water/rock ratios of ~ 3:1 are indicated by the shift in redox state,
283

e~ + + + +
o

¢.

0 0 0 0

e~

0
0

-d

~ 0 ~

% + + + +

0 0 0 0 0 0
~-~<

e~
0

~ 0 ~
Lt~
+ + + + + Ob
+ 0
e~
._~
°-

e-

e~ ~ 0 ~
._~ Oq

CU 0

0
g 0

0
¢. N
:.=
C
¢,L

E
~'~
.~ E
E m E
<
O.
284

and values in excess of 30:1 may be deduced from the abundance of quartz
(Kerrich et al., 1977a).

THERMAL CONDITIONS DURING MINERALISATION

Ambient thermal conditions during mineralisation may be estimated from


two independent lines of evidence. The coexistence of abundant chlorite with
almost pure albite (Ab92) in wall rocks adjacent to veins provides an upper
limit of ~ 4 7 0 ° C for the ambient temperature (Liou et al., 1974).
Oxygen-isotope fractionations between coexisting minerals may be used
to calculate the temperature at which the minerals attained isotopic equilib-
rium. Temperatures of mineralisation at the Con Mine, Yellowknife have
been calculated from &quartz--muscovite-values with reference to: (1) the
experimentally determined equations of Clayton et al. (1972) and O'Neil and
Taylor (1969) for quartz--water and muscovite--water, respectively; (2) the
equations derived by Bottinga and Javoy {1973}; and (3) the empirically
calibrated quartz--muscovite equation of Blattner (1975). In addition, tempera.
tures have been calculated from quartz--chlorite fractionations, based on the
equation for chlorite--water given by Wenner and Taylor (1971).
Q u a r t z - m u s c o v i t e fractionations in previous metal-bearing veins exhibit
a narrow range from +3.8 to +3.4%0 (Table V). For these fractionations the
three sets of equations referenced above yield calculated isotopic tempera-
tures of 360--430°C, 400--480°C and 420--450°C, respectively. Q u a r t z -
chlorite fractionations are +5.8 to +6.2%0, representing isotopic tempera-
tures of 440--480°C. Errors in the isotopic temperatures, arising from the
analytical uncertainty of + 0.18%0, amount to about ± 30°C. These calcu-
lated temperatures must be treated with caution because concordant isotopic
equilibrium among co-existing mineral triplets is a necessary condition for ob-
taining reliable temperature estimates. However, the results are consistent
with independent estimates given above.
If the ambient mineralisation temperatures of 400--450°C derived above
are broadly correct, then the calculated 5 ' 8 0 of fluids from which the veins
were precipitated is +8 to +9%o [based on the quartz--water equation of
Clayton et al. (1972)]. These values are consistent with the range in 5180 of
+ 4 to +20%0 for most fluids implicated in metamorphism, and overlap the
primary magmatic range (+6 to +8%0 ) (Taylor, 1974). A magmatic origin of
the fluids is considered unlikely inasmuch as quartz veins in the nearby West-
ern Granodiorite batholith have 5 ' 8 0 quartz of + 8 0 0 and Au abundances of
only 10--100 × background.
In summary, early transformation of epidote--amphibolite metabasalts tu
chlorite schists within the shear zones by deformation is considered to have
taken place under conditions of low water/rock, and the chemical system was
essentially closed except for short-range transport of SiO2 and Na20. Con-
versely, subsequent modification of the chlorite schists to carbonate--sericite
schists along hydrothermal conduits occupied by veins involved large fluxes of
285

reducing metamorphic fluids at 400--450°C, and extensive chemical exchange


between fluids and wall rocks. Based on these results, the origin of the Yellow-
knife deposits conforms to the model proposed by Fyfe and Henley (1973)
for Au-bearing veins, involving focussed discharge of metamorphic fluids along
shear zones. Clearly, the temperatures in the fluid source regions must have
been hotter than in the veins, and could have arisen from the thermal aureole
of the Western Granodiorite batholith intruding into hydrated sea-floor basalts.

DISCUSSION

It is generally accepted that metals and gangue constituents in veins have


been transported by hydrothermal solutions ascending through conduits in
the crust, with precipitation occurring in response to cooling of the solvent
along a geothermobar (Helgeson and Garrels, 1968; Fyfe and Henley, 1973;
Holland, 1967). This premise is entirely consistent with predictions from
solubility data, inasmuch as quartz, together with Au, Ag and W (a metal
association characteristic of these deposits) all have a positive temperature
coefficient of solubility (Fyfe and Henley, 1973; Seward, 1976; Foster,
1977; Holland and Malinin, 1979). Other parameters such as pH and wall-rock
reactions may also influence the solubility of some aqueous species (Meyer
and Hemley, 1967}.
The solubility relations of carbonate minerals are more complex than for
quartz (Holland and Malinin, 1979). In general, Calcite solubility decreases
as a function of increasing temperature (T) at constant PC02, or diminishes
with decreasing PCO2 at constant T. The negative temperature coefficient
of solubility dominates the positive effect of pressure, such that carbonates
become less soluble with respect to increasing temperature, and carbonate
precipitation is not anticipated from cooling fluids under typical crustal P--T
cooling paths. Specialised conditions such as boiling of COs are required for
carbonate deposition from cooling solutions. Thus the abundance of carbonate
and its coprecipitation with quartz in Au-bearing vein deposits requires ex-
planation.
In contrast, the distribution of quartz and carbonate minerals in convec-
tive hydrothermal systems is as predicted from simple temperature--solubility
relations. Carbonate veins are present in the peripheral recharge areas where
surface fluids heat during penetration into the crust; whereas quartz is the
dominant gangue mineral in vein stockworks which represent the locus of
hydrothermal discharge of fluids cooling during return flow to the surface
(Andrews and Fyfe, 1976; Spooner et al., 1977; Fyfe and Lonsdale, 1981;
Meyer and Hemley, 1967). Examples of such hydrothermal systems include
seawater circulation through hot basalts at mid-ocean ridges, and convection
of meteoric fluids in proximity to Cu porphyries (Lowell and GuiJbert, 1970;
Sheppard et al., 1971; Wolery and Sleep, 1976; Fyfe and Lonsdale, 1981).
A similar distinction exists between auriferous chemical sediments on the
one hand, which typically contain abundant carbonate, such as at Timmins,
286

and Red Lake, Ontario, and the Lapa Seca, Brazil; and, on the other hand,
base-metal-bearing sediments in which the carbonate content is not generally
significant.
Several explanations have been advanced to account for the precipitatio~
of carbonate minerals from cooling fluids, as discussed by Holland (1967):
additional processes are discussed in the section on the source of CO:. An
important factor is believed to be boiling of solutions, which will take place
during fluid ascent at the point when the confining pressure becomes less
than the vapour pressure of the solution. This condition will be met in prox-
imity to the terrestrial surface (< ~ 2 km); and boiling of hydrothermal solu-
tions has been identified during stages of calcite deposition at the shallow-
level Providencia Mine, Mexico (Sawkins, 1964).
There is no direct evidence available on the ambient hydraulic pressure, or
crustal depth of mineralisation at Yellowknife. However, based on the premise,
that fluids were derived by outgassing reactions under greenschist-facies
conditions, or at the greenschist--amphibolite transition, and assuming a geo-
thermal gradient of 50°C kin-~ with/)fluid :>Plithostatic, fluid pressures and
depths of 8 kin, 2 kbar (200 MPa) to 10 km, 2.5 kbar (250 MPa) are indi-
cated. These conditions are well above the vapour curve for 6-wt. ¢2~NaCI-
H:O--CO: solutions (Takenouchi and Kennedy, 1965). Furthermore,
inspection of primary fluid inclusions in vein quartz from the Con Mine has
not revealed evidence for boiling of the hydrothermal fluids (this work).
In silicate host-rocks, reactions involving hydrogen metasomatism (Hemle)
and Jones, 1964) may well influence the deposition of carbonate via their
effect on the pH of the hydrothermal solutions. Many of the important
reactions in the alteration envelopes around hydrothermal veins involve the
loss of H ÷ from the solutions and its exchange for K ÷, Na ÷, Ca ~ and Mg 2÷ m
the c o u n t r y rock, as exemplified by the alteration described above at Yellow-
knife, or as Holland I1967} observed:
"Such a reduction in the H* concentration leads to an increase in the CO: concentra
tion at a given total-carbon-species concentration, and can therefore lead to the precip~
itation of carbonate minerals."

The source o f CO:

During normal prograde regional or burial metamorphism, fluid pressures


are in the same order of magnitude as lithostatic pressures (Fyfe et al., 1978).
Except in environments where partial melting occurs, total fluid pressure is
close to the sum of partial pressures of HsO and CO:, the dominant volatiles.
If the mixtures are considered ideal, the partial pressures are directly related
to the mole fractions. As the average crustal content of carbon is around 200
ppm while the water c o n t e n t is close to 20,000 ppm, it is clear that water is
the d o m i n a n t volatile in metamorphism (Fyfe et al., 1978).
For given lithostatic pressures and temperatures, studies of the chemistry
287

of fluids (see Ellis and Mahon, 1977), and experimental studies of carbonate
and hydrate equilibria (see Winkler, 1976; Fyfe et al., 1978} clearly show that
for most temperatures up to the amphibolite facies PH~O is much greater than
PC02. Only under extreme conditions as for the granulite facies (Janardhan
et al., 1979) does CO2 become a major or even dominant volatile species.
This observation is in accord with the great thermal stability of relatively
pure carbonate rocks. Most work on metamorphic rocks shows that given the
very low porosity and permeability of rocks at depth, massive degassing pro-
ceeds only where the geothermal gradient intersects the equilibrium vapor
pressure curves when the latter approach the lithostatic pressure. Thus for pure
dolomite or siliceous marbles temperatures in excess of 800°C (migmatite
temperatures) are required to achieve conditions PCO~ ~ Plithostatic-
The two major situations where massive fluid motion and chemical trans-
port occur in the crust (Fyfe, 1978) involve the cooling of igneous bodies by
convective circulation of surface waters (continental or submarine), and during
metamorphic dehydration reactions. In the former case, the mass of fluid is
controlled by the energy of the system and the ratio of water to rock is large
(10--100, see Wolery and Sleep, 1976). In the metamorphic case, the fluid
volume is simply a function of the water content of the rocks undergoing
regional metamorphism, and in general the water/rock ratio will not exceed
0.1. Frequently, and particularly in Archaean terrains, the cooling of igneous
rocks and metamorphic processes may be closely related in space and time and
may lead to the closely-associated compressional--tensional tectonic patterns
associated with crack formation, vein filling processes and ore formation.
Whether or not large-scale carbonate veins form, or massive-carbonate-
forming reactions with wall rocks occur, ultimately depends on the CO2 con-
tent of the metamorphic rocks at depth. A few examples suffice to illustrate
the process. If we consider the simple case of fluid circulation through a
siliceous limestone in the system CaCO3--SiO2, the CO2 pressure in the car-
bonate rock will be buffered by the calcite ~ wollastonite reaction and PCO2
will be at low values (less than 100 bar) until temperatures exceed 500°C. For
a surface-water-convecting system (Pfluid -~ 1/3 Plithostatic) the mole fraction
of COs in the discharge fluid will be low and will depend on the degree of
dilution by the circulating fluid and its COs content in the environment above
and below the limestone. If it is flowing through rocks with little carbonate,
it may, or may not, precipitate carbonates on the discharge path, depending
on the degree of dilution. If the host-rocks contain dispersed carbonate
(calcite or dolomite) at the percent level, PCO~ will always be buffered by
appropriate reactions with silicates and when this fluid discharges it must
precipitate carbonates, the COs-forming reactions must be reversed by wall-
rock interaction. The two situations are well represented by the carbonate
veins described here, or the characteristic, almost pure, quartz veins of the
Otago greenschists described by Hutton (1940). In both cases Au concentra-
tion may occur.
Another typical situation occurs where seawater convects through basaltic
288

rocks of the sea-floor crust. It is now well established that carbonates are
precipitated during recharge and C02 is lost from the input fluids (Fyfe and
Lonsdale~ 1981). When this fluid (high water/rock ratios) discharges after
reaction with almost carbonate-free, deeper, basaltic materials, it is unlikeb
that carbonate veins will form and simple quartz veins will dominate the d i s
charge systems [one of the authors (W.S.F.} has noted such discharge veins
in Liguria, Italy and the Oman ophiolites].
In the general case of regional metamorphism and dewatering of a pile ol
complex lithology including marine sediments and spilitic rocks, the primar~
carbonate content is likely to be significantly greater than 200 pt)m. In
Archaean sections where altered submarine volcm~ic rocks are dominant,
again the CO2 content is likely to be large via fixation of marine carbonate
into cooling basalts. Jolly (1974) indicates that calcite is ubiquitous in the
low-grade meta-volcanies of the Archaean Abitibi belt in Ontario. Aumento
et al. (1976) indicate that CO2 may reach 3% in heavily-altered sea-floor
basalts, and Hallberg et al. (1976) report CO2 contents of 0.1--0.3 wt.',
even for relatively unaltered Archaean basalts.
In such environments, given thermal gradients in the order of 30°C km '.
PH~O will rise ~ 280 bar km ~. At 300°C (10 km), Pfluid will be in the order
of 2800 bar while PCo~ will be no more than 100 bar. At 500°C (17 kin).
Pfluid will approach 5000 bar with PCO~ in the 1000-bar range. At Y e l l o ~
knife the presence of sphene in epidote---amphibolite facies metabasalts.
formed via the reaction:

CaTiSiOs + CO2 ~ CaCO, + TiO2 + SiO2


sets the xCO2 of metamorphic fluids to Q 0.5 at 500°C and 2 kbar ( Hunt and
Kerrick, 1977).
It is clear that if this metamorphic fluid moves upwards and remains m
equilibrium with Ca--Fe--Mg-silicates, it will precipitate carbonates by wall-
rock alteration until virtually all CO2 is removed. Inspection of carbonate
equilibria (Winkler, 1976; Fyfe et al., 1978; Turner, 1981) indicates that
maximum carbonate formation will occur in the greenschist-facies region if
the source-rocks have moderate CO2 contents. In summary, Au and quartz
precipitation reflects diminished solubility related to temperature, while
carbonate precipitation is related to the above-equilibrium CO2 pressures
relative to wall-rock alteration equilibria (a similar situation to reduction
reactions related to excess H2 pressures discussed above). Some typical car-
bonate-forming reactions that may proceed via the hydrolysis of silicates in
wall rocks are listed in Table VII.
Compilations of the disposition of metamorphic facies in greenstone
belts (Jolly, 1978) reveal that greenschist-grade rocks predominate whereas
rocks at amphibolite facies are rare. These observations signify that the
metamorphic fluids implicated in Archaean lode Au mineralisation (Fyfe
and Henley, 1973; Kerrieh, 1980; Kerrich and Fryer, 1981; this paper} may
289

TABLE VII

Hydration and/or carbonation reactions proceeding within the shear zones

3 N a A I S i 3 0 . + K* + 2 H " ~ KAI3Si~OIo(OH) ~ +6SiO 2 +3Na*


albite muscovite

3(Mg,Fe)4Al:Al:Si:Oto(Oti)~ +6Ca2AI~Si~O:2(OH)+ 6 S i O : + 2 4 C O : + 10 K ÷
aluminous chlorite epidote

1 0 K A I , S i , O , , , ( O H ) 2 + 12Ca(Mg,Fe)(CO~)~. + 10H*
muscovite ferroan dolomite

be derived from dehydration reactions within the greenschist facies as well


as at the greenschist--amphibolite transition.

Significance of potassic alteration

Massive introduction of K is a salient feature in the domains of alteration


enveloping lode Au deposits as discussed above, and also in the high-tempera-
ture potassic zones of Cu porphyries. By contrast, Fe--Mg addition dominates
in the lower-temperature footwall alteration of base-metal massive sulphide
deposits, where K fixation is not pronounced.
If the metamorphic fluids implicated in Archaean lode Au deposits were
at temperatures in the order of 400--450+°C, then the source regions were
almost certainly within amphibolite facies conditions. In rocks undergoing
amphibolite facies metamorphism an almost pure K phase {muscovite) is
stable, but a pure Na phase (albite) is not present; rather Na is incorporated
in amphiboles and plagioclase. A fluid outgasscd under amphibolite facies
conditions will not precipitate a pure Na phase until there is substantial
cooling, but it can precipitate muscovite or biotite. Thus when Al is released
from the hydrolysis of Fe-, Mg- and Ca-silicates during alteration by COz-rich
aqueous fluids, the K/Na ratio will initially favour formation of muscovite
because the potassium level is at greater supersaturation with respect to
muscovite than the Na concentration is in relation to paragonite or albite.
In accord with the above reasoning are data from analyses of liquid inclu-
sions (Roedder, 1972) which indicates that high-temperature fluids possess
K/Na ratios much larger than for seawater: the latter with Na/K ~ 28, favours
albite formation at temperatures above 150°C (Munha et al., 1980). Data for
auriferous hydrothermal solutions are available from studies conducted at
the Lamaque Mine, Quebec, where the Na/K of liquid inclusions rises as Au-
bearing fluids pass through the alteration zone (Ritter, 1971), and Na is
taken into solution via the hydrolysis of albite with concomitant fixation
of Kaqueou s into muscovite. Further evidence has been obtained from the
O'Brien gold mine, Quebec, where Krupka et al. {1977) report Na/K ~ 1 for
liquid inclusions in quartz associated with the main stage of mineralisation.
290

A possible corollary to the typical potassium alteration in lode Au deposits


is that in sufficiently low-temperature regions of the system, when fluids have
become enriched in Na and depleted in K, as discussed above, the Na/K ratio
of fluids may rise to a level where albite becomes stable. Limited observa-
tional evidence for this exists in the upper levels of the Con Mine, Yellowknife
and Lamaque Mine, Val d'Or (this work).

Gold deposits in ultrama[ic rocks

It is interesting in the context of the foregoing discussion to consider the


question of Au deposits located in rocks of ultramafic compositions. Pyke
{1975, 1976) has noted a relationship of Au mineralisation and ultramafic
volcanic rocks in the Porcupine camp, Ontario, claiming that ultramafic rocks
are favourable "source beds" because of their intrinsically elevated Au abtlll-
dance. Analysis of Au by sensitive m o d e m techniques reveals that unaltered
ultramafic rocks have abundances of 0.5--2 ppb Au, in the same rmlge as
other primary igneous rocks (Tilling et al., 1973; Anhaeusser et al.. 1975:
Kwong and Crocket, 1978}: these data contraindicate the "source-bed"
concept.
There are two further aspects to the source-bed problem; namely the
regional geological association and local causal factors. Although prominent
Au deposits are physically related to ultramafic rocks, a case can be made
for Au associations with a multitude of "special" rock types. In the Porcupim,
and Val d'Or-Malartic camps, and at Agnico Eagle, Quebec, Au is related to
Na-enriched felsic intrusive-extrusive igneous rocks. At the Kirkland Lake.
Larder Lake and Matachewan camps, Ontario, Au is hosted by syenites.
Similarly, the presence of Au in granodiorites is notable at Lamaque, Quebec:
and Dryden, Ontario. However, as Hutchinson (1976) has pointed out, the
overwhelming preponderance of Au produced from Archaean lode deposits
is from mines located within mafic igneous rocks. There is thus no geological
basis for claiming a special relationship of Au with ultramafic rocks.
At the local scale, ultramafic rocks may provide favourable chemical en-
vironments for Au precipitation. In a typical granite, basalt and ultramafic
rock, the sum of cations i Fe, Mn, Mg and Ca) which may combine with CO,
to form carbonates is -- 4, 21, and 32 at. %, respectively. From this it is clear
that ultramafic rocks have the greatest capacity to undergo intense carbonate
replacement alteration, and thus to exert the greatest integrated chemical
effect on fluids exchanging with the rocks via scavenging of CO2. If wall-rock
reactions may locally mediate precipitation of Au [ for which there is abun-
dant empirical evidence, cf. Bain (1933)l, in addition to temperature (Fyfc
and Henley, 1973), then it is possible that some of the Au may be transported
as a carbonyl (CO) or carbonate complex (CO23-), with carbonate-forming
reactions triggering Au precipitation by reduction of PCO, and {or) decrease
of acidity. This aspect of the problem is currently under investigation, but it
is well known (see Ephraim, 1948, pp. 776--810) that many metals capable ol
291

forming carbonyl and carbonyl--chloride complexes are also enriched in lode


Au deposits. The presence of "metamorphic" Au-bearing quartz veins in
greenstone belts may also partly reflect the high capacity of mafic and ultra-
mafic rocks to store volatiles during early sea-floor spilitisation reactions, tl~e
volatiles being released subsequently during crustal heating, as discussed above.

ACKNOWLEDGMENTS

The authors thank Cominco Limited for providing access to the Con Mine,
and D. McMurdo and D. Myers for cooperation given during collection of
material for this study. We express our thanks to B.J. Fryer, B.E. Gorman,
R.W. Hodder, and W. Nesbitt for critical reading of a provisional draft, and for
suggesting improvements. Analytical assistance provided by M.G. Peirce, Y.C.
Cheng, J. Forth and B.J. McKinnon is acknowledged with gratitude. Dr. L. Orloci
kindly assisted with statistical treatment of the data. Both authors received sup-
port from Natural Research Council of Canada grants.

REFERENCES

Andrews, A.J. and Fyfe, W.S., 1976. Metamorphism and massive sulphide generation in
oceanic crust. Geosci. Can., 3: 84--94.
Anhaeusser, C.K., Fritze, K., Fyfe, W.S. and Gill, R.C.O., 1975. Gold in "primitive"
Archaean volcanics. Chem. Geol., 1 6 : 1 2 9 - - 1 3 5 .
Aumento, F., Mitchell, W.S. and Fratta, M., 1976. Interaction between sea water and
oceanic layer two as a function of depth and time, I. Field evidence. Can. Miner., 14:
269--290.
Bain, G.W., 1933. Wall-rock mineralisation along Ontario gold deposits. Econ. Geol., 18:
705-743.
Barnes, I., O'Neil, J.R., Rapp, J.B. and White, D.E., 1973. Silica--carbonate alteration of
serpentine: wall rock alteration in mercury deposits of the California Coast Ranges.
Econ. Geol., 68: 3 8 8 - 3 9 8 .
Blattner, P., 1975. Oxygen isotopic composition of fissure-grown quartz, adularia, and
calcite from Broadlands geothermal field, New Zealand. Am. J. Sci., 273: 785--800.
Block, S. and Bischoff, J.L., 1979. The effect of low-temperature alteration of basalt
on the ocean budget of potassium. Geology, 1 : 193--196.
Bottinga, Y. and Javoy, M., 1973. Comments on oxygen isotope thermometry. Earth
Planet. Sci. Lett., 20: 2 5 0 - 2 6 5 .
Boyle, R.W., 1961. The geology, geochemistry, and origin of the gold deposits of the
Yellowknife district. Geol. Surv. Can., Mem. No. 3 1 0 , 1 9 3 pp.
Boyle, R.W., 1976. Mineralisation processes in Archean greenstone and sedimentary belts.
Geol. Surv., Can. Pap., 75-15.
Boyle, R.W., 1979. The geochemistry of gold and its deposits. Geol. Surv. Can. Bull. No.
280, 584 pp.
Breakey, A.R., 1975. A mineralogical study of the gold quartz lenses in the Campbell
shear, Con Mine, Yellowknife, N.W.T.M.Sc. Thesis, McGill University, Montreal, Que.
Clayton, R.N., O'Neil, J.R. and Mayeda, T.K., 1972. Oxygen isotope exchange between
quartz and water. J. Geophys. Res., 77 : 3057--3067.
Coward, M.P., 1976. Strain within ductile shear zones. Tectonophysics, 34: 181--197.
Crocket, J.H., Cowen, P. and Kusmirski, R.T.M., 1980. Gold-content of volcanic hosted
interflow sedimentary rocks in the Red Lake area: implications on ore genesis at the
Dickenson Mine. In: R.G. Roberts (Editor), Genesis of Archean, Volcanic-hosted Gold
Deposits. Waterloo, March 1980. Ont. Geol. Surv., Open File Rep., 5293: 66--91.
292

Ellis, A.J. and Mahon, W.A., 1977. Chemistry and Geothermal Systems. Academic Press.
New York, N.Y., 392 pp.
Ephraim, F., 1948. Inorganic Chemistry. Gurney and Jackson, London.
Foster, R.P., 1977. Solubility of scheelite in hydrothermal chloride solutions. Chem. Geoi
20:27 43.
Fryer, B.J., Kerrich, R., Hutchinson, R.W., Peirce, M.G. and Rogers, D.S., 1979. Archaeat~
precious metal hydrothermal systems, Dome Mine, Abitibi greenstone belt. II. REE and
oxygen isotope relations. Can. J. Earth Sci., 16: 421--439.
Fyfe, W.S., 1978. Large scale crust h y d r o s p h e r e interaction: large scale ore deposits. N Z
Dep. Sci. Ind. Res., Bull., 2 1 5 : 7 1 75.
Fyfe, W.S. and Henley, R.W., 1973. Some thoughts on chemical transport processes with
particular reference to gold. Min. Sci. Eng., 5:295--303.
Fyfe, W.S. and Lonsdale, P., 1981. Submarine hydrothermal activity. In: C. Emiliani
(Editor), The Seas, 7. Wiley Interscience, New York, N.Y. (in press).
Fyfe, W.S., Price, N.J. and Thompson, A.B., 1978. Fluids in the Earth's Crust. Elsevic,l-.
Amsterdam, 393 pp.
Fyon, J.A., Schwarcz, It.P. aad Crocket, J.H., 1980. Carbon and oxygen isotope geo-
chemistry of replacement carbonates from the T i m m i n s P o r c u p i n e Gold Camp. in:
E.G. P y e ( E d i t o r ) , G e o s c i e n c e G r a n t Programe. Summary of Research 1979 1980.
Ont. Geol. Surv., Misc. Pap., 9 3 : 7 7 82.
Gresens, R.l,., 1967. Composition volume relations of metasomatism. Chem. t;eol.. 2
,17--65.
Hallherg, J.A., Johnston, C. and Bye, S.M., 1976. The Archaean Marda igneous complex.
Western Australia. Precambrian Res., 3 : 1 1 1 - 1 3 6 .
Helgeson, H.C. and Garrels, R.M., 1968. Hydrothermal transport and deposition of ~old
Econ. Geol., 6 3 : 6 2 2 63:-).
Henderson, J.F. and Brown, I.C., 1966. Geology and structure of the Yellowknilc ( ; r e c u
stone Belt, District of Mackenzie. Geol. Surv. Can., Bull. No. 141, 87 pp.
Holland, H.D., 1967. Gangue minerals in hydrothermal deposits. In: H.L. Barnes (Editor),
Geochemistry of Hydrothermal Ore Deposits. Holt, Rinehart and Winston, New York,
N.Y., Ch. 9, p. 382.
Holland, ll.D. and Malinln, S.D., 1979. The solubility and occurrence of non-ore minerals.
In: H.L. Barnes (Editor), Geochemistry of Hydrothermal Ore Deposits. Wiley', New
York, N.Y., pp. 461 .-~08
ltunt, J.A. and Kerrick, D.M., 1977. The stability of sphene; experimental redetermina-
tion and geologic implications. Geochim. Cosmochim. Acta, 4 1 : 2 7 9 -288.
Hutchinson, R.W., 1976. Lode gold deposits: the case for volcanogenic derivation, in:
Proc. Vol. Pac. Northw. Min. Metals Conf., Portland, Oreg., 1975. Oreg. Dep. Geol.
Mineral. Ind., pp. 64--105.
Hutton, C.O., 1940. Metamorphism in the LakeWakatipu region. N.Z. Dep. Sci. lnd Res.
Geol. Mere. No. 5, 8,t pp.
Hynes, A., 1980. Carbonatisation and mobility of Ti, Y and Zr in Ascot Formation mcl~l-
basalts, SE Quebec. Contrib. Mineral. Petrol., 75: 79- 87.
Janardhan, A.S., Newton, R.C. and Smith, J.V., 1979. Ancient crustal metamorphism at
lowPH O: charnockite formation at Kabbaldurga, South India. Nature (Londorl).
278:5~1 514.
Jolly, W.T., 1974. Regional metamorphic zonation as an aid in study of Archean terrains:
Abitibi region, Ontario. Can. Mineral,, 12: 4 9 9 - 5 0 8 .
Jolly, W.T., 1978. Metamorphic history of the Archean Abitibi belt. In: Metamorphism in
the Canadian Shield. Geol. Surv. Can., Pap., 7 8 - 1 0 : 6 3 7 8 .
Kerrich, R., 1980. Archaean gold-bearing chemical sediments and veins: a synthesis o|
stable isotope and geochemical relations. In : R.G. Roberts (Editor), Genesis of Archean.
Volcanic-hosted Gold Deposits. Waterloo, March 1980. Ont. Geol. Surv., Open File Rep.
5293: 137--211.
Kerrich, R. and Allison, I., 1978. Vein geometry and hydrostatics during Yellowknife
mineralisation. Can. J. Earth Sci., 1 5 : 1 1 5 3 1 1 6 0 .
293

Kerrich, R. and Fryer, B.J., 1981. The separation of rare elements from abundant base
metals in Archean lode gold deposits: implication of low water/rock source regions.
Econ. Geol. (in press).
Kerrich, R., Fyfe, W.S. and Allison, I., 1977a. Iron reduction around gold--quartz veins,
Yellowknife district, North West Territories, Canada. Econ. Geol., 72: 657.
Kerrich, R., Fyfe, W.S., Gorman, B.E. and Allison, I., 1977b. Local modification of rock
chemistry by deformation. Contrib. Mineral. Petrol., 65: 183--190.
Kerrich, R., Fryer, B.J., Milner, K. and Peirce, M.G., 1981. The geochemistry of gold-
bearing chemical sediments, Dickenson Mine, Red Lake, Ontario. Can. J. Earth Sci.
(in press).
Knopf, A., 1906. An alteration of Coast Range serpentine. Univ. Calif. Dep. Geol. Bull.,
4: 425--430.
Knopf, A., 1929. The Mother Lode system of California. U.S. Geol. Surv., Prof. Pap. 157.
Krupka, K.M., Ohmoto, H. and Wickman, F.E., 1977. A new technique in neutron activa-
tion analysis of Na/K ratios of fluid inclusions and its application to the gold--quartz
veins at the O'Brien Mine, Quebec, Canada. Can. J. Earth Sci., 14: 2760--2770.
Kwong, Y.T.J. and Crocket, J.H., 1978. Background and anomalous gold in rocks of an
Archaean greenstone assemblage. Kakagi Lake area, northwestern Ontario. Econ. Geol.,
73: 50--63.
Lindgren, W., 1905. Metasomatic processes in the gold deposits of Western Australia.
Econ. Geol., 1 : 530---544.
Liou, J.G., Kuniyoshi, S. and Ito, K., 1974. Experimental studies of the phase relations
between greenschist and amphibolite in a basaltic system. Am. J. Sci., 174: 613--632.
Lowell, J.D. and Guilbert, J.M., 1970. Lateral and vertical alteration mineralization zoning
in porphyry ore deposits. Econ. Geol., 65: 373--408.
Meyer, C. and Hemley, J.J., 1967. Wall rock alteration. In: H.L. Barnes (Editor), Geo-
chemistry of Hydrothermal Ore Deposits. Holt, Rinehart and Winston, New York, N.Y.,
pp. 166 -235.
Morrison, D.F., 1967. Multivariate Statistical Analysis. McGraw-Hill, New York, N.Y.,
388 pp.
Munha, J., Fyfe, W.S. and Kerrich, R., 1980. Adularia, the characteristic mineral of felsic
spilites. Contrib. Mineral. Petrol., 7 5 : 1 5 - - 1 9 .
O'Neil, J.R. and Taylor, H.P., 1969. Oxygen isotope equilibrium between quartz and water.
J. Geophys. Res., 74: 6012--6022.
Pirie, J., 1980. Regional geological setting of gold mineralization in the Red Lake Area,
N.W. Ontario. In: R.G. Roberts (Editor), Genesis of Archean, Volcanic-hosted Gold
Deposits. Waterloo, March 1980. Ont. Geol. Surv., Open File Rep.. 5293: 303--358.
Pyke, D.R., 1975. On the relationship of gold mineralisation and ultramafic volcanic rocks
in the Timmins area. Ont. Div. Mines, Misc. Pap. 12, 23 pp.
Pyke, D.R., 1976. On the relationship of gold mineralisation and ultramafic volcanic rocks
in the Timmins area. Can. Inst. Min. Metall. Bull., 69: 79--87.
Ramsay, J.G. and Graham, R.H., 1970. Strain variation in shear belts. Can. J. Earth Sci., 7:
786--813.
Ritter, C.J., 1971. Trace elements of gold-bearing quartz veins of the Lamaque Mine,
Bourlemaque, Quebec. Ph.D. Thesis, University of Michigan, Ann Arbor, Mich.
Roedder, E., 1972. Composition of fluid inclusions. In: Data of Geochemistry, Ch. JJ.
U.S. Geol. Surv., Prof. Pap. 440-JJ.
Rose, A.W. and Burt, D.M., 1979. Hydrothermal alteration. In. H.L. Barnes (Editor}, Geo-
chemistry of Hydrothermal Ore Deposits, Wiley, New York, N.Y., 2nd ed., pp. 173---235.
Sawkins, F.J., 1964. Lead--zinc ore deposition in the light of fluid inclusion studies,
Providencia Mine, Zacatecas, Mexico. Econ. Geol., 5 9 : 8 8 3 - - 9 1 9 .
Seward, T.M., 1976. The stability of chloride complexes of silver in hydrothermal solu-
tions up to 350°C. Geochim. Cosmochim. Acta, 40: 1329.
Sheppard, S.M.F., Nielsen, R.L. and Taylor, H.P., 1971. Hydrogen and oxygen isotope
ratios in minerals from porphyry copper deposits. Econ. Geol., 66: 515--542.
294

Spooner, E.T.C., Chapman, H.J. and Sinewing, J.D., 1977. Strontium isotope contamina-
tion and oxidation of the ophiolitic rocks of the Troodos Massif Cyprus. Geochim.
Cosmochim. Acta, 41: 873--890.
Takenouchi, S. and Kennedy, A.C., 1964. The binary system H 2O - C O : at high tempera
tures and pressures. Am. J. Sci., 212: 1055--1074.
Takenouchi, S. and Kennedy, A.C., 1965. The solubility of carbon dioxide in CaCI solu-
tions at high temperatures and pressures. Am. J. Sci., 263: 445--454.
Taylor, H.P., 1968. The oxygen isotope geochemistry of igneous rocks. Contrib. Mineral.
Petrol., 19: 1--17.
Tilling, R.I., Gottfried, D. and Rowe, J.J., 1973. Gold abundance in igneous rocks: bearing
on gold mineralisation. Econ. Geol., 68: 168--186.
Turner, F.J., 1981. Metamorphic Petrology. McGraw-Hill, New York, N.Y., 524 pp.
Wenner, D.B. and Taylor, H.P., 1971. Temperatures of serpentinisation of ultramafic rocks
based on ' ~O] ~sO fractionations between co-existing serpentine and magnetite. Contrib
Mineral. Petrol., 3 2 : 1 6 5 - - 1 8 5 .
Whitehead, R.E.S., Davies, J.F. and Cameron, R.A., 1980. Carbonate and alkali alteration
patterns in the Timmins Gold Mining area. In: E.G. Pye (Editor), Ont. Geol. Surv., Misc
Pap., 93: 244--256.
Winkler, H.G.F., 1976. Petrogenesis of Metamorphic Rocks. Springer, New York, N.k~..
334 pp.
Wolery, T.J. and Sleep, N.H., 1976. Hydrothermal circulation and geochemical flux at
mid-ocean ridges. J. Geol., 84: 249--275.

You might also like