You are on page 1of 26

Finding deeply buried deposits using geochemistry

Eion M. Cameron1, Stewart M. Hamilton2, Matthew I. Leybourne3, Gwendy E.M. Hall4


& M. Beth McClenaghan4
1

Eion Cameron Geochemical Inc., 865 Spruce Ridge Road, Carp, Ontario, K0A 1L0, Canada
2
Ontario Geological Survey, Sudbury, Ontario, P3E 6B5, Canada
3
Department of Geosciences, University of Texas at Dallas, Richardson, TX 75083-0688, USA
4
Geological Survey of Canada, 601 Booth Street, Ottawa, Ontario, K1A 0E8, Canada
ABSTRACT: It has become increasingly common for geologists to drill through

100 m or more of cover in search for buried mineral deposits. Geochemistry is one
tool applied to this search, using a variety of approaches, including selective leaching
of soils to extract the mobile component of elements, and the measurement of
inorganic and organic gases. This paper provides an overview of some of the work
carried out by the project Deep-Penetrating Geochemistry, sponsored by the
Canadian Mining Industry Research Organization (CAMIRO), and supported by
26 Canadian and international companies and by the Ontario Geological Survey and
the Canadian Geological Survey. The objective was to provide the mining industry
with information relating to processes that may form anomalies at surface over
buried deposits and to provide comparative data on methods used to detect these
anomalies.
Phase I of the project considered the theoretical and experimental framework for
the movement of material from deeply buried deposits to the surface; much of this
information has come from research on the containment of buried nuclear waste. In
arid or semi-arid terrain, with a thick vadose zone, advective transport, which is the
mass transfer of groundwater or air along with their dissolved or gaseous
constituents, is the only known viable means of moving elements to the surface;
diusion of ions in water or gases in air is orders of magnitude slower. Examples of
advective transport are pumping of mineralized groundwater to the surface during
seismic activity and the extraction of air plus gas by barometric pumping. Both
mechanisms require fractured rock and the interpretation of the derived anomalies
requires consideration of neotectonic structures. In wetter climates, where water lies
close to the surface, a variety of mechanisms have been proposed for creating
anomalies at the surface. Diusion-based models again suer from slow rates of
migration. Electrochemical models show a cathodic zone at the top of a buried
sulphide conductor. Cations are attracted to the cathode, rather than to the surface,
yet metals that most commonly migrate as cations are found to form anomalies at the
surface.
Phase II of the CAMIRO study involved field studies at ten test sites. The test sites
included buried porphyry deposits in northern Chile, a goldcopper deposit in the
Carlin district of Nevada, and volcanogenic massive sulphide bodies covered by
glacial sediments in the Abitibi greenstone belt of Ontario. In all cases anomalies
were found in soils above buried mineralization. It is suggested that anomaly
formation is an episodic and cyclic process, in which batches of metal in
water-soluble form are introduced and the metal is then progressively incorporated
with time into the secondary minerals of soil. Selective leaches have been developed
to dissolve specific phases in the soil to detect these anomalies. We have compared
the results for five selective leaches that are available from commercial laboratories:
deionized water, ammonium acetate, hydroxylamine hydrochloride, Enzyme Leach
and Mobile Metal Ion (MMI) plus one non-selective decomposition, aqua regia. In
addition, the Institute of Geophysical and Geochemical Exploration laboratory in
China has supplied data for four sequential selective leaches: water-extractable,
adsorbed, organic-bound and iron- and manganese-bound. The weakest leaches
dissolve mainly the most recently introduced metals that remain in water-soluble
Geochemistry: Exploration, Environment, Analysis, Vol. 4 2004, pp. 732

1467-7873/04/$15.00  2004 AEG/Geological Society of London

E.M. Cameron et al.


form. Other leaches dissolve specific secondary minerals, such as carbonates, or iron
and manganese oxides, which contain the introduced metals. The usefulness of
leaches that dissolve secondary minerals depends on the ratio of introduced
(exogenic) metal that the minerals contain relative to that of endogenic origin derived
from the primary minerals of soils. Our results indicate that this ratio is variable from
site to site, so that there is no universal best leach for dissolving secondary minerals
in exploration surveys. For the test sites in Chile and Nevada, anomalies may have
formed incrementally over a period of a million years or more, which permitted
metals of exogenic origin to become incorporated into many secondary minerals. For
these sites, some anomalies can be detected by aqua regia, although the
anomaly/background contrast is less than for selective leaches. For the test sites in
Ontario, only a few thousand years have elapsed since glacial sediments were
deposited to conceal mineralization. Over this short period, metal of exogenic origin
has been incorporated into only the most labile of secondary minerals and it is the
leaches that dissolve these labile minerals that can successfully identify anomalies. At
the two sites where the most detailed studies have been carried out, the Spence
deposit in Chile and Cross Lake near Timmins, we have found that the optimum
sampling depth in soils is critical to detecting anomalies.
KEYWORDS: exploration, geochemistry, buried deposits, selective leaches, Chile, Nevada, Ontario

INTRODUCTION
The most frequently used geochemical method to identify
buried mineral deposits is selective leaching of soil samples.
These leaches remove only a fraction of the metal that might be
dissolved by aqua regia or by total dissolution, in the expectation that this fraction represents a more readily dissolved
mobile phase, perhaps derived from an ore deposit. Leaches
include the proprietary reagents, Enzyme Leach and Mobile
Metal Ion (MMI), and others that have been developed,
principally by soil scientists, to dissolve specific secondary
minerals in soils, such as carbonates, or Fe and Mn oxides. The
leach solutions are most frequently analysed by inductively
coupled plasma mass spectrometry (ICP-MS). This technique
has shown progressive evolution, so that detection limits of rare
elements have been lowered to levels thought unobtainable a
few years ago. This has had the eect of increasing the number
of elements that can be measured with precision; now 50 or
more are being reported. This benefit also presents a challenge.
With 50 elements at hand and the many ratios that may be
derived from these primary data, it is not too dicult to find an
anomaly over every known deeply buried deposit. The ease
with which data may be obtained has led to the widespread
application of selective leaches, with some successes, but many
dry holes and a poor understanding of the processes that might
be involved in the creation of real or false anomalies.
This uncertainty caused a number of companies to propose
a scoping study to the Canadian Mining Industry Research
Organization (CAMIRO) to see how they might better understand the processes that have caused anomalies to form above
deeply buried deposits. This scoping study, Phase I of this
study, was carried out by Cameron over a six-month period
from October 1997. It was generously supported by 28
companies, both Canadian and international. Several of these
companies provided data from their own orientation studies
over known deposits.
During the development of selective leach geochemistry,
ideas have emerged about processes for the dispersion of
elements from buried targets, the most prominent being electrochemical dispersion. Research on dispersion mechanisms by
the exploration community, including researchers in universities

and government laboratories, has been limited by the available


funding. In contrast, over the past two decades, tens or possibly
hundreds of millions of dollars have been spent examining
processes that might cause dispersion of toxic materials from
buried nuclear waste. These processes are entirely analogous to
those that might be involved in the dispersion of metals from
buried mineral deposits. Moreover, a substantial portion of the
nuclear studies in the United States have been carried out in the
arid terrain of the Southwest, similar to regions where
companies are targeting buried deposits.
The results of the Phase I scoping study encouraged a
full-scale study to test methods and better understand processes. This was launched in 1999 as Deep-Penetrating Geochemistry, Phase II. The project was again sponsored by
CAMIRO and supported by 26 Canadian and international
companies. Test sites were selected in three regions, where
companies are actively exploring for deeply buried deposits: the
Abitibi greenstone belt of Ontario, where thick glacial sediments cover prospective rocks; the Carlin belt of Nevada,
where the most productive cluster of Au deposits in North
America has prospective areas of basement covered by the
Carlin Formation; and thirdly, northern Chile, where basement
rocks containing porphyry mineralization are covered by thick
piedmont gravels. We looked at ten sites. In this account we will
discuss five: (a) the Spence deposit, Chile, where 400 Mt of 1%
Cu is covered by piedmont gravel; (b) the Gaby Sur deposit,
Chile, where 400 Mt of 0.54% Cu is also covered by piedmont
gravels; (c) the Mansa Mina deposit, where piedmont gravels
conceal 325 Mt of 1% Cu; (d) the Mike deposit, Nevada, where
240 m of Carlin Formation conceals a coppergold deposit; and
(e) the Cross Lake prospect near Timmins where Zn-rich VMS
mineralization of Archean age is covered by 30 to 50 m of
glaciolacustrine clay and sand.
For the Phase II studies Stew Hamilton of the Ontario
Geological Survey (OGS) assumed the major role of working
on the Abitibi sites and came with substantial funding from that
organization that permitted drilling of the overburden. Beth
McClenaghan of the Geological Survey of Canada provided
expertise as a Quaternary geologist. Before the study commenced, it was apparent that groundwater might have played an
important role in the formation of surface anomalies in

Finding deeply buried deposits using geochemistry


northern Chile. Matthew Leybourne of the University of Texas
at Dallas led the studies of groundwater geochemistry. Mary
Doherty of BHP Minerals and International Geochemical
Consultants participated in the studies at Mike. The project
involved comparing dierent leaches, and Gwendy Hall of the
Geological Survey of Canada took responsibility for the
analytical programme. We were fortunate to get the support of
all major commercial analysts in Canada, who provided analyses
gratis and also communicated much of their knowledge: John
Gravel of Acme Laboratories, Eric Homan of Actlabs, Brenda
Caughlin and Pat Highsmith of ALS-Chemex, Claude Massie of
Bondar-Clegg, Robert Ellis of Gedex and Hugh De Souza of
XRAL. Moreover, Xie Xuejing of the Institute of Geophysical
and Geochemical Exploration (IGGE) in China provided
analyses by methods typically used in that country. Thus, for
many of the samples there are nine sets of analyses to compare.
Starting in 1999, 38 reports on the Phase II project were
released to sponsors on topics as dierent as biogeochemistry,
hydrology and Pb isotopes at the Abitibi sites, stable isotopes in
groundwater and soils from Chile and soil gas tests in Nevada.
This article touches on only a few of these topics and includes
only information for which a two-year confidentiality limit has
expired.
PROCESSES FOR THE VERTICAL MIGRATION OF
ELEMENTS
Migration of dissolved solids in water
Solids dissolved in water may migrate within the water mass
diusion or may migrate as part of a moving water mass
advection. Diusion, the first of these to be discussed, may
occur as a result of a chemical or electrical gradients.
Diusion through the vadose zone. For all but the wettest of areas,
vertical transport of elements from a buried deposit to the
surface must include passage through the vadose or undersaturated zone. Except for hyper-arid climates, the vadose zone is
not dry. There is a film of water around mineral grains of the
rock. This film of water is in downward motion and serves to
recharge to groundwater the water from precipitation remaining
after evaporation and run-o. The rate at which this water film
moves can be estimated by measuring depth profiles of
elements and isotopes in the water film extracted from drill
core. Since it is a conservative element, Cl is widely used. Rain
contains Cl, the content of which, per volume of precipitation,
is relatively constant over time for a given area. With increased
aridity and therefore evapotranspiration, the Cl content of soil
water increases and the downward flux of moisture recharging
to the groundwater decreases.
Figure 1 shows data from core of the top of the unsaturated
zone from the Hueco Bolson region, near El Paso, Texas. This
alluvial plain slopes 1 to 1.5% to the Rio Grande. Creosote and
mesquite are common and root to 1 to 5 m depths. Mean
annual precipitation is 28 cm. The unsaturated zone comprises
0 to 15 m of silty to gravel loam, underlain by 140 m of clay
with interbedded sand and silt. A discontinuous layer of caliche
occurs at 1 to 2 m. Chloride was measured after adding
deionized water to dried core samples. The content is low from
the surface to 0.3 m, the result of downward leaching by rain.
Below this, Cl increases as a result of evapotranspiration,
forming a bulge with maximum concentrations at depths of 1.3
to 4.6 m for dierent cores. This bulge occurs throughout the
southwestern United States, just below the root zone (Phillips
1994). It corresponds to a minimum in the moisture flux, i.e. a
maximum in evapotranspiration, which represents a cessation

Fig. 1. Chloride content and estimates of moisture flux for profiles


of unsaturated soils, Hueco Bolson, Texas. Scanlon (1991).

of recharge to the water table about 13,000 to 15,000 years ago,


when the climate began to shift from cooler and wetter to
the current arid conditions. When precipitation is greater,
profiles move slowly downwards as water and Cl are added
incrementally at the top, thus providing a view of changing
moisture fluxes with time. The 15 m profile shown is estimated to represent 30,000 years. Below the bulge, the
decreasing Cl contents indicate an increasing moisture flux
prior to 15,000 years.
Mathematical methods of relating Cl concentration to
moisture flux are given in Scanlon (1991). An advection
diusion equation can be fitted to the data to evaluate the
relative importance of these two factors on the mobility of Cl.
The very large concentration gradients represented by the bulge
provide a driving force for diusion, which will tend to smooth
the bulge. Above the peak in the bulge, the upward diusive
fluxes of chloride are 103 to 104 mm a1. Below the peak,
the downward fluxes are 103 to 105 mm a1. The downward advective flux is 101 to 102 mm a1, or two to three
orders of magnitude greater than the diusive flux.
The moisture content has a direct eect on the migration of
dissolved material by diusion. As the ground becomes drier,
liquid flow paths become more tortuous and the degree of
liquid interconnection decreases. Some flow paths may dead
end. Whereas water molecules may move in the vapour phase
across air-filled pores, involatile solute molecules must follow
tortuous connected liquid pathways. There will be flow into
dead-end paths, since water continues to move in the vapour
phase. Non-volatile dissolved molecules can only escape from
dead-end paths by diusion back out against the flow direction.
Thus, involatile dissolved molecules, such as Cl, move more
slowly than water. This has been demonstrated by isotope
studies, including 36Cl and 3H, released during atomic bomb
tests. 36Cl was generated during sub-sea tests and fallout peaked
in 195556. 3H was generated during atmospheric tests and
peaked later, in 196465. Once incorporated into soil moisture,
these isotopes behave dierently. 36Cl is non-volatile and is
restricted to liquid phase flow, whereas tritiated water, like
normal water, can move in either liquid or vapour phases.
Scanlon (1992) measured the penetration of 36Cl and 3H into
the soils of the same region, Hueco Bolson, referred to above.
The 36Cl/Cl ratio reaches a maximum at 0.5 m depth (Fig. 2).
Although 3H peak fallout was eight years later than 36Cl, tritium
penetrated deeper, below 1 m, indicating that faster migration
is possible in the vapour phase compared to dissolved
components.
Electrochemical dispersion. Electrochemical cells are generated
around most sulphide bodies as a result of their oxidation and

10

E.M. Cameron et al.

Fig. 3. Supergene alteration caused by electrochemical oxidation of


pyrrhotitepentlandite ore in arid terrain. From Blain & Brotherton
(1975); Thornber (1975).

Fig. 2. Vertical profiles of 3H and 36Cl/Cl for unsaturated soils,


Hueco Bolson, Texas. Scanlon (1992).

are often detectable by negative self-potential (SP) anomalies


above the bodies. A sulphide body can be considered a
conductor immersed in an electrolyte, groundwater. Redox
dierences exist between the more oxidized upper (cathodic)
portion of the body and the more reduced deeper (anodic) part.
Anodic corrosion takes place at depth between sulphide minerals and the electrolyte, releasing electrons: MeS < M2+(aq) +
S + 2e. Electrons travel, usually upwards, through the
conductive sulphide body to locations where oxygen or other
oxidants are available and take part in its cathodic reduction: O2
+ 4H+ + 4e < 2H2O. Because of the electrical potential
established between the cathodic and anodic sites, there is
current flow, carried by ions in the groundwater. Cations
migrate to the cathode and anions to the anode. Cathodic
reactions are usually shallow, close to the supply of oxidants
near the water table. Water is required as a source of dissolved
oxidants and to permit the migration of dissolved cations and
anions, which carry the current. One well-described cell is that
produced by the massive nickel sulphide deposits at Kambalda,
Western Australia (Fig. 3).
Much of the pioneer work in exploration geochemistry on
electrochemical cells involved sulphide bodies that reach close
to the surface. Air and infiltrating rainwater, saturated in
oxygen, bathe the cathodic zone. Bolviken & Logn (1974)
described an electrochemical cell formed by the Joma deposit,
Norway, which is covered by 0.5 to 1.2 m of glacial sediments.
There is a well-developed SP anomaly detectable at surface.
Samples of till, taken as close as possible to bedrock, showed
elevated concentrations of metals over the mineralization and
display higher conductivity in watertill slurries. Increased
conductivity supports the contention that free mobile ions
should be more abundant near the upper cathodic zone. Similar

Fig. 4. Conceptual model by for the electrochemically induced


dispersion in arid terrain (Elura deposit, Australia). For Elura, the
sulphide has been completely oxidized to a depth of 98 m and the
water table lies between 80 and 115 m. Diagram simplified from
Govett et al. (1984).

results were obtained by Govett (1975) for a variety of


locations, with unambiguous conductivity anomalies found over
deposits with shallow cover.
Smee (1983); Govett et al. (1984) developed similar models
for the formation of twin or rabbit ear anomalies at the
surface above an oxidizing sulphide deposit. Whereas
Govetts model was for transfer through the vadose zone in arid
terrain, Smees was for migration through saturated clay. In
Govett et al.s model, cations following the current path
accumulate on either side of the sulphide body in a zone of high
current density near the base of the conductive weathered zone
(Fig. 4), which is near to the water table. From these zones of
high concentration, cations move to the surface by chemical
diusion. But, as described in an earlier section, chemical
diusion within the water film coating grains in the vadose zone
is much slower than the downward advective flow of the water
film.

Finding deeply buried deposits using geochemistry

11

water table there is ready access to O2, permitting oxidation of


the reduced material. This will produce a number of eects,
including a decrease in pH and the concomitant dissolution of
carbonate (Fig. 5). The reduced column, pH lows and dissolution of carbonate predicted by this model have been documented at the Cross Lake site discussed below and at other
locations in the Abitibi greenstone belt. And elements present
in the mineralization, such as Zn and Cd, show clear anomalies
in the soils above. There are also negative SP anomalies,
indicative of current movement in the conductor. There are
considerable technical diculties in fully evaluating the processes that occur within the reduced column. To create the
features seen at the surface at Cross Lake, where 30 m or more
of clay lie above the sulphide lens, and within a period of 8,000
years since the clays were deposited, requires rates of migration
well beyond that of chemical diusion.

Fig. 5. Conceptual model from Hamilton (1999) showing the eects


of the dispersion of a reduced column of metal and other
constituents upwards through clay overburden from sulphide mineralization exposed at the subcrop. The reduced material reacts with
oxygen at and above the water table to form acid, which dissolves
and redistributes carbonate.

Smees model is for the Abitibi greenstone belt, where thick,


>30 m, generally saturated, glaciolacustrine clays often cover
mineralization. Cations that concentrate at the base of the clays
diuse upward. Diusion coecients of 1x1010 and 1x1011
m2 s1 are typical ranges for non-reactive chemical species in
clay. A diusion front representing only one percent of the
original concentration would reach 10 m out from source after
10,000 years. This is for one-dimensional diusion, whereas
dispersion from an orebody would be radial, resulting in lower
concentrations at a given distance from the deposit. Clays may
act as a semi-permeable membrane, resisting the diusion of
ions (Schwartz 1974) or clay mineral surfaces may adsorb metal
ions. The limited nature eect of either advective or diusive
flow in thick glaciolacustrine clays since their time of deposition
is illustrated by the empirical data of Remenda et al. (1994). They
found that at depths of 20 to 30 m in clay from various sites in
the southern Canadian Shield, interstitial waters retained 18O
compositions of 24 to 25 characteristic of surface
waters at the time of deposition. These waters have been static,
showing no evidence of equilibration with modern waters in the
area of 13 to 14. Thus, neither advection or diusion
has been eective in moving these waters over the last c. 10 ka.
Although electrochemical processes are an important means for
releasing ions from a buried sulphide body, there is diculty in
explaining how the ions so released travel to the surface.
A dierent model was proposed by Hamilton (1998, 1999,
2000) for saturated overburden in glaciated terrain where the
sulphide subcrop is covered by tens of metres of glacial
sediments. This model (Fig. 5) invokes transport along redox
gradients. Consumption of oxidants at the sulphide/sediment
interface creates a reduced environment. Continued oxidation
can only occur if reduced species are removed. Redox dierences between the top of the body and the water table near the
surface, where oxygen is in contact with groundwater, produce
a vertical electrochemical gradient. Reduced species, such as
HS and Fe2+, migrate upward along this gradient and
reactions between oxidized species moving in the opposite
direction dissipate charge away from the sulphide. As the
limited number of oxidizing agents between the top of the body
and the water table are consumed, a reduced column is
postulated to form between the body and the water table. At the

Advective flow. Water and its dissolved contents typically flow


down-gradient, not upwards. However, there are some
exceptions. One of these is the eusion of groundwater at the
surface during earthquakes. In the Phase I scoping report by
Cameron (1998), this mechanism was given less than a page,
because it was considered unlikely to produce significant surface
anomalies. However, in the subsequent field programme in
Chile this appeared as an important process for anomaly
formation.
Earthquake-induced surface flooding by groundwater has
long been recognized. Sibson (1981), p. 593) described groundwater movements along fault lines after earthquakes: In arid
terrain particularly, there are reports of changes in well water
level, spring flow and occasional dramatic eusions of groundwater immediately following moderate to large shallow earthquakes. Surface flows occurred in a desert area of Iran during
earthquakes in 1903 and 1923 (Tchalenko 1973). During the
magnitude 7.3 Salmas earthquake of 1930 in NW Iran, water
was expelled along the main fault trace and a broad area of
alluvial plain on the downthrow side became fissured and
waterlogged (Tchalenko & Berberian 1974). Nur (1974)
noted outpourings of warm groundwater accompanying the
Matsuchiro earthquake in Japan. Following the Hebgen Lake
earthquake of Montana in 1959, three rivers increased in flow
by c. 50%, the increases continuing for several weeks through
dry weather (Muir-Wood 1994). The Kern County, California,
magnitude 7.5 earthquake of July 1952 was followed by two
months of drought. There were outpourings of groundwater in
the vicinity of faults, and increases in spring flow and well water
level, which continued for two months (Briggs & Troxell 1955).
Sibson et al. (1975) used the term seismic pumping to explain
this process. Pre-seismic extension produces fractures in the
brittle upper crust that provide pathways for groundwater
migration and storage. During earthquakes, stress fields become
compressional, closing fractures and forcing groundwater to the
surface along faults. Given the low permeability of basement
rocks, groundwater is slow to migrate to faults, resulting in
continued eusion of water for weeks following an earthquake.
Groundwaters expelled from depth are typically old and saline,
but may mix near-surface with less saline waters.
There are other mechanisms for the advective transfer of
groundwater and solutes to the surface. One that was applied
early in the history of exploration geochemistry is advective
transfer by deeply rooted plants. Analysis of the plant material
provides information on the composition of deep groundwater.
In some arid areas, plants have been found to root as deep as
100 m (Cannon & Starrett 1958). Also, capillarity can draw
groundwater from the water table to the surface, where it
evaporates. However, the water table must be relatively close to

12

E.M. Cameron et al.

Fig. 6. Rainier Mesa, Nevada, showing the location of a cavity for a


1kt explosion. Bottles of SF6 gas and 3He gas were set close to the
explosion. SF6 first reached the surface at site OS6 50 days after the
detonation during a barometric low. 3He arrived only after 325 days.
From Carrigan et al. (1996).

the surface, since the height over which the water can be drawn
is dependent on the parameter known to soil scientists as soil
suction. Soil suction depends on the nature of the soil and the
degree of dryness. Clay soils exert greater suction than sand or
silt and dry soils more than damp. Fontes et al. (1986) describe
a site in the Sahara Desert where capillarity draws water to the
surface from a depth of 10 m. They suggest that 20 m may be
a limiting depth for this process.
Migration of gases and the eects
Lovell et al. (1983); McCarthy et al. (1986) and others have
shown that gases such as carbon dioxide and methane are more
abundant in soils above some deposits, whereas oxygen is
depleted. Clark et al. (1997) have argued for the migration of
halogens and volatile metal compounds to the surface. Chinese
geochemists (e.g. Xie et al. 1997) have suggested that a variety of
elements, Au, As and Sb, are carried to the surface in mobile
form by gases, including very fine particulates, in the submicrometre to nanometre range. Hydrocarbons are currently
being measured in soils. The SGH method of Actlabs Ltd
(Ancaster, ON, Canada) desorbs and measures hydrocarbons in
the C5 to C17 range from B-horizon soils. A similar method,
soil desorption pyrolysis (SDP, St. Lucia, Queensland,
Australia), has also been developed for hydrocarbon gases in
soils. Similar to dissolved solids in water, gases may migrate by
diusion through air or water, or advectively, as a result of the
movement of a mass of air or water containing the gases. For
gases, as for dissolved solids in water, advective flow is a far
more rapid transport mechanism than diusion.
Barometric pumping. Carrigan et al. (1996) carried out a simulated
underground nuclear test. A charge of 1.3  106 kg of chemical
explosives, equivalent to a 1 kt nuclear charge, was detonated at
a depth of 400 m in bedded tu at Rainier Mesa in the Nevada
Test Site (NTS) (Fig. 6). Two bottles of gas were placed near to
the charge. One contained 3He and the other SF6. The amounts
of gas involved were not large, for 3He, 1.3 m3 and for SF6,
8 m3. The detonation chamber was close to a fault. After the
explosion, sampling sites were established at the surface to
detect these gases. The first gas to be detected was SF6 along
the fault at site OS6, 50 days after detonation, during a strong
barometric depression. The fault along where OS6 is sited is not
the one that runs close to the detonation cavity. Thereafter, SF6

was detected at sites OS1, OS2 and OS3. 3He was first detected
at the surface at OS6, 375 days after detonation.
These empirical observations are entirely contrary to migration by gaseous diusion. 3He has a much higher diusivity than
SF6 and, if diusion was the dominant mechanism, should
reach the surface long before the other gas. The diusivity of
SF6 is such that it would require 10 s to 100 s of years to reach
the surface, yet it has happened in days. Why? The reason is that
gaseous diusion is overridden by a much faster mechanism,
barometric pumping. The description of barometric pumping
given here is largely derived from Nilson et al. (1991), work
carried out at the NTS.
Barometric pumping refers to the process where cycles of
high and low barometric pressure first force air into the earth
and then withdraw a mixture of the air plus gases that were in
rock. For all practical purposes, barometric pumping applies
only to fractured rock. For reasons discussed below, barometric
pumping is not a significant process for non-indurated, unfractured material, even if permeable. When the permeability of
NTS alluvium is measured on small samples in the laboratory or
large volumes are measured in situ in the field by borehole
injection, results are similar, in the range 1 to 15 darcy (D).
When the permeability of NTS volcanic rock is measured on
core samples in the laboratory and again by borehole injection,
results are very dierent, only 106 to 102 D on core, but 1
to 15 D for borehole injection, the same as for the alluvium.
The higher permeability of the bulk rocks is because fractures,
not present in the small laboratory core samples, control the
permeability.
High barometric pressure forces air down fractures and into
pore space in the rock around the fractures. Gases within the
rock, such as CO2 or hydrocarbons, mix with air by molecular
diusion. When the barometric pressure drops, air in the
porous rock, now containing the gases, returns to the fracture
and, after several cycles of high and low pressure, reaches the
surface. Pumping occurs because the volume of air entering
rock porosity is much greater than the volume of air present in
the fractures. The rock porosity provides the breathing volume
that permits large vertical movements during high pressure
inhalation and low pressure exhalation. By contrast, in
non-fractured permeable media, such as alluvium, air movements are piston-like and nearly reversible, so that there is little
upward transport of a gas of deep origin. For a fractured
permeable medium, barometric transport can be several orders
of magnitude greater than molecular diusion, whereas for
unfractured soil and alluvium, molecular diusion is more
important.
Using data from NTS volcanic rock, simulations were carried
out on the removal by barometric pumping of a deep gas. With
the condition of fresh air from the surface to 200 m, then air
plus radioactive gas from 200 m to the water table at 500 m,
10% of the contaminated gas was removed in one year. In terms
of geological processes this is rapid. The overall eciency of
deep gas transport to the surface is critically dependent on the
spacing of fractures. For closely spaced fractures, say only a few
centimetres or less, each fracture has only a small breathing
volume of porous rock around it. The transport eciency
increases with increasing fracture spacing up to a few metres.
But after about 10 m spacing there is no incremental benefit
because the air can only penetrate a few metres in the half
period of about 100 hours that is usual between high and low
barometric phases. Maximum transport eciencies are reached
for spacing in the range 2 to 10 m. Although it may initially
appear counterintuitive, transport eciencies are decreased
with increasing molecular diusivities of gases. This is because
a gas with high molecular diusivity can more readily diuse out

Finding deeply buried deposits using geochemistry

Fig. 7. Heavy metal enrichments in the marine atmosphere. Enrichment is measured relative to crustal abundance normalized to Al.
Data from Rahn (1976) and for Hg from Crozat et al. (1973).

of the upward-moving air/gas mixture in the fractures into


porous wall rock. This explains why in the bomb-simulation
experiments SF6 arrived at the surface more speedily than 3H,
which has a higher molecular diusivity. Diusion of hydrocarbons out of the upwardly rising air/gas mixture into the
surrounding wall rock may explain a reduced column of rock
above a leaking hydrocarbon reservoir.
Seasonal exhaustion of air and gas from the vadose zone. Yucca
Mountain, Nevada, rises 250 m above Solitario Canyon. In
winter, air within the mountain is warm. This warm air rises and
is replaced by cold, denser air entering at the base of the scarp.
Air exhausts continuously, at a relatively high velocity, typically
3 m s1, from wells drilled near the crest of the mountain
(Weeks 1987). In summer, the wells alternately intake and
exhaust air several times a day, but at lower velocity. A farmer
in Idaho drilled a 30 m dry well into the Snake River basalts,
near the Snake River Gorge. He built a greenhouse over the
well, with the air exhausting from the well maintaining plant
growth through the winter. This process for the advective
movement of air plus gas has been described by Rose & Gow
(1995).
Migration as bubbles and bubble-generated aerosols. Whereas barometric pumping provides a mechanism for the advective
transfer of gases to the surface in arid terrains with a thick,
fractured vadose zone, bubbles provide a means for advective
movement through the saturated zone. Kristiansson &
Malmqvist (1982) found that the radon migrated to the surface
much faster than can be accounted for by diusion in groundwater. They suggested that radon was included in microscopic
bubbles composed of nitrogen, argon, oxygen and methane.
Bubbles and bubble-generated aerosols may be involved in
the upward transport of the dissolved constituents of groundwater. The clue for this is metal-rich aerosols over the oceans.
Some elements are greatly enriched in the marine atmosphere
relative to their crustal abundance (Fig. 7). When waves break,
air is trapped in seawater, which rises as bubbles. Surface-active
substances are attracted to the bubblewater interface. These
include bacteria, colloids, soluble organic molecules and dissolved heavy metals (Piotrowicz et al. 1979a; Piotrowicz et al.
1979b). As bubbles rise they scrub the water of these
substances. At the surface of the ocean there is a surface
microlayer, enriched in the same surface-active materials. When
a bubble bursts, aerosols are formed from the bubble boundary
layer and from the surface microlayer. Experimental studies of
aerosol generation show strong fractionation of elements.
Piotrowicz et al. (1979) found that Fe was not enriched, whereas
Cu, Cd and Zn were enriched by a factor of 200. Lead had the

13

greatest enrichment with a factor of c. 2000. Fractionation is


increased with increasing length of the water column through
which the bubbles pass. Metals are most strongly enriched in
sub-micrometre aerosols (Arimoto et al. 1990). Aerosols in the
size range 0.1 to 10 m have a residence time in the lower
atmosphere of about a week. The small size of these particles
favours transfers from the solid to vapour phase and reactions
between the gases and solids, which may form volatile compounds
What are the possible relationships between oceanic aerosols
and ore deposits? Gases may perhaps be generated during the
oxidation of sulphides or gases from deeper sources, e.g. the
mantle, and may rise up faults along which deposits lie. Some
samples of groundwater taken from the Spence copper
porphyry deposit, Chile, were found to be saturated with CO2.
The degassed CO2 can rise as bubbles through groundwater,
scrubbing metals from the water en route. Mineralized groundwaters have concentrations of metals orders of magnitude
greater than sea water and bubbles are likely to scrub a greater
depth of water than the one or two meters typical of wavegenerated bubbles. Barometric pumping may serve to extract
sub-micrometre aerosols from the water table, but this has not
been demonstrated. Even where aerosol generation is minimal,
bubbles, with a high concentration of metals in the boundary
layer, may deliver metals to the water table. This may be of
significance in wet areas, where the water table is close to the
surface.
The Osborne deposit, in northern Queensland is a copper
gold replacement deposit in lower Proterozoic metamorphic
rocks, covered by 30 to 60 m of fractured sedimentary rock of
Mesozoic age. Placer Australia Pty sponsored a series of
geochemical tests over the deposit prior to mining, including
pSirogas collectors (Rutherford et al. in press). A plastic
collector containing a polystyrene film was buried in the soil at
a depth of 50 cm. A hole in the base of the container allowed
soil air to circulate over the polystyrene film. After 30 to
60 days, the collectors were recovered and the surface of
the films analysed by Particle Induced X-ray Emission
Spectrometry (PIXE) for a variety of trace metals. The surfaces
of the films were also examined by Scanning Electron
Microscopy (SEM). These scans show circular areas of precipitate, 20 to 30 m in diameter, which they attribute to the
impingement, then evaporation, of water or aerosol droplets.
Most of the geochemical data from Osborne are erratic between
adjacent sites (spikey), perhaps reflecting transport up
fractures, with the pSirogas results showing this strongly,
including the strongly anomalous results for Cu (Fig. 8).
Mineralogical change in soils by gases. The passage of gas from a deep
source through soils may cause changes in the soils composition or mineralogy. A possible, if extreme, example is
shown by analyses of soils above underground nuclear test
explosions in Nevada (Hall et al. 1997). In these tests, detonations were at the bottom of a vertical shaft that had been
back-filled and cemented. For the described sites, explosions
took place at depths of 293 to 350 m, well above the water
table, at 450 to 530 m. Data shown on Figure 9 are from soil
samples taken across the Ville test site, where the explosion
caused a surface slump crater 120 m in diameter and 17 m deep.
Analyses after a hydroxylamine extraction show a strong
anomaly for I, extending for 300 m on either side of ground
zero. Iodine is an element that forms a variety of volatile
compounds. Arsenic shows a weaker anomaly and also can
form volatile compounds. The presence of anomalies for
volatile elements suggests the passage of these plus possible
carrier gases through the soil. But also note the strong depletion

14

E.M. Cameron et al.

Fig. 8. Cu in soil air over Osborne


deposit, Line 21737N, as measured by
CSIROGAS collectors. Deposit lies
below 11200E to 11600E. Modified
from Rutherford et al. (in press).

in the amount of Fe extracted. Analyses for the total contents


of Fe and other elements showed a flat pattern: there has been
no change in the bulk composition of the soils above the
detonation site; only the small fraction of selected elements that
can be dissolved by a selective leach. The depletion in Fe shown
by the selective extraction may be the result of gas changing the
mineralogical state of the Fe oxide minerals, such that Fe was
less easily dissolved by the hydroxylamine leach. The possibility
that the observed changes may be the result of contamination
during drilling have been discounted by more recent unpublished work where the nuclear device was placed using a
horizontal adit. Soils over the Ruby Star copper skarn deposit,
Arizona (Fig. 10) are also depleted in Fe, when measured by a
selective leach. Until more detailed studies are carried out, the
causes for these depletions must remain speculative.

including a mineral deposit, are termed the exogenic phase. The


exogenic phase is initially added to the soil in water-soluble
form; as a result of soil-forming processes, this phase is
progressively incorporated into secondary minerals (Fig. 11).
Each secondary mineral contains elements of both endogenic
and exogenic origin. The relative amounts of these phases vary
between minerals and, for a given mineral, from place to place.
Given that the exogenic phase must first enter a soil in a
water-soluble form, one approach is to use a weak leach that
does not attack any minerals, but dissolves water-soluble salts
and elements loosely adsorbed to mineral surfaces. Pure water,
i.e. deionized water, is the most simple leach for this purpose,
but suers from poor analytical reproducibility. An alternative is
the Enzyme Leach. In our experience, this leach dissolves little

Measurement of elements and gases in soils


Selective leaches. Most applications of geochemistry to exploration
for buried deposits involve the analysis of soils by selective
leaches (Hall, 1998). The term selective leach was first applied
by soil scientists to reagents that could dissolve particular
mineral phases, with minimal eect on other minerals. In
exploration geochemistry, selective leaches are used to dissolve
minerals that may include a high proportion of the mobile
phase of elements, including material derived from deposits.
Soil-forming processes convert primary minerals derived from
rocks into minerals stable at low temperature. The element
fraction that comes from primary minerals is here considered
the internal or endogenic phase. Elements from external sources,

Fig. 9. Contents of Fe, I and As in soils collected along a traverse


crossing ground zero (0 m on horizontal scale) above the Ville atomic
bomb test, Nevada. Soils were extracted with hot hydroxylamine
(HX Fe). From Hall et al. (1997).

Fig. 10. Iron and Cu by Enzyme Leach in soils over the Ruby Star
copper deposit, Arizona. Data are moving average of three sites.
Deposit contains 100 Mt 1% Cu and consists of massive slump
blocks enclosed within fanglomerate and covered by 40 to 300 m of
gravel. Data from Kelley (1995).

15

Finding deeply buried deposits using geochemistry

Fig. 11. Conceptual model showing the


distribution of exogenic and endogenic
phases between soil-forming minerals.

more than deionized water (Fig. 12), but the analytical reproducibility is better. The other main proprietary leach, Mobile
Metal Ion (MMI), comes in several forms, two of which are an
acidic MMI-A and a basic MMI-B. The MMI-A leach (targeting
Zn, Cu, Pb, Cd) dissolves more metal than the Enzyme Leach,
which indicates that it is dissolving secondary minerals. However, the formulations of the MMI solutions have not been
published and, because major elements are not determined by
the laboratories carrying out MMI analyses, it is not possible to
interpret which mineral phases are being dissolved.
The second approach is to use a leach that dissolves one or
more secondary minerals that contain a favourable ratio of
exogenic to endogenic material at a particular site. Ammonium
acetate at pH 5 dissolves carbonate minerals. For oxide
minerals, hydroxylamine hydrochloride is widely used. Cold
(room temperature) hydroxylamine in an acidic solution, pH c.
1.5, (HX Mn) dissolves Mn oxides, whereas hot (60 oC)
hydroxylamine in a more strongly acidic solution (HX Fe)
dissolves both Mn and Fe oxides. Because both hydroxylamine
leach solutions are acidic, they will also dissolve carbonate, so
that the element phases extracted are the sum of those present
in carbonates and oxide minerals, plus, of course, that present
in water-soluble form. In general, the more acidic the leaches,
the less selective they become. For organic material, such as
humus, sodium pyrophosphate is the most commonly used
leach, although it tends to dissolve large amounts of elements of
endogenic origin, unrelated to mineralization. Aqua regia should
also be mentioned. This is not a selective leach, but dissolves
most secondary minerals and partially extracts some silicates.
Because it dissolves a high proportion of endogenic phases, the
anomaly to background contrast is generally low. However, it
provides useful information on the overall composition of the
soils, which assists in the interpretation of the selective leach
data. And it may give good results for mature anomalies, where
elements of external origin are no longer being introduced.
There are complexities involved in the use of selective
leaches, which require consideration during survey planning,
analysis and interpretation. When significant amounts of carbonate are present in soils, this can partially neutralize leach
solutions that contain acid, so that their leaching capacity is
degraded. For example, 1215% CaCO3 in samples can increase
the pH of a cold hydroxylamine (HX Mn) solution from c.1.5 to
5.5 and MMI-A from c. 2.5 to 6. Element contents that might
be extracted in the absence of carbonate may not be fully

extracted or are reprecipitated or adsorbed as the pH increases.


Thus analytical laboratories should measure and report the final
pH of the leach solution. The ICP-MS is subject to spectral and
non-spectral interferences, which may aect the determination
of numerous elements, particularly those at low concentration
(Hall 1992). For example, mass 105Pd is commonly used to
measure Pd, but this is also the mass for the molecular band
105
(SrOH). Palladium is usually present in soils in the low ppb
range and its signal can be overwhelmed by SrOH formed in the
plasma during analysis of soils of moderate to high Sr content.
Quality control on analyses is mandatory for surveys, using
hidden duplicates and standards, the latter being samples that
have been analysed in previous surveys. Analytical duplicates
are separate vials of the same prepared sample, given dierent
numbers and placed at distant positions in the submitted
analytical batch. Field duplicates are separate samples collected
at the same site, but 2 to 5 m apart from each other. After
analysis, the analytical and field duplicates are separated into
groups and the standard deviation, s, and relative standard
deviation, RSD, are computed for each:
s2 = ~&~xi1  xi2!2!/2N
RSD = 100 s/X where the squares of the dierences between
the duplicate pairs are summed, then divided by 2N, N being
the number of pairs, to produce the variance estimate s2. The
mean of all duplicates, X, is then used to calculate RSD.
In Figure 13 RSDs are shown for 29 elements as determined
by the Enzyme Leach. The RSDs were computed from ten pairs
of analytical duplicates of B-horizon soils from the Abitibi test
sites. The median RSD is also shown. The median RSD of the
analytical duplicates for each leach provides a useful comparison of analytical precision between dierent leaches. In Table 1,
median RSDs for five leaches have been compiled. In general,
the stronger the leach, extracting more of each element, the
lower is the RSD. This is because it is more dicult to
consistently extract a small fraction of the element present than
a large fraction. The amounts of elements extracted by the
dierent leaches, as shown in Figure 12, correlate with the
RSDs given in Table 1. Other than deionized water, the RSDs
obtained from the selective leaches are satisfactory. The variance estimate, s2, for the field duplicates is the sum of the
analytical variance plus that due to sampling. Thus the RSDs for
field duplicates are higher than those for analytical duplicates

16

E.M. Cameron et al.

Fig. 13. Relative standard deviation (RSD) for ten pairs of analytical
duplicates of B-horizon soils from the Abitibi belt analysed by
Enzyme Leach.

every 10oC increase in temperature and is also influenced by soil


moisture and the content of organic matter (Amundson &
Davidson 1990). Carbon dioxide dissolves in water, forming
H2CO3, HCO3 and CO32. Thus the presence or absence of
moisture can substantially influence the concentration of this
gas. An alternative, that reduces temporal variations, is to
measure the integrated flux, such as is done with radon by
burying alpha track film in the soil for days or weeks or
adsorbing gaseous Hg on Au film. For metals in soil air, CSIRO
in Australia and BRGM in France place collectors in the ground
for c.100 days. In the BRGM method (Pauwels et al. 1999) the
metals are collected on activated carbon, whereas in the CSIRO
method (Rutherford et al. in press) solids and vapour impinge
on a thin plastic film. Organic gases are absorbed on clay
minerals. By desorbing these gases, then measuring their relative
abundance, clues may be obtained to buried deposits. This is the
basis of the SGH (Soil Gas Hydrocarbons) technique of Actlabs
and the SDP (Soil Desorption Pyrolysis) method of the
University of Queensland.
Fig. 12. Box-whisker plots showing the amounts of element
extracted by deionized water, Enzyme Leach, MMI-A, ammonium
acetate at pH 5 (AA5), cold hydroxylamine (HX Mn) and aqua regia.
Top of the diagram, Cu in 63 samples of soil from a traverse across
the Spence copper deposit. Bottom of diagram, Zn in 54 samples of
B-horizon soils from two traverses across VMS mineralization at
Cross Lake, Ontario.

(Table 1). Only in the case of the deionized water leach is the
RSD for the analytical duplicates so high that it approaches the
RSD of the field duplicates.
Measuring gases. Soil gases are commonly measured directly in the
field using instrumentation. This confers the advantage that
anomalies may be investigated immediately. The disadvantage is
that the instantaneous flux of gas so measured is subject to
many environmental variables that may change the estimated
concentration. In a study of radon activities from sites in
Pennsylvania, Rose et al. (1990) found that readings from depths
below 70 cm varied by factors of 3 to 10 during the year for
dierent sites. At shallower depths, seasonal variability is even
greater. Temperature and moisture variations, rain, wind and
barometric change may all aect the measured concentration.
Carbon dioxide shows variation due to seasonal, sampling
depth and environmental changes similar to that of radon. The
flux of CO2 in soils increases by between 1.5 to 3 times for

FIELD TRIALS
Study areas and methods
Phase II of the CAMIRO Deep-Penetrating Geochemistry
project was to carry out studies at test sites where known
mineralization is covered by 30 to 240 m of rock or consolidated or unconsolidated overburden. Sites were chosen in the
Abitibi greenstone belt of Ontario, the Carlin district of
Nevada, and the Atacama Desert of northern Chile. The
characteristics of these sites are summarized in Table 2 and the
individual studies carried out at each site are listed in Tables 3
and 4. These three regions were chosen because of their
geological and climatic contrasts. The Atacama is the driest
large desert area of the world and is host to some of the
largest and lowest-cost Cu deposits. Carlin is the most productive district for Au in North America and has a semi-arid
climate. The Abitibi, a glaciated region with a boreal cool
humid climate, contains many major lode Au and volcanogenic
massive sulphide (VMS) deposits.
Work in Chile and Nevada was entirely funded by company
sponsors through CAMIRO. Much of the work in Ontario was
funded by the Ontario Geological Survey (OGS). Reports on
individual studies were provided to the company sponsors on
an ongoing basis starting in 1999. In November 2001, a
CD-ROM containing 31 reports and seven database files was
distributed. These results were reviewed at a meeting held in

17

Finding deeply buried deposits using geochemistry

Table 1. Relative Standard Deviations (RSD) for ten pairs of analytical duplicates and ten pairs of field duplicates taken from three test areas in the Abitibi greenstone belt in 1999
Analytical method

Number of elements for median RSD Median RSD for analytical duplicates

Aqua regia
Cold hydroxylamine
Enzyme Leach
Deionized water leach
MMI

52
35
29
49
7

Median RSD for field duplicates

7
11
21
56
(14)*

11
22
38
60
(53)*

Samples of B-horizon soils are composites of sub-samples from five auger holes within a 2 m radius. Duplicate composite samples were collected at a 4 m distance
from the first sample.
*
RSDs for MMI are not directly comparable with that of the other methods. MMI samples are single (not composite) samples taken by auger at a fixed interval
1025 cm below the Ao-horizon. Analytical RSDs for the MMI samples are derived from in-house laboratory duplicates, not blind duplicates as with other
methods.
Table 2. Summary of sites studied during CAMIRO Deep-Penetrating Geochemistry Phase II project, 19992001
Location

Deposit type

Age

Chile
Spence
Gaby Sur
Mansa Mina
Tamarugal

Cu Porphyry
Cu Porphyry
Cu Porphyry
False Anomaly

Paleocene
Oligocene
Oligocene
n/a

Nevada
Mike

Cu-Au

Pre-Tertiary

Abitibi
Cross Lake Line 6
Cross Lake Line 40
Half-Moon Lake
Marsh Zone
Tillex

VMS
VMS
VMS
Gold
Copper

Archean
Archean
Archean
Archean
Archean

Grade

Cover

400 Mt 1.0% Cu
400 Mt 0.54% Cu
325 Mt 1.0% Cu
n/a

150 Mt 0.25%
Cu, 0.71 g/t Au
n/a
n/a
n/a
n/a
1.38 Mt 1.6% Cu

Piedmont
Piedmont
Piedmont
Piedmont

Cover thickness

Cover age

gravel
gravel
gravel
gravel

3180 m
2040 m
50>300 m
300 m

Miocene
Miocene
Miocene
Miocene

Carlin formation

150250 m

Miocene

Clay
Clay and sand
Clay
Clay and peat
Clay, till, peat

30 m
52 m
1215 m
1027 m
30 m

Quaternary
Quaternary
Quaternary
Quaternary
Quaternary

Table 3. Analytical methods for B-horizon and equivalent soils


Location
Chile
Spence
Gaby Sur
Mansa Mina
Tamarugal
Nevada
Mike
Abitibi
Cross Lake Line 6
Cross Lake Line 40
Half-Moon Lake
Marsh Zone
Tillex

Deionized
water

Enzyme
Leach

MMI-A

U
U
U

U
U
U
U

U
U
U
U

U
U
U
U
U

U
U
U
U

Amm. Hydroxylamine Aqua regia


acetate
U
U

U
U

WEM

AEM

OBM

FMM

U
U
U

U
U
U
U

U
U
U

U
U
U

U
U
U

U
U
U

U
U
U
U
U

U
U
U
U
U

U
U
U
U

U
U
U
U

U
U
U
U

U
U
U
U

WEM=Water-extractable, AEM=adsorbed extractable metal, OBM=organic-bound metal, FMM=iron- and manganese-bound metal

Toronto in December 2001 when sponsors approved further


work. The additional results were distributed as seven reports in
October 2002, including isotope studies in Chile, detailed trench
sampling at Cross Lake, and data provided by IGGE in Beijing.
Project results are subject to a two-year confidentiality period
after initial distribution to sponsors. Because of this restriction,
some of the results from the November 2001 release and none
of the data from the October 2002 release could be considered
for inclusion in this overview report. Work funded by the OGS
was not subject to the confidentiality agreement and these
results have been presented in OGS publications and elsewhere,
which are listed in the references. Because of the volume of
information obtained during Phase II of the project, only
selected parts of the work are described here.

Analytical methods. Analysis of soils was on material sieved to <80


mesh (<0.177 mm). For the deionized water extraction, 20 ml of
water were added to 2.5 g sample and tumbled in a roller for
3 min every 15 min for 1 hour. After settling for 10 min, the
leach was centrifuged for 8 min at 2500 rpm, then decanted.
Enzyme Leach and MMI-A are patented and proprietary leaches,
respectively, described by Hall (1998). For ammonium acetate at
pH 5 (AA5),which dissolves carbonate minerals, 1 g sample was
digested by 40 ml of 1.0M NH4OAc/pH 5.0 for two hours,
then centrifuged and the solution diluted with water 50:1. Cold
hydroxylamine in weakly acidic solution dissolves Mn oxides (HX
Mn). For this, 1 g of sample is treated with 25 ml of a solution
with 0.1M NH2OH:HCl in 0.04M HNO3 for 30 min at
room temperature. For a limited number of samples involving

18

E.M. Cameron et al.

Table 4. Other methods used at the test sites


Location

Chile
Spence
Gaby Sur
Mansa Mina
Tamarugal
Nevada
Mike
Abitibi
Cross Lake Line 6
Cross Lake Line 40
Half-Moon Lake
Marsh Zone
Tillex

Groundwater

Soil gas
CO2, O2

Soil:
stable
isotopes

U
U

Soil:
Metals in Soil pH, Soil CO3 Soil SO4 Biogeochemistry
Pb
soil gas
Cond
isotopes

U
U
U

U
U

U
U
U
U

U
U

U
U
U
U

U
U
U
U

Humus,
Na
pyrophosphate

Humus,
aqua regia

U
U
U
U

U
U
U
U

U
U
U
U

U
U
U
U

U
U

sequential leaches at Spence and Cross Lake, we also used


ammonium acetate at pH 7 (AA7) and a more aggressive hydroxylamine leach (HX Fe), which dissolves Fe as well as Mn oxides:
hot 0.25M NH2OH:HCl in 0.25M HCl. An aqua regia leach was
also used; this dissolves all secondary minerals in soils. After
dissolution all leach solutions were analysed by ICP-MS and, for
elements at high concentrations, by ICP-ES. It should be noted
that because the MMI-A and hydroxylamine are acidic, they also
dissolve carbonates. Thus, in a non-sequential leach mode, HX
Mn dissolves the water-soluble salts and elements within
carbonates, as well as elements present in Mn oxides.
The WEM (water-extractable metal), AEM (adsorbed
extractable metal), OBM (organic-bound metal) and FMM
(iron- and manganese-bound metal) methods were carried out
at the IGGE in Beijing. These leaches were applied sequentially.
After the first leach, the residue is treated with the second leach
and this is repeated through the third and fourth leaches. The
first leach (WEM) is deionized water. The next leach (AEM) is
ammonium citrate, which removes adsorbed elements. The
third leach (OBM) is sodium pyrophosphate to extract organicbound material. The final leach (FMM) is a mixture of
ammonium citrate and hydroxylamine hydrochloride to dissolve
Fe and Mn oxyhydroxides.
Studies in Chile
Study areas in Chile are shown in Figure 14. Spence, Gaby Sur
and Mansa Mina are deposits with substantial reserves of
copper, which are being prepared for mining or are the subject
of feasibility studies. All are covered by piedmont gravels, which
are consolidated rocks of Miocene age. At all sites, there are
distinct geochemical anomalies in soils above the deposits. The
fourth study area is Tamarugal, 50 km NW of Chuquicamata.
Here there is a extensive anomaly, >100 km2, with high
contents of porphyry indicator elements, Cu, Mo, Re, As and
Se. Drilling by Noranda through 300 m of piedmont gravels
showed that the basement rocks below the gravels are barren.
The study of this false anomaly was undertaken to understand
the processes that led to its formation. In this overview paper
we describe selected aspects of our study at the Spence, Gaby
Sur and Mansa Mina sites.
Spence Deposit, Chile. The Spence deposit was discovered by
RioChilex in 199697 by grid drilling through piedmont gravels
of Miocene age. The deposit displays a typical supergeneenriched sequence and is associated with three dacite porphyry
stocks (Fig. 15) intruded along a NE axis into andesite. Sulphide

Fig. 14. Map of northern Chile showing the location of major


porphyry copper and other copper deposits. The sites studied where
porphyry copper deposits are covered by piedmont gravels are
Spence, Gaby Sur, and Mansa Mina. Another site examined is
Tamarugal, 50 km NW of Chuquicamata, where a strong anomaly in
soils lies above barren basement covered by 300 m of gravel.

Finding deeply buried deposits using geochemistry

Fig. 15. Outline map of the basement geology of the Spence deposit,
based on drilling carried out by RioChilex. The basement is covered
by 30 to 180 m of piedmont gravels. On the soil sampling traverse,
the gap immediately to the east of the deposit is occupied by a road
and an aqueduct. UTM coordinates are in metres.

mineralization was formed at 57 Ma, followed by supergene


alteration over a long interval from 44 to 28 Ma (Rowland &
Clark 2001). Reserves recoverable by open-pit mining are 50 Mt
of oxide ore with 1.4% Cu, 200 Mt of enriched sulphide ore
with 1.3% Cu, and 150 Mt of primary sulphide ore with 0.6%
Cu. The irregular deposit surface is covered by 30180 m of
gravels.
Soil samples were collected at 1020 cm depth along an
undisturbed eastwest traverse over the centre of the deposit
(Fig. 15), where gravels are c. 100 m in thickness. Measurements
made in the field during sampling on soil deionized water
slurries identified two zones of high conductivity, one directly
over the Spence deposit and another 1 km to the east, shown in
Figure 15 as the Eastern fracture zone. High conductivity is due
to NaCl (Na is shown in Fig. 16), which elsewhere on the
traverse is present in only trace concentrations. The high
conductivity zones also contain higher levels of CaCO3 and
many elements that are abundant in groundwaters of the region,
including Br, I, Li, S and Sr plus mobile indicators of porphyry
mineralization that dissolve as anions: As, Mo, Se and Re.
Caliche is developed in the soils over the Eastern fracture zone,
but is otherwise absent along the traverse. Trenches dug in the
gravels of the high conductivity zones showed that they are
fractured, with north-trending and east-trending sets of vertical
fractures. At a background, low conductivity site, west of the
deposit, gravels are unfractured. The fracture zones are interpreted to represent the upward propagation of basement faults
during the c. 10 Ma interval since the piedmont gravels were
deposited (Cameron et al. 2002). Pampa over the deposit is bare
of vegetation and slopes gently to the west; there are no

19

depressions to explain the saline zones. Sodium chloride is


involatile, so the most reasonable explanation for its presence in
soils together with other constituents of local groundwater is
that mineralized groundwater reached the surface through the
fracture zones. Where the soil sampling traverse crosses the
deposit, the water table is at 60 to 70 m depth. During
earthquakes in this seismically active region, mineralized
groundwaters were pumped up the fracture zones and flooded
the surface (Cameron et al. 2002).
Sampling of groundwaters from 25 drill holes showed that
waters from the deposit are saline (Leybourne & Cameron
2000a). Waters flowing into the deposit from the east have low
contents of Cu, <50 ppb, whereas in the deposit, contents can
exceed 1000 ppb (Leybourne & Cameron 2000b). Downflow,
Cu content falls to <20 ppb (Fig. 17). High contents of Cu in
groundwater resulting from dissolution of ore minerals are
confined to the area of the deposit because Cu2+ is readily
adsorbed by colloids, limiting its mobility. Copper in soils
(Fig. 16) shows a similarly restricted spatial distribution, with
anomalous amounts occurring in soils above the Spence fracture zone but not over the Eastern fracture zone. By contrast,
porphyry indicator elements that migrate as anions are not
adsorbed by colloids. Selenium, for example, which amounts to
<50 ppb in waters upflow from the deposit, increases to as
much as 700 ppb in the deposit, as a result of oxidation of
sulphide minerals of magmatic origin. This high level continues
in the downflow groundwater plume from the deposit (Fig. 17).
Sulphur isotope data on the groundwater show that sulphides
are actively oxidizing (Leybourne & Cameron 2000a). The
energy thus released appears to be stimulating bacterial activity,
possibly including methanogenesis, since there is an unusually
wide range in 13C for dissolved inorganic carbon from 28 to
+9 (Leybourne & Cameron 2000a; Cameron & Leybourne
2002). Although groundwater sampling is rarely used in mineral
exploration, Figure 17 indicates the possible utility of this
method using the far-travelling nature of porphyry indicator
oxyanions.
The ratio of filtered to unfiltered groundwater samples
(Fig. 18) is an eective measure of the mobility of elements in
groundwater. Iron and Al readily precipitate as hydroxides to
form suspended colloidal particles, which are removed along
with adsorbed ions during filtering. These colloids have a
negative charge, which attracts metal cations such as Pb, Cu, Ce
and Y, fixing these cations and restricting their movement.
Other metals Re, Mo, V, As, Se form oxyanions, with
negative charge, that are not attracted to the negatively charged
suspensates. These metals are mobile as anions, along with
conservative cations, such as Na, Ca and Sr. The right-hand side
of Figure 18 is dominated by cations, which either form
colloidal particles or are fixed to these particles, limiting their
mobility. The left-hand side of the figure comprises metals that
form oxyanions or by conservative cations, e.g. Ca2+ and Na2+,
neither of which are adsorbed by negatively charged suspended
colloidal particles.
The dierent behaviours of cations and anions are also
evident at the surface, after flooding by mineralized groundwaters. The rare rains of the region redistributed elements that
arrived with the groundwaters. Copper as Cu2+ is readily
adsorbed, then incorporated into soil minerals; it is retained in
the top 20 cm of a vertical soil profile over the deposit
(Cameron et al. 2002), whereas elements that form anions, such
as As, and conservative cations, like Na, are removed to depth
by the rain (Fig. 19).
Figure 20 shows the proportions of Cu contained in dierent
mineral phases within the same soil profile as shown in Figure
19. Samples were treated sequentially by four leaches. The first

20

E.M. Cameron et al.

Fig. 16. Plots of Na extracted by deionized water and Cu by deionized water, Enzyme Leach, MMI, ammonium acetate, hydroxylamine (HX
Mn) and aqua regia from an eastwest traverse across the Spence copper deposit, Chile. Horizontal scale (eastings) in metres. The background
concentration of Cu away from the deposit represents the endogenic component and that over the deposit is a mixture of the exogenic and
endogenic phases. The aqua regia leach shows the highest proportion of the endogenic phase and, hence, the poorest anomaly to background
contrast.

was ammonium acetate (AA5), to dissolve Cu present in


carbonate. The filtered residues from AA5 were treated with
cold hydroxylamine (HX Mn) to dissolve Mn oxides. The
residues from HX Mn were treated with hot hydroxylamine
(HX Fe) for Fe oxides. The final leach in the sequence was aqua
regia. Anomalous quantities of Cu are confined to the top
20 cm of the profile; the total concentrations of Cu extracted
from samples below 20 cm are similar to the contents of Cu by
aqua regia in the 1020 cm interval in background areas away
from the deposit (Fig. 16). Separate, i.e. non-sequential, analyses
of the samples were done by the Enzyme Leach to provide a
measurement of water-soluble Cu. The amounts extracted by
the Enzyme Leach are small relative to Cu sequestered in
minerals, but also show the enrichment of water-soluble Cu in
the upper part of the profile. This profile, like most others
observed at Spence, shows no clear dierentiation into horizons
either by colour or composition; the soils are reddish-brown
throughout and are not layered with calcrete. The plots to the
right of Figure 20 represent the amounts of the dierent host
phases extracted during the sequential leaching: Ca representing
carbonate, Mn representing manganese oxides an Fe representing iron oxides. Carbonate and iron oxides increase down the

profile, whereas manganese decreases. These plots indicate that


the higher contents of Cu present in the upper 20 cm of the
profile are unrelated the abundance of the host phases. The
results of the sequential leaching experiment is an illustration of
the conceptual model of Figure 11, where Cu derived from the
deposit via the intermediary of groundwater has become
distributed through a variety of secondary minerals present in
the upper 20 cm of soil, including those only soluble in aqua
regia. In the upper 20 cm of the profile all measured phases are
enriched in Cu.
Gaby Sur. Codelcos Gaby Sur porphyry deposit, containing
400 Mt at 0.54% Cu, is 43 km east of Spence. It lies at an
altitude of 2700 m in a broad valley. Like Spence, the climate is
hyper-arid. Supergene-enriched copper oxides, up to 180 m
thick, are underlain by hypogene sulphides (Fig. 21). The
deposit lies in a graben, delimited by high-angle boundary faults
and is covered by up to 40 m of gravel, deposited at c. 9.6 Ma
(Camus 2001). The upper 20 m of the gravel is cemented by
calcrete, making it impermeable to water where unfractured.
Water is found only below the basement unconformity, i.e. at
>40 m depth, where drill holes intersect major faults.

Finding deeply buried deposits using geochemistry

21

Fig. 17. Copper and Se contents of


groundwaters from within and near the
Spence deposit, Chile. Squares are drill
holes where groundwater was sampled,
with analyses as shown.

Fig. 18. Ratio of concentrations of elements in filtered to unfiltered


groundwaters, Spence deposit.

Soils were sampled along an undisturbed eastwest traverse


across the deposit, with samples at 50 m intervals over the
deposit and 100 m intervals on the flanks. The soils are more

complex than those at Spence due to the presence of calcrete


and gypcrete. The sampling procedure was dierent from
Spence and followed the Codelco procedure of taking
red-coloured, Fe-rich soil from within the top 1030 cm of the
profile. In the soils over both boundary faults there are distinct
anomalies for water-soluble Na (Fig. 21) and Cl and for
elements similar to those found to be anomalous at Spence: As,
I, Mo, Se and Re. In soil pits, the calcrete-cemented gravels are
fractured in these saline zones, also similar to Spence. These
features are attributed to pumping of mineralized groundwater
to the surface up the boundary faults and through the overlying
fractured gravels during earthquake activity, followed by
evaporation (Cameron et al. 2002).
Analyses for Cu by dierent extractions are shown in
Figure 21. The results by the weakest leaches, deionized water
and the Enzyme Leach, show clear anomalies over both
boundary faults, with a high anomaly to background contrast.
These anomalies are coincident with the peaks for Na (and Cl).
The similarity in the amounts of Cu extracted by deionized

Fig. 19. Distribution of Cu, Na and As in a vertical soil profile over the Spence deposit from the surface to 100 cm depth. There are no readily
discernible soil horizons. Extraction by hydroxylamine (HX Mn). Prior to sampling the profiles, the top few centimetres of the surface was
removed so as not to include any contaminated material.

22

E.M. Cameron et al.

Fig. 20. Vertical soil profile directly above the Spence deposit. On the left are the amounts of Cu in ppm extracted sequentially by: (1) ammonium
acetate (AA5) dissolving carbonate, (2) cold hydroxylamine (HX Mn) dissolving manganese oxides, (3) hot hydroxylamine (HX Fe) dissolving
Fe oxides, (4) aqua regia. In the centre is Cu in ppb by Enzyme Leach, extracting water-soluble material. On the right is (a) Ca extracted by
ammonium acetate (AA5), (b) Mn extracted by cold hydroxylamine (HX Mn), (c) Fe extracted by hot hydroxylamine (HX Fe).

water and Enzyme Leach indicates that the latter is extracting


mainly water-soluble Cu. The anomalous zones above the faults
that bound the deposit are poorly defined by the stronger
leaches. Ammonium acetate at pH 5 (AA5) dissolves Cu present
in carbonate minerals. Background concentrations of Cu by the
AA5 leach are an order of magnitude higher than for the two
weaker leaches that extract water-soluble material only and
these additional amounts of Cu obscure the Cu derived from
the mineralized groundwaters that reached the surface. Copper
present in carbonate is thus largely of endogenic origin, derived
from primary minerals in the gravel soils. The cold hydroxylamine leach (HX Mn) dissolves Mn oxides, plus carbonate
minerals, plus water-soluble salts. Cumulatively the first two of
these phases have a high content of Cu of endogenic origin,
giving a background concentration of Cu approximately 50
times higher than Cu extracted by deionized water or the
Enzyme Leach. Again, this obscures the Cu of exogenic origin.
For the aqua regia extraction, the background concentration of
Cu reaches 2000 times that of the deionized water and the
Enzyme Leach, again obscuring Cu derived from the deposit via
mineralized groundwaters. Note, however, that one sample of
soil situated at 517450E, above the west boundary fault, has
high contents of Cu by both hydroxylamine (HX Mn) and aqua
regia. This high concentration suggests that, like Spence, there
has been a long period of pumping of mineralized groundwaters
to the surface and the incorporation of Cu of exogenic origin
into secondary minerals.
Mansa Mina. Mansa Mina is a faulted slice of porphyry
mineralization, possibly from the nearby Chuquicamata deposit

(Fig. 22). The fault that cuts the deposit, and forms its eastern
boundary, is the West Fault, a major strike-slip fault of regional
extent that may have had an influence in localizing the
important cluster of porphyry deposits in the Chuquicamata
area (Fig. 22). Mansa Mina is an elongate, steeply dipping
deposit (Fig. 22), containing a mainly hypogene assemblage.
Sulphides account for 300 Mt at 0.95% Cu and oxides 25 Mt at
1.11% Cu. The deposit is up to 300 m in width and over 1000 m
in vertical extent and comprises four steeply dipping panels that
are delimited by faults subsidiary to the West Fault (Sillitoe
et al. 1996). The easternmost panel is barren, comprising
chloritized and pyritized andesitic flows. Panels 2 and 3 are of
sericitized granodiorite containing pyrite-poor, copper porphyry
mineralization. Panel 4 is also highly altered granodiorite, with
sericite and advanced argillic alteration, with a enargite- and
bornite-rich high sulphidation assemblage. There has been
supergene alteration, forming a leached zone and a partially
oxidized zone above the sulphide zone. The upper part of the
sulphide zone has been enriched by chalcocite. The deposit is
entirely covered by Miocene gravels, the minimum thickness of
these being 50 m on the west side of the West Fault and several
hundred metres on the east side. Depth to the water table is not
known.
Mansa Mina lies between the towns of Calama and
Chuquicamata and immediately east of the highway linking
these centres. Only a limited length of undisturbed ground was
available for a sampling traverse across the deposit (Figs 22, 23)
between the highway and gravel dumps to the east. Samples
were taken at irregular intervals to avoid vehicle tracks. The
ground slopes gently from north to south. Soils are marked by

Finding deeply buried deposits using geochemistry

23

Fig. 21. Plots of Na extracted by deionized water and Cu by deionized water, Enzyme Leach, ammonium acetate, hydroxylamine and aqua regia
for soils from an eastwest traverse across the Gaby Sur copper deposit, Chile. Horizontal scale (eastings) in metres.

strong development of both gypcrete (near the surface) and


calcrete, which first occurs at depths between 20 and 40 cm.
Samples were taken immediately above the calcrete horizon to
minimize inclusion of gypcrete. Results of Enzyme Leach
analyses show a strong spike anomaly for Cl (Fig. 23) and Na
in the soil immediately above the West Fault. There is a
coincident spike anomaly for Se (Fig. 23). All samples contain
S, because of the presence of gypcrete, but the gypcrete has a
low Se content, and the median ratio for Se1000/S across
the entire traverse is 0.07, in the range of ocean water. For the
sample directly above the West Fault the ratio is 1.9, in the
range of the oxidation products of igneous sulphide. The same
sample is anomalous in Re (31 ppb by Enzyme Leach, 48 ppb
by aqua regia), an indicator element for copper and copper
molybdenum porphyries, since it mainly occurs as an impurity
in molybdenite. Other elements that are strongly anomalous in
this sample are Br and I, typical constituents of groundwaters of
the region, and K, Rb and Tl, which are abundant in sericitic
alteration. An anomaly for Mo (Fig. 23) and Cu are displaced
60 m to the east. The results suggest that mineralized groundwater has been pumped up the West Fault during seismic
activity, followed by its evaporation at the surface.

Fig. 22. On the left is an eastwest cross-section across the Mansa


Mina deposit after Sillitoe et al. (1996). The unpatterned area between
the sulphide zone and the fault is barren rock. On the right is a map
showing the spatial relationship between the West Fault and the
cluster of copper porphyry deposits in the Chuquicamata area. For
location of area shown, see Figure 15.

24

E.M. Cameron et al.

Fig. 24. Map showing the Carlin Trend (shaded) and some of the 40
gold deposits found along this trend.

localized by structural features. The evidence is consistent with


groundwater having been forced to the surface up fracture
zones by earthquake-induced (seismic) pumping.
Studies in Nevada
More than 40 Au occurrences within the Carlin Trend in
northeastern Nevada (Fig. 24) form the largest accumulation of
mineable Au in North America. One of these is the Mike
deposit.

Fig. 23. Analyses of Cl, Se and Mo by Enzyme Leach in soil samples


from a westeast traverse across the Mansa Mina deposit at
7526500N.

Summary of observations and interpretation at Chilean sites. At all three


sites there is evidence of saline groundwater having reached the
surface and evaporated. In addition to Na and Cl, zones in soils
are enriched in elements typical of groundwater in the region, I
and Br, and porphyry indicator elements, Se, Re, Mo and Cu.
Over the Spence deposit enrichment occurs above a fracture
zone in the gravels, and above a water table at 6070 m depth.
At Gaby Sur the groundwater signature is seen in soils above
the boundary faults of the deposit; the water table is at >40 m
depth. At Mansa Mina the groundwater signature occurs
directly above a major fault in the basement below the gravel
cover. The depths of the present water table at Spence and
Gaby Sur are too great to permit groundwater to be raised to
the surface by capillarity. In a coarse-grained medium, such as
gravel, soil suction can draw water up only a few metres at best.
Moreover, given the permeable nature of gravel, groundwater
raised by capillarity would extend over a wide area and not be

Mike Deposit. This deposit, hosted by sedimentary rocks of


Palaeozoic age, was discovered by Newmont Corporation in
1989 while drilling on the predicted extension of the NWtrending Good Hope fault (Fig. 25) into an area with a thick
cover of Carlin Formation of Tertiary age. The intersection of
this fault with NE-trending faults was the primary structural
control for the previously discovered Tusk and Gold Quarry
deposits. The Good Hope fault is the boundary between two
portions of the Mike deposit: the Main Mike, containing
43.2 Mt of 0.034 oz/t Au and 76 Mt of 0.22% Cu; and the West
Mike containing 110 Mt of 0.025 oz/t Au and 74 Mt of 0.28%
Cu.
Copper mineralization consists of a sub-horizontal supergene
oxide and sulphide blanket up to 120 m thick that mainly
underlies, but also overlaps, Au mineralization. Supergene
processes may have caused the mobilization of Cu to greater
depths than Au. Teal & Branham (1997) note enrichments in
the ore of Cu, Au, Zn, Ag, Bi, Mo and Te; in addition, Cd and
Se may be enriched. The deposit is covered by up to 240 m of
Carlin Formation of Eocene age comprising piedmont gravel,
finer clastic sediments, waterlain tu and a basal conglomerate
and regolith that contains mineralized (oxidized) clasts. A
perched water table in the Carlin Formation is c.50 m below the
surface and a second aquifer is in the Palaeozoic rocks (Jackson,
2000).
The simplest mechanism for generating geochemical
anomalies through 240 m of post-mineral cover is by the

Finding deeply buried deposits using geochemistry

Fig. 25. Distribution of copper mineralization in the Mike deposit in


relation to major faults in the basement, shown by dashed lines.
Information courtesy R.G. Jackson and Newmont Corporation. Solid
NNE lines show two interpreted Pleistocene faults that cut the Carlin
formation and intersect the sampled Line 3 (see Fig. 27).

movement of fluids or gases up faults in this cover.


Dohrenwend & Moring (1991) carried out photo-geological
interpretations of recent faulting in this region. Of several
criteria used to distinguish faulting, the most relevant to the
Mike area is: prominent alignments of linear drainageways,
ridges and swales, active springs or spring deposits, and linear
discontinuities of structure, rock type, and vegetation. This
faulting, which they assign to early to middle Pleistocene time
(0.13 to 1.5 Ma), with a mean orientation of 028o, results in
dissection of the surface along faults of greater than 10 m. The
topography of the Mike area (Fig. 26) shows deeply incised dry
stream beds with an orientation close to 028o. Based on this
topography, two faults in the Carlin Formation are interpreted
to intersect the sampled Line 3. The fault to the NW, Channel
A, is marked by a stream and floodplain. Channel B was dry
during our visits. Its east slope is precipitous, which may
indicate a fault scarp.
Climate is semi-arid, with sparse sagebrush vegetation. Soils
were collected along Line 3 from sites at 30 m intervals; at each
site, samples were taken from a depth of 4050 cm from five
holes dug within a radius of 1.5 to 3 m. The sampling depth was
selected on the basis of earlier studies by Jackson (2000). The
five samples were mixed to form a composite sample. Soils are
immature with a weak B-horizon below 1530 cm. Relatively
little carbonate was found along this line and soils are mostly
low in organic material, except in the alluvial soils around the
stream that marks Channel A.
Figure 27 shows results for Cd, Cu, Mn and Au by aqua
regia. There are anomalies for Cd in the alluvial soils centred
over Channel A. However, the strongest anomaly for Cd is at
Channel B, located in residual soils of the steep west-facing
(scarp?) slope of the dry valley, not in the vicinity of the
stream-bed. Copper and Au are anomalous only on Channel B.
The highest value for Au by aqua regia, at 17 ppb, is 9 times the
median value of 1.9 ppb, the median being a reasonable proxy
for the local background. Selective leaches were also used. The
results by hydroxylamine (HX Mn) are mostly below the
detection limit of 0.05 ppb Au. The MMI maximum value of
6.2 ppb Au over Channel B is 14 times the median. Many of the

25

Fig. 26. Topography of the immediate Mike area, with contours at


25 ft intervals. The dashed lines show interpreted Pleistocene faults
in the Carlin Formation, based on the criteria of Dohrenwend &
Moring (1991), who defined an orientation of 028o, as shown by
the arrow. These interpreted faults intersect the sampled Line 3 at
sites A and B.

Enzyme Leach results for Au for this line are below the
detection limit, but this extraction also shows an anomaly on
the east side of Channel B. For Cu, the results by the four
extractions are shown in Figure 28. The Channel B anomaly is
shown clearly by the three leaches that extract the most Cu
aqua regia, hydroxylamine (HX Mn) and MMI but the
anomaly is not apparent in the Enzyme Leach results.
In an earlier section we have discussed anomaly formation as
an incremental process, with metal being introduced in watersoluble form and then progressively incorporated into the
secondary minerals of the soils. If the anomaly associated with
Channel B is related to Pleistocene faulting of the Carlin
Formation, then the anomaly has had up to 1.5 Ma to form.
This time span may account for the strong response shown by
the aqua regia extraction, which dissolves most secondary
minerals in the soil. Anomalies can also, presumably, stop
developing when new batches of elements in water-soluble
form are no longer introduced. This may account for the
anomaly response of the Enzyme Leach, which extracts watersoluble metal, being lesser than that of the stronger leaches that
dissolve secondary minerals.
Other elements that show strong anomalies on the steep
west-facing slope of Channel B include Mo by aqua regia, Cd by
hydroxylamine (HX Mn), Zn by MMI and by aqua regia, and
Hg by aqua regia. Elements that are anomalous in the soils are
those enriched in the ore, so that their ultimate source in the
deep mineralization is a reasonable supposition. Elements that
are not components of the ore, such as rare earth elements, do
not have anomalous patterns in the soil. Jackson (2000), after
collecting samples on a wide-interval grid over Mike, found that
several elements were depleted across the top of the deposit,
notably Mn; this was confirmed by our sampling (Fig. 27).
Jackson suggested that elements may be transported to the
surface laterally and vertically around the deposit to create
a central low with flanking highs. Manganese oxide is a
redox-sensitive mineral that may be dissolved in a reducing
environment. A sulphide deposit is a large mass of reduced
material, which, when oxidized, must reduce an equivalent mass

26

E.M. Cameron et al.

Fig. 27. Analyses for Mn, Cu, Au and


Cd by aqua regia in soils from Line 3
across the Mike deposit. Note that the
horizontal axis is in feet, corresponding
to the original Newmont survey.

Fig. 28. Analyses for Cu by aqua regia,


hydroxylamine, MMI-A and Enzyme
Leach in soils from Line 3 across the
Mike deposit. Note that the horizontal
axis is in feet, corresponding to the
original Newmont survey.

of oxidized material. So a possibility exists that the Mn low over


the deposit is a remote product of the oxidation of Mike
sulphides, i.e. a reduced chimney. But elements that are not
redox sensitive also show a low over the deposit: Cs, Ti, Zr and
Al. So facies change in the Carlin Formation to a sandier
sediment overlying the deposit, and more enriched in resistate
(heavy) minerals, may explain some of the geochemical results.
In addition to the analyses of the soils by selective leaches,
other studies were carried out. In July 1999, a CO2 O2 soil gas
survey was carried out at 159 sites on three lines by Patrick
Highsmith and Mary Doherty. The sampling depth for the
probe was c. 40 cm and the principal sampling interval was
30 m. Repeated measurements at a base station showed that soil
gas concentrations were stable during the sampling period.
Delta values were calculated relative to atmospheric concentrations for these gases. For CO2 the range was 0.08 to 0.93%,
with a median of 0.46% and for O2 the range was 0.1 to
0.8%, with a median of 0.35%. There is an inverse
correlation between CO2 and O2, which is consistent with
the consumption of O2 to produce CO2. The amount of CO2
is lower than values reported by Lovell & Hale (1983) for soil
gas over a variety of massive sulphide mineralization, but is

similar to data obtained over the Carlin Trend by McCarthy &


McGuire (1998). Higher CO2 values and corresponding lower
values for O2 were found both over the deposit and in areas
remote from it, but the survey was not extended to where Line
3 intersects Channel B, the area of strongest metal response.
The low delta values and anomalous values away from the
deposit, make it is dicult to judge whether variation is due to
deep mineralization, or near-surface features, or normal variations of gas within soils.
Pauwels et al. (1999) described a method for the collection of
metals in soil air using activated carbon held inside a bag of
Gore-Tex within in a cylindrical plastic container. They carried
out orientation surveys in southern Spain, where measurements
of CO2, He and Rn showed that gas is emanating from fractures
that cut ore bodies in the Iberian Pyrite Belt. Collectors were
installed in the soils and recovered after 100 days. Elution and
analysis of the activated carbon from the Iberian Belt showed
metal anomalies coincident with CO2 anomalies. Courtesy of
Philippe Freyssinet of BRGM, similar collectors were installed
on Line 3 at the Mike deposit in May 2000 and retrieved in
August 2000. The collectors were placed at the top of a 50 cm
length of PVC pipe inserted into auger holes, leaving the

Finding deeply buried deposits using geochemistry

27

Fig. 29. The western portion of the Abitibi greenstone belt (shaded)
showing the location of the four study sites. Much of the area to the
west of Rouyn-Noranda and Matagami has a discontinuous cover of
fine-grained glaciolacustrine sediments.

collectors above ground. Forty-four collectors, including


duplicate sites, were installed and three blank collectors were
retained within plastic bags. Two of these blanks were kept in
the field and one in the laboratory. After shipping to BRGM,
the carbon was eluted and the solution analysed by ICP-MS.
Contents above the detection limit were obtained for As, Cr,
Co, Cu, Ni, Pb and Zn. However, in all cases, the contents
obtained from the collectors installed in the field were no
dierent from the blanks, i.e. no measurable amounts of metal
were detected in soil air. Similar collectors were installed in the
piedmont gravel soils of the fracture zone above the Spence
deposit and at a background location west of the deposit. These
were retrieved after one year. Elution and analysis of the carbon
showed that metals were no more abundant at the sites over the
deposit than at the background sites.
Note added in proof
Newmont carried out a major drilling program at the Mike
deposit from 1997 to 2000 and the results have been described
by Norby and Orobona (2002). The principal soil anomalies at
location B (Fig. 27) are now known to occur at the surface
intersection of the Nebulous fault, a NNE-trending post-Carlin
fault. In addition to cutting the Carlin Formation, it cuts and
displaces the goldcopper oxide zone, and a lower sulphide
zone at c. 500 m depth now described by Norby and Orobona.
The top of the sulphide zone is distinguished by a 60 m thick
blanket of sphalerite-rich rock. Cadmium is a common trace
constitutent of sphalerite. This explains why Cd shows such a
clear anomalous pattern (Fig. 27) above a goldcopper deposit.
There is localized oxidation of the sulphide zone where it is
intersected by the Nebulous fault. Other elements present in the
sulphide zone, including Zn, Ag, As and Ni, are also anomalous
in the soils along the scarp slope of the Nebulous fault. The
elements are interpreted to have moved 500 m up the fault from
the sulphide zone. Recent isotopic studies by Dublyansky et al.
(2003) at the Yucca Mountain nuclear waste disposal site, also
in Nevada, have shown that fluids of deep-seated origin have
moved up several hundred metres through a thick vadose zone
along a permeable fault.
Studies in Ontario
Four sites were studied in the Abitibi greenstone belt of Ontario
(Fig. 29). This belt is host to a number of world-scale deposits,

Fig. 30. Geological cross-section along Line 6 at Cross Lake, based


on diamond drill hole data by Cross Lake Minerals Ltd and
overburden drilling by Ontario Geological Survey. Superimposed are
a spontaneous potential (SP), ORP measurements of redox conditions at 1.5 m below water table with a two-minute reading,
carbonate content of B-horizon soils, and Zn content by MMI
analysis of soils taken at a depth of 10 to 25 cm below the base of
decomposing vegetation.

such as the Kidd Creek volcanogenic massive sulphide (VMS)


deposit near Timmins, the Hollinger, Dome and McIntyre Au
deposits at Timmins and the Lake Shore and Kerr-Addison Au
deposits near Kirkland Lake. Much of the belt has a discontinuous cover of glaciolacustrine clays and other finegrained glacial sediments (Fig. 29). Most of the known deposits
are exposed in windows in this glacial cover. There is reason to
expect that undiscovered deposits lie hidden beneath glacial
sediments. We here discuss some of the results for one of the
sites, Cross Lake.
Cross Lake. The Cross Lake VMS mineralization of Archean
age, located 50 km SE of Timmins, is similar in mineralogy,
composition and age to the well-known Kidd Creek ZnCuAg
deposit. Host rocks are felsic pyroclastic rocks, intruded by
feldspar porphyry and diabase dykes. The immediate host is tu,
lapilli tu and chert, with sericitic and chloritic alteration. On
Line 6 (Fig. 30), sulphides in the mineralized lens are, in order
of abundance, pyrite, honey-coloured sphalerite, chalcopyrite
and galena. Zinc dominates over Cu with assays as high as 26%
compared to a maximum of 1.3% Cu. The width of the VMS
subcrop underlying Line 6 is about 25 m and is covered by 30 m
of varved clay with silty laminae that was deposited in a glacial
lake c. 8,000 years. The area is wet, with groundwater close to
or at the surface in low areas.

28

E.M. Cameron et al.

Fig. 31. Analyses for Zn and Cd by


Enzyme Leach and Ca by ammonium
acetate on soil samples taken from the
upper 10 cm of B-horizon and the
1020 cm interval of B-horizon along
Line 6 at Cross Lake. Also shown are
pH values for soil-water slurries. The
VMS subcrop is beneath 30 m of silty
clay and extends from 175 m to 200 m
along Line 6. Locations of samples A,
B and C for the 010 cm interval of the
B-horizon (see lower left plot) relate to
Figure 32.

A spontaneous potential (SP) survey showed a low of 15 mV


directly over the sulphides (Fig. 30). Initial sampling at sites
25 m apart of soils taken from a thick 20 cm interval of the
B-horizon and also of humus (H), followed by analysis by a
variety of methods, gave only weak base metal anomalies over
the mineralization. But there is a pH low for soil-water slurries
over the mineralization and strong peaks for Ca, Mg and CO3
on the flanks of the pH low. Unweathered clay contains
carbonate, but this is removed during the formation of
B-horizon soils. Thus the carbonate peaks in the soils flanking
the pH low is unusual for B-horizon soils in this clay terrain.
Separate soil samples were collected specifically for MMI
analysis over a constant interval 10 to 25 cm below decaying
vegetation. These samples were mainly B-horizon with an
admixture of Ae and/or Ah soil at some sites. They showed
distinct Zn anomalies over the mineralization that were absent
in the B-horizon samples collected over a thicker and deeper
interval from that horizon. These disparate results for two sets
of samples collected at the same time from the same sites
suggested a depth control on Zn contents in the soils.
Follow-up sampling over a restricted length of the traverse,
centred on the mineralized interval, collected (a) the upper
010 cm interval of the B-horizon and (b) the 1020 cm interval
of the B-horizon. The results (Fig. 31) show that anomalous
levels of Zn are confined to the uppermost part of the
B-horizon, being absent in samples taken immediately below.
Sequential leach analyses were carried out by the Geological
Survey of Canada on selected samples from the 010 cm
interval of the B-horizon. These samples (Fig. 32, with locations
shown on Fig. 31) comprise sample A from a site away from
the mineralization and samples B and C directly above the
mineralized subcrop. The first leach in the sequence was
ammonium acetate at pH 7 (AA7). This dissolves carbonate
minerals plus water-soluble elements. There is very little carbonate in the samples, as shown by total CO3 concentrations of
ranging from 300 to 800 ppm for the three samples. Nevertheless, this leach dissolves what carbonate is present, as shown by
analyses for Ca (Table 5).

The location of the samples is shown in Figure 32. Sample A


is from a background area, whereas samples B and C overlie the
mineralization. The abbreviated headings for the dierent
sequential leaches are explained in the text.
The Zn analyses by this leach show a strong anomaly/
background contrast between samples B and C over the
mineralized subcrop compared to background sample A. After
leaching with AA7, the sample residues were treated with
ammonium acetate at pH 5 (AA5). This is a stronger extractant
for carbonate, but little, if any, carbonate is inferred to have
remained after the previous treatment, since the amounts of Ca
extracted are low. This leach extracts more Zn that AA7 and the
anomaly/background for AA5 is greater than for AA7. The
next leach, cold hydroxylamine (HX Mn), extracts Mn oxides.
This leach dissolves more Zn from samples B and C than from
background sample A, but the anomaly/background contrast is
less than either of the acetate leaches (Table 5). The next leach,
hot hydroxylamine (HX Fe) dissolves Fe oxides. This extracts
substantially more Zn than any of the preceding leaches, but
this is endogenic material, not derived from the mineralization,
and no anomaly is apparent over the mineralized zone. The fifth
leach, aqua regia, presents a similar pattern, with much Zn
extracted, but no anomaly.
Comparative data are given in Table 5 for separate (nonsequential) analyses by the Enzyme Leach and MMI-A. Enzyme
Leach extracts less Ca, somewhat less Mn, and an equivalent
amount of Fe to leach AA7. Less Zn is extracted by the Enzyme
Leach than AA7, but the anomaly/background contrast is
high. No major elements are reported with MMI-A analyses.
However, the anomaly/background contrast for Zn by MMI-A
is excellent. The data shown in Table 5 and Figure 32 illustrate
the eectiveness of weak leaches in selectively extracting the
exogenic phase of an indicator element and separating this
from the much more abundant endogenic phase that is derived
from the primary minerals of the clay sediment, the parent of
the soil.
Concurrent with the soil sampling, drilling was carried out by
the Ontario Geological Survey. Shallow boreholes were drilled

29

Finding deeply buried deposits using geochemistry

Fig. 32. Zinc in ppb by sequential leaches of background sample A and samples B and C over a VMS subcrop below 30 m of clay. Samples
are from the upper 10 cm of the B-horizon, Line 6 at Cross Lake. Locations of samples are shown in Figure 31, Zn plot. Upward projection
of the VMS subcrop is indicated by the hatched block.

to 9 m depth to characterize the overburden and subsurface


geochemical, pH and redox conditions. Lower redox values, as
measured by ORP, outline a reduced column in the clay
substrate above the mineralization (Fig. 33). This feature is
replicated at the Marsh Zone site. Below the water table, pH is
relatively constant in a neutral to weakly alkaline range, e.g., at
1.5 m below the water table the pH ranges from 7.1 to 7.8. But
at the water table and above there is a substantial drop in pH
directly over the mineralization, to values as low as 5.0.
Fieldwork at Cross Lake provided an opportunity to test
the model of Hamilton (2000). The empirical observations
summarized in Figure 31 show a close similarity to the
theoretical predictions summarized in Figure 5. Oxidation of
the sulphides at depth, suggested by SP lows, releases a rising
column of reduced material. Below the water table, access to
oxygen and other oxidized species is restricted, but where the
column intersects the water table, the reduced species are

oxidized, e.g.: Fe2+ + YO2 + 2H2O h Fe(OH)3 + H+.


Formation of hydrogen ions lowers the pH of the soils at and
above the water table, dissolving CaCO3, which migrates
laterally and precipitates on the flanks of the reduced zone in
areas of higher pH. The reduced column must include base
metals present in the sulphide mineralization, most notably Zn
and Cd, which show as anomalous concentrations in a narrow
zone at the top of the B-horizon (Fig. 31). The width of the
acidic zone and anomalous levels of Zn and Cd in the upper
B-horizon are 65 to 70 m, with sharp boundaries (Fig. 31).
Given the estimated width of 25 m for the VMS lens at the
subcrop beneath 30 m of clay, the upward dispersion of
material is fairly narrowly constrained. Substantial amounts of
reduced material are involved to create the zones of low pH and
dissolve carbonate at the surface. Lead extracted from the soils
above the mineralization is less radiogenic than that sampled
away from the mineralization (Cameron et al. 2001). This

Table 5. Data for Ca, Mn, Fe and Zn for a series of sequential leaches and for two single leaches of samples from the 010 cm interval of the B-horizon on Line 6 at Cross Lake.
Sequential leaches
Sample
Ca (ppm)
A
B
C
Mn (ppm)
A
B
C
Fe (ppm)
A
B
C
Zn (ppb)
A
B
C

AA7
1710
560
490
1.2
4.1
2.2
2.2
14
10
42
435
317

AA5
288
106
84
5.1
5.2
2.5

HX Mn
654
62
213
48
3.5
2.0

Single leaches
HX Fe

Aqua regia

294
65
135

1513
1039
1032

64
46
39

100
79
76

1.1
2.1
0.9

2.7
15
10

23
164
135

118
344
252

2800
4500
3600

11400
7900
7200

103
1500
1410

303
970
590

8200
7700
7400

25000
19100
18700

Enzyme Leach
205
142
99

7
270
186

MMI-A

64
1930
1670

The location of the samples is shown in Figure 32. Sample A is from a background area, whereas samples B and C overlie the mineralization. The abbreviated
headings for the dierent sequential leaches are explained in the text.

30

E.M. Cameron et al.

Fig. 33. Model for development of anomalies through clay cover


incorporating empirical observations from the Cross Lake test site.
There is a reduced column rising above the sulphide subcrop through
the clay cover. At and above the water table, metal ions in this
reduced column are oxidized by infiltrating oxygen, which generates
hydrogen ions. The resulting low pH environment dissolves carbonates from C-horizon clay soils and this carbonate is reprecipitated on
the flanks of the zone of low pH. The Zn2+ and Cd2+ ions are
inferred to have migrated upwards in the reduced column because
anomalous amounts have accumulated in a narrow zone at the top of
the B-horizon.

suggests that Pb2+, as well as Zn2+ and Cd2+ are being


transferred upwards in the reduced column, because of the
non-radiogenic nature of Pb present in VMS, compared to that
of clay. Trenching showed that tree roots do not penetrate
deeper than the water table, indicating that roots are not
involved in the advective transfer of mineralized groundwater
from depth to the surface. Given the short length of time since
the glacial sediments were deposited, c. 8,000 years, and the low
rates cited above for chemical diusion, diusion cannot be the
principal means for moving reduced material to the surface.
Despite the close coincidence between the Hamilton
theoretical model and empirical observations, the high Zn and
Cd concentrations in the soils above the reduced column
cannot be easily accounted for by electrochemical processes.
These elements are usually mobile as Zn2+ and Cd2+, which are
not redox-active; in this form their migration should not
determined by redox gradients. Hamilton et al. (2001, 2002)
reported further intriguing data from the drilling studies at
Cross Lake. Over the mineralized zone there is an upward
bulge in the piezometric surface of close to 50 cm on both
lines, despite downward hydraulic gradients between the overburden and bedrock piezometers. Near-surface groundwater
also shows an increase of c. 1C over the mineralized zone;
measurements that have been replicated at dierent times.
Similar features have been noted at the Marsh Zone and other
sites where reduced columns are documented within otherwise
oxidized overburden. Much remains to be done to understand
transport mechanisms and to better define changes in temperature and the piezometric surface that may be related to
mineralization. This is being undertaken as part of a successor
project at Cross Lake led by G.E.M. Hall and funded by the
Ontario Mineral Exploration Technology programme and the
Geological Survey of Canada.
CONCLUSION
In the Deep-Penetrating Geochemistry project we have
considered the theoretical basis for the upward movement of
material from buried deposits and have carried out field tests in
Chile, Nevada and Ontario to examine transport processes and
compare methods for detecting anomalies. For arid or semi-arid
areas, with a thick vadose zone, upward diusion of dissolved

elements is orders of magnitude slower than the downward


movement of water films. The only possibilities for the upward
transport of elements are mass (advective) transfer of water plus
dissolved constituents, or air plus indicator gases or aerosols.
The most striking example of advective transport is in
Chile where the earthquake-induced (seismic) pumping of
mineralized groundwater to the surface through fracture zones
is recognized. The assemblage of elements found in the soils
above fracture zones is similar to that found in groundwaters
and, at Spence, Cu in soils and Cu in groundwaters at
60 m depth are only anomalous in the vicinity of the deposit.
Faults or fracture zones are an essential conduit for bringing
mineralized groundwater to the surface and we find a spatial
correlation between anomalies in soils and neotectonic
structures, which appear to represent reactivation of more
ancient faults. In Nevada, at the Mike deposit, covered by
240 m of Carlin Formation, we also find strong anomalies
associated with interpreted neotectonic structures, but here we
have no evidence of the transport mechanism.
The Abitibi region stands in diametric contrast to Chile and
Nevada: the glacial clay and sand that cover bedrock are only a
few thousand years old, lack fracturing, and are water-saturated
to near the surface. Here too soils show clear signs of buried
mineralization and the empirical observations are remarkably
similar to those predicted by the model of Hamilton. This
involves oxidation of the buried sulphide mass and the upward
transfer of reduced material, including base metals. At the water
table this reduced material reacts with O2, creating an acidic
environment that dissolves and redistributes carbonate
originally present in the clay. The substantial flux of reduced
substances to the surface, the elevated concentrations of Zn and
Cd in the soils over mineralization, bulges in the piezometric
surface, and temperature increases of groundwater at the same
locations cannot be accommodated by diusion processes;
additional mechanisms are required to explain the empirical
data.
Barometric pumping provides a means for the rapid
transport of air plus gas or aerosols. At the Mike deposit,
Nevada, tests of the metal content of soil air using collectors
containing activated carbon failed to detect greater amounts of
metal than blank collectors sealed in plastic bags. Similar tests in
the piedmont gravel soils on the fracture zone above the Spence
deposit also failed to detect metal concentrations that were
greater than a nearby background area. At Mike, CO2 and O2
measurements in soil air showed variation, but it was not
possible to conclude that these were related to the deposit.
There is a need to carry out tests on air/gas mixtures released
from fractured rock over known mineralization to identify the
constituents present.
By far the most widely used geochemical method for the
discovery of buried deposits is selective leach analyses of soils;
this study has focused on these methods. We argue that
geochemical anomalies in soils over buried mineralization form
incrementally. Metals of external, exogenous origin enter the
soil in water-soluble form and are progressively incorporated
into secondary minerals, such as carbonates, and iron and
manganese oxides. In the case of the Chile and Nevada test
sites, the time available for anomaly formation may have been
a million years, or longer. At these sites there has been time for
metal of exogenous origin to be incorporated in a variety of
secondary minerals. Thus the anomalies can be detected by a
strong leach, aqua regia, which dissolves all secondary minerals,
although the anomaly/background contrast is less than for
selective leaches that target specific secondary minerals. In the
clay-covered terrain of the Abitibi region, several thousand
years only have been available for anomaly formation. In this

31

Finding deeply buried deposits using geochemistry


case metals from the mineralization have only been incorporated into the most labile secondary minerals and anomalies over
mineralization cannot be recognized by strong leaches.
There are two strategies for the choice of leaches: (a) measure
the water-soluble phase before it has entered secondary minerals; or (b) selectively dissolve one or more secondary minerals.
A deionized water extraction or the Enzyme Leach both
measure the water-soluble phase, but the latter gives better
reproducibility. The amounts of metals extracted by either leach
are similar. To dissolve secondary minerals we have tested
ammonium acetate for carbonates, cold hydroxylamine (HX
Mn) for Mn oxides, hot hydroxylamine (HX Fe) for Fe oxides,
and MMI-A for undetermined minerals. The usefulness of these
methods for exploration purposes depends on the ratio of
exogenic metal to endogenic metal in the target mineral. Where
the ratio is high, a good anomaly/background contrast is
obtained, as for Cu in carbonates in Spence soils or for Zn
associated with carbonates at Cross Lake. In another case, at
Gaby Sur, the amount of Cu of endogenic origin in carbonate
is too high and obscures the Cu derived from the deposit. Other
than the leaches that target the water-soluble phase, leaches
usually measure the cumulative content of two or more phases.
Ammonium acetate at pH 5 measures metal present in carbonate plus that which is water-soluble. Hydroxylamine and
MMI-A are acidic leaches, so dissolve carbonate as well as, for
example, the Fe and/or Mn oxides which hydroxylamine is
targeted to dissolve. Neutralization of leach acidity by carbonate
reduces their ability to dissolve and hold metals in solution,
whereas ammonium acetate is buered at constant pH.
The depth at which soil samples are taken is critical. At
Spence, the principal indicator element, Cu, is confined to the
top 20 cm of the profile, whereas other porphyry indicator
elements, such as As and Re, which dissolve as anions, have been
removed by rainfall below 40 cm. In the Abitibi region, at Cross
Lake, only sampling that includes a critical metal-accumulation
horizon near the top of the B-horizon proved eective.
As will be apparent to the reader, this study could not have been
carried out without the cooperation of many people in the mining
industry and in government. We thank Ollie Bonham and Jack Currie
of RioChilex and Gordon Gray and Kelly OConnor of Rio Algom
for many courtesies during our work at Spence. Sampling at Gaby
Sur was done collaboratively with Ricardo Venegas and Aldo
Vnegas of Codelco. E.M.C. and M.I.L. were assisted in the field in
Chile with enthusiasm and humour by Daniel Salinas and Alexi
Ramirez. George Steele of Rio Tinto kindly gave logistical support
for the Chile work. At Mike, Robert Jackson of Newmont Inc.
provided a great deal of help and advice. Our sampling at Mike was
carried out together with Mary Doherty and Kevin Creel. The CO2
gas sampling at Mike was by Patrick Highsmith and the collectors
containing activated carbon were provided by Philippe Freyssinet,
who also provided the analyses of the carbon. For the Cross Lake
studies we are indebted to Robert Middleton and Ian Millar-Tate of
Cross Lake Minerals for providing maps and drill hole information.
Brian Polk and Devin Cranston worked with us throughout the
sampling programme in the Abitibi. Colin Dunn carried out the
biogeochemical studies at Cross Lake on a tight time frame near
the end of the project. We thank Richard Alcock, Research Director
of CAMIRO; Bill Coker, chair of the CAMIRO Geochemistry
Committee; and 26 company sponsors for their support and encouragement throughout the project. The fruitful collaboration with
the Ontario Geological Survey would not have been possible without the enthusiastic support of Cam Baker. The participation of
Geological Survey of Canada sta and facilities is gratefully acknowledged. Sample preparation was by Overburden Drilling
Management. Analyses were provided at no charge by John
Gravel of Acme Laboratories, Eric Homan of Actlabs, Patrick
Highsmith of ALS-Chemex, Claude Massie of Bondar-Clegg,
Robert Ellis of Gedex and Hugh DeSouza of XRAL and Xie
Xuejing of IGGE.

REFERENCES
AMUNDSON, R.G. & DAVIDSON, E.A. 1990. Carbon dioxide and nitrogenous
gases in the soil atmosphere. Journal of Geochemical Exploration, 38, 1341.
ARIMOTO, R., RAY, B.J., DUCE, R.A., HEWITT, A.D., BOLDI, R. & HUDSON, A.
1990. Concentrations, sources, and fluxes of trace elements in the remote
marine atmosphere of New Zealand. Journal of Geophysical Research, 95,
2238922405.
BLAIN, C.F. & BROTHERTON, R.L. 1975. Self-potentials in relation to oxidation
of nickel sulphide bodies within semi-arid climatic terrains. Institute Mining
Metallurgy Transactions, Section B, 84, 123127.
BOLVIKEN, B. & LOGN, O. 1974. An electrochemical model for element
distribution around sulphide bodies. In: ELLIOT, I.L. & FLETCHER, W.K.
(eds) Geochemical Exploration 1974. Special Publication, 2. Association of
Exploration Geochemists, 631648.
BRIGGS, R.C. & TROXELL, H.C. 1955. Eect of Arvin-Tehachapi earthquake on
spring and stream flow. In: OAKESHOTT, G.B. (ed.) Earthquakes in Kern
County, California, during 1952. Part 1. Division of Mines Bulletin, 171.
Department of Natural Resources, California, 8198.
CAMERON, E.M. 1998. Deep-Penetrating Geochemistry. Report. Canadian Mining
Industry Research Organization (CAMIRO).
CAMERON, E.M. & LEYBOURNE, M.I. 2002. Isotopic compositions of oxygen and
carbon from carbonate and sulphur from sulphate in soils from the Spence and Gaby Sur
porphyry deposits and from Tamarugal, northern Chile. Report. Canadian Mining
Industry Research Organization (CAMIRO).
CAMERON, E.M., COUSENS, B.L. & HALL, G.E.M. 2001. A pilot study for the use
of lead isotopes to determine the genesis of anomalies above buried VMS mineralization,
Cross Lake. Report. Canadian Mining Industry Research Organization
(CAMIRO).
CAMERON, E.M., LEYBOURNE, M.I. & KELLEY, D.L. 2002. Exploring for
deeply-covered mineral deposits: formation of geochemical anomalies in
northern Chile by earthquake-induced surface flooding of mineralized
groundwaters. Geology, 30, 10071010.
CAMUS, F. 2001. Descubrimiento de porfidos de cobre en zonas cubiertas: El ejemplo de
Gaby Sur. 20th International Geochemical Exploration Symposium.
CANNON, H.L. & STARRETT, W.H. 1958. Botanical prospecting for uranium on
La Ventura Mesa. Sandoval County, New Mexico. United States Geological Survey
Bulletin, 1009M, 391407.
CARRIGAN, C.R., HEINLE, R.A., HUDSON, G.B., NITAO, J.J. & ZUCCA, J.J. 1996.
Trace gas emissions on geological faults as indicators of underground
nuclear testing. Nature, 382, 528531.
CLARK, J.R., YEAGER, J.R., ROGERS, P. & HOFFMAN, E.L. 1997. Innovative enzyme
leach provides cost-eective overburden/bedrock penetration. Abstract,. Exploration
97.
CROZAT, G., DOMERGUE, J.L. & BOGUI, V. 1973. Etude de laerosol atmospherique en Cote dIvoire et dans le Golfe de Guinee. Atmosphere and
Environment, 7, 11031116.
DOHRENWEND, J.C. & MORING, B.C. 1991. Reconnaissance photogeological map of
young faults in the Winnemucca 1o by 2o quadrangle, Nevada. Miscellaneous Field
Studies Map MF-2175. United States Geological Survey.
DUBLYANSKY, Y.V., SMIRNOV, S.Z. & PASHENKO, S.E. 2003. Identification of
the deep-seated component in paleo fluids circulating through a potential
nuclear waste disposal site: Yucca Mountain, Nevada, USA. Journal of
Geochemical Exploration, 78, 3943.
FONTES, J.Ch., YOUSFI, M. & ALLISON, G.B. 1986. Estimation of long-term,
diuse groundwater discharge in the northern Sahara using stable isotope
profiles in soil water. Journal of Hydrology, 86, 315327.
GOVETT, G.J.S. 1974. Soil conductivities: assessment of an electrochemical
technique. In: ELLIOT, I.L. & FLETCHER, W.K. (eds) Geochemical Exploration.
Special Publication, 2. Association of Exploration Geochemists, 101118.
GOVETT, G.J.S., DUNLOP, A.C. & ATHERDEN, P.R. 1984. Electrogeochemical
techniques in deeply weathered terrain in Australia. Journal of Geochemical
Exploration, 21, 311331.
HALL, G.E.M. 1992. Inductively coupled plasma mass spectrometry in
geoanalysis. Journal of Geochemical Exploration, 44, 201249.
HALL, G.E.M. 1998. Analytical perspectives on trace element species of
interest in exploration. Journal of Geochemical Exploration, 61, 120.
HALL, G.E.M., VAIVE, J.E. & BUTTON, P. 1997. Detection of past underground
nuclear events by geochemical signatures in soils. Journal of Geochemical
Exploration, 59, 145162.
HAMILTON, S.M. 1998. Electrochemical mass transport in overburden: a new
model to account for the formation of selective leach geochemical
anomalies in glacial terrain. Journal of Geochemical Exploration, 63, 155172.
HAMILTON, S.M. 1999. Summary of Fieldwork and Other Activities, 1999. Open File
Report, 6000. Ontario Geological Survey, Ontario, 421426.
HAMILTON, S.M. 2000. Spontaneous potentials and electrochemical cells. In:
HALE, M. (ed.) Geochemical Remote Sensing of the Subsurface. Handbook of
Exploration Geochemistry. Elsevier, Amsterdam, 6, 81119.
HAMILTON, S.M., CAMERON, E.M., MCCLENAGHAN, M.B. & HALL, G.E.M. 2001.
Deep-Penetrating Geochemistry: Cross Lake Final Report. Report. Canadian
Mining Industry Research Organization (CAMIRO) Report.

32

E.M. Cameron et al.

HAMILTON, S.M., CAMERON, E.M., MCCLENAGHAN, M.B., HALL, G.E.M.,


LEYBOURNE, M.A., SADER, J.A. & CRANSTON, D.R. 2002. Thick Overburden
Geochemical Methods: Studies over Volcanogenic Massive Sulphide
Mineralization and Kimberlite. In: Summary of Fieldwork and Other Activities,
2002. Open File Report, 6100. Ontario Geological Survey, Ontario,
2712717.
JACKSON, R.G. 2000. Mobile Ion Geochemistry Mike Epithermal Gold Deposit,
Nevada, USA. Report. Canadian Mining Industry Research Organization
(CAMIRO).
KELLEY, D.L. 1995. An investigation of piedmont geochemical exploration techniques for
buried porphyry Cu deposits. Part 1. Soil Geochemistry and Biogeochemistry. Report.
BHP Minerals.
KRISTIANSSON, K. & MALMQVIST, L. 1982. Evidence of non-diusive transport
of Rn-222 in the ground and a new physical model for the transport.
Geophysics, 47, 14441452.
LEYBOURNE, M.I. & CAMERON, E.M. 2000. Composition of groundwaters at the
Spence deposit: Part A: Major elements and stable isotopes. Report. Canadian
Mining Industry Research Organization (CAMIRO).
LEYBOURNE, M.I. & CAMERON, E.M. 2000. Composition of groundwaters at the
Spence deposit: Part BMetals associated with supergene Cu mineralization. Report.
Canadian Mining Industry Research Organization (CAMIRO).
LOVELL, J. & HALE, M. 1983. Application of soil-air carbon dioxide and
oxygen measurements as a guide to mineral exploration. Transactions Institute
of Mining and Metallurgy, 92, 2832.
LOVELL, J.S., HALE, M. & WEBB, J.S. 1983. Soil air carbon dioxide and oxygen
measurements as a guide to concealed mineralizations in semi-arid and arid
regions. Journal of Geochemical Exploration, 19, 305317.
MCCARTHY, H. & MCGUIRE, E. 1998. Soil gas studies along the Carlin Trend. Eureka
and Elko Counties, Nevada. Open File Report, 98-338. United States
Geological Survey, 243250.
MCCARTHY, J.H., LAMBE, R.N. & DIETRICH, J.A. 1986. A case study of soil
gases as an exploration guide in glaciated terrain: Crandon massive sulphide
deposit, Wisconsin. Economic Geology, 81, 408420.
MUIR-WOOD, R. 1994. Earthquakes, strain-cycling and the mobilization of
fluids. In: PARNELL, J. (ed.) Geofluids: Origin, Migration and Evolution of Fluids
in Sedimentary Basins. Special Publication, 78. Geological Society, London,
8598.
NILSON, R.H., PETERSON, E.W., LIE, K.H., BURKHARD, N.R. & HEARST, J.R.
1991. Atmospheric pumping: a mechanism causing vertical transport of
contaminated gases through fractured permeable media. Journal of
Geophysical Research, 96, 2193321948.
NORBY, J.W. & OROBONA, M.J.T. 2002. Geology and mineral systems of
the Mike deposit. In: THOMPSON, T.B., TEAL, L. & MEEUWIG, R.O. (eds)
Gold Deposits of the Carlin Trend. Nevada Bureau of Mines and Geology
Bulletin 111, 5970.
NUR, A. 1974. Matsushiro, Japan, earthquake swarm: confirmation of
dilatancy-fluid diusion model. Geology, 2, 217221.
PAUWELS, H., BAUBRON, J.C., FREYSSINET, P. & CHESNEAU, M. 1999. Sorption
of metallic compounds on activated carbon: application to exploration for
concealed deposits in southern Spain. Journal of Geochemical Exploration, 66,
115133.
PHILLIPS, F.M. 1994. Environmental tracers for water movement in desert
soils of the American southwest. Journal of Soil Science Society of America, 58,
1524.
PIOTROWICZ, S.R., DUCE, R.A., FASCHING, J.L. & WEISEL, C.P. 1979a. Bursting
bubbles and their eect on the sea to air transport of Cd, Cu, Fe, Pb and
Zn. EOS, 60, 276.
PIOTROWICZ, S.R., DUCE, R.A., FASCHING, J.L. & WEISEL, C.P. 1979b. Bursting
bubbles and their eect on the sea to air transport of Fe, Cu and Zn. Marine
Chemistry, 7, 307324.

RAHN, K. 1976. The chemical composition of the atmospheric aerosol. Technical


Report. University of Rhode Island, USA.
REMENDA, V.H., CHERRY, J.A. & EDWARDS, T.W.D. 1994. Isotopic composition of old ground water from Lake Agassiz: Implications for Late
Pleistocene climate. Science, 266, 19751978.
ROSE, A.W. & GOW, W. 1995. Thermal convection of soil air on hillsides.
Environmental Geology, 25, 258262.
ROSE, A.W., HUTTER, A.R. & WASHINGTON, J.W. 1990. Sampling variability of
radon in soil gases. Journal of Geochemical Exploration, 38, 173191.
ROWLAND, M.D. & CLARK, A.H. 2001. Temporal overlap of supergene alteration and
high-sulphidation mineralization in the Spence porphyry Cu deposit, II Region, Chile,
Abstract. Annual Meeting. Geological Society of America.
RUTHERFORD, N.F., JOHNSON, C., GIBLIN, A.M., GRIFFIN, W.L., RYAN, C.J. &
SUTER, G.F. in press. The pSirogas research project: an assessment of the
Geogas exploration method in Australian terrains. Geochemistry: Exploration,
Environment, Analysis.
SCANLON, B.R. 1991. Evaluation of moisture flux from chloride data in desert
soils. Journal of Hydrology, 128, 137156.
SCANLON, B.R. 1992. Evaluation of liquid and vapor water flow in desert soils
based on chlorine 36 and tritium tracers and nonisothermal flow
simulations. Water Resources Research, 28, 285297.
SCHWARTZ, F.W. 1974. The origin of chemical variations in groundwaters from
a small watershed in southwestern Ontario. Canadian Journal of Earth Science,
11, 893904.
SIBSON, R.H. 1981. Fluid flow accompanying faulting: Field evidence and
models. In: SIMPSON, D.W. & RICHARDS, P.G. (eds) Earthquake Predication:
An International Review. Maurice Ewing Series, 4. American Geophysical
Union, 593603.
SIBSON, R.H., MOORE, J.M. & RANKIN, A.H. 1975. Seismic pumping; a
hydrothermal fluid transport mechanism. Journal Geological Society London
Special Publication, 131, 653659.
SILLITOE, R.H., MARQUARDT, J.C., RAMIREZ, F., BECERRA, H. & GOMEZ, M.
1996. Geology of the concealed MM porphyry Cu deposit, Chuquicamata
District, northern Chile. In: CAMUS, F., SILLITOE, R.H. & PETERSEN, R. (eds)
Andean Cu Deposits: New discoveries, Mineralization, Styles and Metallogeny.
Special Publication, 5. Society of Economic Geology, 5970.
SMEE, B.W. 1983. Laboratory and field evidence in support of the electrochemically enhanced migration of ions through glaciolacustrine sediment.
Journal of Geochemical Exploration, 19, 277304.
TCHALENKO, J.S. 1973. The Kashmar (Turshiz) 1903 and Torbat-e Heidariyeh
(South) 1923 earthquakes in Central Khorassan (Iran). Annali di Geofisica,
26, 2940.
TCHALENKO, J.S. & BERBERIAN, M. 1974. The Salmas earthquake of May 6th,
1930. Annali di Geofisica, 27, 151212.
TEAL, L. & BRANHAM, A. 1997. Geology of the Mike gold-Cu deposit, Eureka
County, Nevada. In: VIKRE, P. et al. (ed.) Carlin-Type Gold Deposits Field
Conference. Guidebook Series, 28. Society Economic Geology, 257276.
THORNBER, M.R. 1975. Supergene alteration of sulphides, I. A chemical model
based on massive nickel sulphide deposits at Kambalda, Western Australia.
Chemical Geology, 15, 114.
WEEKS, E.P. 1987. Eect of topography on gas flow in unsaturated fractured
rock: concepts and observations. In: EVANS, D.D. & NICHOLSON, T.J. (eds)
Flow and Transport Through Unsaturated Fractured Rock. Geophysical
Monograph, 42. American Geophysical Institute, 165170.
XIE, X., WANG, X., XU, L., KREMENETSKY, A.A. & KHEIFETS, V.K. 1999.
Regional orientation of strategic deep penetration geochemical methods in
the central Kyzylkum desert terrain, Uzbekistan. Journal of Geochemical
Exploration, 66, 135143.

You might also like