Sei sulla pagina 1di 445

H A N D B O O K OF E X P L O R A T I O N G E O C H E M I S T R Y

G.J.SGOVETT (Editor)

1. ANALYTICAL METHODS IN GEOCHEMICAL PROSPECTING


2. STATISTICS AND DATA ANALYSIS IN GEOCHEMICAL PROSPECTING
3. ROCK GEOCHEMISTRY IN MINERAL EXPLORATION
4. REGOLITH EXPLORATION GEOCHEMISTRY IN TROPICAL AND
SUB-TROPICAL TERRAINS
5. REGOLITH EXPLORATION GEOCHEMISTRY IN ARCTIC AND
TEMPERATE TERRAINS
Handbook of Exploration Geochemistry

VOLUME 5
Regolith Exploration Geochemistry
in Arctic and Temperate Terrains

KALEVI KAURANNE
Geological Survey of Finland, SF-02150 Espoo, Finland

with

REIJO SALMINEN and KARIN ERIKSSON

Assisted by

Nils Gustavsson, Pentti Noras and Veli-Pekka Salonen

ELSEVIER
Amsterdam - London - New York - Tokyo 1992
ELSEVIER SCIENCE PUBLISHERS B.V.
Sara Burgerhartstraat25
P.O. Box 211,1000 AE Amsterdam, The Netherlands

Library of Congress C a t a l o g i n g - i n - P u b l i c a t i o n Data

R e g o l i t h e x p l o r a t i o n g e o c h e m i s t r y i n a r c t i c and t e m p e r a t e t e r r a i n s
e d i t e d by L . K . K a u r a n n e , R. S a l m i n e n , and K a r i n E r i k s s o n .
p. cm. — ( H a n d b o o k o f e x p l o r a t i o n g e o c h e m i s t r y ; v . 5 )
I n c l u d e s b i b l i o g r a p h i c a l r e f e r e n c e s and i n d e x .
ISBN 0 - 4 4 4 - 8 9 1 5 4 - 4 o i k . paper)
1. Geochemical p r o s p e c t i n g — S c a n d i n a v i a . 2. Geochemical
prospecting—Arctic regions. 3. S o i l chemistry—Scandinavia.
4. S o i l c h e m i s t r y — A r c t i c regions. 5. Geology, S t r a t i g r a p h i c -
-Quaternary. I . K a u r a n n e , L . K. I I . Salminen, Reijo.
I I I . E r i k s s o n , K a r i n , 1937- . IV. Series.
TN270.R447 1992
622'.13'0948--dc20 92-11570
CIP

ISBN 0-444-89154-4

© 1992 Elsevier Science Publishers B.V. All rights reserved

No part of this publication may be reproduced, stored in a retrieval system or transmitted in any
form or by any means, electronic, mechanical, photocopying, recording or otherwise, without the
prior written permission of the publisher, Elsevier Science Publishers B.V., Copyright and Per-
missions Department, P.O. Box 521,1000 AM Amsterdam, The Netherlands.

Special regulations for readers in the U.S.A.-This publication has been registered with the Copy-
right Clearance Center Inc. (CCC), Salem, Massachusetts. Information can be obtained from the
CCC about conditions under which photocopies of parts of this publication may be made in the
U.S.A. All other copyright questions, including photocopying outside of the U.S.A., should be
referred to the publisher.

No responsibility is assumed by the publisher for any injury and/or damage to persons or proper-
ty as a matter of products liability, negligence or otherwise, or from any use or operation of any
methods, products, instructions or ideas contained in the material herein.

Although all advertising material is expected to conform to ethical (medical) standards, inclusion
in this publication does not constitute a guarantee or endorsement of the quality or value of such
product or of the claims made of it by its manufacturer.

This book is printed on acid-free paper.

Printed in The Netherlands


V

EDITOR'S FOREWORD

Regolith Exploration Geochemistry in Arctic and Temperate Terrains is the


fifth volume in the Handbook of Exploration series which, together with
the volume Regolith Exploration Geochemistry in Tropical and Sub-tropical
Terrains covers the use of overburden in exploration geochemistry. This
book, as with other volumes in the series, is designed to be of practical
assistance to the field geologist as well as providing a comprehensive review
of the subject that will be a reference source for research workers.
It is axiomatic that those best qualified to produce a work such as this
book are also too busy to undertake it. It was, therefore, with considerable
trepidation that I asked Professor L.K. Kauranne to be senior editor of the
book. I was aware that Professor Kauranne was Director of what I believe
to be the largest Geological Survey in northern Europe; I was also aware
that he had founded and nurtured one of the world's foremost exploration
geochemistry departments at the Geological Survey of Finland. It was the
latter experience I wished to capture for the Handbook series.
Whereas I was successful, Professor Kauranne was rather less lucky
in obtaining support from other busy people, e.g. in North America. My
original concept of this volume was that it would deal with till and soil
geochemistry in the glaciated and temperate zones of the northern hemi-
sphere, but because of the lack of input from North American geochemists,
this book is confined to Europe — and dominantly Fennoscandia.
During the time I worked as a geochemist in glaciated terrain I largely
avoided the problems of till by concentrating on rock geochemistry. This
book is an elegant example of the old adage that necessity is the mother of
invention — the blanket of till that smothers the land surface is regarded by
the Fennoscandinavians as a rare gift in providing a homogenized average
sample of the underlying bedrock.
As in all exploration geochemistry, the exquisitely simple principles
prove agonizingly difficult to implement in practice. The key to successful
applications of exploration geochemistry to surficial materials in glaciated
terrain is to thoroughly understand the processes that gave rise to the
overburden. For this reason, the first five chapters of this volume are largely
devoted to Quaternary geology. The importance of these chapters cannot be
VI Editor's foreword

over-emphasized; the history of exploration geochemistry is littered with


examples of "failures" that, with hindsight, are clearly attributable to an
inadequate understanding of surface processes.
In this, as in the other volumes in the Handbook series, adequate space
has been given to allow ideas and procedures to be explained in sufficient
depth to permit the practising exploration geologist to understand —
and apply — the principles. The practical aspects are emphasized by the
inclusion of chapters on field methods, case histories, analytical techniques,
and data interpretation.
This volume specifically addresses the use of till in exploration geochem-
istry, but many of the principles and techniques can be applied to other
sample media. The imaginative data display techniques should prove to be
especially useful to exploration geochemists and geologists in all geological
and geochemical environments.

G.J.S. GOVETT
Helsinki, Finland and Sydney, Australia
VII

ACKNOWLEDGEMENTS

We are indebted to Professor Gerry J.S. Govett who in asking us to write


this part of the Handbook of Exploration Geochemistry not only expressed
his trust in us but also supported us with his own enormous enthusiasm.
The secretaries Sirkka-Liisa Ollikainen and Asta Sainio retyped all the
problem manuscripts, some of them several times; geologists Sinikka Roos
and especially Erna Kuusisto painstakingly compiled the texts, compiled
the indexes and assisted in innumerable other ways. Heli Moberg with
the assistance of Soili Ahava and Pirkko Kurki drafted the hundreds of
figures. We would like to express our sincere gratitude to all of them.
Kathleen Ahonen deserves a special word of thanks for battling with our
often mysterious use of English, improving both the fluency and logical
flow of the texts. If in the end something remains amiss in either the texts
or illustrations, please do not blame the wonderful ladies; the fault lies
entirely with the writers.

KALEVI KAURANNE
IX

THE AUTHORS

Mr. Nils Gustavsson, mathematician, Geological Survey of Finland, Beton-


imiehenkuja 4, 02150 Espoo, Finland
Dr. Karin Eriksson, geologist, Ängsvägen 6, 77600 Hedemora, Sweden,
formerly geochemist of Boliden Co.
Dr. Kalevi Kauranne, professor, Satukuja 1 F 35, 02230 Espoo, Finland,
formerly director of Geological Survey of Finland
Mr. Pentti Noras, chemist, Geological Survey of Finland, Betonimiehenkuja
4, 02150 Espoo, Finland
Dr. Reijo Salminen, professor, Head of Geochemistry Department, Geologi-
cal Survey of Finland, Betonimiehenkuja 4, 02150 Espoo, Finland
Dr. Veli-Pekka Salonen, act. professor, University of Turku, Institute of
Geology, 20500 Turku, Finland

Authors of the case histories prepared for this book

Mr. Erkki Ilvonen, geochemist, Outokumpu Finnmines Co., Lapin Malmi,


Kivikatu 6, 96400 Rovaniemi, Finland
Mr. Terho Koivisto, free contractor, Smedsby, 02400 Kirkkonummi, Finland,
formerly geochemist with the Geological Survey of Finland
Mr. Jouko Kokko, inspector, Social Board of Finland, Siltasaarenkatu 18,
00530 Helsinki, Finland, formerly geologist with the Geological Survey
of Finland
Mr. Martti Kokkola, geochemist, Outokumpu Finnmines Co., Exploration
Vammala mine, 38200 Vammala, Finland
Mr. Esko Kontas, chemist, Geological Survey of Finland, Lähteentie 2,
96100 Rovaniemi, Finland
Dr. Pasi Lehmuspelto, geochemist, Geological Survey of Finland, Lähteentie
2, 96100 Rovaniemi, Finland
Dr. Pekka Lestinen, geochemist, Geological Survey of Finland, Neu-
laniementie 5, 70700 Kuopio, Finland
X Editor's foreword

Dr. Maria Nikkarinen, geochemist, Geological Survey of Finland, Neu-


laniementie 5, 70700 Kuopio, Finland
Dr. Vesa Peuraniemi, lecturer, University of Oulu, Linnanmaa, 90570 Oulu,
Finland
Dr. Matti Äyräs, geochemist, Geological Survey of Finland, Lähteentie 2,
96100 Rovaniemi, Finland
1

Chapter 1

INTRODUCTION

SCOPE AND CONTENTS OF THIS VOLUME

The purpose of this volume of the Handbook of Exploration Geochemistry


is to assist the geologist in the use of overburden materials in geochemical
exploration. The properties of the materials to be assessed, sampling,
analysis and the interpretation and presentation of results are all described.
Geographically the present volume is limited to arctic and temperate
regions. A companion volume covering the soil geochemistry of tropical and
subtropical regions will be appearing about the same time.
Together these volumes describe the geochemical analysis of the Earth's
unconsolidated overburden, both the in situ residuum of weathering and
the transported mineral drift. Other volumes of the Handbook are devoted
to the use of rocks, stream and lake sediments, organic materials, water and
air in prospecting (Fig. 1-1). As chemical analysis and data interpretation
are also treated in separate volumes of this series, they are touched upon
here only briefly.
Chapters 2, 3 and 4 describe the characteristic materials of the over-
burden in arctic and temperate regions, their origin, structure and geo-
chemical character. A general description of the weathered bedrock, glacial
till, glacifluvial formations and more recent alluvial sediments and organic
deposits is presented, together with the types and properties of soil. Re-
gional and local scale geochemical dispersion of elements, and the details
of different transport mechanisms, are discussed in Chapters 5 and 6;
understanding of the dispersion is the key to the interpretation of results.
Chapters 7 and 8 provide a short introduction to field work — to the
different sampling, analytical and data processing methods appropriate for
the different scales of study and to the equipment and measurements that
need to be made in the field. Chapters 9 about chemical analysis and 10
about statistical treatment are included to assist the reader acquainted
with these subjects and partly to help him/her in the necessary discussions
with chemists and mathematicians.
The case histories presented in Chapter 11 have been chosen to illustrate
a wide variety of situations: topographies from peneplain to mountains,
2 Introduction

"V**

TILL ON SLOPES

TILLY HILLS
SHALLOW BOGGY TERRAIN = JÄNKA = MUSKEG

Fig. 1-1. Different types of overburden formations and materials in arctic and temperate
glaciated terrain, (photo Peter Johansson)
History of soil geochemical prospecting 3

climate from arctic to temperate, and as many types of overburden as


possible. It is to be hoped that from among the many case examples the
reader will find at least one closely resembling his or her own research
object, and find instruction in how others have proceeded and succeeded,
or failed. Besides explaining their particular studies, the authors tell what
else should be done and what conclusions can be drawn.
The final Chapter 12 "Focal aspects of soil geochemistry applied in arctic
and temperate regions" is not a summary but some kind of overview of
the state of art and it is hoped that recommendations given in it will help
geologist in the geochemical stage of her/his usually urgent prospecting
task.

HISTORY OF SOIL GEOCHEMICAL PROSPECTING

The use of metals, their exploration and the beginning of geochemistry


go back a long way Five thousand years ago, village blacksmiths in India,
Iran, Mesopotamia and Greece were acquainted with copper and were able
to smelt it with tin to make bronze. Iron was introduced to Greece about
3100 BP and its use had spread to northern Europe by about 2500 BP.
How did our forefathers find their ores before the days of sampling
nets and sophisticated analytical instruments? Ancient prospectors were
practising a form of geochemical exploration of soil when they searched for
copper stain and iron rust. An old tale from Sweden relates how a farmer
in Falun was led to an iron ore by a goat that came home with its hoofs
coloured red. In the Finnish national epic Kalevala it is observed "the birth
of iron is of rust".
Georgius Agricola colourfully describes the use of "geochemical" methods
of prospecting in his famous book De Re Metallica (1556):
Now I will discuss that kind of minerals for which it is not necessary to dig, because the
force of water carries them out of veins. Of these there are two kinds, minerals — and their
fragments — and juices. When there are springs at the outcrop of veins from which, as
I have already said, the above-mentioned products are emitted, the miner should consider
these first, to see whether there are metals or gems mixed with the sand, or whether the
waters discharged are filled with juices (quoted by Boyle, 1967).

Jakob Forsskäl, Finnish state prospector in the early eighteenth century


advises in his book Om malmers kännande och efterletande. Underrättelse
for allmogen och gemene man (1736), how overburden, water and air should
all be taken note of:
Overburden itself can tell about the ores below or in the vicinity. The weight and colour of
such a soil, sand or clay, when carefully studied, usually tell much: green soil or drift points
to copper ore, black, red and brown sand or soil to iron, varyingly yellow, white or bluish
sand or clay to silver or lead.
4 Introduction

He also notes that water which in confined places takes on colour and
taste from the surrounding soil "reflects the metalliferous rock through
which it is running by its nasty taste and mean flavour," and as the final
test "kills the animals drinking it".
Daniel Tilas, another Finnish state geologist, reported in 1743 that ore
floats are situated to the southeast side of ore suboutcrops in Finland, so
hinting for the first time at the method of boulder tracing. This method
of exploration was taken to North America as early as 1747-1751 by Pehr
Kalm during his botanical excursions (Goldthwait, 1982). A century later
the direct correlation between the chemistry of the overburden and the
bedrock below was taken careful note of by Hjalmar Lundbohm (1887)
in his studies of Swedish marble formations and the Ca content of the
overlying soil.
It was not hard to guess at the origin of ore boulders in talus up on the
valley side. As early as 1802 Playfair deduced the activity of valley glaciers
in transporting boulders along the valley, and in 1832 Bernhardi traced the
source of the big porphyrite and granite erratics in northern Germany back
to Scandinavia (Goldthwait, 1982). Nils Nordenskiöld published his map
of glacial striae in Finland in 1863. Moving into our own century, boulder
tracing for glaciological (Helmersen, 1882; Hedström, 1894; Hausen, 1912)
and later for explorational purposes became popular and important not
only in Finland but in all northern countries including Canada and the
United States (Sauramo, 1924; Högbom, 1931; Lundqvist, 1948; Flint,
1947; Holmes, 1952; Grip, 1953; Aurola, 1955).
Minerals were sought and mined from sediments long before Agricola
— for example, by panning of gold and precious stones. Since the gold
rushes of last century, the heavy minerals in sediments have been studied
worldwide both for prospecting purposes and "mapping" of the underlaying
bedrock (e.g., Raeburn and Millner, 1927; Kivekäs, 1946; Mertie, 1954;
Theobald, 1957; Theobald and Thompson, 1959; Lee, 1971).
Mineralogical analysis is based on relatively few identified grains that
hopefully are typical. If thousands were counted the process would be too
time consuming and tedious. A sample of finer fraction analyzed chemically
gives more reliable results. The first attempts to use soil chemistry
specifically for prospecting were made in the Soviet Union in the 1920s by
Vernadskij, Vinogradov and Fersman; the method was called metallometry,
and according to the metal sought, cuprometry, ferrometry etc. (Sergeev
and Solovov, 1937).
With little delay the chemical method was applied in Norway (Gold-
schmidt, 1934; Vogt, 1939), Finland (Rankama, 1940; Kauranne, 1951),
Canada (Chisholm, 1950) and the United States (Hawkes and Lakin, 1949;
Huff, 1951). Rapidly pedogeochemistry, as it was called in the West, spread
throughout the world, being successfully applied in a variety of geological
environments.
General principles of element distribution 5

As further examples of earlier case histories in pedogeochemical prospect-


ing, from warmer to cold climate, the following may be mentioned: Hawkes
(1952, 1954); Holman and Webb (1957); Webb (1958); Govett (1960);
Armour-Brown and Nichol (1970) — Sergeev (1941); Fulton (1950); Huff
(1952); Bloom (1955) — White and Allen (1954); Warren and Delavault
(1956); Ermengen (1957); Boyle and Cragg (1957); Kauranne (1958, 1959);
Dreimanis (1960) — Pitulko (1968); Garrett (1971); Gleeson and Cormier
(1971); Shilts (1971); Cameron (1977); DiLabio (1981).
The new method proved successful in tropical and nonglaciated tem-
perate areas, whereas in glaciated terrain there were both successes and
failures. The haphazard sampling of glacial overburden produced results of
no value to exploration and the financing of pedogeochemical studies ran
into difficulties in Soviet Karelia (A.V Sidorenko, pers. commun., 1978) and
Scandinavia. One of the first attempts to apply the methods in Finland was
in connection with the search for the source of the Vihanti zinc-bearing
boulders, but no obvious anomalies were found (O. Joensuu, unpubl. report,
1947). Later on, the till at Vihanti was shown to be highly complex in struc-
ture and material (L.-M. Kauranne, 1979). Failures, of course, are seldom
reported in the literature, but plenty of the successful cases of geochemical
exploration are described, for example Harbaugh (1953); Ginzburg (1960);
Hawkes and Webb (1962); Kvalheim (1967); Levinson (1974); Nichol and
Björklund (1973); Siegel (1974); Beus and Grigorian (1977). Other examples
are given in the many compilations of international symposia arranged by
the Association of Exploration Geochemists and the Institute of Mining
and Metallurgy (see works edited by Bradshaw, 1975; Kauranne, 1976a;
Björklund, 1984).
Traditions in science and different geological conditions have combined
to produce a number of "schools" of geochemistry, e.g., the "French-Soviet"
type of geochemistry, centring around the International Association of
Geochemistry and Cosmochemistry and including e.g., A.E Solovov, A.I.
Perelman, VV Polikarpochkin, J. Barbier and E. Wilhelm, and the "Anglo-
Saxon" type of geochemistry, represented by J. Webb, H. Warren, H.E.
Hawkes, R.W Boyle. A.W Rose, E. Cameron, I. Nichol and G.J.S. Govett.
Then there is the "glacigeological" group of geochemists A. Dreimanis, W
Shilts, W Coker, R. DiLabio and most Fennoscandian researchers. Each
group has had its own interests and methods of study and has developed
its own terminology. For example, lithogeochemistry in the Soviet school
covers all materials consisting of mineral grains, both nonconsolidated
overburden and tough rock, while lithogeochemistry in the Anglo-Saxon
school is equivalent to rock geochemistry. The joint symposia organized by
the various groups have helped to construct a common "language" without
which no comparison of achievements is possible.
6 Introduction

GENERAL PRINCIPLES OF ELEMENT DISTRIBUTION

The variations in the concentration of elements in the regolith depend


on the primary and secondary differentiation processes. The metal content
of an economic deposit is typically ten to a thousand times the average
content (clarke) of the same element in nonmineralized rocks. Almost every
mineralization in the bedrock is surrounded by a zone of anomalously
high concentrations of ore elements. Sometimes such a halo or aureole is
asymmetrical, with light elements predominating in the upper part and
heavier elements in the lower part of the halo (e.g., Solovov, 1959). A halo
is also found in all the secondary materials in the vicinity (e.g., Nurmi
and Isohanni, 1984), but the ore elements are usually present in different
concentrations and proportions in each material, and the concentrations
are different to those in the bedrock.
Every rock type is characterized by a specific chemical composition — a
geochemical spectrum specific as to contents of elements, their variation
and proportions. This spectrum can be used as a "fingerprint" in the
localizing and identifying of rocks by geochemical mapping. There are
also regional differences in concentrations of trace elements in bedrock
lithologies. The reflection of bedrock composition into the overburden
varies widely from area to area. Some areas are characterized by higher
than average concentrations of certain elements in the overburden. These
zones, which may or may not coincide with the lithological or stratigraphical
zones of the bedrock, show the syngenetic metallogenic provinces of the
bedrock.
In the first stage of geochemical exploration, large scale sparse sampling
is essential for determining the background levels and detecting and
localizing anomalies. Analysis needs to be made for a wide selection of
elements in order to reveal the regional variation. Simultaneously the type
and direction of transport of the material and the mode of occurrences of
the ore elements should be studied to allow interpretation of the anomalies.
Reliable mineralogical and sedimentological methods are available for
transport studies. Rapid and cheap semi-automatic and automatic methods
of chemical analysis allow a sensitive and accurate total element or partial
leach determination of a variety of elements; computer facilities exist for the
manipulation and storage of large amounts of data and their presentation
in informative ways. Nevertheless there are still situations for which tools
are not available and where the very low concentrations of some elements
or their intractable mineralogical character set limitations on geochemical
exploration.
No equipment is adequate, on the other hand, if the exploration geo-
chemist does not possess sufficient creative imagination. No element ven-
tures out in isolation, but each has companions in the deposit; typical
paragenetic associations of elements exist for each ore. If the other ele-
General principles of element distribution 7

ments are easier to analyze or are present in greater concentration than


the element of particular interest, then they may be used as pathfinders.
The resolution may be improved mathematically by adding the contents of
the paragenetic elements together, or multiplying the contents and then
dividing the product by the product for elements that occur in negative
correlation with them. Procedures like these are used in discrimination
analysis.
Intractable elements may of course be explored by other semi-geochemical
methods. For example, gold, which occurs in extremely low concentrations,
is easily enriched by panning and identified in the heavy mineral con-
centrate by its intense yellow colour. From a bucket of soil, gold nuggets
weighing less than 0.1 mg can be separated by panning and counted, which
is equivalent to a sensitivity of 10 ppb for the analysis (Nichol, 1986).
Radioactive minerals are easily found by gamma spectrometry (Osterlund,
1982). Similarly, the tungsten minerals are identified by their fluorescence
in ultraviolet light. Some minerals generate gases when weathering, and
dogs can be trained to detect them. Some minerals are easily separated on
the basis of their different grain form (Halonen, 1967).
The one requirement for exploration geochemistry is a clear contrast in
some metal content between the mineralization and the country rock. If
that is the case in bedrock, then a clear anomaly may exist in secondary
formations as well. Mostly the search is for higher than average concentra-
tions (positive anomalies). But lower than average concentrations (negative
anomalies) are sometimes used too — in the localization of granite massives
for example. The resolution of the methods for finding negative anomalies
may not be as good as for delineating positive anomalies.
In geochemical mapping aimed at a general study of the bedrock a
determination of the total contents of elements is preferable. But in
prospecting for mineralization the greatest contrast is obtained if only
the metal contents of the particular minerals forming the ore body are
determined. When searching for a sulphide mineralization, for example,
leaching of the sample material with weak mineral acids is more effective
than total solution, because with acids sulphides go into solution, whereas
oxides and silicates mainly do not.
In overburden, sulphides may be partly weathered and their heavy
metals absorbed into oxyhydrate colloids. Such metals go into solution even
upon cold extraction with organic acids or their salt solutions. In drift, the
weakest minerals, such as sulphides, are ground to the finest grain-size
fractions, where also the free chemically attackable mineral surface is
largest. Analysis of these fractions thus gives the greatest concentrations
as well as variations and therefore the best results in a search for deposits
of such minerals.
The most suitable secondary material for geochemical study is one that
was transported mechanically and only once, or one where the ions of the
8 Introduction

mineralization were transported chemically but in only one direction. The


material chosen for assay should cover the main part of the study area
and be uniformly distributed. Seldom is the case that simple, however.
Typically the occurrence of material is uneven and patchy. Moreover,
multiple transports, superimposed clastic and chemical anomalies and
different modes of occurrence of the elements complicate the interpretation
as well as tracing of the anomaly back to the source. Sequential leaching,
multi-elemental analysis of different fractions and microscopical or electron
microscopical studies of the material may be needed.
The extent and intensity of the anomalies of different elements reflect
the size of the outcrop and the contrast between it and the country rock.
But they tell little about the volume of the orebody or its grade. Only after
the deposits of a region have been mined out it is possible to calculate ratios
between the anomalies in the secondary materials and the dimensions of
the orebodies in bedrock and obtain some rough regional constants. Such
constants have been used in predictions of ore potential in the Soviet Union
(Solovov, 1959; Rundqvist, 1979) and elsewhere.
If no geological or geophysical maps are available for the region, the
geochemical research should be started with low-density sampling over a
larger area. From each sampling point a representative sample of different
materials should be taken and analyzed for as many elements as possible.
Stream and lake sediments, humus and peat, stream moss, water and
glacial till have been shown to be suitable in the internordic Nordkalott
project (Bölviken et al., 1986).
In the anomalous areas identified in the first stage, one or two materials
can then be sampled more densely and the samples analyzed only for the
more promising elements. This second stage is like a pilot prospecting
phase and should include the necessary tests for finding the best material
and method for analysis. It is followed up by a detailed study of the most
informative type of overburden material at different depths, preferably also
of the surface of the bedrock, and the analysis of the optimal grain size
fraction for only a few selected elements.
If geological or geophysical maps already exist for a region, the study may
begin from the second (pilot) stage and be concentrated on the geologically
or geophysically most promising subareas. With geochemical methods
it is also possible to discriminate the ore critical geophysical or geological
anomalies from the less promising ones. However, since geological maps and
especially geophysical maps give only indirect indication of mineralization,
it is always worthwhile to begin the geochemical mapping with regional
low-density sampling.
Overburden of arctic and temperate regions 9

OVERBURDEN OF ARCTIC AND TEMPERATE REGIONS

The northernmost parts of the Earth are still covered by ice and
surrounded by a zone of permafrost. Similar conditions are met with in the
southern hemisphere. A much larger part of the Earth was frozen and even
glaciated several times during the Pleistocene. The preglacial weathered
crust was abraded and is today found in situ only in isolated places such as
in areas of ancient ice divides and in bedrock depressions. Upon this regolith
were deposited vast amounts of glacial till and glacifluvial sediments, now
partially covered by postglacial mineral and organic formations similar
to those forming the overburden of the nonglaciated temperate zone.
The different materials of the overburden as a target for geochemical
exploration are shown in Fig. 1-2.
The distribution of elements in a material — its geochemical character
— depends on the origin and geological evolution of the material and is
specific for just that material. Without a sound understanding of this fact it
is impossible to interpret correctly the results of geochemical studies based
on the material. It cannot be emphasized too strongly that mixing of results
obtained from the assay of different materials leads to fatally erroneous
conclusions.
The primary dispersion of elements occurred during the magmatic dif-
ferentiation of the Earth, when major, minor and trace elements were

Λ STREAM & LAKESEDIMENT -

Fig. 1-2. The different materials of the overburden as a target for geochemical exploration.
10 Introduction

concentrated into their characteristic minerals and rocks. Regional and


contact metamorphism later altered the original concentrations and inter-
relations of the elements in the rock, and new minerals were formed.
But much more radical changes occurred and continue to occur in
weathering and possible transport and sorting and sedimentation of the
materials.
Minerals crystallized at high temperature and pressure are usually
not stable at prevailing temperatures and pressures or moisture levels,
and begin to disintegrate. Some of the ions are dissolved in water and
transported away by gravity or capillarity, leaving the residual material
depleted in these elements and enriched in those not leached. More stable
minerals may be formed, like kaolin from feldspar, with new chemical
and physical properties. Typically a tough rock is transformed through
weathering to a loose mass easily eroded by running water and flowing ice.
In the surficial part of the bedrock and the in situ regolith, weathering
leads to chemical zonality not unlike the epigenetic wall rock alteration lay-
ers found close to hydrothermal ore deposits. Elements tend to concentrate
in specific zones, but the same element may appear in different forms in
different zones, e.g., copper as sulphide, native copper, oxide or carbonate
depending on the Eh/pH conditions. The contents and internal relations
of elements in these zones naturally deviate from those of the original
rock.
The rock and mineral particles are comminuted when transported by
ice, water or wind. Mixed to form a diamicton, they are simultaneously
separated, according to their durability, into different grain-size fractions.
Fragile minerals are quickly ground and concentrated into the finest
fraction; the tough minerals are resistant and therefore transported farther
than others. In sorted sediments, particles are classified according to weight
into layers, so that minerals of the same grain size are separated and the
same element may appear in different mineral form in different layers. For
example, in the light mineral layer iron occurs mainly in mica but in the
heavy mineral layer in magnetite or garnet.
No density sorting has occurred in glacial till, which is a mixture of all
grain sizes, densities and lithologies. Some separation occurs in connection
with grinding. The concentrations and proportions of elements vary much
less in the surface till than they do in the bottom layers and especially in
different lithologies of the bedrock below. The vertical transition in element
concentrations in a till bed is gradual, whereas in sorted sediments the
geochemistry tends to vary sharply from layer to layer. It has nevertheless
been observed that the regional mean concentrations of elements in till,
sorted sediments and bedrock are nearly the same (Kauranne et al.,
1977).
Geochemical anomalies in the overburden 11

SOILS OF ARCTIC AND TEMPERATE REGIONS

The uppermost "ploughable" part of the overburden, rich in organic


material and suitable for plant growth, is called soil. Rainwater, oxygen and
organic acids play an essential role in the weathering process that leads
to soil formation. Humus enhances the leaching power of water and binds
many heavy metals as well. Of the many different soil types in the world,
podsol is the most common in temperate and arctic regions and chernosem
(on calcareous ground, rendsina) the next most common.
As far as possible, and in good-humoured defiance of the title chosen
for us by our editor, in this volume we generally reserve the term soil
for the thin, altered surface layer of the overburden. In our opinion it is
confusing to refer to the whole bed of unconsolidated overburden as soil,
particularly in glaciated regions where there is a sharp boundary between
the solid bedrock and the almost unweathered drift. The term is more
justified in regions with in situ regolith where the transition from bedrock
to overburden and to soil is much more gradual.

GEOCHEMICAL ANOMALIES IN THE OVERBURDEN

Grains of ore minerals and anomalously high concentrations of ore


metals will usually be present in the overburden above or near by ore
deposits. The true size and intensity of these anomalies will depend on the
type of overburden — whether it consists primarily of in situ residuum or
of transported nonsorted or sorted material — and on its thickness.
Elements leached and transported by water may be adsorbed or pre-
cipitated in sediments to form hydromorphic anomalies. Electrochemical
diffusion induced by Eh differences, i.e., the electrostatic potential round a
suboutcropping conductive ore body, increases the heavy metal concentra-
tions especially of the humus layer, as does the capillary rise of water or the
suction activity of tree roots.
The geochemical anomalies in till and sorted sediments in arctic and
temperate regions have originally been formed by clastic transport of
ore mineral grains. The transport direction and distance of fine materials
resembles closely that of boulders why a chapter about it is included. Deeper
down in the nonoxidating ground-water regime the ore minerals have been
preserved through the ten thousand years of the Quaternary almost
intact. Minerals higher up in the overburden will partly have oxidized and
the metals gone into water solution. In the upper levels, therefore, the
anomalies of mechanical transport and chemical transport are intermixed
and sometimes difficult to interpret; the element associations ("spectrum")
may differ strikingly from those of the mother mineralization. Thus,
without comprehensive study of the mode of occurrence of the ore metals
12 Introduction

in the anomaly, follow-up exploration may be tedious and the identification


of the source difficult. The modes of occurrence can be studied either
microscopically or chemically, the identification of source by isotopic, trace
element or fluid inclusion analysis of certain minerals — chapters of
chemical analysis and data processing may help in planning of these
studies. Happily, not all cases are so complex.
13

Chapter 2

GLACIGENIC DEPOSITS

GLACIATIONS AND INTERGLACIALS

The Pleistocene epoch, which began 2-3 million years ago, was dominated
by glaciations. Large parts of the northern and southern hemispheres, as
well as the highest mountain ranges were entirely covered by ice (Fig. 2-1)
(Flint, 1967; Lundqvist, 1980; Ehlers, 1982; Rogerson, 1982; Michelson et
a l , 1983).

Fig. 2-1. Areas of the northern hemisphere covered by the Pleistocene ice sheets.
MAXIMUM EXTENT OF THE PLEISTOCENE GLACJATIONS,
AREAS INUNDATED BY SEA OR LAKES AND 25>

THE STOP \H MELTING AT SALPAUSSELKÄ

8
500 km 3

ft*
•8
o
CO

Fig. 2-2. Areas in Europe covered by the Pleistocene glaciations and by water during deglaciation.
Glaciations and interglacials 15

Northern Europe was glaciated and deglaciated at least three times,


and a number of formations have been found to suggest still other, older
glaciations. Beginning with the oldest, the glacials were Elster, Saale and
Weichsel. In the Alps and in North America four different glacials have been
distinguished: Günz, Mindel, Riss and Wurm in the Alps and Nebrascan,
Kansan, Illinoian and Wisconsin in North America. Each of the glaciers
abraded the surface of older deposits and covered them with new deposits,
especially in the melting phase. Most important for the geochemist are the
formations of the youngest glacials: Weichsel, Wurm and Wisconsin (Figs.
2-2 and 2-3).
Warmer climatic periods intervened giving rise to ice-free interglacials
between the glaciations; the climate gradually cooling as the new glacial
approached. But the climate fluctuated during the glacials as well. During
the Weichsel glaciation, for example, the ice in Fennoscandia melted almost
totally away for a shorter interstadial period (Korpela, 1969) (Fig. 2-4).
Thus in Fennoscandia it is the final stage of the Weichsel glaciation that is
responsible for the most visible signs of the whole glaciation: most of the
striae on roches moutonnees, and most of the till and the sorted sediments
date from this stage. Older marks are partially obscured by youngest striae
and materials from older stages are mixed in and hidden under these
formations.
The final retreat of the Weichsel glacier in Europe took some 4000
years. By about 8500 years B.P., most of the ice had melted, leaving only
Greenland, the Arctic and the highest mountains still covered. Extensive
areas, which during the long glacial period had been weighted down by ice,
were left under water and therefore covered by postglacial sediments. With
later land uplift the sediments emerged from the water, only to be partly
eroded and redeposited by the waves on ancient beaches.
The dominant water body in North Europe is the Baltic Sea. During the
last 10,000 years it has passed through several brackish and fresh water
stages, each producing clay formations of characteristic geochemistry, grain
size and texture (Fig. 2-5). A comprehensive description of the geology
of the Gulf of Finland, its glacial and postglacial sediments underlain by
Precambrian crystalline rocks of the Fennoscandian Shield and bordered
by Phanerozoic sedimentary rocks on the Estonian coast (East European
plateau) was recently published by Raukas and Hyvärinen (1992).
The glacial history as well as the overburden of North America closely
resemble those of northern Europe. The dominant water body of the
Canadian Shield is Hudson Bay, which like the Great Lakes covered a
much larger area during the melting of the glacier (Fig. 2-3). Glaciations,
interglacials and corresponding formations in Europe and North America
are shown in Fig. 2-4.
The glacial ice did not cover as coherent an area in northern Asia as
in Europe or North America; rather the glaciers appeared as separate
16 Glacigenic deposits

Fig. 2-3. Maximum extent of Pleistocene glaciations in North America and areas of upheaval.

bodies. Likewise in the Himalayas, Alps, Cordilleras, Andes, and even the
Apennines the valleys were filled with glaciers while the highest tops were
left barren.
Glacial processes 17

PLEISTOCENE GLACIATIONS
CANADA FENNOSCANDIA EUROPE
(Shilts - 8 4 ) (Sibrava - 8 6 )
YEARS BP
PEAT and HUMUS BEDS
CENE
I HOLO-

MARINE ^GRAVED/SEDIMENT S
9000 TYRREL SEA
VARVES, GLACIAL LAKES
AGASSIZ etc.
WISCONSINAN

^/CLAY
^ ^ - ^ ^ ^ TILL WEICHSEL III Z>
KIPLING TILL ^"""^^ (VALDAI)^>
VARVED CLAY
FRIDAY CREEK (^GÖTAÄLV

TILL (?) WEICHSEL II ^)


S~*~~ PERÄ-
76000 FAWN RIVER GRAVEL V^_^ POHJOLA
WEICHSEL I ^ )
ADAM TILL
VARVES with PEAT and
SANGA- I
MONIAN

WOOD LEVEÄ- EEM


PEAT and FOREST BEDS NIEMI S0KLI
(MIKULINO)
v. ig
MARINE V ^ ^ ^ ^ S E D I M E N T S
BELL SEA
1 3 0 00C WARTHE
I

DARK GREY TILL (MOSCOW)


S*~ TREENE
SAND and GRAVEL ^ (ODINTSOVO)
ILLINOIAN AND OLDER

SAALE
DARK GREY TILL (DNIEPR)
f "DÖMNITZ
SAND, GRAVEL, CLASTS
V (ROMNY)
DISCONTINUOUS VARVES FUHNE
(PRONYA)
DARK RED TILL S ^ HOLSTEIN
( (LIKHVIN)
LIGHT GREY TILL
ELSTER I&II
(OKA)

ζ^~ CROMER
^ ^ _ _ i g

HELME 1,11,III
(ILOVAI)

/ ^ ARTERN;
^^___ig
MENAPIAN

Fig. 2-4. Glaciations, interglacials and the corresponding formations in North America and
Europe. Compiled from Shilts (1984a) and Sibrava (1986).
18 Glacigenic deposits

Baltic Sea
clay

Litorina Sea
muddy clay

Ancylus Lake
stratified clay

Yoldia Sea
clayey silt

Baltic Ice Lake


laminar silt

IVBED*OCI

Fig. 2-5. Late-glacial and postglacial formations at the bottom of the northern Baltic Sea.

GLACIAL PROCESSES

Erosion

Glacial action has profoundly influenced the landscape of glaciated


regions, first by abrading and polishing the hills and then by filling the
valleys. Large U-shaped valleys are characteristic of glaciated mountain
terrain, while streamlined tilly hills — drumlins with their rock cores,
smoothed stoss sides and plucked tails — lend a striped appearance to
glaciated lowlands. The smooth-surfaced rock exposures known as roches
moutonnees are grooved and striated in the direction of ice movement (Fig.
2-6) and even the polished surfaces of quartz grains are engraved with
microscopic striae.
The basal ice temperature, increasing with the velocity of the flow and
thickness of the ice, controls the processes occurring in the sole of the
Glacial processes 19

Fig. 2-6. Roches moutonnees. If the rock is fine-grained and resistant to weathering, grooves,
striae, facets, trains of crescentic marks and traces of scratching and polishing will be visible
on the surface (photo Kalevi Kauranne).

glacier. The critical temperature is the pressure melting point. In a "cold-


base" glacier, rocks and overburden materials are abraded and the debris
is incorporated into the moving glacier. In a "warm-base" glacier, material
is deposited at the rock-ice interface. As the glacier grows, a critical load is
eventually achieved and the viscosity decreases, leading to the lodgement
of till.

Transport
Two types of ice flow have been distinguished: compressive and extend-
ing. In compressive flow the slip surfaces favour upward movement of
material, in extending flow downward movement. The former leads to long-
distance transport, the latter allows mixing of the supraglacial material
into local lodgement till.
Debris may be transported beneath or in the glacier, on the surface or in
front of it. The comminution, mixing and orienting of the material mainly
depend on where and how it is transported. The mode of transport is im-
printed into the lithology of different fractions, into grain size distribution,
into orientation of grains and into the form and texture of the formation.
20 Glacigenic deposits

Likewise mode of transport affects the geochemistry of till and cannot be


neglected in the interpretation of geochemical anomalies.
Glaciers not only grow and flow but waste away by melting and sub-
limation. At the beginning of a glaciation the ice balance is positive, in
the deglaciation phase it is negative. In winter glaciers accumulate, but
in summer the volume of the ice mass diminishes as vast amounts of
meltwater are released, carrying along the stony material incorporated in
the ice, sorting it and sedimenting it into tunnels and crevasses of the ice
or outside the glacier. The sediments sorted by meltwater form glacifluvial,
glacilacustrine or glacimarine formations depending on the mode of sedi-
mentation, in other words the velocity and salt content of the water at the
place of deposition. Different types of glacial sediments are shown in Figs.
2-7 and 2-8.

BASAL BASAL WATERLAIN GLACIMARINE


DEBRIS TILL TILL SEDIMENTS

Fig. 2-7. Glacier margin and glacigenic formations during the melting off.

Fig. 2-8. Ice marginal and glacifluvial deposits formed during deglaciation: till, esker gravel
and sand, beach sand, laminated silt and clay, annual moraines. Redrawn with permission from
Magnusson et al., 1949.
Glacial tills 21

Deposition

Deposition of glacial debris gives rise, as defined by the INQUA Com-


mission on the Genesis of Tills, to subglacial, supraglacial and ice-marginal
till formations. The primary processes leading to the deposition are lodge-
ment, melting-out and sublimation. Secondary processes such as sliding by
solifluction, and deformation and translocation by new glaciation rework
the form, structure and texture of the till in many places.
The primary processes, which produce ortho-tills, have been defined by
Dreimanis (1982, pp. 25-26): "is the deposition of till from the sliding base
of a dynamically active glacier by pressure melting and/or other mechanical
processes. Melting-out is the deposition of till by a slow release of glacial
debris from ice by melting and/or sublimation, without sliding or deforming
internally. Most melting-out takes place either at the surface or at the base
of the glacier, while sublimation may occur at the surface only in arid polar
climate."
The secondary processes, which produce allo-tills, take place subglacially,
supraglacially or postglacially Tills deposited beneath a moving glacier may
be deformed and translocated, and tills deposited on sloping ground and
oversaturated with water may slump and slide.
Meltwater flowing on the surface or in tunnels or crevasses of the ice
also transports mineral material, which settles out grain size by grain
size as the stream velocity slackens (Fig. 2-8). Boulders are dragged along
the bottom and usually transported only a short distance, whereas stony
material and sand may be transported several kilometres and deposited as
eskers. The finest fractions are sedimented only in standing water — silts
and clays with graded bedding in fresh water and homogeneous silty clays
in salty water.

GLACIAL TILLS

Characteristics

The erosion, transport and deposition of stony material by glaciers lead


to the formation of material called till. As Dreimanis (1982) defines it, "Till
is a sediment that has been transported and deposited by or from glacier
ice with little or no sorting by water." (Fig. 2-9)
In glaciated terrain till covers the bedrock almost entirely, itself covered
in places by glacial or postglacial sediments. As a mixture of the bedrock
lithologies, it is the most important sampling medium for geochemical
exploration in the Arctic and over most of the temperate zones (Shuts et al.,
1987). Before till can be used successfully, however, its main characteristics
must be known.
22 Glacigenic deposits

DERIVATION
of
BY SUBGLACIAL PROCESSES BY E X T R A G L A C I A L PROCESSES
glacial debris

P o s i t i o n of d e b r i s
IN GLACIER
BENEATH
in TRANSPORT SUPRAGLACIAL
GLACIER
BASAL r ~ ^ ENGLACIAL
before deposition

DEPOSITION
SUPRAGLACIAL
SUBGLACIAL
place AND I C E - M A R G I N A L

a, primary
T
1 LODGEMENT | | J ~ ~ ~ MELTING-OUT ( AND SUBLIMATION )

DEFORMATION LIMITED
MASS
secondary AND F R E E FALL
TRANSLOCATION THROUGH WATER MOVEMENTS

G L A C I A L DEPOSITS
(GENETIC TERMS)

a) general SUBGLACIAL TILL SUPRAGLACIAL TILL

PRIMARY
OR ORTHO-TILLS
h t—r 3
SECONDARY
|| OR A L L O - T I L L S
SECONDARY
OR A L L O - T I L L S
PRIMARY
OR ORTHO-TILLS
n
SUB- MASS MASS
P LODGE- MELT- MOVE-
MELT-OUT
AQUATIC MOVEMENTS
b) specific MENT OUT
o MENT TILLS TILL
varieties MELT- TILLS
OUT -SUBAERIAL ( INCLUDING
or eg. -SUBAQUATIC
TILL TILL SUBLIMATION
la- FLOW e.g. TILL)
in
a TILL FLOW TILL

\, ————4J

Fig. 2-9. Formation of different types of tills. Reproduced with permission from Dreimanis,
1982.

Texture
Till consists of an unsorted mixture of rock and mineral fragments
from boulders to clay size. The particle size distribution depends on
the lithological composition and the amount of preglacial sediments or
weathered material intermixed. Rocks are rather easily broken down into
their natural mineral grain size, or "terminal grade", which commonly lies
in the sand or silt fraction (Fig. 2-10). Further comminution depends on the
abrasion resistance of the minerals; micas and sulphides are easily ground
to clay size, while quartz and feldspars are tougher and concentrated in the
sand fraction (Dreimanis and Vagners, 1971). Glacial milling thus results
in the differentiation of minerals and, simultaneously, the elements.
The grain-size distribution is often bi- or multimodal, with different max-
ima for rock fragments and the principal mineral grains. Short-transported
Glacial tills 23

0 10

0 0.001 0.004 0.016 0.062 0.25 mm

Fig. 2-10. Terminal grain sizes (modes) of selected minerals in tills. Redrawn with permission
from Dreimanis and Vagners, 1971.

till contains more coarse fractions than till of remote origin. Mica-rich and
schistose rocks are more easily comminuted and generate more fine mate-
rial than quartz- or amphibole-rich igneous rocks. Classification of tills on
the basis of grain size distribution is done by sieving the material finer than
stones. Grain size distribution of the fines is analyzed by sedimentation
method.
The finer the material the more surface there is in a weight or volume
unit and the more opportunity for fixing of ions transported in ground-
water solution. The finer fractions of till may contain greater amounts of
heavy metals of hydromorphic origin than of glacigenic origin, whereas the
coarser fractions contain mainly clastically transported heavy metals.
Boulders freshly broken from bedrock are angular in shape. During
glacial transport the weaker parts are worn off and the boulders become
somewhat rounded. Clasts of finer grain lithologies are sometimes striated,
obtaining a bullet-like shape. If the material is transported in water it be-
comes rounded and polished and preserves this shape if again incorporated
in till and transported.

Structure
Most tills are homogeneous and massive, without layering or other
distinct structural features. The circumstances occurring during and even
after the deposition affect the structure, the grain size, the lithology and
thus the geochemistry of tills.
24 Glacigenic deposits

Strong ice pressure generates fissility and high compactness in lodgement


tills (Alley et al., 1986), whereas melting processes are responsible for the
laminae and irregular streaks of sand in melt-out tills. The clasts in the
upper part of till formations have a fine coating of sand on their lower side
due to repeated freezing and thawing. This kind of seasonal activity lifts
boulders towards the surface millimetre by millimetre, resulting in places
in boulder fields or polygonal stone rings. The abundance of clasts in the
surficial parts tends to hinder geochemical sampling from till.
Sediment lenses and flow structures which have destroyed the original
fabric characterize the flow tills. Folding, faulting and other tectonic
features are common in deformation till. During periglacial times, and in
temperate regions today, frost wedges reaching down to a depth of 2-3
metres have mixed the material in a vertical direction in spot-like points,
just as mud boils do in permafrost areas today (Shuts, 1978) (Fig. 2-11).
Each successive glaciation produces its own till bed (Eriksson, 1983), often
with a boulder-rich surface, possibly separated from other beds by layers of
sorted sediments or soil profile.

Fig. 2-11. Vertical movement of mineral material during periglacial time: an ice wedge in
a gravel deposit in southern Finland. This is analogous to a mudboil in till (photo Kalevi
Kauranne).
Glacial tills 25

ORIEN-
TYPE OF TILLS, TYPE AND AGE OF INTERLAYER
TATION

° f 6 0 t f o Q f t o Q ^ ° OQoQo9oo feoO^J^P^Oo ο 0 ο θ θ ΰ o °08 ° ^Q<aoQ


400
jSILT AND CLAYI^-T- ^-^=Τ^.|45800 t |gg|

^ V ^ ^ - ^ O v iff? ΐ>, °7TP


^
SANDY
500
14
Fig. 2-12. Till layers with sand and mull interlayers (drawing P Hakala, C determinations T.
Kankainen).

Fabric
During transport the elongated grains of till tend to be oriented in
the direction of ice movement, the very long ones transverse to it. The
orientation, the striation and the bullet shape of the clasts and stones are
due to the laminar movement of the stone/ice material. Stone orientation
is usual for lodgement tills and may be found in melt-out tills as well (Figs.
2-12 and 2-13). Flow of the till partly or totally reorients the material, so
that the predominant orientation of stones may be perpendicular to the
flow (Lundqvist, 1948; Virkkala, 1958; Kauranne, 1959).
Orientation can be determined by compass and clinometer in the field
(Glen, et. al, 1957; Haldorsen, 1983), under the microscope (Seifert, 1954) or
from the anisotropy of magnetic susceptibility (Puranen, 1977; Pulkkinen
et al., 1980). The long axes can be measured either in a plane or three-
dimensionally (Boulton, 1971). Usually the axes tilt gently upwards in the
direction of ice movement.

Lithology
The rock material of till is mainly of local origin, although some stones
and boulders may be transported several, even hundreds of kilometres
(Fig. 2-14). Among factors influencing the lithology of till are the variable
abrasion resistance of rocks, terrain forms, the thickness and velocity of
the ice and any preglacial materials included, especially at the till/bedrock
interface.
26 Glacigenic deposits

Stage
S t a g e II
youngest
»//>
NORWAY

mL0
* *J**i 'fe USSR

USSR

SWEDEN

Fig. 2-13. Flow directions of Weichselian and earlier glaciations in northern Finland. Reproduced
with permission from Hirvas et al., 1977.
Glacial tills 27

%
80-,

Fig. 2-14. Transport of basic volcanic rock material in different grain sizes. Redrawn with
permission from Perttunen, 1977.

The bottom part of a till sheet is somewhat coarser in average grain size
than the surficial part (Fig. 2-15) and the mineral material down there is
mainly local in both the fine and coarse fractions. Geochemically, therefore,
the bottom part is more inhomogeneous. The upper part of a till sheet,
representing mixed material of more remote origin is lithologically more
homogeneous (Hyvärinen et al., 1973) (Fig. 2-16).
Tills may contain material that was abraded, transported and even
sorted in some earlier glacial or glacifluvial process. Such material may
have been transported in different directions than the bulk of the till
material, complicating the interpretation of geochemical results.
Interpretation becomes especially difficult if the intermixed material
was previously sorted and enriched in heavy metals. Distinct leaching
and enriching may have occurred during weathering of the surface of the
bedrock or older soil deposits. Heavy minerals and metals may have been
enriched during transport by flowing water or by wave action, and heavy
metal ions in water solution may have attached themselves to the surfaces
of mineral grains. The "glacial mill" also separates minerals and elements
into different grain size fractions during transport.
Dreimanis (1983) summarizes the main characteristics of tills as follows:
— Tills are usually poorly sorted diamictons.
— They consist of a variety of rocks and minerals, some of them distantly
transported.
28 Glacigenic deposits

Fig. 2-15. Three difFerent till sheets superimposed. Note the distribution and roundness of the
stones (photo Robert Lagerbäck).

— The surface of many of the basally or englacially transported clasts


is glacially abraded showing percussion and traction marks, even of brittle
sand grains.
— The fabric and structure of tills deposited directly by the glacier are
Glacial tills 29

0 0

^ 0 0 0

0 0 Ü
0
0

% % c·
0 Ü

0
[äT 0 0

®
0
0
0
"% % 0

0 % % 0
0
0
%%
-[35
®
o 39 (61 0 o
[37 4θ1(6θ) 0

® % %
o 0 66)
0 % %
|ϊοο c
0
88 (12
•[57
® Δ Δ
Δ Δ
-[βδ1
0
•[TO
®
ΛΛ


®
0
0
0

Δ Δ

Δ Δ

53
©
50 100 140 m
I _J

Till 31 1 Percentage of (Jg\ Percentage of


local rocks ^-^ long-transported rocks
Till with weathered rock
Weathered bedrock Drill-holes

Fig. 2-16. Proportions of local and remote rock material in till at different depths, northern
Finland. Redrawn with permission from Hyvärinen et al., 1973.

laterally consistent and related to the direction of glacial stress.


— Primary (ortho-) tills are usually underlain by glacially striated and/or
deformed substratum.

Classification of tills

Tills can be divided into two main types — subglacial and supraglacial
till — according to their location in the glacier during transport and
deposition. This was for many years the preferred classification, although
earlier the names used were basal till and ablation till.
The INQUA Commission on the Genesis of Tills has more recently
recommended a new broader classification into primary and secondary till,
according to whether the till was deposited directly by or merely from the
glacier. The former is also known as ortho-till and the latter as allo-till
(Dreimanis, 1982). The terms primary and secondary till are used in the
following.
30 Glacigenic deposits

Primary tills
"Primary tills are deposited directly by glacier ice by lodgement, melting-
out, sublimation or subglacial deformation and traction without noticeable
disaggregation and resedimentation"(Dreimanis, 1982).
Lodgement till is characterized by its tight compaction and fissility
(lamination), its abraded, often striated clasts and its silty major matrix
fraction. Under the clasts in surficial parts there may be a fine sand bed
1-2 mm thick and on the surface of the bedrock a similar thin fine sand
layer, sometimes called rock flour. No other signs of sorting are seen. The
stone orientation is usually well developed and the clasts have obtained a
bullet-like stoss-lee shape. The material is mainly local.
Subglacial melt-out till is deposited from stagnant ice by melting and
closely resembles lodgement till. The fabric usually shows the earlier flow
direction and the degree of compaction is high. Fine lenses of sand or silt
are included, and clasts may be coated with fine sand. The material is
local or more remote depending on whether it was transported at the base
(subglacially) or higher up (englacially) in the glacier (Fig. 2-7).
Supraglacial melt-out till has been transported either englacially or at
the surface (supraglacially). In mountain valleys the material may even be
of extraglacial origin -talus material from mountain slopes. The lithology
is thus heterogeneous and stones are partly oriented parallel and partly
transverse to the direction of glacier movement. The clasts at the surface
are angular if not rounded by meltwater streams and are subrounded if
they have been transported englacially. The grain size is coarser on average
than in the other primary tills, and some sorted materials deriving from
meltwater washing activity are present.
Deformation till is folded, faulted and sometimes thrust off from its
place of sedimentation. The orientation of stones may be disturbed and a
new orientation emplaced. Deformation till might equally well be included
among the secondary tills.

Secondary tills
"Secondary tills are formed from glacial debris which has undergone
redeposition shortly after its release from glacier ice. They are deposited
from glacier ice rather than by glacier ice, in most cases by mass movement
or a short limited free fall of glacial debris subaerially or subaquatically,
with little or no sorting"(Dreimanis, 1982).
Subglacial subaquatic till (water laid till) is deposited by dropping of
mineral particles from the glacier sole through shallow water beneath
the glacier. Poorly sorted, it represents a transition between till and
glaciolacustrine sediments.
Subglacial flow till forms on the lee side of rocks occurring as obstruc-
tions, as in drumlin tails. Flow structures of sorted mineral material
inter layers are typical.
Morainic landforms 31

Supraglacial flow till is formed on dipping ground from till flowing


down either subaquatically or, when inner friction and cohesion decrease
critically upon oversaturation, subaerially A new orientation is assumed
during the mass movement, and the structure may be stratified. Beds of
sorted sediments may be slumped with and intermixed in till.
Tills are identified and classified according to their colour, grain size
composition, grain orientation, roundness of clasts and stones, lithology,
heavy mineral composition and of course chemistry. Pollen and diatomacae
can be used for identification and differentiation in cases where pre- or
interglacial materials have been intermixed, and the weathering degree of
minerals when preglacial crust is present (see, e.g., Dreimanis et al., 1957;
Kauranne, 1960b; Eriksson, 1973, 1983; Shuts, 1976; Perttunen, 1977;
Ehlers, 1982 and Geddes, 1982).
Tills have been widely used in exploration pedogeochemistry in the
Arctic and temperate zone. Case histories are published frequently by AEG
and IMM (see, e.g., Kvalheim, 1967; Bradshaw, 1975; Kauranne, 1976a;
Björklund, 1984; Shilts, 1984b; Gleeson and Nichol, 1987; Rogers, 1988;
DiLabio and Coker, 1989).

MORAINIC LANDFORMS

Distinct glacigenic formations consisting mainly of till are called


moraines. Many classifications have been suggested and there is a vast
literature on their structure and origin (Okko, 1941; Prest, 1968; Aario,
1977; Lundqvist, 1980).
The INQUA Workgroup on Landforms has developed a system for
classification of glacial tilly formations within four main groups: subglacial
landforms parallel to ice movement, subglacial landforms transverse to
ice movement, unoriented landforms and ice-marginal landforms. The tilly
parts of moraines have been transported by ice and geochemical anomalies
found in those parts will be mainly glacigenic; the sorted parts have been
transported or deposited in water and the geochemical anomalies they
exhibit are likely to have a more complex genesis.

Landforms parallel to ice movement


These streamlined moraines are deposited from basally or englacially
transported debris, lodgement or basal melt-out till. The shapes of forma-
tions are determined by alternating erosion and deposition. The four main
types of these landforms are drumlin, drumlinoid, fluting and lee-side cone.
Drumlins are whale-back-shaped, elliptically elongated hills with height
maximum near the stoss side and a gently sloping ridge as a tail (Fig.
2-17). Commonly they enclose a rock core. Drumlinoids are larger cigar- or
32 Glacigenic deposits

Fig. 2-17. Streamlined forms of tilly hills known as drumlins. Reproduced with permission from
Gluckert, 1976.

spindle-shaped elevations. Fluting appears as a series of low ridges parallel


to the direction of ice movement and covers broad fields. The ridges are
fairly easily recognized from air photos. Lee-side cones, or crags and tails,
are formed behind a bedrock knob and vary considerably in size.
The material of drumlins is typically fine-grained till at the proximal
end and coarse, even partly sorted material at the distal downstream end,
at the lee of the bedrock knob. The flutes and drumlinoids consist of
lodgement till, the lee-side cones of lenses of tilly and sorted materials
with interbedded boulders. The material of all these formations has been
transported with the ice and mineral separation has occurred only through
grinding, except in tails where some small scale sorting by meltwater has
occurred. Because of their simple genesis and relatively short transport, all
the above materials are suitable for geochemical mapping and exploration.

Landforms transverse to ice movement


Washboard and ribbed moraines which occur perpendicularly to the
striae are formed under the ice. Their material has been transported by ice
and the almost negligible separation of minerals has occurred mainly by
grinding. Deposition has been partly by lodging, and partly by flowing, and
in the latter case water has also acted as a separator of grain sizes. The
ridges are irregular both in width and height.
Washboard moraines, also known as De Geer moraines, are parallel
sets of various-sized ridges clearly perpendicular to the ice movement.
Their material is lodgement till (De Geer 1940). Ribbed moraines, or
Rogen moraines, are more irregular ridges, resembling fluting in covering
broad fields. Often the stoss side and crest of the ridges are fluted. Ribbed
moraines consist of mixed lenses of till and sorted material. Thrust moraines
Morainic landforms 33

resemble ribbed moraines in form but are larger and include more single
ridges; they consist of lodgement till but also contain large rock slabs.
Probably they were formed in front of an oscillating ice lobe. Like that of
the parallel landforms the material of transverse landforms is suitable for
geochemical studies.

Unoriented landforms
The material of unoriented moraines has been deposited by lodging and
the geochemical anomalies found in them, as in the parallel and transverse
formations, are mainly glacigenic. The irregular shape is due to moulding
of the surficial part by freezing and thawing and the rise of boulders. In
submerged areas the surface may be washed poor in fines as well. Often
the main bed of lodgement till is covered by sorted melt-out till of irregular
thickness.
Cover moraine is a thin even layer of lodgement till following the surface
of the bedrock. Hummocky moraine forms a more irregular partly melt-out
till layer without orientation, with an average thickness of a few metres.
Till plain is a basal till formation, thicker in valleys and thinner above
bedrock surface elevations. Often it consists of multiple till beds from
several glaciations, which complicates its use for geochemical studies.

Ice-marginal landforms
Ice-marginal moraines are formed in the melting phase of the glaciation
at the ice front or at the sides of an ice lobe (Fig. 2-18). In summer
the ice edge retreats by melting, and in winter it may push forward a
little. Although the material was originally emplaced by the glacier as till,
meltwaters have manipulated part of it before deposition. Because of the
complex transport of the material, these formations are much less suitable
for geochemical prospecting than the other moraines.
The end moraines formed along the front of the retreating ice consist on
their stoss side of basal till lenses mixed with sorted material and on their
distal side of outwash layers of gravel, sand and silt. The annual moraines,
each formed during a single summer, are only a few decimetres in height.
Terminal moraine is formed at the edge of an ice lobe over a period of years.
Where the glacier paused longer, for tens of years as e.g., at Salpausselkä
near Lahti, Finland, deposits may exceed 100 m in thickness (Fig. 2-19).
Terminal moraines are often hummocky consisting of supraglacial melt-out
till and boulders.
Crescentic ridges of push moraine (thrust moraines) are formed of till
or sorted materials by the winter thrust of the ice edge. Käme moraines
are formed by the melting of dead ice, so that this kind of knob and kettle
terrain is partly covered by till and partly by sand and clasts.
34 Glacigenic deposits

ESKERS AND MARGINAL DEPOSITS


* - -
\ ^
"> ' / !
I
^

\\ \ I (A ICE,LOBE

w \r&

scale km

Fig. 2-18. Eskers showing the directions of crevasses and tunnels in the glacier and ice-lobe
margins, usually almost parallel to the earlier directions of ice movement, and marginal deposits
showing the different positions of the glacier rand. Map from southern Finland with the great
Salpausselkä end moraines.

N
Proximal s
Distal

140

120

100

80

60

100 200 400 m

mixture of till, silt llljll silty-clay


and gravel layers and blocks
|?V§j fine sand
till |·';·;-;| sand and gravel

Fig. 2-19. Cross-section through Salpausselkä I end moraine at Lahti. Redrawn with permission
from Hyyppä, 1966.
Sorted glacial sediments 35

Lateral moraines are formed between ice lobes and grade into end
moraines at the ice front. They consist of lodgement till and sorted
material transported from either side of the formation. Crevasse fillings
consist mainly of sorted sediments and resemble radial eskers. Ablation
moraines are very similar to crevasse fillings but formed at the front of the
melting glacier or around melting dead ice lumps; they represent a gradual
transition to glacifluvial formations, and contain lenses of till and layers
of gravel, sand, silt and even clay Suppas, deep conical depressions in the
landscape, some still filled with water, mark the former sites of dead ice
lumps.

GLACIFLUVIAL PROCESSES

The bedrock surface of formerly glaciated regions is covered by till except


for high mountain tops or where flowing water or wave action has washed it
away In areas that have been under water at some point after the melting
of the ice, the till is covered by glacial sorted fine-grained sediments.

Erosion
Meltwater erosion is defined by the ice. Meltwater comes into contact
with bedrock and older overburden formations only at the sides of valley
glaciers, in front of land ice or at the bottom of crevasses and tunnels in the
ice. Mainly it erodes the mineral material already transported by the ice.

Transport, sorting and deposition


Transport by water may occur in one stage or several successive stages.
During the transport the material is sorted into different grain sizes
according to the velocity of water flow and the weight and shape of the
particles. The coarser grains are sedimented first and the finer grains
transported further. The finest particles are sedimented only in standing
water; in fresh water they are somewhat separated according to particle
form but in salty water flocculated and deposited as such.

SORTED GLACIAL SEDIMENTS

Characteristics
Texture
The narrow range of grain sizes is the most striking feature of sorted
sediments, though the degree of sorting varies widely, being highest in
coarse fractions and lowest in the finest fractions. Sorted sediments where
36 Glacigenic deposits

Fig. 2-20. The surface forms, grain size and sorting of glacifluvial sediments depend on whether
they were deposited above or under water, i.e. above or under the highest coast line.

one grain size dominates thus contrast sharply with till where all particle
sizes from boulders to clay flakes are represented. In glacifluvial and
glacilacustrine formations, layers of different grain sizes are superposed
one above another (Fig. 2-20).
The grains, especially clasts of sorted sediments, have characteristically
been rounded by running water or by wave action. Only the very finest
mineral particles are not rounded and polished. The roundness depends on
stream velocity or wave activity, length of time and distance of transport,
and the abrasion resistance of the rock.

Structure
Sediments formed in close contact with the rim of the retreating ice are
coarser than those sedimented farther from it. Beds of gravel, formed in
front of retreating ice typically lie under beds of sand, silt and clay, though
re-advancing of the ice may cause coarse material to be emplaced on top of
fine sediments. Boulders from floating and melting icebergs are sometimes
dropped and buried into the surficial clay of glacimarine and glacilacustrine
deposits.
In fresh and brackish water, the negative surface charge of silicate
minerals keeps the clay flakes apart from one another; silt sedimentates
rapidly but clay continues to sedimentate throughout the winter. In salty
water, the cations of the water "solution" become fixed on the surfaces
of grains and neutralize the repellent forces between mineral particles,
enabling them to flocculate and settle with greater speed.
The summer layer of glacilacustrine clays consists mainly of silt and is
thick, whereas the thinner and darker layer, deposited in winter when no
more material is entering the basin, consists solely of clay-size particles.
Formations of sorted sediments 37

The silts and clays of fresh and brackish water formations show graded
bedding, while those of salty water are massive.

Fabric
Elongated particles are typically oriented perpendicular to the wa-
ter movement, that is, transverse to the esker ridge or parallel to the
beach. Later processes, like mass movements or overriding glacier, may
re-orientate the particles as well as cause folding and faulting of the beds.

Lithology
Glacifluvial deposits are typically polymict, having obtained their mate-
rial both from the till in the ice and from older deposits under the ice. In
smaller formations the material tends to be local; in larger formations it
is of more remote origin. The finer materials have been transported longer
distances. Quartz and feldspars dominate in sand and silt layers, the micas
in clay layers. The great amount of fine materials in glaciated areas in
comparison to the coarser ones strongly suggests that a large part of the
clay originates from preglacial clay formations.
If formations of sorted sediments are sampled by drilling, the different
layers may become mixed and the sample till-like in appearance. The
chemical and lithological contents of sorted sediments are nevertheless
different from those of till from the same locality, owing to the different
transport and separation processes. Utmost care must thus be taken in
sampling formations where mixing of materials may happen.
Till should always be preferred over sorted sediments in geochemical
exploration. The provenance of sorted sediments is much more difficult
to find, because the transport has occurred by at least two means and in
two different directions with separation of original constituents from each
other. The use of sorted glacial sediments in geochemical exploration is still
very much at the experimental stage (Lee, 1965; Shilts, 1972; Kauranne
et al., 1977; Smee, 1983; Coker and DiLabio, 1989). The great economic
value of sorted sediments for other purposes has been evaluated in several
publications (e.g. Lüttig, 1990; Königsson, 1992).

FORMATIONS OF SORTED SEDIMENTS

Like the type of material, so the size and shape of glacifluvial formations
depend on the place of sedimentation. The sedimentation may occur inside,
in contact with or near by the glacier ice. Because the purpose of this book
is to assist in geochemical exploration, only the main types of formations
are described, and only to the extent necessary. Until now, assaying of
glacifluvial sediments for purposes of geochemical exploration has not given
encouraging results. They may be used in regional mapping but not in
target prospecting.
38 Glacigenic deposits

Subglacial meltwater formations


Subglacial meltwater formations, which together can be called eskers,
typically form where a sharp escarpment in the bedrock surface causes
faulting in the glacier, giving rise to a crevasse or gorge. Formations
sedimented at such places are classified as beaded eskers, engorged eskers,
squeeze-up eskers, esker chains and crevasse-filling eskers according to
their mode of appearance (INQUA).
Eskers vary in height, width and length. Minor ones may be as small as
two metres high, ten metres wide and a few hundred metres long. Glacial
or postglacial clays may cover them entirely. By contrast major eskers may
grow to over 100 m in height, two or three km in width and a few hundred
km in length.
Since eskers were deposited in subglacial tunnels or ice crevasses, their
form and material are influenced by the original shape and inclination of
the channel. The final shape may have been determined by the slumping
of material at the sides, by sedimentation of till or clay on the top, or by
erosion of the flanks by waves on ancient beaches.
Most eskers were formed subaquatically by the deposition of sorted
materials at the mouth of a subglacial tunnel. A hummock of gravel is
formed in spring and covered by sand in autumn (Fig. 2-8) and varved
silt in winter. Several hummocks together form a chain-like esker. Eskers
formed in crevasses show a somewhat similar structure, with alternating
coarse and fine material layers dipping at the sides and somewhat away
from the glacier.
Often eskers are pitted with dead ice depressions (suppa in Finnish),
and if the eskers were subaquatically formed, the bottom of the suppas are
covered with clay. The sorted material of eskers is easily eroded and the
waves of an ancient lake or sea may have levelled the crest or cut terraces
into the flanks. The stages of the later land uplift can be measured from
the height of these ancient beaches.
Eskers formed subglacially in ice-bottom tunnels where water was under
pressure may run up as well as down valley slopes, though most often
eskers run parallel to them. The till that frequently covers this type of
esker may deceive the careless sampler into assuming the whole formation
is tilly.

Ice-marginal meltwater formations

Glacial mounds consisting of a mixture of till, gravel and sand with


possible silt cover are deposited in close contact with the rim of melting
ice. Upon melting, dead-ice lumps hidden in the sediments and till cause
slumping and sliding of the original strata. These kinds of formations
are called kames and include moulin kames, käme plateaus, käme terraces,
Conclusions 39

and käme deltas. All are difficult to use as material for geochemical
exploration.

Proglacial meltwater formations


Deltas and sandurs
Glacifluvial deltas and sandurs consisting of sand and somewhat coarser
and finer materials form immediately in front of the melting glacier. Deltas
are formed subaquatically, and the sorting of materials and layering of
different fractions is somewhat sharper than in the supra-aquatic sandurs
deposited above the water surface. Above the highest coastlines, all sandy
formations are poorly sorted sandurs, with the exception of those that were
deposited in ice-dammed lakes.
The sandurs along the slopes of valleys are often pitted with dead-ice
hollows and the surface does not form such a smooth terrace as that
of valley deltas. Deltas and sandurs on valley bottoms are often covered
by postglacial beach or river sediments. It is important to distinguish
between glacifluvial and postglacial outwash sediments when sampling for
geochemical purposes; postglacial sediments typically have been formed by
spring floods and contain material from the upper stream valley, whereas
the glacifluvial ones tend to contain mostly washed till originating from the
sides of the valley. Flood deposits are sometimes called overbank sediments.

Glacilacustrine and glacimarine silts, clays and muds


All areas below the highest coast-line will have been covered by fine
meltwater sediments of glacilacustrine silt and clay. In freshwater clays,
annual variation in the melting process causes a varved structure. The
thickness of the lighter coarser summer varves depends on the warmth
of the particular summer. The annual retreat of the ice can be followed
by identifying the successions of similar varve patterns and counting the
individual varves.
Where changes in the salinity of water bodies have occurred, laminated
glacilacustrine sediments may be covered by varveless glacimarine sedi-
ments. The massive structure of marine clays, such as found in the Baltic
Sea, has been enhanced by the flocculating effect of salts and humus in the
warm sea water of the first thousand years after the melting of the ice.

CONCLUSIONS

Glacigenic deposits are direct products of glacial activity: abrasion,


transport, comminution, separation and sedimentation. There are two
main types of deposits: (1) glacial till, metamict material deposited by or
40 Glacigenic deposits

from ice, and (2) glacifluvial deposits transported and sorted by water
and deposited either in running water (glacifluvial proper) or in standing
(glacilacustrine) fresh or brackish water.
Till consists of fresh or weathered rocks newly torn from bedrock and
mixed with earlier deposited materials of overburden. Till is a mixture of
all possible grain sizes and lithologies, where every grain size fraction has
its own lithological and geochemical character, varying with depth.
Till represents the lithologies of the bedrock beneath, having been trans-
ported on average only a few hundred metres. Material at the till-bedrock
interface is local, with the surficial parts representing further transported
material. Occasionally, distinctive lithologies have been identified as far
away as 1000 km from their source (Schuddebeurs, 1981). Grains gouged
out from an ore suboutcrop form a plumelike anomaly in till (Drake, 1983),
narrow and strong at the apex and widening and weakening with distance
from the source. Typically the anomaly touches the surface two or three
hundred metres from the mother lode (Kauranne, 1976a; Lehmuspelto,
1987). Metal concentrations are lower and more homogeneous at the sur-
face and highest close to the bedrock. Standard deviation is also greatest
and anomaly/background contrast strongest at the bedrock interface.
Glacifluvial deposits are stratified and characterized by a distinct layering
according to grain size. The coarse fractions consist of multimict lithology,
whereas the finer fractions are more monomineralic. The intensity of
fractionating has been proportional to the ratio of grain mass to water
velocity. Under favourable conditions, layers of ore minerals are formed.
If such a layer consists of an economically valuable mineral (garnet,
magnetite, cassiterite or native gold) it is called a placer deposit and is, of
course, geochemically highly anomalous. Silt and clay sedimented in fresh
water are distinctly laminated but are, if deposited in salty water, with no
annual layering.
Material of glacifluvial deposits has been transported several kilometres
on average, the larger clasts and heaviest minerals a shorter distance than
the clay-size particles of light minerals like mica. Tracing of clastically
transported grains in sand and gravel back to the source may be of utmost
difficulty but can be solved by sophisticated chemical, isotopic (Gulson and
Vaasjoki, 1987) and mathematic methods (Granath, 1983). There may be
considerable amounts of hydromorphically transported heavy metal ions
precipitated on the surfaces of silt and clay grains. Water moves very slowly
in these fine grained materials but given time may form strong anomalies.
The provenance of such anomalies may be tedious to discover, but easier
than for clastically formed anomalies in sorted sediments.
41

Chapter 3

NONGLACIAL OVERBURDEN

MINERAL FORMATIONS

Introduction
In the mountainous part of arctic and temperate regions rock expo-
sures are common but in flatter areas the bedrock is mostly covered by
overburden. The overburden consists of either regolith formed in situ or
transported mineral formations. Both may be covered by organic deposits
or water.
In nonglaciated areas the overburden is mainly formed by weathering in
situ and there is a gradual transition from the unconsolidated overburden
to the hard bedrock. The transported sediments have a similar distribution
to the postglacial sediments in glaciated terrain.
Even within glaciated terrain, however, there are considerable areas that
have never been worked over by glaciers and where no glacigenic drift
exists. The loose deposits found in the depressions on highest mountain
tops, which during glaciation appeared as nunataks, are mainly sorted
sediments, ranging from talus cones to stream gravels, and including of
course slabs and gravel of the in situ weathered crust. Isolated remnants
of the preglacial weathered crust are also preserved in glaciated lowlands
where ice abrasion was minimal, not only at the ice divide and in the
bottoms of valleys perpendicular to the ice movement, but almost anywhere
in depressions as thin patchy layers. Nonglacial formations can also be
found as inter layers between glacial or glacifluvial sediments.
Dune sand and loess silt, found both in and outside glaciated terrain gen-
erally have been derived from glacial deposits and are thus semiglacigenic.
During the glacial period or soon thereafter they were formed by the
action of strong winds on mineral deposits not yet sufficiently sheltered by
vegetation.
The geochemist assaying secondary materials needs to know how closely
the overburden resembles the bedrock beneath — whether till is a direct
crushing product of bedrock and whether sorted sediments have been
derived by direct sorting of till.
42 Nonglacial overburden

The inventoried or calculated amounts of different types of sorted


material in Finland (338,000 km2) are 13 Gm3 gravel, 36 Gm 3 sand and 48
Gm 3 silty clay. The ratio of these materials in the overburden of Finland
is thus about 1:3:4. The ratio of the same fractions in common (sandy)
till is 2 : 5 : 3 , and in a single crushing round by jaw crusher from a
medium-grained granite 8 : 9 : 1 .
In Makola, in flat central Finland, the thickness of a nickel ore abraded
from the ore subexposure by the last glaciation was only 60 cm as calculated
from the nickel contents of the till bed, 4 m thick and the surface area and
grade of the ore suboutcrop (Kauranne, 1957). Yet the mean thickness of
the till in Finland is approximately 6 m and the total amount of till in the
country about 2000 Gm3. This would correspond to about 1400 Gm 3 solid
rock and a mean thickness of abrasion 4 m rather than 60 cm.
These figures show clearly that neither the till nor the sorted sediments
in Finland can be the product of crushing of rock alone, but must contain
a portion of preglacial weathering crust, preglacial sorted sediments and
materials from previous glaciations and interglacials. The contribution
of these other possible components must always be kept in mind when
interpreting geochemical anomalies found in the fine fraction of till.
Evaluations of the amount of preglacial material in till have further
been done by grain size analysis (Kauranne, 1960b) and chemical analysis
(Salminen, 1976).

Weathering
Rocks formed deep down in the crust are labile and sensitive to the
attack of mechanical and chemical forces when brought to the surface.
Especially high up on mountains, barren, steep rock faces shatter under
the stress of rapid temperature changes. Temperature expands minerals
differently, setting up tensional stresses and causing the rock to crack along
mineral contacts or planes of weakness in the mineral lattice. The repeated
freezing and thawing of free water in joints and capillary water in pores
has the same effect.
Final products of the mechanical disintegration consist of monomineralic
grains, the size and form being characteristic of the mineral in question.
Material is then dispersed mechanically with separation of constituents
along the transport route, usually with the help of running water. Waves
and wind assist in the task. Finally, through sedimentation, the materials
are classified according to their physical characteristics. The parent rock,
topography and climate thus all have their effect on the course of this
evolution.
Physically weathered rock has much more surface available for chemical
interaction with water than has the sound original rock. Even rain water,
which is slightly acidic, reacts on contact with silicate surfaces, as shown
by the increase in pH of water absorbed by quartz (pH 6-7), micas (pH
Mineral formations 43

7-9), feldspars (pH 8-9), carbonates (pH 8-10) and amphiboles (pH 10-11).
The presence of carbon dioxide increases the leaching capacity of water;
calcite, for example, then goes totally into solution in the form of calcium
bicarbonate. The common ions from the minerals of granite are leached
during weathering in the order Ca, Mg, Na, Ba, K, Si, Fe, Mn, Ti, Al. Final
products of chemical disintegration are ions in water solution and secondary
minerals that are stable in surficial conditions, e.g., clay minerals.
A more common process of chemical weathering than by direct dissolu-
tion is hydrolysis, where in the reaction between water and mineral both
acid and base are formed. As an example of hydrolysis, the change of potash
feldspar occurs as follows:
2KAlSi 3 0 8 + 6 H 2 0 + C0 2 -> Al2Si205(OH)4 + 4SiO(OH) 2 + K 2 C0 3
Before hydrolysis some minerals become hydrated (e.g., CaS04 + H 2 0 —»
CaS04H 2 0, anhydrite is transformed to gypsum).
Yet another common weathering reaction is ion exchange, or replacement
of an ion in the mineral lattice by another ion from the interstitial water.
Minerals in the mica group readily exchange ions lying between silica oxide
tetrahedron and aluminium oxide octahedron layers. Plants may contribute
to this reaction.
The most striking changes of all take place in oxidation; in the case
of iron, for example, there is a drastic change of colourless ferrous iron
solution of ground water to brown red ferric iron limonite precipitate (Fig.
3-1). Sulphides are easily oxidized and leached away, and their space in the
rock is either left empty or filled with material such as limonite. Often the
products of oxidation and hydration are of greater volume than the original
minerals and produce stresses sufficient to disintegrate the rock (Rankama
and Sahama, 1950).
Plants and animals play their part in loosening the surface of the
overburden, creating pores for water and air to intrude. Organic tissue
usually decays by oxidation, with the production of new acidic compounds
able to attack minerals. Organisms also increase the C0 2 content of soil
water, and thereby its corrosivity Some weathering of organic matter also
occurs in reducing environments, e.g., through the activity of anaerobic
bacteria.
Dense and fine-grained, homogeneous, silica-rich, unoriented rocks are
more resistant against weathering than coarse, jointed and schistose,
multi-mineral, Fe- and Mg-rich rocks. In general, high physical strength
also means greater chemical resistivity

Erosion and transport


Mechanically and chemically disintegrated rocks are easily eroded and
moved by ice, water and even air. The primary forces behind erosion and
transport are solar energy and gravity. The sun evaporates water from the
44 Nonglacial overburden

H
ACID P ALKALINE

Fig. 3-1. Solution and precipitation of the elements, e.g., iron, mainly depends on redox potential
(Eh) and hydrogen ion concentration (pH) of the solvent. Conditions in ground water of humid
and arid regions as well as in rain water are redrawn after Friedman and Sanders (1978).

surface of the Earth, which then precipitates down, filling streams and
piling up snow on glaciers. The bottom parts of a moving glacier are filled
with a mixture of stone material. Boulders transported by ice multiply its
abrasional force. Meltwater on a glacier carries material from nunataks
and wind blown dust (DiLabio and Shuts, 1978). Streams transport loose
material either in solution, in suspension or as bottom load. Stones dragged
along the stream add its erosional effect. Stone material is comminuted
in connection with transport by ice or running water. The sun also raises
the wind, which then carries fine particles, and waves which then remove,
crush and classify materials on beaches.
Gravity determines the predominant direction of flow On steep valley
sides material is continually rolling, slumping or creeping downwards.
Water flows down by the force of gravity whether on the surface or inside
the Earth. At the same time, independent water molecules in the voids
of overburden above the ground-water table strive upwards by capillarity.
Likewise water as steam, and radon, argon, helium, oxygen, nitrogen,
carbon dioxide, methane and the other soil gas molecules, "climb" upwards
in the soil, transporting up the heavy metal ions adhering to them.
Mineral formations 45

Sedimentation
Layers formed solely by gravitational transport are poorly classified; no
seasonal layering builds up, even though the slumping of material occurs
seasonally, in spring when the soil loses its strength, cohesion and inner
friction through melting of the frozen water in pores and joints, and in
autumn when the same thing occurs through oversaturation by infiltrating
rainwater.
When sedimentating again in unstratified deposits the different grain
sizes and lithologies of layers become intensively intermixed, although
geochemically each separate grain size fraction may be rather homogeneous.
Particles transported by surface water settle out along watercourses ac-
cording to weight, the coarsest fluvial sediments settling on the upper, hilly
part of the stream where gradients are steepest and the velocity greatest.
Destruction of primary bedding by new flood erosion and resedimentation
tends to lead to still greater order and sharper classification, but further
downriver vast areas are covered by mixed mineral and organic material.
The balance between erosion and sedimentation depends mainly on the
water velocity, as illustrated by the well-known Hjulström diagram (Fig.
3-2). Flood sediments, sometimes called "overbank sediments", can well be
used for regional geochemistry.
Mineral grains of sand size may be classified further according to their
density by waving water on beaches. In littoral sediments minerals like
magnetite and garnet form distinct layers between feldspar and quartz-rich
layers. The geochemical character of different layers varies distinctly.
The slower the water flows the smaller the particles settling, and in
fresh water the sharper the grain weight classification. Silt- and clay-sized
particles are sedimented in lake or sea where water is almost standing. In
fresh, cold water a few metres deep and close to glaciers, not only summer
and winter layers but day and night layers are formed and may later be
discriminated.
In sea water the natural negative charge of clay flakes, which keeps
other flakes away, is satiated by adsorbed positive cations or organic
colloids. The clay and silt particles are flocculated and no seasonal strat-
ification occurs. Geochemically salt water clays are rather homogeneous.
The chemical/physical character of marine clays depends on the type of
cations in it and the electrostatic forces strengthening the structure caused
by them — the cations can easily be exchanged, but if totally leached out
the structure collapses. Sedimentation may be disturbed and layers already
formed damaged by alternating or turbulent bottom currents.
Ground water carries material of colloidal and ion size. Sedimentation
of the material transported by ground water occurs when the chemi-
cal environment changes. For example, limonite and lime are precipitated
46 Nonglacial overburden

256

126

■64

32

16

TRANSPORT Hjulström

Settling rate SEDIM ENTATION


in still water

0.5
0
O.OCK 0.015 0.06 0.25 1 16 6Λ mm
Clay Silt Sand Gravel Stones

Fig. 3-2. Erosion of sediments, transport and sedimentation of suspended material depend on
grain size and the speed of water flow. The well-known Hjulström (1939) nomogram has been
improved by Sundborg (1956) and is presented here modified from Friedman and Sanders
(1978).

when seeping water meets an oxygen-rich environment at the soil surface


or in springs (Fig. 3-3).
Water is a "soft" element, yet becomes so powerful at high velocity that it
can move giant boulders. Air is even gentler, yet wind too has abrasive force
and, given time, can transport huge amounts of material over open areas.
The most typical aeolian formations, dunes consisting of fine sand-size
material, are formed in treeless areas — on beaches, on the foreland of
glaciers and in arid deserts. Thick beds of silt-size dust called loess cover
vast areas of the Soviet Union, China and the United States just outside
the maximum extension of Pleistocene glaciation. Similarly materials of
volcanic eruptions are transported by the wind, even half way round the
globe. Close to a volcano, cobble-size "bombs" and stone-size "lapilli" may
be found, while further away there are extensive, thin layers of ash-size
"tephra". The grain-size composition of selected sediments is shown in Fig.
3-4A and B.
Fig. 3-3. Hydromorphic transport of elements in surface and ground-waters. Conditions deeper in the ground are reducing, but when
ground-water seeps up to the surface, coming into contact with air, soluble ferrous ion is oxidized to ferric form and precipitated as slimy
red ferri-oxy hydrate. The oxyhydrate changes into limonite, which may later form goethitic iron pan.
48 Nonglacial overburden

99.5

loess Kansas
dust Kansas
river sand
beach sand
stream sand
glacial till

2
0.004 0.015 0.06 0.25 16 64 mm

Clay Silt | Sand Gravel Stones

-99.5

-99

-95

-80
-60
-Λ0
-20

-5

-1

I-0.5

-2 0 10 0
64 16 4 1 0.25 0.06 0.015 0.004 0.001 mm

Gravel Sand Silt Clay

Fig. 3-4. Examples of grain-size composition of selected sediments presented in the European
way (A) and in the American way (B) on log/probability paper where the Gaussian distribution
curve of grain composition of a single crushing product forms a straight line. Modified from
Friedman and Sanders, 1978.
Mineral formations 49

In situ deposits
In situ deposits are on their original site; the rocks are physically and/or
chemically altered but not moved. Chemical composition of the secondary
material is not the same as that of the underlying primary source rock, since
concentrations of some elements have been increased and concentrations
of others diminished.
On the top of mountains the overburden typically consists of cobbles
and slabs loosened from the bedrock beneath. Exposed to great and rapid
temperature variations, the material has merely been broken loose and is
unchanged mineralogically or chemically, forming a type of resistate. More
common resistates are those formed by chemical alteration. In temperate
climates weathering leads to silica-rich products. When sulphides or other
easily soluble minerals have been leached out and transported away, rock
consisting of quartz or quartz-feldspar skeleton with open voids, sometimes
called "sugar rock" is formed. More often the voids are filled with limonite,
rich in lead, cobalt or other metals coprecipitated from sulphides. Such
weathered material easily collapses downslope, forming residual quartzose
or arkosic sand.
These limonitic fillings are in fact in situ oxidates. Though small in
amount they can form striking geochemical anomalies by absorbing heavy
metals. The limonite film around silicate grains acts similarly by capturing
heavy metals from the ground-water solution and is therefore important in
geochemical prospecting. Larger formations, where limonite is cementing
silicate grains are familiar as gossan or "iron cap".
The true clays formed in warm humid climate, even in the Nordic
countries during the Tertiary, are a product of partial leaching and re-
arrangement of crystal lattices. Enriched in aluminium, potassium and
sometimes iron, chemically they are hydrolysates. Their lattices are built up
of aluminium oxide octahedron and silicon tetrahedron layers with cation
interlayers. Some of them, like montmorillonite, expand, when they capture
water and large cations. Because of their heavy metal scavenging activity
they play a greater part in geochemistry than the other rock forming miner-
als. In places, considerable portions of such preglacial clays are intermixed
with till, causing "false" anomalies and confusion in the interpretation of
the mainly low glacigenic anomalies. The really false anomalies are caused
by either analytic error or contamination during the sampling.
Geochemical anomalies in the in situ soils lie almost directly above their
source, but the element contents may be higher or lower and certainly in
different ratios than those in the original rock. The type and magnitude
of anomaly does not directly indicate the type, grade or volume of the
mineralization beneath. Yet the concentrations of metals in overburden
have in some countries widely been used in predicting the ore potential of
the different types of regions (Solovov, 1987).
50 Nonglacial overburden

Transported deposits
Transported nonglacial deposits have been moved by gravity, water or
wind in one or more stages from their place of origin. The type of transport
must be known when tracing their source.

Gravitational deposits
The simplest transported formations are those formed by gravity. In
talus cones the material has been rolling almost directly down the hillside,
with slight sorting only according to size and form of the particles: the
smaller and more isometric are transported further down and the larger
and slablike a shorter distance. The size and form of particles depend to
some extent, of course, on the rock type.
Friction soils, especially in spring when waterlogged and underlain by
a frozen layer, tend to creep downslope by simple gravity, in a slow
process called solifluction. Originating in the repeated freezing and thawing,
solifluction destroys stratification, stone orientation and other original
textures.
Cohesive soils behave differently, they slump rapidly when the pore water
pressure exceeds the cohesion between particles and the stresses exceed the
shearing strength of the soil. The slumping of greater masses that takes
place along a concave shearing surface is called a landslide.
The geochemical anomalies found in these gravity formations all point
upwards. However the overburden material of the slopes often has earlier
been transported along the valley, so that the gravity anomalies do not
necessarily point to the original source of the material. Such anomalies
are thus "second generation". Since gravity anomalies are mainly formed
by relatively short mechanical transport and of clastic material it should
be easy to trace their source by combined geochemical, geological and
geophysical studies.

Fluvial deposits
Running water as transporting force and medium is the common de-
nominator for fluvial deposits, which may consist of pure mineral, pure
organic or mixed material. A separate volume of the Handbook deals with
these important materials of present watercourses. We nevertheless include
some mention of the geochemical character of fluvial deposits, because
ancient fluvial deposits are widely encountered and glacifluvial deposits in
particular cover large areas outside now existing rivers.
Present-day rivers transport 13,695 million tonnes of solid and 3,600
million tonnes of dissolved material down to the sea annually. By way
of comparison, 1,000-1,500 km 3 ice, containing on average 1.6% mineral
deposits and generating 35,000-50,000 million tonnes of sediment, melts
each year in the Antarctic (Friedman and Sanders, 1978; Fyfe, 1982). Thus
Mineral formations 51

enormous amounts of material are transported both mechanically and in


solution even today
Present brooks, streams and rivers get most of their sedimented mineral
material as well as organic material from the older deposits over which
they run and which they are constantly abrading (Wennervirta, 1968). In
the upper courses of streams material from barren rock may be swept along
either as grains or as dissolved (e.g., uranium is easily leached, Salo and
Voipio, 1973). The dissolved material mainly originates from ground water
seeping into brooks from their banks. Waters from deeper parts of joints
in bedrock may have very high concentrations of different salts and gases
(Nurmi et al., 1988).
Where stream flow is rapid, only coarse mineral material settles to the
bottom. Grains may be coated with a thin film of iron and manganese
hydrous oxides, sometimes containing organic complexes in colloidal form
or clay. Such films readily absorb heavy metals from water, the finer the
material the more abundantly (finer fractions contain more surface area
per unit weight than the coarser fractions). Pebble coatings tend to have
a great anomaly/background contrast and exhibit a longer dispersion train
than does the fine mineral sediment itself (Hale et al., 1984). Trace metals
are especially enriched in fine "active" organic clayey detritus, which has
settled in leeward places, e.g., around boulders in the stream bottom or in
pools where the flow velocity is small.
In flooding rivers, bottom sediment materials may be eroded and resed-
imentated on banks as a mixture of organic and mineral material. These
so-called overbanks often exhibit graded bedding where a layer from a
specific year can be identified if necessary. Overbank sediments have
successfully been assessed for regional geochemical mapping, e.g. in Nor-
way. The use of these layers for geochemical prospecting is nevertheless
hampered by their inhomogeneity
Lower in the river course the whole valley bottom may be filled with
this kind of mixed alluvium, continuously moved and resettled by the
meandering flow At the mouth of the river the water velocity abruptly
minimizes and the suspended material settles into a delta formation. The
layers of a delta exhibit the same structure as flood deposits.
In clay most of the base metals are absorbed on or into phyllosilicates
or by various colloids; in sand they are mostly bound in surface coating
of separate mineral grains. For instance in soil at pH below 7 adsorption-
desorption reactions and at pH above 7 precipitation-dissolution reactions
determine the fixation of Zn (Brummer et al., 1983).
Geochemical anomalies in fluvial sediments are formed both by clastic
transport of ore mineral grains and by precipitation of dissolved metals
from the water. Ground water seeping down or laterally, or rising by
capillarity in the sediment transports ions, which may be fixed on the
surfaces of mineral grains in oxidizing conditions. The differences in the
52 Nonglacial overburden

TABLE 3-1
The effect of grain size on the geochemical properties of sediment, Bouma (1987)

Sediment Cation exchange capacity, Field capacity Permeability


dry mat. (meq/100 g) (cm) when satured (cm/d)
Fine (clay) 30 20 5
Medium (silt, loam) 20 16 50
Coarse (sand) 5 6 200
Humus 200

fixing of heavy metals depend largely on grain size of the sediment but also
on its mineralogy and humus content of the fluvial sediments (Table 3-1).
Adsorption capacity increases in the order CaC0 3 (0.44), bentonite (44),
humic acid (842), amorphous Fe- and Al-oxides (1,190-1,300) and M n 0 2
(1,540) in /miole/g in CaC0 3 system (Brummer et al., 1983).
Fluvial deposits have been transported along at least two paths, first
by gravity or ice and then by water, and this makes it very difficult if
not impossible to interpret their provenance. The same can be said about
possible clast anomalies (placers) in them, which have later still been
reformed by solution and precipitation by ground water. Knowledge of the
mode of occurrence of the ore elements in the anomaly is of great help in
the difficult interpretation. Geochemical character of certain transported
minerals like garnet (Hyvärinen, 1969) or magnetite (Granath, 1983) may
guide the prospector to the source deposit.

Littoral and manne deposits


Deltas are formed at the mouths of rivers, in lakes and seas having a
more or less stable water level. The material is subsequently spread by
coastal currents and waves along the beaches, the coarser grades to the
proximity of the shore and the finer fractions to the more distal and deep
zones of the lake basin or sea. A delta-like deposit just over the sea water
level is called a sandur.
Sand banks parallel to the sea shore, built up by tidal and wave action,
form a wall for a semi-closed basin where silty material rapidly sediments.
The basin is gradually closed, the mineral bottom sediments gradually
become finer and much more organic material (humus, diatomacae) co-
settles in the standing water (Fig. 3-5). Likewise in eutrophic lakes the
bottom sediment contains large amounts of organic material, forming mud
(in Swedish gyttja).
In shore deposits the classification of grains is very sharp and different
populations are easily discriminated from the mix (sand, etc.) (Friedman
and Sanders, 1978, p. 72). As a result of the different densities of the
monomineralic grains, sorting according to weight leads to bi- or multi-
Mineral formations 53

Fig. 3-5. Formation of sulphur-rich black mud on the bottom of a partly closed basin. Resembles
the Baltic Sea and some of its haffes (semiclosed estuaries). Original figure by Ström (1937),
modified from Friedman and Sanders (1978).

modality of grain composition. Repeated classification of sand material on


beaches by waves results in stratification where grains of different minerals
are separated from each other (Friedrich, 1974). Very common are the
black streaks of magnetite and red streaks of garnet grains. In coastal areas
with rocks containing cassiterite, chromite and other heavy and physically
resistant minerals, placer deposits of these minerals are likely.
With the complex mode of transport the anomalies in sand and silt may
be either clastic or hydromorphic in origin, and the mother lode may for
that reason be difficult to find. The trace metal spectrum of a mineral
like magnetite in beach sand, like that in till, may indicate from which
rock massive it originates (Granath, 1983). The mineralogical composition,
especially certain typical minerals, may reveal the provenance of the
sediment in question.(e.g., Friedman and Sanders, 1978). The case will be
more complex if the "fingerprint" minerals derive from different sources.
For identification of the source and for discrimination of superimposed
geochemical anomalies, Kinnunen (1979) has successfully applied a fluid
inclusion study of quartz.
The above described marine deposits are exposed by land upheaval.
A wide zone (30-300 km) of former sea bottom covered by clay has
been exposed on the west coast of Finland. Here the deeper clay layers,
which were formed during the melting of the continental ice, are laminar
and silty, while the upper postglacial layers are homogenous and contain
organic material. The uppermost layer, which sedimented during the warm
Littorina period, consists of muddy clay high in sulphur. In these low-lying
areas, ground-water rising by capillarity brings iron and sulphur to the
surface, causing either red limonite or white aluna precipitations to form
in the bottom of ditches. There have been very few successful uses of
these deep standing water sediments in geochemical exploration (Smee,
1983; Lalonde and Beaumier, 1984). In general, littoral and marine deposits
cannot be considered suitable materials for geochemical exploration.
54 Nonglacial overburden

Aeolian deposits
Abrasive wind action is strongest on vegetation-free beaches, mountains,
deserts, and in the proximity of glaciers. Wind erosion is continued until
the finer material is blown off and only the larger stones — typically worn
and polished "dreikanter" — are left in place, where they form a protective
pavement against further erosion. The blown material is sedimented in the
lee of the wind, in pits and grooves and behind trees, boulders, hills and
dunes.
The dunes, most eye-catching of the aeolian deposits, are found close to
present or ancient coastlines, in deserts like the Sahara and in areas of
ancient glaciation. Dunes are variously named according to their size, form
and geographical location but all consist of fine sand. Dunes eventually
become chained to the site by surface vegetation, but while bare they shift
slowly in the direction of the prevailing winds (Fig. 3-6).
The finer material is transported further to form the loess. One might
expect that such a soft transporting medium as air would effect a sharp
separation of size fractions. This is not the case, however; dust called loess
typically contains a wide range of grain sizes and mineral compositions
deposited together (Friedman and Sanders, 1978). The mineral particles of
loess are angular silt-size grains of quartz, feldspar, mica and calcite typi-
cally coated with iron oxyhydrate. The common concretions are cemented
with CaC0 3 precipitated from bicarbonate-rich capillary or hygroscopic
water.
The dusty material of aeolian deposits has been clastically transported.
Although in theory it should thus be possible to discover the mother lode
of ore mineral grains, in practice it is exceedingly difficult to trace back the
long route of transport, even where prevailing wind directions are known.
Occasionally, rare earth element spectrum analysis has successfully been
used for tracing the region of provenance of loess, e.g. in China. The

ORIGINAL SURFACE

WIND —► β£0οο DUNE


o°oo°„° o
SY*K° If Pf***
• o ° °ft^"üo6 ' · ον· V o · oo·
v
o.· ° ···· ·
Fig. 3-6. Erosion, transport, sorting and deposition of drift by wind close to beaches. Also
encountered close to ancient coasts and identifiable as vegetation-covered dunes, polished stones
(dreikanter) and surfaces varnished with pebbles.
Mineral formations 55

material of the "warm" loess deposits of western China is now known to


have come from rocks in the Gobi desert and the material of the "cold"
loess deposits of the black soil region in southern Soviet Russia has been
traced back to tills at the margins of ancient Fennoscandian glaciation.
The capillary movement of water responsible for hydromorphic anomalies
in loess depends on the porosity and size of voids in the material. Capillary
rise is high but very slow in clay because the permeability is almost nil,
and is quick but low in sand where the voids are large. Capillarity in dune
sand is so weak that no anomalies of hydromorphic origin are possible at
the surface.
Evidently in the silty loess deposits of White Russia the capillarity is
capable of carrying heavy metals through thick formations. There, chemical
anomalies in loess have been shown to reflect either mineralization beneath
or, more often, pollution from nearby industry, settlement or farming.
Lukashev (1983) has used artificial sorbents, buried for a certain time in
soil, in a successful study of pollution as well as the geochemical migration
of elements.

Chemical deposits
Chemical deposits can be classified into oxidates, redusates, precipitates
and evaporates according to the reaction leading to their sedimentation
from water solution (Rankama and Sahama, 1950). Water percolating in soil
brings elements into the soil; some of them become fixated in humus and
in other superficial layers, while others go deeper, together with elements
washed out from upper layers (Soveri, 1985).
Most typical of the recent chemical sediments are oxidates precipitated
from metal-rich solutions. Deeper down in water-saturated overburden,
reducing conditions prevail and the iron there is in readily soluble ferrous
state. Seeping into ditches or lakes, this ferrous ion is oxidized into ferric
state and settles out as water-rich red limonite gel. Later, when the water
content diminishes, the limonite hardens as a goethite crust on the bottom
(Blain and Andrew, 1977).
Limonite material transported by ground water to the surface from a
mineralization beneath cements the mineral grains of drift or of in situ
weathered material into deposits of gossan (Wilhelm et al., 1979; Wilhelm
and Kosakevitch, 1979). Through measurement of the isotopic composition
of lead it can be decided whether or not the parent rock of the gossan
represents an economic type of deposit (Vaasjoki and Gulson, 1985, see
Table 3-II). If the isotopic ratios in the gossan coincide with those of known
economic ore deposits in the region, follow-up studies, including diamond
drilling, can be recommended. The same isotopic method can be used for
identification of the source of lead-bearing erratics, soil and ground water.
A similar dissolution/precipitation process occurs in podsol profile: heavy
metals are leached from the A-horizon and precipitated to the B-horizon
56 Nonglacial overburden

TABLE 3-Π
Isotopic composition of lead of ore mineralization and gossan at Lady Loretta, Australia
(Vaasjoki and Gulson, 1985)

Ratio Mineralization Gossan


208pb/206pb 2.220- 2.206 2.2145 ± 0.0030
207Pb/206pb 0.957-0.950 0.9540 ± 0.0011
206 Pb/ 204 Pb 1 6 1 6 _ 1 6 30 1 6 220 ± 0.027

where pores become filled with rust. Often rust forms a compact layer (iron
pan) at the surface of the ground-water table, too.
Sizeable deposits of nickel-rich, manganiferous rust nodules are formed
on the sea bottom at levels where redox conditions are favourable. Rust
from lake bottoms was once exploited as iron ore in central Finland. Some
of the nodules may contain economic contents of other base metals besides
iron. Sulphur from sulphide rich mud rises by capillary action up to the
surface and white aluminium sulphate is precipitated.
Reducing conditions encourage the formation of sediments rich in iron,
sulphur and carbon. Such redusates form in closed ponds or at the bottom
of depressions in the sea (e.g., the Gotland Deep in the Baltic Sea). The
postglacial muddy clays contain large amounts of black iron sulphide, which
is easily oxidized to sulphate if exposed, e.g., by ditching, and the surface
waters become acidified.
Uranium minerals are easily dissolved in oxidizing environments such
as found on hillsides. The uranium is then carried by ground water to
valley bottoms where deeper down in the overburden the conditions are
reducing and uranium is fixed to form anomalies (Väänänen, 1976). During
transport the uranium will usually be separated from its daughter elements,
so that in individual samples the correlation is poor between gamma- or
alpha-intensity and the uranium concentrations.
The term precipitate is used for minerals such as Si0 2 and CaCOa sinter
(chert and travertine) precipitating on the Earth's surface or in the pores
of overburden from warm water springs. Also siderite (FeC0 3 ), vivianite
(Fe3(P0 4 )2-8H 2 0) and dopplerite (humus) layers in bogs belong to this
category, and the silt concretions in loess and glacial clays cemented with
calcium carbonate or marcasite (FeS2).
Evaporites are formed when an oversaturated solution of salts (brine) is
evaporated and Ca, Na, K and Mg sulphates, chlorides or carbonates are
left behind. Some brines, and the corresponding evaporites, are rich enough
in heavy metals to be exploited as ores.
The various chemical sediments themselves can be said to be hydromor-
phic anomalies. Although the volume and areal distribution of chemical
sediments are insignificant compared with other sediments, their impor-
tance for geochemical exploration is considerable.
Mineral formations 57

The percolating rainwater leaches minerals of the soil and carries the
dissolved elements deeper into the overburden along the ground-water
stream, towards valleys. When such happens in a material bearing a clastic
anomaly, superimposed geochemical anomalies will be formed. In certain
conditions the clastic and hydromorphic transports may even occur in
opposite directions (Paskolahti, Fig. 3-7).

LEAD IN TILL
0 . 0 6 mm 7.25 M N H 0 3
PASKOLAHTI BAY
KIIHTELYSVAARA

S a m p l i n g point

Pb in t i l l , ppm

^ Boundary of g l a c i -
morphic Pb anomaly

Height above sea


level

200

Fig. 3-7. Dispersion of lead (1) clastically by glacial transport uphill towards SE and (2)
hydromorphically in ground water solution downhill back to the shore and over the suboutcrop
of the Pb mineralization. Reproduced with permission from Mäkinen and Lestinen (1990).
58 Nonglacial overburden

Another type of mixed anomaly is formed when a clastic anomaly of


resistant minerals becomes "polluted" by hydromorphic transport of heavy
metals from another source. This occurs, for example, when uranium is
transported in solution to valleys transverse to ancient glacial transport,
or when lead is lifted up to the A-horizon by capillary water. Lead very
easily adopts the insoluble forms PbO and PbSÜ4 if oxygen and sulphur are
available.
In cases where clastic transport of the constituents of an anomaly can
be traced, this should be done first since it is easier and more reliable.
The more difficult hunt for the source of the hydromorphic portion of
the anomaly can then be undertaken. In most cases the source of the
hydromorphically and clastically transported material will nevertheless be
the same.

ORGANIC FORMATIONS

Climate, topography, bedrock and type of overburden control the de-


velopment of organic deposits. Temperature and moisture, erosion and
infiltration of water and fertility of the soil determine what plants will
thrive and how they will decay and form deposits. Though most organic
deposits remain where they have been formed, some of them have been
transported by water and perhaps mixed together with mineral material to
form gyttja (mud).
Almost all organic deposits of glaciated terrain — humus, peat and gyttja
— are postglacial in age. In rare instances, nevertheless, thin compressed
beds of organic deposits are found between tills (Korpela, 1969; Hirvas and
Nenonen, 1987) and can prove useful for dating and correlation of the tills.
In nonglaciated terrain organic deposits of any age may be found both at
the surface and buried under mineral sediments.
Although the use of organic materials in geochemical prospecting will
be properly treated in the volume of the Handbook on Biogeochemistry
in Geochemical Exploration, organic formations are briefly touched upon
here as well because of the intimate connections between mineral and
organic deposits, and especially their geochemistries. The laws regulating
heavy metal distribution in organic residue are almost the same as in
the mineral overburden. Organic colloids in ground water penetrating
a mineral formation surrender a part of their cations to the mineral
formation, affecting the geochemical behaviour in that way. Likewise
ground water seeping from a mineral formation into an organic deposit
carries a chemical signal from one to the other.
Living organisms have established themselves almost everywhere — high
on mountains, deep in ocean basins, in the tropics and on arctic ice. Some
parts of organisms are unfit for food and also resist weathering better
Organic formations 59

than others: for example the stems of coral, mussel shells, the skeletons of
diatoms, plant pollen, insect shards and the bones of animals. Soft parts
of the tissue will be preserved only in favourable reducing conditions such
as found in peat bogs. For deposits to form it is necessary that the growth
of tissue be abundant and that the tissue not be totally decayed after the
death of the plant or animal consumed.

Plant residue
Humus
Lichen, grass, leaves of trees, chitinozoan parts of insects and other
remnants of living organisms are constantly accumulating on the surface
of the overburden. Nevertheless organic material decays at such a rate
that the thickness of the topsoil remains almost constant through the
years. The slow oxidation and disintegration of the organic compounds
is called humification and the mature material humus. The nature and
constituents of humus depend on the vegetation, which in turn depends on
the conditions set by the climate and the overburden.
Organic material has been found to enrich heavy metals by binding them
in chelates and other complex salts of humic and fulvic acids (Tenhola,
1988). Humus concentrates different amounts of heavy metals according
to the type of forest (Fig. 3-8). Part of the heavy metal content of humus,
maybe that part which is bound to fulvic acids, is in dynamic equilibrium
with the surroundings (Fig. 3-9); metals enriched during the dry season are
leached off by rains but perhaps fixed in the B-horizon below.

B bog peat
C conifer humus
D deciduous tree humus

Fig. 3-8. Distribution of bulk contents of extractable heavy metals determined by dithizone
titration in different types of forest humus. Reproduced from Kauranne (1967).
60 Nonglacial overburden

ppm Ni

200
m,ca mafic intrusion
(V\YI | x x j
(mineralized)
I \ \ \\ gneiss
150 f
| Δ Δ | till W$M sand

| O Q | boulder layer

Sampling

Fig. 3-9. Seasonal variation in nickel content in humus over mineralized bedrock and country
rock at Enonkoski Ni deposit. The sand interlayer may cut off the capillary rise. Redrawn with
permission from Salminen and Kokkola (1984).

Humus reflects rather well the mineralization of bedrock as well as the


heavy metal anomalies in the overburden (Govett, 1973; Nuutilainen and
Peuraniemi, 1977; Äyräs, 1979; Toverud, 1984). Ores even at a depth of
over 200 m can be detected by analyzing the heavy metal content bound
in humates (Fig. 3-10, Antropova, 1975 in Lukashev, 1983). The anomalies
in humus and in the drift beneath do not necessarily lie directly above one
another (Figs. 3-11 and 3-12).
There has been much discussion about the mode of origin of the
anomalies in humus (Fig. 3-13). Perhaps the first explanation was given by
Goldschmidt (1937) who proposed that the high heavy metal contents were
Organic formations 61

06 FULVIC ACIDS

0 _i I i_ , I i ■ I L_. I . , I . i

04 I
0.8 %
03
HUMIC ACIDS
02

01

04-1
0.8 %
03-

HUMATES
02

01

1 1 i I 1 1 1 1 .1
3i HCL e x t r a c t

0 i 1 i i 1 , II 1 1 1111 1i1 i 1 i 1 i
100τ

TOTAL Zn in PEAT

50

rfrf77777* τζππτπ.
k AT "7 T X X
^ "

PEAT SAND HUMUS ROCK


3SCE
FAULT ORE
[m]

Fig. 3-10. The amount of zinc occurring in different modes in peat, modified after Antropova
(1975) and Lukashev (1983). The greatest contrast between anomaly and background is found
in humates. Antropova recommends the leaching out and analysis only of humates in humus
geochemistry.

the result of metal uptake by tree roots and concentration into the falling
leaves (see also Vogt and Bergh, 1943). There are great variations in the
heavy metal contents between different parts of the plant and still greater
between different species. It has been noticed, for example that some plants
(e.g., "copper plant", Viscaria alpina) tolerate much higher concentrations
62 Nonglacial overburden

Till 0 - 0 . 0 6 mm Humus C x H m
Cg§\ magn.anom.
i:':'y':'i 1 9 0 - 2 3 5 ppm 4 - 5 ml
^ ^ Ni ore exp.
fT:--:-::l
Ιϊ<<<>>^ > 2 3 5 ppm > 6 ml Ni boulder

Fig. 3-11. Anomalous contents of bulk cxHm in humus correlate positively with nickel anomaly
in the fines of till, but do not lie directly above the anomaly in till. Reproduced Kauranne
(1976b).

of toxic elements than others, whereas other plants even have a repelling
mechanism towards toxic elements (Figs. 3-14 and 3-15). As an example of
selective intake (Duchaufour, 1982, p. 93) notes needles of conifers growing
on dune sand which displayed an Al: Fe ratio of 8-10, while the ratio in the
sand underneath was less than 1.
Organic formations 63

Skarn with
scheelite

Scheelite W content in heavy


grains in concentrate of till
sample [PPm] W in humus [ppm]

•50 > 1400 > 6400 >5 >20

Fig. 3-12. A method lying between boulder tracing and geochemical prospecting is the tracing of
heavy minerals. The method is suitable for tungsten exploration, for example, because scheelite
and wolframite are easily concentrated by panning and amounts can be semi-quantitively
determined by the unaided eye under ultraviolet light. The anomaly in humus shows the
difference in mode of origin — glacigenic and hydromorphic transport. Reproduced with
permission from Toverud (1984).

Drake (1983) has advanced a theory of diapiric capillary rise of metallif-


erous ground-water solutions up to the surface (Fig. 3-16). In turn Cazalet
(1973) and Malmqvist and Kristiansson (1983) have argued that emanating
gases carry heavy metal ions up to humus. An upward flow of cations by the
self-potential force surrounding ores was proposed by Govett (1973), backed
up by Bölviken and Logn (1975) and confirmed in laboratory experiments
bySmee(1983).
Despite drawbacks such as the inhomogeneity of the material and
the seasonal variation of the heavy metal contents, humus is a valuable
prospecting material. It is easily sampled and, if ashed before analysis,
64 Nonglacial overburden

' / / DIFFUSION \ \ \
CLAV
♦ */ ^ \ Mi
\ΤΤΤ TILL
Δ Λ "^" //// Δ

~Γ X WEATHERE
SUBEXPOSURE \ \

Fig. 3-13. Processes causing upward dispersion of heavy metals from the suboutcrop of
mineralization into humus.

metal concentrations will be high enough to allow analysis by AAS,


colorimetry or other simple methods (Table 3-III). Even bulk analysis of
the cold extractable heavy metal with dithizone directly at the sampling
site is applicable.

TABLE 3-III
Mean concentrations of zinc and nickel in rock, soil (ppm) and in plants (ppm in ash) growing
on them (Lounamaa, 1967)

Bedrock: Siliceous Ultrabasic Calcareous


Element: Zn Ni Zn Ni Zn Ni
Conifers:
needles 1500 34 950 1200 860 16
twigs 1600 38 1200 620 1200 21
Deciduous trees:
leaves 2100 43 1200 1200 1000 18
twigs 4100 42 2700 1100 1700 15
Sou 320 91 270 1200 220 60
Rock 180 38 190 1500 230 64
Organic formations 65

1000

100

10000

Element concentration in soil [clarke]

Fig. 3-14. Relation between element contents of plants and soils. Plants growing above
mineralized bedrock contain anomalous amounts of the bedrock elements. The elevated
concentrations of metals in the substratum are reflected as elevated concentrations in plants,
in some without obvious limitations, in some only to a certain maximum. Kovalevskii (1984)
classified plants into four groups according to their ability to restrict metal uptake. Some plants
die if the concentrations of toxic metals increase above a certain limit. Others are not affected,
or even flourish with an extra supply. Viscaria alpina, for example, tolerates copper and is used
as a Cu-indicator plant. Modified after Kovalevskii (1984).

C o n c e n t r a t i o n in solution

Fig. 3-15. Plant metabolism depends on certain elements and suffers when others are present.
Kabata-Pendias and Pendias (1984) observed that plants selectively uptake elements from water
solution. Boron and also zinc in small amounts are good for plants, whereas lead and cadmium
are poisonous. Reproduced with permission from Kabata-Pendias and Pendias (1984).

Peat deposits
In regions of arctic climate, flat moist areas are widely covered by the
peat-forming plants of moss, sedge and horsetail. In peat all pores are water-
filled, no air is circulating and oxidation is not possible; the environment
66 Nonglacial overburden

diapiric ribbon-type
anomaly anomaly plume

Ore Country rock


Fig. 3-16. In glacial terrain the anomalies in drift are mainly clastic, even though easily
weathered ore minerals develop around themselves a hydromorphic halo, which may be
transported further by ground-water flow or diapirically upwards by capillarity to form
anomalies in humus. Modified after Drake (1983).

favours very slow decomposition of tissues instead. Boggy flat areas called
muskeg by North American Indians and jänkä by the Lapps cover extensive
regions of the Arctic. There are countless types of peaty formations — bog,
fen, marsh, mire, moor, quagmire, slough, swamp, etc.; their differences are
not discussed here as the geochemical behavior is affected in only a minor
way by the type of plants forming the peat. Geologically a peat deposit is
called a bog when the thickness of the humified peat layer exceeds 50 cm.
In the tropics the peat layer may be as much as 40 to 60 m in thickness, in
temperate region it rarely exceeds 10 m and in the arctic 1 m.
As determined by 14C and palynological studies the Finnish raised bogs
were growing at a rate of 0.5 mm a year during the warm period about
6,000 years ago, and have grown some less a year since then. The same
is true of Canada where growth rates of 0.7 mm to 0.3 mm have been
reported (DiLabio and Coker, 1982; Ovenden, 1989). Age determinations of
peat bogs have also been established with the uranium-thorium and amino
acid methods.
Three types of bogs can be differentiated on the basis of their water
balance and corresponding vegetation: (1) raised bogs, in plain areas with
only rainwater supply; (2) slope bogs, on sloping ground or beach with
down-seeping ground water; and (3) floating bogs, which are ponds, bays
or rivers overgrown by moss with continuous water supply from beneath.
Raised bogs, which are the thickest, are built up primarily of Sphagnum
moss. Sedge (Carex) is the usual peat-forming plant in floating bogs, which
also may be of considerable thickness, especially if the bottom mud is
included. Slope bogs tend to be rather thin and have a varying vegetation
of moss, grass, bush and small sized pine or spruce trees.
The degree of humification and the amounts of humic and fulvic com-
pounds depend somewhat on the original vegetation, i.e. on the type of
bog. Geochemically, mature peats behave about the same whatever their
original plant composition and fibrous structure. Ore metal anomalies tend
Organic formations 67

to be distributed laterally in peat layers, although some vertical diffusion


occurs as well.
In raised bogs, most of the metal content of peat has been deposited
from rain and wind, which makes them of greater use for air pollution
studies than for geochemical exploration. Thin ash layers originating in the
volcanic eruptions of Mount Hekla and Mount St. Helens, and radioactive
fallout from Chernobyl, have been traced in the surface of raised bogs in
Finland.
Bogs on slopes are minerotrophic, i.e. they rest on ground water and from
there get the main part of their heavy metals. Organic material easily binds
heavy metals, and anomalous concentrations are typically found at the rims
and bottom of such bogs. Sampling for regional geochemical exploration is
easily and effectively done from that side of the bog where ground water is
seeping into the peat. Eriksson and Eriksson (1976) found that anomalies
at bog rims can frequently be traced in the follow-up stage to mineralized
boulders and further to the mother lode (Fig. 3-17). Lateral diffusion in
peat is clearly easier than vertical diffusion and in many cases an anomaly
can be followed right across a bog (Fig. 3-18,; see also Armands, 1967,
p. 141).
Some metals are typically enriched at the top and some at the bottom
of the peat layer. Salmi (1967) in one of his classical peat geochemical
studies found, for example, that iron tends to concentrate at the surface
and titanium at the bottom (Fig. 3-19); lead occurs in anomalous amounts
close to the surface, nickel and copper close to the bottom (Fig. 3-20). The
total amounts of heavy metals in peat naturally depend on the bedrock
beneath (Fig. 3-21, see also Salmi, 1967, p. 115). Similar observations have
been made in bogs of permafrost area, where the concentrations of Cu, Mo,
Fe and especially U increase with depth, Mn is a notable exception (DiLabio
andCoker, 1982).
Floating bogs grow downwards toward the bottom of the basin, gradually
filling the water volume beneath until only narrow paths of running water
remain. Heavy metals moving in that solution are trapped by the living
plants or by the organic complexes of humified peat. There is a dynamic
balance between the concentrations in water and in peat varying with the
season.
There is one more type of peat very important for geochemical ex-
ploration: the pieces, fibres and colloidal gels transported by water and
accumulated in places where stream flow is negligible. These deposits are
very active in fixing heavy metals into mineral bottom sediments. Under
favourable conditions this accumulation peat intermixed with mineral ma-
terial will build up into thick deposits of mud. In lakes and bottoms of
floating bogs of temperate regions the thickness of mud layers sometimes
exceeds 10 m, though the peat layer proper may be only e.g., 4 m. The
average heavy metal concentrations in peat and humus in mineralized
68 Nonglacial overburden

Fig. 3-17. Distribution of metals and ore float in till and peat. Ground water seeps into a peat
bog from the mineral drift on hillsides, transporting heavy metals which are fixed in peat
as organometallic complexes. Easily sampled peat from the bog rim can be used in regional
prospecting. Redrawn with permission from Eriksson and Eriksson (1976).

exploration target areas vary considerably. Table 3-IV gives an idea of the
enriching factors of different materials.
The heavy metal binding capacity of organic material is so great that
the detection of anomalies is possible even with simple chemical meth-
ods. A portion of the heavy metal content in peat is loosely bound
and in dynamic equilibrium with water solution, in which the concen-
Organic formations 69

Fig. 3-18. Distribution of Mo in till and peat. Heavy metal transport in peat occurs both laterally
and vertically as shown by a molybdenum anomaly in peat over a glacigenic molybdenum
anomaly in till. Compiled after Smith and Gallagher (1975).

TABLE 3-IV
Average and maximum concentrations (ppm) of Mo, Cu, Zn and Co in humus, peat, till and
bedrock at Aittojärvi (Kokkola and Penttilä, 1976)

Mo Cu Zn Co

Humus:
mean 14 97 335 22
maximum 613 338 1520 263
Peat:
mean 67 79 153 35
maximum 1732 347 1103 579
Till:
mean 9 36 41 n.a.
maximum 78 128 104 n.a.
Bedrock:
mean 54 47 52 20
maximum 2382 346 1230 71

trations vary according to the season. With methods of selective disso-


lution any significant anomalies can be discriminated from the natural
variation.
70 Nonglacial overburden

O Fe,Ti,V-ore suboutcrops

Fig. 3-19. Differential dispersion of metals in peat. Heavy metals in peat behave like ions in
a Chromatographie column — some are concentrated near the top, some near the bottom. An
example from titaniferous iron ore. Redrawn with permission of Salmi (1967).

Diatomaceous earth
Muddy deposits (gyttja) often contain large amounts of diatom frustules.
Though areally and by volume small, these deposits of diatomaceous earth
are geochemically important because of their pure siliceous composition
but unusually great capacity for absorption of e.g., heavy metals.

Animal residue
Shell deposits
Clam, mussel and mollusc shell layers are virtually the only fossiliferous
(animal) formations forming in arctic and temperate climate. Commonly
they occur as pocket-like deposits in ancient basins in areas of land
Organic formations 71

V NiCrCu
0.1

a
Φ
Ό

0.5

1.0
ppm ppm
125 175 Ni V 150 250 350 Zn
150 250 Cu Cr 30 50 70 Pb
60 95 Co

Fig. 3-20. A statistical study of heavy metals in peat throughout Finnish Lapland shows how
most metals are concentrated near the bottom whereas lead is concentrated at the surface.
Redrawn with permission from Tanskanen (1976).

Cu
[ppm]

1000
OO O ·

#· · · · •
500
o ° # · · • •
oo •
o o° · · ·
ο<%^ q? °o · ·
1000 2000 Ni [ppm]

O peat over gabbro · pea t over peridotite

Fig. 3-21. Illustration of the strong relation of heavy metal contents between peat and underlying
bedrock lithology in Lapland. Redrawn with permission of Tanskanen (1976).

upheaval. Because the shells are composed almost totally of calcium


carbonate, the main composition of the deposits deviates sharply from the
surrounding siliceous formations.
The trace element spectra also deviate, because some molluscs can in
some degree select their heavy metal intake, and some are very resistant to
toxic elements. Mussels, in fact, are exceptional in enriching elements that
kill many other species and are finding an interesting use in monitoring
pollution.
Like trees and glacial clay, clams undergo seasonal growth and their
age and growth conditions can be studied from the shell laminae. Their
72 Nonglacial overburden

geological age can be determined from the radioactive carbon isotope ( C)


concentration and their habitat (e.g., salinity and temperature of water)
from the stable carbon isotope (12C, 13C) and oxygen isotope ( 16 0, 18 0)
composition. Light carbon isotope prevails in fresh water and light oxygen
isotope is enriched in cold climate. The same geochemical methods that are
used in studying the provenance and residence time of ground water in
bedrock can thus be applied for the study of shell formations.

Human impact
Human cultures have affected the environment as long as they have
existed, changing the trace element contents of the soil.
Older polluted formations, like the lake sediments in central Sweden
polluted by mine smelters during the seventeenth to nineteenth centuries
(Vuorinen, et al, 1988), have in places been buried by younger sediments,
and could mislead the unwary geochemist working in such areas. The
present pollution by agriculture, settlements, traffic, industry and especially
ore smelting pervasively affects the topsoil and water sediments (Fig. 3-22).
A prospector assessing topsoil and water sediments must thus be constantly
alert to the possible effects of pollution on the results (Ek et a l , 1988; Table
3-V).
Buried bones form a relatively open system with the surrounding soil
so that part of the lead content of bones may have originated from the
surrounding soil or alternatively a part may have been leached away. The
probable source of the highly anomalous lead accumulated in e.g., mediaeval
man or soil was water pipes, glaze on earthenware pots, and lead added as
"spice" to wine (Table 3-VI).
Qvarfort (1977) reports that wind-blown dust from 400-year-old dumps
of exhausted lead mines polluted the soil only within a 350 m radius
and only down to the B-horizon (50 cm) in till. The pollution of lake
and stream sediments in the same area was much worse and extended
several kilometres along watercourses below the old waste heaps (Table
3-VII). In Finland such pollution, depending on the metal, extends as

TABLE 3-V

Average concentrations of lead in garden soil, house dust and in hair of children 2-4 years old,
in a village near by a closed lead mine in Derbyshire (Thornton, 1988)

Soil (ppm) Dust (ppm) Hair (ppm) Observation


420 530 8 at the mine village
3,390 1,560 11 at the mine village
13,960 2,580 29 at the mine village
265 560 - averages for Great Britain
5,600 1,870 - averages for Derbyshire
Organic formations 73

Fig. 3-22. Distribution of lead in lake sediments. Lead concentrations in lake sediments have
been found highest beside highways and in boat harbours. Improvements in motors and fuel
show a trend towards decreasing pollution. Reproduced with permission from Pirttisalo and
Räisänen (1988).

much as 3-10 km downstream along watercourses (L.-M. Kauranne, 1979;


Salminen, 1980), in Jinzu river, Japan even 40-50 km severely (Kobayashi,
1972). Beside the tailings pond at Outokumpu mine, polluted waters even

TABLE 3-VI

Relative exposure to lead in Britain (Waldron, 1988). Lead content of neolithic bones taken as
standard
Neolithic 1 Mediaeval 13
Iron Age 3.5 18-19th century 10
Roman-British 7.0 Present day 4
74 Nonglaeial overburden

TABLE 3-VII

Heavy metals (kg/year) leached by rain-water from waste heaps (Qvarfort, 1979)

Mine Cu Pb Zn Cd
Bersbo 5,000 55 30,000 40
Falun 15,000 400 433,000 300
Garpenberg 375 125 26,010 35
Saxberget 595 120 14,255 25

penetrated a 200-300 m wide sand esker (Kauranne, unpubl. report, 1949).


Heavy industrial pollution directly form chimneys is usually concentrated
fairly narrowly (Kobayashi, 1972) as well as pollution by traffic along roads.
In treeless tundra it is more widely distributed, however, and has killed
vegetation within a radius of tens of kilometres, as sulphur and nickel have
done in Petchenga, Kola (formerly Finnish Petsamo). Building of roads of
mining wastes distributes heavy metals more widely.
The soil buffering capacity varies (Nuotio et al., 1985) but with a heavy
continuous load it is broken and the insoluble noxious metals become mo-
bile. In such areas the use of geochemistry becomes seriously complicated.
Everybody applying geochemical prospecting methods in populated areas
and especially in abandoned mining areas must carefully select the methods
of sampling and analysis to avoid erroneous results.
The wide use of artificial fertilizers, which nowadays contain any number
of trace elements advantageous for plants and animals, is rapidly changing
the heavy metal spectrum of topsoil of cultivated fields and forests. The
effect does not extend deeply in nonpermeable formations (clay, till) but
may be marked in highly permeable formations (silt, sand). This promises
to make the use of the top layers of sorted sediments for geochemical
exploration still more difficult in future than today. Pollution can be
detected and measured by using selective extraction methods in analysis
(Sippola and Tares, 1978; Sontag, 1981) and geochemistry can be used even
in old mining areas (Steward and van Hees, 1982). New statistical methods
for the discrimination of natural and pollutional components and for the
interpretation of complex geochemical anomalies need to be developed.

CONCLUSIONS

What are the experiences obtained from the use of sorted mineral
sediments and organic deposits for geochemical exploration? Gravel is so
coarse grained that it is difficult to get a representative sample for analysis.
Sand is sorted both by grain size and mineralogically, and it consists
mainly of quartz and feldspars. It is highly permeable but its capillarity
Conclusions 75

Fig. 3-23. The distribution of the zinc concentrations in the Pihlajanneva mire shows both
the location of the zinc mineralization as well as that of the clastic zinc anomaly in till. With
permission from Virtanen (1991).
76 Nonglacial overburden

is low. Typical geochemical anomalies in sand are clastically formed placer


deposits of heavy minerals, and gold.
Silt and clay have also been mineralogically separated and shown
to contain large amounts of feldspar, mica and clay minerals. Their
capillarity is high but permeability low. They do not have clastically formed
geochemical anomalies although there is a certain variation in chemistry
between the layers of laminated silty clays. Anomalies in silt and clay are
hydromorphic.
Composite samples of various sorted mineral formations over a large
area give the same picture of the geochemical character over the area as the
average chemical composition of till samples from the same area (Kauranne
et al., 1977). The use of sorted mineral sediments for geochemical explo-
ration, with the exception of those of present watercourses, has proved to
be of questionable value (Coker and DiLabio, 1989).
Assaying of organic deposits, humus and especially peat, on the other
hand, has given good and easily interpretable results (Eriksson, 1976, see
also Fig. 3-23). In populated areas the increasing pollution of near-surface
materials hampers geochemical exploration based on topsoil, however.
Both chemical analysis and statistical methods have been successfully
used for the discrimination of natural and man-made anomalies. Through
selective sequential dissolution, the easily soluble portion of heavy metals
which contains most of the pollution component can be separated from
the natural component. Over 70% of the lead accumulated along highways
in snow, lichen and humus is in exchangeable form, soluble in dilute
ammonium acetate (Vuorinen, et al, 1988). By calculating element ratios it
may be possible to identify the portion of natural heavy metal distribution
populations in the total dispersion spectrum.
Assaying of the margins of peat bogs can be recommended for regional
geochemical surveys and the analysis of humus for localization of the
mother lode of anomalies. The sorted mineral sediments cannot be rec-
ommended for anything more than delineating broad regional geochemical
zones.
77

Chapter 4

SOIL TYPES

SOIL FORMATION

Soil in the narrow sense used here is the chemically differentiated upper
layer of the overburden, typically a black humus surface layer underlain by
a pale A-horizon deficient in Al, Fe and clay, a brown B-horizon enriched in
Al, Fe and Mn, and a C-horizon of unconsolidated parent mineral material.
Soil-forming processes influence the secondary dispersion of metals, shaping
anew the anomalies and making their interpretation more difficult.
Soils are the product of weathering — the physical, chemical and
biological forces which begin their disintegrating and discriminating work
immediately fresh mineral material is exposed. The longer the time for
development, the more extreme the temperature variations, the higher the
humidity, the more profuse the organic life, the greater the porosity of the
parent mineral formation and the weaker the minerals, the deeper will
be the weathering effect and the more distinct the soil horizons that are
developed.
The global zoning of vegetation follows both the latitude and the height
above sea level (Figs. 4-1 and 4-2). Local factors affecting the vegetation
include warmth and water and the time elapsed, but also the mineral
composition and looseness of drift, exposure to winds and surficial water
streams, and human activity.
In a warm dry climate the capillary rise of water is greater than the
downward seeping of seasonal rainwater, and alkali and other soluble salts
move up to the surface and are precipitated. When soil is alkaline, clays
with high base-exchange capacity, rich in montmorillonite, are formed.
Such clay minerals tend to capture and enrich heavy metals from the
ground-water solution.
By contrast, the humid climate of arctic and temperate latitudes favours
the formation of acid soils. Clays with low base-exchange capacity minerals
like kaolinite through illite are developed. Although they adsorb heavy
metals, their scavenging ability is weaker than that of the expanding clay
minerals. However, the organic colloids in the soil enrich heavy metals by
00

TUNDRA, . COOL CONIF- ΓΤΤΤΓ 1 TEMPERATE SAVANNAH,


NO FOREST ER0US FOREST 11111 J MIXED FOREST
lilllii DRY FORES!

WARM MOIST DESERT,


MOIST TROPICAL | ^ B ^ ^ ^ _ IKUKIUAL M U I » I
FOREST llllllllll
lllllllll MONSOON FOREST ^ ^ ^ ^ M EVERGREEN FOR. 1 1 NO FOREST

Fig. 4-1. Distribution of the main forest zones of the world. Vegetation reflects the temperature and humidity as well as grain size S
composition and nutritive state of the overburden, all important factors in soil formation and the secondary dispersion of elements. The Q-
type and thickness of the important organic top layer Ao of the overburden depend on the vegetation. ^
Soil formation 79

ZONES VEGETATION

Fig. 4-2. Relation between vegetation and altitude. The height above sea level has almost the
same effect on vegetation, and thereby on soil formations, as does latitude. Redrawn with
permission from Duchaufour (1982).

capturing them into insoluble chelates. The most common types of soil in
arctic and temperate regions are podsol and chernosem. Both exhibit a
distinct and prominent zoning if the parent material is permeable enough.
In places of impeded drainage, lack of aeration prevents the oxidation
required for normal soil formation and gley — a bluish-grey horizon with
ferrous iron — is produced instead. A fluctuating water table may cause
formation of rusty concretions or streaks in the gley Deeper down in porous
sediments, at the surface of the ground water, an almost impermeable hard
iron pan is sometimes precipitated.
In the high Arctic the permafrost hinders drainage and soil formation.
Close to glaciers the surface of the overburden is newly exposed and
immature and the soil profile is very thin. The same is true of high
mountains and in arid deserts where the lack of water delays or prevents
soil formation.
As soils mature a distinct layering tends to develop. From bottom to top
the main layers of the idealized soil profile (Fig. 4-3) are: the unaltered
C-horizon containing fresh mineral material; the B-horizon, brownish due
to precipitated ferric iron and containing oxides of Mn and Al; the ashy pale
A-horizon where an overall weathering of minerals and leaching of elements
other than Si has occurred; and uppermost, a brownish black A0-horizon of
humified organic material, which supports the living vegetation.
The upper parts of the B-horizon also contain some downward trans-
ported clay particles, kaolinite or montmorillonite, which through hydroxyl-
groups bind Fe leached from the more weathered A-surface. The B-horizon
can thus be divided into a lower Bs-horizon containing complex salts of
fulvic acids, a middle B h -horizon containing the complex salts of larger
molecular humic acids (Duchaufour, 1982, p. 318), and an uppermost
Bfhorizon rich in clay. The A-horizon can also be subdivided: into the
deeper mineralic A2-horizon, which forms a rather sharp boundary with
80 Soil types

A2
B,
Bh

Bs

Fig. 4-3. Schematic podsol profile. The most characteristic soil profile for arctic and temperate
zones is podsol.

the B-horizon, but changes gradually upwards, and the Ax-horizon, which
contains abundant organic material and is therefore dark or greyish in
colour.
The above-described idealized soil profile is called podsol and tends to be
acidic. The lowest pH is measured at the top and differences are significant
between the various horizons, which demonstrates that there is some
buffer capacity in soil layers in calcareous and even in granitic areas (see
e.g., Nuotio et al., 1985; Vuorinen and Lahermo, 1986). Acid rain easily
breaks it down, however, causing fixed ions like Al to go into solution,
with toxic effects on vegetation. Typical examples of pH and heavy metal
concentrations in different layers of the podsol profile are shown in Table
4-1, for till with a soil profile 60 cm thick and silt with a profile 5 cm thick.
The rate of humification, decaying of the plant tissue and producing
organic acids for the weathering of minerals and development of the soil

TABLE 4-1
pH and heavy metal concentrations (ppm, determined from a 25% HNO3 leach), in different
horizons of podsol profile in till and silt of the background area in Viitasaari, Finland (Kauranne,
1967)

pH Tül (<0.05 mm) Silt


Cu Zn Ni Cu Zn Pb Ni
Humus 4.4 _ _ _ _ _ _ _
A2 5.1 5.1 28.8 3.6 9.7 55.9 18.1 13.5
B 5.6 6.9 45.6 3.3 14.9 49.1 16.9 18.8
C 6.0 12.0 37.2 7.5 14.4 43.4 15.6 18.6
Soil formation 81

profile strongly depend on the type of vegetation. The nitrogen content in


plants varies. Plant tissues with low C/N ratio (<25) such as grass and
the leaves of deciduous trees decompose quickly, relative for example to the
needles of conifers and stems of heather (C/N > 60) (Duchaufour, 1982).
Climate has varied since the Pleistocene glaciation, and with it the
vegetation. Ten thousand years before present, tundra plants covered
Europe. Then until 9,500 BP, during the period of preboreal climate, pine
and birch were the dominant trees. During the boreal climate, which lasted
until 7,500 BI^ hazel became important and during the warmest period of
the atlantic climate, until 5,000 BI^ oak and other noble deciduous trees
flourished even up in Fennoscandia. Since then pine, spruce and birch have
dominated in the north, and beech and fir further south. Figure 4-4 shows
the development of forests in Finland as an example of the changes in
vegetation over time.
A similar succession of species has prevailed in the glaciated areas of
North and South America and Asia. In warmer climate deciduous trees
with deeper penetrating roots prevail, the production of organic litter
is greater, the humification faster and greater amounts of heavy metals
become enriched in humus.
The idealized soil profile described above is typical for most parts of arctic
and temperate regions. The thickness varies according to the permeability
of the overburden: the more porous the deposit the thicker is the soil
profile. In sand the thickness of the soil profile may be a couple of metres,
in till a few decimetres and in clay a matter of millimetres. The thinner
the profile the sharper are the horizontal boundaries. Development of the
profile is prevented if the voids are filled with water, as under a peat layer
where the ground is wet all the year. The order of horizons may be changed

SPRUCE

^r/ //////// u /// / n-m > , > > / 1 iALDER


r
-f} / / / / ' / / l-^r ELM
HAZEL

TTTTTTi
mm PINE

νΤΤΤΐΎΤΤΊ^ΤΆ BIRCH

Fig. 4-4. Relative distribution of genera in Finnish forests over the last 10,000 years. The
development of Finnish forests after melting of the continental ice, as studied by pollen analysis,
reflects the variations in climate, weathering and the composition of clay, mud and peat deposits.
Reproduced with permission from Alhonen (1988).
82 Soil types

if evaporation and the corresponding capillary rise exceed the amount of


downward infiltrating water; accumulation then occurs at the surface and
leaching is from below.

SOIL DIFFERENTIATION

Soil science, or pedology, has been developed mainly for the needs of
agriculture, as cultivation practices depend heavily on the type of soil.
Though factors affecting the development of soils may be few, their local
and seasonal variations lead to so many different soil varieties grading into
one another that, as yet, no single internationally accepted classification of
soils exists.
The weathering processes causing soil formation, varying in intensity
according to climate and other factors, include dissolution, hydration,
hydrolysis, oxidation, reduction and carbonation. Organic and mineral
material is added and removed, translocated and transformed, resulting in
vertical discrimination and horizontal layering. The processes are partly
reversible and a dynamic balance is established, depending on the changes
in climatic and chemical conditions (Figs. 4-5 and 4-6) leading to a type of
horizontal layering, the soil profile.
Geochemical exploration proceeds through chemical analysis of natural
materials in which the mineralizations of bedrock are reflected as anoma-
lously high contents. These anomalies are only discoverable, however, from
samples that are entirely comparable. Within one study, samples absolutely
must be taken from the same material — and in the case of soil from the
same soil horizon.
There is no need for the exploration geochemist to have a thorough
knowledge of all soil types — it is sufficient to know what happens to

Fig. 4-5. Variation of different properties of soil in the vertical (podsol) profile. Reproduced with
permission from Hawkes and Webb (1962).
Soil differentiation 83

TUNDRA PODSOL CHERNOSEM BOG

FeQ(OH) ■
s^f^^^^ä
Fe2 +

PERMA
Ca rich layer
FROST

TILL SAND SILT

dry humid semiarid humid

C O L D T E M P E R A T E

Fig. 4-6. Weathering and vegetation both modify and change the overburden, producing
different soil types. The commonest soil types in the arctic and temperate zones are podsol and
chernosem.

the elemental concentrations of the parent rock during the process of soil
formation, and what is the best horizon to assay if the virgin C-horizon is
not accessible or available.
Concentrations are nearest to the original in the C-horizon and anomalies
found there are easiest to interpret. There may of course be great variation
between the upper and lower parts of the C-horizon depending on the
type of formation and the level of ground-water table. The A-horizon is
depleted in most of the heavy metals; only lead tends to be enriched. The
B-horizon is enriched in most of the heavy metals, especially those that are
coprecipitated or adsorbed by sesquioxides, as are Co and Zn. Humus (A0)
is enriched in almost all the heavy metals, but especially in those like Sb,
Co, Ni, Cu and Cr which have a great affinity for organic compounds.
In the following section the properties of podsolic and chernosem soils
— the soils most common in arctic and temperate regions — are described
and a few other less common types are briefly characterized. In the main,
the classification is that proposed by Duchaufour (1982). Soil zones of the
world according to Simonson (Hawkes and Webb, 1962) are presented in
Fig. 4-7.

Soil types
Immature soils are encountered where climatic and biologic forces have
not operated long enough for soil development. In newly formed deposits
like sandurs or dunes, submerged areas recently exposed by land upheaval
and areas released from melting glaciers, vegetation has not had time to
OO

g>
Fig. 4-7. Distribution of the most common soil types after Simonson (1957) in Hawkes and Webb (1962).
Soil differentiation 85

grow, decay and alter the mineral material of the surface of the regolith.
In very arid or very cold environments vegetation has been too sparse
and weathering too slow for the development of a soil profile during the
Quaternary.
Erosion by water, wind and landslide in places cuts into older deposits,
exposing new surfaces and burying or dispersing the layers of the old soil
profile. The development of a new soil profile requires time.
Immature soils are typical in the Arctic and on high mountains (Figs. 4-1
and 4-2). According to the FAO/UNESCO soil map of the world, immature
soils occupy 12.50% of the total land area of the world (Kabata-Pendias and
Pendias, 1984). This map differs somewhat from that of Simonson (Hawkes
and Webb, 1962) in Fig. 4-7. In immature soils no secondary processes have
affected the clastically formed geochemical anomalies, also the formation of
hydromorphic anomalies is delayed, making these soils simple to sample,
analyze and interpret.
The aluminium- and iron-rich soils are sometimes collectively called
pedalfer and the calcium-rich ones pedocal. The most important of the
mature aluminium- and iron-rich soils of the arctic and temperate regions
is podsol. Cool boreal climate, but also a permeable surface and quartzose
parent material, favour its formation. Typically the forest of this zone is
coniferous or coniferous mixed with deciduous trees. Podsol is also the main
soil at high altitudes in alpine and subalpine climate and is locally very
common in humid atlantic climate. On limestone bedrock the development
of podsol is delayed by the buffering capacity of calcium carbonate.
Podsol consists of distinct layers, which in till from the top downwards
are as follows: Ao is a mor type of humus (mor is a poor and mull is a rich
type of humus), biologically it is not very active but releases abundantly
chemically active, water-soluble compounds when rainwater penetrates it
(Fig. 4-9). It varies in thickness between 1 and 10 cm. Ai is a grey horizon
of mixed organic and mineral matter 1 to 5 cm thick. A2 is the ashy
white layer that has given its name to this soil type. Consisting mainly
of fine and friable quartz grains and varying in thickness from 1 to 10
cm, it grades upwards gradually and downwards rather sharply, unlike the
A2-horizon in brunisol, another soil variety met with in temperate climate.
Further down, the Bh-layer is sometimes well developed as a black horizon
several centimetres thick with an aggregate structure rich in flocculated
amorphous humic salts. The material changes gradually downwards to the
Bs-horizon, which is a spodic accumulation 5 to 30 cm or more thick, often
rich in red-brown ferric iron, and sometimes forming a hard iron pan. In
its deeper parts there occur colourless or pale fulvic salts of aluminium
and other metals. In atlantic climate a clay-rich B t -horizon sometimes
accumulates below this, but more often clay is accumulated above the Bh-
layer. The boundary of the B-horizon towards the virgin C-horizon below
is transitional. The total thickness of the podsol profile varies according to
86 Soil types

HUMIC ACIDS

FULVIC A C I D S

ALTERED MINERALS

CLAY A C C U M U L A T I O N

—* OXIDE PRECIPITATION

Fig. 4-8. Soil formation involves the differentiation of elements in the surncial part of the
overburden. This leads to chemically different horizontal layers. Reproduced with permission
from Duchaufour (1982).

0.8
Mn

0.6 \
\ Mn
\
o
\
2+ \ Fe3 +
HI Fe 2 + \
0.2 \
\
\
5 6 7
* pH

Fig. 4-9. Effect of Eh and pH on ionic species of iron and manganese. Because of their amount
and strong colours the modes of occurrence of iron and manganese are the most conspicuous
in the soil profile. Reproduced with permission from Duchaufour (1982).

the permeability of the deposit, from a few millimetres in clay to a couple


of metres in sand.
The typical components and composition of the different layers of podsol
soil are shown in Fig. 4-8. The upper profile is very acidic (A0 pH 3-3.5
and Ai pH 4.5). According to Delecour et al. (1974), about 30-50% H + and
50-70% Al 3+ and thereby also Fe, Al and Mn are readily soluble (Fig. 4-9)
(Duchaufour, 1982).
The soil on densely vegetated steppes like in the Black Earth Zone of
Russia and in other loessic deposits along the outer front area of Pleistocene
glaciation is usually chemosem, where the thick black A-horizon lies directly
Soil differentiation 87

SPARSE FOREST DENSE FOREST

CaCOß

Fig. 4-10. There are variations in structure and chemical composition of all soil types, mainly
due to the lithology and porosity of the parent sediment. Here two types of chernosem are
presented.

above a streaky C-horizon rich in calcium carbonate. There may or may not
be a weakly developed clay-rich B-horizon (Fig. 4-10). The slight differences
in geochemistry of podsol and chernosem profiles are shown in Table 4-IL
The black cotton soil of tropical latitudes resembles chernosem. In more
arid areas where the soil is not so black and not so thick it is called
burosem and in semi-deserts its grey salty variety is known as sierosem.
In somewhat warmer and more humid prairie and pampa the soil has a
thinner A-horizon (due to repeated fires), a little better developed argillic
B-horizon and, unlike chernosem soils, white CaC0 3 precipitations are not
found in the C-horizon.
All the different horizons of the soil profiles mentioned have been
successfully assayed for geochemical exploration; a vast literature exists on

TABLE 4-II
Trace elements in organic material of soils (in ppm dry weight), according to Stepanova, 1976
(Kabata-Pendias and Pendias, 1984)

Sou Element Total Humic Fulvic


compounds compounds
Chernosem Cu 90 3.6 29.4
Zn 116 3.4 38.1
Mn 1110 trace 254.0
Mo 5 0.8 0.9
Podsol Cu 44 1.2 16.7
Zn 80 15.6 29.1
Mn 1830 44.0 267.0
Mo 3 0.2 0.5
88 Soil types

all of them. The concentrations of the common base metals are highest in
humus, Ao, on average, but vary seasonally (Salminen and Kokkola, 1984)
(see Fig. 3-9); they are lowest in Ai, but in B concentrations usually are
somewhat higher than in C (Kauranne, 1967).
The composition of humus varies with the type of soil; either small
molecular fulvic acids or large molecular humic acids prevail.
The calcimagnesian soils with only A- and C-horizons, most notably the
rendsinas, are typical for grasslands over calcareous bedrock. Under the
crumb structured "stony", humus-rich A-horizon is a pale calcium-rich
C-horizon. In areas with trees the B-horizon is developed as well, and the
soil is then described as brunified.
Bog is here considered to be an organic overburden deposit consisting of
peat. Bog could well be classified as a type of soil too. Peat bogs are formed
in perpetually wet places from slowly decaying plant rests, the deeper parts
of the deposit being more humified, the surface layer still fibrous like the
living plants (Fig. 4-11).
Brown soil with biologically active mull humus occurs in atlantic and
semi-continental climate. In theory, gravitational water and the water rising
by capillarity carry equal amounts of sesquioxides up and down so that
the A- and B-horizons are mixed. Brown soils have an ABtC-profile, where
the clay-rich B-horizon is light in colour because of its calcite content.
These soils are common in central Europe and other temperate areas with
deciduous forests. Luvisols are clay-rich brown soils.
Hydromorphic soils called gleys are found in lowlands where the water
table is high, the environment is reducing and iron is in ferro-form. In the
typical case there are in gleys only a brownish black Ai-horizon and bluish
grey sticky A2-horizon with possible rusty patches due to seasonal lowering
of the water table.
Saline soils, with sodium and calcium chloride or sulphate accumulations
due to evaporation of ground water, and black, iron sulphide-rich marine
mud deposits below, occur along marshy coasts, as on raised beaches of the
Baltic sea.

CONCLUSIONS

The most common soil types of arctic and temperate regions are podsol
and chernosem. In these, as well as in other soil types, horizontal lay-
ering has developed through leaching and precipitation; iron, aluminium,
calcium, manganese and other cations and organic complexes and clay min-
erals have removed from one depth layer and accumulated in another. The
physical as well as the chemical characteristics vary between the horizons.
The order and number of the horizons depend on moisture, direction of
water movement and the lithology of the overburden.
Conclusions 89

Fig. 4-11. Profile of a drained peatbog in Ireland, utilized as fuel peat. The top layer is raw,
while the deeper part is mature, strongly humified peat mass (photo L.K. Kauranne).

Under the topmost layer of plant litter there is the organic black
humus layer in which usually high concentration of heavy metals are met
with. Concentrations vary greatly according to the season. Humus covers
almost the entire area, it is easily sampled and therefore very much used
in geochemical research. Sometimes only the humic compound of it is
analyzed (Lukashev, 1983).
Underlying the dark humus layer is the ashy mineralic horizon Αχ, which
90 Soil types

is rich in organic material and has large ion exchange capacity. Lead tends
to concentrate here. The Ai-horizon is sensitive to pollution. Colorimetric
analysis of it is hampered by the presence of colloids. The seasonal variation
of heavy metal concentrations is strong, but sampling is easy. The pale A2
is not particularly rich in organic materials or heavy metals. Very often it is
impossible to discriminate the different A-horizons. Assessing of A-horizon
has also been popular in geochemistry.
In the rusty B-horizon many of the base metals are coprecipitated with
iron or aluminium or scavenged by clay minerals or organic complexes. Two
or three different B-horizons can be delineated in well-developed mature
soil. Anomaly/background ratios tend to be highest here. The enrichment
factor of the heavy metals varies widely. The B-horizon is easily sampled,

P b i n B n o r i z o n tj|1
• sampling point \ striae
A^ jllfllll 110-250 ppm
llllllllllllllll 250- ppm
0 500 m

Fig. 4-12. Distribution of lead in B- and C-horizons of till in Ireland. Compiled after Miller and
Cazalet (1979).
Conclusions 91

and therefore it has been perhaps the most popular sampling medium
in explorational geochemistry. Pollution sometimes reaches B-horizon and
colorimetric analysis is often hampered by colloids. As an example of the
assay of the B-horizon see Fig. 4-12, lead in B- and C-horizons of till in
Ireland (Miller and Cazalet, 1979).
The C-horizon is untouched by soil forming processes and constituents
are almost in their original state, especially below the ground-water table.
Elements like copper and nickel usually appear in highest concentrations
here. In water permeable formations the C-horizon is deep down, but in
less permeable formations it is usually accessible at a depth of 0.5-2.0 m.
It is very seldom polluted by human activity. It is widely used both in
geochemical mapping as well as in geochemical exploration.
Opinions about the suitability of different horizons of soil for geochemical
studies vary. All horizons reflect somehow the chemistry of the underlying
bedrock: "the vertical distribution in the overburden profile does not
appear to be significantly affected by soil processes" (Lahti, 1971, in
extreme, says for Canada). The topmost horizons Ai and A2 are better used
for environmental monitoring than for geochemical exploration, but the
Ao-horizon (humus) is better for exploration. The rusty B-horizon consists
of several sub horizons with varying enrichment factors, and is not ideal for
geochemical exploration either. The almost virgin C-horizon soil material,
especially if samples can be taken under the ground-water table, is strongly
recommended for geochemical exploration
Trace element concentrations gradually increase towards depth in C-
horizon if it consists of only one single bed. If there are several intercalated
beds above each other the case is more complex and must be studied more
closely before deciding which layer in the C-horizon of overburden should
be assayed.
93

Chapter 5

GEOCHEMICAL DISPERSION IN THE SECONDARY


ENVIRONMENT

INTRODUCTION

Geochemical methods of prospecting are based upon a systematic study


of the dispersion of chemical elements in the natural materials surrounding
an ore body. Traditionally (Hawkes, 1957; Hawkes and Webb, 1962; Rose et
al., 1979) this dispersion has been related to the geochemical cycle, which is
considered to consist of two parts: one embracing the deep-seated processes
of metamorphism and magmatic crystallization and gravity differentiation,
and the other the superficial processes of weathering, transportation and
sedimentation at the surface of the earth. Correspondingly, the dispersion
pattern related to the former is called the primary dispersion pattern,
or primary halo or aureole, and that related to the latter is called the
secondary dispersion pattern, or secondary halo or aureole.
The boundary between "primary" and "secondary" is not always clear-
cut, however. For example, patterns in the walls of orebodies formed by
syngenetic sedimentary processes are not primary dispersion patterns in the
strict sense but are fossil secondary patterns (James, 1967). For simplicity,
a geochemical dispersion pattern found in unweathered bedrock is called
here a primary pattern, and a dispersion pattern found in the overburden
(weathered bedrock, till, sand, peat, plants, etc.) is called a secondary one.
In practice, it is more useful to know whether a particular dispersion
pattern represents a chemical or clastic anomaly; such knowledge is vital
in revealing the suboutcrop of an orebody.

Primary dispersion
Primary geochemical dispersion is usually associated with high temper-
atures and pressures (Rose et al., 1979; Beus and Grigorian, 1977) and
involves crystallization from magma and hydrothermal liquids and the
metasomatic migration of elements. The zone formed as a result of primary
dispersion and within which the content of valuable elements diminishes
to background values is called the primary halo or aureole. The width of
94 Geochemical dispersion in the secondary environment

the aureole varies with factors such as lithology, tectonics, P-T conditions,
availability of water and the particular element involved.
There may have been several successive stages in the primary dis-
persion. The differentiation of materials and dispersion of elements that
happened long ago during weathering and transport and during sedimen-
tation of recent sedimentary and metamorphic rocks gave rise to fossil
secondary dispersion patterns, which form a transition between primary
and secondary dispersion patterns.

Secondary dispersion
Secondary dispersion commences when rock begins to disintegrate by
weathering. The constituents of the rock, either as mineral grains or in
aqueous solution, are transported by ice, air and water seeking to reduce
differences in gravity or electrochemical potential. The relevant factors in
the secondary dispersion of elements are thus the type of material, the
transporting agent and the mode of dispersion, all of which are closely
interrelated (Kauranne, 1976a).
Weathering is the process of changing of mineral material at or near the
earth's surface to products that, in the prevailing chemical and physical
conditions, are in better equilibrium with their environment. The forces
influencing the weathering process are, in principal, physical or chemical
and very often a combination of the two. Even weathering processes
that seem to be very complicated usually prove to be based on a simple
conformity of physical or chemical laws (Oilier, 1969).
Physical weathering is mechanical comminution of rocks and minerals
caused by several separate processes — partly originating from the rock
itself, partly depending on the physicochemical conditions prevailing in
the surroundings. The most important processes are fracturing caused by
tension forces in the rock and freezing and melting of water in the fissures;
even changes in temperature may cause the rock to break mechanically.
Typically the first stage in the weathering process, physical weathering
facilitates the progression of chemical weathering by enlarging the total
surface of rocks and minerals.
Chemical weathering can be defined as the chemical reactions between
rocks and minerals and constituents of air and water at or near the
earth's surface. In this environment oxygen, carbon dioxide and water
are abundant, and temperatures and pressures are low in comparison
with the conditions of the deep-seated primary environment. Rocks and
minerals that were stable in the primary environment are unstable in the
secondary environment, and chemical changes occur in their attempt to
reach equilibrium. As a result, elements bound together in rock forming
minerals may become separated in the secondary environment because
their behaviour is now governed by an entirely new set of parameters, or by
Introduction 95

1 1 1 1 1 1
Soluble
cations
•Cs

• Rb
. K • Ba
Elements of
hydrolysates
• Sr
'fRare .Th
•Na • ca y earths •u
• M n / • Sc • Zr
. Fe • Nb, Ta
• Li . Fe • Ti
y Mg * Ga
• Mn
" Λ1
_'Si
• P .S
/ • Be
Soluble complex
• B
• C . N anions
1 1 1 1 1 1
0 1 2 3 ^ 5 6 7
Ionic charge

Fig. 5-1. Geochemical separation of some important elements on the basis of their ionic
potential. Reproduced with permission from Mason (1966).

different values of the old parameters, which result in different dispersion


characteristics for each element. The pH of solutions is the controlling
factor in most of the chemical reactions involved in weathering processes.
Closely connected with H + -ion concentration is the redox potential which
also regulates chemical reactions and the weathering process. The third
important factor in chemical weathering is the ionic potential. Elements
that have low ionic potential (e.g., Na, K, Mg) remain in solution; elements
with average ionic potential precipitate through hydrolysis; and elements
with high ionic potential form soluble complex anions (Fig. 5-1). Thus,
depending on their ionic potential, some elements remain in residual
material while others are transported away in solution.
Rock-forming minerals react to chemical weathering differently accord-
ing to their chemical composition. The order of resistance can generally be
defined as oxides > silicates > carbonates and sulphides. The final products
of chemical weathering are: (1) soluble constituents, (2) insoluble secondary
minerals and (3) insoluble residual primary minerals.
The variables in the chemical weathering process are usually so numer-
ous that to predict changes and reactions is impossible. The transport of
weathering products away from the site allows the weathering process to
continue; without transport the result would be a closed system in chemical
equilibrium, and weathering would cease.
The products of physical weathering and the insoluble minerals produced
by chemical weathering are clastically dispersed by ice, water, wind and
96 Geochemical dispersion in the secondary environment

gravity In areas of glaciated terrain the most important of these agents


is continental ice; the other agents contribute to the total process but are
not of great relevance to geochemical prospecting. In mountainous areas,
soil creep and rolling of talus will be of importance in the transport of
weathered bedrock and till downslope.
As well as clastic dispersion, dispersion in solution due to chemical
weathering takes place in the glacial overburden. The two processes are
at least partly simultaneous, as each grain in the glacial till continues its
chemical weathering in its new physico-chemical environment after glacial
transport.
During mechanical transport of the materials the grains are finely
ground and, depending on the transporting medium (ice, water, air), are
either spread over a wide area or are concentrated in a few places. Grains
in drift undergoing clastic transport may also be dissolved in water and
transported further as ions. The solubility of the mineral grains of the drift,
as of the minerals of the original rock, varies according to local conditions:
acidity, oxidation potential, temperature, pressure, degree of saturation of
the water solution, etc.
When the conditions change, some of the dissolved elements will be
precipitated. Organisms, organic and inorganic colloids, and mineral grains,
especially clay minerals with an expanding lattice, scavenge ions and cause
their fixation and sedimentation. Eventually these ions may be leached
once again and moved still further on, so that there are several successive
stages in the secondary dispersion as well.
In areas of glaciated terrain secondary dispersion is mainly clastic in
nature, as the material of primary aureoles is abraded and scattered by the
continental ice. The contributing roles of gravity, water and wind are small.
Chemical dispersion in ionic or molecular form was probably significant
during weathering of the crust before the glaciation, which disrupted the
process. Since deglaciation, a new cycle of weathering and transport has
been proceeding, but this represents a very short period in the geological
time scale.

CLASTIC DISPERSION

Glacigenic clastic dispersion

General
Clastic dispersion in glaciated terrain is the transporting of physically
weathered or crushed material by glacier (Fig. 5-2), and, to a lesser extent
by water, wind and gravity. Chemical alteration of the material is minor,
and the main part of the material is transported in the form of rock
fragments and minerals.
Clastic dispersion 97

Fig. 5-2. A schematic impression of clastic dispersion in drift.

Goldthwait (1971, p. 3) wrote that "Till is the only sediment stemming


directly and solely from glacial ice". But he continues, "Till has more
variations than any other sediment with a single name" (p. 4). Dreimanis
(1983) classified tills according to their formation and genesis into nine
different types, providing (p. 301) a succinct definition of till as "a sediment
that has been transported and deposited by or from glacier ice with little or
no sorting by water" (see Chapter 2).
To correctly interpret the results of a geochemical study it is important
to know the type of till sampled, although it is not necessary to be able
to recognize separately all of the nine different till types presented by
Dreimanis. Usually division into two different classes is sufficient: (1)
basal till containing moderately or abundantly finest fractions, which is
transported a short distance and almost totally unsorted, and (2) ablation
till washed by the melt waters of ice, usually containing only small amounts
of fine fractions and in most cases transported a long distance. Although
the various types of basal till differ in genesis, this does not usually give
rise to essential differences or problems in interpreting the results of a
geochemical till study The amount of water operating on ablation till
during the sedimentation phase, especially on the finest fractions, will in
certain cases change the metal contents and metal ratios.
There are a number of other factors connected with the genesis of till,
whose influence on the interpretation of results is much more important
than the way in which the till was deposited. In areas where there was a
thick pre-glaciation weathering profile, weathering products are prominent
in the till, especially where the glacial erosion was weak; and geochemically
the till resembles weathered bedrock. Because the weathering products
reflect the nature of the weathering process, it is useful to know the
conditions that prevailed during the weathering process before glaciation
98 Geochemical dispersion in the secondary environment

and the chemical changes that took place in the rock at that time. The
imprint of these processes is often still visible in present day till.
In areas covered by thick waterlain stratified unconsolidated sediments
before the glaciation, the influence of those sediments on the geochemistry
of the till is dominant, even where glacial erosion was strong. Correspond-
ingly, the bedrock is reflected more weakly in the till compared with areas
where sediment layers were thin or absent.
The influence of factors prevailing before glaciation on metal contents of
the finest fractions of till is illustrated in Table 5-1. Deposits of preglacial
weathered bedrock tens of metres thick can be found under the till in
Finnish Lapland in the area of the last ice divide where glacial erosion
was very weak. The till cover is thin and itself contains abundant material
from the weathered bedrock. The metal contents of the till are high
and in some cases even higher than in the unweathered rock beneath
due to ionic dispersion and enrichment during the weathering process.
Large amounts of preglacial fine-grained waterlain sediments were mixed
with the till in western Finland, so that the proportions of the finest
fractions of till are greater than average there. During development of
these waterlain sediments in a number of successive stages, many metals
were concentrated or leached and transported away, causing higher or
lower metal concentrations in the till than in the source rock. The original,
unaltered material from bedrock comprises the main part of the till in
eastern Finland. The dissolving and sorting influence of glacifluvial and

TABLE 5-1
Some average metal concentrations of till in areas, where the material of till differs by its
preglacial state. Analyzed by I CAP after Aqua Regia leach

Northern Finland 3 Eastern Finlandb Western Finland0


Al(%) 1.53 0.74 1.14
Ba (ppm) 43.3 55.4 79.5
Cr (ppm) 45.9 22.8 30.8
Cu (ppm) 23.7 17.8 22.2
Fe (%) 2.10 1.04 2.20
U (ppm) 9.56 8.33 15.9
Mn (ppm) 163 37.8 277
n 105 692 995
Share of <0.06 mm
fractiond (%) 30 20-25 38-50
Area (km2) 25,000 1,500 4,000
a
Till consists of an abundance of weathered bedrock.
b
Till is composed of fresh bedrock.
c
Till consists of an abundance of preglacial fine-grained sediments.
d
Data compiled from different sources.
Clastic dispersion 99

fluvial waters has had little effect and the chemical composition of the till
is similar to that of the underlying bedrock.

Transport direction of till


In the area of the ice divide the movement of ice and the transport
of material is minimal, and because the ice divide also tends to fluctuate
a little, even during the same glacial stage, it is difficult to say anything
about the direction of transport there. On the other hand, since the distance
of transport was in any case very short, the direction is of no essential
importance to the interpretation of the results of geochemical till studies in
these areas.
A comprehensive study on the transport direction and distance is
essential in all other areas of glaciated terrain where geochemical till
studies are carried out. The general direction of ice movement is from
the ice divide towards the ice margin, although local differences may
be appreciable in some cases. The greatest deviations from the overall
direction are encountered in areas where glacial lobes were active during
the deglaciation. During the main phase of glaciation the ice flow was more
or less stable, but many local factors come in to play as well at the newly
activated ice margin of the melting phase.
The general consensus today is that till is mainly created and deposited
during the deglaciation phase. When the ice lobe is already thinned, small
differences in the topography of the terrain will easily produce a change
in the flow direction. Changes in the ice flow direction are also often
associated with oscillations of the retreating ice margin, and in such places
determination of the exact transport direction of till may be difficult.
The best evidence of ice flow direction is provided by roches moutonnaes
(Fig. 2-6). Larger forms, like the troughs and grooves in an outcrop and the
shape of the outcrop, as a rule are created by strong glacial erosion. Smaller
abrasion marks, e.g., striae, are generated throughout the glaciation, but
evidently most of them during the deposition of till. The striae are very
sensitive in showing the latest direction of the ice flow — as well as the
variations in flow occurring close to the margins of ice lobes and valley
glaciers (Fig. 5-3).
Because local variations in the topography may change the direction of
the ice movement appreciably — a change of 15° was once observed on a
single stria in a 30 m range — it is not reasonable to measure the direction
of striae with greater accuracy than 5°. If the rock type is coarse grained
or otherwise easily weathered, postglacial weathering may have destroyed
the striae. The best places to find good striae are then new road cuttings
and other areas where the bedrock is freshly exposed. Shore cliffs are also
good places to look for striae, although care must be taken to distinguish
glacial striae from striae generated postglacially by winter ice. Where there
are neither outcrops in the study area nor measurable striae on outcrops,
100 Geochemical dispersion in the secondary environment

Fig. 5-3. Radial streamlines of flow in the marginal part of a glacier lobe, seen in ground plan
showing: a, direction of streamline when lobe, A, is fully extended; 6, streamline when lobe,
B, is shrinking, Sets of striations formed parallel with such streamlines would have different
trends. Modified after Flint (1967).

the direction of ice movement and the direction of the source of the till can
be determined from the orientation of stones. When till is in motion, the
long axes of the most contained stones align themselves in the direction
of the transport. As long as postglacial processes (such as frost heave or
solifluction) have not disturbed the orientation, counts of pebble alignment
can provide a statistical direction for the ice movement.
Information about the transport direction of till needed to interpret the
results of a geochemical till study must always be obtained from the same
till bed and for the depth at which samples will be taken.

Transport distance of till


The distance of glacial transport and the length of the geochemical
anomaly are the most important features of clastic glacial dispersion. The
length of the geochemical anomaly is the distance from the source area to
the farthest point at which material derived from the source area can be
recognized by the applied sampling and analytical methods. Thus length
depends not only on the particular material assayed and its transport
distance, but also on the sensitivity of the method of analysis. Moreover,
the greater the chemical difference between the mineralization and the
country rock, the greater is the anomaly/background contrast and the more
visible the anomaly.
A simple and reliable way to determine the transport distance of till is
to investigate how far material has spread from some well-known bedrock
formation in the study area. Stone counts and chemical analyses at different
depths along a line in the direction of ice flow can be used to determine the
proportion of the material derived from the source area (Fig. 5-4). Material
derived from a particular suboutcrop rises obliquely in the till blanket
towards the surface, while at the same time spreading out and becoming
diluted. Thus, the results obtained for the transport distance are different
at different depths: the distance measured in the surface part of the till
blanket is greater than that measured in the bottom part, and the thicker
Clastic dispersion 101

4 km

Percentage of s c h i s t s Bedroc k
n stones of t i l l

111 >40 %


Archean granitoids

30 - 40 % AmphiboJite

•'.\·.·'.■'.·
25 - 30 % Mica schist

4 km

Zn in till

> 80ppm 60 - 80 ppm 40 -60 ppm

Fig. 5-4. The percentage of schists in the 0.8-2.0 cm till fraction (upper figure) and the zinc
content in the <0.06 mm till fraction (lower figure) in Parissavaara, Ilomantsi, Finland. The
direction of ice movement is from left to right (320°). Modified after Salminen and Hartikainen
(1985).

the till blanket, the greater is the difference in transport distances between
the surface and the bottom. This means that the transport distance must
be measured from the particular till bed and depth that will be sampled
during the geochemical study. Moreover, since the transport distance of
different fractions may vary as well, measurement should be made for the
102 Geochemical dispersion in the secondary environment

actual fraction that will be used in geochemical analysis. Only in this way
can the optimal cost effective sampling grid be designed to provide the
desired scientific results.
According to the transport distance, tills can roughly be divided into two
groups: (1) those transported a long distance (1 km or more) and (2) those
transported a short distance (tens or hundreds of metres). Tills transported
a short distance are common in areas where the glacier was relatively
inactive during the deglaciation. Correspondingly, tills transported a long
distance are common where the glacier was active during the deglaciation
(the marks of active ice lobes are usually recognizable). Tills transported
long and short distances may exist side by side, often with a sharp boundary.
In areas of active ice lobes the transport distance of till and the length of a
geochemical anomaly may be several kilometres. The transport distance of
individual mineral grains or boulders will sometimes be much longer than
the average, however. Boulders derived from the Outokumpu ore eastern
Finland, for example, have been detected at a distances of more than 100
km from the deposit.
Between the active ice lobes, where there are wide areas of passive ice,
the till is local and strongly reflects the underlying bedrock. In most cases
the transport distance is only some tens or hundreds of metres, and the
lithological composition of the till is more or less homogeneous. Areas like
this, where long and complex glacial transport does not complicate the
interpretation of results, are ideal for geochemical studies.
The partition into areas of active and passive ice is equally seen in the
lengths of geochemical anomalies. Of 34 geochemical anomalies reported
from Canada (Bradshaw, 1975) and Scandinavia (Kauranne, 1976a), 14
were less than 400 m in length (Fig. 5-5). The secondary maximum in the
frequency distribution of the lengths of these 34 anomalies, lying between
1.2 and 1.5 km, reflects the distant transported tills.
Few geochemical anomalies have been reported for tills in areas of active
ice; this is apparently because, in the course of extended transport, material
from different bedrock types has been added and mixed into the till making
the results more difficult to interpret. The metal ratios change and contents
are strongly diluted along the anomaly, and other kinds of changes may
occur as well, depending on the circumstances and particular metals.

Patterns of clastic anomalies in till


A geochemical anomaly in till derived from a particular source (ore,
minor mineralization, particular rock type, etc.) is three dimensional. In
the ideal case the anomaly pattern looks like a smoke plume drifting
laterally in the direction of ice movement from the suboutcrop of the ore
towards the surface of the glacial overburden (Drake, 1983) (Fig. 3-16).
The surface area of the ore suboutcrop is an appreciable factor in
determining how widely the anomaly plume can spread. Long transport
Clastic dispersion 103

n Scandinavia

Canada

800 1600
j
2400
400 1200 2000 2800 3 2 °° m
Fig. 5-5. The length of geochemical anomalies in Scandinavia (Kauranne, 1976a) and in Canada.
Compiled from Bradshaw (1975).

distance causes dilution of the ore material in till because material from
other rock types over a wider area is intermixed. Although the anomaly
can still be detected far away from the source area, it is weak and no
longer reflects the source well (Fig. 5-6). The extensive dilution of the
anomalous material places special demands on sampling (samples should
be homogeneous and sufficiently representative) and on the sensitivity of
assaying methods.
The width of the ore suboutcrop (or of rock formation, or mineralization)
in the direction of ice movement has a decisive influence on the shape and
size of the anomaly pattern. The wider the source area the more a glacier
can abrade it, and hence the greater the amount of anomalous material
that can move up into the glacier, where it enjoys longer transport than
at the bottom of the glacier. The anomalies derived from the serpentinites
in the Outokumpu region provide a good example of this (Fig. 5-7).
The serpentinites lie in the bedrock as folded bands 0.5-2 km wide and
hundreds of kilometres long. In the Kokka and Losomäki areas where the
bands are oriented in the direction of the ice movement, the glacier has
abraded the serpentinites over a range of 7-8 km, and the resultant Ni
anomaly is clearly detectable 20 km from the distal contact. In Outokumpu,
on the other hand, the band-like serpentinite lies transverse to the ice
movement, and the abrading action of the glacier has been effective over
1 km only; here the Ni anomaly disappears into the background 5-6 km
from the distal contact of the serpentinite. A few kilometres to the north
104 Geochemical dispersion in the secondary environment

Rocks of Outokumpu-association Black t schist Mica schist

Fig. 5-6. The summed percentage of mica schist, serpentinite and skarn at different depths
(upper figure) and the content of nickel (ppm) in the <0.06 mm till fraction (lower figure)
in the pits at Miihkali, eastern Finland. The direction of ice movement is from left to right.
Reproduced from Salminen and Hartikainen (1985).

east near Horsmanaho the same band becomes still narrower, and the
Ni anomaly shortens to about one kilometre. The surface areas of these
serpentinite outcrops are much greater than those of most ore outcrops,
but the relationship between the width in the direction of ice flow and
the length of the anomaly is the same. Roughly it can be said that the
detectable length of a metal anomaly in the fine fractions of till is 2-4 times
the width of the source area in the direction of ice movement.
In addition to the width of the suboutcrop, weathering and the resistance
to abrasion of the bedrock influence the anomaly/background contrast. The
glacier is able to loosen more material from softer and more weathered
bedrock than from hard and unweathered bedrock. The host rock of the
Outokumpu Cu ore is quartzite, a rock type resistant to abrasion, and a
common ore mineral is pyrite, a mineral harder to weather than many other
sulphides. Although the copper content of the ore is 3.8%, the anomaly in
till derived from the ore does not differ appreciably from the anomalies
Clastic dispersion

£} High Ni anomaly

■■' ) Low Ni anomaly

Zone of the rocks


of Outokumpu
association

I Boundary between
long and short

Fig. 5-7. The anomaly pattern of Ni in the <0.06 mm till fraction, Losomäki-Outokumpu
area, eastern Finland. The anomaly is caused mainly by the serpentinites of the Outokumpu
association. The difference between the areas of long and short transport distance is clearly
visible.

derived from the nearby black schists which have a Cu content only one-
hundredth of that of the ore (Fig. 5-8). This is because the width of the
black schist layer in the ice movement direction is 2-3 times the width
of the ore outcrop, and the black schist is much softer and more easily
weathered than the ore.
The grain size of the original bedrock has a further effect on the
formation of an anomaly. In the case of two rock types having the same
weathering properties and resistance to abrasion, more material from the
fine-grained than from the coarser-grained bedrock gets into the finest
fractions of till typically analyzed in geochemical prospecting. Material
is easily ground to its original grain size (Dreimanis and Vagners, 1971;
Haldorsen, 1983). So it is possible — and in fact quite common — that a
geochemical anomaly in till derived from a rock type with low metal content
but occupying a large area is as high as the anomaly derived from a rich
but small ore outcrop.
106 Geochemical dispersion in the secondary environment

Fig. 5-8. The anomaly pattern of Cu near by the Outokumpu ore outcrop, Outokumpu, Finland.

Although in theory a geochemical till anomaly is a simple matter and the


majority of the work has been done by taking into account the transport
distance of till and the surface area and resistance to erosion of the source
rock, there remain several other complicating factors. Chemical weathering
of the till continues after the transport and deposition, and at different
rates above and below the groundwater table where the oxidation-reduction
conditions are different. As some of the products of this weathering are
carried as ions and molecules by groundwater in the direction of the
gradient, changes will be effected in the shape of the original anomaly
pattern.
Unevenness of the terrain, which changes the ice flow direction, will
cause modifications in the anomaly pattern of till during the transport
stage. Later, the deposition of the till may be disturbed by the unevenness
Clastic dispersion 107

of ground, and again the anomaly pattern will be modified. A steep ridge
in the bedrock tens of metres high causes till to pile up in front of the
ridge, while only a little (most of it very local) is deposited on the ridge
itself. More till is again deposited behind the ridge, but it is now intermixed
with material from the proximal side of the ridge transported inside the
glacier (long transport, Fig. 5-9). The same rules hold for all fractions of
till — the geochemical anomaly pattern detected in the finest fractions is
often interrupted directly above the threshold of the ridge but is again
strengthened on the distal side (Fig. 5-10).
In places multiple transport of till has occurred, caused by two or more
separate ice flows during the oscillation of the ice margin. Often a later ice

Rocks of Rocks of Prekarelidic


Kalevian Jatulian Group basement
Group (mainly rocks
(mainly mica quartzites and
schists) volcanites)
>/o •
1km
30
^ ^ \
SO

40 b \^
20-

I I

Rocks of Kalevian Group


Rocks of Jatulian Group
Prekarelidic basement rocks

Fig. 5-9. Distribution of surface stones in Viesimo. Reproduced from Salminen and Hartikainen
(1985). (a) Topography of the study line, (b) Percentage of Karelidic rocks along the study line,
(c) Percentage of different rock groups along the study line.
108 Geochemical dispersion in the secondary environment

i l l Mica schist [jwÄ) Black schist [%%] Quartzite f£§) Rocks of Outokumpu- ~^" Sampling site
association (tractor excavator)

~~7T Sampling site


Anomalous Ni-content ( > 9 0 ppm) X~ Probable surface of bedrock (percussion drill)

Fig. 5-10. Nickel content in the <0.06 mm till fraction in Petäinen, Finland. Reproduced from
Salminen and Hartikainen (1985).

flow will have abraded the till and totally obliterated the marks of an earlier
flow. But in some cases the marks of separate ice flows are still discernable,
especially if the directions of the ice movement differed. The anomaly
pattern in till is the cumulative effect of successive periods of transport and
is elongated in accordance with the separate ice flow directions. Usually,
however, the erosion of the last ice movement was so strong that it renewed
the anomaly pattern totally; at least this can be assumed if only one till bed
is found in the study area.
Boulders strongly resistant to glacial erosion have often been transported
by more than one ice flow. This can be recognized in the shape of boulder
fans or in the location of a boulder from a known ore that cannot be
explained by a single transport direction. More commonly boulders lying
on the surface of a till bed have been involved in complex transport, in
which case a thorough study of the glaciation history of the study area is
needed.

Clastic dispersion by water and wind


In areas of glaciated terrain the most common drift material and
hence the most important secondary geochemical environment is till. Drift
formations deposited by water, and especially those by wind, are small
Clastic dispersion 109

in area compared with till-covered areas, although they may also occur
as a thin layer covering the till. In most cases the material of these
water- and wind-transported formations originated from till and as such
represents a secondary geochemical environment of second generation.
Tracing the source of geochemical anomalies in this kind of material is in
most cases impossible. The anomalies studied in glacifluvial material are
usually mineral anomalies — not anomalies of a single element.
Sulphide grains and other easily weather able minerals will seldom be
found in glacifluvial material. However, certain resistant minerals such as
cassiterite, magnetite and gold may even be concentrated by glacifluvial
processes, to the extent of forming economic mineral deposits (placer
deposits).
As most glacifluvial material has come from till, two different transport
mechanisms have had their effect on the dispersion of the material. The
total transport distance may be some kilometres or even tens of kilometres,
and to discover the actual transport route is difficult.
Silts and other fine-grained waterlain sediments which were deposited in
ice margin lakes or in the sea in front of the terminus of the continental
ice cover a much wider area than other glacifluvial deposits. This kind of
material has not found much use in geochemical prospecting. Although
it is quite possible to find geochemical anomalies in clay sediments, it is
virtually impossible to trace them back to the bedrock and particularly to a
specific ore deposit in the bedrock.
Transport by waves takes place along the coastline, restricted to the
depth of the wave action. Concentrations of heavy minerals, like magnetite
and garnet, occur along both present and ancient lake and sea coasts. The
material of beach sand may be derived from the drift or rock exposure on
the shoreline or transported to the coast by streams.
Boulders found on ancient shorelines in places where glaciers could
not possibly have deposited them may have been transported by icebergs.
Famous examples are the Finnish rapakivi boulders which have been found
as far away as in Germany and Czechoslovakia — a distance of more than
1,000 km.
Single boulders caught up in the bottom of an iceberg and deposited
on the sea bottom are called dropstones. In fine-grained sediments these
stones are easily recognized as transported by an iceberg, but proving an
iceberg transport mechanism for an ore boulder found at the surface of till
bed, amid glacially transported boulders, presents a challenge.
The wind carries mineral material over barren areas, and in the process
separates minerals according to their abrasion resistance and specific
gravity. Anomaly patterns created by wind have been reported in the Sahara
and Kalahari Deserts (Rose et al., 1979), but it would be unreasonable to
seek wind-created geochemical anomalies in areas of glaciated terrain. Even
if they were found, it would be exceedingly difficult to interpret them.
110 Geochemical dispersion in the secondary environment

The melting of the continental ice produced enormous amounts of water-


transported material, not only along present river valleys but elsewhere;
glacifluvial materials now form the overburden over wide areas of formerly
glaciated terrain. The geochemical anomalies in far-transported drift —
including the gravel and sand of eskers and shore deposits, sand and
silt of delta formations and dunes, and the silt and clay of sea or lake-
bottom sediments — tend to be subtle and difficult to interpret. In
general, therefore, the sediments of existing streams are better suited
for geochemical exploration and indeed have found wider use in routine
prospecting.
Clastic dispersion by water has been described here only very briefly
as another volume in this series (Drainage Geochemistry in Mineral Explo-
ration) specifically deals with the application of stream and lake sediments
in prospecting.

Clastic dispersion by gravity


Downslope migration of the mantle of rock-waste and associated dis-
placements of bedrock are commonly referred to as surface erosion or mass
movements (Holmes 1972). The movements range from falls of materials
and landslides resulting from slips, to soil creep resulting from various
types of flow. The most dramatic mass movements in glaciated terrain are
those associated with periglacial conditions or "peripheral to a glacier" as
Flint (1967) defined the term.
Solifluction is a seasonal, progressive downslope movement, mainly by
viscous flows of water-saturated drift. The movement differs from soil
creep in that, although frost heave plays a part, movement by true flow is
an important component. Solifluction is the most widespread and obvious
process of mass movement in high latitude areas not covered with perennial
ice and snow. It is most abundant in regions of perennially frozen ground,
with summer thawing limited to a depth of one metre or less (Flint, 1967).
Only a very gentle (2°) slope is needed. A young, wet oversaturated till is
particularly sensitive to solifluction.
The shape of a primarily glacigenic anomaly as well as the transport
direction and distance of the anomaly may be severely altered by soil creep
or solifluction. The action of the solifluction is strongest in the upper part
of the till layer where the results of a geochemical till study cannot be
interpreted without taking it into account.
Whereas solifluction mostly takes place in periglacial conditions, soil
creep, or rolling transport of material downslope, is a continuous process
in mountainous terrain. A talus cone is formed by weathering products
and other material falling or rolling down under the action of gravity.
Although this process can change a glacigenic geochemical anomaly, more
often it creates its own unique anomaly (Fig. 5-11). The source area of a
Chemical dispersion 111

Fig. 5-11. Clastic dispersion pattern caused by gravity in talus.

talus anomaly is found upslope where bare rocks are often outcropping.
The transportation of material may be complicated by later movements,
however, making results difficult to interpret.

CHEMICAL DISPERSION

General
Chemical dispersion is transport of material in ionic or molecular form,
mainly in water solution (Fig. 5-12). Gaseous dispersion and dispersion due
to electrochemical forces are special forms of chemical dispersion. The usual
prerequisite for chemical dispersion is weathering of the rocks. The main
factors controlling the dispersion are the acidity and redox potential of the
environment, the solubility of the minerals and other chemical compounds
in the rock, and the adsorption of ions and ion exchange on colloidal
particles. In the course of the dispersion the elements are released from
the primary minerals and transported in water solution, in aerosol form or
by electrochemical forces. Fixation of elements by chemical reaction into
112 Geoehemical dispersion in the secondary environment

Ice flow
(
Chemical dispersion
a) BY GROUND WATER Glacigenic clastic
b) ELECTROCHEMICALLY dispersion
C) POSTGLACIAL WEATHERING

<\
Ground water Jx_.<0 '/L· . '^Α ό O

T
Λ

flow ^<^^>j.\:vV/# ^
"7" M — X ^ X X
Bedrock

Clastically dispersed material


iOREf
Chemically dispersed material

Fig. 5-12. Schematic illustration of chemical and clastic dispersion in the overburden.

new (secondary) minerals, or adsorption on mineral surfaces or in colloids


terminates the chemical dispersion.
The first stage of chemical dispersion or the redistribution of elements
in the surficial environment is weathering of the fresh rock. Weathering
consists of a number of processes that gradually break down the fresh
solid rock into an aggregate of loose material; some constituents go into
solution, while others succumb to chemical changes and still others remain
unchanged in composition. The agents of the decomposition are both
chemical and physical.
The presence of water — rain, surface and ground waters and of
the solids and gases dissolved in them — is the basic requirement for
all chemical weathering as well as for chemical dispersion. The various
materials dissolved in water determine its acidity and redox potential,
which in turn determine the destructive and dissolving properties of water
on the surfaces it comes into contact with. Oxygen, carbon dioxide, nitric
and sulphuric acids, humic compounds, ammonia and chlorides are the
most important among those substances dissolved in natural waters that
promote decomposition (Rankama and Sahama, 1950). An understanding
of the chemical reactions and equilibria that determine the behaviour of
the elements in the surficial environment is clearly essential for effective
geoehemical prospecting (Rose et al., 1979).
The products of weathering at the Earth's surface are partitioned
between the relatively immobile solid phase that constitutes the regolith
and the mobile fluid phase made up of free-flowing ground and surface
waters. Weathering products that can be readily suspended and swept along
by flowing water are rapidly dispersed as ions or molecules.
Surface and ground waters leach chemical elements from weathering
products as well as from bedrock exposures and the overburden. However,
Chemical dispersion 113

the mineral grains in drift and in weathered bedrock are much more
vulnerable to leaching than those of the barren bedrock or outcrops because
of the greater surface area. Hence, except in mountainous areas, the main
part of the elements dissolved in water have probably been derived from
drift and the weathered bedrock rather than from the unaltered bedrock.
Because the weathered crust in situ is much thinner in glaciated areas
than elsewhere, the common geochemical rules of chemical equilibrium in
weathering processes must be applied for the special environment of glacial
overburden.

Acidity, redox potential and solubility of elements


Probably the most important factor governing the solubility of a given
element in water is the relationship between the concentration of hydrogen
ions (pH) in the solution, and the oxidation-reduction potential (Eh).
Most natural waters have a pH between about 4 (acid) and 9 (alkaline).
In glacial overburden, however, the range, is commonly only 5-8. The
concentration of sulphuric acid near oxidizing sulphides may be high
enough to produce a pH of 2 or even less. Most metallic elements are highly
soluble in acid solution but will precipitate as oxides or hydroxides when
the pH is increased. For alkali and most alkaline-earth elements, the pH of
precipitation is above the range of natural waters, but many higher-valent
and transition elements are precipitated in the pH range of natural waters.
Many elements occur in more than one oxidation state (valence) in
minerals and solution, and the character of the elements changes markedly
from one valence to another. For instance, Fe occurs in minerals and natural
waters as both Fe 2+ and Fe 3+ , and the minerals containing these two forms
differ from each other in almost all chemical and physical properties.
At the surface of the earth, atmospheric oxygen is the most abundant
oxidizing agent. Environments with access to the atmosphere or to down-
ward percolating waters saturated with oxygen are normally oxidizing. As
access to free or dissolved oxygen decreases with depth below the sur-
face or with impeded flow of oxygenated surface waters, the environment
becomes progressively less oxidizing and more reducing, although it may
still be oxidizing for one element and reducing for another. In glacial drift
the ground-water table usually forms a boundary: under the water table
conditions are more reducing, whereas above they are more oxidizing.
This change in Eh conditions at the water table in some cases leads
to marked differences in the metal content of the finest fraction of till
above and below the boundary. The extent of this difference depends on
the solubility of the minerals in the till. For example, sulphides are easily
leached in the oxidizing conditions above the water table and many metal
ions thus liberated are transported away. Even in areas where there is no
excess of sulphides in till, the difference in metal contents above and below
114 Geochemical dispersion in the secondary environment

TABLE 5-II
The average contents (ppm) of selected metals in the <60 urn fraction of till above (A) and
below (B) the water table. Analyzed by atomic absorption method after Aqua Regia leach. Data
from 14 pits in Polvijärvi, eastern Finland

Co Cu Mn Ni Pb Zn Fe Number of
samples
A 7.2 27.1 80 16.9 5.5 22.1 7400 59
B 8.0 37.6 114 20.2 6.2 31.7 9300 39

the table is sometimes considerable, as is shown in Table 5-II. Although


a marked difference in the metal contents of till would not be expected
everywhere, the feature is so common that it is always better to take
till samples for geochemical prospecting from underneath the water table
where conditions have remained more stable since the last glaciation.
Carbonaceous material produced by plant and animal life furnishes the
principal reductant in the surficial environment. Where organic matter
has accumulated and access to the atmosphere is restricted, conditions are
strongly reducing. Glacial till and glacifluvial sediments seldom contain
that much organic material, but other fine grained sediments do.
Because Eh and pH are such important variables, many features of the
mobility of elements become evident in Eh-pH diagrams, such as those
shown in Fig. 5-13. The Eh and pH of some typical natural environments
are summarized in Fig. 5-14, where the upper oblique line marks conditions
oxidizing enough to decompose water to O2 (1 atm) and H + and the lower
line is a similar limit for conversion of H + to H 2 (1 atm). In general, the
measurement of Eh in most natural waters is fraught with such great
difficulties (Morris and Stumm, 1967), that usually it is not reasonable to
ask the prospector to study the Eh of soil waters in his everyday work.
The solubility of minerals is an important determinant of the mobility
of elements. If perfect chemical equilibrium always existed, the solubility
would be the absolute upper limit for the concentration of an element in a
natural water. In fact, supersaturation and related non-equilibrium states
can lead to concentrations higher than the equilibrium values. A saturation
state in moving waters is seldom reached. Supersaturation is most evident
for minerals of very low solubility.
Oxides and hydroxides are relatively soluble, but a wide variety of other
minerals limit the mobility of elements from them. Sulphides are easily
leached but highly insoluble in reducing environments. In the case of trace
elements the least soluble mineral may be unknown. Yet a knowledge of
possible solids is necessary to predict the limits on mobility. The true limit
of equilibrium solubility may be far lower than the inferred one if the
correct mineral phase is not known, but the least soluble known mineral
does give an upper limit.
Chemical dispersion 115

2.0-, Eh (V) 2r. 0

1.8 i \- i.e

0 1 2 3 U 5 6 7 8 9 10 11 12 13 14
pH

Fig. 5-13. Behaviour of the reduction-oxidation potential (Eh) of several valencies of the
elements arsenic, iron, cobalt, manganese, nickel and vanadium as a function of the hydrogen
ion concentration (pH) at normal conditions (298.16°K, 1 atm). Modified after Rosier and
Lange (1972).

Although most metallic elements occur in acid solution as simple ions


(e.g., Cu 2+ , Pb 2 + , Ag + ), a wide variety of complex hydroxy and oxy ions are
formed in more alkaline solutions. Complexes may be formed in solutions
containing other anions or organic compounds with which the cation forms
a compound of low solubility.
The formation of complexes in ground water of glacial till is not
significant for geochemical prospecting because the time elapsed since
the last glaciation is too short, and the anions and organic molecules
needed for forming complexes are not common enough in till. The role
of complex formation is much more important in the geochemistry of
116 Geochemical dispersion in the secondary environment

-0.8

-1.0
10 U
pH
Fig. 5-14. Approximate position of some natural environments as characterized by Eh and pH.
Reproduced with permission from Garrels and Christ (1965).

waterlain sediments, even in glaciated areas, and in the geochemistry of


weathered bedrock and of course biogeochemistry all over the world.

Adsorption and ion exchange on colloidal particles


Besides solubility, another major process governing the mobility of trace
elements in the surficial environment is adsorption and ion exchange
onto colloidal particles, which, with diameters in the range 10-10,000 Ä
(10~7 to 10" 4 cm), have a large surface for adsorption. Iron and manganese
oxides and hydroxides, organic matter, clays and silica are the most common
natural materials occurring as colloidal particles. Initially, the ions adsorbed
on the surface of colloids are in active equilibrium with the surrounding
solution, but with time they tend to migrate into the crystal lattice or
Chemical dispersion 111

become overgrown by the solid crystal. Under favourable conditions, small


amounts of colloidal material can scavenge important amounts of dissolved
elements from solution. Upon flocculation of the colloidal particles, the
originally dissolved elements are immobilized into soil or stream sediment
as a coating on a mineral grain (Carpenter et al., 1975).
For the most part, the adsorption and exchange are driven by electrical
charges on the surface or in the lattices of colloid-size particles. The surface
charge of oxides, hydroxides, organic materials and other types of colloids
with constant surface potential results from the ionization of surface atoms
or from the chemical adsorption of dissolved ions onto the surface. An
adsorbed ion can occupy several types of sites in or near a charged colloidal
particle. In solution, most ions are surrounded by a shell of water molecules
held by electrostatic attraction between the ion and the H2O dipole.
Charged colloidal particles are surrounded by a cloud of oppositely charged
hydrated ions (counterions) attracted by the charged surface. Exchange of
these electrostatically attracted ions with the surrounding solution takes
place rapidly and easily, with only minor discrimination between ions. The
attraction of such ions by predominantly electrostatic forces is relatively
constant.
Two types of material may be distinguished according to the charge
distribution: material having fixed charge, primarily the clay minerals, and
material of fixed surface potential such as the Fe and Mn oxides. Metal
ions are adsorbed from ground water onto the surfaces of minerals in the
clay fraction (<2 μηι) through the whole till blanket. Studies by DiLabio
and Shilts (1979), DiLabio (1979), Nikkarinen et al. (1984) and Salminen
(1980) on the variation of metal contents in different grain size fractions
reveal a marked enrichment of most metals in the finer fractions (Fig.
5-15). Although the enrichment varies from metal to metal and from place
to place, the metal content of the clay fraction is usually much higher than
that of the sand and gravel fractions because of adsorption.
Fe and Mn precipitates, common in the B-horizon of the podsol profile
in till and sometimes also in the zone where the water table fluctuates,
are available to scavenge large amounts of Zn, Co, and Ni ions. Fe-Mn
oxide coatings on the pebbles of till have been found in the same horizons.
In practice, the significance of Fe-Mn precipitates and Fe-Mn coatings for
prospecting is not as great in glacial till as in areas of older weathered
bedrock outside glaciated areas.
According to Rose et al. (1979), the amount of adsorbed metal ions is
dependent on the following factors:
(1) The activity of metal ions in the surrounding solution (taking account
of any complexing).
(2) The pH of the surrounding solution.
(3) The energy of bonding, as expressed by the equilibrium constant.
(4) The number or density of adsorption sites.
118 Geochemical dispersion in the secondary environment

Cu
Kaasila Maaselkä Talvivaara
(0 10.0 "i
(0 ·\Λ·
υ
(A ;:-
• / I · ·/· ·
o
*/ ·
Φ 1.0 •V • ·
c
o ^7>
Ü
Φ

(0
£ o.i _
2000 500 64 <2 2000 500 64 <2 2000 500 64 <2 |im

Zn
Kaasila Maaselkä Talvivaara
W 10.0
• #
(0
•/
•*··/.
o
(0
J· ·
0)
o ·/·.·

k
c
<D 1.0 %·* -S—**£ *· ·
•*.v··
c
.·:#·.·'
o
o • •
Φ

•(0>
~
2 o.i
2000 500 64 2000 500 64 <2 2000 500 64 <2 μπΐ
Fig. 5-15. Relative Cu and Zn contents in different grain-size fractions compared with the
content of the total till sample. In Talvivaara only three samples contained enough clay for
analysis. Reproduced with permission from Nikkarinen et al. (1984).

(5) In more complex cases, the activity of other ions competing for the
sites.
Geologically, glacial till is a very young material and many of the
processes described above have not had time enough to reach a chemical
equilibrium. It should be emphasized, therefore, that the rules described in
textbooks will not work at all times and in all places.

Electrochemical dispersion
Within and around an orebody with high electrical conductivity, an
electrical potential field is created by the different states of oxidation in
the environment of its upper and lower parts. The orebody and the ground
water of the surrounding country rock could thus be considered as a
galvanic cell through which electrical currents flow, carried by electrons
within the ore and by ions in the ground water (Fig. 5-16). These currents
Chemical dispersion 119

Oxidation
of ore

Fig. 5-16. Schematic model of the upper part of a sulphide ore body outcropping under glacial
till. Thin lines: equipotential surfaces. Heavy lines with arrows: current direction of positive
electricity (H + ) in the electrolyte. p&, ρψ, Ρτ> Ρο>: electrical resistivity in country rock, oxidized
zone, glacial till and ore, respectively. Upper curves indicate paths to be expected from various
dispersion mechanisms in the ground water of the overburden samples close to the bedrock.
Modified after Bölviken and Logn (1975).

may produce chemical distribution patterns distinct from those produced by


diffusion and ground-water movement (Govett, 1973, 1975, 1976; Bölviken,
1979; Bölviken and Logn, 1975).
A typical anomaly pattern in soil on a line crossing such an orebody
is a two-peaked "rabbit ear" anomaly, where the ears show the locations
of the contacts of the ore. This kind of pattern is caused by an ore in
the vertical position (see Fig. 5-17), while a different pattern is produced
if the orebody is lying obliquely (Fig. 5-18). The most striking anomalies
generated by electrochemical forces are reported for the humus layer where
cations are bound to organic compounds (Nuutilainen and Peuraniemi,
1977). Although this type of anomaly pattern may also exist in till (Govett
and Chork, 1977), the chemical process needed for the formation proceeds
more slowly.
In certain circumstances a prospecting method based on electrochemical
dispersion may be useful in delineating a sulphide ore body found by other
methods; all that is required is measuring of the self-potential generated by
the ore, and time-consuming sampling and analysis can be dispensed with.
120 Geochemical dispersion in the secondary environment

Zone of maximum stress due to


Organic carbon in soil high electron current density

Surface

H + in soil due to H"1" in soil due to oxidation of


electrochemical dispersion sulphide body

Surface

Surface self.

Surface

Surface
^Λθ<Τ

Equipotential lines, mV [ ] Sulphi de A Anode


► Current flow C Cathode Electron flow

Fig. 5-17. Schematic illustration of conceptual electrochemical model and current distribution
around conducting sulphide body, and resulting distribution of H + and organic carbon in surface
soils. Reproduced with permission from Govett and Chork (1977).

Gaseous dispersion

Gases are continuously migrating up through the bedrock and the


overburden (Cazalet, 1982; Malmqvist and Kristiansson, 1983, 1984). These
include He and Rn formed in the radioactive decay of elements, noble gases
Chemical dispersion 121

Cu ppm Cu/C

25 0.025

15 + 0.015
Cu

Cu/C
OlO

Fe% Fe/C
4.5 + 18

2.5110 Fe/C

Fe
OlO

mV
450+

225

SP profile

-75

Ia Δ l Overburden
ffltitti Monzonite
I I Skarn and amphibolite
I■:■: I Quartzite
I H Mineralization

\ Drill hole

0 200 m
Fig. 5-18. Some metal concentrations and metal/carbon ratio in humus samples along a
study line over a sulphide mineralization, Laurinoja, Finland. Modified after Nuutilainen and
Peuraniemi (1977).

and hydrocarbons, and a wide range of sulphur gases. Metal ions are
liberated from certain types of ores and are dispersed in the vicinity of
the orebody as far as conditions and time allow. What is occurring, in fact,
is chemical dispersion in the gaseous phase. It can be supposed that the
oxidation of sulphides to sulphates is the chemical reaction producing a
122 Geochemical dispersion in the secondary environment

Anomaly formed by
gaseous dispersion

Fracture
zones

Fig. 5-19. The formation of a geochemical anomaly in the overburden by gaseous dispersion.

part of the gas anomalies detected in the overburden. A sufficient porosity


of bedrock and overburden is a prerequisite for the gaseous dispersion
of elements from the orebody. Although both overburden and weathered
bedrock are sufficiently porous for gaseous dispersion, in unweathered
bedrock the cracks, fissures and fracture zones are the only possible routes
for the dispersion. A geochemical anomaly created by the gas flowing
through those crack zones may be situated far away from its source (Fig.
5-19).
Measurement of the concentration of Hg and Rn gases in the overburden
has been used as a prospecting method for the last two decades. Lead and
zinc ores contain a small amount of Hg which is liberated little by little
and can be detected as anomalous Hg concentrations about the orebodies
(Bradshaw and Koksoy, 1968; Robbins, 1973). In prospecting for uranium,
radon liberated as a product of the fission of uranium is measured by
radiation detectors (alpha-detectors) buried in the ground (Dyck, 1969a
and 1969b; McCorkell et a l , 1981).
Dogs trained to smell the sulphur compounds liberated from sulphide
ores were first used in Finland in the 1960's (Kahma, 1965; Kahma et
al., 1975), and have successfully been used in boulder tracing to discover
sulphide suboutcrops and boulders hidden under the overburden.
McCarthy et al. (1986) recently studied a sulphide ore using C0 2 , CH4
and 0 2 measurements, contending that gas geochemistry can even be used
in distinguishing massive sulphide bodies from other conductors. Methods
originally developed for detecting hydrocarbons in oil exploration have also
been applied in ore prospecting.
The various prospecting methods described above work as well — or as
badly — in glaciated terrain as outside it. For the equipment (and dogs) to
function well, the overburden should rather be dry and the weather fine.
Chemical dispersion 123

Although the methods are technically well developed and easily applied,
wet soil and rainfall will reduce their effectiveness or even prevent their
application. Gas geochemistry is treated in detail in another volume of this
series, Volatile Elements in Mineral Exploration. A short description was
included here because the possible transport of heavy metals by carrier
gases must be taken into account when tracing the source of an anomaly in
soil.

Discrimination of different anomaly types


Secondary geochemical anomalies can be divided into two separate groups
according to the dispersion: 1) clastic anomalies and 2) chemical anomalies.
A clastic anomaly is formed from material transported as mineral grains.
The force causing the transportation can be streaming surface waters,
wind, glacier or gravity on steep mountain slopes. The route of clastic
transport is relatively easy to trace, even though repeated transport may
cause difficulties for the interpreter. The material forming a chemical
anomaly has been transported in ionic or molecular form in water solution
or in gaseous form. Thus, ground water and hydroscopic water usually play
a critical role in the formation of a secondary chemical anomaly.
Anomalies encountered in the field will almost always be a composite of
these two main types. Chemical dispersion proceeds strongly if the bedrock
is weathering, and pure chemical anomalies may be developed. During
glaciation or as a result of landslides or soil creep, however, this chemically
dispersed anomaly becomes elongated, mechanically transported and it
easily assumes the shape of a pure clastic anomaly; the essential chemical
character — the mode of occurrence of the metal ions — remains the
same, however. Correspondingly, a primarily clastic anomaly may continue
its development chemically as the material abraded from the bedrock by
e.g., a glacier weathers in drift; ground water transports away the material
thus liberated and precipitates it again on the surface of mineral grains
and other suitable places.
Only if the dispersion type and transport mechanism of a secondary
geochemical anomaly are known it is possible to find the orebody or
mineralization that has caused it. A clastic anomaly usually declares the
type of ore from which it is derived as the material is mostly in the
form of the original mineral grains. The transport direction (whether
the transporting agent was a glacier, streaming water, wind, gravity or
something else) can usually be determined, and the transport distance can
be roughly estimated. Although a chemically generated anomaly usually
lies closer to its source, it is not always so easy to define the direction from
which the material has come. Moreover, the data obtained from a chemically
dispersed anomaly does not usually point to a single type of source. None
of the original ore minerals are preserved, and even the ratios of the
124 Geochemical dispersion in the secondary environment

elements will have changed in the course of dissolution and precipitation


processes. The situation becomes more complicated when chemical and
clastic components are usually present in the same anomaly.
The contributions of the clastic and chemical components of an anomaly
can often be differentiated through the application of analytical methods
in defined sequence (Chao, 1972; Nurmi, 1976; Gatehouse et al., 1977;
Hoffman and Fletcher, 1979; Sandström, 1984). Material transported in
ionic or molecular form is usually more weakly bound and will respond
more readily than primary mineral grains to the attack of weak agents.
Weak acids in neutral salt solutions are typically used to dissolve chemically
dispersed material, and strong mineral acids or total analysis methods are
used to dissolve clastically dispersed material. The most useful methods for
prospecting, it may be added, dissolve ore minerals but not metals bound
in gangue minerals. On the basis of combination of analytical methods it is
possible to determine the share of clastic and chemical dispersion in a single
anomaly. The several separate sequential analyses cause extra expense, of
course, and the procedure should be reserved for special cases.
Although much less so than in non-glaciated terrain, chemical migration
in certain cases plays a substantial role in transport and in the genesis of a
geochemical anomaly in the overburden of glaciated areas. This possibility
has always to be kept in mind if results are to be correctly interpreted,
and one must be prepared to apply the appropriate analytical methods if
necessary. Fortunately for the prospector, the clastic anomaly consisting
of material transported a short distance — the easiest anomaly to inter-
pret — is the situation most often prevailing in glaciated terrain.

CONCLUSIONS

Because they are broader, secondary aureoles often are easier than
primary aureoles, as targets of geochemical exploration. Where ore material
has been dispersed by several different agents, there will be hints of the ore
over a much wider area than the area of the orebody itself.
Of the several different agents causing dispersion of ore material in
the secondary environment, the most important in glaciated terrain is
glacigenic dispersion. If the Quaternary geological features of the study
area are sufficiently well known, both the direction and distance of the
glacial dispersion can be determined. Then also the source of a geochemical
anomaly — the ore deposit — is easy to locate. Likewise, anomalies due
to gravity, though uncommon, are easy to trace back to their source. The
effect of other agents — like wind and water — is much more difficult to
interpret.
There are two main types of geochemical dispersion in the secondary en-
vironment: clastic and chemical. In areas of glaciated terrain, geochemical
Conclusions 125

anomalies are usually the product of both of these but with the glacigenic
clastic dispersion dominant. Outside of glaciated terrain in temperate zone
chemical dispersion is more important.
The amount, direction and distance of chemical dispersion depend on a
variety of factors — pH, Eh, solubility of ore material, grain size, adsorption
properties, ion exchange capacity, etc. — which means that the task of the
exploration geochemist is more difficult in areas of anomalies formed by
chemical dispersion than by clastic dispersion. Some lesser known types
of dispersion such as electrochemical dispersion and dispersion in gaseous
phase are closely connected with chemical dispersion.
As a rule one needs to know the chemical and clastic contributions to
the anomaly and to have a fair knowledge of the Quaternary and glacial
geology of the study area, to be able to interpret the results correctly and
locate the source of the anomaly — the orebody.
127

Chapter 6

GLACIGENIC DISPERSION OF COARSE TILL FRAGMENTS

INTRODUCTION

Clastic glacigenic dispersion means the movement of bedrock fragments


by glacial action. The coarsest till material (cobbles to boulders) typically is
dispersed in the form of trains, fans or irregular patterns (Rose et al., 1979).
Because glacial deposits commonly mask primary ore outcrops, they have
often been considered a hindrance to exploration. From another perspective,
however, the overburden is very welcome in mineral exploration since the
dispersal patterns of an ore may form a considerably larger target area
than the ore outcrop itself.
The relationship between individual cobbles and boulders must be
established in order to interpret their dispersal pattern and determine
their primary source. With cobbles and boulders as with geochemical
anomalies, the 'tail· has to be identified and the 'head' to be found (Shuts,
1976). Boulders are useful in dispersal studies because their lithology is
easily identified in the field. Boulder tracing provides fast and reliable
information, which can be used as a basis for more detailed methods for ore
exploration.
The variable compositional relationships between glacial products and
the local bedrock have been described on many occasions (Sauramo, 1924;
Dreimanis, 1956; Repo, 1957; Dreimanis and Vagners, 1971; Hyvärinen et
al., 1973; Bradshaw, 1975; Kauranne, 1976a; Shilts, 1976; DiLabio, 1981;
Saarnisto and Taipale, 1985; Perttunen, 1989). Because of differences in
local glacial history, similar glacial sediments do not always have the same
relationship to the bedrock or mineralization (see Nichol and Björklund,
1973, Salminen and Hartikainen, 1985).
Nevertheless, some general laws of clastic glacigenic dispersion have
been formulated, and applied in determining the transport distance of till
materials (Puranen, 1988). Krumbein (1937) observed that debris dispersed
by a glacier occurs with maximum frequency close to its provenance and
in exponentially decreasing amounts in the direction of glacial transport
(Shilts 1976). The half-distance method of calculating the transport distance
of clasts of a particular lithology is based on this observation.
128 Glacigenic dispersion of coarse till fragments

Starting from the same assumption of exponential decrease Salonen


(1986) and Bouchard and Salonen (1990) have suggested a different method.
In this transport distance distribution (TDD) method, a sample of a number
of boulders (or smaller fragments) of different lithologic type is treated
instead of a single indicator rock type. The true distance to the provenance
area measured for every lithologic type upstream along the glacial flow
vector is plotted on a cumulative curve. Plotted on log-normal scale, the
distribution can usually be fitted with a straight line that allows estimation
of such transport distance statistics as geometric mean (G.M.) of the
distance to provenance for a given boulder population, and its standard
deviation (SD). The statistics are in harmony with the depositional morainic
landforms at the sample sites.
In this chapter, the factors affecting the length of the clastic dispersion
are discussed, with examples from shield areas of Scandinavia and North
America that were located in the central parts of the last Pleistocene
glaciation. The main emphasis is on the half-distance determinations of
coarse-grained till material, pebbles and boulders. In a first section the
frequency distribution parameters are defined and the factors controlling
these parameters are briefly discussed. Typical dispersal trains of boulders
are then described, and the TDD method is introduced and applied to
boulder transport in Finland. Finally, some general considerations of glacial
dispersal of boulders and the usefulness of boulders in exploration are
discussed.

FREQUENCY DISTRIBUTION PARAMETERS

From any exposure of a lithology, a certain amount of rock material


is abraded, crushed, and mixed in the existing glacial debris to form a
fixed percentage of fragments. As the fragments are comminuted along the
direction of glacial transport, the percentage of the lithology in the total
debris decreases with distance until the terminal grade level is reached
(Dreimanis and Vagners, 1971). The percentage of coarser till fractions of
the lithology correlates negatively with the transport length and positively
with the abrasion resistance of the lithology. The comminution in the
"glacial mill" is not only autogenous (depending on the lithology itself) but
also depends on the hardness of the other lithologies in the till.
Krumbein (1937) applied a negative exponential function to the distance
distribution of boulders in a boulder fan when he determined the coefficient
(a) for the dispersion caused by glacial action. The decrease in frequencies
of lithologies obeys the function:

y = yo e **
Frequency distribution parameters 129

D i r e c t i o n of g l a c i a l f l o w

x x
\ o (km) —
\ proximal contact

Fig. 6-1. Determination of the frequency distribution parameters K, MD and a for indicator
particles. The provenance rock is shown in black and the dots indicate the frequency of indicator
particles in stone counts.

(Fig. 6-1), where yo = frequency (x0), a = coefficient of particle distribution,


sndx = distance (km) from the observation point of the provenance (XQ).
The half-distance value (#1/2) is the distance to the point where the
frequency of indicator particles is half its starting value. This distance can
be calculated from the coefficient (a) using the equation:
*i/2 = 0.693/a
(Gillberg, 1965). The half-distance value can be calculated from anywhere
along the distribution curve (Fig. 6-1) if the frequency observations are
representative enough.
The transport distance K (Fig. 6-1) is measured as the displacement of
the maximum abundance peak of the distinctive rock fragments studied
from their nearest upglacier source area (Lee, 1965; Salonen and Palmu,
1989). Drake (1983) studied the K distances in till. The same displacement
has been determined quantitatively for gravity anomalies in till, and called
the "shift distance" by Puranen and Kivekäs (1979). Measurement of K is
important in mineral exploration because it is approximately equal to the
length of the geochemical anomaly in the fine fraction of till (Salminen,
1980). K is larger the more resistant to erosion is the source rock (Bouchard
etal. 1984).
The parameter MD (Fig. 6-1) shows the distance at which the abundance
of fragments of a given lithology decreases to a level of 1% (Bouchard et
al., 1984). This can be considered as the length of a boulder train and it is
related to the size of the source rock exposure and, again, to the resistance
of the rock to comminution.
The parameter a describes the dilution of a given lithology in relation
to other lithologies in the glacial transport. It depends on several fac-
130 Glacigenic dispersion of coarse till fragments

tors (Krumbein, 1937; Gillberg, 1965, 1967; Perttunen, 1977), the most
important being the abrasion resistance of the rocks and minerals to com-
minution, the topography and the mode of transport. The parameter is
thus understood as an index of the comminution process. During the glacial
transport the particles are crushed and fractured, their number increases
and, down to the terminal grade, their size and the percentage decrease.
The crushing of large fragments during glacial transport changes the size
distribution. In other words, till becomes more mature during the transport,
and the characteristic terminal grades for each lithologic type are reached
(Dreimanis and Vagners, 1971). There is a typical mode representing rock
fragments (usually granules of 2-4 mm in diameter) and another for single
mineral particles (0 < 0.06 mm). Hence the value of a decreases with
decreasing particle size (and the half-distance value increases).
To summarize, the factors controlling the frequency distribution parame-
ters of coarse grained till fragments are grain-size effect, bedrock lithology,
size of the source area, mixing with older deposits and mode of glacial
behaviour. These factors are briefly discussed in the sections that follow.

Grain-size effect
The transport distance of clasts varies with their size. The distances
of transport for different grain fractions have been studied by Szabo et
al. (1975), Perttunen (1977) and Salminen and Hartikainen (1985). It has
been shown that the contribution of the underlying bedrock to the stone
count greatly depends on the size fraction examined. In studies in New
Brunswick, Canada, Szabo et al. (1975) found a contribution of 90% at
about 13 km from the contact in the 1.9-3.8 cm fraction, but within 3 km
in the 0 > 5.1 cm fraction. Boulders from the underlying bedrock have
been found to constitute as much as 80% of the total count just a short
distance from the proximal contact (see also Dreimanis and Vagners, 1971).
Szabo et al. (1975) argued that investigation of the distribution of
indicator pebbles in drift provides the quickest and easiest way to determine
the general pattern of the transport. Earlier Repo (1957) found that the
pebble fraction reflects most clearly the general lithologic variability of the
underlying bedrock.
Boulton (1975) has described how the preferential transport occurs in
response to different debris sizes. When particles are smaller than 0.008
mm, regelation flow occurs around the individual particles, resulting in only
limited entrainment. At the opposite extreme, when particles are greater
than 16 mm in diameter, plastic deformation of ice may begin to operate,
which means that large particles will not be significantly entrained either.
On that basis, Eyles and Menzies (1983) concluded that debris ranging in
size from silt to pebbles is actively transported, and that the finest till
fraction has only a short glacial transport and reflects the local bedrock
Frequency distribution parameters 131

more clearly than the other fractions. This has also been demonstrated in
field observations. Salminen (1980) emphasizes that when the location of a
mineralization causing an anomaly is being estimated, the investigation of
the transport distance must always be conducted on the fine fraction rather
than the pebbles of the till.
Salminen and Hartikainen (1985) found further that boulders on the
surface of till blanket have been transported farther than the finer fractions
of till, which means that the source area of a surface ore boulder is usually
harder to find than that of a geochemical anomaly. However, in the core
areas of the last continental ice sheets, surface boulders in hummocky and
Rogen moraine areas have a close relationship with the underlying bedrock
(Minell, 1980; Bouchard, 1986; Salonen, 1986).
The different fractions of till thus reflect the system of glacial transport
in different ways. Geochemical anomalies best reflect the local variations
in the transport system. The length of the anomaly not only shows the
length of glacial transport but the combined dilution effect of the mixed
lithologies by comminution. Surface boulders will have been transported a
short or long distance depending on the behaviour of the glacier at that
place during deglaciation. And finally, the general lithologic variability of
the glacial transport is best preserved in the sand (mineralogic) or pebble
(lithologic) fraction.

Lithologic effect
Studies on the effect of bedrock type on till lithology (Gillberg, 1967;
Linden, 1975; Perttunen, 1977; Bouchard et al. 1984; Salonen, 1986) show
that the average length of half-distance in the pebble-boulder fraction
varies with the rock type. Limestones and quartzites have the longest
average values of half-distance, and gabbros and mica schists the shortest.
The lithologic effect in boulder transport has been verified in numerous
field studies. Holmes (1952) observed how weak black shales were abraded
to smaller size so that no fragments of pebble or larger size were observed
more than 6.4 km from the outcrop. On the other hand, the decrease
in the amount of quartzitic rocks was so gradual that they constituted a
conspicuous element in glacial drift as far as 130 km from the outcrop.
The quartzites were the toughest of the abundant drift materials. Gillberg
(1965) noted that a dilution by softer rocks, e.g., Cambro-Silurian limestone
and shale, in till will result in a lower value of (a) for more resistant rocks,
e.g., sandstone and dolerite, than for the same rocks within a gneiss-granitic
till.
In laboratory tests designed to study the abrasion and impact strength of
different rock types, Kauranne (1970) demonstrated that the most brittle
rocks are limestone, pegmatite, quartzite and rapakivi granite. According
to Swedish impact tests, the impact resistant rocks are phyllite, leptite,
132 Glacigenic dispersion of coarse till fragments

ROCK TYPE

Gabbros I · 1 (n-3)

Mica schists · 1 (n-9)

Metavolcanics · 1 (n = 8)

Granitoides · , (n = 18)

Sandstones l · 1 (n-3)

Limestones i · 1 (n-3)

Quartzites . · , (n = 7)

5 10 15 20 25 30 km
HALF-DISTANCE VALUE
Fig. 6-2. Half-distance values of different indicator rocks based on data from Gillberg (1967),
Linden (1975), Perttunen (1977), Bouchard et al. (1984) and Salonen (1986). The circle indicates
the mean value and the bar its standard deviation. Fraction: cobbles-boulders.

diabase, greenstone and amphibolite. Thus, as Fig. 6-2 shows, the impact
resistant rocks tend to have the shortest transport distances, and the most
brittle rocks, quartzites and limestones, the longest.
The observations above reveal a relationship with glacial processes.
According to Drewry (1986), hard brittle rocks such as granites and certain
metamorphic rocks are particularly susceptible to crushing, fracturing
and abrasion. Softer strata, exemplified by shales, thin bedded sandstones
and some volcanic rocks, are readily abraded and also fractured and
crushed. Very soft, argillaceous rocks like limestones, marls and slates are
dominantly eroded by the ploughing component of abrasion and hardly
affected by fracture.
Lithologic factors are important in controlling the length of glacial
transport, and, as Shuts (1976) points out, dispersion curves should always
be defined for particular rock and mineral types. The wide variation within
a single lithology (Fig. 6-2, standard deviations), particularly for granitoids
and volcanic rocks, should be noted, however.

Width of the outcrop


The half-distance value (together with other dispersion parameters) is
also related to the size of the outcrop. In a study of this relationship
Peltoniemi (1985) has shown that, for fragments smaller than boulders,
the size of the provenance outcrop is the main factor affecting the value
of the half-distance. However, analysis (Salonen, 1986) of the data of
52 half-distance traverses for the cobble and boulder fraction (Gillberg,
1965; Linden, 1975; Marcussen, 1973; Bouchard et al., 1984; Bouchard
and Martinau, 1984) gave a value of only 0.565 for the correlation logx 1/2
Frequency distribution parameters 133

Fig. 6-3. Relation between the width of the bedrock area and the half-distance value. Fraction:
cobbles-boulders. Data as in Fig. 6-2.

versus the logarithmic diameter of the outcrop (Fig. 6-3). The correlation is
quite poor: the regression function explains only 32% of the total variance.
Clearly, the larger the provenance outcrop is, the greater the value of
the half-distance can be expected to be. However, this relation is not alone
sufficient. Besides the gross length of the outcrop there are other important
physical factors relating to the provenance outcrop — factors such as the
direction of fracture systems in the bedrock in relation to the direction of
glacial flow and the topographic setting of the source area (see Chapter 5,
this volume).

Topography
The temperature of the basal ice plays a key role in determining the
erosional capacity of glaciers. The thermal boundary conditions at the
ice-bedrock interface are some of the most important in the whole of glacial
geology (Boulton, 1974). If the basal ice is cold, below the pressure melting
point, erosion and sedimentation processes may be severely affected or
totally inhibited. Under cold basal conditions, low velocities reduce the
outward flux of entrained sediments in traction of the bed or higher
level transport. In a warm-based wet glacier, on the other hand, debris
is deposited, underlying strata are abraded, and particles are crushed and
sheared.
Where the surface of the bedrock is uneven, a mosaic of secondary,
deflecting ice movements may develop at the bottom of the ice sheet, result-
ing in small-scale variations in basal temperatures. On the downstream,
freezing-on side of a bed protuberance, clastic material is entrained through
a quarrying-type of bedrock erosion. Melting occurs on the upstream side
of hummocks, leading to net deposition of debris. This process is at work
134 Glacigenic dispersion of coarse till fragments

especially during deglaciation under fluctuating basal thermal conditions


close to the margin of glaciers (Minell, 1980; Bouchard et al., 1984).
The topographic effect on frequency distribution parameters has been
studied by Gillberg (1965), Perttunen (1977) and lately by Salminen and
Hartikainen (1985). Gillberg (1967) explained a change in the gradient of
distribution curves as being solely due to topographical factors.

Mode of glacial behaviour


Sampling is the major problem in connecting distributions of transported
particles with glacial activity. Glacial transport is typically investigated
using long traverses in the direction of the glacial flow. However, the glacial
environment changes, even over short distances, owing to the thermal
mosaic prevailing during the deposition, so that samples from different
sites may not easily be seen to be connected with the same depositional
environment.
Samples may also represent materials affected by more than one glacial
cycle. In measurements of half-distance on samples taken by tractor
excavator from identified stratigraphic units, Hirvas et al. (1977) observed
till beds with different lengths of glacial transport. Clearly, the regularity
in the glacial dispersion will be disturbed if different till units are mixed in
sampling.
The relation between glacial deposition and length of half-distance
values has been studied by Salonen (1986). The largest values of xi/2 were
associated with drumlin areas, and small ones with boulder-rich hummocky
areas. A mixing effect was found on an observation traverse extending
from a hummocky moraine area to a drumlin field. The gradient of the
distribution curve changed so that the value of xi/2 was larger in the
drumlin area than in the hummocky moraine area.
It can be concluded that the local mode of glacial behaviour affects
the frequency distributions. As an example, Salminen and Hartikainen
(1985) found that the transport distance was conspicuously different on
the proximal and distal sides of the big Salpausselkä marginal formation in
Finland. The ice lobe was active on the proximal side of the formation, and
during the deglaciation a partition in the glacial dispersion configuration
developed; as a result, dispersal patterns became long on the active proximal
side of the formation and short on the distal periphery.

General conclusions about glacial transport distance


Despite the many different factors affecting dispersal curves, some
general conclusions about the length of glacial transport can be drawn. The
relation between transport distance and grain size has been documented
in many studies (e.g., Kauranne, 1976a). Compiling data on the length
Dispersal trains 135

M - 4 . 1 0 km
n = 52

0.5 1 2 3 4 5 10 20 3040 50 km
VALUE OF H A L F - D I S T A N C E

Fig. 6-4. Frequency distribution of half-distance transport for cobble and boulder fractions. The
half-distance varies from 0.2 to 50 km.

of geochemical anomalies in the finest fraction of till, Salminen and


Hartikainen (1985) found the median length of 34 anomaly patterns to be
only 1.0 km. For the gravel fraction Gillberg (1967) calculated an average
half-distance value of 10.0 km from the arithmetic mean of 110 distribution
lines. He regarded the measure as an approximate value of the dispersion
constant of till. For boulder and cobbles (Fig. 6-4) there is information
on 52 half-distance traverses (Fig. 6-2). The variability is great, but the
average value (4.1 km) clearly lies between the average values for the fine
fraction and gravel.

DISPERSAL TRAINS

Dispersion parameters (a = coefficient of particle distribution, MD =


distance to the 1% level and K = displacement of maximum abundance
peak) describe the regularities of glacial transport. Though partly theo-
retical, they can be related to the observed properties of dispersion trains
of boulders (Fig. 6-5). The apical boulder is the first observed (surface)
boulder downstream from the source. It is located close to the point of
the maximum frequency (transport distance K), especially when a is high
in value. The parameter a bears a close relationship to the length of
boulder trains (Salonen, 1986), and MD indicates the distance at which the
influence of the studied lithology in till is no longer of importance.
Hirvas (1980b) introduced the concept of "topographic control" in glacial
dispersal studies. In areas where there are two till beds of different ages,
the older bed occupies the lower level. Moreover, dispersal trains of younger
beds have their provenance in topographic highs, and those of older till
136 Glacigenic dispersion of coarse till fragments

Length of boulder fan

Edges of boulder fan

Fig. 6-5. Schematic illustration of an indicator train.

beds originate from lower parts of the terrain. In the study of the pattern
of dispersal of granite blocks in Gaspe, Quebec, David and Bedard (1986)
observed that those lithologies that formed positive relief and projected
well above the general level of the surrounding ground surface were eroded
most and transported farthest.
Dispersal trains are typically finger- or ribbon-shaped (Shuts, 1976).
They are narrow, and often long in mountainous areas and where ice flow
has occupied tectonic trough valleys. Fan shapes may result where trains
are formed with radical spreading in lowlands or near an ice front. Multiple
transport also tends to create wide-angled boulder fans. Dispersal trains
are commonly narrow and short close to the marginal zone of the retreating
glacier.
Boulder fans characteristically coincide in orientation with the last or
most pronounced direction of glacial flow, occupying a sector of approxi-
mately 10°. But this angle may be considerably wider if the fan is a product
of a number of ice flows of varying directions; fans opening up to angles
as much as 90° have been described (Hyvärinen et al., 1973). Close to
ice divide zones, as in Finnish Lapland (Tanner, 1915) and in the Gaspe
Peninsula, Canada (David and Bedard, 1986), the dispersal is often the
combined result of many shifting consecutive glacial phases.
Boulder fans are normally some 1-5 km in length, but variation is wide.
The longest examples described to date measure 50-100 km, though some
individual boulders appear to have been carried as much as several hundred
kilometres from their place of origin by glacial action. At the other extreme,
one finds local boulder trains that are only some tens or hundreds of metres
in length.
Transport distance distributions 137

Nearly 500 boulder dispersal trains have been identified in Finland


(Salonen, 1986). The median value reported for the length is 3.0 km, but
there is also a minor mode comprising dispersal fans 40-100 km in length.
These lengths must only be considered as informative, however, since there
is always a subjective element involved in the measurement. Accurate
determination of the dimensions of boulder trains is not possible, although
semiquantitative data are available on uranium boulder trains (Gustavsson
and Minell, 1977).
The distance of the apical boulder of a fan from the proximal contact
of the outcrop depends on the thickness of the till cover (Fig. 6-5): the
thicker the till cover, the farther the apical boulder from the outcrop (see
Puranen, 1988). The location of the outcrop can be estimated from the
angle of climb, which is about 10° for basal lodgement till but only 1-2° for
drumlin areas. In other words, the average distance from apical boulder to
proximal contact is 50-200 m in areas of basal tills, and 500-1,000 m in
drumlin areas.
For a single basal lodgement till bed, the rule is that the local material
lies below. Even an overlying layer of ablation mantle does not alter the
situation: if it is supra- or englacial melt-out till it usually has a long
history of glacial transport behind it. In the central parts of continental
ice, however, the situation is not always so straightforward: in the areas of
Rogen and active hummocky moraines, the surficial material may be more
local than the deeper material (Minell, 1980; Bouchard, 1986).
The angle of climb is equivalent to the angle of climb in ripple laminated
bed resulting from fluvial sedimentation. Angle of climb is the ratio of the
rate (flux) of transport to the rate (flux) of sedimentation-accumulation.
Investigation of the behaviour of this parameter in till may be the most
interesting area of future glacial dispersal research (see Puranen, 1988).

TRANSPORT DISTANCE DISTRIBUTIONS

In 1961 Matisto determined lateral limits for each bedrock type up ice
with respect to glacial flow and with this technique discussed the effects of
complex transport. Salonen (1986) discerned a negative exponential func-
tion (Fig. 6-1) in the form of cumulative frequency curves. In this transport
distance distribution (TDD) method, a sample of boulders comprising many
lithologies is investigated, rather than a single indicator rock type, and the
transport distance is measured in the opposite way than in the half-distance
method. In the TDD method the many lithologies of one sample are con-
nected with up-ice bedrock sources, whereas in the half-distance method,
many samples of the same lithology are identified and located downstream
in the glacial flow (Fig. 6-6).
138 Glacigenic dispersion of coarse till fragments

B e s t - f i t t i n g curve for midpoints


of provenance areas

- 1 0 1 2 log km
x - a x i s : Distance to provenances
Fig. 6-6. Method of estimating the transport distance statistics GM (geometric mean) and SD
(standard deviation) using log-normal probability paper.

The requirements for using the TDD method are:


— Good knowledge of the local bedrock: the provenances of all or nearly
all of the boulders in the sample must be exactly located.
— Suitability of bedrock: pronounced lithologic variations perpendicular
to the ice flow direction. Too high lithologic contrasts must be avoided (e.g.,
quartz ites/limestone).
— An abundance of surface boulders.
If these requirements are met, the proportion of each bedrock type
in the boulder fraction can be plotted on log-normal probability paper,
where the y-axis is the cumulative frequency and the x-axis is the distance
to source (Fig. 6-6). The transport distance distribution (TDD) of glacial
boulders can normally be depicted with a straight line, and transport
distance statistics GM (geometric mean) and SD (standard deviation) can
be estimated graphically (Salonen, 1986). The geometric mean is distance
/too and standard deviation {η%± — nie)/2, where nx is the xth percentage on
a cumulative scale (Sinclair, 1976).
The TDD method allows an individual boulder to be connected with a
specific glacial transport population. Estimation of the probable transport
characteristics for the population associated with an ore boulder — trans-
port distance (GM), transport distance K, and maximum distance (MD)
— can be of crucial importance in local scale exploration (Salonen, 1986).
This information is also important when interpreting the source of a till
anomaly.
A further advantage of the TDD method is that the size of provenance
outcrop and the bedrock type are not as important factors as in the
half-distance method. With the TDD method the most important factor
controlling the variation in transport distance statistics is the glacial
behaviour (Salonen, 1986). From this it follows that the statistics are
dependent on the different glacial geomorphologic landforms. In field
measurements the geometric mean for the transport distance was found to
vary between 0.4 and 3.0 km for active ice hummocky moraine areas, and
Conclusions 139

between 0.8 and 10.0 km in cover moraine areas. For the surface boulders
of drumlins, the geometric mean was 5-17 km.
Mixed transport populations have been encountered in areas of thick till
beds especially in northern Finland. Typically the deposits consist of one
or more lodgement till beds. Rock material may have been transported by
several glacial cycles and the transport distance (GM) of surface boulders
may vary within a wide range (1-25 km) (Salonen, 1986).
The observations on the transport distance distributions of surface
boulders show that there are two major glacial processes whose possible
existence needs to be kept in mind when studying the length of boulder
transport: the till renewal process and the intermixing of materials from
earlier deposits. The first is important when the ice sheet behaves as
a slow-moving conveyor system, eroding the bedrock during deglaciation
and transferring it. The process is controlled by lobate flow, in which
rests of the ice sheet and various flow regimes are interchanging. In
this situation, boulders tend to form short and narrow indicator trains.
A variable proportion (approximately 1-20%) of boulders is indicative of
the second process — intermixing of materials of earlier glacial cycles.
These boulders may have a long and complicated transport history which
is difficult to unravel. Multistage transport is common especially in areas
where the ice was stagnant during the deglaciation (Punkari, 1984), and
where till stratigraphy tends to be complex (Hirvas and Nenonen, 1987).
There is information on almost 500 boulder trains in Finland, and,
boulder transport has been measured in 41 half-distance traverses and 111
TTD determinations (Salonen, 1986). On the basis of these data, the known
areal variability of glacial deposition (Kujansuu and Niemelä 1984) and
general trends in till stratigraphy (Hirvas and Nenonen, 1987), it is possible
to summarize the dispersion for the boulder fraction as shown in Fig. 6-7.
In the dark grey areas boulder trains tend to be short and narrow. These
areas are most suitable for indicator tracing. In the light grey areas the
lobate flow pattern predominates in clast transport and boulder trains may
be long, but narrow, and strictly follow the direction of glacial flow. In the
white areas with short average transport distance local material dominates
in surface boulders, but part of them may have been transported by earlier
glacial flows. In Finland, this kind of transport distribution is mostly found
in the north, where it is associated with the ice divide zone. White areas with
long average transport distance are those of complicated boulder transport
Complex stratigraphy with multistage glacial transport characterizes the
till material. Boulder tracing is difficult and many valuable ore boulders
have never been connected to their mother lode. In several cases a follow-up
study with geochemical and mineralogical methods has led to a positive
result (e.g. Hyvärinen, 1969).
140 Glacigenic dispersion of coarse nil fragments

LEGEND

Glacier flow line, and


average transport distance

0 50 100 km

Fig. 6-7. Prediction of the average length and dispersion (mixing rate) for boulder transport in
Finland based on transport distance parameters.
Conclusions 141

CONCLUSIONS

The parameters describing the glacial distribution of till material are


controlled by such factors as the type of indicator rock, the size of the
provenance area, local topography and the grain size. Shortest glacial
transport is associated with the fine fraction of till, the average length
of geochemical anomalies being just 1.0 km (Salminen and Hartikainen,
1985). The transport distance of the pebble fraction is longest; using the
half-distance method Gillberg (1967) suggested an average of 10.0 km. In
the case of the boulder fraction, Salonen (1986, 1987) found an intermediate
distance of glacial transport; depending on the method of measuring, the
median for this fraction varies from 3.4 to 4.1 km. Thus indicator particles
of the same lithology are transported different distances in different grain
size fractions.
Shilts (1984a) emphasizes that the dispersal trains in glaciated terrain
are overlapping and they occur at a variety of scales:
— continental scale (100-1,000 km)
— regional scale (10-100 km)
— local scale (1-10 km)
— small scale (0-1 km).
The scale, and thereby the method of study of clastic glacigenic disper-
sion, need to be selected according to the purpose of the study. Continental
scale questions of glacial dynamics can be investigated by observing errat-
ics and dispersal trains over large, distinctive provenance areas (cf. Shilts,
1984b; Donner, 1986; Prest and Nielsen, 1987). Regional scale problems,
like the identification of glacial lobes and their flow configuration, can be
studied with pebble counts (Repo,1957; Shetsen, 1984). In that connection
the number of counting sites needs to be really representative and great
enough to allow multivariate statistical analysis.
At the local scale, which is the scale of ore boulder prospecting, the
explanation of dispersal patterns and search for the primary outcrops call
for detailed boulder tracing and half-distance determinations. In areas
where the bedrock is well known, the transport distance distribution (TDD)
method can be used to obtain a general view of the boulder transport, from
which the glacial transport of single indicators can be deduced.
Small-scale dispersal studies require semiquantitative boulder tracing
supported by till geochemistry and geophysical measurements. Estimating
of K values helps in pinpointing the locality of the sought for provenance.
Explaining the genesis of glacial deposits and determining the angle of
climb are important tasks in this final phase of boulder tracing as well as
in till geochemistry.
143

Chapter 7

SCALE OF GEOCHEMICAL SURVEYS

GENERAL

Geochemical studies vary enormously in area. At one extreme geochem-


ical maps are prepared for whole countries based on a couple of thousand
samples, while at the other highly detailed maps based on several thousands
of samples are drawn for the vicinity of a narrow dike.
Studies at different scales differ considerably in the way they are carried
out. Not only sampling density, but sampling material, sampling depth,
sampling equipment, analytical methods and methods of data processing
essentially depend on the aim of the study, the size of area to be studied,
the objects to be recognized and the contrast between the anomaly and
the surrounding area. The sources of the anomalies recognized by different
sampling densities may be of totally different character.

SAMPLING MATERIAL

Minerogenic stream sediments are the traditional sampling medium in


small-scale regional geochemical mapping. In areas of residual overburden
they have proven to be extremely useful, providing data from a wide
drainage area where the stream has been in contact with the bedrock.
Geochemical mapping based on stream sediments is as its most effective
in areas of temperate climate where rivers are draining in situ weathered
bedrock. It is also very effective in mountainous areas where the bedrock is
widely exposed.
In glaciated areas the stream is usually disconnected from the bedrock
and from possible mineralization by till and the stream sediments mainly
reflect the variations in metal contents in the underlying till. The interpre-
tation of results is difficult, and the information level lower. Despite of that,
in till-covered mountainous areas with steep topography, as in Scotland
(Plant et al., 1984) and Norway (Wennervirta et al., 1971), quite useful
results may be obtained by stream sediment geochemistry.
Till has conventionally been exploited as a sampling material only
144 Scale of geochemical surveys

in local-scale exploration studies. In the case of regional-scale studies,


the greater cost of such sampling has encouraged many to imagine that
stream sediment could be as informative in areas of glaciated terrain as
in areas of residual overburden. Now, however, results from Scandinavia
— the Geochemical Atlas of Finland and the geochemical mapping of the
Nordkalott project (Bölviken et al., 1986) — show beyond doubt that
highly informative and easily interpreted results can be obtained from till
geochemistry practised also on a regional or a reconnaissance scale.

SAMPLING DENSITY

Ginzburg (1960) divided geochemical mapping into three classes accord-


ing to scale: (1) reconnaissance, (2) prospecting, and (3) detailed mapping.
Rather than specifying sampling densities, Ginzburg described the phases
logically included in geochemical studies: the reconnaissance phase, which
reveals the areas where geochemical studies can reasonably be carried out,
the prospecting phase, which reveals mineralized areas and zones where
economic ores can be supposed present; and the detailed phase, which
reveals ore suboutcrops.
A decade after Ginzburg, Bradshaw et al. (1972) proposed a division
of geochemical mapping into three classes specifically according to the
sampling density: (1) one sample/40-80 sq. miles (about 150 km 2 ), (2)
one sample/5-20 sq. miles (about 30 km 2 ), and (3) one sample/1-2 sq.
miles (about 4 km 2 ). In their view, each class is specific for a certain type
of geological formation and thus reflects different geochemical features.
Probably they are right, though even the densest of their sampling networks
will not unambiguously reveal a single mineralization or ore when till is
the sampling material. Conversely, a much lower density than one sample
per 40-80 sq. miles can produce much very useful regional geochemical
data from till.
The sequence of geochemical exploration developed in Finland during
recent years likewise consists of three phases (Salminen and Hartikainen,
1986). The regional phase, where the sampling density is 1 sample/4 km 2 ,
corresponds to Ginzburg's reconnaissance phase. The anomalies detected
in this phase are typically hundreds of km 2 in extent. In the next phase
— the local phase — the sampling density varies from 10 to 30 samples/
km 2 , and mineralized areas and zones several km 2 in size are delineated.
The final and most expensive phase is the detailed study phase aimed at
finding anomalies of some hundreds of m 2 in area where the sampling
density may be as high as 400 to 1000 samples/km 2 (Fig. 7-1). Additional
to these three phases, a geochemical survey of the whole of Finland (Fig.
7-2) has been done using a sampling density as low as 1 sample/300
km 2 (Geochemical Atlas of Finland). Even such very low density sampling
Sampling grid 145

REGIONAL PHASE

LOCAL PHASE

DETAILED PHASE

0.01 0.1 1 10 100 1000 Sampling density


sites/km 2
2
Fig. 7-1. Sampling density (sampling sites/km ) for different scales of geochemical studies.

can produce coherent geochemical information which is of great value for


the exploration manager of a mining company planning programmes and
defining areas — though not of any great value for the prospector.

SAMPLING GRID

The selection of the sampling grid depends to a large degree on the


expected character of the anomaly and the scale of the study. Traditionally
the sampling grid in glaciated terrain was designed on the assumption that
geochemical anomalies are spread by the glacier to the distal side of the
source area. Further, it was supposed that the anomaly is totally created
by glacigenic clastic dispersion and that the contribution of chemical
dispersion — both before and after the glaciation — is negligible.
In the ideal case of pure glacigenic clastic dispersion, the geochemical
anomaly should rise with a gentle slope, gradually weakening up to the
surface of the overburden (Halonen, 1967; Kauranne, 1958; Nurmi 1977;
Salminen and Hartikainen, 1985). The three-dimensional shape of the
anomaly resembles a drop or a plume, as Drake (1983) has described it (Fig.
7-3); it is elongated in the transport direction of the glacier, and a surface
section is ellipsoidal (Fig. 7-4). In the ideal case then — if the transport
distance is known — it should be possible to calculate mathematically a
sampling grid, or net, that will include a sufficient number of anomalous
samples to define the anomaly (Savinskii, 1965; Gustavsson, 1983). In this
type of grid the sites are situated along lines transverse to the direction of
ice movement and to the longitudinal direction of the geological formations.
The distance between lines is usually many times greater than the distance
between the sampling sites along the line.
The mathematically optimum grid can be successfully applied only when
the size and shape of the suboutcrop are known, the direction of glacial
transport was linear and the transport distance was considerable (at least
146 Scale of geochemical surveys

Fig. 7-2. Distribution of aqua regia soluble barium concentrations in the <0.06 mm till fraction,
Finland. Reproduced from the Geochemical Atlas of Finland.
Sampling grid 147

Fig. 7-3. The three-dimensional shape of an anomaly resembling a smoke plume in the
overburden. Reproduced with permission from Drake (1983).

Fig. 7-4. Three-dimensional pattern of a glacially dispersed anomaly in the overburden.

a few kilometres). Such conditions are met in areas where ice lobes were
active during the deglaciation. However, there are many areas where
anomalies have no clearly defined length, e.g., areas where the glacial
erosion was weak and the transport of material slight. Anomalies are then
local, and it is more reasonable to use a sampling grid that is regular in
every direction.
148 Scale of geochemical surveys

The transport distance is shorter at the bottom of a till bed than at the
surface. Also, the metal and mineral concentrations derived from an ore
suboutcrop increase in the deeper parts of the overburden. Thus a sampling
grid regular in every direction will be preferable for deep sampling.
The design of the sampling grid will be more complicated when two or
more glaciers have transported the till. If the transport directions have
differed and if separate till beds have been laid down, sampling must be
done from different depths at every sampling site. Chemical dispersion
in the till may also reform the anomaly. In cases like these, preliminary
studies are necessary before the final form of the sampling grid can be
decided upon.
Besides the way in which an anomaly was created and the conditions
prevailing where it was formed, the scale of the study has a decisive
influence on the sampling grid. In regional-scale geochemical mapping
(Fig. 7-1) there is no reason to use sampling lines. The distance between
two neighbouring lines would, in any event, be so great that even the
longest glacigenic anomalies would fit between the lines. Also, for reasons
of economy, the sampling sites should be located in accessible places such
as along roads. In regional-scale studies, the optimum sampling grid is an
irregular one (Fig. 7-5A) with variable distances between sites.

REGIONAL PHASE

B
\

Road
\ \
X
V^
, χ

I
x""

1 km
/

Fig. 7-5. Sampling patterns in regional phase geochemical studies. (A) Irregular sampling grid;
(B) irregular sampling, where the sampling density is controlled by the 4 km 2 grid.
Regional geochemical mapping 149

As the sampling density increases, more attention must be paid to


the regularity of the sampling. Cost considerations still demand irregular
sampling, but the study area may now be divided into cells, with one
sampling site chosen from each cell (Fig. 7-5B). For the purpose of locating
the actual sampling site inside the cell, geographic features (the existence
of roads, streams, ease of access to the site) as well as geological factors
should be taken into account.
In local-scale and detailed studies the sampling grid should always
be regular. Because the distance between neighbouring sampling sites is
usually only tens of metres and never more than a few hundred metres,
determining the exact coordinates of a site is very important. A reasonable
way is to determine the coordinates of every sampling site beforehand. The
sampling grid will then be exactly as was planned.
In principle, a regular sampling grid can be of two different types: (1) a
grid regular in every direction, and (2) a grid in which the distance between
sampling sites along lines is much shorter than the distance between lines.
Choosing the most economical sampling grid and sampling depth demands
some preliminary study. Factors that should be taken into account include
the direction of the ice movement and the ice transport distance, the
strike of the bedrock, especially in schist areas, and the stratigraphy of the
overburden.
If preliminary studies indicate that the dimensions of an anomaly are
small and that the variations in different directions will not be great
(usually when the till and the geochemical anomaly are local), the use of
a sampling net regular in every direction will give the best results. In
areas of long glacial transport (active ice lobes) the use of sampling lines
will be more economical. However, if the distance between lines is many
times greater than the distance between the sampling sites along the lines,
problems may arise in the interpretation of results. It may be difficult to
connect anomalous points on neighbouring lines with each other.
Many of the computer programs available for drawing derivative anomaly
maps give a successful interpretation only if the sampling grid is regular
in every direction. If faulty interpretation in the early phase of the work
leads to the location of follow-up studies in the wrong area, the costs could
be great.

REGIONAL GEOCHEMICAL MAPPING

Mineral stream sediment, which is commonly used as sampling material


in continental Europe and the United Kingdom, is not the best choice
in glaciated terrain. In areas of residual overburden, streams are usually
in direct contact with the bedrock and the mineralization in it, but
in areas of glaciated terrain this contact is broken by till and other
150 Scale of geochemical surveys

sediments transported in from elsewhere. The mineral grains and the


metal ions adsorbed on the mineral grains in till and in glacial and
postglacial sediments will often have been transported long distances in
several separate phases. The material of till is more local than stream
sediments, and in composition it resembles the bedrock more closely than
stream sediments do. Thus to recognize the type of bedrock that caused an
anomaly is easier in till than in stream sediment. In stream sediments in
areas of glaciated terrain the anomaly/background contrast is weakened,
and the results are less easily interpreted.
Till was first used as a sampling medium in regional-scale studies in
Finland as early as in the 1950's (Tavela, 1957) and in country-wide
geochemical mapping in the early 1970's (Kauranne, 1976a). In Canada
till was used somewhat later as sampling material in some regional-scale
prospecting projects for reconnaissance purposes (Shuts, 1976). A highly
ambitious regional-scale till study, and one which produced positive results,
was the recent mapping of the whole of northern Fennoscandia in the
Nordkalo tt project. Six different materials were collected with a density
of only 1/30 km 2 , all samples being composite samples. This project was
a joint effort of the Geological Surveys of Norway, Sweden and Finland
(Bölviken et al., 1986).
Till showed to be the best sampling medium in geochemical studies
over wide areas. Even extremely low density sampling revealed large
units in the bedrock if the contrast with the surroundings was great
enough. Sampling done for the Geochemical Atlas of Finland (sampling
density 1/300 km 2 , composite samples from the surface part of till), for
instance, detected rapakivi massifs tens of kilometres in diameter (Fig.
7-6). Likewise, it revealed the existence of geochemical anomalies hundreds
of square kilometres in area, reflecting both the lithology of the bedrock
as well as the tectonic features in it. These anomalies in till extend
across the contacts of lithological units and even across the contacts
between geological formations of different age, reflecting the fracture and
shear zones of bedrock, and possibly zones where ore formation is taking
or has taken place. In certain cases they may even reflect geological
formations deeper in the bedrock in the same way as anomalies in humus
do (Antropova, 1975), or as hydrocarbons in the overburden (Malmqvist
and Kristiansson, 1984; Cazalet, 1982; McCarthy et al., 1986) may do.
Because fracture and shear zones have, as a rule, been active for a long
time (even on the geological time-scale) and as such will have favoured
dispersion in gas and liquid phases in the area, anomalous amounts of metal
ions can be expected in solution or crystallized on shear planes and other
suitable surfaces. However, it may also be that in these geologically active
zones chemical dispersion in gas or liquid phase has occurred in postglacial
time as well. Indeed, dispersion may be a continuing phenomenon.
Regional geochemical mapping 151

100 km

Geological Survey of Finland


Geochemistry department

Y
I I I I 1 I I I
$4:
N
.p». 01 σ> - i ontog O
Y (ppm)
/·♦ ·
fi · · ·
mm · . .
,#···.
>ß · · · # ·
• ·. · · · £ * · · .

§· *···♦· · · ···>
i ············· ·· ·^ ·· ·ν

Fig. 7-6. Distribution of yttrium in the <0.06 mm fraction of till, sampling density 1/300 km 2 ,
Finland. The highest yttrium anomalies reflect areas of rapakivi granite. (From the Geochemical
Atlas of Finland.)
152 Scale of geochemical surveys

Regional geochemical mapping reveals large areas in the bedrock dif-


fering from each other geochemically. Although discrete mineral deposits
are not recognized at this scale, wider zones where the mineralization is
more common than in the surrounding areas are well reflected. The main
sulphide ore belt cutting through central Finland and extending over the
Gulf of Bothnia to Northern Sweden contains an appreciable number of sul-
phide ore bodies and smaller mineralizations (Kahma, 1973a, 1973b). This
zone appears as a striking anomaly in the maps of the Geochemical Atlas of
Finland (Fig. 7-2). There had earlier been considerable argument about the
existence of this zone because no clear geophysical (before newly done deep
seismic soundings) or even geological indications could be demonstrated.
However, the geochemical till study, carried out using very low density
sampling, has proved that the main sulphide ore belt does exist as a distinct
chemical zone in the Earth's crust.

Sampling for regional-scale studies


The sampling grid in regional geochemical mapping is widely spaced
— typically one sampling site per several km 2 . The choice of the sampling
density will be determined by the size, the topography and the road system
of the study area and, of course, the available funds. Conversely, the geology
of the study area, the stratigraphical features of the overburden and other
such factors are of no great relevance for the choice of the sampling density
and material.
As the surface part of the till blanket has usually been transported
farther than the lower parts, and as such represents a wider source area,
sampling needs to be concentrated at the surface (0.5-1.0 m) in regional
geochemical mapping. To eliminate the influence of postglacial chemical
dispersion and soil-forming processes like podsolization on the anomaly
patterns, the sampling medium should be till that has been altered as
little as possible, preferably occurring under the ground-water table. To
minimize costs, sampling sites should be easily accessible. Very suitable
to the purpose are sampling sites along roads or, in roadless areas, along
the shores of rivers and lakes. The features of the separate sampling
sites (topography, moisture of ground, stratigraphy of the overburden, etc.)
should be similar. For this reason, sampling from a very thin overburden,
where the influence of the local bedrock may be overstressed, should be
avoided. On the other hand, results for samples from exceptionally thick
overburden may be difficult to interpret. To smooth the local differences it is
always reasonable to use composite samples made up of several subsamples.
Depending on the sampling density, the distance between the subsamples
may vary from tens to several hundreds of metres.
Regional geochemical mapping 153

Analysis for regional-scale studies

Because samples will usually be taken from the surface layer of till which
has been transported a longer distance, and where the metal concentrations
will be considerably diluted the range of concentrations is smaller than
in the bedrock of the source area. The analytical method should therefore
be sensitive, accurate and precise enough to detect the relatively weak
contrast between the anomaly and the background. The aim of regional
mapping is not to search for a specific type of mineralization but for all
potential mineralized areas and targets, and to determine the variations in
the regional distribution of elements. Thus, the analytical method should
allow measurement of as wide an elemental spectrum as possible: major,
minor and trace elements — reasonably 30-40 elements from every sample.
Often more than one analytical method must be used, with a consequent
high analytical cost per sample; this is offset, however, by the relatively
small number of samples. Hence the analytical cost per km 2 is modest.
In addition to ore elements, the most useful elements to determine in
regional geochemical mapping are pathfinder elements occurring in close
association with the elements of interest but more easily found because
they form a broader halo and are more readily detected by analytical
methods. Examples are arsenic for gold and sulphur for sulphide ores.
Another group of elements important to assay in the regional phase are
key elements or key element parageneses, which are typical for certain
ore types. Their concentrations are high in association with ore bodies but
elsewhere quite low; their dispersion halo may be wider than that of the
metal being sought and their range in the sampling material is usually
quite wide. Unfortunately many of them are quite difficult to analyze at the
concentration level in which they occur in drift. Most important of these
key elements are boron, barium, bismuth, antimony and phosphorus.

Data processing for regional-scale studies

Statistical methods and computing power are indispensable to handle


the large amounts of data generated in regional-scale geochemical studies
where the concentrations of as many as 50 elements may be measured from
every sample. The results are usually presented as concentration maps
with one to three elements in one sampling material on a single map. The
concentrations are classified and symbols specific for each class are marked
on the map (Fig. 7-7A). A system has also been developed where the size of
symbol (a dot) exactly represents the concentrations of a sample (Björklund
and Gustavsson, 1987). This type of map allows the geologist freely to
interpret the results in a manner appropriate to his use.
Maps may also be drawn with some computer interpretation of the data
already done. From the original data a new, denser observation net can be
154 Scale of geochemical surveys

llomantsi
Au in till
f r a c t i o n - 0 . 0 6 mm
GAAS-analysis

■ 1 ^ i '
Au ppb
·····# ··
'..... 1.0
·..... 2.0

. · · '...._ 8.0
: . . ; · . · •--16.0
• · · . •....32.0

a-.:·::; · · · · _64.0
7m. KUJTT.ILA .

:fc ■..■:.:
• Pulkir^arkat
5 km

31 10'

llomantsi
Au in till
f r a c t i o n - 0 . 0 6 mm
AAS-analysis

Γ
3 1 " 10'
Fig. 7-7. A. A geochemical map with concentrations divided into different levels and each level
indicated by a dot. B. Same data as in A, but the interpolated data are presented as grey tones.
Local scale studies 155

interpolated and maps produced as coloured surface or grey tone presen-


tations (Fig. 7-7B). With statistical methods like factor analysis, principal
component analysis and discriminant analysis, the correlations between
elements can be visualized, the paragenesis of some specific elements may
be discovered and the data of several elements can be combined in one
map. The variations in metal concentrations with geological background
can be eliminated by measuring the deviation of a single result from the
moving average or median; these residual anomalies often are indicative
in exploration. Innumerable methods of visualizing geochemical data exist
— the problem for the geologist is to identify the most illustrative method
for his specific task.
It needs to be stressed once again that regional geochemical mapping
does not aim at revealing any single ore or mineralized occurrence. Only
wider zones and provinces containing mineralized formations that are
worthy of more detailed study are identified. The geochemical maps of large
areas reflect the lithology of the bedrock and some tectonic features in the
bedrock. A natural way to proceed with small-scale geochemical mapping is
first to get data from areas where more detailed geological or geophysical
data are not available. However, as the Geochemical Atlas of Finland has
shown (Fig. 7-2), it is also possible to obtain new knowledge from areas
that are geologically already well known.

LOCAL-SCALE STUDIES

Local scale geochemical research is aimed at supplementing information


obtained from earlier regional geochemical mapping, by other geological
methods, or from glacial erratics of ore material. In contrast to the regional-
scale study, in planning and carrying out local-scale geochemical studies
one must always be aware of the exact goal, for example the discovery of a
specific ore type or only the delineation of a potential ore area.
Throughout the world unconsolidated overburden has been the most
common sampling medium in this phase of geochemical studies. In areas
of residual soil, stream sediments can be used as well; however, because
the sampling sites are bound to the stream network, the sampling density
seldom is adequate. The sampling density should in this phase be 10- to
100-fold compared with the regional phase (Fig. 7-1).
Relative to the area covered, costs will inevitably be quite high in local-
scale studies. A typical sampling group (usually two men and one drilling
machine) can take till samples from 1.5 m depth from 10 sites during a
working day. Proper selection of the sampling density and delimiting of the
study area are thus of considerable importance, making it wise to carry
out preliminary studies, determining the direction and distance of glacial
transport and clarifying the main features of the soil stratigraphy. The
156 Scale of geochemical surveys

sampling can then be effectively designed on the basis of these data, and
on any information on the bedrock, information from earlier geochemical
studies, and geophysical measurements.
The main aim of the local-scale geochemical study is to identify areas
where the occurrence of ore deposits is a strong possibility. Areas to be
delineated are usually of several hundred metres wide. Follow-up studies
— further geochemical sampling, digging of traverses or diamond drilling
— can then be concentrated in these areas.
In addition to the analytical data, stone counts in the coarser fraction
(e.g., >4 mm grains) of the samples easily provide information about the
lithology of the bedrock; they may even give evidence of ore minerals and
ore mineral associations (Fig. 7-8). The heavy mineral study of soil is often
combined with this local-scale phase of geochemical study.
If the surface of the bedrock and the material of the overburden are
not too strongly weathered, a good picture of the lithology and structure
of the underlying bedrock is obtained on the basis of the analytical and
mineralogical data of the overburden. In principle, all the chemical changes
that have occurred in the bedrock are reflected in the overburden. Further,
transport of the material either mechanically or chemically in solution will
have diluted element contents and changed the original element ratios.
However, if the samples are taken from the bottom part of the overburden
— as they should be in local surveys — the contents will not be excessively
diluted nor will the material be mixed too much.
The anomaly patterns of local phase studies mainly reflect the lithology of
the bedrock (Fig. 7-9A and C), but quite commonly they also extend across
the contacts of different rock types. Some anomalies may be associated
with fracture zones where metal ions are easily able to migrate as fluids,
in water solution or in gas bubbles, up the contacts. Others may be due to
metasomatically altered aureoles around ore bodies (Fig. 7-9B). In this way,
information is obtained not only about the surface of the bedrock but about
deeper parts, and about the chemical changes that have occurred there.
The best results in local-scale geochemical studies are obtained by com-
bining geochemical, geophysical and geological methods in the particular
manner the case demands.

Sampling in local-scale studies


Whereas in regional-scale studies the sampling pattern tends to be
irregular, in local-scale geochemical studies it should be as regular as
possible — preferably a rectangular grid equally spaced in either direction.
If the transport of material, for example by glacier, was long, this can be
taken into account by making the distance between sampling sites longer
in the direction of transport (Fig. 7-10). The data processing is easier,
however, if the sampling intervals are all equal. Because the number of
Local scale studies 157

| | Mainly archean granitoids

III! Mainly archean schists

— The boundary of schists and granitoids

The boundary of schists and granitoids uncertain

The percentage of schists in the 2-20 mm till fraction > 25 %

Site of stone count (50-100 stones or boulders)

Fig. 7-8. The percentage of schist pebbles in the 2-20 mm till fraction in the Korentovaara study
area, Ilomantsi, Finland compared with the lithology of the bedrock. Bedrock after Lavikainen
(1973). Modified after Salminen and Hartikainen (1985).
158 Scale of geochemical surveys

•S·*.·»«
Ilomantsi

·#¥· » ··■··. . . · . · . .Tl


Ni in till
f r a c t i o n < 0 . 0 6 mm
OES-analysis

ih Ni p p m
;iir;;:-;:-;ji
fe-sa?-···*··:·:
A ··::*:::::,
• · ·· ■ · · · · ·

62°_
45' ::::i:?:::::
::·.:·::·.::*:

■ Α * ; ; : : Λ : : : Ϊ : Ϊ : Ϊ : · " : :.'::?:·: 3 km

31° 1 0 '

Ilomantsi
:::?::?:::: : · : : : : ! Μ Ι
?ϋ:>::?1? ·.: K in till
!Ü f r a c t i o n < 0 . 0 6 mm
OES-analysi s

? K
■:.:.'.:::::fi!:· ····■··■ ··■·· ·
:."!".;r.t i"; # . . · . · · · · · · β ·.·· ·
. . . . . . ·· . ; .· ^ . . · . . . ·: . :. · . ; ; · ■ · . ·· . : ; , . : . :
r.S"-.- ··■.■.·.::::::::. : : ^ : : : : : : J
i· ■ · ^^ ..." ;* e r . . . . . . . 1
aktaL·* ..·:'.·:
•F·•'W"·· ·· ■·.·"·· ·
.:..·;.:.--- i
" . : ·:· :·■. ;· ·: : · .'.*·'.;
- ' - > · · · ■ * · · · · . ·· · ·· . · · · ·«
i · · · · ! · « · · . . · · . · · ··.···■;.·■:::·:

•v:sfei3iti»>Kv;v;·;*:;^
■ * ■ · ··

62°-
45'

:
·::·:Λ::·.: τϊ: ϊ:
*as
»r·:*· 3 km

;*!:
Ί
31° 1 0 '

Fig. 7-9. A. Nickel contents of till fraction <0.06 mm, sampling density 16/km2. The Ni
anomalies reflect the schist belts (see Fig. 7-9C). B. A metasomatic halo shown by K anomalies
in <0.06 mm till fraction, sampling density 16/km2. Lithology of the bedrock is shown in Fig.
7-9C.
Local scale studies 159

C x ^
X X X X
x ^ x x
χ X X 1 x/ X X X
x
x
x x/
As
if X x x
x / ' X
x XY/ Y X

\f
X X

X X
X X

Wj Y 7
X V\\ f x x
x x
x\ X YYT X
X X
X X X
X

x
Kx

/ x
/ x
X

X
X

X
X

0
1
II x
γ 3

χΛ X X X km x
' X

x K\
x / \ Granitic and granodioritic r o c k s

Y Tonalite

Archaean s c h i s t s , mainly g r a y w a c k e s

Iron f o r m a t i o n

Fig. 7-9 (continued). C. The main lithologic features of the area presented in Figs. 7-9A and B.

Sampling
• sites
# ·
Ice flow · * #

■* * · i ·
direction · ·

Fig. 7-10. A schematic presentation of a sampling grid to be used in areas of long glacial
transport of drift.

samples per km 2 is high in local-scale studies, the delineation of the study


area must he done very carefully to minimize costs.
Another difference between regional and local-scale geochemical studies
is the sampling depth. In the local phase the samples should be collected
as close to the bedrock as possible rather than at the till surface. In
160 Scale of geochemical surveys

the bottom part of the drift the material more closely resembles the
underlying bedrock because of the shorter transport. According to recent
studies in Scandinavia, the material derived from a smaller ore suboutcrop
is detectable by chemical and mathematical methods available today only
within a distance of a few hundred metres. The sampling interval selected
should thus be less than the distance over which the anomaly is detectable.
Light drilling equipment is usually sufficient in regional mapping, but
heavier drilling machines are necessary in the local phase. Portable percus-
sion drills are adequate if the average thickness of the overburden is less
than 10 m. Thicker overburden demands heavier hydraulic or pneumatic
machines. Because the samples are taken from the bottom part of the
overburden, one must be careful that bedrock material — weathered or
unweathered — does not get mixed with the till or other material in the
overburden. The bedrock sample from beneath the overburden is valuable
in itself, but combining the results from different sampling materials is
never permissible in geochemical studies. A sample from the overburden,
especially a till sample, gives information from a much wider area than a
single sample from the bedrock.

Analysis for local-scale studies


The specific goal of local phase investigations, which is to find a certain
ore type, must be taken into account in planning the analytical program.
There is no sense in analyzing all possible elements that the laboratory can
handle, but only those elements that are relevant to identifying the type of
ore in question. In this way costs are reduced and the labour of the geologist
in processing and interpreting the data is kept to a minimum. Information
about the lithology is obtained by analyzing some main elements as well.
In the first place, however, the analytical method should be such that all
major, minor and trace elements characteristic of the ore type being sought
are analyzed. A semiquantitative method may sometimes be adequate for
the elements.

Data processing for local-scale studies


The number of samples in a local-scale study is typically several hun-
dreds, sometimes even thousands, so that computer processing of data is
in practice essential. Manual processing is reasonable if there are only
100-200 samples and 2-3 elements being analyzed. However, most of the
routine data processing methods (factor analysis, discriminant analysis,
gliding mean, residual anomalies, etc.) do not take adequate account of
the many geological factors relevant in interpreting geochemical results
— such factors as weathering, the amount of weathered bedrock in till, the
amount and type of glacial erosion, secondary processes and the location of
Detailed studies 161

the sampling point relative to the water table. These can vary much over
the target area, affecting concentrations and ratios of elements, so that
the results from separate sampling sites differ in a way that sophisticated
statistical methods cannot tolerate. If the variation in the geological na-
ture of the target area is not known and the area cannot be divided into
sub-areas that are statistically homogeneous, the mathematical results that
are obtained may be meaningless or misleading. In such cases the easily
controlled, simple parameters like metal ratios are very useful. In inter-
preting results it is essential to keep in mind the type of ore being sought
and how this type of ore is typically reflected in the metal concentrations
of the overburden. Whereas use of statistical methods and a computer has
obvious benefits, common sense must prevail in their application. In these
days of the widespread availability of specific statistical software packages,
the temptation to mindlessly feed geochemical data into a computer for
commercial routine methods of processing must be overcome.

DETAILED STUDIES

The goal of detailed geochemical studies is to pinpoint the ore suboutcrop.


Studies at this stage are not usually carried out alone but are part of a
broader package of geophysical and geological methods of exploration. At
this stage the object sought is well defined and methods can be selected
that are specific for the ore type expected.
The bedrock — either fresh or weathered — is usually the main
target of sampling, and soil sampling is only complementary. In areas of
very thick overburden it may be sampled from several depths. This, and
the bedrock sampling beneath, usually demand the use of heavy drilling
equipment. Accordingly, the target area should be as small as it can be on
the basis of knowledge available. Besides chemical analysis, mineralogical
and petrological studies should be done; this is especially important for
the bedrock samples. Also, the mineralogical composition of the samples
from the bottom part of the till may give valuable information about the
underlying bedrock and ore deposits.

Sampling in detailed surveys


No special rule can be given for the sampling grid in detailed studies.
If the strike of the bedrock is known, a reasonable approach is to sample
along lines or traverses perpendicular to the strike. Although an equally
spaced grid can be used, it becomes expensive because the space between
neighbouring sites is then usually less than 10 m. In areas of thinner
overburden the use of a tractor excavator for digging trenches is eco-
nomical. The bedrock, weathered or unweathered, is exposed and reliable
162 Scale of geochemical surveys

sampling done, and direct observations can be made of the bedrock. The
representativeness of a sample becomes very important in this phase. In
the lower density sampling, when samples are taken from the overburden,
and especially from the surface layer of the overburden, even a very small
sample is homogeneous and representative enough. But for detailed study
a single sample from the bottom of the overburden, and especially from the
bedrock under the overburden, is not representative enough. The quality
can be improved by using composite samples from the till or taking large
(> 1 kg) samples from the bedrock.
In areas of thicker (>5 m) overburden, where the bedrock cannot
be reached by excavator, heavy drilling machines must be used. The
equipment should be large enough to allow penetration of as much as 30
m of overburden, and to sample the bedrock underneath. There is one
big difficulty in bedrock sampling through the overburden: the sample
could unknowingly be from a narrow vein or unrepresentative type of
bedrock. Caution about the representativeness of samples needs to be
exercised at all times. It would be helpful to take a sample from the
bottom part of the overburden as well, since it reflects — especially in
glaciated terrain — bedrock from a wider area than the sample from a
point in the bedrock. The representativeness can also be improved by
taking a core sample a few metres long by diamond drill. Unfortunately, the
sample from the overburden, the sample from the surface of the bedrock
(crushed or powdered) and the core sample all require a different type of
sampler; taking all types of sample may be very expensive despite modern
drilling machines that make it possible to use the same engine for different
samplers.

Analysis in detailed surveys


In detailed-scale geochemical studies the ore type sought is well known
and the number of elements to be investigated small. Often it is sufficient
to measure concentrations of the metal of interest alone. On the other
hand, there are usually different sample media to analyze — till, bedrock,
weathered bedrock — all of which require unique handling. This makes the
analysis both more laborious and more expensive. Also, the way in which
the metal is bound, e.g., in primary or secondary minerals, must be taken
into account. The sample should be treated so that only the portion of the
metal that is of interest is assayed.
Because sampling in the detailed phase is always expensive, laboratory
results should be made available as quickly as possible so that the sampling
can be directed in an optimal way and the number of useless samples
minimized. Quick analysis also allows the sampling to be completed in one
operation; transport of drilling machines to and from the target area is
expensive, as are start-up and wrap-up of the operation.
Conclusions 163

Data processing in detailed surveys

The computer plays a less important role in processing the data and
results of the detailed study phase than in the regional and local-scale
studies. Usually mineralogical studies account for a substantial share of
the data interpretation at this stage. Multivariate statistical methods are
not required, because trends in elemental concentrations and their internal
relations will suffice. Often the raw data will be sufficient for the geologist.
If the number of samples is less than a hundred, or at most a few hundred,
as often is the case, it may also be more economical to prepare the maps
and diagrams for the report by hand.
In contrast to the regional and local-scale studies, what is important at
this stage is that the geologist in charge be constantly in close touch with
the work in field, laboratory and office. Things usually happen fast and
interpretations and decisions need to be made promptly if the operation is
to be cost effective.

CONCLUSIONS

The scale of a geochemical study depends on the aim of the study and,
according to the aim, geochemical studies are divided into three groups: (1)
regional, (2) local, and (3) detailed.
In regional-scale studies (i.e., regional geochemical mapping), geochem-
ical provinces several tens of square kilometres in area are delineated. A
study of this scale is cheap because of the low sampling density (1.0-0.01
sites/km 2 ) and easy sampling (samples are taken close to the surface of the
overburden). The result is broad and reasonably reliable information about
the geochemistry of the study area. It is also possible, on the basis of the
regional-scale study, to evaluate whether geochemical prospecting methods
are practical in the study area. Analysis for as many elements as possible
and interpretation of the results by multivariate statistical methods are
typical in this phase.
In local-scale studies, zones some hundreds of metres wide, which
are critical for prospecting, are delineated. The sampling density varies
between 4 and 40 samples/km 2 . The best sampling material is basal till,
typically transported some tens or hundreds of metres by glacier. This
till represents a fairly wide source area. As a rule, the ore type sought
is sufficiently well known that analytical methods can be selected on that
basis. Fewer elements than in regional-scale studies need to be analyzed.
However, sampling depth needs to be greater, which makes the sampling
more expensive and limits the feasible area of study to a maximum, say
100 km 2 . The results are used for planning the next phase (detailed) of
geochemical studies and sometimes even diamond drilling.
164 Scale of geochemical surveys

The goal of the detailed studies is finally to locate the ore outcrop or to
indicate sites for diamond drilling. The sampling density is tens or even
hundreds of samples/km 2 . Samples should be taken from the bottom of the
overburden close to the bedrock. These studies are expensive and must
be restricted to small areas only. Optimally they are used together with
geophysical and geological methods.
A geochemical study proceeding step by step from regional mapping via
local-scale studies to detailed geochemical prospecting is very useful in
areas where geological knowledge is scant. It is then possible to proceed
from large geochemical provinces via ore critical zones to the final locating
of an ore body.
165

Chapter 8

FIELD METHODS

INTRODUCTION

Factors to be considered in planning sampling grids were discussed


in Chapter 7 where it was pointed out that the decisions concerning
sample density, line and sample spacing, and depth of sampling depend
upon the scale of the survey Once these decisions are made, the choice
of sampling equipment becomes important. Sampling depth and soil will
largely determine the selection to be made from among the wide variety of
equipment now available. And this in turn will have a decisive impact on
overall costs. The survey strategy must thus be developed as a totality, in
which the whole field operation stage, particularly the choice of sampling
equipment, is in harmony with the objectives and cost limitations of the
study.
The size of sample is often a key factor in the selection of the sampling
method. When the fine fraction of till is studied for base metals, a sample
of 100-200 g is usually sufficient. However, if gold, which as "nuggets"
is heterogeneously distributed in the soil at very low concentrations,
is the exploration target, the sample needs to be many times larger,
even if the final grain size fraction to be analyzed is of the same size.
If coarse fractions are used as the study media, as in heavy mineral
surveys, the sample size required increases in relation to the grain size.
The representativeness of the sample is a crucial factor in geochemical
exploration, and under no conditions should representativeness or other
qualities of the sample be sacrified even if the result is an increase in overall
costs.
Experience has shown that a sample composed of several (2-5) subsam-
ples is invariably more representative than a single sample. Subsamples
collected at intervals of a few metres naturally raise the cost of sampling
considerably, especially if the samples are taken with a drill; however, good
quality is always worth the expense. The number of subsamples must, of
course, be optimized in terms of cost. When samples are taken with a drill
from the basal part of thick overburden a couple of subsamples are often
166 Field methods

sufficient. But, if a markedly less expensive sampling method can be used,


four or five subsamples are preferable.
The representativeness of the samples should be checked at each stage
by taking duplicate samples and the analytical error by making replicate
analyses. The natural heterogeneity of till inevitably causes variation in
the analytical data. Though often falsely referred to as sample error, this is
only variation due to natural heterogeneity. To check for representativeness,
duplicate samples should be taken at appropriate intervals. For example,
5% of the total amount of samples should be duplicates. The element
abundances in the sample and in the duplicate sample can be compared as
a way of assessing the natural heterogeneity of the soil and, at the same
time, the representativeness of the samples.

REQUIRED FIELD OBSERVATIONS

The sampling material of the overburden — especially organic soil — is


continuously subject to a variety of natural processes. These processes may
seasonally change the abundances and ratios of the elements in material
why they must be taken into account when interpreting the data. Thus
certain observations of the environment should be made during sampling.
The most common practice is to enter details about relevant environ-
mental factors on a data form at the sampling site. If portable computers
(Fig. 8-1) are available the data can be stored equally well in the field; in
fact this is recommended, as one source of error — that due to the transfer
of data from forms to computer — is thus eliminated.
Since the observations required will depend on the survey strategy, study
area, sampling medium and other local factors, there is no point here in
giving a comprehensive list of items that should be observed. Factors that,
in general, may affect the study results include morphology, topography,
vegetation, moisture (the position of the ground-water table in relation
to sampling depth), soil type and its properties such as grain size and
stoniness. There may also be some important local factors to be observed
and recorded. Information about the type of till or soil from which the
sample was taken and the location of the sampling point in till or soil
stratigraphy are the minimum items to be recorded.
Perhaps the most crucial observation to be made at any sampling site is
the location of the sampling point. The accuracy required of the location
data will depend on the sampling density. In normal cases, coordinates
measured from a map are sufficient for regional and local surveys. The
actual sampling point can be freely chosen by the sample taker. In detailed
studies, however, in which the sample spacing is usually only some tens of
metres or even no more than a few metres, the location of the sampling
point must be known accurately. The best results are then obtained by
Required field observations 167

Fig. 8-1. A portable computer is recommended for recording the field data. Photo Pekka
Virtanen.

staking out lines at the sampling target and tying the sampling points to
them.
The need to stake out lines beforehand to obtain the location data
greatly increases the total cost of sampling. Often, however, the same line
system can be used for other studies, such as geophysical ground surveys
and geological observations, thus reducing geochemistry's share of the
total expenses. Once automatic locating satellite systems are available at a
reasonable price, as they soon will be, determining the location and storage
of the data will be considerably more rapid, easy and accurate.
168 Field methods

SAMPLING EQUIPMENT

The wide range of sampling equipment on the market puts the planner
of a survey project in quite a quandary, at least in theory. Sampling
(the whole field stage) accounts for 50-80% of geochemical survey costs.
Thus it is not just the survey strategy but also the funds available that
decisively affect the choice of sampling method and sampling equipment.
The sampling equipment used for till and residual sediments can be divided
into two groups with very different rates and depths of penetration: (1)
sampling without flushing, and (2) sampling with drills provided with air
or water flushing. Drills with flushing can penetrate tens of metres of
minerogenic overburden; the best units even permit sampling of underlying
fresh bedrock. Without flushing, in contrast, the sampling depth in till of
light drills is no more than a few metres. Flushing, of course, tends to
remove the finer grain size fractions.

Sampling without flushing

In areas where the overburden is shallow (less than 5 m) or if it has been


decided that samples from the superficial soil will suffice (regional mapping),

Fig. 8-2. Sampling from a pit dug by a tractor excavator is often the most cost-effective
procedure. Photo Peter Johansson.
Sampling equipment 169

Fig. 8-3. Blasting of the sample from bedrock gives samples with a fully satisfactory size and
representativeness. Photo Pekka Virtanen.

the most cost-effective procedure in many cases is to take samples from


pits dug with spade or excavator (Fig. 8-2). Terrain permitting, the use of a
tractor excavator is recommended, since sufficiently big and representative
samples can then be obtained whatever the conditions, at the same time as
reliable observations can be made on till and soil stratigraphy in the pit.
In many cases, the tractor excavator has also turned out to be the most
economical sampling equipment. If cheap labour is available and sampling
depth of less than one metre is acceptable, the pits can, of course, be dug
with spade. The representativeness and the size of the sample are the same
as those with the excavator, but the economical sampling depth remains
fairly low.
Sometimes it will be necessary to blast a sample from frozen or tightly
cemented (residual) sediment or from bedrock (Fig. 8-3). In these cases,
too, the size and representativeness of sample are often fully satisfac-
tory, although blasting is not without its risks. Metallic detonators may
contaminate the samples.
In the event that digging exploration pits for sampling is out of the
question, the choice is between the great number of light-weight portable
drilling units on the market. The simplest is a manual impact drill (Fig.
8-4), which guarantees a sampling depth of about 2 m when conditions are
170 Field methods

Fig. 8-4. A simple hand-held auger drill guarantees a sampling depth of about 2 m provided the
conditions are good. Photo Martti Kokkola.

good. The sample is obtained with an open auger bit. A manual impact
tube (Irish) sampler is still better. With these samplers there is always a
danger of contamination, therefore particular caution should be exercised
during sampling. The method has the advantage that the equipment is
light-weight and can be taken anywhere accessible to the sample taker.
Portable percussion drills (mechanical or hydraulic) driven by combus-
tion engines are a more advanced and efficient type of drill (Fig. 8-5). The
hydraulic ones, mounted on sledge, have shown to be very effective. They
penetrate more than 10 m in mineral soils. The stoniness and boulders
hamper the use of this equipment. Nevertheless, samples can be taken from
substantially greater depths than with the above sampling techniques, and
contamination can be avoided by using an appropriate sampler. Although
in principle the equipment is portable, so many rods and other accessories
have to be taken along that its transport by human force is not economical.
A snowmobile is a convenient transport vehicle in winter and a crawler in
summer.

Drilling with flushing

Drilling techniques that make use of flushing are notably more effective
than those that do not. The flushing can be done with compressed air
or water or even with both together. In either case the sampling unit is
a complex, heavy and awkward piece of equipment to move around (Fig.
8-6). The primary requirement for the application of water flushing is, of
course, that sufficient water be available and that a supplementary pump
Sampling equipment

Fig. 8-5. A portable percussion drill driven by a combustion engine is a more advanced and
efficient type of drill. Photo Pekka Virtanen.

and water line be installed at the sampling site. Indeed, the method tends
to be so costly that it is seldom used for till and soil sampling and is
better reserved for bedrock sampling through the overburden, and even
then mainly for delineating a formation already discovered. In terms of
cost, this type of sampling approaches that of diamond drilling, which it
also resembles technically although the hard metal crowned drilling bits
are cheaper. Under difficult conditions, it may turn out to be even more
costly.
Drilling equipment relying on compressed air flushing can usually be
mounted on a self-propelled base, which makes for easy mobility (Fig. 8-7).
However, the unit is large and clumsy, requires fairly passable terrain and
is often useless on boulder-strewn terrain or steep slopes. In operating cost
172 Field methods

Fig. 8-6. A drilling machine using water flushing is effective but so costly and so difficult to
move from one sampling site to another that it is seldom used for till and soil sampling. Photo
Pekka Virtanen.

it does not differ significantly from water-flushed drills, and when very
thick overburden (several tens of metres) is being drilled the penetration
capacity is not quite as good as that of water-flushed drills. On the other
hand, the higher mobility of the equipment permits sampling on a coarser
grid, provided, however, that the terrain is not too rough for the heavy unit.
The usual procedure in both water-flush and compressed-air flush drilling
is that, once the desired depth has been reached, the drill bit is replaced by a
special sampling bit. This requires hoisting up the string of rods. Although
this means extra work, the risk of sample contamination is nonexistent
and, in general, the sample obtained is of good quality and the right size.
Reverse circulation drilling (e.g., Strauss et al., 1989) permits continuous
Sampling equipment 173

Fig. 8-7. Drilling equipment relying on compressed air flushing is fairly mobile, but the unit is
large and clumsy and requires fairly passable terrain. Photo Martti Kokkola.

sampling without rod hoisting and bit changing. In this procedure, the
flushing water is returned to the surface along with the loosened till or
soil. Collecting the sludge gives a continuous sample large enough for most
purposes. Although some of the finest material is lost with the flushing
water, experience indicates that sample mixing is very slight and that the
depth information is unbiased. Like the other drilling units with flushing,
174 Field methods

this, too, is expensive, complicated and cumbersome to move and is totally


out of the question for a large number of samples over a wide area,
which is geochemical exploration at its most typical. However, in detailed
studies with a sample spacing of only a few metres, drilling by flushing
is well justified, provided that conditions do not permit cheaper sampling
techniques.

Rods
The drilling method and the source of power it requires generally
determine the type of rod. Given the large selection of rods available,
there is some danger of choosing an inappropriate one. It is normally the
rods close to the coupling sleeves that fail. Therefore both the rods and
coupling sleeves must be of strong enough material to withstand the blows
associated with drilling. Even so, metal fatigue is inevitable and the rods
will deteriorate in operation. (When the drill stem is extended the oldest
and weakest rods should be placed lowermost, because fewer rods will then
be lost if the stem breaks and remains partly embedded in soil). Steel is
the most common material. The aluminium rods employed in diamond core
drilling are no good in soil sampling, not even if flushing is used, because
they are soft and easily damaged in stony soil.
Solid rods can be used in lighter drilling units. These need external
coupling sleeves when the stem is extended, however, and this hampers not
only the penetration of the rods but in particular their hoisting.
The heavier drilling units, which use thicker and stronger rods, usually
employ hollow rods with internal coupling. The drill stem is then smooth
and does not resist penetration.

Samplers
A number of samplers have been developed for soil studies, most of
which have also been tested for geochemical soil sampling. Experience has
taught, however, that to work reliably in hard and stony soil the sampler
must be as simple as possible. Piston and rod type drills or the like are
thus out of the question. Three types of sampler have established their
position: the tube sampler operating on a flow-through principle, the closed
tube sampler and the open auger bit. The closed tube sampler (Fig. 8-8) is
used mainly in drilling by flushing. Sometimes (mainly when compressed
air flushing is used) the sampler is attached to the rods only for the actual
taking of samples; a different cutting head is used for penetrating the soil
layers.
The auger bit (Fig. 8-9) facilitates penetration of the drill stem and is
often used without percussion merely by rotating the drill. However, the
recovery of a sample with an open bit is not very good, and in dry soil
Sampling equipment 175

Fig. 8-8. The closed tube sampler is used mainly with water-flushed drills. Photo Martti
Kokkola.

the sample tends to slip off the bit before it reaches the surface. The risk
of contamination from the upper layers is also considerable. It is often
possible, however, to use fairly wide bits. The size of the sample is then
big enough for most purposes, and the internal parts of the thread contain
material from the desired depth.
The tube bit operating on the flow-through principle (Fig. 8-10) has
turned out to be reliable and easy to handle. However, when the bit is
176 Field methods

Fig. 8-9. The auger bit is usually merely rotated when it penetrates the fine-grained overburden
fairly well, but the recovery of the sample with an open bit is not very good, especially in dry
soil. Photo Martti Kokkola.

coupled to a hand held percussive drill the size of the sample will be small
and usually only sufficient for chemical analysis; it is definitely too small
for mineralogical and lithological studies. In quality, however, the sample
is good and rarely contaminated. This type of bit is not only the cheapest
possible sampler but is simple in construction and therefore reliable.
Sampling equipment 111

Fig. 8-10. The tube bit (Irish type) operating on the flow-through principlehas proved to be
reliable and easy to handle. Photo Martti Kokkola.

Transport vehicles
The choice of transport vehicles used in sampling depends on many
factors. The sampling sites are often far from roads, at least during
the final stages. Only regional geochemical mapping can rely on roads.
Transport by car is then fast and fairly cheap, and sampling units, samples
and other supplies can easily be carried.
178 Field methods

Fig. 8-11. A cross-country motor bike is often useful for moving outside the road network.
Photo Henry Vallius.

In areas of flat topography and not too lush vegetation it is often possible
to move outside the road network with all-purpose vehicles or at least
with cross-country motor bikes or various amphibious and other all-terrain
vehicles (Fig. 8-11). Heavy sampling units cannot be transported by motor
bike, however. Special logistics are required for the heavier equipment
required in detailed studies.
An unbroken cover of snow in winter makes it easier to move and
transport drilling units of all kinds. Pulled by motor sledges or heavier
crawler-based vehicles (Fig. 8-12), even the heaviest drilling units can
rapidly be moved over quagmires and lakes to take the shortest route to
the sampling site. Likewise, daily trips from the sampling site to the road
are feasible, even from distances of tens of kilometres. Deep snow prevents
the freezing of ground, deep frost hampers sampling by drilling.
In summer, use can be made of the waterways in otherwise inaccessible
terrain, although transport of the heaviest drilling units by boat may be
laborious and even dangerous (Fig. 8-13).
Aircraft, and helicopters in particular, make the transport of drilling
equipment feasible to practically any site. Expensive though it may be, use
of aircraft is often the only way to undertake studies economically at targets
Sampling equipment 179

^i 4

Fig. 8-12. Snow in winter makes it easy to move and transport drilling equipment by motor
sledge. Photo Jari Nenonen.

Fig. 8-13. Boats can be used for transporting drilling units through waterways in otherwise
inaccessible terrain. Rock geochemical sampling with portable diamond drill. Photo Tapani
Taipale.
180 Field methods

Fig. 8-14. Use of aircraft is often the only way to reach targets in the trackless wilderness
economically. Photo Pekka Virtanen.

in the trackless wilderness (Fig. 8-14). With light-weight sampling units


it is sometimes even worth moving from one sampling site to another by
helicopter. This is particularly reasonable in geochemical mapping, where
sampling sites can be several kilometres apart.
The choice of logistics has a marked effect on the overall costs of the
study. The transport vehicles and the sampling equipment are related to
each other: a heavy sampling unit requires special transport, whereas a
light unit moves easily. The final decision of course depends on the resources
available. And the experience of the person responsible for the planning
and supervision of the study plays a crucial role in the accumulation of total
costs. Field operations always constitute the principal source of the cost of
an investigation, accounting for 50-80% of overall expenditure, depending
on the equipment.

FIELD MEASUREMENTS

Measurements and determinations made in the field are a vital part of


geochemical exploration. Observations on the geological features of soil and
bedrock are a natural component of any explorational activity. However, in
the course of geochemical studies it is also possible to carry out elemental
Field measurements 181

analyses and determine certain chemical and physical parameters right at


the sampling site. Field laboratories have the capacity for fairly demanding
chemical analyses.

Physical measurements
Chemical dispersion that takes place in the gaseous state can be mea-
sured with counters, spectrometers and photographic films (Dyck, 1969a;
McCorkell et al., 1981). The measurement data are then instantly available
in the field, or at least in the field laboratory. Measurements have been
made of radon, helium, hydrocarbons and gaseous mercury. Gases also
carry heavy metal ions. Ions migrating in the pore solutions of soil can be
collected by allowing them to accumulate in filters, ion-exchange resins, ion
traps, etc. for a given period — generally a few weeks — before analysis.
The analyses often must be made in more sophisticated laboratories than
those applicable in the field, however.
Some purely physical methods, such as the self-potential method and
determination of magnetic susceptibility, can be used in conjunction with
geochemical studies. Under certain conditions these are useful, particularly
because the measurement data are available instantly without the lag of up
to several weeks with laboratory procedures. Susceptibility values permit
rapid determination of the variation in magnetite abundances at the study
target. The use of self-potential measurements, though, requires not only
sound knowledge of the measuring technique and its underlying principles
but also sufficient information about the type of target being sought. Only
then can the right conclusions be drawn from the survey data describing
the target's potential field.
It is often important to know the thickness of the overburden, particu-
larly in glaciated areas, as this is vital for the interpretation of geochemical
data. If the available sampling equipment is heavy enough the information
about the thickness of the overburden can be obtained directly in the course
of sampling. When lighter-weight equipment is used, however, the thickness
data must be obtained by other methods. An accurate method is seismic
sounding, but its high operational costs mean that it is rarely applied in
geochemical studies. A markedly cheaper approach is ground-penetrating
radar. True, it too has its limitations, particularly when the stratigraphy is
complex, but it can often be used successfully and even cost-effectively.

Chemical measurements
Earlier, and particularly in locations far from population centres, anal-
yses were made for many elements right in the field. These included
element-specific spot-test analyses, one of which was the qualitative de-
termination of nickel with dimethylglyoxime. With the improvement of
182 Field methods

laboratory methods and the introduction of rapid methods permitting si-


multaneous determination of several elements, the application of spot-test
analyses has declined. The other widely used method is the bulk analysis
of heavy metals by dithizone titration, which was very popular during the
1950's. These methods are still useful under certain conditions and should
not be allowed to become obsolete.
A more advanced form of spot-test analysis than with chemicals is the use
of ion-specific electrodes, which is often feasible even in field laboratories.
Similarly, portable XRF analyzers have made great strides recently and are,
with all their limitations, quite useful.
Sometimes the pH and Eh of soil water are measured in conjunction
with soil sampling. Under certain conditions, the pH measurements may
have some value, but the unambiguous interpretation of Eh data is almost
impossible and measurements are useless in ordinary exploration.

Mineralogical studies
Several heavy-mineral separation methods, which are mainly applicable
to field operations, are in use, particularly in gold exploration. Should
the gold occur as coarse enough nuggets (grain size over 30 μιη), these
methods would often give a more reliable result than chemical analysis.
Several mechanical explorational devices have been developed to separate
gold. However, if the gold is very fine-grained, as it unfortunately often is,
the numerous and complex devices whose operation is based on the density
difference between grains are not satisfactory. The devises all require the
hand of a skilled user if satisfactory results are to be achieved.

CONCLUSIONS

Immediately the geochemist begins to work with a new exploration


target, he/she faces a large number of choices. The scale of the study
will be determined according to the strategy of the study. Then the
sampling material and the sampling equipment will need to be selected.
The situations, when and exactly what to select are boundless. Fortunately
or unfortunately, many of the alternatives are effectively eliminated by lack
of money!
The study strategy determines the type of sampling equipment. In
regional geochemical mapping, when the samples are taken from the
surface part of the overburden and when a sparse sampling net is used,
light portable drilling equipment is preferable. But in detailed studies
when samples should be taken through a thicker overburden, much heavier
drilling machines are needed. A wide variety of good equipment for every
purpose is on the market, and the final selection will often be based on
Conclusions 183

secondary factors: what is available at the moment and how large the
budget is.
Besides sampling, a variety of observations and measurements can be
made during geochemical studies in the field. Observations such as of the
location (coordinates) of the sampling site are essential. But several other
observations appropriate to the aim and scale of the study can be made
as well. Naturally, the most important of these deal with the geology of
the bedrock and overburden. Studies concerning the character, amount and
direction of geochemical dispersion are particularly useful.
Measurements of radiation are used in prospecting for radioactive ele-
ments. Some indicator and also ore elements can be studied in gaseous
phase by special measuring equipment. Purely physical measurements such
as the measurement of the self-potential field are sometimes made in the
course of ordinary geochemical studies.
Certain laboratory work can be done, and in remote places even should be
done, in field conditions. Simpler analyses and certain mineralogical studies,
especially in heavy mineral investigations, are appropriately carried out in
field laboratories.
185

Chapter 9

ANALYTICAL METHODS

INTRODUCTION

Modern analytical techniques enable the production of multi-element


data for large batches of samples at affordable cost. Analytical determina-
tions of most elements bound in silicates, sulphides and oxides can be done
by relatively direct procedures and with automated, but often expensive
instruments. Only a few elements — usually geochemical pathfinder ele-
ments with low crustal abundances — still need to be chemically separated
and preconcentrated. Since the present level of automation allows easy col-
lection of data from instruments and further processing and presentation
of the data in an almost unlimited number of ways, the bottlenecks of geo-
chemical study are the sampling and manual handling of sample materials
in the field and laboratory. Today, laboratory work accounts for perhaps
10-25% of the total expenses of a geochemical study, and approximately 1/3
ofthat is incurred by sample logistics and pretreatment.
This chapter introduces various aspects of chemical analysis: criteria
for method performance, sample pretreatment, sample preparation and
instrumental techniques of element determination. Special attention is
paid to the analytical performance characteristics, to the preparation of
solution mode analytes and to instrumentation suitable for the analysis
of solutions. In addition to mechanical dispersion, the mobilization and
sorption of elements in the overburden depend on the characteristics of the
natural solutions in the secondary environment. By using the appropriate
extraction agents for samples one can simulate these natural phenomena,
though in the reverse order, and obtain information on primary mineralogy,
modes of occurrences of elements and the character of anomalies.
None of the concepts or techniques touched upon is used solely in the
analysis of materials of arctic and temperate regions, but the evaluation
of various analytical tools for the analysis of these samples receives
prominence. The chapter is intended to provide, in a nutshell, relevant
analytical information for geochemists planning geochemical studies and
interpreting analytical data. Most of the practices described are ones used
in the chemical laboratories of the Geological Survey of Finland.
186 Analytical methods

Production of geochemical data useful for environmental studies will


be increasingly important in the future. Many of the analytical methods
described here can be applied as such or slightly modified in environmental
studies; however, no specific methods for environmental analysis have been
considered in this chapter.
Truly comprehensive surveys of analytical methods available for varied
geochemical studies can be found in Volume 1 of this Handbook, Analytical
Methods in Geochemical Prospecting (Fletcher, 1981), as well as in the
textbooks of Schroll (1975), Johnson and Maxwell (1981), Potts (1987) and
Van Loon and Barefoot (1989).

PERFORMANCE OF ANALYTICAL METHODS

General
Informed selection of analytical procedures is essential to a successful
geochemical study Until recently, the selection of procedures largely de-
pended on experience and intuition, but today demonstrated performance
is the only acceptable criterion. The problem is to identify the procedure
or procedures that correspond best to the exigencies of the application.
It is well enough known how "poor quality" analyses may undermine
the objectives of a study. But it is often forgotten that choosing proce-
dures on the criteria of greatest precision, lowest limits of detection, etc.,
may lead to such low throughput and high costs that the study becomes
prohibitively expensive. Ensuring the economic viability of an extensive
mapping programme involves a very delicate balancing of diverse criteria.
In general, the performance characteristics (or "figures of merit") of
an analytical procedure can be divided into technical characteristics and
economic considerations. The technical characteristics comprise precision,
accuracy, reliability, sensitivity, limit of detection, selectivity and informa-
tion capacity, while the economic considerations involve cost and time, viz.,
throughput and turnaround. Most of these factors have been discussed in
detail by Massart et al. (1978). In the following each is touched upon briefly
from the point of view of exploration geochemistry. The characteristics are
considerably interrelated, and optimization with respect to one parameter
may weaken another. For example, improving the precision by a factor n
may increase the costs of an element determination by the same factor. In
this case the interrelation is straightforward; but most often it is obscure
(Fig. 9-1).

Technical characteristics
Precision is a measure of random errors and is by definition closely
related to repeatability, reproducibility and laboratory bias (interlaboratory
Performance of analytical methods 187

RELIABILITY

SENSITIVITY SELECTIVITY

ACCURACY

DETECTION LIMIT— INFORMATION CAPACITY >RECISION

COSTS

— EFFICIENCY
STAFF

Fig. 9-1. Interrelations between analytical performance characteristics.

precision). Analytical precision should be adjusted to pair with the sam-


pling precision and with natural geochemical relief. Thus a high analytical
precision and minimizing of sampling errors is required for weak relief,
while lower sampling and analytical precision can be allowed for stronger
geochemical relief. To maintain economy, analytical precision should not
substantially surpass the sampling precision, but the two should be bal-
anced. A lack of any anomaly pattern on geochemical maps may indicate
an inadequate precision of analysis. In terms of the analysis of variance
(ANOVA) the combined analytical and sampling variances should account
for not more than 1/4 of the total variance due to the geochemical relief.
The analytical variance should then preferably comprise less than 5-10%
of the total variance. An analytical precision of roughly 5% (indicated
as relative standard deviation at the level of threshold) is suitable for a
method used for glacial drift materials. Much poorer precision must be
accepted for routine analysis of elements like Au, Sn and W, which are
non-uniformly distributed in laboratory (and field) samples and appear in
low concentration.
Precision is usually a function of concentration: standard deviations
tend to increase towards lower limits of detection, and often towards
higher limits of detection as well. The precision of X-ray fluorescence
measurements, however, improves steadily up to 100% concentration.
Accuracy refers to systematic error or difference between the experimen-
tal mean and the "true" value. Poor accuracy in analyses causes particular
problems in mapping programmes carried out over a long period of time
(see Reliability below). Likewise, integrated processing of geochemical data
produced by two or more different methods or laboratories is always risky
because of possible differences in the accuracies. Differences on geochemical
maps due to the analytical method or to laboratory practice may prevent
recognition of real relief.
Some lack of accuracy, or mere "consistency" of data, can be accepted
within limited geographical areas or areas of strong geochemical relief.
188 Analytical methods

A good rule of thumb is to strive towards an accuracy that matches the


precision. Thus, for most sulphide-forming metals in till, a relative accuracy
of 5 to 10% should be acceptable; for the total concentrations of the main
components an accuracy of better than 5% is often required.
Accuracy is not a relevant criterion for procedures based on partial
leaching of samples. For these methods the constancy of method parameters
must be guaranteed to produce consistent data.
Reliability refers to the capability of a method to maintain accuracy and
precision over time. The notion of drift is related to reliability and is defined
as a systematic trend in the results as a function of time. Poor reliability
(together with poor selectivity, see below) is a fundamental danger in
extended exploration surveys. For example, time-dependent shifts in the
degree of accuracy may appear on maps as a "checkerboard" pattern or else
as a lack of clear pattern altogether. The former problem could arise for
samples analyzed in the sequence of sampling order, the latter for samples
analyzed in random order.
Unreliability and drift are often said to be characteristic of automated
methods, e.g., simultaneous optical emission and X-ray fluorescence spec-
trometry. In fact, automated methods are not more prone to these effects
than manual methods, but the larger series of determinations carried out
with the automated methods makes the time-dependent problems more
visible. Such methods of quality control as control charts have been used
to reveal drift. Reliability of a method can be improved through more
thorough calibration of the procedure.
Sensitivity is defined in quantitative analysis as the slope, dy/dc, of the
analytical calibration function, y = f(c). The expression "sensitive" is often
erroneously used in the sense of low limit of detection. The sensitivity
of the method is high if small changes in concentration (c) give large
differences in measured signal (e.g., voltage units) responses (y). The range
of concentrations over which the sensitivity can be considered measurable
and even constant — the "linear dynamic range" — is bounded by lower
and upper limits of detection. The upper limit of detection can easily
be altered without significant change in the procedure, for example by
diluting the analyte. However, the lower limit of detection, "the detection
limit" as it is generally called, is a property of the analytical procedure.
Wide dynamic ranges, as in plasma emission methods, enable rapid access
to the variable concentrations that often apply to geochemical samples.
Satisfactory methods should be highly sensitive (and precise), especially at
threshold level. Sensitivity of some methods can be improved, in absorption
spectrometric measurements, for example, by increasing the optical path
length.
Limit of detection is defined in terms of a calibration function and
precision. Literally, it is to be understood as the concentration below which
the detection of a signal from the analyte is impossible. Unfortunately,
Performance of analytical methods 189

P(y) UNRELIABLE DETECTION DETECTION DETERMINATION


\ ^.
QUALITATIVE QUANTITATIVE ANALYSIS
ANALYSIS ^τ^
/ ^

y L L
blank decision "-detection determination y
Fig. 9-2. Illustration of normally distributed blank, signals with increasing confidence and
critical limits. Shaded areas represent errors of the second type (/?), that is, a decision that the
signal (element) is absent when it is present. Beta has been given here a value of 0.5 at the
decision limit and a value of 0.05 at the detection limit. The standard deviations a b j a n k and
^signal are considered to be equal, which is the case for small concentrations.

there is no single definition for detection limit. Currie (1968) introduced


the concepts of decision limit, detection limit and determination limit
(L), which are expressed in terms of the measurable property (y) in the
expression:
=
L y blank + ^ b l a n k

where ah\&nk represents the standard deviation of the blank. The choice of k
is arbitrary and depends on the confidence required to answer the question
whether the analyte component is present (decision), can be detected or can
be determined (Fig. 9-2). For detection limit, many authors have chosen a
value of 3 for k. The three-sigma detection limits are obtained in practice by
taking the noise on the background as sigma. However, in some analytical
techniques, as in atomic absorption, the background noise is comparable to
the noise of the analyte signal, whereas in others as in plasma emission, it
is much lower. Thus 3σ limits are not more than indicative for elements
in geosamples measured by plasma emission spectrometry. The real limits
of detection are much higher. A more useful approach may be to define
the limit of detection as the concentration where the relative standard
deviation of the result reaches, say, 50% without regard to the blank or
background noise at all (Thompson and Howarth, 1978).
Very low detection limits are typically required for the study of elements
like Au, most of its pathfinder elements, and PGE, Ag, Mo and Pb.
Too high detection limits and imprecision show up as an absence of
any clearly delineated pattern on geochemical maps. Detection limits
are typically lowered by implementing preconcentration and separation
procedures before measurement.
The selectivity of an analytical method is related to how free the cal-
ibration function, y = f(c), is from interfering influences of concomitant
190 Analytical methods

components, in the matrix. Complete specificity is an ideal feature sel-


dom encountered, although approached in plasma-mass spectrometry Poor
selectivity may be an inherent property that is difficult to control (some
spectrophotometric procedures) or one that can be managed, as often in
X-ray fluorescence spectrometry
Poor selectivity in geochemical data may be seen when there is interfer-
ence from one type of sample matrix and not from another. For example, the
determination of Sn by arc/spark emission spectrometry is difficult in till
enriched in Fe oxides. On a geochemical map and in factor analysis, it may
be difficult to separate these seeming interrelations of elements from real
geochemical correlations. Mathematical correction, matrix modification and
chemical separation can be applied to overcome such effects, but misleading
interferences may still remain.
The information capacity of a method, as introduced by Belyaev and
Koveshnikova (1975), is a useful concept, especially for simultaneous multi-
element methods like emission and X-ray fluorescence spectrometries. The
information capacity (7) of a single-element method is defined as:
7 (in bits) = 2 log7Q
where K{ is the number of distinguishable concentration gradations within
the dynamic range of the method. Thus, 7 is strongly related to the
precision of the method. The information capacity of a simultaneous n-
element method is calculated by summing over i = 1 to n, which means
that 7, and the amount of information from a sample, increases more
rapidly with the number of simultaneously determined elements than with
the number of distinguished concentration gradations. High information
capacity may be considered a very desirable feature for multipurpose
geochemical mapping, but is less important when considerable information
already exists about samples, as in the follow-up phases of exploration.

Economic considerations
Costs encompass the monetary value of labour, instruments, materials
and housing for all functions of the laboratory Unit prices should be
estimated for all operations from sample logistics and pretreatment to
reporting of the results, and include the R&D required to bring a procedure
into service. Cost-benefit analyses are seldom done explicitly, but are
usually involved implicitly in evaluations such as "cost-effective" and
"adequate". Labour is by far the most important cost-determining factor in
laboratories, and the following generalizations can be made about costs per
analysis.
When classical methods, cheap instruments and much labour are used,
the cost per determination is largely independent of the number of deter-
minations carried out. But with automated instruments, determinations
Performance of analytical methods 191

TABLE 9-1

Estimated costs of analysis (U.S.$) at different daily throughputs. Aqua regia soluble contents
of some 30 elements measured by simultaneous plasma emission spectrometer (ICP-AES) at
the Geological Survey of Finland

Item Daily throughput (number of samples)


1 10 100
Investments and maintenance 24 82 280
Materials 3 22 170
Salaries 67 230 780
Overhead 27 89 310
Total 121 423 1540
Cost/sample (U.S.$) 121 42 15

are produced at an almost constant price per time unit, so that up to the
maximum throughput of the instrument the unit price of determination
or analysis decreases (Table 9-1). The cost criterion is relevant for almost
all technical features of a method. High throughput of services is a nor-
mal requirement of an applied geochemical laboratory and purchase of
automated and simultaneous instruments with high information capacity
is often economically justified. Sometimes, however, non-economic criteria,
like extreme requirements for detection limit, will necessitate the use of
costly procedures.
Maximum throughput or production capacity is a desirable characteristic
of an exploration analytical procedure. In general, and unlike assaying
laboratories, exploration geochemical laboratories should be organized for
high productivity. High throughput can be achieved by attention to the
following:
— Rationalizing the pretreatment of samples (use of manifold sieve
shakers and grinding equipment, etc.).
— Rapid methods of sample decomposition (typically partial leaches in
test tubes).
— Avoidance of chemical preconcentrations (low dilution factors, instru-
ments capable of low detection limits).
— Rapid determination of elements (simultaneous or "multichannel"
instruments with autosamplers instead of sequential ones).
— Computerized laboratory information management systems (LIMS)
including uniform identification of samples and rapid data retrieval.
— Close cooperation between laboratory staff, those who submit the
samples and those responsible for the interpretation, to allow efficient
scheduling of analytical requests.
Cooperation is the most effective means of reducing dead time of samples
in the field and data in the files. The achievement of maximum throughput
is a matter of proper laboratory management and R&D methods. Too often,
192 Analytical methods

throughput is regarded as a lesser scientific concern and its importance is


undervalued.
The turnaround, or delivery time, is closely related to the throughput of
a method, and the tools needed to achieve good turnaround are much the
same as for throughput. Shortest turnarounds are required for sequential
investigations where analytical data of one phase are needed for planning
of subsequent operations. This often applies to the detailed stage of
a phased geochemical study. Regional mapping programmes are more
dependent on high throughput than rapid turnaround, as data are typically
presented in extensive geographical blocks comprising thousands of sample
stations. A rapid turnaround needed for directing the expensive sampling
by heavy machines in the final stage of exploration or inventory of reserves
may, if often demanded, considerably limit the long term production
and R&D of the methods and hinder efforts towards cost-effectiveness.
"Blitz" requests should be authorized only by the management of an
organization. Rapid turnaround is easiest to accomplish with simultaneous
instruments that presume minimal amounts of sample preparation (e.g.,
simultaneous arc/spark optical emission spectrometers and "direct readers"
in general).

PRETREATMENT OF SAMPLES

Sample pretreatment is typically regarded as a tedious necessity rather


than as a scientifically important concern. In earlier days, a trainee or
some lower-level employee was given the task, usually in a remote corner
of the laboratory building. Underestimation of the pretreatment phase
has undermined the value of many geochemical efforts. Close attention
indeed should be paid to this first step in laboratory operations as it is
the foundation for good or poor quality analyses. Being a labour-intensive
operation, presenting certain health risks and often not subject to any
considerable automation, the pretreatment of samples is a bottleneck in
many modern laboratories.
The objective of the pretreatment step is, through maintaining or in-
creasing the representativeness of the submitted sample (the gross sample),
to provide a few tens of grammes of homogeneous sample. The pretreat-
ment comprises drying and sieving or crushing and grinding with an
adequate number of reducing and homogenizing steps. A suitably fine-
grained product ensures that the small subsample taken for actual analysis
(the test portion, typical size 0.005-1 g) will be representative. Moreover,
a fine-grained sample has a large surface area for reaction with decom-
position agents. The principal problem is how to reduce contamination of
samples during pretreatment and how to control unavoidable contamina-
tion. Materials worn from sampling equipment sometimes cause significant
Pretreatment of samples 193

TABLE 9-II

Geochemically important contaminants worn from different grinding materials. In parentheses


elements of secondary importance (based on the data by Sturhahn and Otto, 1974)

Material Critical compositional elements


Agate (Si0 2 ) Si, (Al, Na, K, Ca, Mg, Mn, Fe)
Zirconium oxide (Zr0 2 ) Zr, Hf, (Ca, Mg, AI, Fe, Ti,Si)
Sintered corundum (A1203) AI, Si,(Mg, Fe, Na, K)
Hardened steel (No. 1191) Fe, Mn, (C, Si, P, S)
Hardened chromium steel (No. 2080) Fe, Cr, C, (Si, Mn, I> S)
Stainless chromium-nickel steel (No. 4301) Fe, Cr, Ni, Mn, (Si, S, C)
Hard metal ("Widia", tungsten carbide, WC) W, C, Co, Ta, (Ti, Fe)

contamination of samples. A useful rule of thumb is to maintain the


nonavoidable "contamination line" and not to add new associations of con-
taminants during further treatment of the samples. Typical compositions
of some grinding materials are given in Table 9-II. In addition, soldered
joints are always suspicious, lubricants may contain Mo or Li and certain
plastics include considerable amounts of metallic additives (e.g., Zn, Cd, Pb,
Sb, Sn and Ba).
Glacial overburden, especially till, is a homogeneous material in com-
parison to the original bedrock. The representativeness of samples may be
increased still more by using the fine fraction of till (e.g. <0.06 mm grain
size) for the main analysis.
At the Geological Survey of Finland, three fractions of till are prepared
in the following steps:
— Drying (about 100 g of gross sample) in opened sample envelopes
made of kraft paper. Large heating cabinets (e.g., 500 1) are used with
temperature set at 70°C.
— Light disaggregation of samples in the envelope with a rubber hammer
(clay-size fraction cements during drying).
— Sieving in manifold shakers to produce fractions, e.g., —0.06 mm,
+0.06-0.50 mm, and +2.0 mm. Contamination during screening is mini-
mized by using nylon meshes mounted in nylon frames.
Stream sediment samples are pretreated in much the same way.
Conventional test sieves made of brass or stainless steel with possible
solderings must not be used for other than grain-size analyses. Operators
should not wear gold rings, as these easily contaminate samples analyzed for
gold in the ppb range (Kontas, 1991). Great care is needed to maintain an
overall cleanliness of the pretreatment area and to keep samples in proper
sequence when not in their containers, as during screening. Cleaning of
the pretreatment equipment and area is traditionally done by blowing and
suction of the dust, but washing and drying is preferable because of less
overall contamination of the laboratory air.
194 Analytical methods

For rapid dry sieving of samples an aperture of about 0.06 mm is a


practical minimum. Sieving to <0.06 mm grain size is also convenient
because such material can be used without further grinding for chemical
decomposing of the sample or exciting it in arc/spark emission spectrometry
Finer grained material is needed for X-ray fluorescence measurements,
however. For producing samples within a range of 0.005 to 0.1 mm,
ultrasonic sieving units are available where sieving is performed in media
like water or white spirit. Generally, fine fractions are favoured in regional
studies and coarser fractions are needed for identification of till lithologies.
Some investigators prefer total samples (whole sample crushed and/or
ground) for ore exploration at a detailed scale.
If the grain size of the original sample exceeds 10 mm and determinations
on the total sample are requested, samples are pretreated using the
standard procedure for bedrock samples. Jaw crushers with jaws and side
walls made of hardened steel are used to obtain 1-5 mm grain sizes.
The samples are then sized in a riffle sizer and ground to analytical
fineness in a swing mill (vibrating cup mill) in a vessel made of plain
carbon steel. Vessels made of Cr steel or hard metal should normally be
avoided because of the danger of contamination. Overloading of vessels is
a frequent mistake which produces sorted material and material too coarse
for analysis. The working area should be kept well ventilated and clean
to avoid cross contamination. Cleaning of equipment by grinding quartz
between samples is recommended. After crushing and grinding of assay
samples, a very careful cleaning is necessary; or, preferably, assay samples
should be prepared separately from geochemical samples.
Small sample quantities (0.1-10 ml) are difficult to grind. Specially
designed ball mills ("mixer mills" or "micro mills") with small grinding
cups may be useful.
The size reduction of organic samples, subsequent to drying, is ac-
complished by a very rapidly rotating impact mill or a cutting mill. For
incidental needs, a hand mortar or an inexpensive kitchen blender may
suffice.

DECOMPOSITION OF SAMPLES

General

Decomposition of samples or of test portions (sample preparation) pro-


duces sample solutions in which elements are liberated from their minerals
or compounds either totally or partially. A wide variety of reagents are
available for decomposing samples. No individual procedure is either abso-
lutely total for all minerals or totally specific for a certain mineral, or group
of minerals or compounds.
Decomposition of samples 195

Many analytical methods of interest including spectrophotometry, atomic


absorption spectrometry and plasma source spectrometry make use of
solution mode samples. A solution sample has the favourable properties of
generally providing a representative subsample. Using partial extractions
allows element determinations of certain phases of a sample.
Solid mode samples are typically used in arc/spark emission spectrome-
try, X-ray fluorescence spectrometry and neutron activation analysis. These
methods easily suffer from sample heterogeneity and enable only determi-
nations of total concentrations of elements. However, the heterogeneity of
samples may be overcome by analyzing solid solution preparations instead
of raw powders of samples.
It is also important to remember that although element determinations
can be carried out successfully by many alternative techniques, the concen-
tration level may be strongly dependent on the method of decomposition.
Thus it is more important to specify that, for instance, Cu was extracted by
aqua regia solution than that it was determined by AAS.
Whatever procedure of decomposition or other preparation of samples
is chosen, this tends to be the most labour-intensive step in the entire
process of analysis and the step that consumes the most chemicals. Thus,
the cost-effectiveness of a geochemical laboratory may easily be improved
by control of this step and the previous sample pretreatment.
Decomposition procedures for various inorganic and organic materials
are reviewed in the excellent textbooks of Sulcek and Povondra (1989) and
Bock (1972).

Total decomposition
Total methods of decomposition are used at least as a preliminary step
in determinations of the main components of silicates and some minor
and trace elements occurring in more resistant minerals. Two practical,
essentially total methods are (in order of effectiveness) fusion with borates
and digestion with hydrofluoric acid (HF) mixed with strongly oxidizing
acids.
Borate solutions are advantageous for atomic absorption and plasma
emission spectrometries, for example, lithium metaborate (LiB0 2 ) is a
powerful alkaline flux that forms solid solutions with most minerals, and
the solidified melt is easily dissolved in dilute nitric acid. The procedure
allows the use of rapid automated "fluxer" devices, but the costs of
operation are considerable because of the platinum crucibles required.
A typical fusion procedure as used in the GSF laboratory involves
weighing 100 mg of sample powder and 500 mg of LiB0 2 , homogenizing the
mixture, fusing for 5 minutes in the fluxer flame and then pouring the melt
into a beaker with 5% H N 0 3 solution. The melt is dissolved by mixing for
15 minutes and the solution is diluted with water to 250 ml (total dilution
196 Analytical methods

factor of 2500). A daily throughput of 30-50 sample solutions per operator


is achievable.
Decompositions based on HF digestion with various oxidizing acids are
usually not as complete as fusions with LiB0 2 but they are easier to carry
out. Use of hydrofluoric or boric acids with aqua regia as an oxidizing agent
is rapid and cost-effective for large batches of geochemical samples. The HF
dissolves most silicates, the aqua regia a great many sulphides, carbonates
and phosphates, and the B(OH)3 serves to dissolve the primary reaction
products and to neutralize the final solution. This procedure is ineffective,
however, for spinel group and Al 2 Si0 5 minerals and minerals like zircon,
cassiterite, rutile and staurolite. Thus, recoveries of Al, Cr, Fe, Si, Sm, Ti,
V and Zr may be low when refractory mineral phases are present. The
solution allows easy access to main components (viz. Si, AI, Fe, Mg, Ca, Na,
and K) and to at least 12 minor and trace elements (e.g., Ba, Co, Cu, Cr, Li,
Mn, Ni, Rb, Sr, Ti, V and Zn) through flame AAS or ICP-AES.
In the GSF laboratory we weigh 200 mg of sample powder into an
inexpensive polyethylene bottle, digesting with 2 ml of freshly prepared
aqua regia (2 h at 90°C), then digest with 3 ml of 30% HF (18 h at
room temperature), neutralize with 45 ml of saturated B(OH)3 solution
and dilute with 50 ml of H 2 0 to produce 100 ml of sample solution (total
dilution factor 500). A daily throughput of 80-100 solutions per operator is
feasible with the HF procedure.
Especially where multi-user geochemical mapping is carried out by a
government agency, some analysts insist that only the total contents of
elements should be provided since only they are verifiable (C. Frick,
personal communication, 1985). Such would be of very limited value, e.g.
for environmental studies.
Materials that are essentially organic (vegetation, humus, peat) are ashed
prior to analysis. Dry ashing, for 10-20 h at temperatures of 400-600°C, is
simple but causes matrix dependent losses of several elements of interest,
such as Ag, alkali metals, As, Au, Br, Co, Cu, Ge, Hg, Mo, Pb, Sb, Se, Te,
and Zn. The organic matter present in samples that are mainly mineralic
(stream and lake bottom sediments) can be destroyed by slow combustion
at 250°C without any serious losses, followed by acid leaching and analysis
by AAS (O'Leary and Viets, 1986).
Wet ashing ensures total recoveries of most elements in samples contain-
ing organic matter. Dried and ground material is digested with strongly
oxidizing acid solutions (e.g., H N 0 3 / H 2 0 2 or HN0 3 /H 2 S0 4 /HC104) in ei-
ther open or closed (pressure) vessels. With use of a microwave oven,
digestion times can be reduced, contamination minimized and acid fumes
collected.
Decomposition of samples 197

Partial extractions

Partial extractions make use of a wide variety of decomposition reagents.


As defined here, these dissolve, selectively or non-selectively, particular
phases of minerals or organic constituents of a sample. There are no
sharply defined differences between the selective, non-selective or even
total methods of decomposition, but the grade of selectivity is dependent on
sample type, element and matrix.
Partial extraction procedures need to be carefully defined operations
since the concentrations obtained are highly dependent on such method
parameters as:
— reagents and their concentrations,
— reaction temperature, time and mixing efficiency,
— grain size distribution of the sample, and
— mineralogy of the sample.
Unlike total concentration data, partial extractions often enhance geo-
chemical contrasts and reveal geographically broader anomaly patterns.
Enhancing contrasts in this way is advantageous if the anomalies are
known to be related to a particular source, e.g., mineral occurrences, but
otherwise may encourage faulty interpretation. Partial extractions — par-
ticularly the selective procedures — should never be used as an end in
themselves or before pilot studies.
Even selective extraction reagents are never specific to an individual phase.
However, the character of an anomaly may be appraised by choosing ap-
propriate selective reagents and particularly by applying them sequentially.
Selective methods typically liberate elements from ion-exchange positions,
organic matter, or secondary Fe-Mn oxides and sulphides. Some of the
common selective extraction reagents are listed in Table 9-III. Selective
extractions are used mostly for detailed-scale studies and are usually sup-
ported by other means of decomposition. Environmental studies especially
need both local and regional data on availability of elements, for instance,
elements in ion-exchange positions.
Sequential extraction systems are used for research purposes. Least
aggressive reagents are applied first to remove and estimate the amount
of organics and elements associated with secondary oxides. Then more
aggressive reagents are used to determine elements associated in residual
primary mineral phases. Fletcher (1981) presents an overview of useful
sequential extraction schemes. Sandström (1984) describes an application
to stream sediment samples.
Non-selective extractions are frequently employed in practical exploration.
In till studies at a regional scale, for example, use of hot, concentrated
strong mineral acid is preferred to total decompositions because sharper
contrasts and more homogeneous anomaly patterns are achieved; also,
elements such as Al and some of the Fe and Ca which would interfere
198 Analytical methods

TABLE 9-III
Association of trace elements with different phases of overburden samples and typical extraction
agents used

Phase Extraction agent


Soil solution water
Exchangeable ammonium acetate (buffered or non-buffered)
Organic hydrogen peroxide
sodium hypochlorite
EDTA (humus complexes)
Secondary iron and manganese oxides hydroxylamine hydrochloride
ammonium oxalate (Fe oxides)
sodium dithionite (crystalline Fe oxides)
Definite minerals potassium chlorate/hydrochloric acid (sulphides)
bromine/methanol (sulphides)
Silicates hydrofluoric acid
Refractory silicates and oxides alkaline fusions
non-destructive techniques (XRF or INAA)

^ ^ ^ ^ ^ ^ ^ ^ ^ ^ W \ |
H^^^^^^M^fi^^^^^ 9 ·H·· · · . ·
RΑ Ν Ι Τ Ε
[iVc ^:-^^Ζνά\ ''T<'& Y?-£v\? ς/
ρ Ι Ι β ^ Ι Ι ^ ·· ' ■
•• ·· · .· * · *

f iVJMICA SCHlSJ^'^^y^^^Y · · ·

^$S^^$00^M · j ^ / # . · · · ·
• · ·
^^^^^00M^0ips. · · RAPAKIVI# A

^ ^ ^ ^ ^ ^ ^ ^ ^ ^ ^ ^ ^ ^ > A *
e
Ky^/^^^^^^^^GRANODIORiTE^^^e · . Ä · LAPI#ÄRV*
, · · ·
• · ·
\///////////// · • • ·

WMMMMMMM: ■
·

i^^^^^e^' ·


• . · · · . *

10 km

Fig. 9-3. Total concentration of Ba (HF-aqua-regia dissolution) in till samples (-0.06-mm


fraction) in Askola map sheet area, southeastern Finland (30 x 40 km). High Ba is associated
with the rapakivi granitic area. As a diadochic substitute for K, Ba is mostly incorporated in
potassium minerals of which K-feldspar is abundant in rapakivi.
Decomposition of samples 199

. ·

^GRANITEx

#
b~JMICA SCHIST -V^-'c''/V^7>-^Ä/ · * #

RAPAKIVI

; GRANODIORITE
LAPINJÄRVt

1 0 km

Fig. 9-4. Aqua-regia soluble concentration of Ba in the same samples and area as in Fig 9-3.
High Ba is associated with the mica schist area because Fe-rich biotite — the major source
of Ba — is very abundant in the schist and easily soluble in aqua regia. Here Ba contents in
the rapakivi area are relatively low due to very poor solubility of K-feldspar in aqua regia. Of
course there are small amounts of biotite in rapakivi and microcline in mica schist.

with the analysis of trace elements are largely separated into the insoluble
residue. Traditionally, partial extraction data have been sought for elements
forming sulphides; however, main component data for silicates may give
useful background information for interpretation (Noras and Kontas,
1989). Essentially total (HF/aqua regia decomposed) and partial (aqua
regia extracted) concentrations of barium are compared in Figs. 9-3 and
9-4. The aqua regia extraction has proved to be a useful (and "universal")
method for many purposes and sample types. The efficiencies of different
acid digestion procedures for the extraction of elements from silicates are
discussed in Foster (1974).
As applied in the GSF laboratory, the aqua regia procedure involves
weighing 500 mg of sample powder into a test tube, adding 2 ml of cone.
HC1 and 1 ml of cone. HN0 3 , digesting at 90°C for 2 h with frequent
agitation, and diluting to 20 ml with water. The element determinations
are done by flame atomic absorption (Ag, Cd, Co, Cr, Cu, Fe, Mn, Mo, Ni,
Pb and Zn) or plasma emission (the above elements plus Al, As, Ba, Ca, K,
La, Li, Mg, Na, I? Sc, Sr, S, Th, Ti, V and Y). A daily output of 150 sample
solutions per operator is achievable.
200 Analytical methods

DETERMINATION OF ELEMENTS

General

There is no simple answer to the question which analytical method


should be selected for which element. Not only will the performance of a
technique vary from one element to another, but the concentration range
and matrix of the element, differing from one sample to the next, may
be important to the choice of method. Furthermore, the availability of
some special facility, such as a nuclear reactor for activation analyses, may
determine the choice of complementary techniques.
In a study of the frequencies of analytical techniques used for the
analysis of some international standard reference rocks in 1970/80's, Potts
(1987) divided the routinely determined trace elements into the following
methodological categories:
(1) A core of trace elements, including Ba, Co, Cr, Cu, Ni, Rb, Sr, Th, \ζ
Zr, are overlappingly determined by XRF, NAA, AES and AAS.
(2) The rare earth elements are normally determined by NAA or ICP-
AES. Higher contents of the light REEs are less frequently deter-
mined by XRF.
(3) Certain groups of elements are mostly determined by one technique
only:
XRF: Zr, Nb, Ga, Pb, U, Y
NAA: Ta, Hf, Cs, Sc, U
ICP-AES: Be, Y, Zr, Ga, Sc
Ion-specific electrode: F
(4) A large group of trace elements are not determined routinely at all:
Ge, Se, Br, Mo, PGE, Ag, Cd, In, Sn, Sb, Te, I, W, Re, Au, Hg, T1 and
Bi.
In the case of glacial overburden samples (and geochemical studies in
general), the analytical requirements may modify the categories consid-
erably. For instance, many of the elements in category 4 are frequently
requested and are then typically determined by flameless atomic absorption
spectrometry or, more recently, by plasma-mass spectrometry.
Almost any multi-element geochemical study requires the use of a
multi-element analytical method complemented with one or more other
methods. Analyses for regional multi-purpose mapping programmes should
always be based on a multi-method, multi-element approach. Inductively
coupled plasma emission spectrometry (ICP-AES) is the single most ver-
satile multi-element method, providing reasonably low detection limits for
many interesting trace elements and good dynamic ranges for the main
components. If the capabilities of ICP-AES are joined with flameless atomic
absorption spectrometry, most elements and abundance ranges encountered
Determination of elements 201

TABLE 9-IV

Evaluation of capital outlays (thousands of U.S.$; "other" includes sample preparation equip-
ment, etc.) for acquiring different analytical techniques, annual throughputs (thousands of
element determinations) and unit prices (U.S.$) of commercial services

Technique Capital outlay Annual Commercial prices


throughput (U.S.$)
instrument other
(1000U.S.$) (lOOOdet.'s)

Spectrophotometry 5-10 very low 5-15 3-15/element


Atomic absorption
flame 40-100 low 100 1-5/element
flameless 80-130 low-interm. 10-25 3-10/element
Emission spectrometry
ICP-AES 150-250 interm.-high 500 10-25/30-element package
Plasma-mass spec- 350-500 high 50-100 40-50/20-element package
trometry
X-ray fluorescence 300-400 interm.-high 100-500 40-50/25-element package
spectrometry
Neutron activation thousands very high 200-500 20-30/30-element package
analysis

will be covered. Other useful multi-element techniques besides ICP-AES


are instrumental neutron activation (INAA), plasma-mass spectrometry
(ICP-MS) and X-ray fluorescence spectrometry (XRF). The choice of INAA
depends on the availability of production oriented and reasonably priced
reactor services. However INAA fails to determine some indispensable
elements. ICP-MS is unrivalled for most sub-ppm trace elements, including
the precious metals, REE, and many "non-spectroscopic" elements, and
is capable of determining isotope ratios. The standard XRF methods are
limited to higher detection limits, but XRF has some advantages in tol-
erating extreme sample matrices and refractory minerals, and if desired,
non-destructive sample preparates. Some economic considerations of useful
instrumental techniques are evaluated in Table 9-IV
In general, it is of principal importance that the laboratory specialize
in at least one of the multi-element techniques and the related sample
preparation, chemical separation and preconcentration methods, so that it
exploits available technical capabilities. This especially applies to smaller
laboratories that are only equipped with, say, an atomic absorption spec-
trometer and a spectrophotometer. In the hands of skilled laboratory staff,
even these minimal instruments are useful and certainly economical tools
for many geochemical studies.
The most widely used instrumental techniques and their relevant ca-
pabilities are described in the following. Some individual methods not
202 Analytical methods

touched upon here are used for measurements of fluoride, chloride and
pH (ion-selective electrode measurement), and sulphur, carbon and water
(combustion-infrared absorption measurement). Microbeam and surface
analytical techniques such as electron microprobe, ion probe (SIMS), laser
probe and particle induced X-ray emission (PIXE) are used in special cir-
cumstances only and likewise are not included in this overview. Except for
ICP-MS, the use of mass spectrometric techniques is still far too expensive
to have any practical role in the analysis of exploration samples.

Spectrophotometry and fluonmetry


Spectrophotometry, or rather colorimetry, was the prevailing method
for exploration geochemical laboratories from the early 1950's to the late
1960's. Even now after the introduction of atomic absorption spectrometry
and multi-element methods, certain elements are still determined by these
single-element techniques. Although the instrumentation is generally low
priced, the running costs may be considerable. High throughput is only
achieved with the simplest procedures.

Principles
Since most colorimetry indicators are non-specific for a particular el-
ement, much effort must be directed to minimizing interferences from
matrix elements. Sample solution cells for instruments are of interchange-
able type or fixed flow cells, the latter being recommended for large batches
of samples or when several procedures are used. Relatively broad molecular
absorption bands are measured, which means that high optical resolution
is not required of the instruments. Spectrophotometric procedures involve
seven steps: decomposition of the sample, transfer of an aliquot of the
sample solution, addition of a buffering agent, addition of a masking agent,
addition of a colorimetric indicator, extraction of the coloured compound
and measurement of the intensity of colour (colorimetry) or the absorbance
of its complementary colour (spectrophotometry). The element concentra-
tions are obtained by comparing the colour intensity against a series of
standards (colorimetry) or by applying Beer's law (spectrophotometry).

Applications
Relative to the capabilities of other instrumental methods, spectrophoto-
metric procedures continue to be useful for certain elements as summarized
in Table 9-V
Fluorometry measures the intensity of fluorescence emitted by some an-
alyte compounds after their exposure to UV radiation. The only important
application of fluorometry in exploration geochemistry is the determination
of uranium. Neutron activation with delayed neutron counting for U is less
sensitive to matrix effects and should be more cost-effective, if available.
Determination of elements 203

TABLE 9-V

Selected spectrophotometric procedures for geochemical overburden samples

Element Colorimetric indicator Reference


Arsenic Mercuric chloride (Gutzeit test) Stanton(1976)
Ag diethyldithiocarbamate Welsh (1979)
Copper 2,2/-biquinolene Stanton(1966)
"Heavy metals" Dithizone Stanton (1976)
Niobium Thiocyanate Ward et al. (1963)
Tin Gallein Stanton (1976)
Dithiol Cogger(1974)
Tungsten Dithiol Welsh (1983)
Thiocyanate Ward et al. (1963)

However, the fluorometric method allows easy use of partial extractions. A


comparison of the two methods for uranium can be found in Garrett and
Lynch (1976).
Recent developments in spectrophotometric methods have primarily
involved new indicators, separation schemes and automation. Flow injection
analysis (FIA) is now being applied for high throughput duties; Fuge and
Andrews (1985), for example, present an automated method for fluorine.
The manuals written by Ward et al. (1963) and Stanton (1966, 1976) give
detailed spectrophotometric procedures for many applications of interest.

Atomic absorption spectrometry


Atomic absorption spectrometry (AAS) is the principal analytical tech-
nique of most laboratories involved with exploration geochemistry. AAS is
characteristically a rapid method for analyzing large numbers of samples,
but for a limited range of elements. However, slower procedures and tech-
nical options are available for most elements of the periodic table. In the
1960's AAS took over many of the earlier duties of colorimetry because
of its usefulness in determining elements of then prevailing interest (i.e.,
sulphide-forming metals). The new technique was easily adopted since it
utilizes similar sample solutions to colorimetry. In the largest exploration
programmes, plasma atomic emission spectrometry has replaced flame-
mode AAS in some of its traditional tasks. Flameless AAS, which continues
to undergo rapid development, is unsurpassed for determining economically
very low concentrations of numerous elements.

Principle
All free, ground state atoms are capable of absorbing energies of charac-
teristic optical wavelengths. The AAS instrument basically comprises (a) a
204 Analytical methods

modulated light source, usually a hollow cathode lamp, which emits light of
characteristic wavelengths; (b) an atomizer which releases free atoms from
sample solutions; (c) a monochromator which isolates the resonant wave-
lengths and (d) a detector (photomultiplier) which measures the absorption,
proportional to the concentration of an element (Beer's law). A notable
technical limitation of the method is the need to use an individual light
source with characteristic spectral lines for each element to be determined.
The availability of a continuum light source, still not completely solved,
would make the AAS a multi-element method.
Three types of atomizers are in general use: the gas flame, the hydride
generation cell and the graphite furnace.

Applications
Flame atomic absorption spectrometry (FAAS) is a widely used funda-
mental method. Based on solution injection into a laminar flow nebulizer
and atomizing in a gaseous flame, the flame techniques are straightforward
to apply, though not the ideal means of atomizing. The initial cost of
the instrumentation is relatively low, and simply prepared solutions can
be used for important ranges of elements (most of the sulphide-forming
elements and all of the main silicate-forming elements). Two applications
are of special interest for overburden samples.
(1) Hot, strong mineral acid leaching of samples (e.g., aqua regia) enables
daily batches of up to 100 sample solutions with determinations of Ag,
Cd, Co, Cu, Fe, Mn, Ni, Pb and Zn in an air-acetylene flame, and of Cr
and Mo in a nitrogen oxide-acetylene flame. All are easily detected at their
normal levels of background in till samples, excepting Ag (detection limit
1.5 ppm), Cd (1 ppm), Pb (15 ppm) and Mo (1 ppm). Unfortunately, these
latter elements are also prone to interferences in emission spectroscopy
(2) Hydrofluoric and boric acid solutions, as described in the section
"Total decomposition", provide a useful matrix for determinations of Si, Al,
Ti, Fe, Mg, Ca, Sr and Ba in nitrogen oxide-acetylene flame (Sr in emission
measurement mode) and Mn, Na, K, Li and Rb in air-acetylene flame.
As compared with ICP-AES determination of these elements from similar
solutions, the FAAS shows better precision for Si and the alkali metals.
Expanding the use of the flame methods to other elements generally calls
for chemical separations and enrichment, most conveniently by solvent
extraction. Methyl isobutyl ketone (MIBK) has proved by far the most
useful organic extractant for applied geochemical analysis. O'Leary and
Viets (1986), for example, have presented a group separation approach
involving a hydrochloric acid-hydrogen peroxide leach and extraction into
Aliquat 336-MIBK. Then, up to nine elements are determined in the
organic phase by FAAS, viz. Ag, Cd (detection limit 0.05 ppm); Mo, Cu,
Pb, Zn (0.5 ppm); Sb, Bi (1 ppm) and As (5 ppm). Gold is determined (0.2
ppm) in samples large enough to be representative (10 g or more) after
Determination of elements 205

digestion with aqua regia-KBrOß and extraction into MIBK (Noras, 1992).
The determination of Au in MIBK can be done by graphite furnace AAS
to achieve a detection limit of below 0.001 ppm, as presented by Meier
(1980). Other well-established procedures making use of MIBK have been
presented for In, Te, Tl (detection limit 0.2 ppm) and Au (0.1 ppm) in aqua
regia solution (Hubert and Chao, 1985), and for Sn (2 ppm) in MIBK-TOPO
phase after condensation with ammonium iodide (Welsh and Chao, 1976).
Hydride generation atomic absorption spectrometry (HGAAS) can be
applied for volatile, covalent hydride-forming elements (viz. Ge, As, Sb,
Bi, Se, Te, Sn and Pb) that are only sparingly soluble in water. Use of a
hydride generation cell involves the selective separation of elements and an
efficient means of atomizing compared with normal nebulizing. Attractive
lower limits of detection are offered in a typical range of 0.05-0.5 ppm. Un-
fortunately, the generation reactions (usually with sodium borohydride) are
prone to interference from many transition group elements which suppress
the yield. Hydride generation is used for preconcentrating of Se prior to
determination by graphite furnace AAS (Willie et al., 1986). The HGAAS
method (and HGICP-AES) allows automation and implementation of flow
injection techniques which are sure to become important in the future.
Guo et al. (1989) have used hydride generation prior to determinations by
atomic fluorescence spectrometry (HGAFS). Mercury has been determined
after reduction to non-volatile form by several well-established flameless
methods closely resembling HGAAS. A review of these procedures is pre-
sented in Ure (1975). Portable Hg determinators are available for field use
(Scintrex Ltd, Concord, Ont., Canada).
Graphite furnace, or electrothermal atomic absorption spectrometry
(GFAAS), has gained attention for attaining lower detection limits, often
by one or two orders of magnitude, than those attainable by FAAS and
other commonly used techniques in exploration geochemistry. The essence
of GFAAS is total vaporization of a small quantity of sample solution
into an electrothermal furnace atomizer. Transient absorption signals are
measured rather than constant signals as in FAAS. Direct injection of multi-
element solutions is typical for FAAS, whereas GFAAS usually requires
prior separations and preconcentratings for single elements. Interferences
arise mainly from non-isothermal atomization and are due to matrix
elements. Overcoming of interferences typically requires spike calibration,
matrix modification, special treatment of graphite, platform technique
(i.e., isothermal atomization) and/or the use of integrated (peak area)
absorbances. Thus, working with GFAAS calls for more professional skill,
and the operating and capital costs are higher.
In view of the other methods available, the elements most usefully
determined by GFAAS for exploration purposes would appear to be Au, Pd,
Se and Te. Kontas (1981) has presented a powerful procedure based on
relatively simple aqua regia digestion and coprecipitation of the analytes
206 Analytical methods

with SnCl2-Hg for Au (detection limit 0.0001 ppm). The method is rapid —
daily throughput can be up to 100 determinations — and it has successfully
been applied to more than 200,000 till samples in the Nordic countries. The
procedure has recently been modified to allow measurements of trace levels
of Ag, Pd, Pt, Rh, Se and Te (Niskavaara and Kontas, 1990). Selenium
(detection limit 0.05 ppm) can be determined after nitric, perchloric and
hydrofluoric acid decomposition and extraction into toluene (Sanzalone
and Chao, 1981); and Te (detection limit 0.004 ppm) can be determined
by hydrobromic acid-bromium digestion prior to extraction with MIBK
(Chao et al., 1978). Other useful single element procedures have recently
been reported for Ga (Anderson et al., 1985) and In (Zhou et al., 1984).
Many separation and preconcentration procedures developed for FAAS are
applicable for GFAAS, too. The separation and preconcentration methods
are reviewed in the textbooks of Stary (1964) and Minczewski et al. (1982);
practical procedures for geological samples are given in Stoch and Dixon
(1983).
Slavin (1984) has published a comprehensive source book on GFAAS
methods. Complete bibliographies and state of the art developments,
including all AAS techniques, are presented in the Atomic Spectroscopy
handbook (published by Perkin-Elmer Corp., Norwalk, CT, U.S.A.).

Emission spectrometry
Optical emission methods with excitation of atoms in a direct current arc,
have been applied for a wide range of geochemical materials over the past
four decades. As a replacement for the traditional photographic recording
of spectra (optical emission spectrography, OES), convenient direct reading
spectrometers with several photomultiplier "channels" were introduced in
the late 1960's. Then, in the late 1970's the electrical arc/spark sources of
emission gave way to plasma sources (either inductively coupled or direct
current plasmas). In larger geochemical laboratories today the commonest
workhorse among instruments is the inductively coupled plasma atomic
emission spectrometer (ICP-AES) which is much preferred to the direct
current plasma spectrometer (DCP-AES).
Emission spectroscopy, from spectrography to plasma spectrometry, offers
a rapid, virtually simultaneous multi-element technique — highly valuable
for a geo-analytical instrument. Extra cost effectiveness is provided by the
minimal requirement for sample preparation, though this does not always
apply to plasma AES.

Principles
Atomic absorption and atomic emission involve changes in the energy
levels of an atom's outer electrons. In the emission spectrometer, ground
state electrons of an atom are excited in the source to higher energy
Determination of elements 207
TABLE 9-VI

Common sources for atomization and excitation in atomic spectroscopy, abbreviations, and
epithets of instrumental arrangements used

Atomization/ Emission Absorption Fluorescence


Excitation spectro- spectrometer, spectrometer, spectrometer spectrometer
graph monochromator polychromator
Flame - flame - FAAS AFD
photometer
Arc (spark) arc OES DR-OES - -
Plasma - ICP-AES ICP-AES - ICP-AFS
Graphite
furnace - - - GFAAS* -
Hydride
generation - HGICP-AES - HGAAS HGAFS
Epithets manual, film sequential, simultaneous, * nameless or
recording scanning direct reading, electrothermal
multi-channel, AAS
Quantometer

levels, which soon collapse back to lower energies with an emission of


photons. Each element has its own characteristic line spectrum, which is
reproduced by a grating. The concentration of the element is calculated
from the intensity of one of its emission wavelengths. Because the emission
spectra of geological materials are extremely complex, often containing
thousands of lines, expertise is required in the operation of spectrographs,
or else sophisticated spectrometers. Several categories of emission and
other atomic spectrometers can be distinguished (see Table 9-VI).

Applications
The DC arc spectrography (DC arc OES) system comprises a gitter
spectrograph, photograph processing facilities and a microdensitometer for
measuring line intensities. For rapid, semiquantitative analysis, lines can
be visually compared with those of standard films or plates. Owing to poor
total cost effectiveness, poorer limits of detection, small size of sample
analyzed and non-manageable spectral and matrix interferences, the DC
arc OES does not compete with recently available techniques. With direct
excitation of geosamples, reasonable limits of detection (2-5 ppm) are
achieved for Be, Sc, V, Cr, Co, Ni, Cu, Ga, Rb, Y, Ag, Sn, Yb and Pb. Without
special procedures the method fails in determining important elements like
F, Zn, As, Mo, Cd, Sb, La, W, Bi and U at their background levels. The DC
arc OES is especially used for determining certain elements associated with
resistant minerals, e.g., Be, B, Nb and Sn. Spectrography is still frequently
used in the the former U.S.S.R. Much of the geochemical mapping carried
208 Analytical methods

out in China is based on the recently improved spectrographic method of


Shen (1989). Authoritative textbooks on emission spectrography are those
by Ahrens and Taylor (1961) and Mitchell (1964a).
Simultaneous spectrometers (direct readers) with arc, spark or mixed
excitation sources are widely used in the metal and cement industries and
for geochemical applications where speed is the overriding consideration.
Separate photomultipliers are typically mounted on the focal circle for 20-
30 carefully chosen element lines, and automatic feeding of samples into
the source has been developed. Hence, simultaneous spectrometers possess
excellent information capacity and are easy to use if the types of sample
materials and suite of requested elements remain constant. Care is needed
in managing the inherent spectral and matrix interferences, and especially
analytical drift and rapid background shifts associated with instabilities of
the source. Simultaneous spectrometers should therefore be supported by
computerized performance control. Unfortunately, with electrical emission
sources and sample powders there are a number of serious interferences
not amenable to correction.
During 1973-1989 about ten million determinations on 500,000 samples
have been made with a simultaneous emission spectrometer at the GSF
chemical laboratory. The instrument is tailored for fine fractions (—0.06
mm) of till and is equipped with an automated sample feeder (tape machine
from Danielsson et al., 1959), an interrupted arc source, a polychromator for
30 elements and a microcomputer for versatile concentration calculations
and performance control of the analyses. Tape feeding would not be
satisfactory for less homogeneous sample types since less than 1 mg of
powder is burned (actually analyzed) in the source.
ICP-AES will certainly remain the major emission technique for explo-
ration laboratories in the 1990's. Unlike conventional emission methods,
but like AAS, analysis is primarily of nebulized solutions, which ensures
homogeneity of the analyte solution. Solid sample injection through nebu-
lization of slurries, laser ablation and ramp heating has been studied, and
these techniques may have a role in the future, in ICP-AES and partic-
ularly plasma-mass spectrometry Compared with conventional excitation,
the ICP offers better reproducibility in sample introduction, easy prepa-
ration of matrix-matched standard solutions, better stability of excitation,
and wider determination ranges. In this way, when equipped with advanced
correcting systems, the ICP-AES is subject to fewer non-amenable inter-
ferences than DC arc OES. Modern ICP-AES instrumentation incorporates
integrated computer controls and software developed from earlier direct
readers. In many of its features the ICP-AES may actually be superior to
all other analytical methods, but even then it is no universal panacea for
exploration geochemistry. Direct determination of important elements such
as Pb, Mo, Ag, U, \\ζ Cd, Sb, Bi and As is not straightforward at back-
ground levels, mainly due to the high spectral background continuum from
Determination of elements 209

the main components, especially Al. Most important, any nebulizer design
still suffers from blockage and salting-up problems which are difficult to
overcome.
Simultaneous ICP-AES instruments, which readily allow multi-element
analyses and high sample throughput, are most appropriate for the explo-
ration laboratory.
A typical polychromator assembly, as employed in the GSF laboratory,
comprises channels for 33 elements: Ag, Al, As, Ba, B, Ca, Cd, Co, Cr,
Cu, Fe, K, La, Li, Mg, Mn, Mo, Na, Ni, P, Pb, Sb, Sc, Si, Sr, Th, U, V, W,
Y, Yb, Zn and Zr, and one scanning channel. Sulphur can be determined,
but a vacuum path spectrometer is then required instead of the normal,
air path instrument. Although not a full response to client demand, the
33 channels were installed because they exploit most technical capabilities
of the spectrometer in direct analysis of samples. The two methods of
sample preparation comprise HF-aqua regia-B(OH)3 decomposition (which
is essentially total) and aqua regia extraction. By the latter procedure, most
of the interfering matrix elements in ICP-AES (e.g., Al, Ca and Fe) are
avoided, since they are largely left in the insoluble residue. Recently, there
has been much interest in extractability data such as ratios of aqua regia
soluble content to total content of an element, which are easily obtainable
by simultaneous ICP-AES but with difficulty by any other instrumental
technique.
Sequential ICP-AES is slower, mechanically more complex and requires
more computing time than the simultaneous instrument; but it enables
flexible measurement of an unlimited number of spectral lines and thus
any element. Sequential ICP-AES may be the choice of smaller and less
production-oriented laboratories (or for owners of a simultaneous unit) that
deal with widely varying geochemical problems.
Direct analysis of solutions by ICP-AES is a well established method;
current developments include new methods to separate and preconcenträte
low abundance elements. Above all, procedures designed to separate groups
of elements are invaluable for ICP-AES. Useful procedures have been
presented for hydride-forming elements (Thompson et al., 1978; Halicz
and Rüssel, 1986), for REEs (Walsh et al., 1981) and for PGEs plus Ag,
Ta, Th and U (Barnes and Diallo, 1985). Some of the separation and
preconcentration procedures developed for AAS can be applied as such or
modified for ICP-AES, too.
A comprehensive ICP Handbook with emphasis on geochemical samples
has been published by Thompson and Walsh (1983). Complete annual
bibliographies and state of the art developments are presented in the ICP
Information Newsletter (edited by R.M. Barnes).
210 Analytical methods

Plasma-mass spectrometry
The first inductively coupled plasma-mass spectrometer (ICP-MS) was
brought to a reliable operating state by Date and Gray in 1983. The
technique is considered to be an invaluable alternative to such conventional
techniques as AAS and AES, which are prone to the common interferences
in all atomic optical spectra. As a multi-element, wide dynamic range
method, ICP-MS resembles ICP-AES, but in providing very low detection
limits it approaches some capabilities of GFAAS, and of INAA in particular.
The sample changeover of about one sample/minute is as rapid as in
ICP-AES. The mass analyzer enables for the first time "affordable" access
to isotope ratios. Many larger geochemical laboratories that today rely
on multi-channel ICP-AES, extensive use of GFAAS and contracted INAA
services will be equipped with ICP-MS during the 1990's even if it still is
rather expensive. Changing over from AAS through ICP-AES to ICP-MS is
logical and relatively easy because the sample pretreatment and preparation
requirements are similar for all three techniques. Some applications of ICP-
MS, however, call for clean room arrangements.

Principles
The ICP is not only an efficient source of photons as in ICP-AES,
but also a highly useful ionization source. In the ICP-MS, subsequent
to nebulizing of sample solutions, ions are extracted from the central
channel of the plasma through a sampling aperture into a quadrupole
mass analyzer. Ions are separated by their mass-to-charge ratios prior to
detection by an electron multiplier and accumulation in a multi-channel
analyzer. In contrast to atomic optical spectra, mass spectra are relatively
simple because most ions occur only as singly charged and mono-atomic.
However, interference due to polyatomic ions generated from solution
chemicals or plasma gases, or by formation of refractory oxides is possible.
Doubly charged ions are observed only for elements with low second
ionization potentials. Most interferences are manageable by appropriate
data processing.

Applications
Comparison of approximate detection limits obtainable with other major
analytical techniques shows that ICP-MS may be a very competitive
alternative, especially for geochemical background levels of Sb, Bi, B, Cd,
Ga, Ge, Hf, I, In, Mo, Pb, Nb, Rb, Ag, Ta, Te, Th, Tl, Sn, W and U.
For direct determination of the complete REE group and Y, the ICP-MS
method is straightforward down to 10 times chondrite abundances. With
well-established separation and preconcentration by ion exchange, levels
down to 0.01 times chondrite can be achieved (Jarvis, 1988). Gold and PGE
can be determined by using lead or nickel sulphide as collector. Using a Pb
Determination of elements 211

button, Denoyer et al. (1989) have obtained detection limits as follows: Au,
Ru and Rh (1 ppb), Pt and Pd (0.1-1 ppb) and Ir (0.2 ppb).
Applications to isotope ratio studies have been presented for B, Os and
Pb, and also for Sm, Sr, Tl, U and Zn. Determination of isotope ratios
for Pb has attracted most interest (Longerich et al., 1987). In addition,
isotope dilution procedures offer the best way of calibrating ICP-MS (Hall
et al., 1987). Nevertheless, the precision of isotope ratio measurements
does not usually rival thermal ionization mass spectrometry (TIMS) used
in geochronological studies.
As with ICP-AES, the efficiency of sample introduction into the ICP is the
critical factor in the performance of ICP-MS. Thus, mainly to achieve lower
detection limits, alternative methods of sample introduction — electrother-
mal vaporization, laser ablation, arc nebulization, direct sample insertion
and nebulization of slurries — are being developed (Riddle et al., (1988).
The methodology of ICP-MS for geochemical application has been re-
viewed by Riddle et al. (1988), but being at a stage of development new
methods are continually being published.

X-ray fluorescence spectrometry


X-ray fluorescence spectrometry (XRF) is a non-destructive instrumental
technique, which has been commercially available since the late 1950's. XRF
produces inherently total concentration data on powder, pressed pellet or
fused bead mode analytes; use of solutions is possible but impractical. Some
spectrometers are designed for simultaneous multi-element analysis of
utmost rapidity. In general, high capital outlay, the need for a solid analyte
and special sample pretreatment requirements together with relatively poor
detection limits mean that XRF is employed, jointly with other techniques,
mostly for large multi-element geochemical mapping programmes and less
for selected, incidental studies.

Principles
The X-ray fluorescence phenomenon results when the analyte is irradi-
ated with X-rays (primary radiation), which are resonantly absorbed by an
atom, causing transitions in its inner electron shells. An amount of energy
equal to the difference between the original and final states of the electrons
is released as fluorescent X-ray energies or wavelengths (secondary radi-
ation) which are characteristic of the particular element. The intensities
of the secondary lines are proportional to the element concentration. The
primary radiation is produced by an X-ray tube, a radioisotopic source or
electrons (electron microprobe). The spectrometer separates the secondary
radiation on the basis of its wavelength (analyzing crystal) or energy (e.g.,
Li-drifted Si detector). The intensity of the secondary radiation is measured
by a detector or as the amount of energy stored in a multichannel analyzer.
212 Analytical methods

Flat analyzing crystals are used in sequential XRF (two changeable detec-
tors), and focusing crystals in simultaneous XRF (up to 30 fixed detectors).
Wavelength dispersive spectrometers (WD-XRF) are mechanically more
delicate, expensive and less compact than energy dispersive spectrometers
(ED-XRF), but they give better resolution of lines and lower detection than
ED-XRF.
The penetration of X-rays into the analyte is typically a fraction of a
millimetre, which requires high representativeness of the prepared sur-
face (compared with solution analyte techniques and neutron activation).
Quantitative analyses require the comparison of fluorescent intensities of
the sample with those of similar standard materials. By using multiple re-
gression for matrix corrections, the number of these comparative materials
can be considerably diminished. As the physical phenomena in the XRF
analysis are easily predicted, the concentrations can be calculated merely
by theory without use of standards (i.e., fundamental parameters method).
This is a significant advantage of the XRF technique.

Applications
Despite its limitations, simultaneous WD-XRF is an efficient multi-
element instrument for systematic long-term surveys. For most elements
the lower limit of detection ranges from 5 to 20 ppm — always much higher
for light elements, and somewhat higher for simultaneous than sequen-
tial instruments because of the limited possibility for matrix correction in
"fixed channel" mode. Today, the more flexible sequential spectrometers are
approaching the speed of simultaneous instruments; "hybrid" units (com-
bined simultaneous/sequential) have been developed, too. Pressed pellets
with wax binding or sample powders (5-50 g) are used for exploration ap-
plications; better accuracy and precision but poorer detection are achieved
by using fused beads (e.g. solid solutions with borates). The relatively
large sample weight compensates for the shallow penetration of X-rays
into the sample. In comparison with other instrumental techniques, the
non-destructive feature of WD-XRF is advantageous for determining Cr,
La, Nb, Sn, Ta, V, W and Zr which tend to be found in resistant minerals.
Low detection limits may be an asset for Cl, (Nb, Ta) and U. Simultaneous
determination of other elements (e.g., As, Br, Cu, Co, Cs, Mo, Ni, Pb,
Rb, S, Sb, Sr, Th, Y and Zn) is convenient but possible only at non-trace
level. Some serious interferences that reduce the quality of trace element
determinations in AES methods are not a problem in the XRF method.
As most of the above elements represent higher atomic numbers and can
be determined with the ED-XRF (Ag tube), this instrument may be a
cost-effective alternative to simultaneous or sequential WD-XRF (Potts et
a l , 1985).
XRF allows extensive automation from sample preparations to data
readout and can easily be operated unattended, which decreases running
Determination of elements 213

costs considerably. A throughput of up to 600 samples and 15,000 element


determinations/day with a simultaneous spectrometer has been reported
by Willis (1985). In production-oriented laboratories the high capital outlay
for such an instrument is frequently recovered within one or two years
of operation. The fundamental methodology of XRF has been described
thoroughly by Jenkins et al. (1981) and reviewed, for instance, in Markowicz
and Van Grieken (1986).

Neutron activation analysis


Neutron activation analysis (NAA) is the only radioanalytical technique
extensively used in exploration geochemistry. Not only are its capabilities
for determining low abundance elements almost infinite, but it is a powerful
multi-element technique. For economic reasons, use of direct instrumental
methods (INAA) is favoured. The INAA has enjoyed wide use in geochem-
istry since the late 1970's. Mostly it is employed selectively for sample types
and elements where other techniques cannot achieve the detection limits
attainable with INAA. Like those of XRF, the NAA standard procedures do
not allow solution mode samples. The availability of NAA is almost entirely
dependent on government research centres equipped with nuclear reactors
operating on the 235 U fission chain. There are about 140 such research
centres in 50 countries, operating one or more irradiation facilities.

Principles
Neutrons interact with the nucleus of an element to produce a beta
decayable radioactive isotope. Following the beta decay the residual nucleus
is in an excited state, which then collapses by emitting characteristic
gamma ray energies. The measurement of these energies is based on
gamma spectrometry using Ge(Li) or low energy photon detectors and
multichannel analyzers for storing the information. Different irradiation
times, decay and counting intervals are used for particular groups of
elements. The two analytically important energy ranges of the fission
neutron energy spectrum are the thermal (below 0.5 eV) and the epithermal
(0.5 eV-0.01 MeV). As both neutrons and gamma rays (100-3,000 keV) are
highly penetrating, representativeness of the analyte is achieved with less
effort than in XRF methods.

Applications
With the sequential, relatively expensive and slow procedures of thermal
INAA, it is possible to determine up to 30-35 elements in exploration
samples. Laboratories typically offer analyses in element packages, the
composition of which is determined by measurement practices. Because of
the expense, a need for only one or two of the elements included in the
package rarely justifies the use of an INAA method.
214 Analytical methods

A more favourable and cost-effective method for exploration geochemistry


is epithermal NAA (ENAA). Although the epithermal neutron flux is much
lower than the thermal flux, epithermal neutrons have definite advantages
stemming from the interference of main components in geological matrixes.
Relative to other, non-NAA methods as well, the ENAA allows very
satisfactory lower limits of detection for Cs (0.6 ppm), Sc (0.5 ppm), La
(1.5 ppm), Ta (0.5 ppm), Au (3 ppb), As (1 ppm), Sb (0.1 ppm), Br (0.6
ppm), Sm (0.05 ppm), Th (0.4 ppm) and U (0.3 ppm). The thermal method
offers better detection only for Sc, Hf, Ag and part of the REE group.
Unfortunately, some geochemically important elements (e.g. Cd, Cu, Pb,
and S) cannot be determined by NAA but combined use of other techniques
is required.
An advantage of the NAA method (a disadvantage for elements like Au
with particle effects) is the normally small sample quantities needed (less
than 0.5 g) and the minimum amount of preparation (powders suffice). The
latter feature is often desirable in the analysis of organic materials. In
general, after allowing some time for decay, the activated samples can be
used for other analyses.
The method of delayed neutron counting (DNC) enables accurate, very
rapid and fairly selective determination of low abundances of U (and Th
with special techniques) in exploration samples. The basis of the method is
the property of some heavy nuclides to produce delayed neutron precursors
under irradiation. These are beta decayable with short half-lives, and
the residual nuclei are neutron emitters (rather than "normal" gamma
emitters). Actually, when automated as illustrated by Rosenberg et al.
(1977), the DNC determination of U (from 0.5 ppm level) is inexpensive
and may be less costly than the fluorometric determination.
Determination of the entire groups of REE or PGE has often been
carried out by radiochemical NAA (RNAA), followed by laborious and costly
separation and enrichment procedures. The method is usually not suitably
applied for large batches of samples and may soon be replaced by ICP-MS
(REE and PGE) or GAAS (PGE). The introduction of new separation
systems might make the RNAA more practicable for geochemical work.
Prompt gamma-ray NAA (PGNAA) has been used for determination of
trace level B in samples from Au exploration.
An overall review of the general application of NAA is presented by
Bruneetal. (1984).

FUTURE NEEDS

Today's laboratory techniques allow almost any parameter to be mea-


sured either qualitatively or quantitatively. With a suitable combination of
the latest techniques, most sample types can he handled and most elements
Recommendations 215

determined with high throughput, low detection limits and reasonable


cost. Besides this, laboratory information management has been automated
from registration of samples, through scheduling of requests to reporting
of analytical data.
Some areas of laboratory operation nevertheless remain in need of
development:
(1) Sample pre treatment. This stage currently accounts for too large a
share of the total costs and also presents certain worker risks. At least part
of the pretreatment stage should be automated.
(2) Sample preparation. Selective extractions and methods capable of
differentiating the modes of occurrences of elements are needed for better
understanding of the character of anomalies and for environmental studies.
Alternative methods of sample introduction promise to lower the present
detection limits of ICP-based methods considerably.
(3) Sample logistics. Data management has often been automated but
much labour and loss of time is still spent in sample materials management.
(4) Quality assurance. Much efforts is being directed toward the creation
of systematic quality control programmes and, in general, good laboratory
practice (GLP). In particular, those laboratories involved in environmental
studies will be obliged to adhere to national and international quality
system standards, such as the European EN-45000 series.
(5) Analyst-geochemist cooperation. Sophisticated analytical techniques
require most analysts to specialize. Better interaction is needed between
those responsible for planning of geochemical programmes, for the labora-
tory and for data interpretation.
(6) Modifying of existing methods and acquiring special methods for
environmental studies. Many of the multitask geochemical mapping pro-
grammes initiated in the 1990's will be concerned with environmental
questions.

RECOMMENDATIONS

Depending on their objective, geochemical studies can be devided into


three sectors: (1) regional mapping, (2) exploration and (3) environmental
monitoring. Each sector prefers its own types of material and chemical
analysis.
For regional geochemical/geological mapping, widely distributed materi-
als that record the compositon of the bedrock are recommended. Stream
and lake sediments and the surficial layer of till have been used. The
total concentrations of elements best reflect bedrock composition. However,
partial leach methods may be more succesful in the delineation of miner-
alized formations. The analytical method applied should have a rather low
detection limit for most of the major, minor and trace elements and should
216 Analytical methods

be capable of producing reliable data over a long period. IPC-AES would be


ideal as would also OES complemented with AAS.
In local and target exploration, materials closely related to the under-
lying bedrock are analysed to find and delineate anomalies caused by ore
mineralizations and thus to be establish drilling targets. Deep till, weath-
ered bedrock, gossan and, occasionally, humus have been used as sampling
materials. Selective sequential methods that attack only the minerals of
the ores searched for or the secondary minerals formed from them in
weathering processes are recommended. ICP-AES and ICP-MS are the best
instruments although AAS with its many modifications is also appropriate.
In some cases (e.g., if electricity is not available) colorimetric methods can
also be used.
Monitoring aimed at measuring the human impact on the environment
is best done by sampling surficial materials such as humus, water and
vegetation. Partial extraction and methods with a very low detection limit
are to be preferred. INAA, ICP-AES and ICP-MS meet these requirements
but AAS, especially GFAAS, is also feasible.
217

Chapter 10

RECOGNITION OF ANOMALY PATTERNS IN REGIONAL


GEOCHEMICAL INVESTIGATIONS

INTRODUCTION

The primary objective of data interpretation in geochemical exploration


is anomaly recognition. In easy cases the presentation of data on a map
is sufficient for identification of anomalous patterns. Often, however, high
background variation camouflages the anomalies, with the result that
during the last decades much research has been focused on statistical
methods for separating anomalies from the background. Many of these
methods are based on idealized statistical models and strategies believed to
most effectively isolate the samples indicating ore mineralizations from the
background.
This chapter examines anomaly recognition methods in regional geo-
chemistry from the practical point of view and, as such, it should be
considered complementary to Volume 2 of the Handbook of Exploration
Geochemistry, Statistics and Data Analysis in Geochemical Prospecting. Most
of the presented material is based on experience with the regional map-
ping programmes at the Geological Survey of Finland. Attention is paid
to statistical evaluation of the homogeneity of materials, quality control
of chemical analyses, univariate and multivariate data description and
anomaly recognition, data presentation on maps and, finally, strategies for
selecting methods for anomaly recognition.
Although the concept of geochemical anomaly is widely discussed in
the literature, a precise and exhaustive definition of this everyday term is
hard to come by. Most geochemists familiar with statistics associate the
anomaly with the upper tail of a univariate distribution curve indicating
the highest detected concentrations of a metal. Similarly, exceptionally low
element contents can be considered anomalous. Geochemists talk about the
anomaly threshold as a critical separator between the anomaly and the less
interesting background. A variety of more or less successful methods have
been employed to yield this threshold, but none of them has obtained the
status of a standard nor been accepted as a firm rule of thumb.
The regional behaviour, or pattern, of the high concentration values lends
another dimension to the concept of anomaly, which has recently assumed
218 Recognition of anomaly patterns in regional geochemical investigations

more importance with the significant progress in automatic map drawing


and data presentation techniques. In a good map presentation, anomaly
thresholds are unnecessary if the variation in the element contents is
adequately shown at all relevant levels on the map. Then both local features
and the geochemical relief show up as regional patterns on the same map.
In detailed prospecting it may, however, be useful to distinguish certain
classes of concentration values on the map using traditional presentation
with "hard classes".
Many map drawing techniques require a preliminary manipulation of the
data such as interpolation and smoothing. This is necessary, for example,
if coloured surface maps are desired. Image processing methods provide
useful tools for the geochemical anomaly as well. In fact, edge detection,
basin detection, high- and low-pass filtering and so forth can be employed
to delineate regional anomaly patterns.
Multivariate statistical pattern recognition models are commonly used
to classify samples into anomaly and background classes when taking into
account the statistical distributions and regional paragenetic patterns of
several elements simultaneously. Some widely applied mineral resource
assessment methods such as characteristic analysis can be simplified and
utilized as geochemical recognition methods.
The present concept of a geochemical anomaly is such that no unique
general definition can be offered for it. Because the purely geochemical
information of an area of investigation is gained only from chemical
analyses of disparate samples, the "natural anomaly" is necessarily blurred
and distorted despite the use of an acceptable recognition method.
In regional geochemistry the difficulty is to resolve complex multivariate
behaviour from a limited amount of data. The quality and content of
the information carried by the collected data cannot be improved by any
method; it can only be presented more clearly to the user. These realities
lead to the conclusion that the geochemical anomaly seen by an interpreter
will be unique for every (1) objective of the investigation, (2) sampling
method, (3) analytical method, (4) anomaly recognition method, (5) data
presentation method and finally but not least, (6) interpreter. Mistakes
at any one of these stages may prevent the detection of a fine "natural
anomaly" in the area under investigation. Some mistakes yield spurious
anomalies which are difficult and tedious to discard. The amount and
quality of prior information and knowledge of the area to be investigated
may drastically influence the choice of statistical methods. The more that
is known the less that is asked for.
The choice of methods also depends on the available computing facilities:
both hardware and software. A wide range of general purpose statistical and
graphical program packages are available for the most common computers
including personal computers (PCs). Most statistical problems, even quite
complicated ones, can be solved by these programs in the standard way, but
Statistical aspects of quality control 219

it is assumed that the user has a good statistical education and does not
fail in selecting the right method. A few supervising and guiding statistical
expert system programs already exist. But still the user has to be an expert
himself.
Modern user-friendly programs have removed the barrier between the
geologist and the computer. Moderate prices of powerful PCs have made
the computer an everyday tool for the geologist. Very large data sets still
require processing on a mainframe computer, especially in production-
type work like map drawing. The question of which geochemical data
analysis can be done on PCs and which on mainframe computers must
be answered specifically for every organization and type of investigation.
The answer should be based at least on (1) the general computing policy
of the organization, (2) the available resources (qualified staff, computers,
programs, data bases, auxiliary files, etc.), (3) the size and life-time of
the data set to be created and treated, and (4) the importance the data
set may have for later use. These aspects are only briefly discussed here;
some recent comments about geochemical data processing can be found in
Garrett and Leymarie (1989).
Basic statistical concepts are not defined here but can be looked up in
the referenced books. No attempt is made to present the statistical methods
completely.

STATISTICAL ASPECTS OF QUALITY CONTROL

The interpretation of regional geochemical data is based on the infor-


mation achieved from chemical analyses of samples processed in several
stages. The quality of the resultant data depends on the quality of the
data production in these stages. Random variation is an unavoidable in-
herent component of the information relayed by geochemical samples. The
randomness in sampling is due to lack of regional representativeness of the
samples and lack of local representativeness of single sampling units. These
experimental "errors" are determined by the sampling grid and the type
and size of single samples. Any stage of sample treatment following the
sampling, such as storage, preparation, dissolution and analysis, will add
additional components of randomness to the final information. The image
of the natural patterns and distributions of elements is always more or less
distorted. Figure 10-1 shows schematically how the measured concentration
of an element may vary around the real level owing to fluctuation in sam-
pling and analysis. Here the fluctuation is caused by normally distributed
random numbers and each component accounts for an error of 10% relative
to the real level. In general, the error components cannot be totally removed
but they can be minimized according to the circumstances. The investigator
should at least be aware of the presence of error components in his data.
220 Recognition of anomaly patterns in regional geochemical investigations

(Concentration

Horizontal line

Fig. 10-1. Simulated variation of sampled element concentrations around the "natural" level
(black curve) caused by randomness of sampling (dark grey) and analysis (light grey).

Geochemical investigations are usually focused on the natural variation


or heterogeneity imbedded in measurements of field samples. The regional
heterogeneity, or the geochemical relief, is typically the object of regional
investigations. Regional investigations are targeted at large-scale patterns
and detailed studies at local patterns. The smaller the local features that
need to be revealed, the more homogeneous samples and more statistically
representative data need to be available for an area.

Sampling grid
In the planning of the sampling strategy the objectives and aims of the
investigation are considered, and the resources available for the sampling.
Any preliminary information giving a clue to the heterogeneity of the area
should be utilized in designing the density and interval of the sampling
points. If the grid, or sampling scheme, is too sparse then few points
will hit anomalies of the desired size. On the other hand, if the grid
is overly dense money will be wasted on the collection of superfluous
information. A sampling grid can be optimal only in relation to the aims of
the investigation and the available auxiliary information. Universal rules
for grid design therefore cannot be stated.
Unexpected technical restraints may be so limiting that a planned
optimal grid cannot be used. For example, if till sampling is performed and
significant areas are without till then the regional optimization is weak.
A grid for stream sediments must be designed with respect to the pattern
of streams, and the data are not then areally equally representative. In
the case of stream sediment samples collected in the lower courses of
Statistical aspects of quality control 221

drainage basins (Bölviken et al., 1986), samples correspond to areas of


variable size and thus the regional representativeness varies from sample
to sample. Ordinary data of stream sediments can nevertheless be worked
up to describe average element concentrations in drainage basins instead of
single samples. This improves the regional representativeness of the data
and justifies a regional map presentation.
If the objective of the investigation is to study the regional geochemical
relief with no prerequirements on the shape of patterns, then a pilot study
with a sparse grid is recommended to achieve preliminary information
about the geochemical variation in the area. This variation can be inferred
from univariate maps of the pilot data and a denser grid designed on
the basis of this pilot information. Quality control of the pilot data is
recommended to gain valuable information at an early stage. The regional
variation and statistical behaviour are specific to each variable and the
variables may behave quite differently. One may show heterogeneity and
the other homogeneity over the same area. Thus, the final grid must be a
compromise designed on the basis of those target variables considered most
important.
Sampling strategies for purely exploration purposes over a relatively
unknown region will be more economical than overall dense sampling, if
they involve several successive sampling stages with increasing sampling
density in critically selected subareas. This stagewise strategy will be
successful if the anomalies are located in zones or if large patterns are
detected in the initial sampling stages. If neither criterion is satisfied then
the stagewise strategy is risky, because significant anomalies may be lost in
the initial stage and not revealed in a later stage.
If the investigation aims at the detection of local anomalies the grid
design can be tailored to be effective for anomalies of expected or desired
size and shape. Assuming elliptical anomalies (in overburden) with given
smallest size and approximate direction an optimal grid design can be
chosen from the tables given by Savinskii (1965). The regular rectangular
grid design offered in these tables is optimal, however, only if the premises
really are satisfied and the anomalies are elliptic all over the area of
investigation. If several target variables are involved then a compromise
must be made concerning the size and shape of the desired anomalies.
If the preliminary information is weak and nothing can be assumed
about the direction or shape of the anomalies then the grid pattern should
be random or systematic, without directional emphasis. In both cases
the samples are evenly distributed over the area, and the designs are
statistically almost equivalent. The sampling design can affect the results
of regional interpolation or trend surface fitting. One design may be more
appropriate than another for a specific interpolation method. The Kriging-
interpolation method, for example, yields more reliable estimates with an
equilateral triangle design, where the area is divided into equal equilateral
222 Recognition of anomaly patterns in regional geochemical investigations

triangles and the grid points are located at corners of the triangles, than
with a hexagonal or quadratic design (Yfantis et al., 1987).
Variation of sampling density in regional investigations should be avoided
whatever interpretation method is applied. Otherwise the interpretation
and final conclusions will not necessarily hold for the entire area.
The effect of local heterogeneity, or sampling error, can be reduced by
clustered sampling in which several samples are taken near or at the
same site and composited to a single sample. Composite samples reflect the
local average level of element contents better than single samples and are
consequently locally more representative. This procedure is recommended
particularly for regional investigations, where large scale anomaly patterns
shown on a sparse grid are of interest. Compared with simple sampling,
compositing reduces the need for analyses without drastic increase in
sampling costs or significant loss in information. Compositing has been
studied theoretically and with simulation models by Garrett and Sinding-
Larsen (1983).

Size of the sampling unit


Besides the sampling grid design, the size of the single sample is critical for
the representativeness of the data set. The significance of the sample size
varies with the structure and homogeneity of the material to be sampled. A
relatively small sample will be representative if the grain size is small and
the target material, i.e., those grains or minerals containing the elements
of interest, are evenly distributed. Examples of this type of material are the
fine fraction of till and water. However, if the grains are coarse or the target
material is unevenly distributed, good representativeness will require a
relatively large sampling unit. Such a situation is met in sampling for gold
in till or in lithogeochemical sampling from strongly stratified bedrock.
Representative sampling is especially difficult in lithogeochemistry where
the results are strongly influenced by both the sampling design and sample
size. Random lithogeochemical sampling through overburden may hit any
of several rock types on the bedrock surface, some of them hard and raised
above the general rock surface and others soft and deeply worn. Thus, many
of the sampling units will represent rock types not typical of the bedrock
being investigated. Fatal mistakes can be made in the interpretation of
lithogeochemical data if the sampling conditions are not considered.
Problems in choosing an optimal size of sampling unit, given certain
features of the material (size and shape of grains), have been treated by Gy
(1982) and Minkkinen (1987). Minkkinen has even worked out a computer
program for selecting optimal samples of particulate materials considering,
among other things, grain size and shape, Minkkinen (1989).
Representative sampling does not guarantee representative results be-
cause subsequent stages in the analytical process may lead to "new
Statistical aspects of quality control 223

sampling" or selection of the material to be fed to the instrument. Although


homogeneity of samples is usually preserved in laboratory routines, some-
times a particular stage of the routine, e.g., transportation or the analytical
method itself, may be selective, especially when particulate samples are not
dissolved and when only a fraction of the original sample is analyzed at a
time. Any estimation of the representativeness and quality of results must
pay attention to all phases of treatment from sampling to analysis. The
phases must also be known by those who use the data, which requires that
a rigorous documentation be done for each data set.
It bears emphasis that weak representativeness and quality of samples
cannot be improved by good and expensive methods of analysis nor
statistical methods. Poor samples provide poor information yielding poor
geochemical inference.

Effects of information processing


The quality of geochemical data is also affected by factors inherent to
computer processing factors such as precision and presentation of data fed
out from the instrument, round-off errors, precision and presentation at
storage and, finally, the data presentation within the application program
employed for interpretation. These factors are relevant to both manual
and automatic data registrations and tend to cause problems especially in
those extraordinary situations when data are of poor quality. Presenting
data precise only to within ± 5 ppm is not necessarily catastrophic for the
data quality, but the geochemist has to know about the round off, because
such data are not continuous but discrete, and this may result in spurious
statistics. This discretization error may become more severe if the output
results are further rounded off and increased by a dilution coefficient.
It is therefore recommended to avoid round off procedures, despite the
known instrumental error, and to store the data in as unchanged form as
possible.
Data storage devices are nowadays so low cost that the data can be stored
as exactly as outputted from the instrument without the need to reduce
the information for cost reasons. The data should be checked for errors
and corrected before storage into a data base or file. It is worthwhile to
spend resources, especially computer resources, on diagnostics at the input
stage because errors that have reached the data base are difficult to identify
and are usually revealed later only by chance, if at all. In the worst case
erroneous results in the data base will be widely used, just like the correct
ones. Errors are not necessarily detected in statistical resumes and may
well distort them.
Good laboratory practice and a good geochemical data base system provide
the user with information about data quality and conditions affecting
the quality of any data set. The user should consult this information
224 Recognition of anomaly patterns in regional geochemical investigations

before attempting an interpretation. Single measurement values can be


equipped with flags indicating poor quality if they lie below the detection
limit or outside the calibration range. General method-dependent quality
information such as accuracy and precision should be included in the above-
mentioned documentation (see Chapter 9). The date of analysis should be
stored for each sample and each event of the analysis. This is particularly
important when measurements contain drift or bias varying with time
and require correction. The date of analysis is also an important label
on measurements to identify results analyzed before and after substantial
changes in the analytical procedure during the life-time of the data base.
Such changes are unavoidable because analytical facilities must now and
then be modernized.
The quality of the information processing becomes extremely important
if the data base is to be used for diverse purposes and by people of varying
expertise. The problem of opening geochemical data bases to research
workers of different disciplines resides here. How can we guarantee that
a research worker not acquainted with geochemistry or exactly that type
of data, or a journalist, will use the data properly? At least the quality
information has to be available, especially if several data sets with diverse
methodological backgrounds are being compared or integrated. These
problems need to be solved for any data base utilized by several users or
organizations or any public geochemical data base. One solution is to draw
up a detailed manual or guide for the data base, containing full information
on data quality and restrictions. With this in hand, the user alone will be
responsible for his/her inferences.

Measures of quality
As mentioned earlier, many factors at different stages of sample treat-
ment from sampling to analytical procedures and data processing, may
distort the quality of the final data. Not all can be measured or controlled,
because some are unknown or not observable. Average effects can be es-
timated by statistical methods, but then additional information must be
gathered at each stage. Single sample analyses are not sufficient for quality
control and estimation of variation components. The quality of measure-
ments is characterized by statistical parameters indicating reproducibility
or stability and these parameters are usually computed from replicates
obtained from the various stages of treatment. Some characteristic pa-
rameters are specific for a method or an instrument and do not need to
be computed separately for each batch. The detection limit of an atomic
absorption spectrophotometer, for instance, is the same for a particular
element irrespective of what samples are analyzed. But the detection limit
of measurements done by neutron activation analysis varies from sample
to sample and is usually separately estimated for each sample.
Statistical aspects of quality control 225

Laboratories usually report the reproducibility and detection limit of


a particular analytical method (see Chapter 9), but these measure only
the quality at the stage of analysis (possibly for selected samples and
under ideal conditions), not the quality for all samples and all stages of
the treatment. Procedures for analytical quality control can be found in
Thompson (1983). Any information that the laboratory offers about quality
is useful and should be utilized to avoid embarrassing surprises after the
report is published. The selection of an analytical method or methods for
an investigation should be based on the existing information on the quality.
That should guarantee sufficiently good analytical data for the objectives of
the investigation.
The components of variation, or errors, not only depend on the methods
applied at different stages but may be specific to sets of samples or batches.
One set of samples may, for example, show a moderate local variation
while another set shows significant local heterogeneity. This means that
an analytical method may be sufficiently good for the heterogeneous batch
but not for the other. One way of attacking this problem is to make a pilot
study, including quality control, over the investigation area and then adjust
the existing measures of quality to correspond to the current measures.
Results from the pilot sampling can be used as additional information when
estimating the quality of the final data set.
Useful data for quality control is obtained through replicate sampling
and through replicate analysis of samples randomly scattered over the
area and over the duration of the analytical procedure. One could then
expect the measures of quality to be specific to current methods (sampling,
preparation and analysis), area and time. To prepare replicates for all
stages of treatment where error components are expected, would be
expensive and sometimes even impossible. Thus replicates are usually
inserted only in the sampling and analytical stages, resulting in just two
estimated error components. The information obtainable from replicates
can be fully exploited only if the sampling design is hierarchic. In the
two-stage model this means that the replicate analyses are done on samples
replicated in the field. If the replicate design is not hierarchic then the error
components cannot be estimated (see below Analysis of variance). Figure
10-2 shows a general hierarchic symmetrical design with one replicate
(two branches) at each stage. Such a design is usually called balanced.
Increasing the number of replicates improves the statistical stability and
confidence of the estimates, but in practice and for cost reasons usually
only one or two replicates are included at each stage. The cheapest and
still statistically satisfactory design for replicates is the twofold unbalanced
design shown in Fig. 10-3. To yield satisfactory data, any design must
be statistically representative over the investigation area and analytical
time interval. The regional representativeness is enhanced if the replicated
samples are systematically or randomly spread over the investigation area.
226 Recognition of anomaly patterns in regional geochemical investigations

SAMPLE

PHASE 1

PHASE 2

Fig. 10-2. Hierarchic replicates at two stages (a balanced nested design).

ANALYSIS Ö

Fig. 10-3. Unbalanced twofold nested design of replicates with sufficient information and
niinimal costs for monitoring variation at sampling and analysis.

The replicates also should be analyzed in random order within the batch.
Otherwise trends and systematic variation in analyses may cause too
good estimates of reproducibility! If, for example, replicates of a sample
are analyzed one after the other and the instrument tends to produce
exceptionally high values that day, then the replicates capture only the
random not the total variation.
To avoid spurious dependence between measurements caused by system-
atic variation in analysis, it is recommended that even the unreplicated
samples be analyzed in random order. Randomization spreads the sys-
Statistical aspects of quality control 227

tematic error throughout the investigation area, and thereby reduces the
probability of getting false regional anomalies raised by that error type.
The randomization is tedious because the sampling units must be reordered
physically. The randomized analytical order of samples within the batch
can of course be generated by computer, but the manual work is difficult to
eliminate.
It is almost impossible to randomize and analyze in one batch samples
collected in large projects spanning several years. Usually the samples are
grouped into smaller batches, which can then be randomized internally,
but systematic errors between batches will remain. Such between-batch
discordance is sometimes appalling on maps that cover large data sets.
Such a map would scarcely gain the confidence of the critical user. One way
of avoiding between-batch discordance is to make the map presentation
sufficiently coarse to hide the systematic variation along batch borders.
Unfortunately this method also deletes low-contrast features within batches
that might just be correct.
Systematic variation is nowadays reduced, at least to some extent, by
frequent calibration and monitoring in the laboratory. If the laboratory
from time to time can provide unbiased data then the randomization
process is less important and can be ignored. Despite good laboratory
practice and modern techniques, however, some kind of bias may slip in due
to minor changes in the analytical procedure (new electronic components,
new calibration programs, adjustments of calibration curves, calibration
samples replaced by new ones, etc.). The distortion of new data relative to
old can be very difficult to estimate. So, since the analytical bias hardly
can be completely avoided, it is important to inform the geochemical
investigator how serious the bias is and what means are available to
reduce it.
The number of replicates at the various stages can be optimized within
the economic restraints, but in practice the number of replicate sites in
the field will be roughly proportional to the size of the investigation area
and the number of sampling sites. A relatively dense grid easily offers
representative sites for replicates. But if the area is large and the grid
sparse, the proportion of replicates should be high. Clearly, too, the number
of replicate sites will depend on the geochemical complexity of the area
(much variation requires dense sampling and a dense grid of replicates) and
on the goals of the investigation (is the interest details or large zones?). The
more details one desires, the more samples and replicates must be taken to
catch them and evaluate their significance. Universal rules of thumb are
difficult to recommend.
The regional geochemical till sampling programme at the Geological
Survey of Finland is being carried out with 1 sample/4 km 2 . The whole
country (338,000 km 2 ) was covered within a few years. The samples
were treated in batches of 4,800 samples each (19,200 km 2 ). Replicates
228 Recognition of anomaly patterns in regional geochemical investigations

were taken at every 30th sampling site, yielding about 160 replicates for
each batch. The cheapest twofold unbalanced design (Fig. 10-3) has given
satisfactory results for monitoring of quality. This scheme reduces the "net
analytical capacity" by only 7% compared with no replicates at all (and no
information about quality).
Replicates require considerable practical organization (e.g., marking
samples) and systematic treatment, but the result will be invaluable
information about the data set.
The data gained through replicates can be viewed in scatter diagrams
and the error components can be tested by significance tests by analysis
of variance. From now on, only the simplest twofold unbalanced replicate
design is considered (Fig. 10-3).

Graphical tools for quality assessment


A useful tool for viewing replicate data (2 values per sample) is the
scatter diagram or point diagram where the two measurements of each
sample are plotted against each other, showing how well the replicates
correlate. The axes represent the two concentration values of an element
and each dot on the plot corresponds to a single sample. Data from regional
geochemical till sampling are considered in the following example. Again
only two stages are monitored: the sampling and analysis. This means that
the analytical variation contains components of all stages after sampling
(e.g., sample preparation). If the replication of analysis starts with the
same solution, for an atomic absorption determination, for example, the
variation describes only the internal heterogeneity of the liquid (nil) and
the analytical event.
The scatter diagram is a tool for eyeball estimation. If the dots fall closely
along the diagonal then the reproducibility is good; otherwise it is bad.
Garrett (1969) has presented a rule of thumb: the data is good if the length
of the dot cloud (along the diagonal) is more than four times the width.
It is important to note that the scale has an important role in shaping
the dot clouds. Of course, scales and axes must be identical for two
axes in the same diagram. But the same is true when plots of the same
element in different data sets are to be compared (e.g., different areas).
Scaling functions can stretch or squeeze the dot cloud to emphasize
desired ranges of measurements. This is an instance where one can lie
with statistics, intentionally or unintentionally, by scaling. Figure 10-4
shows the scattering of measurements for replicate samples. That variation
contains all variation components between and including sampling and
analysis. Figure 10-5 shows the dispersion of replicate analyses of the same
data set. Comparison of Figs. 10-4 and 10-5 shows that the dot cloud
is clearly narrower for analytical replicates than for sampling replicates.
This indicates that the analysis is good relative to the field variation at
Statistical aspects of quality control 229

Replicate samples» Cu

1000

Fig. 10-4. Scatter diagram of replicate samples (3.4% of samples duplicate). Concentrations in
ppm.

sampling. Here the logarithmic scale is selected to avoid problems with


possible large values. The logarithmic scale is liable to exaggerate variation
at low concentrations, however — a fact that needs to be kept in mind
when making eyeball estimations of dot cloud shapes.
Another graphical tool for viewing data quality is a diagram showing the
relative variation estimated from replicates in relation to the concentration
level. In such a diagram the absolute difference between replicate values,
|*i — *2|/(*i + *2)> is presented on the vertical axes and their average,
(xi + #2)/2> on the horizontal axes (xi and x^ stand for the first and the
replicated value, respectively). These diagrams tend to show high relative
error at low levels (near or below the detection limit) and at extremely high
levels.
230 Recognition of anomaly patterns in regional geochemical investigations

Rep 1 i c a U RNRLYSES, Cu

#
·*
1
··
i*
:
ß
iff- ·
x

β

""

j . 1
10 100 1000

Fig. 10-5. Scatter diagram of replicate analyses (each duplicate sample re-analyzed). Concentra-
tions in ppm.

Graphical tools are practical because no statistical theory with uncom-


fortable assumptions is needed. Graphical presentations are insensitive to
outliers (if scaling is properly handled) and do not depend on any statistical
distribution laws. Graphical tools are used to show, simultaneously, many
kinds of patterns and behaviour in the data, which are difficult to detect
by parametric estimation alone. Such patterns in the scatter diagram are
dot clusters, which may be a symptom of some analytical or sampling phe-
nomenon. Sudden jumps in the relative error plot may indicate problems in
the calibration curve. But graphical tools do not allow direct quantitative
estimation of the error components. In graphical displays, error compo-
nents can only be evaluated in relation to each other by eyeball. Exhaustive
Statistical aspects of quality control 231

utilization of diagrams of the types discussed demands some experience on


the part of the user.

Analysis of variance
The analysis of variance is a statistical method often employed for
statistical quality control and, in particular, as a test of the significance
of the error components. Traditionally, various forms of the analysis of
variance have been used in manufacturing quality control, to test the
variation in product quality over time. The analysis of variance is usually
based on replicate sampling or measurements arranged according to a
scheme suitable for the problem.
The simplest analysis of variance scheme or design is the one-way design,
where all samples (measurements) are grouped into disjoint classes and
each class contains one or more measurements. The one-way analysis of
variance consists of testing the significance of differences between the
classes relative to the variation of repeated measurements within the
classes.
In all its forms, the analysis of variance is based on the assumption
that the distributions of measurements within classes are normal or
transformed to normal and, rigorously, their variances should be equal.
The significance test measures the spread of the means of the classes in
relation to the average width of within class distributions. Closely located
means imply equal distributions, but even one clearly deviating mean
implies unequal distributions. It is very important to note that narrow
distributions may imply significantly different means, but the same means
can be insignificantly different for wider distributions. In short, separation
of small differences requires narrow distributions!
The expression for the one-way analysis of variance model (with fixed
effects between classes) is:
xy = μ + at + €ij, i = 1, . . . , /i; j = 1, nt (1)
where xy is measurement number j of class number i, μ the expectation
(population mean) of all classes, oti the deviation of the expectation of class
number i from μ, and ey the deviation of measurement j in class i from
the within class expectation /i + ftj, ni is the number of measurements
within class i. Here μι are assumed fixed (fixed effects model) and ey are
assumed to be normally distributed. The F-test is used to find out whether
the variation of the α^ is significant in relation to the variation of the
€ij. This simple one-way analysis of variance is useful in situations where
the difference between clearly defined classes (analytical methods, devices,
sample types, etc.) is investigated and where replicate measurements are
available. This model is also applicable to homogeneity tests of large bulk
samples, which are divided into subsamples, which in turn are repeatedly
232 Recognition of anomaly patterns in regional geochemical investigations

analyzed. Then the significance of the differences between subsamples is


tested.
The twofold unbalanced nested analysis of variance design for replicate
sampling and analysis is formulated as (Graybill, 1961):
*yk = M + αί + ßij + €jjk> i = 1, . . . , n; j = 1, n f ; Ä = 1, ny (2)
where xy^ is the &th replicate of the jth replicated sample at the ith
sampling site, μ stands for the total expectation of the sampled population,
Oi{ is the deviation between the expectation at the ith sampling site and μ,
ßij is the deviation between the j t h replicate sample and μ + aj, and ey^ is
the deviation between the ßth replicate analysis and μ + α^ + ßy. In this
model a\ and ßy are considered random and normally distributed additive
components with expectation 0 and specific variances. Thus the variance of
OLi can be interpreted as the variance of the geochemical relief (between-site
variance) and the variance of ßj is the variance of the local sampling.
Finally, the variance of ey^ can be interpreted as the analytical (including
pretreatment) variance. The statistical significance of these components
can be tested by the F-test assuming that normal distributions rule.
Formula (2) is incomplete because the interaction between sampling and
analysis is omitted. Such interaction may exist when the error components
correlate, and that may happen if the samples are not randomized and
analytical-time-dependent bias exists. If the interaction component is in-
cluded in the model then the significance tests become more complicated.
If the interaction component can be omitted (or tested insignificant) then
the variance components can be quantitatively estimated. This means, for
example, that the standard deviations of sampling and analytical errors
can be expressed in ppm. Such information is naturally very useful in
anomaly interpretation. The unbalanced hierarchic analysis of variance is
rarely presented in the literature but basic concepts and formulas can be
found in Graybill (1961). Applications of these methods for quality control
in geochemistry are described by Garrett (1979).
The normal distribution plays a central role in classical analysis of
variance procedures. If the normal law cannot be accepted, then a data
transformation yielding normal variates should be employed. Frequently
used transformations are the logarithmic transformation and the Box-Cox
power transformation (Box and Cox, 1964). If the distribution does not
behave normally after transformations then a nonparametric analysis of
variance can be applied. The advantage of nonparametric methods is their
independence of distribution laws. These usually operate on ordered or
classified data containing less information than chemical analyses typically
do. Nonparametric methods are surprisingly rarely used in geochemistry,
possibly because the lognormal law is so widely accepted and because non-
parametric methods were not included in the early commercial statistical
program packages. Recent versions of most program packages now contain
The univariate approach to anomaly recognition 233

both classical analysis of variance methods and their nonparametric coun-


terparts. Nonparametric methods are described in Maritz (1981). A useful
and handy collection of statistical methods, including analysis of variance
is found in Pollard (1977).
When interpreting results from analysis of variance one should bear in
mind that the results reflect the overall features of the whole replicated
data. Consequently, the results may not be representative for subareas of
the investigation area. Assume, for example, that 80% of the investigation
area is highly homogeneous with little regional or local variation in, say, Cu,
while 20% of the area is highly heterogeneous locally (but not regionally)
in Cu. In this case the result of the analysis of variance may show that the
local variation is high relative to the geochemical relief and the analytical
variation. Extreme variation in subareas may accordingly distort the overall
result. The replicated sampling sites must be regionally evenly distributed
because the interdistances between sites are not taken into account in the
analysis of variance method or in scatter diagrams. Thus, the analysis of
variance of replicate data can be used to evaluate the average quality of a
data set, but errors in measurement of single sampling units or subareas
cannot be detected by these methods.
Graphical presentations, together with the analysis of variance, are
powerful tools for quality assessment and detection of the various error
types measured by replicates. Correction of these errors is another matter.
If the detected error is such that the concentration level is systematically
wrong (bias) and the affected measurements can be identified, then some
kind of correction can be made: for example, multiply the erroneous level
with a proper coefficient or add a constant. Adjustments of this type involve
risks, because the true behaviour of the error is seldom exactly known. If
the bias changes with the time of analysis then multiplication by a single
coefficient may distort the data even worse. If the quality of data is low
for critical elements in identified samples, a new analysis with improved
monitoring is possibly the safest way to proceed. One should be sure that
the new measurements are unbiased, however; otherwise the reanalysis is
meaningless.
If the low quality of measurements cannot be computationally improved
nor the problematic samples reanalyzed, the only course is to take the
information alerting to poor data into account when making maps and
drawing statistical and geochemical conclusions from the data.

THE UNIVARIATE APPROACH TO ANOMALY RECOGNITION

Geochemical data usually consist of measurements of element concentra-


tions in regionally distributed samples. The variation in the measurements
depends on the natural variation in the sampled material, statistical fea-
234 Recognition of anomaly patterns in regional geochemical investigations

tures of the sampling grid, variation in sampling (depth, habits of the


person doing sampling, etc.), sample preparation and analysis. The usual
objective of a geochemical regional investigation is to pick up the natural
variation from noisy data and study its geochemical character. Any other
variation is disturbing and should be identified and removed. The natural
variation is generally considered to be composed of two main components:
(1) the background (not anomalous), and (2) the anomalous part of the data.
Usually the investigation aims to separate these components and locate the
anomalies.
When the goal of the geochemical investigation is to find single element
anomalies in the area, univariate procedures are employed to inspect
the data. Two approaches need to be distinguished: (1) inspection of the
statistical distribution of the concentration values, and (2) inspection of the
regional behaviour and patterns of the concentration values.
Inspection of the statistical distribution starts from the assumption
that the prevailing concentration values of an element in nature follow
a distribution law different from the law that governs anomalous values.
Thus, inspection of the statistical distribution can reveal the existence of
anomalies and even offer a rough estimate of their level (deviation from
background) and significance.
The regional distribution of element concentrations is commonly dis-
played in map form, in which locally and regionally significant patterns
stand out. The concentration levels in these patterns may be low in com-
parison with the overall levels, but anomalous relative to the local level.
Such local anomalies are not necessarily revealed through a statistical
distribution alone.
Because modern mapping techniques have provided the geochemist with
visually effective maps, little interest has been focused on the anomalies
revealed by statistical distributions. There is strong demand for geochemical
maps that distinctly show relevant patterns and that are easy to compare
with other maps, like geophysical or geological maps.
Anomaly detection from statistical distributions is based on assumptions
about distribution laws that are difficult to verify. Anomalous data identified
in the distribution do not necessarily correlate with regional anomaly
patterns and are not relevant to the ore prospector. Situations may occur
in which values only slightly higher than the background on the map
are much more important than their statistical significance suggests.
A regional anomaly does not necessarily cause an anomalous statistical
distribution and a statistical anomaly does not necessarily show the regional
characteristics of an anomaly. Thus, these two anomaly concepts are not
equivalent.
The statistical distribution is nevertheless useful for obtaining an
overview of the whole data set and for detecting features not revealed
in a map presentation such as:
The univariate approach to anomaly recognition 235

— sudden jumps in measurements caused by analytical errors or by


inherent but unknown effects (high or low frequencies at discrete values
due, for example, to dilution effects),
— the effect of the detection limit (the shape of the frequency distribution
function may be distorted below the detection limit at the lower tail),
— effects dependent upon data processing decisions or rules (data
presentation at registration and storage which involves round off).
These are in fact features that should be considered before constructing
the map. The statistical distribution should always be calculated at an
early stage of the investigation, with the distribution function described
graphically and with characteristic parameters. Descriptions of univariate
distributions are also needed when multivariate anomalies are searched for,
because irrational behaviour of the data may be element specific and not
detected by ordinary multivariate procedures.
The next two sections deal with methods aimed at the detection of
a single element anomaly through statistical distributions and regional
patterns, most vital for a prospector.

Univariate statistical anomalies


The recognition of geochemical univariate anomalies was earlier and
still is often based on the statistical distribution only. The background
and anomalous values are assumed to originate from samples of two
different populations. Presenting the total distribution of the observed
values graphically, either as a histogram or a cumulative distribution curve,
allows the difference between the two subpopulations to be revealed. Figure
10-6 shows a histogram of an ideal data set generated from two normal
populations. Statistically the anomaly detection in this case is equivalent
to decomposing two (or more) mixed distributions into subdistributions.
The task is not difficult if one can assume that the subpopulations follow
well-known distribution laws. Hosmer (1973) has presented an algorithm
and Agha and Ibrahim (1984) have developed a computer program to
resolve the mixture for some common distribution laws. The decomposition
problem can be expressed as
k k
a 1
g(x) = j2 jfj^ E r (3)
1 1
where g(x) is the frequency function at x of the composed distribution,
fj is the j t h population distribution function, aj is the weight of the j t h
subpopulation, and k is the number of subpopulations in the mixture.
Solving the problem means to estimate the weights aj and the parameters
(expectation and variance) of the frequency functions fj. Large data sets
require considerable computing, which may be why the estimation in the
past was often done visually from graphs.
236 Recognition of anomaly patterns in regional geochemical investigations

Q 2 8 14 20 26 32 3B 44 50 56 62 6B 74 80 86 92 9B 104 110 116


Concentration
Fig. 10-6. Simulated bimodal histogram for normal background and anomaly.

Perhaps the best known graphical tool for presenting the frequency
distribution of measured values is the histogram. The histogram shows the
frequency distribution as a sequence of bars, or frequency classes, where
the height corresponds to the frequency and the width to the class width.
In such a figure the background and anomaly ideally appear as distinct
modes or "hills" or as excessive positive skewness (stretched upper tail).
Again, the reasoning must be based on certain premisses:
(1) If the lognormal (or normal) distribution law is satisfied then modes
or skewness should be detected for distinct anomalies.
(2) If the distribution law is not satisfied then very little can be said, at
least about the meaning of skewness.
Clearly a histogram presentation is useful for viewing the data, but one
should not forget that the class width, number of classes and scale of
presentation are subjectively chosen. Different class widths and numbers
of classes would result in far different histograms of the same data. It is
also sometimes hard to judge whether one bar is significantly higher than
another and an anomaly threshold, visually estimated between the two hills
of background and anomaly, may therefore be very uncertain.
The cumulative frequency curve (probability plot) provides a more objec-
tive presentation of the distribution of measured values than the histogram,
because it shows the distribution unclassified and with the precision of a
single value. The curve can also be drawn for classified data, but the unclas-
sified more objective version should be preferred. Often the curve is drawn
The univariate approach to anomaly recognition 237

C u m u l a t i v e f r e q u e n c y for Zn (511P)

i «**\ 99.7
.:::::
; ip"" O
97 C
95 g
90 £
po

::? ao a
70 t
60 H,
40 £
30 C
: : :· ■ CD
20 tJ
.::;="" o
10^
4 3
.:!="' 2
S 0.7
0.4
0.2
_J_ 0.1
Ca <l CD <£>>-*
o
o o o og o
ppm

Fig. 10-7. Multimodal cumulative frequency curve in a probability plot.

on probability paper where the frequency scale is Gaussian. Gaussian scal-


ing transforms the distribution curve of (1) a normal distribution on an
arithmetic scale, and (2) a lognormal distribution on a logarithmic scale
to a straight line. This presentation can be used to graphically estimate
percentiles, such as medians and other quartiles (Fig. 10-7).
Assuming that the background and anomaly are normally distributed,
the curve will appear as two line segments and the anomaly threshold
may be considered to lie at their intersection. Attempts have been made
visually to decompose the cumulative frequency curve into separate curves
for the subpopulations using formula (3) (Bölviken, 1971). The curves of the
subpopulations are conjectured assuming normality (or lognormality), and
the weights in formula (3) are experimentally estimated by guessing good
initial values. This procedure involves a sequence of guesses, however, and
fails until a fairly good curve has been computed back from the estimated
subdistributions.
Sinclair (1983) has presented a detailed discussion of univariate dis-
tributions, histograms and probability plots. Stanley and Sinclair (1989)
have recently discussed the detection of anomaly thresholds from different
assumptions, and Miesch (1981) has considered the problem of estimating
geochemical anomaly thresholds.
The lognormal distribution law is often favoured for geochemical data;
Ahrens (1953) was probably the one who introduced the idea. The statistical
characteristics of the lognormal distribution have been more deeply studied
238 Recognition of anomaly patterns in regional geochemical investigations

by Rodionov (1971) and Crow and Shimizu (1988). But arguments for the
normal law have been presented as well (cf. Gubac's, 1986, discussion of the
physical grounds). The fans of lognormality draw their graphs on log-scales
and the doubters use the arithmetic scale. A compromise is to draw a
dual scale plot showing both curves on the same frequency scale and then
select the one behaving more linearly (or piecewise linearly) for further
inspection.
Much attention has been paid to the graphical presentation of the
cumulative frequency distribution curve, especially on probability paper.
Lepeltier (1969) has asked whether the values should be cumulative from
the smallest towards the highest or the other way around. This is a
relevant question because the highest value cannot directly be presented
on the Gaussian scale (100% is at infinity). If the greatest value must be
presented then one solution is to cumulate backward from the greatest to
the lowest value and miss the lowest one on the plot. Another alternative is
slightly to adjust the cumulative frequency of each value by the expression:
F(xi) = (i- 0.375)/(n + 0.25) (4)
where xi is the ith. observed value in ascending order, F{%i) is the cumulative
frequency function at x^ and n is the number of observations (Everitt,
1978). In fact, in some sense, this formula (4) yields optimal estimates of
the frequencies for graphical estimation.
In practice the greatest value is not necessarily the most important one
if the data set is large and regional anomalies are of interest. One should
also bear in mind that the quality of extreme values usually is low because
they tend to be far away from the range where the measurements are tuned
to be good both absolutely and relatively. Again, when using cumulative
graphs and drawing conclusions about data behaviour, the detection limit
should be considered. Data below the detection limit do not necessarily
follow the distribution law of good data and thus cause distortion of the
curve. Special care is required when the detection limit is relatively high
and a large portion of the values fall below it. Then the low values falsely
may be interpreted as the background and the background as the anomaly.
Cumulative frequency curves often show gradually changing slopes when
passing the detection limit, which indicates that the change of quality is not
abrupt and illustrates the nature of the subjectively determined detection
limit (Fig. 10-8). Low values are stretched on the logarithmic scale, which
may also cause odd behaviour of the cumulative frequency curve as a whole.
The most serious problem in the interpretation of cumulative frequency
curves is the inherent effect of the Gaussian scale. If neither normality
nor lognormality is satisfied then the curves are likely to behave in an
unpredictable manner. Different subpopulations may follow different laws
and sometimes the subpopulations are so many and overlapping that no
distinct thresholds can be found. Regional geochemical data is typically
The univariate approach to anomaly recognition 239

Cumulative frequency for Zr (511P)

_99.7
;:: -w
o
A·' _97 £
-95 3
£ „90 E
.80 £
-70 Jj
-60 „,
_40J?
.30 C
_soS
O
.10^
A. Ϊ*.

:::::■'
. . . : : =: : : ■

Jt
o oo o o
en ^
ppm

Fig. 10-8. Cumulative frequency distribution curve when most measurements are below the
detection limit ( « 2 0 ppm).

stratified over different backgrounds, each containing different anoma-


lies, and visual decomposition of the wrinkled curve may be impossible
(Fig. 10-7).
Another serious problem with cumulative frequency curves is the statis-
tical representativeness of the background and anomaly. If the data set is
large and the anomalous samples are very few, the anomalies may cause
such a small change in the curve that they may go unnoticed (10 anomalous
samples among 10,000 is equal to 1 per mil in frequency). Conversely, if
anomalous areas are heavily represented in the data set, the background
may be lost and not detected in the curve. This is particularly liable to
happen in detailed investigations. Strong stratification of the background,
in turn, will tend to group the data and may hide relatively weak (but not
irrelevant) anomalies within strata. This is especially a danger in large data
sets over large areas.
Anomalous values occurring in the data can be viewed as statistical
curiosities or outliers. Outliers in statistical data have been extensively
treated by many workers and a very comprehensive presentation of methods
for interpreting them can be found in a book by Barnett and Lewis (1978).
Most of these methods involve testing the significance of single extreme
values relative to the other values in the data set. Again assumptions about
distribution laws must be made. One should also know how many samples
belong to the anomaly population; for when one value is tested then
one needs to know if the next largest value belongs to the (background)
240 Recognition of anomaly patterns in regional geochemical investigations

population or not. The outliers can be grouped and the significance of the
group can be tested but this must be done under the same assumptions.
It is important that outliers be identified, and they should be treated with
care in later stages of interpretation. If one wants to wash off the effect
of "dirty" outliers in the statistical parameters describing the distribution,
then one of several adjustments can be applied. Such adjustments make the
parameters more stable and less sensitive to errors due to extreme values.
One method is trimming where an equal number of values is simply omitted
from both tails of the distribution. A more gentle method is symmetrical
winsorizing where the extremes are replaced by upper and lower threshold
values. In this case extremes are not dropped but replaced with less extreme
values. The formula for the winsorized average is

Z r w r = (rx(r+i) + * (r+ D + · · · + X(n-r) + rx{n_r))/n (5)

where X™r is the winsorized average, r is the number of data items to be


replaced in the upper and lower tails by respective thresholds and n is the
number of observations.
Trimming and winsorizing offer robust estimates of statistical parameters.
Robust methods yield stable estimates despite slight deviations of some
values from the assumed distribution law (Zhou, 1987). Some robust
methods reduce the influence of extreme values through a weight of
all values. Robust methods have been widely proposed for many common
statistical procedures which are optimized for the underlying (often normal)
distribution and that are excessively sensitive to minor deviations from the
assumptions. A successful application of robust methods to multivariate
geochemical exploration data is recently reported by Chork (1990). A
comprehensive overview of robust methods can be found in Huber (1981). A
program package for various robust estimators has been reported by Rock
(1987).
Also, nonparametric methods which are distribution-free are now more
often being used in geochemistry. These methods use information coded
to coarse statistical measurement scales, usually the nominal or ordinal
scales. When using nonparametric methods for the analysis of typically
proportionally scaled concentration data, one has to transform the data to
a coarser scale. In geochemistry this means that information paid for in the
form of expensive analyses is lost. Reduction of assumptions (distribution
laws) is done at the expense of lost information. The fewer the assumptions,
the less information can be extracted from the data.
Because statistical methods often require substantial knowledge of sta-
tistical assumptions, many descriptive and especially graphical methods,
some of which have been described above, are often preferred for data
viewing. Descriptive methods of this type are commonly called exploratory
data analysis (EDA), a concept introduced by Tukey (1977).
The univariate approach to anomaly recognition 241

Background Anomaly

ppm

b o o
0 · · • o
·
0 · , 0 0 Oo
°o OO,

0
0
· ·
°0
• °° P *°
One large anomaly
\ · 1Γ · 1
Many small Single points

Fig. 10-9. This bimodal frequency curve may be the result of three different regional anomaly
patterns.

Univariate regional anomalies


The nature of geochemical anomalies cannot be exhaustively described
by statistical distribution curves or histograms because the regional infor-
mation of distances between sampling points is not included. The schematic
bimodal distribution shown in Fig. 10-9 as a frequency function could have
originated from at least three regionally diverse anomaly patterns. An
anomaly may even be a random event, a sum of random errors. The real
nature is not revealed until an informative map of the concentrations
is drawn, displaying a pattern and being complemented with auxiliary
information.

Univariate maps
Detection and evaluation of anomalies from maps is a visual process,
which relies on experience. The map should display the variation of
concentration levels forming large and small scale patterns of interest to
the geochemist. The patterns can be at low or high concentration levels
and their contrast with the surrounding is not necessarily high. Here the
242 Recognition of anomaly patterns in regional geochemical investigations

problem of a geochemical map is faced: how to display concentrations


on a map so that relevant differences are shown for both high and low
levels. If the map is used for diverse purposes (geochemical investigation,
environmental studies, exploration, bedrock geology) then requirements
might well be contradictory and impossible to satisfy in a single map. The
general characteristics of a high quality geochemical map are as follows:
(1) Large areal features such as zones and large anomalies should show
up, even when the contrast between anomaly and background (local sur-
rounding) is low, so long as it is systematic (weak lineaments, geochemical
indicators of crush-zones, etc.).
(2) Regionally small anomaly peaks should be clearly shown (if reliable!).
(3) The presentation should be visually clear so that the high and low
levels are easily identified without frequent consulting of the legend.
Howarth (1983) has discussed various map presentation techniques
focusing on symbol techniques, where either symbol size or shape indicates
the concentration level at the sampling site. Symbol maps are objective in
the sense that they show the concentration value exactly at the sampling
point, allowing the reader to see the position of sampling sites in the grid
as a whole. Another advantage is that the symbols usually show the true
values, and not interpolated or smoothed "artificial" values over unsampled
areas. The symbol map also allows one to emphasize single anomalous
points, for exploration purposes for example.
Symbols are not easily designed, however, so as to clearly show concen-
tration levels and variation. The location and distribution of sites heavily
influence the visual message received. Low contrast anomalies are difficult
to see if the grid is sparse and neighbouring symbols are distant from each
other. The classification of the concentration values into a few categories
may fail and then the large areal features may not show up (characteristic
(1) above). Symbol maps are valuable nevertheless because they show the
original measurement values only slightly modified (due to classification)
or "manipulated."
The visual clarity of symbol maps has been discussed by Björklund and
Gustavsson (1985). An apparently good result is obtained if the following
guidelines are observed:
— The size of the symbols should be the major indicator, not the shape.
— The dot size should vary continuously with concentration and not
stepwise (to avoid at least partially the problem of categorizing data).
— The dot size should not vary below the detection limit.
— The dot size should vary distinctly only for reliable concentration
values (not for lower and higher values).
— The dots should be black to show high visual contrast for regional
patterns and local peaks.
— Dots should be large, arranged by overlapping so that underlying dots
are not lost (done by raster techniques, with white frames).
The univariate approach to anomaly recognition 243

Figure 10-10 shows a dot map constructed according to these guidlines.


A scatter diagram showing the replicate data is included as an insert. This
type of map presentation is clear even if the language is not known and it
exhibits characteristics (2) and (3), but maybe not (1).
The raster technique, where the plot area is divided into tiny picture
elements (pixels) and where each pixel can be addressed and coloured,
allows a dot or other symbol map to be combined with other types of
maps such as topographical, geological or cadastral maps. This advantage
of combination makes the raster technique superior to pen plotting. No
matter which technique is used, the quality of the data and the objectives
of the investigation should be taken into account in the map design.
If desired, regional information can be treated by interpolation and
smoothing to yield the regular grid that is necessary for a coloured
surface map. Each square representing a grid point can be displayed as
a coloured picture element on the map. The colour map is really useful
only if the picture elements form a contiguous surface, which is why
a regular grid must be computed. There are a variety of interpolation
and smoothing methods available in commercial program packages and
presented in the literature. The choice must be based on the particular aim
of the presentation. Some kind of smoothing together with interpolation
is justified because geochemical data always contain noise. Interpolation
methods can roughly be categorized into those that attempt optimally to
imitate the raw data (e.g., point kriging) and those that in one way or
other smooth the data (moving weighted average, moving weighted median,
block kriging). Kriging interpolation is widely used in geostatistics for
ore reserve estimation. Kriging interpolation works with moving averages
where the weights are determined using a variogram that indicates the
variation between measured values of samples at different distances. The
variogram is a function of the distance between samples and the orientation
of that distance. Kriging interpolation possesses the statistically beautiful
feature of being a "best least squares unbiased estimator." These methods
are rarely used in regional geochemistry, however, possibly because the
variogram is not representative of a large heterogeneous area. Large data
sets also require much computing time, which may be another reason for
spare utilization of kriging in geochemistry. In some situations in regional
geochemistry kriging has been applied for anomaly detection or detection
of unusual patterns. Myers (1988) defined "unusual regions" as those
where the error of the kriging estimate exceeded two kriging standard
deviations.
More often than Kriging, simpler methods such as moving average and
various trendsurface applications (splines, polynomial trend surfaces) are
employed for interpolation and smoothing. The moving weighted average
methods usually work within a moving circular window stepping from
one grid point to the next. Only points within the window are used for
Geologian tutkimuskeskus
Geokemian osasto
244

Geokemiallinen kartta

MOREENI
Zn

Esikäsittely: seulonta
Analyyailajite (pro): —62
Iiuotus: kuuma HNOa+HCl
Analyysilaite: ICP
Laboratorio; GTK
Näytteenottovuosi: 19B3
Analyyaivuosi: 1985
Pistelukumäära: 2061

Symbolikoko pitoisuusarvojen
funktiona (^—) ja pitoisuuksien
kumulatiivinen jakauma ( «ass»)

PUrretlf: 08.09. I860


Recognition of anomaly patterns in regional geochemical investigations

Fig. 10-10. A dot map of Zn in till with dot size expressed as a continuous function of concentration; measurements of duplicate samples are
shown on the scatter diagram.
The univariate approach to anomaly recognition 245

interpolation. The points are weighted according to their distance from the
window centre; typically high weights are given to points near the window
centre and low weights to points at the periphery. The weighting function
can be tailored according to the particular effects being sought.
Besides the weighted average, a weighted median can be computed
within the window (Björklund and Lummaa, 1983). The weighted median
is especially suitable for noisy geochemical data because it is insensitive to
outliers and the values obtain a just representation in the final result as
the weighting affects the frequencies not the values. Briefly, the method
delivers votes to every value within the window, where the vote is the
computed distance-dependent weight. The value that has the same number
of votes above and below is the weighted median. The result is influenced
by many control parameters, such as window size, point density, shape
of weighting function etc. These control parameters must be adjusted
purposely for each data set. Most maps in the Geochemical Atlas of Finland
have been computed using the moving weighted median.
Recommendations for plotting coloured surface maps are difficult to state
concisely. But a couple of things are clear. First, the final pixel size (the grid
square) should not exceed 1 mm 2 ; otherwise the chess board effect becomes
too disturbing. Secondly, as in dot maps, the quality of the data must be
considered and the colour scale must be tied to the concentration scale in a
meaningful way. The colour scale must be easy to use and visually decisive.
The most common and possibly the most natural colour scale is the one
picking up colours of the rainbow, where low values are associated with blue
and high values with red (cold-warm —> low-high). Many colour classes
should be used on regional geochemical maps and adjacent colours need
not differ drastically from each other. Soft fuzzy effects are then achieved
and the anomalies are not sharply outlined by contours that might lead the
reader astray.
Large-scale anomalies, and low contrast anomalies in particular, are
usually easier to detect from coloured surface maps than dot maps. There
is a risk nevertheless that an anomaly pattern becomes hidden within a
single colour class. When this is likely to happen, features can be visualized
by employing hill-shading as a complementary method. Hill-shading is a
method where each pixel is presented as if illuminated by a directed light
beam and the pixel reflects the light with an intensity dependent on its
slopes towards the source and the viewer. Shades are very powerful for
detecting features that depend only on slopes (changes) and not on absolute
concentration level. Combining a coloured surface map and a hill-shading
map offers a visually clear presentation of a complicated concentration
surface. Such maps (Koljonen et al., 1989) are strikingly informative.
Geochemical data that has been transformed to a regular dense grid
can also be treated by image processing methods. The traditional image
processing methods are filtering techniques: low-pass filters remove high-
246 Recognition of anomaly patterns in regional geochemical investigations

frequency variation and high-pass filters remove low-frequency variation;


some filters outline homogeneous basins and others sharpen contours
(sharp contrasts) giving emphasis to low contrast features (Gonzales and
Wintz, 1987). It is important to note that image processing can be used
for combining geochemical maps with each other or with maps presenting
other geodata (geophysics, geology, topography, etc.). Image processing will
certainly become more commonly used in geochemistry, as in the other
geosciences, with declining prices of equipment and programs, increasing
power and graphical performance of microcomputers and personal worksta-
tions, and increasing systematic storage and retrieval of image data (image
databases).

THE MULTIVARIATE APPROACH

A thorough geochemical investigation of an area often requires a multi-


variate view. Many geochemical features are recognized only through their
multivariate behaviour; in this case a univariate inspection of one element
at a time is insufficient. Sometimes no element alone is anomalous, but a
particular combination of paragenetic elements is rare and thus anomalous.
Such anomalies are difficult if not impossible to find by scanning through
a succession of univariate maps, no matter how well they are designed
and prepared. Human brains are capable of handling only a few variables
simultaneously! To obtain a holistic multivariate picture from, say, 20
elemental maps is humanly impossible. If the geochemical data could be
presented sufficiently readily and uniquely and if the data always were to
follow the same rules in the same situations, then rules for recognition of
situations could be stated and learned. But that is not so, possibly because
each geochemical data set is to some extent unique.
Geochemical regional investigations today tend to be multivariate in
nature, not only because instruments are capable of analyzing many
elements but because the combined information often is invaluable when
interpreting anomalies of the target variables.
The statistical analysis of several variables at once is performed by
multivariate methods, the general goal of which is to reveal and estimate
the multivariate data structure. In this context the data structure refers
to the grouping of sampling units according to element contents and the
grouping of element contents according to statistical dependence between
variables. A whole battery of multivariate methods are available for solving
these problems. Managing the methods requires hardware, software and
know-how.
Although many commercial program packages, both for micros and
mainframes, offer easy-to-use versions of multivariate methods, the user
still needs some theoretical background in order to apply them properly.
The multivanate approach 247

Methods can often be mechanically applied to data that does not satisfy
the requirements of the method, and in that case the output may well be
nonsense. Any worker intending to use multivariate methods in a serious
way is certainly advised to improve his or her knowledge through study. For
the more casual user, commercial programs already exist — expert systems
— advising on the selection of methods, both uni- and multivariate, given
the type of data and the goal of the study (Brent, 1989). Such guidance
is valuable even for the professional statistician, who does not always
remember the pitfalls.

Multivariate data structure


The statistical dependence between variables is a basic concept in
multivariate methods. The statistical dependence is defined in probability
theory: if the distribution of one variable depends on the values of another
variable then the two variables are (statistically) dependent. Statistical
dependence is not equivalent to causal dependence: statistically dependent
variables are not necessarily causally related, but their values have related
occurrence. Detected statistical dependencies may nevertheless alert the
observer to a causal dependence in the data, which can be explained by
known geochemical processes.
Many multivariate methods are based on correlation measures that
indicate statistical dependence. There are many variants of correlation
measures for different measurement scales, but here emphasis is put on the
Pearson product moment correlation coefficient, which is most commonly
used for numerical data like geochemical concentration values. If the
observed values of variables x and y almost perfectly satisfy the equation
of a straight line through the origin: y — ax (\a\ > 0), then the variables
are strongly correlated. This linear type of correlation is measured by
the Pearson coefficient (r). Negative r indicates negative correlation and
a falling line, while positive r indicates positive correlation and a rising
line. If the values do not follow a straight line then x and y are not
linearly correlated but may be nonlinearly correlated (Fig. 10-11a). A
nonlinear correlation can sometimes be forced into a linear correlation
through a transformation of variables. The linear correlation is invariant to
linear transformations (multiplying with coefficients and translating with
constants) and thus scale independent.
Particularly in geochemistry there are situations where the correlation
coefficient (r) can give misleading information about the true dependence
between variables, e.g. in an anomaly consisting of clastic and hydromorphic
parts. In looking at a scatter diagram of x versus y, the geochemist should
be aware of the behaviour of r at least in the following cases:
(1) The dots form internally uncorrelated groups located along a straight
line; r indicates strong correlation whether or not correlations prevail
248 Recognition of anomaly patterns in regional geochemical investigations

Fig. 10-11. Bivariate point patterns that could result in spurious correlation; D indicates the
detection limit.

within groups (Fig. 10-lib).


(2) The dots form groups that are internally strongly correlated but are
scattered randomly; r is small even though the within-group correlation is
strong (Fig. 10-llc).
(3) One or more dots drastically deviate from the correlated pattern
of the total data set; r is too small because the coefficient is sensitive to
outliers (Fig. 10-lld).
(4) One or both of the variables have relatively many values below the
detection limit and the values above the limit are strongly correlated while
the values below are weakly correlated; r is small and not representative of
the qualified data (Fig. 10-1 le and f).
(5) Both variables are simultaneously disturbed by a common phe-
nomenon, analytical bias for example, which dominates the variation and
causes a virtual correlation; r is great but the variables are not correlated
in nature.
These properties of the data need to be identified to guarantee a proper
interpretation of a multivariate method based on correlation coefficients.
Case (5) is hard to detect because the correlation coefficient does not of
course reveal the reason for correlation. The virtual correlation can be
detected and the reason identified from the quality control data monitoring
bias. Case (4) can be avoided by converting to ordered measurement scale,
The multivariate approach 249

putting the values below the limit into the first rank, classifying the
rest of the data into reasonably many classes, and using rank correlation
coefficients: Spearmans p, Kendall's r, etc. (Maritz, 1981). Ranking causes
loss of information (interdistances between values), but may be the only
way to go in extreme cases. Rank correlations can be used as input into some
multivariate methods, but ranked data cannot be subjected to arithmetic
operations, and this is an important caution when using methods that
try to reproduce the original values. Cases (1) to (3) may be displayed in
simple scatter diagrams and univariate distributions. If each variable is
plotted against all others then n{n — l)/2 scatter diagrams must be drawn
for n variables. It is not unusual to have 20 variables in a geochemical
investigation and if all pairs are inspected 190 diagrams will be required!
The alternative to a large number of scatter diagrams is a multivariate
graphical presentation of all variables and samples on a single plot,
showing grouping and single multivariate outliers in a satisfactory way.
The plot displays the Andrews' curves (Andrews, 1973), which describe
each sampling unit as a trigonometric polynomial where the coefficients
correspond to element concentrations. Each sampling unit is represented
by a wrinkled curve where each element appears as the amplitude at a
given wave length. The formula for an Andrews' curve is:
F(t) = Xijy/2 + x2 sin(i) + x3 cos(i) + x 4 sin(2£) + x 5 cos(2£) + · · ·, (6)
0 < t < 2π
Formula (6) represents the curve of one sampling unit, where xj is the
value of the jth. variable. Unique combinations of element concentrations
give unique curves. Grouped samples appear as a uniform band of curves
for each group and single outliers appear as "wild" curves with no similarity
to any band. Dominating wavelengths of a band can be used to identify
the dominating elements in the group. The element concentrations must
be standardized (subtracted by mean and divided by standard deviation)
to obtain comparable scales of amplitude. Like all graphical presentations
the Andrews' plot has some drawbacks. A large amount of data can cause a
virtual mess of curves impossible to distinguish. The effect of elements on
the curve depends on the order of the elements. A new ordering of elements
will result in new shapes of curves, though the bands will normally be
preserved. The first elements of expression (6) are more influential because
they cause longer waves than those at the end of the polynomial. Despite
these disadvantages the Andrews' plot has the advantage of being fast to
compute and easier to handle than a host of scatter diagrams. Figure 10-12
shows a typical set of Andrews' curves with clusters as bands and outliers
as wild curves. Each curve can be identified from the sample label plotted
at the edge of the plot.
The data structure of multivariate data can be examined in two ways: in
terms of (1) the interdependence between variables using factor analysis,
250 Recognition of anomaly patterns in regional geochemical investigations

Fig. 10-12. Typical Andrews' curves showing clustering and single outliers in the multivariate
data.

and (2) the grouping or classification of sampling units as indicated by


the variables. Different methods are used in the two cases, although the
methods can support each other (results from one method can be input to
another). Thus it is good to know the geological dependence of variables
when studying the grouping of sampling units. And in the same way it is
useful to know the grouping behaviour of sampling units when studying the
dependence of variables (geochemistry). The examination of the structure
of multivariate data by factor analysis and classification is discussed in the
following two sections.

Factor analysis

Factor analysis (FA), along with principal components analysis (PCA), is


one of the main methods used to uncover the statistical dependence between
more than two variates. A geologically oriented discussion and examples
of factor analysis can be found in Jöreskog et al. (1976). Starting from
the covariance or correlation matrix, these methods of analysis attempt to
explain the variation of the original variates by a few new variables, called
factors, without losing significant information. Generally, it is desired that
the new variables be uncorrelated in order to simplify the data structure.
In special cases, when uncorrelated factors are unsatisfactory, correlated
The multivariate approach 251

(oblique factor models) factors are applied instead. Uncorrelated factors are
easier to interpret than the correlated ones and they possess nice statistical
features that are advantageous if the factors are input to other procedures.
Factor analysis does not assume the normal law if no hypothesis tests
are performed. It must be noticed that the correlation coefficients reflect
only linear dependence and thus the factors can in turn explain only linear
dependencies. Nonlinearly dependent variables should be transformed to
yield linear dependence, mostly possible. The initial settings must be
carefully checked because, although factor analysis can almost always be
carried out, the result can be relied upon only if the conditions are satisfied.
There are many variants of FA and here only the most commonly used
method is briefly introduced from a practical point of view. FA is based
on a correlation matrix that is invariant to scales of variates, unlike PCA,
which starts from the covariance matrix, and depends on original scaling.
The FA model includes factors that are unique to a variate, while the PCA
model contains common components only. The FA model is linear and can
be expressed as:
*i = anFi + a12F2 + alsF3 + · · · 4- almFm + diUx
x2 = a2\Fi + a22F2 + a2SFs + · · · + a2mFm + d2U2
(7)
:
Xn = amFi + a>n2F2 + anSF3 + · · · + anmFm + dnUn
where X{ is the ith original standardized variate, Fj is the jth common
factor, t/fc is the unique factor of the ith variate, ay and d{ are loadings
reflecting the dependence between the variates and factors, n is the number
of variates and m is the number of common factors. The factors are by
definition standardized variates with mean 0 and standard deviation 1.
The loadings can be interpreted as correlation coefficients between variates
and factors (not true for PCA). The loadings are subject to interpretation
and the combination of elements with high loadings on a factor can give a
clue as to what natural process is explained by that factor. The degree of
explanation for each variate is the communality expressed by loadings as:
m
^= E4 1
<8>
The communalities are always < 1 , because m < n and the variates are
standardized (Var(xj) = 1). Sometimes the communality is multiplied by
100 to obtain a percentage. The rest of the variation 1 — hi is explained by
the unique factor. The proportion of variation explained by a single factor
can be computed from the eigenvalues of the FA problem. The eigenvalue
corresponds to the variance of a factor and the sum of eigenvalues
corresponds to the total amount of variance explained by the whole factor
252 Recognition of anomaly patterns in regional geochemical investigations

model (Lawley and Maxwell, 1971). The proportion of variation explained


by thejth factor can then be formulated as:
m

where λ^ is the ith. eigenvalue, pj is often expressed in % of the total


variation explained. The FA model can be fit to the data in many ways and
there are in fact infinitely many solutions. Therefore additional criteria are
needed to select the solution that is best for the situation in question. The
solution is reached in stages starting with an initial model, which defines
the required number of factors or the dimension of the model. Often the
initial model is difficult to interpret because it tries to explain all variation
in the data with one or two factors. Rotation procedures are applied to the
initial model to make the loadings easier to interpret.
Most commonly the orthogonal VARIMAX criterion is used for rotation.
VARIMAX attempts to reduce the number of absolute values of high
loadings for each factor (column) and increase the number of low loadings
to enhance the association between factors and variables. Each factor is
then easier to recognize. Once the factor matrix (loadings) is interpreted
one can go on with computing factor scores, which are values of the artificial
factors at the sampling units. A tool for quality control of computations can
now be employed to compute the correlation matrix of the factor scores.
The result should be a diagonal matrix with l's in the diagonal and (almost)
O's elsewhere.
The relevant output from a FA procedure contains:
— the factor matrix (loadings),
— the communalities of the variates,
— the proportions of variation explained by each factor,
— factor scores plotted as maps or scatter diagrams against variates.
Small data sets can be displayed and studied using biplots. A biplot shows
the observed vectors of sampling units and the vectors of variates in a plane
spanned by two factors on the same plot. A biplot may reveal associations
between factors, variates, and samples (Howarth and Sinding-Larsen,
1983). The biplot is discussed as a general graphical multivariate technique
in Gabriel (1981). The associations shown in a biplot are not exact because
the vectors are two-dimensional projections of multidimensional vectors.
Some important things should be noted when interpreting and using FA:
(1) The correlation matrix should be non-singular because otherwise the
result is poor. The matrix is singular in some special situations: if at least
one variate is constant, or if two variates correlate perfectly (r = 1); usually
the program takes care of this, warns and asks for changes in the data set.
(2) The eigenvalues of the initial factor model can be used to select the
number of final factors; a common threshold for eigenvalues of sufficiently
The multivariate approach 253

relevant factors is 1, i.e., all factors having an eigenvalue < 1 are skipped;
in general one should avoid interpreting factors with smaller eigenvalues
because they tend to reflect noise.
(3) If the correlations between variates are all close to 0 only one common
factor will be found per variate; the factor model does not then provide
the interpreter with new information. Almost equal eigenvalues indicate
the same situation. Some programs include a sphericity test to detect this
situation before the attempt is made to fit a meaningless model.
(4) The goodness-of-fit of the model expressed by communalities should
always be checked for each element; low communality, <0.5 say, reveals
that the variate is not explained by the factor model and is mainly explained
by the unique factor.
(5) If only a low proportion of the variation is explained by a factor, the
factor may be reflecting only less important variation — noise perhaps —
and should be interpreted with caution; low proportion does not necessarily
mean a bad factor, because the measure is relative and the less important
factors have already been dropped at the dimensionality choice.

Multivariate classification
Grouping or clustering of data can be investigated by graphical tools
(Andrews' plots or biplots, see above) or by some statistical classification
procedure. The statistical classification methods can be divided into un-
supervised and supervised methods depending on what initial conditions
are set. Unsupervised methods cluster the data points into homogeneous
classes which are unknown beforehand and the result is a classification of
all points into those classes. The classification is then based only on the
interdistance between points and the decision algorithm telling whether a
point belongs to a class or not. Supervised classification methods classify
data points into classes that are known or described beforehand.
Unsupervised methods are usually employed when unknown structures
are searched for. A geochemical example could be the classification of till
samples taken over an area from which very little bedrock information
is available. Then the objective of the classification could be to find any
clustering in the data to identify geochemically homogeneous lithologic
blocks or provinces in the area, aiding, e.g. the geological mapping.
Supervised methods are used when one wants to see which data points
fit known cases and which ones are different. For example, supervised
classification is a powerful approach if one wants to classify unknown
geochemical data using a well-known data bank of anomalies. The result
would show whether or not the data points match any known anomalies,
where they reside and which anomaly they match. The supervising can be
done through training data so that each class is assigned a representative
set of training samples, which are analyzed in the same way as those to be
254 Recognition of anomaly patterns in regional geochemical investigations

Quantitative Classification
Methods

Supervised Unsuper vised


methods methods

Ad hoc—methods Partitioning Methods


Non—linear Methods Hierarchical Methods
* Divisive
Linear Discr. Analysis
* Agglomerative

* Classification rules
and criteria
* Distance measures
Fig. 10-13. Schematic view over quantitative classification methods. Variants of methods are
generated by diverse distance measures and classification criteria.

classified. The computer "learns" from these training data what features
are characteristic to each class, e.g. for Sudbury type of mineralisations.
The classification methods can also be divided into categories depending
on what distance measure, classification criterion (agglomerative or devi-
sive), hierarchic or non-hierarchic structure and estimation method are
used. A schematic view of these methods is shown in Fig. 10-13.

Unsupervised classification

Unsupervised classification methods are often called cluster analysis


methods. In fact, there are many cluster analysis methods and the dilemma
is that most methods yield widely divergent results for the same data set.
Again, the geochemist has to make a choice according to what his goals are.
In cluster analysis one must decide upon:
— The similarity (or distance) measure to be used to indicate the distance
between sampling units: should sampling units be considered equal if their
concentration levels match for most elements (Cul « Cu2, N l « Ni2, etc.)
or only if their concentration levels have the same proportions (Cul/Nil «
Cu2/Ni2)?
— The desired structure: are hierarchic clusters allowed (clusters imbed-
ded in each other) or is an agglomerative structure with non-overlapping
clusters to be sought?
— The criterion to be used in steering the classification: how close should
The multivariate approach 255

points be to existing clusters to be included into one of them and how is the
"nearest cluster" defined?
This list is not complete and the investigator must be very familiar
with his method to avoid blind working and uncontrolled results. Once the
method is selected, the results of the analysis can be viewed as maps where
each sampling point is represented by coloured dot or symbol to indicate
the class index. Hierarchic clustering can be viewed as a dendrogram
showing all divisions from the lowest level where each sample is a single
cluster to the highest level where all samples belong to the same cluster. In
geochemistry a classification map often is the only way of interpreting the
meaning of clusters. Nevertheless, cluster analysis methods are not often
applied in geochemistry, possibly because the geochemical data set tends
to be large and the distance matrices involved require enormous computer
memories. The number of variables can be reduced by factor analysis (R-
mode), but the size of the distance matrix is still the same. Cluster analysis
for geologists is discussed in Davis (1973). Various clustering methods are
described in Anderberg (1973), Duda and Hart (1973), and Späth (1980).
Factor analysis can be used for classification of samples if places of
samples and variables in the raw data matrix are changed. This Q-mode FA
starts with correlation between samples and results in factors that indicate
groups of samples.

Supervised classification
Linear discriminant analysis appears to be the most popular of the
supervised classification methods. One simple reason for this is that it
is included in almost all commercial statistical program packages. The
method is also computationally very effective. Linear discriminant analysis
assumes multinormal distributions within classes and equal covariance
matrices for all classes, and to work properly it requires linearly separable
point patterns in the space spanned by the variables. This means that
the patterns should be separable by hyperplanes (lines in two dimensions
and planes in three dimensions). These assumptions are very restrictive,
however, and prohibit a proper application in many practical cases. At least
the assumptions are tedious to verify and a battery of significance tests
must be employed.
More flexibility and fewer restrictions are offered by the method of
empirical discriminant analysis introduced to geochemistry by Howarth
(1973). Empirical discriminant analysis is a statistical multivariate method
for supervised classification of objects into classes where each is defined
by a representative sample — training data. The variates observed on the
model objects and the unknown objects must have a measurement scale
that allows measurement of distance between points or samples. Thus the
method is not applicable to nominal variables such as rock type. Because
256 Recognition of anomaly patterns in regional geochemical investigations

the model objects — the training data that supervise the classification —
need not be "naturally" clustered in the variable space, the method is more
powerful than most other classification methods in situations where shapes
of classes are complicated. Though not a recent method in geochemistry
(it has been used at least since 1973 at Imperial College, London), it has
only infrequently been applied, probably because of the extended and high
precision computing required. Now that computers are drastically more
powerful and cheaper it becomes highly attractive.
Supervised classification by empirical discriminant analysis encompasses
the following steps:
(1) Learning of classes from a sample of model objects from each class
(the training set) and estimation of class-conditional frequency distribution
functions
(2) Definition of the current decision rule with subjective prior probabil-
ities and a loss (cost on misclassification) function
(3) Testing of the quality of the learning by classifying a known test
set of objects not included in the training set; if the test result is
unsatisfactory the training set must be adjusted and the quality of the
variables reconsidered
(4) Classification of the unknown objects into the learned classes by a
classification rule and display of the obtained results.
The estimation of the class-conditional frequency distribution functions
is nonparametric and no assumptions concerning the distribution law of the
variables need be made. These functions can be estimated using Parzen's
window method in which the error of measurement can be tolerated and
taken into account. The nonparametric feature of the method makes it
safe and flexible in situations where little or nothing is known about the
behaviour of the variables.
The classification rule is usually based on Bayes' decision rule, which
operates on class-conditional frequencies, prior probabilities, and a value
of loss for each class measuring the cost of misclassification. This rule is
optimal in the sense that it minimizes the average overall loss. The rule
can be simplified by the introduction of equal a priori probabilities and
equal values of loss for all classes. A threshold can be employed to screen
out those samples that do not significantly belong to any of the given
classes and should therefore be considered as outliers or unknowns. The
classification results consist of the empirical probabilities of the classes and
the index of the selected class for each object (Gustavsson, 1983).
The results of the classification are shown as maps of various types
presenting the regional distribution of the class labels or probabilities of
individual samples. A case history of this method applied to a large data set
with resultant maps has been presented by Gustavsson and Kontio (1990).
Most classification methods need a distance measure between items.
Various types of multivariate distance measures have been designed, but
The multivariate approach 257

they are typically restricted to one type of variable (nominal, ordinal or


interval scaled). Mixing variables on different statistical scales to form a
distance measure can be done by reducing all scales to the least informative
one among the scales of the variables. Then an appropriate distance
measure can be appplied to that scale. Loss of information due to scaling
according to the worst variable can be avoided in methods based on
classification trees, where the decision tree can be constructed so that only
one variable (in some sense the best one) is involved in each branching
step. All variables are involved in the classification as a whole, but no
multivariate distance measures are needed, Breiman et al. (1984).
All of the above-described classification methods — both unsupervised
and supervised — put the data points into "hard classes" in the sense that
each point must belong to exactly one class or no class at all. Fuzzy-set
theory allows a single point to belong to several "soft classes" at the same
time (Bezdek, 1981). A till sample, for example, is composed of various rock
types and has membership in all of them. It would be unfair to classify it
with just one rock type. An application of fuzzy clustering in geochemistry
and a computer program has been worked out by Granath (1984). Fuzzy
clustering has some obvious appeal in geochemistry, and in geology in
general, but few computer programs are yet available.

Data compression
If the number of variables in the data set is too large to be handled by a
multivariate interpretation procedure, some kind of data compression be-
comes necessary. Typically this situation occurs when there are restrictions
in the employed computer program; numerical problems may result when
there are a lot of variables, say more than 100. Data compression is also
sometimes important when efficient classification rules have been designed
and redundant information is undesired.
One way of compressing data is to fit a factor analysis model to them
and use the most significant factors as new variables representing the
whole data set. Not all variation of the original variables is described by
the factors of course, and some information is inevitably lost. One must
remember that the factor analysis model is optimal for the whole data set
only on average: statistically minor features do not influence the model
and may be lost in the compression. If desired, some of these original
features can be preserved by including the Mahalanobis' distance with the
factors. This measures the multivariate distance (generalized distance) of
a sample to the average sample (Everitt, 1978) and, varying with the
covariance or, alternatively, the correlation matrix of the data set, measures
the rarity of the single sample relative to the data set. Single values
of the Mahalanobis' distance are not easy to interpret, however, because
geochemically different samples may have the same Mahalanobis' distance.
258 Recognition of anomaly patterns in regional geochemical investigations

This method of data compression through factor analysis and Mahalanobis'


distance was applied in the mineral resource assessment of the Nordkalott
project (Sinding-Larsen et al., 1986).

Multivariate statistical anomalies


A multivariate anomaly is an abstract and vague concept. If the term
anomalous is understood as unusual then "unusually high values" is not
sufficient to define a multivariate anomaly. The combination of element
concentration values can be unusual in a sample even though none of the
single elements has unusually high values. Multivariate unusual samples
can be such that some elements are negatively anomalous and others
are positively anomalous. A heterogeneous data set can show unusual
features in almost all samples, because the variants are so many in a high-
dimensional space. Multivariate anomalies can be treated analogously to
univariate anomalies and classified into statistical and regional anomalies.
Statistical anomalies are detected by inspecting multivariate distribution
functions and regional anomalies are displayed on maps.
Statistical multivariate anomalies can be detected and recognized using
the cumulative distribution curve of Mahalanobis' distances mentioned
above. The Mahalanobis' distance has become fairly common for studying
deviations from the background. It offers a tool to present the multivariate
information in very compact form.
The cumulative distribution function of the Mahalanobis' distance can be
presented using values of the inverted x 2 -distribution function as ordinates
in a Chi-square plot (Garrett, 1989). The chi-square plot is constructed
analogously to the univariate Gaussian probability plot mentioned earlier.
If the data behaves multinormally and is ungrouped, the curve should
appear as a straight line on the plot. Irregularities on the curve can be
interpreted as clusters and outliers. A more or less subjective threshold
separating background, or smoothly behaving data, from the anomalies
can sometimes be found on a curve. The next step is then to identify the
samples exceeding the threshold and to check their geochemical character.
Misleading results may occur if the assumption of multinormality is not
satisfied because the chi-square distribution is not valid. This is analogous
to the problem with non-normal distributions on probability paper in the
univariate case. Another difficulty with this way of viewing anomalies
is that the fingerprints of the anomalous samples cannot be seen directly.
Samples at the same distance may be very different in chemical composition.
Nevertheless this method offers the geochemist a means of assessing the
rarity of certain samples compared with the bulk of data. A useful check on
the character of the anomalous samples is provided by Andrews' curves or
a map presenting the regional pattern of the distance values.
The multivariate approach 259

Multivariate regional anomalies


Multivariate regional anomalies are not necessarily the same as the
statistical ones. In the case of regional univariate anomalies overlapping
to form a regional multivariate anomaly, the anomaly is easily detected
by simply comparing and superimposing univariate maps. But anomalous
combinations of elements may also mean low or average concentrations for
some elements and high concentrations for others at the same time.
If a particular anomaly fingerprint is not known, a measure of rarity like
the Mahalanobis' distance can be invoked to detect the most exotic samples
in the data set. These values should be presented on maps to check their
regional patterns and their match to patterns of univariate or auxiliary
data.
If the fingerprints of some of the anomalies are known, a supervised
classification method may reveal regions matching the known anomalies.
This approach allows different backgrounds to be defined as classes, too.
The results of the classification then show regions that match to known
anomalies, e.g. Outokumpu type, and backgrounds and regions that do not.
The latter should be of great interest to the anomaly hunter.

Multivariate maps
Visually clear and readily informative multivariate maps are difficult
to design. Attempts have been made to display several variables simulta-
neously in the form of complex symbols. In such maps the multivariate
information is explicitly included in each symbol but the regional features
are hard to distinguish. Examples of multivariate symbols are Kleiner-
Hartigan trees (branches indicating element and their length indicating
content), polygons or stars (prongs indicating contents), whiskers (length
of whiskers indicating content) and ChernofFs faces (nose length, position
of eyebrows, and shape of mouth as indicators). Colours can be added
to improve the readibility of the symbols, but still the regional features
are difficult to grasp. Advantages and disadvantages of these multivariate
displays are discussed by Tukey and Tukey, (1981).
The regional aspect of the display can be enhanced by using compound
variables instead of displaying many variables. Factors and the Maha-
lanobis' distance are variables that carry multivariate information and can
easily be presented as coloured surface maps or dot maps.
If certain paragenetic combinations of a few elements are desired on the
map, shaded dots (size indicates one variable and the grey-tone another)
can be overlayed on a coloured surface map by the raster technique.
Experiences with such combination maps are reported by Björklund and
Gustavsson (1985).
260 Recognition of anomaly patterns in regional geochemical investigations

Define the goals of


the geochemical
investigation

Choose the area of


inves tiga tion ,
sampling material
and analytical
procedure
satisfying goals

Do pilot Check data Draw cumul


sampling incl. quality; use frequency
replicates for seatterdiagram
quality control and/or ANOVA

Draw dot maps o.


Use existing pilot data, asse;
data to the regional
qua 1i ty,revea1
sampling grid systematic erro
(regional block;

Design
sampling
lines
perpe'ndicula
to anomalies

Store the data


systematically for
Design the easy retrieval;use a
replicate DATA BASE MANAGEMENT
sampling and SYSTEM that is common
analyses (inc
and follow available
standard formats (if
agreed upon)

Draw the raw data Draw


as dot maps, asses frequenc
reanalysis
correction quality,reveal statistical
ste ati ariation and
(r nal blocks)

Compute the basic


statistics of the
eded, ct final data, use Draw
data base and EDA-methods, draw
sure that the cumulative curves dot maps of Contin J
X
neous data are
ectly replaced
and identify
expected and
final data
\ /
unexpected features

Fig. 10-14. Alternative routes through the various stages of regional geochemical data processing
are shown in the diagram.
The multivariate approach 261

Smoothen the
data slightly Compute a
factor

model and
factor scores

i
Draw factor
score maps
and interpret
Smoc then the data Draw colored mpare to relevant the factors
surface and/or her information;
filts hill-shaded
median) to sho maps to show ng tools if
large features details liable

Compress data
(usually needed if
many variables) by
selecting important
factors and
possibly the
Skip variables w Mahalanobis
low quality (low
reproducibility or
high detection
limit) OR rank ALL
variables to
prepare for

methods
Select an
V Unknown unsupervised method
(cluster analysis)
with an appropriate
distance measure

Com pute a rank


/NonparametriX Yes Lation

N. requested? /
structure

No

Detect multivariate
outliers in the
data using
Mahalanobis
distance and
chi-square plots

1
Prepare the final
Remove outliers report incl
temporarily and descriptions of
compute the data base and data
2
product-moment
correlation matrix
\ / quality (not
interpretation
results only)

Fig. 10-14 (continued).


262 Recognition of anomaly patterns in regional geochemical investigations

STRATEGIES FOR SELECTING STATISTICAL PROCEDURES

As has been mentioned frequently, the goals of the investigation and the
quality of data must be kept in mind when selecting statistical methods
to summarize and illustrate the geochemical information. Guidance for the
selection of mathematical methods of analysis in geology has been presented
in the form of a diagram by Agterberg (1974). Decision trees for steps within
a single complicated and multistage method (factor analysis, for example)
can also be found in the litterature. Figure 10-14 suggests statistical and
graphical methods for geochemical investigation, from sampling through
analysis to data display. While this guidance will be of little use to the
experienced geochemist, it encourages the novice to consider the ways
how to benefit from statistics, which, when collecting, analyzing, assessing,
interpreting and storing huge sets of samples and data, forms a crucial part
of the geochemical investigation process.
263

Chapter 11

GEOCHEMICAL EXPLORATION EXAMPLES

Twenty-four case histories, selected to represent a wide variety of soil,


dispersion, topographical and climatic conditions, are presented on the
following pages. Some of the studies are new and some have been taken
from the literature. In the latter case the original texts have been modified
and comments added as appropriate to meet the needs of the Handbook.

LUTSOKURU: Ni AND Cr IN WEATHERED ROCK IN HUMID HILLY


ENVIRONMENT
From: Kokko, J., 1981. Eräiden metallien esiintymisestä moreenissa, humuksessa ja
turpeessa Lutsokurun alueella Kittilässä. Unpublished M.S. Thesis in the archives
of Helsinki University.

The Lutsokuru study area is situated in Finnish Lapland close to the


Swedish border about 180 km north of the Arctic Circle. The local height
differences are considerable: the hills rise up to 370-430 m a.s.l., while
the surrounding swamps are at 300 m height level. The 30° inclination of
some valley flanks allows free drainage, but the wet peat bogs in valleys are
indicative of poor drainage. The climate is cold and moist. The fell tops are
treeless, northern spruce grow on lichen-rich ground on the valley sides,
birch and willow in stream valleys, dwarf birch on bog rims and moss and
sedge and grass in the middle of the swamp.
Bedrock, beginning from the stratigraphical bottom in the west, consists
of granite, amphibolite, N-S elongated serpentinite lenses and, above these,
alternating N-S directed horizons of mica schist and cherty quartzite
with greenstone lenses. The greenstone and amphibolite are volcanogenic,
the quartzite and mica schist sedimentogeneous. Serpentinite contains
pyrrhotite impregnated with pentlandite (0.2-0.3% Ni); between the lenses
there are skarns with more massive sulphides (0.3-0.7% Ni and traces of
Cu). Occasionally the cherts contain sphalerite and the greenstones contain
molybdenite.
The overburden consists of weathered rock regolith in valley bottoms,
deep in association with serpentinite and sulphide-rich quartzite; above
this there lies glacial till of lodgement type, except in the hill saddle where
264 Geochemical exploration examples

it is of melt-out type. Thickness of the weathered rock was not measured;


there is virtually none on hilltops but in valleys it sometimes exceeds 20
m. The average thickness of overburden is about 4-6 m. The maximum
thickness of peat in the middle of bogs is 3 m.
In the 330° oriented valleys one sometimes finds an older till transported
towards SE. Above it two younger till beds transported towards the north
are encountered — the uppermost being loose and yellowish and the
older one laminar tight and grey; both contain a normal amount of stones.
Geochemically the two youngest closely resemble each other. Their material
is local, having been transported only some tens of metres.
Sampling was done in two stages: in the first stage on E-W oriented
lines 200 m apart at 50-m intervals and in a follow-up stage with 50-100
m line spacing at 25-m intervals. Surface till samples were taken from
the B-horizon of the podsol profile with a spade, bottom material was
taken with a percussion drill and an Irish-type sampler from the maximum
obtained depth. Humus and peat were sampled manually from the same
grid.
Of 925 bottom samples 188 were classified as weathered rock, the rest
as till. The till samples were mostly a mixture of far transported and local
materials with local material dominating in the fine fraction as always in
this kind of ancient ice divide area. Both till and weathered bedrock samples
were sieved, the —0.07 mm fraction and the possible coarser fractions were
leached with hot H N 0 3 and Cu, Ni, Co, Zn and Pb were determined in
the solution by AAS. The total Cr was determined in the fine powder by
XRE Mo was analyzed colorimetrically. Humus and peat, after drying, were
ashed at 500°C, leached in HC1 and analyzed by AAS. All the results are
given as ppm in dry material. It was observed that the heavy metals tend
to be enriched into the fine fraction of weathered rock and similarly in till
(see Tables 11-1 and 11-11).
The weathered fine material is enriched in Co, Pb, Mo and Ni, while Cü
and Zn, which are more mobile, have been removed. True clay minerals
have formed in altered rocks, and they and iron oxyhydrate have captured

TABLE 11-1

Average heavy metal concentration (ppm) in different grain size fractions of weathered rock
samples

Fraction Ni Co Cu Zn Pb Mo
-0.06 mm X 1979.9 142.5 145.5 233.4 105.1 23.7
s 1217.3 114.3 162.5 466.0 241.6 60.6
+2 mm X 1025.2 57.2 115.4 208.3 34.9 10.7
s 532.0 25.0 204.0 516.9 98.9 27.2

X = arithmetic mean, s = standard deviation.


Lutsokuru: Ni and Cr in weathered rock in humid hilly environment 265

TABLE 11-11

Average metal concentrations in different sampling materials. All metal contents are given as
ppm in dry material

Material Cr Ni Co Cu Zn Pb Mo
Humus X 2.6 9.4 2.8 6.2 43.5 15.6 0.4
s 5.2 30.0 4.6 5.6 28.5 14.9 0.8
Peat X 6.1 27.8 4.3 8.0 27.2 7.5 3.4
s 12.6 91.7 6.9 13.4 86.8 7.9 8.6
Surface till X n.d. 13.6 6.6 11.6 14.9 8.5 0.6
s n.d. 10.4 3.6 15.3 7.3 3.3 0.5
Bottom material X 480.7 189.2 31.8 69.7 58.8 17.9 3.8
s 677.8 572.0 43.1 71.4 105.9 39.6 12.3

X = mean, s = standard deviation.


Bottom material: 4/5 till, 1/5 weathered rock.

heavy metal ions from water solution. Both clay and limonite are found in
the fine fraction after sieving. The easily friable minerals have been ground
to fine sizes in the till formation process. Bottom samples contained much
higher concentrations of heavy metals than surficial till samples.
Till usually represents all the lithologies of the bedrock beneath. The
surficial parts contain more far-transported material and heavy metal con-
centrations are more dilute than in the bottom parts where concentrations
are higher and variations stronger. The standard deviation of concentra-
tions in weathered bedrock samples is very large in comparison to that in
till. Mafic lithologies and rocks rich in sulphides are easily weathered and
therefore over-represented in the sampled material, with the result that av-
erage metal concentrations of weathered rock samples may be erroneously
high.
Because of these great differences in chemical composition between till
and weathered bedrock the results should have been presented on different
maps, rather than the maximum values for each sampling point being
plotted on the same maps (see Fig. 11-1A and B).
The best anomaly/background contrast was obtained in weathered rock
analysis; anomalies were strong and coherent. The Ni and Cr anomalies in
bottom samples indicated serpentinites containing sulphide mineralization.
Weathered rock seems to be a good material for geochemical exploration if
it covers the area of interest widely, even if every sample of weathered rock
represents only itself.
Geochemical investigations of weathered rock make use of the fact
that different lithologies are characterized by their own heavy metal
associations. Serpentinite bodies can be rather sharply delineated on maps
by plotting the bulk sum of Cr+Ni+Co concentrations and the Ni/Zn ratio.
993

Ni concentrations Cr concentrations
in bottom samples in bottom samples
1 /
1 / Serpentinite
1 / '////
sdjduivxd uopvuojdxd lOoxxudxpod^

Serpentinite body 1/ ■ 0 body


17 170 1700 ppm
m 0 140 5 0 0 2 5 0 0 ppm

Fig. 11-1. Distribution of (A) nickel and (B) chromium in the <0.07 mm fraction of drift (maximum values ppm at each point either in till
■1 t i l ^ ... - j . - . . ; „ n n4- T ,,4-„~1 TV - I T Λ i
Lutsokuru: Ni and Cr in weathered rock in humid hilly environment 267

Fig. 11-2. Nickel/zinc and chromium/nickel concentrations in weathered bedrock at Lutsokuru,


Finnish Lapland.

If the ratio in a till or bottom material sample is over 20 the sample clearly
represents serpentinite, but if the ratio is over 40 and simultaneously the
Cr/Ni ratio is under 2, a nickel mineralization is indicated (see Fig. 11-2).
The usefulness of weathered rock in geochemical exploration can thus be
improved by mathematical treatment of the results. Factor analysis gave
three factors: Cr+Ni, Zn+Cu and Mo. The factor score maps were not so
clear as the maps of simple element ratios.
Cherty quartzite is characterized by the bulk sum of Cu+Zn+Pb+Mo
concentrations. Granite is characterized by low concentrations of all other
heavy metals determined except Mo.
to
00
^sokuru

rtJ rw

Striae
Striae and till
and till orientation
orientation

O CK ^
£>
- ^ Diamond
°* drill hole Diamond ©
O
drill hole
8

Cr concentrations Ni concentrations o
in dry humus 3
in dry humus

Serpentinite
Serpentinite body Ö body 1
4.1 8.6 53.2 ppm
1.4 3.8 ppm
o
Fig. 11-3. Distribution of (A) nickel and (B) chromium in humus (ppm in dry matter) at Lutsokuru, Finnish Lapland.
Mäkärärova: Au in weathered bedrock in humid hilly environment 269

The same rules do not apply to peat. Mo, Cu, Ni and Cr were found to
be absorbed and enriched at the rim of the peat bog, while Co was first
captured further into the bog. What was amazing was that peat as well as
humus also gave a very distinct response to the elemental concentrations
of the bedrock. As seen from the maps (Fig. 11-3A and B) both Ni and Cr
concentrations of humus and peat indicate the southern mineralized end
of the largest serpentinite body, but fail to indicate the southern cluster
of serpentinite bodies. Still, organic materials could be recommended for
geochemical exploration.
Presentation of the assay results for different materials on the same map
is confusing and may lead to erroneous conclusions. This is true for the
organic materials, but still more so for till and weathered rock, especially
when the concentrations in weathered rock are distinctly higher than in
till.
Although the HNO3 leach used in base metal analysis is best in explo-
ration for sulphides and the total content determination best for exploration
of most oxide ores, it would have been good, for reasons of comparison of
different materials, also to have analyzed all materials for all elements with
one and the same method. Vertical profile sampling should have been done
across the anomalies to check the metal distribution in the third dimension
and find out whether the anomalies point to a suboutcrop of ore.

MÄKÄRÄROVA: Au IN WEATHERED BEDROCK IN HUMID HILLY


ENVIRONMENT
Esko Kontas

The Mäkärärova area is situated in the central part of Finnish Lapland


about 120 km north of the Arctic Circle. The area is characterized by local
deposits of bouldery till, with large peat bogs indicative of poor drainage.
The climate is cold and humid. Vegetation consists of northern spruce,
juniper and lichen with birch and willow in valleys and dwarf birch, grass
and moss in bogs.
The dominating lithology of the bedrock in the area is Archaean
gneiss granite with narrow quartzite and amphibolite zones. In places the
gneiss contains garnet and disseminations of hematite and magnetite. In
Mäkärärova itself the bedrock is intersected by numerous quartz-hematite
veins, which also contain some pyrite, chalcopyrite, goethite and a little
gold. The largest known vein is about 0.5 km long and the hematite-rich
portions in it vary from 0.3 to 2.0 m in thickness. Maximum gold contents in
hand specimens of mineralized bedrock were 5-10 ppm, while the average
content found in 34 intersecting drill cores was 2.1 ppm. The next largest
vein, 0.2 km long with similar structure, contained an average of only 0.25
ppm gold in 11 drill holes.
270 Geochemical exploration examples

Gold occurs in pyrite as a fine submicroscopic dust, but in goethite and


limonite pseudomorphs of pyrite there are gold grains as large as 5-40 μιη.
This type of gold occurs near the surface, indicating that it was moved
in solution or as a colloid before precipitation (Tuominen and Mikkonen,
1955; Säynäjärvi, 1958; Ollila, 1976).
The lowlands are surrounded by rounded hills, which like the entire
region are covered by fines-rich sandy till 0.3-3.0 m thick. In the valleys
there are fields of boulders probably lifted up by frost. The study area
itself is covered by a thin layer of sand, which further to the west is
developed into a narrow round-crested gravelly esker. The interface of till
with the weathered bedrock is transitional in places showing that abrasion
was weak on this region of the ancient ice divide and that a great part of
the till is weathered bedrock material. The ice has moved from southwest
to northeast (Rahkola, 1982), but the transport has been very short, e.g.,
the hematite boulders are concentrated in close proximity to the veins.
The thickness of the weathered regolith varies; on hills there may be fresh
bedrock immediately under till, on lowlands the regolith may be several
tens of metres thick. The weathered rock is usually soft enough for manual
digging, but the original lithological structures are preserved and only on
the very surface deformed in the direction of ice movement.
The Mäkärärova area had already been studied for gold by different
methods 2-3 times earlier. The new study was prompted by the discovery
of gold anomalies during regional geochemical mapping and the general
international interest in gold in the early eighties.
Samples were taken at 50 m intervals along lines 400 to 800 m apart and
perpendicular to the vein strike but about parallel to the glacial striae. At
each point, one sample was taken at 1 m depth and another at the maximal
depth reached by percussion drill and Irish sampler (1-10 m). Altogether
488 samples were collected, 309 of them classified as till, 109 as weathered
bedrock and 70 as sorted sediments.
Samples were dried and sieved and the -0.05 mm fraction was analyzed.
The fine fraction was leached with HCI-H2O2 and gold was enriched by
SnCl2 + Hg precipitation, after which gold was determined in HC1 + H 2 0 2
solution by AAS with graphite furnace (Kontas, 1981). The detection limit
of the method is 1-2 ppb.
Gold contents of the sorted sediments were mostly less than 4 ppb. The
analytical results obtained for the fine fraction of till and the fine fraction
of weathered bedrock coincided closely. Obviously the fine fraction of till
almost totally consists of weathered bedrock material, and the transport of
till is almost nil. The results, which thus could be handled together, are
shown in Fig. 11-4.
The maximum gold concentration found in overburden was 950 ppb.
Black hematite or goethite was usually seen in samples with high gold
concentration. A part of the anomalies were obviously derived from the
Mäkärärova: Au in weathered bedrock in humid hilly environment 271

^- 0 1 km

Fig. 11-4. Gold concentrations in weathered rock and till consisting of weathered rock material
at Mäkärärova, Finnish Lapland.

known veins, but some new gold-bearing veins were discovered as well.
Some of the anomalies even point to sources outside the study area. The
continuation of the veinlet towards the northwest and southeast could also
be deduced from the results of the preceding geochemical mapping and of
low altitude airborne geophysical mapping.
Gold occurs in such small quantities and so irregularly in overburden
that normal geochemical search by direct Au analysis can seldom succeed.
The extra sensitive method for analysis employed here, as well as the
usually high Au concentrations in the weathered bedrock, made the study
successful and the method can be recommended for exploration in similar
circumstances.
The report tells nothing about the distribution of the other elements,
such as Cu, Zn, Pb or Ag, nor about the usual pathfinder elements S,
As, Sb or Bi. In this case even the distribution of Fe could have pointed
the way towards promising follow-up targets. Kontas combined the results
of two totally different sampling materials — till and weathered rock.
Although such combination usually leads to fatally erroneous conclusions,
272 Geochemical exploration examples

in this case it could be tolerated because of the particular local circum-


stances.
The direction of the veins and of glacial transport are perpendicular in
the Mäkärärova area, which made it difficult to find an optimal direction
for the sampling lines. The veins were longer than the glacial transport,
however, so the chosen line direction was right. Usually the till transport
is considerably longer than the length of the suboutcrop of explored
mineralization.

AITTOJÄRVI: Mo CONCENTRATIONS OF TILL NEAR AN ARCHAEAN Mo


DEPOSIT IN HUMID PLAIN TERRAIN
Martti Kokkola

The study area is located in eastern Finland, in Kainuu, about 180 km


north of Kajaani. The till cover is gently sloping, clearly reflecting the forms
of the bedrock underneath. The elevation is approximately 240 m and the
relative height differences are 8 m.
The bedrock is basement granite gneiss, migmatite granite, aplite granite
intrusions with a north-south orientation and amphibolite. The molybde-
num mineralization, which also has a north-south orientation, is approx-
imately 1000 m long and 15-20 m broad. In addition to this, smaller
separate molybdenum mineralizations occur in the area.
The study area is partly covered by a 1-3 m thick peat bog. Beneath
the peat there is a 1-5 m thick sandy till layer derived from the latest
movement (270°) of the ice sheet in the area. Further away, striae inscribed
by the older ice sheet that flowed from a direction 310-320° are visible,
but corresponding till layers have not been identified with certainty in the
study area.
The aim of the till study was to clarify the extent and possible extensions
of the Mo mineralization that had been localized by other methods (humus-
peat geochemistry and drillings). Consequently, the sampling lines and sites
do not cover the study area systematically, but are concentrated in areas
selected as critical on the basis of earlier work.
The samples were collected from E-W lines 100 m apart, at intervals of
50-20 m along the lines. A gas-powered Cobra percussion drill equipped
with a flow-through type sampler was used, with the aim of taking samples
from as near as possible to the surface of the bedrock. Consequently, in
most cases varying amounts of weathered bedrock material were mixed
with the sample material proper. Some samples obviously represented this
weathered material nearly entirely. The average depth at the sampling sites
was 3.8 m. The samples were dried and the <200 mesh fraction was sieved
out. From this fraction molybdenum was analyzed colorimetrically with
dithiol.
Aittojärvi: Mo concentrations of till near an Archaean Mo deposit 273

\ Joutensuo

_D CtO— r. m - n - m A ππΐππ

ii

M
jjsi
L 200 m

10 ppm
L Mo in till 5 / Mo
0

Fig. 11-5. Areal distribution of molybdenum in till.

The distribution of the molybdenum content is presented in Fig. 11-5.


The contents of molybdenum in till are completely local; no glacigenic
features were evident as the sampling was concentrated on material
containing partly weathered bedrock material and taken close to the
surface of the bedrock. Too few samples were taken from higher up of the
till bed for conclusions to be drawn, except that the molybdenum content
in the upper part of the till bed is clearly lower.
The maximum Mo content found in the samples was greater than 2700
ppm. The sampling site in question is situated in the southern part of the
area, south of the mineralization discovered by drilling. The result indicates
that at the surface of the bedrock the mineralization has extensions towards
the south, farther than had been supposed, but in both north and south
directions it appears to terminate abruptly. Owing to the uneven coverage
of the sampling, the mineralization found by drilling is not all seen in
the results for till. Nevertheless, till sampling served well to reveal the
extensions of the mineralization at Aittojärvi.
274 Geochemical exploration examples

HARJUNPÄÄ: CHECKING OF MAGNETIC ANOMALIES BY GEOCHEMICAL TILL


STUDY IN HUMID PLAIN REGION

Martti Kokkola

Harjunpää is located about 10 km east-northeast of the town of Pori.


The land is low lying (about 24 m above sea level) and wooded; about 12%
of the area is exposed.
The bedrock of the study area is a peridotite plug occurring in mica
gneiss. An inhomogeneous, weakly orientated granodiorite lies on the
northern side and a migmatized mica gneiss on the southern side of the
peridotite. A weak Ni-Cu mineralization is associated with the peridotite.
Quaternary geological studies in the vicinity of the study area revealed
two till beds covered by younger sand, clay or peat layers. The upper,
younger till bed consists of brownish grey, structureless sandy till, the

5^—
Ni - CONCENTRATION
IN TILL (ppm)

U > 8703
O 135-8703
o 47 - 135
26-47
0-26

c) o Areas of
.<^§)) low magnetic
anomalies

0 100 m
Oi w
<0
o

o ^ -

(
0

c<:3 •
o
o o o
o O
o

LD ))

\\^^^%y o

Fig. 11-6. Nickel concentration in till (ppm).


Harjunpää: checking of magnetic anomalies by geochemical till study 275

Cu - CONCENTRATION
IN TILL (ppm)

rT&^ o °
U > 4352
o O 322 - 4352
o 41 - 322
22-41
0-22

c) β Areas of
^^^) low magnetic
anomalies

oi9
0 100 m

0
o

o ^ -

(
/fCZ_°) ■ςφ
o
0
O

c<=3 °

o
o o
o o
o

<ό >>
Si® °
o
ο-σ=

Fig. 11-7. Copper concentration in till (ppm).

older bed of bluish grey, dense, more stony gravelly till. Sand layers are
sometimes found between these two till beds. From till fabric analyses and
numerous striae observations, it was deduced that the older ice flow was
from about 320°, and the younger flow more from the west, about 285°.
Sampling was carried out across anomalies classified as interesting in an
earlier magnetic study. The intention was that samples should represent
the source of the anomaly as well as possible. The sample lines were
10-30 m apart and the sample site interval was 5-20 m. The samples were
collected from the bottom part of the till bed, close to the bedrock surface,
by a gas-powered Cobra percussion drill equipped with a flow-through type
sampler and extension rods. As a result, the sample consisted not only of
till but of weathered bedrock material. In this way the glacigenic feature of
an anomaly possibly present higher in the soil cover in the two till beds was
eliminated. The average thickness of the soil cover at the sampling sites
was 2.7 m and the maximum thickness 8.2 m.
276 Geochemical exploration exampU

Cr - CONCENTRATION
IN TILL (ppm)

W > 2000
O 1000 - 2000
o 500 - 1000
o 200 - 500
0 - 200

CD
Areas of
<X<y) low magnetic
anomalies

Of
Ό <§>

Ä
<tD

Fig. 11-8. Chromium concentration in till (ppm).

Because of the sampler and the favourable soil conditions, a rock


"button" was usually obtained in addition to the till sample. In most cases
the button sample represented the surface of the bedrock at the sampling
site. Where the till was very stony, however, it could have represented
an internal stone or boulder of the till bed, in which case the evaluation
becomes difficult.
The till samples were dried and the <0.06 mm fraction was separated by
sieving. Analyses for Cu, Zn Ni, Co, Pb and Mn were made by AAS after a
H N 0 3 leach, and Cr was analyzed by XRF.
Strongly anomalous Ni, Cu and Cr contents (maximum Ni content nearly
3%) were associated with the two most southerly magnetic anomalies
studied (Figs. 11-6, 11-7 and 11-8). Rock observations made in conjunction
with the sampling indicated that ultramafic rocks of the peridotite class
were associated with these anomalies, whereas in the more northerly
magnetic anomaly, where Ni and Cu values were low, the rock was
dominantly mica gneiss and graphite gneiss.
Hieronmäki: W, Sn, Cu and Zn concentrations near a scheelite mineralization 277

The low metal contents in the more northerly magnetic anomaly may
mean that the source of the anomaly does not outcrop or that the sampler
was prevented by stones in the till from penetrating to the surface of the
bedrock. In any event, even the low magnetite content of the graphite
gneiss is sufficient explanation of the magnetic anomaly. However, the
percussion drill-method that was used does have its limitations, as noted
above, so that the till study was able to confirm only the more southerly
anomaly as related to mineralization, while the explanation of the more
northerly anomaly remains in some doubt.

HIERONMÄKI: W, Sn, Cu AND Zn CONCENTRATIONS OF TILL NEAR A


SCHEELITE MINERALIZATION IN HILLY REGION
Vesa Peuraniemi

The study area is located in southern Finland, 7 km north of the town


of Hämeenlinna. The local topography is rugged. The absolute height
varies between 90 and 135 m above sea level. Peculiar to the area are the
bedrock-headed hills and valleys filled with clay and silt.
The bedrock of the survey area is part of the Svecokarelian orogenic belt
(Matisto, 1976). The main rock types are mica schists and mafic volcanites.
The mafic volcanites in places contain skarn breccias and skarn interbeds.
These skarns are the host rocks for scheelite and three minor scheelite
mineralizations have been found as a result of prospecting work. The
scheelite skarns contain anomalous contents of tin in places. Some sulphide
dissemination has been found in skarns, too.
One till bed consisting of grey, sandy till, pebble-rich in places has been
detected in the study area. Till covers the bedrock hills as a rather thin
blanket and follows the relief of the bedrock surface. It can be classified as
a cover moraine (Aario, 1977). On the valleys between the hills there are
deposits of late- and post-glacial clays and silts. In places, small patches of
sand occur on top of till.
The average sampling depth was 1.3 m (minimum 0.1 m, maximum
9.0 m). This is probably at the same time the average thickness of the
overburden on the till-covered hills. The thickness of the clay and silt
deposits in the valleys may range from a few metres to some tens of metres
(Virkkala, 1969).
The Second Salpausselkä ice-marginal glacifluvial formation is located
10 km southeast and the zone of the Third Salpausselkä 5 km southwest
of the study area. Thus the area is inside the former Baltic Sea lobe
(Punkari, 1979). During this deglaciation phase the ice flow direction was
from the northwest (320°), against the Salpausselkä marginal formation.
The transport distance of till on the basis of geochemical study is very
short, mostly from nil to some tens of metres.
278 Geochemical exploration examples

Till sampling was done with a light percussion drill, Partner M 100.
The through-flow pit was used as a sampler. The study area was 1.4 x 1.8
km. The interval between the sampling lines was 100 m and the interval
between the sampling points was 20 m. At places of interest the grid was
made denser, with intervals of 50 m and 5 m.
The samples were dried and sieved into three grain size fractions: <0.06
mm, 0.06-0.25 mm and >0.25 mm. The finest fraction was dissolved in
a mixture of perchloric, hydrochloric, nitric and hydrofluoric acids and
analyzed by AAS. The medium, sand (0.06-0.25 mm) and coarse (>0.25
mm) fractions of some anomalous samples were also analyzed, so that the
relationship between metal content and grain size of glacial sediments
could be evaluated. The sand fractions of some anomalous samples were
separated in tetrabromo-ethane (density 2.96 g/cm3). The heavy fraction
was studied under a stereomicroscope, and a number of scheelite grains
were counted in UV-light. Some heavy fractions were mounted in Epofix,
made into polished sections and studied with a polarizing microscope. Some
heavy fractions were also studied with a scanning electron microscope and
by X-ray diffraction and chemical analysis
The statistical distributions of tungsten, tin, copper and zinc according
to sediment type are listed in Table ll-III. The average concentrations of
tungsten and tin were higher in till than in sorted sediments. This is easy
to understand if the tungsten and tin occur in residual minerals and we
take into account that till is the first, sand the second and clay/silt the
third derivative of the bedrock.

TABLE ll-III
The statistical parameters of tungsten, tin, copper and zinc in glacial sediments at Hieronmäki

X M s c n
W (ppm) till 2.3 1.0 13.5 6.0 1773
sand 1.2 1.0 0.8 0.7 53
clay/silt 1.6 1.0 3.1 2.0 198
Sn (ppm) tiU 1.4 1.0 1.5 1.1 1009
sand 1.0 0.5 1.1 1.1 27
clay/silt 0.6 0.5 0.9 0.9 20
Cu (ppm) till 67 60 34 0.5 1773
sand 40 38 21 0.5 53
clay/silt 71 53 40 0.6 198
Zn (ppm) tiU 91 81 36 0.4 1773
sand 73 72 39 0.5 53
clay/silt 113 125 41 0.4 198

X = arithmetic mean, M - median, s = standard deviation, c = coefficient of variation, n -


number of samples.
Hieronmäki: W, Sn, Cu and Zn concentrations near a scheelite mineralization 279

TABLE 11-IV

The distribution of tungsten, tin, copper and zinc according to the grain size

Min Max X n
W (ppm) <0.06 mm 0 370 13.8 91
0.06-0.25 mm 0 199 9.3 60
>0.25 mm 0 1028 70.1 53
Sn (ppm) <0.06 mm 0 14 2.0 90
0.06-0.25 mm 0 10 1.7 60
>0.25 mm 0 20 3.0 53
Cu (ppm) <0.06 mm 25 354 74 91
0.06-0.25 mm 8 96 33 60
>0.25 mm 6 253 37 53
Zn (ppm) <0.06 mm 37 381 100 91
0.06-0.25 mm 25 160 67 60
>0.25 mm 5 800 90 53

Min = minimum content, Max = maximum content, X arithmetric mean, n = number of


samples.

The average concentrations of copper and zinc were higher in clay/silt


than in till and sand. The sand fraction is composed mainly of metal-poor
quartz and feldspars and the silt fraction contains more biotite, pyroxenes
and amphiboles. The higher modes of the distributions in sand and in
clay/silt are in all probability due to the occurrence of copper and zinc in
adsorbed form and to ions adsorbed in clay minerals and Fe-Mn oxides
(Jenny, 1968; Horsnail and Elliott, 1971; Levinson, 1974; Rose et al., 1979;
Wilhelmetal., 1979).
The copper and zinc concentrations of the till are at about the same
level as in till in the volcanite areas of Ostrobothnia and Karelia in Finland
(Kauranne, 1980).
The results of the grain size study are presented in Table 11-IV A
clear bimodality is revealed (cf. DiLabio, 1982). The average contents of all
metals analyzed were highest in the coarsest and finest fractions and lowest
in the sand fraction. The arithmetic means of copper and zinc were highest
in the fine fraction, but those of tungsten and tin were highest in the coarse
fraction. This result is unreliable for tungsten because of contamination
during sample preparation. Tin may occur as lattice-bound tin in calc-
silicate minerals (Kinealy and Eadington, 1984), which are resistant to
chemical weathering and even to strong glacial milling, whereas copper
and zinc occur as brittle sulphides which are readily ground fine under a
glacier.
The results of the mineralogical investigation showed ore minerals of
tungsten, copper and zinc in anomalous samples (Table 11-V). Scheelite,
280 Geochemical exploration examples

TABLE 11-V

Composition of fine (<0.06 mm) and medium (0.06-0.25 mm) fractions of seven till samples

Sample Fine fraction Heavy fraction


W (ppm) Cu (ppm) Zn (ppm) CaWo4 grains W (ppm) Cu (ppm) Zn (ppm)
1 2 354 229 30 60 771 519
2 1 102 381 14 30 87 3060
3 14 175 107 28 70 73 310
4 340 39 52 1000 8120 19 185
5 5 60 91 7 60 31 178
6 4 49 77 14 80 22 224
7 50 152 125 94 130 62 364

chalcopyrite, sphalerite, pyrite and pyrrhotite were all found in the samples.
The anomalous contents of tin are so low that it would be very difficult
to find grains of tin-bearing minerals. Some copper and zinc anomalous
samples contained goethite and lepidocrocite, which have been generated
as weathering products of the sulphides. The other minerals detected in the
till samples were mica, amphibole, chlorite, almandite, plagioclase, quartz,
microcline, apatite and tourmaline.
The anomaly maps of tungsten (Fig. 11-9) and tin (Fig. 11-10) in
the <0.06 mm fraction have been drawn by using the moving average
technique. The size of the window for tungsten was 80 m x 80 m and for
tin 100 m x 100 m. Two tungsten mineralizations (I and III) have been
found in the outcrops before the geochemical sampling was done. Each is
indicated by a clear and intensive W-Sn anomaly in till. Besides these, two
new anomalies were detected. The southerly one was drilled and a new
mineralization (II) was found.
The anomalies of the metals analyzed in this area are mainly clastic and
quite local. The dispersion length has mostly been from nil to some tens of
metres. Perhaps the main reason for short dispersion has been the elevated
location of the survey area.
The geochemistry of glacial sediments was studied in the area of a
scheelite mineralization in southern Finland. On average tungsten and tin
contents were higher in till than in sorted sediments. The mean contents
of copper and zinc were higher in clay/silt than in till and sand. The
distribution of metals according to grain size was bimodal, high values
occurring in both coarse and fine fractions. The anomaly contrast is
strongest in the fine fraction. Dispersion in the area has been mainly
clastic. The anomalies are quite local and therefore a dense sampling grid is
needed. As a result of the geochemical study a new scheelite mineralization
was found.
Hieronmäki: W, Sn, Cu and Zn concentrations near a scheelite mineralization 281

W ppm
+
< 3

* 3 - 4

Ο 5 - Β

Φ 7 - 10

On > 10

^ DRILL HOLE

+
τ+
♦ ♦ ♦ ♦ τ

ττ! +

+ + + + + +
+ + Φ
+ + 4
+ +
+ + +

Φ ♦ ♦ + Φ « *
Φ ♦ + + * Φ ♦

- | - 6774 750 + + Φ Φ +
+ +
<« Φ

♦ Φ + Ο
+ Φ Φ ♦

100 500m
+ Φ *
+ + Φ
♦ + Φ

Fig. 11-9. Tungsten anomaly map. I-III denote the sites of the scheelite mineralizations.
282 Geochemical exploration examples

Sn ppm

+ < 1

o o 1 - 1

o
O 2 - 2

O o 3 - 5
O
o OB > 5

o DRILL HOLE
s*
o

o
O
O

o o o o + ► ■ ♦ ♦

o o - * * -
o O O ♦ ♦ T ♦ ♦ i ♦ ♦ ♦ ♦ ♦ ♦

O o o * ♦ o * * ♦ ♦ ♦ ■♦· · ♦ ♦ *
;
o ^ ♦o o o φθ * ♦ + · ■· · ■♦ *
o o ♦ ♦ 9 J ? ♦ ♦ ♦ ♦ o * *
♦ ♦ ♦ ♦ T | * T * ♦ * ♦ o ♦ -
«■ + + + + + + +
o o *
* * ♦ r ♦ ♦ *
* o o -
♦ ♦ ♦ ♦
* o o OO o o o

+
♦ · ♦ ■ ♦ ♦ ♦

* · SWS!o 0
OO
♦ * o oto

+ 6774.750

100 500m

Fig. 11-10. Tin anomaly map.


Kissanloso: use oflithology and Cu-Mo concentrations of till 283

KISSANLOSO: USE OF LITHOLOGY AND Cu-Mo CONCENTRATIONS OF TILL


TO DETECT THE SOURCE OF Cu ANOMALY IN ARCHEAN PENEPLAIN AREA

Pekka Lestinen

Geochemical and lithological till studies were used to delineate more


accurately a Cu anomaly revealed by regional geochemical till and stream
sediment mapping (Figs. 11-11 and 11-12), to establish its source, and to

Fig. 11-11. Cu concentration in the <0.06 mm till fraction at Kissanloso based on regional
geochemical mapping of eastern Finland. The background lithological map is based on the
bedrock map by Nykänen (1971a).
284 Geochemical exploration examples

Cu ppm
o < 22.0
O 22.0 - 3 6 . 0
MVS
Q 36.0 - 5 0 . 0
• 50.0 - 7 2 . 0

• 79 >72.0

{ 1 km |

GG ARCHEAN
GG Gneissic granitoid

GR I Cataclastic granite
|[s/|\/S| Metavolcanite and
' ' -sediment

PROTEROZOIC
GG KD Metadiabase

MVS Study area


GR ^>
ELESSIIN-J
LAMP! )
LAKE
HAARAJÄRVI

Γ210

_W_

GR
90^°° GR
310

Fig. 11-12. Cu concentration in the dry matter of organic stream sediment at Kissanloso. The
lithological map by Nykänen (1971a).

assess the feasibility of follow-up exploration at the target. The investiga-


tion revealed that the anomaly is due to weakly mineralized metadiabase
dikes.
At the site the terrain slopes westward and the easternmost parts of the
area are at elevations 50 m higher than the western margin. In the middle
of the area, there is a small lake, Elessiinlampi, at 112 m above sea level.
The target lies in the Archaean basement complex. According to the
geological map (Nykänen, 1971a), the area is predominantly cataclastic
granite. At the northwestern end of the target area the rock grades into
Kissanloso: use oflithology and Cu-Mo concentrations of till 285

gneissic granite, which is the dominant lithology in the Archaean basement


complex. Older than the granite, the gneissic granite is quartz diorite or
granodiorite in composition. In the environment there are metavolcanic
rocks and, to a lesser extent, metasedimentary rocks that are older than
the granite. The metavolcanic rocks are mainly mafic, although ultramafic
and acid variants are also encountered. The metasedimentary rocks are
predominantly volcanoclastic quartz-feldspar schists. All these rocks are
intruded by abundant Proterozoic metadiabase dykes (Fig. 11-11).
The most important Cu occurrences in the immediate vicinity of the
target are the Mo-Cu occurrence in the Archaean granitoid massif at
Ilomantsinjärvi (Roos, 1981) and a Cu occurrence in metadiabase in
Hyypiä, Ilomantsi (Pekkarinen, 1976). The former is located about 15 km
northwest of the target and the latter at the same distance from the
target to the southwest. Anomalous copper concentrations are also known
in metadiabase dykes elsewhere in the area of the Archaean basement
(Nykänen, 1971b).
The poorly exposed bedrock in the study area is covered by lodgement till
deposited by glacial flow from west-northwest (Salminen and Hartikainen,
1985). The overburden usually measures 2-6 m, but it may be over 10 m
thick in bedrock depressions, where the till is generally overlain by up to
10 m of sorted sediments. In places the surface of the till-covered bedrock
is weathered (observed at four sampling sites; the soft weathered layer is
no thicker than 0.3 m).
The samples were taken on a 50 x 150-300 m grid from the base of
the till blanket at 2 m intervals upwards, but always from the C-horizon.
Attempts were also made to take samples from the bedrock surface, but
they were successful only at four sites. Sampling was accomplished with a
light-weight percussion drill Irish-type sampler.
Till samples were screened into the following fractions: <0.06 mm,
0.06-0.5 mm, 0.5-2 mm and 2-20 mm. The <0.06 mm fractions of all the
samples were chemically analyzed as were the comminuted pebbles from
the 2-20 mm fractions of 20 samples. The weathered samples collected
from the bedrock surface were treated in the same way as the till samples.
The <0.06 mm fractions from till and weathered bedrock were analyzed
for Co, Cr, Cu and some other elements by FAAS (samples digested in hot
7.25 M HN0 3 ) and for Mo by GAAS (samples dissolved in hot 6 M HC1 +
H 2 0 2 ) at the Geological Survey of Finland. The 2-20 mm fraction of the
weathered bedrock samples was also analyzed for these metals. The pebbles
of the till samples were analyzed only for Cu by the same method as the
fine fraction.
The lithological studies, which were based on pebbles from the 2-20 mm
fraction, were conducted under a binocular microscope. The percentages of
the lithologies were estimated visually.
Most of the anomalous Cu contents in the <0.06 mm till fraction in
286 Geochemical exploration examples

Cu ppm
Δ < 100
+ 100 - 129
□ 130 - 149
U 150 - 188
■ 189 - 208
■ 210 > 208
• Not determined

0.5 km

* Maximum content
in the upper or
middle level of
the till blanket

++ 0-*.+ Ά^Τ

Fig. 11-13. Cu concentration in the <0.06 mm till fraction (maximum concentration at each
sampling point). The classification of the data is based on the values of the 60th, 80th, 90th,
95th and 98th percentiles of the cumulative frequency distribution.

the study area are concentrated south of Elessiinlampi (Fig. 11-13), where
they form a large and coherent anomaly about 1.5 km long, up to 400 m
wide and trending 290°. Vertically the anomaly is distributed so that its
southern margin, excluding the very western end (line 4), is at the base
of the till cover. Towards the northern margin, anomalous values are also
occasionally encountered in the middle and upper layers of the till cover.
The values are highest in the middle of the anomaly, where the copper
Kissanloso: use oflithology and Cu-Mo concentrations of till 287

Fig. 11-14. Mo concentrations in the <0.06 mm till fraction (maximum concentration at each
sampling point). The classification of the data is based on the values of the 60th, 80th, 90th,
95th and 98th percentiles of frequency distribution.

contents reach 450 ppm. Another anomalous zone lies parallel to this zone
and immediately to the south. It is composed of two separate anomalies, of
which the eastern one is in the upper part of the till cover. Elsewhere the
Cu anomaly is less coherent.
The Mo concentrations in the fine till fraction are modest (Fig. 11-14),
the maximum value being 5.7 ppm. However, a well-defined, fairly large
anomalous area occurs southwest of Elessiinlampi. At its southeastern
end the anomalous concentrations are often in the upper part of the till
blanket. As not all the sampling lines in the southeast of the study area
288 Geochemical exploration examples

were analyzed for Mo (lines 5, 7, 9 and 11), it is not easy to compare the
anomaly patterns of molybdenum and copper. It appears, however, that
although the anomalies overlap to some extent, the patterns differ from
each other.
In the 2-20 mm fraction the bedrock fragments show a clear systematic
variation in lithology. In the area southeast of sampling line 5 (Fig. 11-15),
granite and metadiabase predominate, whereas in the northwest gneissic

Fig. 11-15. Metadiabse fragments in the 2-20 mm till fraction. Each sampling point is
represented by the sample with the highest Cu concentration in the <0.06 mm fraction.
Numbers in parentheses refer to samples analyzed for copper (see Table 11-VI).
Kissanloso: use oflithology and Cu-Mo concentrations of till 289

granite and metadiabase prevail although the abundance of metadiabase is


lower than southeast of the line. The southeasternmost samples, in which
the fraction is composed of gneissic granite, derive from sampling line
3. The abundance of gneissic granite fragments is up to 25% and even
higher at several sites in the middle of lines 10 and 12. Metavolcanites
and metasediments occur randomly throughout the area, the abundance
increasing northwestward and towards the surface of the till cover. Quartz-
feldspar schist is the predominant supracrustal rock and occurs as the
southeasternmost lithology on sampling line 10. Like other supracrustal
rocks, however, its abundance exceeds 25% at only a few sites.
No chalcopyrite or other copper minerals were observed in bedrock
fragments; nor was molybdenite. In contrast, iron sulphides or oxides,
possibly weathered sulphides, were found in some granite, gneissic granite,
metadiabase and metasediment fragments. However, there is no clear
correlation between the anomalous copper values in the fine till fraction
and the occurrence of iron sulphides and oxides.
The anomalous Cu values in the fine till fraction correlate clearly with
the abundance of metadiabase fragments, and 87% of the samples in which
the copper abundance exceeds the 95th percentile (208 ppm Cu) have

TABLE 11-VI
Cu concentration of the <0.06 mm fraction and metadiabase fragments in some till sample
from the Kissanloso study area

Sample number FF (ppm) RF (ppm) FF/RF


01 126 39 3.2
02 124 72 1.7
03 82 87 0.9
04 118 51 2.3
05 102 54 1.9
06 105 360 0.3
07 240 117 2.1
08 210 105 2.0
09 141 69 2.1
10 201 147 1.4
11 150 75 2.0
12 126 78 1.6
13 123 63 2.0
14 75 210 0.4
15 69 66 1.0
16 330 150 2.2
17 240 150 1.6
18 450 111 4.1
19 150 108 1.4
20 330 150 2.2
FF = <0.06 mm fraction, RF = metadiabase fragments in the 2-20 mm fraction.
290 Geochemical exploration examples

bedrock fragments of which 75% or more are metadiabase. The correlation


is also evident between the distribution patterns of the copper values in
the fine till fraction and the anomalous metadiabase fragments (Fig. 11-
15). However, anomalous abundances of metadiabase are also encountered
at sites where the Cu values are not anomalous. Furthermore, in the
Cu anomaly south of Elessiinlampi, six samples contained keratophyric,
feldspathic metadiabase material (Fig. 11-15), which, according to Nykänen
(1971b), represents late crystallites occasionally enriched in chalcopyrite.
It is concluded that Cu anomaly in till is caused by metadiabase dikes
containing weakly mineralized portions.
The metadiabase fragments from 20 of the samples with 75% or more
of the fragments were analyzed for copper. The Cu data show a clear
correlation with the Cu abundances in the fine fraction (Table 11-VI). If
samples 6, 14 and 18 with abnormal abundance ratios are omitted, 0.85 is
obtained for the linear correlation coefficient. The abundance ratio (Cu in
fine fraction/Cu in fragments) has a median value of 1.9. It is worth noting
that comminution tests made by Salminen (1980) on similar metadiabases
from the Archaean basement in eastern Finland showed that Cu was
enriched in the <0.06 mm fraction by a factor of 1.8.
The anomalous Mo abundances in the fine fraction do not exhibit the
same lithological correlation as the Cu abundances. However, there seems
to be a weak correlation between the Mo values and the abundance of
granite fragments, as 57% of the samples with Mo values exceeding the
95% percentile (3.0 ppm) had 75% or more granitic pebbles. Similar studies
were not made on the other elements analyzed.
Bedrock could be sampled at only four sites. At each site the sample was
from weathered rock, which at three sites was recognized as metadiabase
but at one site could not be identified due to advanced weathering (Fig.
11-15). The Cu concentration in the <0.06 mm fraction of the weathered
metadiabase samples was 66-151 ppm, with the highest value being
from the southwestern end of sampling line 11 and the lowest from the
northeastern end of the same line.
The occurrence of weakly mineralized metadiabase dykes at Kissanloso
shown by the present study was a major new contribution to understanding
the bedrock in the area. The target does not call for follow-up exploration.

KYLMÄKOSKI: BOULDER-FAN STUDY NEAR A PROTEROZOIC Ni-BEARING


ULTRAMAFIC INTRUSION IN HUMID PLAIN TERRAIN

Martti Kokkola

The study area is located in southwestern Finland, about 40 km south of


Tampere. The terrain is gently sloping and cultivated fields are intermixed
Kylmäkoski: boulder-fan study near a Ni-bearing ultramafic intrusion 291

with woodland (90-100 m above sea level). The height variations are only
a few metres. Exposed bedrock makes up a small proportion of the area.
Information about the bedrock is based principally on drilling results
and on information obtained from Kylmäkoski mine. The bedrock consists
mainly of mica gneiss. The mica gneiss is cut by granites and pegmatites,
and in places there are smaller horizons of amphibolites and ultramafic
rocks. The Kylmäkoski ore is a Ni mineralization associated with differen-
tiated ultramafic rock.
Two till beds of different ages have been identified in the vicinity of
the study area. The younger one, caused by the due westerly flow of the
continental ice sheet, was generally more prominent in the exploration
as the boulder fans were oriented in the same direction (270-285°). The
thickness of the younger bed ranges between 1 and 5 m. Striae etched
by the older ice sheet flowing from a more northern direction (325°) were
encountered, but till beds deposited by it were less commonly identified and
the variations in their thickness are not known.
The study was spurred by the mafic ore boulders found southeast of
the Kylmäkoski mine. Their source from a direction of 280° had not been
proved. As a result of excavator pitting and careful boulder tracing in the
area, nearly 300 different types of ore boulders were found. The till cover
proved to be very thick (more than 7 m) and tightly compacted and to
contain large boulders. On the basis of the structure of the till and stone
counts the deposition was classified as a Rogen moraine. Careful stone
counts were done on the material excavated from the pits and on the
surface stones of the area. By comparing results of the counts with the
bedrock data it was concluded that the transport distance was not more
than 1-2 km.
A total rock analysis by XRF was made for more than 200 separate boul-
ders found in the field and for typical samples taken from the Kylmäkoski
ore. The analytical data were interpreted by discriminant analysis.
The results indicated that the boulders, which even macroscopically
could be divided into types resembling those occurring in the Kylmäkoski
ore, were so closely similar in chemical composition to the ore that, to
a high degree of probability, they could be regarded as derived from this
known ore (Fig. 11-16).
The longitudinal orientation of the Kylmäkoski ore differs only slightly
from the more northern flow direction of the ice sheet. Striae caused
by the flows of the ice sheet from both directions are found on the
surfaces of roches mountonnees occurring close by the open mine pit. These
indicate that the ore outcrop could have been eroded by flows from both
directions.
The results of the study suggest that the Kylmäkoski ore was eroded by
two different flows of the ice sheet. The older one, from the direction 325°,
was able to act on the ore over a longer distance, because the flow direction
292 Geochemical exploration examples

Fig. 11-16. Ore boulders and boulders of mafic rocks at Kylmäkoski.

was about the same as the longitudinal orientation of the ore. Consequently
this flow, the erosional effect of which was also stronger, was able to quarry
a great many boulders from the ore. These were transported approximately
one kilometre towards the southeast. By contrast, the younger flow of
the ice sheet from 280° crossed the ore transversely. Unfortunately, the
land on the eastern side of the ore has been fairly extensively cultivated,
making the comparison more difficult. However, fewer boulders have been
associated with the younger than with the older boulder fan. How a till
Laukunkangas: Ni, Co and Cu concentrations of till at a mafic intrusion 293

formation deposited by the older flow (325°) of the ice and not covered by
younger till could be preserved as a relict on the southeastern side of the
ore is a mystery. The geochemical study of tills had helped in solving the
question.

LAUKUNKANGAS: Ni, Co AND Cu CONCENTRATIONS OF TILL AT A


PROTEROZOIC MINERALIZED MAFIC INTRUSION IN HUMID PLAIN
TERRAIN
Martti Kokkola

Laukunkangas is located in eastern Finland, about 20 km north-


northwest of Savonlinna. The area is wooded hilly terrain with an ele-
vation of about 120 m and relative height differences of about 20-30 m.
The object of the study, the Laukunkangas hill, rises approximately 15
m above the neighbouring peat bogs and small lakes. The landscape is
dominated by drumlins oriented in northwest-southeast direction, which
strongly emphasize the topography of the bedrock of similar orientation.
The bedrock in the Laukunkangas area is composed of a differentiated
intrusion of the olivine-tholeite series in which norite is the dominant
rock. Pyrrhotite and pentlandite, and to some extent chalcopyrite, are
encountered in the ultramafic part of the intrusion.
Two till beds occur in the study area, which at least in places are
separated by gravel and sand layers 1-3 m thick. Most of the mineralization
proper is covered by the thin younger till bed, 1-2 m thick, but where both
till beds are represented the till may be 20-30 m thick. The ice sheet
that deposited the younger till bed flowed from west-northwest (300-310°),
whereas the orientation of the older till bed is considerably more from
the north (350°). Striae corresponding to both flow directions are found in
outcrops of the study area.
The sampling was carried out in 1970, partly with an Auger-type hand-
driven percussion drill and partly by hand excavations of pits about 1 m
deep. Because of the variation in thickness of the overburden, sampling
was done at depths ranging from 0.5 to 1.2 m and was directed only at
the younger till bed. The sampling interval was 40 m in each direction and
the longitudinal direction of the study area followed the flow direction of
the ice sheet. The sample material was fine till. After drying, the sample
was sieved to <200 mesh and nickel, copper and cobalt were determined by
AAS method after H N 0 3 leach.
The results are shown classified into content levels in the maps of
Figs. 11-17, 11-18 and 11-19. The Laukunkangas mineralization is clearly
reflected in the pattern of each metal. The copper anomaly is completely
local, whereas a weak glacigenic feature is observed for nickel and cobalt.
The striking local character of the anomaly is dependent on the thinness
294 Geochemical exploration examples

Ni
in < 0 . 0 6 mm till fraction

X ore outcrop

Fig. 11-17. Nickel concentration in till (ppm).

Cu
in < 0 . 0 6 mm till fraction

ppm
-19
o 19-48
O 48-100
O100-

X ore outcrop
o o

°Vo°o»v;o0o
o
o o

Fig. 11-18. Copper concentration in till (ppm).


Laukunkangas: Ni, Co and Cu concentrations of till at a mafic intrusion 295

Co
in < 0 . 0 6 mm till fraction

ppm
o°o.°„ . -7
r.°„°.°o
olio 0 °o°qrCh°o°o ° o 7-16
O 16-26
θ 2 6 -

° oo° o oA°o o o oo"0 o o χ oreoutcrop


0
o°o°?tto

Fig. 11-19. Cobalt concentration in till (ppm).

of the overburden. Farther to the north where the till cover is 20-30 m
thick, the mineralization is not detected in the till despite the higher metal
contents there. Evidently it is effectively concealed by the thick till deposit.
Moreover, if a glacigenic anomaly of the mineralization did exist it would
lie — because of the more northern flow of the ice sheet — in the same
area as the glacigenic anomaly caused by the younger ice sheet.
A similar till study with a sparser sampling coverage of 200 x 50 m
was carried out earlier in the area. At that time, however, the lines
were disadvantageously located, with the sampling sites nearest to the
mineralization located in the up-ice direction. No anomaly due to the
mineralization was detected there, nor at the nearest down-ice sites,
approximately 200 m away from the mineralization.
It should be mentioned that the Laukunkangas mineralization had
already been localized by other geological methods (drilling, etc.) and the
study described here was carried out for the purpose of clarifying the nature
of the till anomaly by geochemical means. Subsequent drilling based on
integrated use of different prospecting methods, till geochemistry among
others, in 1984 led to the discovery of the economic orebody.
296 Geochemical exploration examples

MAASELKÄ: DISCOVERY OF A Co-Cu MINERALIZATION BY STEP-BY-STEP


PROCEEDING TILL GEOCHEMICAL STUDY IN HUMID PLAIN TERRAIN
From: Pulkkinen, E. and Rossi, S., 1984. Gradually proceeding geochemical exploration,
a case history of the discovery on Maaselkä copper-cobalt-showing in Finnish Lap-
land. In: Prospecting in Areas of Glaciated Terrain, Glasgow, 1984. Institution on
Mining and Metallurgy, London, pp. 1-5.

Maaselkä lies near the northern border of the Karelidic schist belt. The
bedrock is overlain by 0.5-10 m thick glacial drift, mostly till. Preglacial
weathering crust covers the surface of the bedrock, because the area lies in
the ancient ice-divide zone of Central Lapland.
The geochemical investigations were carried out in several stages. In the
reconnaissance stage till and organic stream sediments were used. In the
regional stage till samples were taken at a higher density. In the detailed
stage till and weathered bedrock samples were taken from geochemically
interesting targets and finally the anomalous areas were investigated by
trenching and diamond drilling. As a result a copper-cobalt mineralization
was found.
Maaselkä is situated 170 km northeast of Rovaniemi in northern Finland
(Fig. 11-20). This area lies near the northern contact of the Karelidic
schist belt, in the eastern part of the Central Lapland greenstone belt.
The bedrock of the study area is composed of metasediments: quartzites,
graywacke schist-phyllite and black schist and ultramafic komatiitic vul-
canite (Mutanen, 1976). The rock types are striking in E-W direction and
dipping 10-30° to the north and they are penetrated by gabbro sills.
Maaselkä belongs to the Central Lapland peneplain with elevations
of 280-300 m. The area forms a watershed which is gently sloping to
the south. The ground-water discharges as large springs in the southern
seepage zone of the canyon.

B | Gneiss and schists, old


ED Granullte
ΕΞ3 Karelian schists
1 I Orogenic plutonites

150 km
(*) Maaselkä

Fig. 11-20. The Maaselkä study area, northern Finland.


Maaselkä: discovery of a Co-Cu mineralization by till geochemical study 297

RECONNAISSANCE SAMPLING STAGE


• • • ill
φβ28

• • • •
• • · • . |I • II
• · • · •
• • • · · •
M a a s e•j k a· •
1 φΐ220 ' ·*« Maaselkä •
#
A 254 · • • •
1<-φ -
• • • •
• • • · •
Copper (ppm) in till Cobalt (ppm) in till
• •
• φ2βθ·
(fraction -0.06 mm) • • (fraction -0.06 mm)

• •• • < 88 • < 21
• 88 - 137 • 21 - 29
• •
• • 138 - 175
• · • 30 - 33
• • • 176 - 250 • 34 4 0
• •
φΐ220 > 250 I
o • > 40 o
• · 7520-I #52« · 7520-J
I I Area in the regional sampling stage

Fig. 11-21. The distribution of copper and cobalt in till, sampling density one sample per 4 km 2 .

Till covers the whole area and rock exposures are scarce. The 0.5 to 10
m thick glacial drift consists of local till where it is thin. Two till beds with
a cemented gravel interlayer occur in places with thick drift. In places the
surface of the bedrock is weathered both physically and chemically down
to a depth of 30 m. This preglacial weathering crust is common in the
ice-divide zone of Central Lapland (Penttilä, 1963). The glacial erosion was
rather weak in the Maaselkä area. The transport direction of the youngest
glaciation was from north to south (Hirvas, 1977).
Till, organic stream sediments and weathered bedrock were sampled in
the geochemical investigations. In the reconnaissance stage till samples
were taken with a light percussion drill at a density of one sample per
4 km 2 from an irregular grid (Fig. 11-21). The cemented gravel bed was
unpenetratable for the drill. The stream sediment samples were collected
at 250 m intervals along the streams (Fig. 11-22). In the regional stage till
samples were taken with the drill along lines at a density of 12 samples per
km 2 (Fig. 11-23). In the detailed stage samples of till and weathered bedrock
were taken from geochemically interesting targets with light percussion
and heavy pneumatic drills and also from trenches. These samples were
taken along lines and trenches spaced 50-100 m apart at 1-20 m horizontal
and 0.5 m vertical intervals (Figs. 11-24 and 11-25). After location of the
source of the highest copper and cobalt anomalies, rock samples were taken
with a diamond drill (Fig. 11-26).
298 Geochemical exploration examples

RECONNAISSANCE SAMPLING STAGE


Organic streamsediment

1 * L •

L s
% • : Copper, ppm Cobalt, ppm
< 6 < 21
1
'
11
' :
6-15
16-30
>30
t •

21 - 1 4 0
141-200

'■ v ··..* · II
'' . ■·
«
· ·« · .
• >200


- " ■ * .'
A->,
1 ·

J%ss&*< .**%%**, "


► '·
Maaselkä ^ Maaselk* ™/f j


>'
\
S
.0

f *

λ^
*
9 4

c
..··/ * */* t

[/_
Φ
V O o
l" M
m
Li
Ift
J
7520 7520 J
I I Area In the regional sampling stage

Fig. 11-22. The distribution of copper and cobalt in organic stream sediment, sampling density
one sample per 0.5 km 2 .

Till and weathered bedrock samples were sieved and the <0.06 mm
fraction was analyzed. Organic stream sediment samples were ashed. The
till samples from the first two stages were analyzed by optical emission
spectrometry for 17 elements. The organic stream sediment samples and
all samples from the final stages were analyzed by AAS for 7 elements after
leaching in 6 M HC1 and 7 M HN0 3 , respectively (Gustavsson et al., 1979).
Average concentrations of copper and cobalt in till in the Central Lapland
schist belt around the study area are 105 ppm Cu and 34 ppm Co and in
weathered bedrock 176 ppm Cu and 32 ppm Co (Kauranne et a l , 1977).
In the reconnaissance stage of this study the maximum contents of the
trace metals in till were high, but high values did not form homogeneous
anomalies (Fig. 11-21).
The concentrations of the trace metals in the organic stream sediments
were based on dry weight. The arithmetic mean of the distribution of
copper was 8.53 ppm and the standard deviation is 8.25; for cobalt the
values were 50.8 ppm and 69.8, respectively. The number of samples was
REGIONAL SAMPLING STAGE

Copper (ppm) in till ! ■■ i Cobalt (ppm) in till


(fraction -0.06 mm) (fraction -0.06 mm)
< 180 < 40
• 180 - 795 • 40-60
• > 795 i—s—r • > 60

*Maas*elkä > t «
? i » Maas'.elkä · 3r-
? :

l i t ! !
ii». J i !
i
i- X.
-783I i- - 7532 - J
Γ Η Area in the detailed sampling stage
! i 2 km 9
&
Fig. 11-23. The distribution of copper and cobalt in till, sampling density 12 samples per 1 km 2 .
3
DETAILED SAMPLING STAGE

Copper (ppm) in till Cobalt (ppm) In till e


(fraction -0.06 mm) • (fraction -0.06 mm)
ii o
. < 160 •·· • <39
• 160 - 240 · . · · · • 39-66
• 241 . 400 : • 6 7 - 93
Ί : i l ΐ ■
• > 400 • i ·>. ·.
Ί • »93

* ' §
• M aaselkä A
t :.:! Hi
• ! A 1 * t :
• ! •
• t
• JÄIIII CO
• f •
;· · • · ··· N
l # ·:
HI i jfc
• •
-*«· • ·
#· •
• • to
• CD
· : 8 CO
7,
533.0 J
A__? Diamond drilling profile Northern edge of komatiite layer
1 km

Fig. 11-24. The distribution of copper and cobalt in till, varying sampling density. Each symbol represents the mean concentration of the
metals in an area 50 x 50 m.
o

DETAILED SAMPLING STAGE

Copper (ppm) in Cobalt (ppm) in


weathered bedrock weathered bedrock ι|ι
(fraction -0.06 mm) (fraction -0.06 mm) I
: ■ ; » :
:
< 100 *· * < 34 ' ·? s . ·
• 100- 350 , ♦ «. • 34-63 • II
• 351-1000 • 64-120 • • Ii
• >1000 · • >120 •
M aaselkä At t M aaselkä Aj

: i · i 11: t i I : · ! :
• • t
. t · :
I »



? : • TO
o
TO
1
-7533.0- -7533.0-
TO
A__J Diamond drilling profile ' Northern edge of komatilte layer 1km

Fig. 11-25. The distribution of copper and cobalt in weathered bedrock, varying sampling density. Each symbol represents the mean
concentration of the metals in an area 50 x 50 m. -8
o
3
O
3

I
Maaselkä: discovery of a Co-Cu mineralization by till geochemical study 301

Co ppm
DIAMOND DRILLING PROFILE r300 Tfll
200 □ Woetherod bedrock

Greywacke schist-phyllite
Black schist
Arkosic quartzite, keratophyre
Gabbro
Komatllte
2SS Cu 0.1 -0.5 %
Cu 0.5 · 7.0 %
\ Co 0.02-0.15%
Sulphides are disintegrated
by weathering
Diamond drilling hole
* Stratigraphic bottom

Fig. 11-26. The distribution of copper and cobalt in the mineral deposit of Maaselkä (lower
part) and in the overlying till and weathered bedrock (upper part).

335. Statistical parameters describing the distribution of copper and cobalt


in till and weathered bedrock are given in Table 11-VII. The drainage
area of the river Viuvalonjoki seems to be anomalous for both copper and
cobalt (Fig. 11-22). Copper anomalies of the regional stage are grouped
into two zones following the strike of the rock types at Maaselkä. The
cobalt anomalies are mainly associated with the ultramafic volcanites in
the southern part of the area. The anomaly contrast is distinct especially
for copper (Fig. 11-23).
The detailed stage showed lower concentrations of copper in till than the
previous stage, because each symbol represents the mean concentration of
several samples taken from an area of 50 x 50 m (Fig. 11-24). The two
302 Geochemical exploration examples

TABLE 11-VII

Statistical parameters describing the distribution of copper and cobalt in till and weathered
bedrock
Cu a Co a Cu b Co b Cu c Co c
Arithmetic mean 148 43 392 35 1980 52
Median 134 29 140 26 389 35
Standard deviation 259 30 829 83 5030 52
Range 32- 6530 12-162 13- 14,000 2- 24 1- 70,260 1-430
Number of samples 542 542 966 966 689 689
a
Regional stage (till fraction <0.06 mm).
b
Detailed stage (till fraction <0.06 mm).
c
Detailed stage (weathered bedrock, fraction <0.06 mm).

metals together form four anomalies but the anomalous contents of the
metals separately do not necessarily coincide. With increasing sampling
density and higher threshold level the cobalt anomalies tend to indicate
mineralized areas in the metasediments north of the ultramafic vulcanite.
The anomalous patterns in weathered bedrock are reflected in till with
the exception that the most significant copper anomaly now stretches
eastward under the cemented gravel bed in the eastern part (Fig. 11-26).
To investigate the significance of the eastern anomaly a diamond drilling
program was carried out along the profile A-B (Fig. 11-26). The maximum
copper content in the bedrock increases in places to 7% and the cobalt
content to 0.15%, but the low mean content of the metals and the small
size of the mineralization do not make it an economic ore.
This case study shows that the geochemistry from a low density sampling
grid of surficial materials brings out anomalous areas especially when both
till and organic stream sediments are used. During a follow-up study a
copper-cobalt deposit was found at Maaselkä.

METSOLA: DISCOVERY OF A Cu-Au MINERALIZATION BY STEP-BY-STEP


PROCEEDING TILL GEOCHEMICAL STUDY IN HUMID HILLY PENEPLAIN

Terho Koivisto

The study area (Fig. 11-27) is located in Northern Finland, 30 km


southwest of the city of Rovaniemi, at an elevation varying between 80 and
100 m. The summits of the neighboring hills rise to a height of 220-260 m,
the most noteworthy of these, Pisavaara lying close by the detailed study
area.
The study area is near the middle of the Peräpohjola schist belt, where the
bedrock consists mostly of Jatulian quartzites associated with conformable,
differentiated albite diabase dikes. A narrow greenstone horizon of tuffite
Metsola: discovery of a Cu-Au mineralization by till geochemical study 303

30°

I B Archean gneiss and schists


t·*;·*;';';] Granulite
ΕΞΓ^Ι Karelian schists
1 1 Orogenic plutonites

1 5 0 km

(x) Metsola

Fig. 11-27. Location of the Metsola survey area.

| | Quartzite \:&:\ Albite Diabase


|gggj Greenstone \ Fault
Area of detailed geochemical survey

Fig. 11-28. Bedrock of the survey area according to Peuraniemi (1982) and Perttunen (1989).

origin is seen above a diabase dike in the southeastern part of the area (Fig.
11-28).
At Petäjäskoski, 5 km southeast of Metsola, there is a Cu-Fe formation
associated with the contact of albite diabase and quartzite (Peuraniemi,
304 Geochemical exploration examples

1982). At Kivimaa, 20 km southwest of Metsola, there is a Cu-Au dyke


formation associated with an albite diabase dike cutting a middle-Jatulian
greenstone, and at Vinsa, 10 km northwest of Metsola, a Cu-Au dike
formation of the same type in a differentiated albite diabase. The formation
at Vinsa occurs as a conformable dike in greenstone, beneath which lies
Jatulian quartzite (Rouhunkoski and Isokangas, 1974; Silvennoinen et al.,
1980).
The extensive Cu-Co-Au-bearing mineralized zone associated with albite
diabase at Metsola, was located by geochemical mapping. Pyrite, chal-
copyrite and magnetite are present in the albite diabase as disseminated
deposits. The same ore minerals have been identified in till samples.
Till and peat are the major surficial materials of the overburden in the
study area. Beneath the peat lie sorted materials, mainly postglacial sea and
river sediments. Till in the sampling area occurs predominantly as a flat
lodgement till. In the western and northwestern parts of the area there are
small rogen moraine ridges oriented approximately north-south, whereas in
the southern and southwestern parts of the area the lodgement till around
the bedrock hills is heaped into drumlin-like formations with longitudinal
axes following the southwest-northeast orientation of the bedrock (Fig.
11-29). Drilling and pit observations show the thickness of the overburden
to vary from 0.1 to 14.0 m.

] Peat J Till | Bedrock

Area of detailed geochemical survey

Directions of ice flow

Fig. 11-29. Quaternary deposits of the survey area.


Metsola: discovery of a Cu-Au mineralization by till geochemical study 305

AREA OF RECONNAISSANCE GEOCHEMICAL SURVEY

Jb__
/ ■ <: ■
Polar circle 2 5*30
Rovanie mi Bp" '""""~

v· ^ >''···>
^ -m
( A '· /^·
V- •·
χ·^β^~
(
N

SWEDEN
i.·
\ ' *f* ' *
• V
Φ/
J.
·
'
· /
7 FINLAND «
Ί· '
\
?v ·/·'''''·
i —\
'
11 ^*^ - ^ K e r n i
Cu
iPPm> jn tMI

1
cf^-o * < 68
li \Λ
\
· 68-106
· 107-168
3 0 km
' Φ >168

D Area of regional geochemical survey

Fig. 11-30. The distribution of copper. Sampling density is one sample per 30 km 2 , fraction
<0.06 mm. The Peräpohjola schist area is bordered with a dashed line.

The till cover consists of two beds of different age, with a boulder pave-
ment between. During the deposition of the older till bed the continental
ice flowed from the northwest, and during the deposition of the younger till
bed from the west (Korpela, 1969; Koivisto and Manner, 1981). Within the
study area the older till bed occurs only in deeper depressions and on the
distal sides of bedrock hills. According to fabric analyses of till stones, the
transport direction was 240°-270° (Koivisto and Manner, 1981).
Sampling was carried out stage by stage using a progressively denser
network. The sampling material at every stage was till. The sampling and
analysis of the reconnaissance stage were done as part of the Nordkalott
project. Till samples were then taken by spade from a depth of 0.5 m at
a sampling density at about 1/30 km 2 . Till samples of the reconnaissance
stage were not analyzed for Au. The Cu content map in Fig. 11-30 has been
simplified from the data of the Nordkalott project.
At the regional geochemical mapping stage the samples were taken with
a percussion drill at intervals of two kilometres. The sampling density was
thus 0.25/km2. At each site the samples were taken at depth intervals of 1
m to the greatest possible depth.
The detailed survey consisted of two different stages and included
lithogeochemical sampling with a lightweight compressed air drill. In the
306 Geochemical exploration examples

first stage (detailed geochemical survey 1) samples were taken with a


percussion drill from 69 sampling sites in an area of 1.4 km 2 selected
on the basis of results of the regional survey. The samples were taken
at a sampling point interval of 100 m along lines 100-200 m apart. Two
samples were taken from each sampling site, the higher from a depth of
2 m metres from the surface of the mineral ground and the lower from
the greatest possible depth. The average sampling depths were 2.4 and
4.2 m respectively. In the second stage (detailed geochemical survey 2) till
samples were taken from the most interesting area from the point of view
of the study. The samples were obtained with a percussion drill using a
regular sampling site interval of 10 m (300 m x 300 m area), one sample
from each site (961 samples). At each site the sample was taken from the
greatest possible depth, near the surface of the bedrock.
Between these last two stages of the detailed study, rock samples for
geochemical and lithogeochemical studies were taken with a light-weight
compressed air drill from 34 sites. The length of the sampling line was 480
m and the site interval 10-20 m.
Till and weathered bedrock samples were sieved and the <0.06 mm
fraction was analyzed. Weathered and unweathered bedrock samples were
also analyzed as pulverized total samples. The samples of the first two
stages of the study were analyzed with an optical emission spectrometer
(OES), the samples of the detailed survey stage with a flame AAS after a 7
M H N 0 3 leach. Au was determined after H N 0 3 + HCl leach by HGA 500
graphite furnace method (Kontas, 1981) in till samples of the regional stage
combined site by site and in all samples of the detailed surveys 1 and 2.
Cu (maximum) and Au concentration maps compiled from the results of
regional stage of the study are presented in Figs. ll-31a and b.
Whereas the geochemical results of the reconnaissance survey (Fig. 11-
30) indicate large geological units (in this case the Peräpohjola schist area),
the results of the regional survey (Fig. 11-31) indicate smaller units: in the
western and southern parts of the area mainly mafic volcanites and in the
central and eastern parts differentiated albite diabases and the associated
sulphide mineralizations (cf. Fig. 11-28).
The threshold of the Au anomaly at the regional stage was set at 4 ppb.
On this basis the most distinct and homogeneous anomaly is situated in the
southern part of the area, northwards from the village of Koivu (Fig. 11-
31). Other anomalous Au contents are encountered eastward from Metsola
where there are many albite diabase occurrences associated with quartzites.
In addition to Metsola, anomalous Au is associated with anomalous Cu only
at Koivu.
In the first stage of the detailed geochemical survey a Cu anomalous zone
was delineated in the vicinity of the regional anomaly (Fig. 11-32). The
anomalous pattern obtained from samples taken at higher levels (average
depth 2.4 m) differed a little from that obtained from samples taken at
Metsola: discovery of a Cu-Au mineralization by till geochemical study 307

AREA OF REGIONAL GEOCHEMICAL SURVEY

Cu (ppm) in till
< 85
• 85-120
• 121-181
• 182-283
φ 284-462
A 5 0 0 >462

Area of detailed
geochemical survey 1

10 km

-7340 « ■-

Au (ppb) in till
< 4
• 4-7
• 8-15
φ 16-31
^A >31

h- 1
1 0 km

Fig. 11-31. The distribution of copper and gold. Sampling density is one sample per 4 km 2 .
308 Geochemical exploration examples

AREA OF DETAILED GEOCHEMICAL SURVEY 1 AREA OF DETAILED GEOCHEMICAL SURVEY 1

Cu (ppm) in till Cu (ppm) in till


deep samples low samples
<49
< 49 \ ·
• 49- 99
• 100-219
:; · •

4 9 - 99
100-219
Α I · >219
• >219
Γ^»Β
Cä? 500 m
• ·
Dictions of ^ύήΐβξ> . Directions of
younger)
ice flows \o\ ice flows \o\
r

I I A r e a of d e t a i l e d g e o c h e m i c a l s u r v e y 2 A r e a of d e t a i l e d g e o c h e m i c a l s u r v e y 2

Pneumatic drilling profile Pneumatic drilling profile

Au (ppb) in till Au ( p p b ) in t i l l

f
< 4 < 4
4- 7 . ' · ■ · ' 4· - 7

••
• 8-15 8-15

:T3 S,B II
• 16-31 16-31
• >31 • >31

Fig. 11-32. Copper and gold concentrations of the fine fraction of till in deep samples (left) and
shallow samples (right).

lower levels (average depth 4.2 m). From the difference it is inferred that
the last transport direction of the continental ice sheet in the study area
was from west to east and that the transport distance of the fine fraction
of till was at least 200 m. Cu contents of the lowest samples indicate a
mineralized zone that seems to continue northeastward.
Although more dispersed, the anomalous pattern of Au is not unlike
that of Cu. The highest Au content of the lower samples occurs in the
Cu anomaly zone. The anomaly is stronger (highest content 2100 ppb) in
upper level samples. The dispersion of Au seems to have occurred over a
shorter distance and the direction is more towards the southeast, than that
of obviously more homogeneous and reliable Cu in till.
The anomaly patterns discovered at the second stage of the survey
(detailed geochemical survey 2) directly point out a Cu-Au mineralization in
the local bedrock, since the sampling network is very dense and samples are
taken from near the bedrock surface (Fig. 11-33). Although the anomalies
in till extend beyond the eastern margin, the mineralization apparently
lies within the sampling area. The results of the lithogeochemical sampling
Metsola: discovery of a Cu-Au mineralization by till geochemical study 309

AREA OF DETAILED GEOCHEMICAL SURVEY 2

. . . . . . . · Al

. . . . . . 0 A.
. . · · ·· ·
• ······· . . ^
T . . .· . . . ·. . . Ä . .
*
. . .·«·.·
· · · · · · · · · ' :: *m.' . . . . •
ν:*::;β?
»··· · · · · · · w ·:::::■··■. 4 %
··· · • · · · · · ·· ·
···
Λ · . «Ä· ··
w *w ^ · ·

- 7355.84 · ' · · -
Cu (ppm) at the bedrock-till interface Au (ppb) at the bedrock-till interface
<150 · 500- 999
< 8 • 3 1 - 60
• 150-299 · 1000-1999 ,
50 m 8-15 • 61-110
• 300-499 φ >1999 >110 50 m
16-30

Fig. 11-33. The distribution of copper and gold at the bedrock/till interface.

PNEUMATIC DRILLING PROFILE A-B

Cu ppm Au ppb
1100
500

Fig. 11-34. Copper and gold concentrations of the samples from the drilling line A-B (see Fig.
11-32) over the mineralized zone.
310 Geochemical exploration examples

with a compressed air drill provide evidence of this (Fig. 11-34). Figures
11-32 and 11-33 show that the sampling line of the compressed air drill
holes does not run over the strongest part of the anomaly but about 50
m to the north. The results nevertheless confirm that the most strongly
mineralized zone lies in the middle part of the albite diabase. As Fig. 11-33
shows, there are again clear similarities in the anomaly patterns of Cu and
Au, although the gold anomaly seems to be more widely dispersed and less
coherent.

PALLASTUNTURI: MULTIMETAL (Ni, Cr, Mg, Co, Cu, Zn) INTERPRETATION


OF REGIONAL GEOCHEMICAL MAPPING DATA USING STEP-BY-STEP
PROCEEDING STUDY IN HUMID HILLY REGION

Erkki Ilvonen and Matti Äyräs

The study area lies in northern Finland east of Pallastunturi at the


boundary between the parishes of Kittilä and Enontekiö. The elevation of
the area varies from 230 to 400 m above sea level. The middle and western
parts are strikingly characterized by several mountains and hills between
which lie small areas of bog. The eastern part is characterized by low-lying
bog.
In the eastern part of the study area (Fig. 11-35), the prevailing rocks
are gneissose granite and micaschists, which are interpreted as belonging
to the basement. In the western and northern parts, granite of Hetta
type prevails. Lying between these rocks, in the middle part of the area,
are sedimentogenous rocks — quartzites and hornblende biotite-gneisses
— and volcanic rocks — amphibolites and amphibole-chlorite-schists.
Intrusive rocks are found between the volcanics and sediments and as
larger intrusions in the eastern part of the area. In the middle part they
are serpentinites and in the eastern part gabbros and diorites.
Nickel mineralizations carried by ultramafites are being investigated
at Mertavaara (1), Lutsokuru (2) and Iso-Siettelöjoki (3). The vicinity of
Hotinvaara (4) is the present object of Ni prospecting.
The drift in the area varies in thickness from 1 to 10 m. Till covers
the bedrock in the whole area, with a thickness generally more than 3 m.
Overlying the till in the western part is an extensive and thick layer (1-8
m) of sorted sediments. The surface till unit is a grey or greyish brown
lamellar sandy till with occasional small stones (Fig. 11-36). Underlying
the surface till is a grey or brownish grey sandy till. In some places a
thin sorted layer marks the boundary. The lower till unit is tighter and
more stony and the stones are clearly bigger than in the surface till unit.
Characteristic of the structure is the occurrence of small lenses of sorted
material. Both till units were obviously deposited by the same glaciation.
Locally in the eastern part, the stratigraphy of the drift exceptionally is
Pallastunturi: multimetal interpretation of regional geochemical mapping data 311

1
Pallastunturi*.*.

I I Basement «
»)))))))] Sedimentary r o c k s '
P ^ H Volcanic rocks \
Intrusive r o c k s
» I I Granite
_ ■ Geophysical
^ ™ conductive zone

Fig. 11-35. A simplified bedrock map of the Pallastunturi area, northern Finland. Nickel
mineralizations are marked 1-4.

o o
o Grey or greyish
(i) ~ ^^ b - brown lamellar
o „ « Cc sandy till
j o>V^
Q> ^ ^. r \
£>
o
° -A
I 0
o Grey or brownish
o grey sandy till
(II)
o> o.
Λ
^ o
o c ^
X X X XX Bedrock

5-
(m)

Fig. 11-36. Stratigraphy of drift.


312 Geochemical exploration examples

Direction of
ice flow
(measured in till)
Glacial striae

Fig. 11-37. Directions of ice flow.

from top to bottom two till units, a sorted sediment layer and a third till
unit.
The orientation of flow of the surface till unit is 160-200° and that of
the other till 170-200° (Fig. 11-37). The general direction of ice flow was
from the south, the orientation of most of the observed glacial striae being
170-180°. No study was made of the transport distance, but observations
in other places in northern Finland have shown the fine matter of till to be
very local.
The sampling lines of the regional investigation were perpendicular to
the strike of country rocks and to the direction of ice flow; the lines were
2 km apart and the distance between sampling points along the lines was
1 km. The samples — till, sorted sediments and weathered bedrock if
available — were taken by light percussion drill. The sampling interval in
vertical direction was 1 m to the depth attainable. In the local investigation
the lines were 200 m apart and the distance between sampling points was
50 m; one sample, from the basal till, was taken at each sampling point.
After drying, the till samples were sieved into three fractions: fine <0.06
mm, middle 0.06-0.25 mm and stone >0.25 mm. The stone fraction of
the deepest till samples and/or weathered bedrock was washed and the
lithologies of rocks were identified.
The fine fraction of till was analyzed with an optical emission quantome-
ter (OEQ: Mg, Cr, Ni, Cu, Co, Zn) and an atomic absorption spectropho-
tometer (AAS: Ni, Cu, Co, Zn). For the AAS analysis, samples were leached
with nitric acid and the atomizing flame was air-acetylene. In the local
investigation, chromium was analyzed by XRF and Ni, Cu, Co and Zn by
AAS.
£

o
o

<*}

§
o
55-

I
I'
a.

CO
I-1
CO

Fig. 11-38. Concentrations of magnesium in the fines of till. The shaded areas are anomalous.
314
Geochemical exploration examples

Fig. 11-39. Concentrations of chromium. The shaded areas are anomalous.


1240^\ + £
A + ΔΔ + ΔΔ Δ + + + + D (
+ + D + + + + * ^ ^
2760 /WV$V^\
+ Da +Δ + « D ^^^
Δ Δ Δ Δ ϋ + Δ * + + +

+ Δ + Δ Δ + Δ H+fr 3

Δ Δ + Δ 4^ + + + D 3
+
Ι' ^
φ + Δ + Δ Δ + + +
•3
1
+ Δ + Δ Δ + Δ Δ +++ +D
Ο
ο

δ'

<^>
ο
ο
55-

+ + + Δ Δ A 3
ppm ε.
Α < 5 0 3
+ Δ + Δ Δ + Δ Δ Δ + (
+ 50-205 ν^. + + + + + I
D 205-280 ΔΔ Δ Δ + ++ + + +
ϋ 280-305
I'
Δ Δ 0
■ 305-615 Δ Δ + Δ + + + + + + + + ? + + + + + + ++++
677Β >615 Δ
5 km Δ + Δ + Δ ϋ Δ + Δ + + + + + + + + D D +^0) + + + + + + + + D + CO

en

Fig. 11-40. Concentrations of nickel. The shaded areas are anomalous.


316 Geochemical exploration examples

The results of the regional investigation show that, even with a sparse
sampling network like this (1 sample/2 km 2 ), the ore critical zone, in which
the known mineralization are situated (Fig. 11-35), clearly shows up as
metal anomalies in the till. The area of volcanic rocks in the eastern part
of the study area is also clearly reflected in the till and one can reasonably
assume that there may be mineralization there, too. The metal anomalies
in till in the study area — Mg, Cr and Ni together — are good indicators for
the mafites and ultramafites carrying nickel mineralizations (Figs. 11-38,
11-39 and 11-40).
The local geochemical investigations of Ni and Cr in till provided an es-
sentially more accurate picture than the regional investigations. Due to the
intense alteration of country rocks, the pattern for samples of basal till is
also a good alternative sampling material. The investigations were directed
to prospecting for Ni mineralization, indications of which were sought
not only by till geochemistry but also by stream sediment investigations.
Several mineralized zones were indicated also by geochemical anomalies in
stream sediments and by outcrops in rivers.
According to the anomaly patterns, the occurrence of ultramafites is no-
ticeably sporadic. In the area with few outcrops the Cr and Ni anomalies in
till suggested the locations of several different serpentinites (Fig. 11-41) and

JkuUk
A A

,Λ,,Λ,,Ι,,Μ
a_uur
U m T i A I
* i

Ni + Cr > lOOOppm

I ▲ = lOOOppm

A A 200m

Fig. 11-41. Chromium and nickel anomalies.


Pallastunturi: multimetal interpretation of regional geochemical mapping data 317

Fig. 11-42. Ni/Cr ratios. The shaded area: the ore-critical serpentinites in this area.

the real occurrence of most of these was later verified by diamond drilling.
Ore-critical serpentinites were differentiated from other serpentinites
on the basis of their high Ni/Cr ratio (Fig. 11-42, shaded area). High Ni
concentrations also originate outside of the ultramafic rock formations.
However, this Ni-anomaly group can be differentiated by means of the Ni/
Cu + Zn + Co ratio, so that only the highest values of the ratio indicate Ni-
critical serpentinites (Fig. 11-43). The high Ni concentrations originating
outside of ultramafites are due to the Ni-rich schists, which are not really
Ni critical.
The group of Ni-critical serpentinites located in the above-described man-
ner was investigated in detail by diamond drilling and the interpretation
was shown to be correct. Pyrrhotite and pentlandite were present in small
amount in many of the drilled serpentinites. There may be economically
interesting ores in the area. The maps should be equipped with fixed points
to allow the reader to make comparisons/interpretations, now it is difficult.

Cr

20
1C
/ 90
/ X \eo
3o/
\70
40/X \ 60
i50
\40
X
6 0 X
/*Χχ><
\ 30
*oJ$g^ X
\20
90/^^x7 X
A10
Ni ^ - * — * — * — * ■ J2L_ l i 1£_
—*—*—^ Cu + Z n + Co
10 20 30 40 50 60 70 80 90

Fig. 11-43. Ni/Cu + Zn + Co ratios. The shaded areas: Ni-critical serpentinites in the area.
318 Geochemical exploration examples

SEINÄJOKI: Au IN TILL IN AN Sb ZONE WITHIN ARCHAEAN SUPRACRUSTIC


ROCKS IN HUMID PLAIN TERRAIN

Esko Kontas

The study area is located in western Finland near the town of Seinäjoki.
The topography is gently sloping. In the rather flat area there are some
elevations of 10 m.
The study area is situated in the centre of a wide schist area where
the main rocks are biotite-plagioclase-gneiss, pegmatites and pegmatite-
granites (Neuvonen, 1961). Plagioclase-porphyrites are found together with
more acidic volcanites. Many mineralized occurrences have been found
in the quartz-bearing fractured contact zone of the volcanites, the most
common ore minerals being metallic Sb, Sb 2 S 3 and FeSb2S2, and less
common ones FeS, FeAsS, FeAs2, FeSbS and NiSb (Pääkkönen, 1966; Aho,
1980; Oivanen, 1982). Mineralization contains small amounts of gold: in
drill core analyses the contents generally range between 0.1 and 1.0 ppm
and at maximum reach 2-4 ppm. In the Sb-rich zones of Kalliosalo (Fig.
11-44) gold occurs at least partly as aurostibite, and in the arsenic-bearing
zones of Marttalanniemi as very fine-grained metallic gold (Aho, 1980).
The basal till cover is unevenly distributed. In hill areas it covers the
bedrock only thinly, but becomes thicker in depressions of the basement
(bedrock) and in drumlin-like tails on the distal side of the bedrock hills
(Hirvas, 1980a). The till contains a notable amount of preglacial sand
(Kauranne, 1960b). Regional mapping of surficial deposits (Mölder and
Salmi, 1954) did not reveal any preglacial weatherings in the Seinäjoki
area. However, weathered bedrock as much as 1 m thick was encountered
at Kallionsalo during till geochemical study.
The latest flow of the ice sheet in the area was from 350-360° as indicated
by the orientation of drumlins, striae and moraine. Also the basal till, and
with it ore boulders, was transported mainly by this flow from the north
(Hirvas, 1980a). According to stone counts carried out near the study area,
90% of the boulders are local rocks and the rest transported from further
away (Pääkkönen, 1966).
Samples were taken with a percussion drill at an interval of 10 m along
lines 100 m apart. At each sample site about 200 g material was taken at a
vertical interval of 1 m and, where possible, the deepest sample was taken
right at the surface of the bedrock (till/bedrock interface). The thickness
of overburden varied between 0.5 and 10 m. The two deepest samples
regardless of the material were analyzed for gold. From the samples
analyzed, 470 were classified in the field as till and 250 as weathered
bedrock or as mixed samples of those materials.
After drying, the <0.05 mm fraction was sieved from the till samples for
analysis, and the weathered bedrock/till samples were ground whole.
Seinäjoki: Au in till in an Sb zone within Archaean supracrustic rocks 319

125m

Fig. 11-44. Gold concentrations in the fine fraction of glacial till (numbers left of the line)
and weathered bedrock (numbers right of the line) in the Kallionsalo-Marttalaniemi target,
Seinäjoki. The curve delimits the Sb anomaly in weathered bedrock.

Gold was digested with HC1—H2O2 and separated from the matrix by
reductive coprecipitation using SnCl2 as reducing agent and Hg as collector.
The Hg-Au precipitate was dissolved in HC1-H 2 0 2 . The measurement
was carried out by atomic absorption spectrometry, using a graphite
furnace atomizer, a deuterium-background correction unit and a plotter.
The detection limit was 1-2 ppb (Kontas, 1981).
Because even very low gold concentration in pedogeochemical sample
materials can be significant (Brown and Hilchey, 1975), the lower limit
for a gold anomaly was set at 4 ppb for both the fine fraction of till and
weathered bedrock till. The following justification can be given for this
anomaly threshold:
(1) The average total gold content of crustal rocks is 3.5 ppb (Li and Yio,
1966).
(2) In analytical measurements the signal of 4 ppb is distinct and
320 Geochemical exploration examples

unambiguous, so that the results of 4 ppb or higher can be considered


reliable.
With this threshold, 19% of the results for the fine fraction of till and
29% of the results for weathered bedrock till were anomalous. This fairly
large proportion is natural as the samples were collected from an area of
Sb-As-Au mineralization and at or near the surface of the bedrock.
Because the dispersion of values is wide the anomalous gold contents
are presented on the map without classification (Fig. 11-44). The variation
is due both to the uneven occurrence of gold in the bedrock and by its
tendency to occur as small metallic pieces (nuggets) in the soil.
The high gold contents do not form coherent anomalies extended in the
direction of transport (clastic glacigenic dispersion). This is because the
bottom part of the till cover, as also the weathered bedrock, reflect the
unevenly distributed gold contents of the underlying bedrock. Likewise the
orientation of the antimony anomaly does not follow the flow direction of the
ice sheet but rather the orientation of the bedrock and the mineralization.
In the area of Kalliosalo-Marttalanniemi the fairly low contents of the gold
mineralization are reflected as anomalous contents in the fine fraction of till
and in weathered bedrock. In the former the contents range between 4 and
200 ppb and in the latter between 4 and 1300 ppm. This indicates that the
till geochemistry is suitable for gold exploration and that the fine fraction
of the deeper till is a useful sample material, even if the unusually high
amount of preglacial sorted sediments in the surficial parts of till somewhat
hampers its use. The gold contents in the preglacial weathered rock are
higher than in till why the results of analysis of these two materials should
be presented in separate maps.
Because of the irregular mode of occurrence of gold in any material,
anomalies tend to be difficult to interpret in detail. On the other hand, the
strong contrast between the contents of gold mineralization and the typical
low natural background values allows broad targets to be identified. Gold
is often associated with antimony- and arsenic-bearing zones. Gold critical
areas can therefore be delineated by using these as indicator elements. Why
were then their distributions in mineralized materials not presented?

TOHMAJÄRVI: Li CONCENTRATIONS AND Li MINERALS OF TILL ACROSS


PROTEROZOIC PEGMATITIC Li DEPOSITS IN HUMID PLAIN TERRAIN
Maria Nikkarinen

Earlier geochemical mapping in the Tohmajärvi area revealed anomalous


Li contents in till. High Li contents were found near known Li pegmatites,
but the pegmatites were not necessarily the only source of the anomalous
values. The objective of the investigation described here was to determine
the cause of these high Li contents. New samples were collected in two
Tohmajärvi: Li concentrations and Li minerals of till across Li deposits 321

target areas, Kuukkeli and Oriselkä, on the basis of both Li intensities and
a microscopic study of the earlier till samples.
The Tohmajärvi area is situated in Karelia in eastern Finland, about 60
km southeast of Joensuu. The two target areas lie about 10 km southwest
of the centre of Tohmajärvi. At Oriselkä the terrain is characterized by low
hills and bogs, while at Kuukkeli the relief is gently sloping from north to
south, with a variation in elevation of about 15 m. Both areas lie about 120
m above sea level.
The target area is situated within the Karelian schist belt. In the
Kuukkeli target area the bedrock consists of mica schist and phyllite,
while at Oriselkä amphibolite, black schist and pegmatites are encountered
as well (Fig. 11-45). The pegmatites are thought to be related to the
large granite massive lying south of the target areas (Nykänen, 1968).

I I mica schist

amfibolite

|^| black schist

| = | quartzite

χ known lithium-
pegmatite outcrops

g> known spodumene


bearing pegmatite
lice sampling sites of
movement / the soil samples
c A directions

Oriselkä

Fig. 11-45. Bedrock of the target area and the sampling sites of the soil samples.
322 Geochemical exploration examples

Approximately 80 complex pegmatites have been found in an area of 150


km 2 . About 20 of these are lithium pegmatites, where the Li is most
commonly hosted by triphylite or its alteration products. Spodumene,
montebrasite and lepidolite occur in three deposits (Alviola, 1974). None of
the known complex pegmatites is of economic size.
In both target areas the bedrock is covered by a uniform till bed, thickest
in the southern part of the Kuukkeli area where it reaches 10 m. The
average depth of the till cover is 3 m at Kuukkeli and 2 m at Oriselkä.
Geochemical mapping has demonstrated two distinct ice movements in
the Tohmajärvi area, one 280°-290° and the other 310°-340° in direction.
The more northerly movement is considered older and more important
(Hartikainen and Salminen, 1982). However, most of the samples collected
in the target area were from the younger till blanket. The till blanket
produced by the older ice flow is encountered only in depressions in the
Kuukkeli area.
Erosion and transportation by glacier have been weak, at least in the
Oriselkä area, where weathered bedrock has been found in some sample
pits and the forms of the bedrock bear a distinctive morphology. After
studying the metal concentrations in the finest fraction of till in Suksijoki,
an area near to Oriselkä, Hartikainen and Salminen (1982) concluded that
the glacier moved the drift only 300-500 m.
Till samples were collected along lines across the presumed main direc-
tion of ice movement (Fig. 11-45). The distance between sampling lines was
200 m at Kuukkeli and 100 m at Oriselkä; sampling sites were spaced at
50 m intervals. Samples were taken by percussion drill as deep in the till
as possible, so that the depth of samples varied between 10 and 0.6 m.
At sites where weathered bedrock was encountered, that too was sampled.
Altogether 151 samples were collected.
The samples were dried (80°) and sieved into three fractions. Closest
attention was paid to the 0.06-0.5 mm fraction, a part of which was
analyzed for Li by AAS after dissolution in HF + H3BO3. Part of the
fraction was also extracted with bromoform (d = 2.89 g/cm3) to assist
microscopic study. From one sample the enriched mica fraction (mainly
biotite) was separated and analyzed for Li. In addition some pegmatite and
mica schist fragments from the 2-25 mm fraction were analyzed for Li.
A polarizing microscope and a heating method were used to identify the
Li minerals, (Nikkarinen and Björklund, 1976). After 15 minutes heating
at 1000 °C spodumene microfractured and the grains contained droplike
inclusions — always a good distinguishing mark of spodumene (Hosking,
1957). Montebrasite (amblygonite) fuses upon heating, and the occurrence
of a fused white pliable mass was considered a sign of montebrasite. The
fine fraction of samples with Li contents over 110 ppm were investigated
for lepidolite. Lilac and grey mica were sought in the fine fraction but
no special method for the identification of lepidolite was used. The coarse
Tohmajärvi: Li concentrations and Li minerals of till across Li deposits 323

fraction (over 2 mm) was visually investigated and the number of pegmatite
fragments counted.
Anomalous Li contents in till (0.06-0.5 mm fraction) are most abundant
in the southern part of the Oriselkä area (Fig. ll-46a). Some high Li
contents were also found in the Kuukkeli area, where pegmatites have
not previously been encountered. The highest Li contents were found in
weathered bedrock samples (Fig. ll-46b) whose rock type is mica schist in
the area.
In the mineralogical investigation (0.06-0.5 mm fraction) spodumene
was found in a few samples, but in each case the number of grains was
less than 5. Montebrasite was observed in three samples from the Oriselkä
area, but only 1-2 grains in each (Fig. 11-47). Lepidolite was not verified
in any sample. Instead the samples contained abundant biotite amounting
to 70-90% of the heavy fraction (d > 2.89 g/cm3). A concentrated biotite
fraction from one sample contained 1020 ppm Li, which was almost 4 times
the content of the total 0.06-0.5 mm fraction.
Visual observation of the 2-25 mm fraction of till revealed pegmatite
fragments in samples from the southern part of Oriselkä (Fig. 11-47).
However, no Li mineral was observed in these fragments. The Li contents
of the pegmatite fragments and of selected fresh and weathered mica schist
fragments are collected in Table 11-VIII.
In all likelihood the strongest uniform Li anomaly in the southern part
of Oriselkä does not originate from the known Li pegmatites; for the
nearest known pegmatite does not lie in the direction of prominent ice
movement, and the transport distance from the other known pegmatite is
twice the prevailing transport distance in the area. Evidently the anomaly
originates from an unknown local source. But there is little hint of
economic Li minerals. The samples of highest Li concentration did not
contain spodumene, montebrasite or lepidolite. More likely it is triphylite,
the main host of Li in the known Li pegmatites, which is responsible for
the high Li contents in till. Triphylite was not observed microscopically,
but it is easily weathered and, in till, Li originally fixed in triphylite would
likely be fixed in minerals like micas. High Li contents of one separated
mica fraction indicated the importance of mica as a secondary host of Li.
The enrichment of Li in micas and other clay minerals in the weathering

TABLE 11-VIII
Li contents (ppm) of rock fragments in till (2-25 mm fraction)

Pegmatite Mica schist Weathered mica schist


Range 4-51 74-140 170-340
Mean 25 113 317
Number of samples 8 3 3
324 Geochemical exploration examples

16
83
70

55
RP 33
·«« 61
. 180 &5 0RP
$0 67RP 6^.RP
• 110RP b
200 6
^ 10
U
< ^ - > ^ 16
26 7
16RP £.1

31
67 RP

Oriselkä ·
3«6110RP
• 68RP ^ <*6
3*5 69RP ^ £β g2
β 20 . 35 g8
17 2
. A.200RP ^
67
. -ΛJ 37
Li-pegmatite *,o . 44

Λ . 1 "°
5
28
" ·,13
Ό 20 1 °RP ,-Q .
63
• 2% . 37
39
öiRP - 572 5 iu
UU
. Jg 88 RP 36
73
. .RlfcS 33 35
10*0 28 £2 iu
70
& . 33 .
80 67
& . 53§1RP β100
?

βλ^Λο37 δ3
9*1 *6<* 50
. & 75 δ3
& v,380RP .
57
.^ÖRP . & ,
1 81
1Λ 3°7° · 270 . &
120 100 60
71 3*2 ξθ
JU
#1 3 0 61 « <0
52 6 1
•r · $
U8 Φ *° 6fifton"»
890
55 66^° 66 I

Fig. 11-46. Li contents (ppm) in till samples and in weathered bedrock samples (RP).
Visasaari and Vuonelonoja: use of magnetic susceptibility data 325

Oriselkä

Fig. 11-47. The observed Li minerals in the 0.06-0.5 mm fraction and pegmatite fragments in
the 2-25 mm fraction. S = the sample containing 1-5 grains spodumene; M = the sample
containing 1-2 grains montabrasite; ruled area: zone of the samples containing over 70%
granite or pegmatite fragments in the 2-20 mm fraction.

process has been noted in many investigations (McLaughlin, 1955; Mitchell,


1964b; Billings, 1969). Also in this study, maximum Li contents (490-380
ppm) were found in weathered bedrock samples.
There was no indication of Li pegmatites in the Kuukkeli target area;
samples did not contain any Li minerals or pegmatite fragments, despite
the scattered occurrence of anomalous Li contents.
Although indications of economic Li minerals were few, the results
strongly suggested that there are unknown pegmatites at least in the
Oriselkä area. Thus the fine fraction of the samples should have been
analyzed for other valuable elements like Sn, Nb, Ta, Th, Be and Ce, which
pegmatites are likely to contain.

VISASAARI AND VUONELONOJA: USE OF MAGNETIC SUSCEPTIBILITY DATA


IN INTERPRETATION OF GEOCHEMICAL ANOMALIES IN TILL IN HUMID
PLAIN TERRAIN

From: Pulkkinen, E., Puranen, R. and Lehmuspelto, R, 1980. Interpretation of geochem-


ical anomalies in glacial drift of Finnish Lapland with the aid of magnetic suscepti-
bility data. Geol. Surv. Finl., Rep. Inv., 47, 30pp.

The study areas are located in central (Visasaari) and in eastern


(Vuonelonoja) Finnish Lapland, 130 km north of the Arctic Circle in
supra-aquatic low-lying bog districts with moraine hillocks. In the Visas-
aari area the magnetite-rich source of the material studied is a hillock and
in the Vuonelonoja area a depression.
326 Geochemical exploration examples

The bedrock in the Visasaari area is composed of talc-chlorite schist


containing disseminated magnetite. The magnetite has been enriched in
a small ore deposit on the top of the hill in the centre of the area.
In Vuonelonoja the magnetic anomaly is caused by a plug of peridotite,
which contains magnetite, chromite and sulphides of iron and nickel as
accessories. The anomalous plug of peridotite is surrounded by weakly
magnetic gneiss granites.
Samples from the glacial drift were collected over a 200 x 200 m grid
in Visasaari and over a 100 x 500 m grid in Vuonelonoja. Only till and
weathered bedrock samples were accepted. The susceptibilities of <0.06
mm and 0.06-0.5 mm fractions were measured with the low-field ac-bridge
(Puranen and Puranen, 1977). The magnetite contents were measured with
a saturation magnetization analyzer. Total Mg, Ca, Cr, Cu, Fe and Mn
concentrations were determined by an optical emission spectrometer.
Glacial erosion and transport of material from magnetic source rocks
gave rise to magnetic anomalies in drift downglacier from the sources.
These anomalies show patterns typical to glacigenic clastic dispersion;
magnetite is resistant to postglacial weathering and no chemical dispersion
has occurred. The main mode of formation of the geochemical anomalies
associated with the magnetic sources can be determined through com-
parison with the magnetic anomalies. The present intensities of source
concentrations and glacigenic drift anomalies should be in accordance with
each other if the source composition was not different (weathered) during
glacial erosion.
In the Visasaari area most of the drift samples taken represent the
topmost till layer, which was deposited by the latest glaciation flowing
towards SE in the area. The flow direction is indicated by the situation of
the drift anomaly of magnetite on the SE side of the source hill composed
of talc-chlorite schist (Fig. 11-48). The magnetite anomaly is consistent
with Mg, Ca and Cr verifying their glacigenic origin. Cu, Fe and Mn
anomalies display features of ground water transport. All the glacigenic
drift anomalies (Cu, Mn, Fe, Mg, Ca and Cr) are proportional to the source
anomalies in intensity, which suggests that the preglacial and present
compositions of the source are similar. The source hill was probably free of
preglacial weathering crust during the latest glaciation.
In the Vuonelonoja area the samples were collected from two till layers
deposited by different glacial phases. The younger phase flowed towards NE
and the older towards SE as indicated by two branches of glacigenic anoma-
lies (Fig. 11-49). This suggests that the composition of the Vuonelonoja
peridotite was changed by preglacial weathering, which is supported by the
systematic variations in the dispersion patterns of different elements.
The shift between source and drift anomalies is about 300 m in the
Visasaari area and only 50 m in the Vuonelonoja area, which demonstrates
the short distance of glacial transport in ice divide areas.
Visasaan and Vuonelonoja: use of magnetic susceptibility data 327

Kk (10" 8 m 3 /kg) Kp (10" 8 m 3 /kg) Mg % Ca %


• <156 • <140 • <2.4 • <1.0
0 0 156-383 O 140-238 O 2.4-3.4 # O 1.0-1.3
• # >383 • >238 • >3.4
o .-^ o O · >1.3
° /o V · /o\ o

• • · I* / o
/ • o l · ·
• L °· · • o —-^# A
w #
° o" · φ o o o o · # o
w
o ° o o o ·
• • •
• Cr ppm Cu ppm

Fe % Mn ppm
• <357 • <80 • <10.0 • <1228
# O 357-611 A O 80-89 A O 10.0-1 1.8 A 0 1228-1440
W
O · >89 • . · >11.8 • 0 · <1440
/ • V O
• / ° o
o • / #. ° . / ° o \ o . . /° · \ · ·
• /. * · ·
• • / i ^ · ° • Λ·^ ° * *
o °
• o · · • r#· o · • o ° '
• • o · ' 500 m W
0 ΙΊ
o o
Fig. 11-48. Dispersion of magnetic susceptibility in the 0.06-0.5 mm (Kk) and <0.06 mm (ÜCP)
fractions of glacial drift and dispersion of magnesium (Mg), calcium (Ca), chromium (Cr), copper
(Cu), iron (Fe) and manganese (Mn) in the <0.06 fraction of glacial drift in the Visasaari area.
Number of samples per sampling site varies from 2 to 10. The dashed line outlines the source
area.

Kk (I0" 8 m 3 /kg) Kp (10" 8 m 3 /kg) Fe %


• <18 • <31 • <4.7
O 18-71 O 31-62 O 4.7-6.3
• >71 # >62 • >6.3
••• ••
1 t •
I
o o 8
• •• o X o
% fa i • o o
• 0 o

1'
o o 8

1i * 1
o o
|
§
8
^ 8 o • : ■ 8 °

I
Mn ppm Mg % Cr ppm
• <646 • <1.6 • «413
O 646-877 O 1.6-2.1 O 413-693
• >877 • >2.1 • >693 II
1 §# if
o #

I •
I
o O
o O

i
o

! '* I
° /#* ° •• o
o T i o o
•• •
• o
o o
* *
o • o o o §
8 • 500 m o

Fig. 11-49. Dispersion of magnetic susceptibility in the 0.06-0.5 mm (2fk) and <0.06 mm (K?)
fractions of glacial drift and dispersion of iron (Fe), manganese (Mn), magnesium (Mg) and
chromium (Cr) in the <0.06 mm fraction of glacial drift in the Vuonelonoja area. Additional
explanations as in Fig. 11-48.

The methods presented in this study for interpretation of geochemical


anomalies in drift are mainly suited for basic research of anomalies
associated with magnetically delineated bedrock sources. The method could
328 Geochemical exploration examples

be used in model studies in different topographic conditions and various


provinces of glaciated terrain. Magnetic bedrock sources of different sizes
and shapes and various grain size fractions of drift should be selected for
such investigations. Comparison of magnetite and geochemical anomalies
in drift would provide information about the dominant modes of element
dispersion, hydromorphic or glacimorphic, under various conditions.

CAMP LAKE: HEAVY METALS IN SEDIMENTS OF A DRY PERMAFROST AREA

From: Miller, J.K., 1979. Geochemical dispersion over massive sulphides within the con-
tinuous permafrost zone, Bathurst Norsemines, Canada. In: Prospecting in Areas of
Glaciated Terrain, Dublin, 1979. Institution of Mining and Metallurgy, London, pp.
101-109.

The Camp Lake area is located in the Bathurst mine area in northern
Canada, 480 km northeast of Yellowknife. Flat and swampy lowlands and
gently rolling hills create a subdued topographical relief (20-50 m). The
mean annual temperature of — 12°C and average precipitation of 280 mm a
year are typical of cold desert. Vegetation is sparse: dwarf birch and willow,
grass, moss and lichens. Permafrost reaches down to 500 m, but the surface
is thawed to 2 m depth in summer.
Bedrock consists of Archaean metasediments, silty schists and tuffitic
metavolcanites. Underlying an assemblage of about 1500 m thick epiclastic
metaturbidites is a sequence of calcareous and argillaceous rhyolitic tuffs
with massive stratabound sulphide mineralization, and under these are a
set of andesitic and rhyolitic flows, tuffs, pyroclasts and breccias.
The glacigenic overburden consists of till of perhaps more than one
glaciation, eskers, outwash deposits, kames, drumlins and numerous of
erratics. The two transport directions are towards west-northwest and
southwest. Soil development has been disturbed by cryoturbation: frost
heave, boulder circles, solifluction, mud boils, etc. On the surface there is
a 1-5 cm thick humus layer (referred to as LFH in abbreviation of litter,
fermentation, humification) and under it poorly developed brunisol, regosol
or gleysol.
The purpose of the study was to determine the suitability of geochemical
methods for prospecting in a permafrost area, where permafrost, according
to earlier assumption, prevents hydromorphic transport of metals.
Samples were collected over a 120 x 30 m grid from humus-rich surface
soil and from mineral materials at 0-35 cm depth (layer 1) and 35-64 cm
depth (layer 2). Lake sediments were collected along the watershed. In
addition, a number of profile samples were taken from pits, 5 cm vertical
intervals down to the frost table level. Ground water seeping into pits,
snow meltwater running on the surface and lake water were collected as
well. Samples of humus and the -0.18 mm fraction of the different mineral
Camp Lake: heavy metals in sediments of a dry permafrost area 329

TABLE 11-IX

Parameters of partitioned log-normal Cu, Zn and Pb populations in soil layer 1

Geometric mean Proportion Number of samples


(ppm) (%)
CuA 64 90 244
CuB 10.5 10 37
ZnA 68 95.5 268
ZnB 22.5 4.5 13
Pb A 100 60 118
PbB 58 40 78

materials were decomposed by evaporation to dryness with a mixture (4:1)


of nitric and perchloric acids, leached with HC1 and analyzed by AAS. Water
samples were analyzed as such by AAS.
Heavy metal distribution curves consisted of two distinctly different pop-
ulations: background values (B) and the anomalous values (A) originating
from mineralization. Because of the two distinctly different populations a
standard anomaly threshold, mean concentration plus two deviations was
not reasonable (Table 11-IX). Calculation of mean concentrations for the
whole area, while partly meaningless for the same reason, was used for
comparison of different materials anyway.
The distribution of lead in surface soil (layer 1) and in till (layer 2,
probably C-horizon) is a clear example of glacigenic transport (Fig. 11-50A
and B). The till has been thrown in sheets towards the west, as reflected
in the layered heavy metal concentrations. Apparently the concentrations
have not been diluted by rapid diffusion, by mixing with far-transported
material.
The highest heavy metal concentrations in lake sediment occur close
to the northern end of Camp Lake, having been transported no doubt,
by seeping water and wave washing, from the anomalous till. This shows
that both the weathering and leaching associated with transport in water
solution can occur in permafrost conditions, apparently in the thawed
surficial part of the overburden.
The distribution of metals between different materials is presented in
Table 11-X. The materials can be compared with each other and with the
source rock, e.g., by calculating the elemental ratios, to see what happens
to the elements during weathering and transport.
There is a marked difference in the behaviour of the elements. The
original concentrations in mineralization are Cu 0.4%, Zn 7.5% and Pb
1.5%. The ratio of zinc to lead and especially to copper is decreased in soil
relative to the mineralization, but is strongly increased in lake sediments
along the direction of stream transport (Fig. 11-51). Similarly the ratio of
copper to lead is increased in the lake sediments. Lead seems to be more or
330 Geochemical exploration examples

Fig. 11-50. Lead (ppm) in the <177 μηι grain size fraction: (A) of surface soil, (B) at depth of
35-64 cm at Camp Lake, Canada.

less immobile, copper mobile and zinc highly mobile.


The processes involved remained somewhat obscure despite the study of
heavy mineral separates done to identify minerals binding lead in the over-
burden. Galena was not found. Evidently extensive chemical weathering
Camp Lake: heavy metals in sediments of a dry permafrost area 331

TABLE 11-X

Mean heavy metal concentrations and ratios of contents in different materials


Geometric means (ppb)
Cu Zn Pb
Soil layer 1 59,000 66,000 62,000
Lake sediment 20,000 51,000 13,000
Seepage water 265 460 12
Snow meltwater 207 296 108
Lake water 9 72 -
Concentration ratios
Cu/Pb Zn/Pb Zn/Cu
Mineralization 0.2 5.0 18.8
Soil layer 1 2.3 2.8 1.2
Lake sediment 6.0 8.0 1.3
Lake sediment downstream 8.1 13.0 1.7
Lake sediment further downstream 20.0 58.3 3.0

Zn Cu Pb
Z n , Cu and Pb in lake sediment p - ^ ^ 1998/1462/47

2000/1500/50 ppm
^ " " \ Β α η α η α ^ ν . 647/285/15 κ1
V
/y' -^Jake^i6r ΓΚ 1113/420/9
Zn and Cu in lake water
-^J WZn 18, \ \(Zn<7,
(70, 10) ppb
' \ Cu 2) 742/722/3C\ · I Cu <2)
Bat lake\
^xg) massive sulphides \A32m) | . \2321/653/36

sampling site • / 1161/1995/60

,498/685/182
6264/883/30^ ^£
g, 392/794/20
*
4 8 3 5 / 1 5 1 1 / 2 5 6 / C a mp
§ } 309/808/13
3396/2260/286" Λ χ· l a | < e *\\
642/1421 /302\v l\U m ^y 370/925/9
558/887/239\.
551/637/185^ N . 430/524/179
226/356/46 I *
> / \ ^ (Zn 65, Cu 8)
(Zn 72, CuS^ \~"^^419/241/31

Sunken ^ - ^ - — ^ \
lakesx^x/\i75/59/3
®
(Zn 2 β Ν θ \ v
Cu 2) XN,
(Zn 85, Cu<5)
| 0 1 km
^
Fig. 11-51. Copper, lead and zinc (ppm) in lake sediment and zinc and copper (ppb) in lake
water at Camp Lake, Canada.

had destroyed all the original sulphides in surface soil. Experiments show-
that lead is easily extracted from humus, e.g. by EDTA, but copper and
especially zinc are not.
332 Geochemical exploration examples

This excellent study shows that extensive clastic dispersion trains and
primarily glacigenic geochemical anomalies are formed in the Arctic. Later,
they may be reformed by seeping ground water, weathering and further
hydromorphic transport. During the transport from lake to lake, elements
in water solution have been gradually separated from one another, as seen
in the change of ratios of element concentrations.
The report recommends analysis of lake sediments for regional studies
and analysis of till for detailed exploration. A general conclusion of the
study was that the effects of permafrost on geochemical programmes
are minimal; hydromorphic and clastic dispersion patterns may be even
better developed than in the temperate zone. It was unfortunate that the
sampling of overburden could not be extended deeper than 2 m, the level of
permafrost.

GLOGFAWR: IODINE IN THICK OVERBURDEN B-HORIZON NEAR A


PALAEOZOIC Pb-Zn MINERALIZATION IN HUMID HILLY ENVIRONMENT

From: Andrews, M.J., Bibby, J.M., Fuge, R. and Johnson, C.C., 1984. The distribution of
iodine and chlorine in soils over lead-zinc mineralization, east of Glogfawr, Mid-Wales.
J. Geochem. Explor., 20: 19-32.

The Glogfawr study area is situated in Mid-Wales, 22 km inland from


the west coast. The area is 400 m above sea level, flat and poorly drained
in the northern half and characterized by gently rolling well-drained hills
in the south. Pines populate the forests, while grass and moss dominate in
the marshy lowland areas.
Bedrock consists of Silurian mud, grit and siltstones. Laminated siltstone
with occasional greywacke, fractured and brecciated by a NNE-trending
fault system dominates in the study area. A blind disseminated mineral-
ization with 1% Pb and some Zn and Cu in quartz- and ankerite-cemented
breccia was located by electrical induced polarization (IP) and electromag-
netic very low frequency (VLF) measurements and drilling.
Thickness of the tilly overburden varies from 3 to 5 m, the brown-earth
soil profile being only 20 cm on rocky till-covered ridges but from 50 to 150
cm on sandy well-drained slopes. The long axes of ridges parallel to the
fault zone suggest a glacial transport from NNE towards SSW though the
report does not mention this. Most of the low-lying area is covered by peat
bog. Thickness of the peat is not mentioned.
The purpose of the study was to see whether mobile elements like iodine
and chlorine could be used as pathfinders in locating ores covered by thick
variable overburden. Iodine has been successfully used for such purposes
in the Soviet Union (Krylova, 1977) and in China (Xie et al., 1981).
Approximately 1 kg of B-horizon mineral material was collected from
80 sites in a grid 100 x 50 m, where lines were laid out perpendicular to
Glogfawr: iodine in thick overburden B-horizon near a Pb-Zn mineralization 333

76 76

0 500 m 0 500 m

Fig. 11-52. The distribution of total and water soluble iodine (ppm) in the -120 mesh crushed
fraction of drift at Glogfawr, Wales.

the strike of the mineralized zone. The samples were dried, disaggregated
and sieved. The <200-mesh fraction was analyzed for total iodine from
the water solution of a sample fused with K 2 C0 3 , Na 2 C0 3 and MgO by
the automated catalytic/photometric method of Fuge and Andrews (1985).
Water-soluble iodine was determined after leaching of the sample with
distilled water, shaking and centrifuging. Also F, Cl, As, Pb, Zn, Cu, LOI
and pH were determined by different methods.
Both the total and the water-soluble iodine distribution curves show
distinct populations (Fig. 11-52). The arithmetical mean of background
population concentrations is 5.8 ppm for total I and 0.1 ppm for water
soluble I. Corresponding values for the anomaly populations are Itot 47 ppm
and I aq 1.7 ppm. About 30% of the samples belong to the log normally
distributed background population, the rest to the log normal "anomaly"
population.
Both iodine and chlorine anomalies pick out the known mineralization.
The strongest anomalies lie directly above the suboutcrop of the dissem-
inated zone, and also upon the possible continuation of it towards SSW.
Among the metals Zn seems to be correlated with pH and thus dependent
on drainage, but Pb and Cu pick out the area of known mineralization
(Fig. 11-53).
The highest concentration of total iodine (149 ppm) is found at the
same site as the highest concentration of total chlorine (520 ppm) and the
highest Cu and Pb concentrations. The water-soluble I and especially the
Cl anomalies are elongated towards the southeast. The authors suppose
334 Geochemical exploration examples

50 75 100 50 75 100
" ' i 1

0 500 m 0 500 m
Fig. 11-53. The distribution of lead and zinc (ppm) in the -120 mesh crushed fraction of drift
at Glogfawr, Wales.

that the extremely high I and Cl concentrations have been formed by the
enriching absorption of clay minerals and they do not attach importance to
the organic matter; even if there was a marked positive correlation both
between the total halogens and loss on ignition. The small concentrations
of halogens leached into water show that they are fairly strongly bound.
Anomalous iodine and chlorine concentrations were also found some 200
m east of the main anomaly zone. Glacial transport of drift cannot explain
these anomalies because it parallels the direction of the fracture zone
and the long axis of the main anomaly. Possibly there is another parallel
mineralized zone, as the VLF curves would suggest as well.
There is a strong and wide Pb anomaly and anomalous concentrations
of Cu, I and Cl as well to the west of the "main" anomaly. These were
not investigated more closely but considered to be "false" by the authors
because they had no geophysical equivalent.
The study showed iodine and chlorine to occur in soil in this area in
truly anomalous amounts. The average halogen concentrations in the ores
of Mid-Wales are only 138 ppm Cl and 2.5 ppm I in galena and 290 ppm Cl
and 1.5 ppm I in sphalerite. Maybe in this type of breccia there are iodine-
and chlorine-rich fluid inclusions both in ore and in gangue minerals.
In any case the halogens, as well as the ore metals themselves, formed
geochemical anomalies on the surface of drift above the known disseminated
mineralization 3-5 m below. It would seem that a simple method of
determining water-soluble halogens could be useful in soil geochemical
exploration for base metals.
Hiendelaencina: elements in a mixed sandy and in situ weathered overburden 335

The sampling lines should have been plotted on the maps. The method of
visualizing the results by classifying and generalizing the data into different
concentration classes makes for clarity but does not allow the reader to
draw his or her own conclusions.
It is unfortunate that the different sampling materials and their con-
centration distributions were not discriminated, because the reader cannot
really know now whether any of the conclusions are right — so great,
usually, are the differences between the heavy metal concentrations in till,
sorted sediments and peat. It is regrettable, too, that samples were not
taken vertically through the overburden at some sites to detect possible
movement of materials. Glogfawr is a fine example of experimenting in
geochemical exploration with the determination of exotic elements — some-
thing not often done, but which seems to have succeeded positively in this
case.

HIENDELAENCINA: As, Pb AND OTHER ELEMENTS IN A MIXED SANDY AND


IN SITU WEATHERED OVERBURDEN AT A SILVER VEIN IN PRECAMBRIAN
GNEISS OF WARM DRY MOUNTAIN PLATEAU

From: De Vos, W and Viaene, W, 1980. Geochemical study of soils and metallogenic
implications at Hiendelaencina, Guadalajara, Spain. Mineral. Deposita 15: 87-99.

Hiendelaencina plateau is situated about 100 km northeast of Madrid


at 1000-1100 m above sea level. The report does not say anything about
vegetation but it can be assumed to be sparse.
Bedrock consists of gneiss varying from porphyric (microcline crystals
up to 10 cm) to fine-grained micaceous, the metamorphic degree varying
from greenschist facies to almandine-amphibolite facies. The gneiss dome is
surrounded by Ordovician and Silurian sediments folded during Hercynian
orogeny. The E-W directed faults were filled during the same period with
hydrothermal solutions in 5 cycles. A zone of baryte, quartz and limonite
was mined at Filon Rico between 1844 and 1925. In the east, a 30 cm thick
baryte portion, rich in Zn and Ag, was mined at Nochebuena to a depth of
600 m. The quartz portion of the veins is sterile. The limonite portion in
the west was weathered and also once mined for silver halogenides.
The bedrock surface was weathered during the Miocene and is covered
with sorted material: gravel, sand and clay known as Rana. The rana is
usually less than 0.5 m thick and only occasionally up to 1 m thick; it forms
a gradual transition to the underlying gneiss. On the western and southern
slopes of the plateau the rana is mixed with in situ weathered gneiss as
a result of soil creep. No real soil profile has been developed on this dry
altiplano but the surficial layer contains a lot of organic material.
The purpose of the study was to investigate what kind of geochemical
response the argentinoferous and the sterile veins give in the overburden,
336 Geochemical exploration examples

and especially to see whether the newly discovered continuation of the vein
at a depth of 300-500 m would show up. The study was simultaneously a
follow-up stage to a regional stream sediment survey which had discovered,
besides anomalies caused by contamination from the old tailings, As
anomalies with no obvious interpretation.
Soil sampling was done along lines approximately perpendicular to the
vein and 100 m apart. The sampling interval was about 50 m and the depth
of sampling 5-20 cm beneath the organic rich topsoil. Samples were dried
and sieved, and the <0.18 mm fraction, after digestion with a mixture of
HC1 + H N 0 3 + HF and evaporation to dryness from a water solution, was
analyzed by AAS for Mn, Fe, Ni, Co, Cu, Zn, Pb, Ag and Ba. As and Sb were
determined by AAS with a hydride generator.
The results visualized on the maps were classified according to the
populations discovered from the frequency distribution curve. Filon Rico is
clearly indicated by a 50-100 m wide anomaly which varies longitudinally.
The distribution patterns of Pb, Zn and Cu resemble each other closely.
Highest Mn and Fe concentrations are found at the western end of Filon
Rico, above the limonitized vein. The distributions of Ag and Sb resemble
that of As: there are anomalies above Filon Rico, high anomalies around
old tailing heaps, and large anomalies some 400 m south of Filon Rico
(Fig. 11-54A and B). These last anomalies are perhaps better shown in the
nonclassified concentrations plotted on a N-S directed line (Fig. 11-55);
the abruptness of the Filon Rico anomaly and the width of the southern
anomaly show up particularly well.

" ' 1 I · · 1 1 1
0 100 200 m . o 100 200 m

Fig. 11-54. Distribution of (A) lead and (B) arsenic in drift (in Rana = a mixture of gravel, sand,
clay and weathered bedrock material) <0.18 mm fraction at Hiendelaencina, central Spain.
Hiendelaencina: elements in a mixed sandy and in situ weathered overburden 337

o Ί~ 500
Slope
1000 m
Filon Rico
Fig. 11-55. Distribution of lead, zinc, arsenic and silver in the <0.18 mm fraction of drift along
a line crossing the Filon Rico mineralized vein at Hiendelaencina, central Spain.

It seems that above Filon Rico the overburden is very local; only the
Ag anomaly is slightly shifted to the south. The northern border of
the southern anomaly is sharp indicating, as the authors state, a linear
structure, possibly a mineralized vein. Slightly decreasing concentrations
towards the south in anomaly indicate the width of the source in bedrock
but also suggest soil creeping along the slope. The transport of elements by
groundwater seems unlikely.
It is a pity that no sampling in the overburden was done in the vertical
direction. Even though the layer of loose material, in the area is thin such
an investigation could have been informative about the redistribution of
the elements during weathering and the formation of rana. Mineralogical
and chemical studies of the mode of occurrence of the metals in overburden
could have helped in the interpretation. The source of the southern anomaly
could have been detected simply by digging through the thin overburden
and one wonders why that was not done. The extension of Filon Rico was
338 Geochemical exploration examples

studied in that way and a vein, although sterile at the rock surface, was
detected.
The results were manipulated by statistical methods after removal of
extreme values. Treatment was made difficult by the fact that many of
the Ag concentrations are the same as the detection limit. There is a
high positive correlation between Pb and Zn and between As and Sb, and
the latter exhibits a good correlation with Ag as well. Zinc, in turn, is
clearly correlated with Mn. Factor analysis strengthened the assumption
that the original paragenetic relations in ore have been much altered in the
formation of soil. Geochemical analysis of residual soil and of the sorted
materials originating from it revealed that both are very local and suitable
for exploration use.

PORTE AUX MOINES: Pb AND Zn IN CHEMICAL SEDIMENTS AT A MASSIVE


BASE METAL MINERALIZATION IN PALEOZOIC BLACK SCHISTS
OF TEMPERATE RAINY HILLY REGION

From: Laville-Timsit, L. and Wilhelm, E., 1979. Comportement supergene des metaux
autour du gite sulfure* de Porte-aux-Moines (Cotes-du-Nord). Application ä la prospec-
tion geochimique. Bull. BRGM, Sect. II, 2-3: 195-228.
Laville-Timsit, L., 1986. Neoformation de sulfures de fer et de zinc en milieu reducteur
en aval du gisement Pb, Zn, Cu, Ag de Porte-aux-Moines (Cötes-du-Nord): l'anomalie
geochemique Pb-Zn des sols de Kerouran. Sei. Geol. Bull., 39 (3): 263-275.

The Porte Moines study area lies in the synclinorium of Armoricain


massif, mid-Bretagne, some 30 km southwest of Saint-Brieuc. Rounded well-
drained hills with poorly drained marshy lowlands in between dominate
on this plateau. The climate is temperate oceanic with 1000 mm annual
precipitation. The land has been cultivated for thousands of years, the
fields are ploughed and fertilized and the marshes pastured, so the original
distribution of elements has been changed by human activity.
Bedrock consists of W-E oriented Cambrian quartzites, mica schists
and black schists with intercalated diabases and tuffs, subparallel spilite
and rhyolite sills all cut by transversal faults. A massive polymetallic
mineralization containing 3 million tonnes of 10-12% bulk Zn + Cu + Pb
and 100 g/t Ag, calculated to a depth of 300 m, occurs in association with
shales, black cherts and acidic tuffs. Overlying the mineralization there is
an alteration zone 5-50 m thick, the upper part of which can be described
as gossan because it is formed of iron oxides and oxyhydrates mixed with
relicts of more resistant gangue rocks and minerals.
The overburden of the area consists of in situ weathered rocks, the
thickness of the weathering varying according to the chemical resistivity
of the rocks from a few centimetres to tens of metres. The soil, where not
peat bog, is brown earth type. The main transport of elements has been
Porte aux Moines: Pb and Zn in chemical sediments at a base metal mineralization 339

chemical, with some particles transported by surface waters. On top of the


mineralization in the elevated northeastern part of the area there is a thick
layer of oxidation sediments (gossan, as mentioned) and in the marshy
depression southwest of the hill there are layers of newly formed sulphides,
redusates, in the reducing environment of the peat bog.
The purpose of the study was to check the geochemical anomalies found
by stream sediment geochemistry, to compare the metal concentration of
soils with that of waters and to investigate the ways and means of secondary
dispersion of the elements in the overburden.
Soil samples were taken in the primary stage at a depth of 20 cm from
a sampling net 100 x 100 m; also a profile sampling was done at 10 m
intervals between sites and about 50 cm vertical interval along a 72 m long
N-S directed line (Laville-Timsit and Wilhelm, 1979). In the second stage
(Laville-Timsit, 1986) more samples were taken from the area where a zinc
anomaly was found in the primary stage. The <0.125 fraction of samples
was analyzed for total concentrations of Ca, Mg, Na, K, Pb, Zn, Cu, Ag, Fe
and Al by optical emission spectrometry.
The concentrations in samples taken from the rectangular grid (Fig.
ll-56a and b) show distinct anomalies for Pb and Zn. The main anomaly of
lead lies directly above the mineralization in gossan, but has an elongation
towards the southeast. The main anomaly of zinc, elongated along the N-S
directed swampy Kerouaran valley, lies west of the mineralization. In part
the anomalies of Pb and Zn are overlapping.
It looks as if zinc, whose ratio to Pb is lower in gossan than in the
mineralization, was captured and enriched by the organic compounds in
the acid, reducing environment of the peat bog. In the Kerouaran valley
anomalous zinc occurs in the fines of organic-rich soil in the form of
sulphide, wurtzite. Pb is found in greater than average amounts in a
somewhat coarser fraction than zinc and it appears to be occurring in
fragments of gossan that were transported to valley mechanically.
The location of the 72 m long shallow vertical profile line is not marked
on the published maps, but one can assume that it crosses the main Pb
anomaly and the eastern wing of the Zn anomaly in N-S direction. The
vertical distribution of the element concentrations in soil is presented in
Fig. 11-57. The iron contents (not shown here) tend to increase with depth
down to 2 m, as do Pb, Zn and Cu. The Mn contents decrease.
A deep drill hole penetrating the gossan and the weathered surficial layer
of bedrock down to 32 m shows that all the metals of interest increase in
concentration towards the bottom of the gossan at —23 m. Concentrations
decrease in the layer of weathered bedrock under the gossan, only to
increase again when the unweathered bedrock at a depth of —30 m is
reached (Fig. 11-58).
Microsond analysis showed variations in the element associations with
depth, suggesting different modes of occurrence of the metals. For example,
340 Geochemical exploration examples

Fig. 11-56. Distribution of (a) lead and (b) zinc (ppm) in surface soil at Porte aux Monies,
mid-Bretagne.

Pb seems to associate at greater depth with Fe, Ba and S, somewhat


higher up with Fe, S and K and closer to the surface with Al and P. The
cryptocrystalline ferri-oxyhydrate (goethite) mass contains as a kind of
concretion (from bottom upwards) anglesite, jarosite and plumbogummite.
Lead is relatively firmly bound in sulphates and phosphates in gossan,
but zinc has been partly leached out from it and transported further by
ground water. The mode of occurrence of metals in the drill core was also
studied by selective extraction, with somewhat similar results. Average
Porte aux Moines: Pb and Zn in chemical sediments at a base metal mineralization 341

Fig. 11-56 (continued).

total contents of important metals in assayed materials are presented in


Table 11-XI.
This fine example of geochemical exploration shows excellently the
different behaviour of elements during secondary dispersion. It shows how
in a single area there may be geochemical anomalies of different origin
and in different relation to one and same source mineralization. It also
shows how important it is to combine several methods of study in detailed
work: analysis of several grain size fractions, use of different methods of
extraction and chemical analysis, use of mineralogical studies, etc.
342 Geochemical exploration examples

L
H^

7ΠΓΧ LÜ!
I
I
oo ' oo
ooo oo
ooo oo oo
C9 T- CO i - CO I CM(0
ppm +- ppm 111111 p p m
3H loo o oo IOOO
ooo 10 O
ooo
i-iOO i-*C0
CM

Pb Zn Mn

Fig. 11-57. Concentrations of Pb, Zn and Mn (ppm) in overburden along a 72 m long N-S
directed vertical profile across the lead and zinc anomalies at Porte aux Moines, mid-Bretagne.
concentration

10 100 1000

Fig. 11-58. Concentrations of Pb and Zn (ppm) and Fe (%) in a drill hole through gossan at
Porte aux Moines, mid-Bretagne.
Grudie Burn: Mo in peat of temperate moist hilly terrain 343

TABLE 11-XI

Average total contents of important metals in assayed materials at Porte aux Moines

Drill hole Profile Surface soil Stream sed.


43 samples 50 samples 123 samples 55 samples
Pb (ppm) 1169 800 113 50
Zn (ppm) 87 89 157 135
Cu (ppm) 203 171 54 41
Ag (ppm) 5.6 4.3 0.6 0.2
As (ppm) 353 179 22 11
Mn (ppm) 200 300 1534 1055
Fe (%) 38 14 10 8

Results would have been better visualized by plotting the original


analytical results on the map and showing exact sampling locations.
Clustering of the results into concentration classes simplifies the anomaly
presentation but the possibility for individual interpretation of the reader
is lost.

GRUDIE BURN: Mo IN PEAT OF TEMPERATE MOIST HILLY TERRAIN

From: Smith, R.T. and Gallagher, M.J., 1975. Geochemical dispersion through till and
peat from metalliferous mineralization in Sutherland, Scotland. In: Prospecting in
Areas of Glaciated Terrain, Edinburgh, 1975. Institution of Mining and Metallurgy,
London, pp. 134-148.

The Grudie Burn West target area lies in the eastern Highlands of
Scotland. The region is hilly and mostly freely drained. The elevation of
the more or less even valley flank west of Grudie Burn stream ranges from
200 to 360 m a.s.l. The climate is coolish and moist. Vegetation consists of
conifers, heath, grass and, upon thicker peat bogs, Sphagnum moss.
The bedrock of the area is Precambrian micaceous metamorphosed sand-
stone with some mica schist intruded by Caledonian granite. Molybdenite
occurs sparsely over a larger area, both in sandstone and granite, with
pyrite and fluorite. Rock exposure is poor. In the exposed mineralization,
molybdenum occurs with pyrite in the quartz veins cutting the platy mica
schist.
The average thickness of overburden is 4.3 m, but in places till exceeds
15 m and peat 3 m in thickness. Peat on average 0.8 m thick, covers
the sandy lodgement till with a gleyed podsol soil profile over the entire
area. Some mineralized boulders have been found south of the exposed
Mo mineralization. The correlation between the location of mineralized
erratics and mineralization in bedrock is not clear. The valley glaciers
moved generally towards the south east — in Grudie Burn valley along the
contour lines.
344 Geochemical exploration examples

The purpose of the investigation was to locate the source of anomalies


found during regional geochemical stream sediment mapping.
Samples of till were taken by auger drill at different depths along lines
50-200 m apart. The <0.15 mm sieve fraction was ground and after fusion
total molybdenum was determined colorimetrically. For other metals other
methods of analysis were applied. Peat was sampled from the same points,
and after ashing at 470°C analyzed by the same method.
Molybdenum has been enriched into the finest grain size fractions of till
(cf. Kauranne, 1957). Its mode of occurrence is not totally clear, however,
as no molybdenite was found in mineralogical analysis of the fine fractions.
Molybdenum concentrations in till are high throughout the area, but the
anomaly/background contrast is clear nevertheless. Mo anomaly in till is
composed of two parts and corresponds rather well to the Mo concentrations
found in bedrock, the highest values in overburden being located almost
directly above mineralizations (Fig. ll-59b.) The geochemical anomaly in
till is elongated approximately in the direction of ice movement but is
open at the northern end. The study did not reveal any mineralizations in
bedrock there and yet the anomaly "overflows" the rim of the map. The
report offers no explanation for that.
Still higher molybdenum concentrations are met with in ashed peat
(Fig. ll-59a). These, even if "erratically disposed" as the authors say, form
a coherent anomaly which coincides fairly well with the Mo anomaly in
till. But the correlation of maximum values in peat and mineralizations
in bedrock was not so clear that the results could have been used for
locating diamond drill holes. The anomaly/background contrast is strong,
the background in peat ash being under 6 ppm and the anomaly mainly
over 21 ppm. The highest Mo concentrations found are in basal peat and
accompanied by high Fe and Mn and low ignition loss (LOI 20%). The low
ignition loss shows that the basal peat samples contained lots of mineral
material (i.e. mud).
The point network used for the till and peat sampling was uneven and
samples were collected by several different methods. In the north-western
part of the study area they were taken at 50 m intervals along N-S directed
lines spaced 100 m apart, whereas in other areas they were taken at
intervals of 50-200 m along lines in the E-W direction, 200 to 600 m apart.
It is difficult to get a real idea of the distribution of metals with such a
sampling network. If samples are not taken from a regular grid, the lines
should be directed transverse to the strike or the contacts of geological
formations, or transverse to the clastic or hydromorphic transport. Here
SW-NE lines would best have fulfilled these criteria.
Use of an open-spiral auger makes it highly possible that till samples
taken from the bottom of the hole were contaminated with peat from the
surface.
Holes drilled through the overburden and fixed interval vertical sampling
Grudie Burn: Mo in peat of temperate moist hilly terrain 345

(a)

- Sampling point

Mo in peat ash
6 21 150 PPm

§
(b)

HJ

Sampling point

Mo in till fines

20 35 70 PPm

w 1000 m

Fig. 11-59. Distribution of molybdenum in (a) peat (ppm, in ash) and (b) till <0.15 mm fraction
(ppm) at Grudie Burn, Scottish Highlands.

allowed study of the distribution of metals in vertical direction. A sharp


difference in concentrations on either side of the peat/till interface, and
at the till/bedrock interface is evident. Concentrations of many metals are
seen to increase with depth in both peat and till (Fig. 11-60). Boreholes
have not been connected to form vertical profiles, though this might have
revealed the type of transport of heavy metals. In the search for the source
of anomalies this sort of vertical distribution study is very important.
346 Geochemical exploration examples

Mo ppm Cu ppm U ppm Mn ppm Fe %

Fig. 11-60. Vertical distribution of Mo, Cu, U, Mn and Fe in a drill hole through peat, till and
weathered rock down to bedrock at Grudie Burn, Scottish Highlands.

The authors assume that the primary transport of molybdenum and


the other heavy metals in drift was mechanical but that the present
distribution is partly caused by a secondary transport in water solution.
The concentrations in peat depend on local drainage and Eh-pH conditions
and, in the authors opinion, peat sampling is therefore of limited value in
detailed exploration.

LIETSONSUO: Cu AND Ni IN PEAT OF HUMID COOLISH PLAIN TERRAIN

From: Nieminen, K. and Yliruokanen, I., 1976. Kitee: the copper-nickel anomalies of
Lietsonsuo peat bog. J. Geochem. Explor., 5: 248-253.

Lietsonsuo bog lies in southeastern Finland close to the border with East
Karelia. The undulating topography allows fairly free drainage except in
the silt-paved depressions. Climate is coolish and humid. The forest is a
mixture of deciduous trees and conifers with the common Vaccinium-type
undergrowth. A sparse growth of stunted spruce and pine characterizes
the bog surface. The bog was originally mainly the floating type, but has
developed into raised bog with Sphagnum moss as the main plant.
Bedrock is Proterozoic gneissose mica schist surrounded by pegmatitic
granites. The overburden consists of till, covered by glacial varved clays in
flat areas and bordered by a ridge of coarser glacifluvial formations in the
east. Two transport directions are known, the older from NNW and the
younger from WNW Soils are podsolic. No ore mineralizations have been
discovered.
The purpose of the study was to find whether heavy-metal-rich peat
layers could be mined as ore and to search for the source of the metals in
bedrock.
Lietsonsuo: Cu and Ni in peat of humid coolish plain terrain 347

500-

Ni concentrations
in peat ash

500

fflffls
Cu co ppm co Ni

Fig. 11-61. Distribution of copper and nickel in peat (maximum concentrations in the bog, ppm
in peat ash) and vertical distribution of copper and nickel along a typical line (ppm, in peat
ash) in Lietsonsuo bog, eastern Finland.

The bog was sampled at 100 m intervals along lines 200 m apart; samples
were taken vertically, one every metre from top down to hard bottom. The
peat was dried, ashed at 550°C and analyzed by XRF for copper, nickel,
zinc, lead and uranium.
According to the report the distributions of Cu and Ni are very similar.
The maximum concentrations at every point are presented in map form in
Fig. 11-61. The concentration contour lines are very similar for the two
metals, with highest concentrations approximately at the centre of the bog.
Figure 11-61 also shows the vertical distribution of Cu and Ni along one of
the east-west directed lines.
The concentrations of Cu and Ni are low in the surface and bottom
samples, although most heavy metals tend to concentrate close to the
348 Geochemical exploration examples

bottom (see Salmi, 1967). The highest concentrations in ashed samples from
the middle of the bog are over 2000 ppm for Cu and over 3000 ppm for Ni.
Unfortunately, the report says nothing about the distribution of zinc, lead
and uranium. From other investigations we know that lead and usually also
iron and molybdenum are concentrated in the surficial layers of peat, while
uranium, like titanium and vanadium, has a tendency to concentrate near
the bottom. Zinc resembles copper and nickel in being more indifferent.
It is further regrettable that clay/silt under the peat was not sampled at
all, and that till was sampled only along two lines, with results presented
very unclearly The few samples taken from till revealed weak anomalies
of about 130 ppm Cu and 120 ppm Ni at the place where a rise in the till
surface penetrates the clay bottom paving of the peat layer. The highest
concentrations in peat occur at about the same height level as the surface
of this till elevation.
Although the heavy metal concentrations in mineral formations are
much lower than those in dried peat and especially those in peat ash,
the element distribution pattern in till have been informative about the
transport direction. The heavy metal spectrum of till might also reveal
the type of source mineralization. In peat the original internal ratios
of elements have been distorted and cannot be used as a "fingerprint"
of the mineralization. Peat study can be recommended for regional and
semi-detailed base metal geochemical exploration but, judging from this
example, not for target studies.

MASUGNSBYN: U IN PEAT IN HUMID ARCTIC HILLY TERRAIN

From: Armands, G., 1967. Geochemical prospecting of a uraniferous bog deposit at Ma-
sugnsbyn, northern Sweden. In: A. Kvalheim (Editor), Geochemical Prospecting in
Fennoscandia. Interscience, London, pp. 127-154.

The Masugnsbyn target is located in Sweden north of the Arctic Circle,


not far from the Finnish border. The bog studied lies about 300 m above
sea level in a 2-4 km wide valley flanked with elongated hills rising to
375-400 m above sea level. Drainage is poor. Climate is cold and humid
and some parts of the bogs are permanently frozen. Vegetation consists of
northern-type spruce, birch and dwarf birch, with sedge (Carex) and some
moss (Sphagnum) in the bogs.
Bedrock under the bog consists of a 0.5-1 km wide NW-orientated zone of
quartzite and phyllite of Lapponian volcanosedimentary series, surrounded
by late Karelian granites. Uranium occurs with high probability in the
fissures parallel and close to the pegmatite-rich border zone of the so-called
Linagranite.
Overburden consists of sandy till 3-7 m thick overlain by peat bog 1-2 m
deep, with 1 m deeper depressions filled with mud. Esker-like formations of
Masugnsbyn: U in peat in humid arctic hilly terrain 349

hummocky ablation till occur on either side of the bog. Numerous magnetite
and sulphide ore erratics, but no boulders with marked radioactivity, have
been found. Because the area lies on the ancient ice divide, glacial transport
has been short. Glacial striae point in two distinct directions: the older
towards east-southeast (from 300°) and the younger towards northeast
(from 225°).
The waters and peat are radon and uranium rich. The anomalous area
was found during car-borne scintillation counting on the Kiruna road in
1958. The purpose of the geochemical study was to find the source of the
anomalies.
The highest uranium content in water (1800 /zg/1) was found in one of
the bore holes getting water from fractures in the bedrock. Ground water
in till contained 2-3 /ig/1; natural springs in depressions showed an average
of 6 /ig/1, but some had contents of more than 700 /ig/1. Such a spring would
produce about 6 kg uranium annually. Uranium concentration in streams
varied from 0.2 to 7.2 /^g/l.
Observations made on vegetation showed variable uranium concentra-
tions in different species and in different parts of plants. Amounts were
highest in willow (ppm in ash): twigs 680, leaves 450 and fruit 450. The
activity of willow (Salix) twigs was 2.5-fold that of willow leaves, while the
corresponding value for alder (Alnus) was 1.9, for dwarf birch (Betulanana)
2.5 and for birch (Betula alba) 3.9. The radioactivity measurement shows
how geophysical methods can give good though vague results in uranium
exploration.
No systematic study was made of the uranium concentrations in till, but
there was a positive correlation between the uranium concentrations in till
and in water. There was also a correlation between uranium content of
the water in peat and that of the peat material itself. As expected, neither
the uranium or radon concentrations nor the activity in waters or in peat
were in equilibrium with each other. The correlation between the uranium
concentrations of surface peat and surface vegetation is not at all clear,
although one would think that the only way bushes can get their uranium
is via water from the peat.
Peat, like humus, is a strong concentrator of heavy metals. The highest
uranium concentration found at Masugnsbyn was 3.1% in dry peat; the
average amount in 445 samples was 600 ppm U in dry material. Peat was
impure: 33% of its dry weight was mineral grains. The enriching factor
if attributed to organic material was thus 9000. According to laboratory
experiments done by Szalay (1958), the enriching factor for uranyl ions
between a dilute water system and a peat system could reach 10,000.
Uranium is easily leached from granite or its ores by HCO3-rich water of
pH about 7.5. It is precipitated with manganese at pH 7.4-7.9 and fixed in
peat between pH 3 and 7.
At the surface of the peat bog the uranium anomaly is branched into
350 Geochemical exploration examples

U concentration
in dry peat

100 1000 4 0 0 0 ppm

100 m

Fig. 11-62. Distribution of uranium in surface peat (ppm, in dry peat) at Masugnsbyn, Swedish
Lapland.

Fig. 11-63. The acidity of peat in a vertical profile through the bog, the distribution of uranium
in peat layer and the radioactivity of surface peat at Masugnsbyn, Swedish Lapland.
Masugnsbyn: U in peat in humid arctic hilly terrain 351

U ppm peaf ash


0 1000 2000
I I J

0 U jug/l water 1000

Rn
1000 (emans) 2000

Fig. 11-64. Uranium concentration in peat and in peat water as well as 7-radiation on the
surface of the bog along the N-S profile at Masugnsbyn, Swedish Lapland.

two parts, possibly reflecting the movement of water in the peat layer. The
longer branch is broken by background concentrations (Fig. 11-62). The
highest concentrations at the surface are well over 4000 ppm (0.4% U, in
dry peat). As is very often the case, the highest heavy metal concentrations
occur in the bottom layers of the peat, (Fig. 11-63), although one 1%
U-content in the N-S profile was found at the surface south of the ditch.
The measured radiation (7-activity) on the peat surface does not follow the
distribution of uranium in either peat or peat water (Fig. 11-64).
In the Masugnsbyn case, a very thorough investigation was made of the
uranium in peat and the factors affecting it. The concentrations of U, Rn, Ra
and the radioactivity were measured — however, conclusions could not be
drawn about the partitioning and the causes of imbalance of the daughter
elements of U and Th. Factors affecting the solubility and precipitation of
uranium — acidity (pH), redox potential (Eh), degree of humification (H),
amount of mineral material (ash%) and concentration of manganese (Mn)
— were studied and internal correlations calculated. It was found that
neither pH or Eh could explain the variation of uranium concentrations,
but that Mn and especially the ash% affected the precipitation and fixation.
It was concluded that uranium is leached from bedrock by ground waters,
which seep out to the overburden from several fractures, enriching uranium
in the organic material.
352 Geochemical exploration examples

One very important medium was not properly explored — the mineral
overburden under the peat. Test sampling revealed variation in uranium
concentrations in the till. A fuller picture of the distribution of uranium
in till could have given valuable information on the places where ground
water is seeping into overburden from the fissures of the rock and perhaps
on where the richest mineralizations are located in the bedrock.
Analysis of peat is a useful tool for regional or semi-detailed uranium
exploration, but it cannot be recommended for target studies with which
sampling of bedrock (diamond drill holes) must be planned.

ALLEBUOUDA: W IN HUMUS OF HUMID MOUNTAINOUS ENVIRONMENT

From: Toverud, Ö., 1979. Humus: a new sampling medium in geochemical prospecting
for tungsten in Sweden. In: Prospecting in Areas of Glaciated Terrain, Dublin, 1979.
Institution of Mining and Metallurgy, London, pp. 74-79.

The Allebuouda prospect lies a little south of the Arctic Circle in northern
Sweden, where the climate is cold and moist. The Allebuouda fell, rising
180 m above the surface of lake Björnträsk, is typical of the rounded,
well-drained hills characterizing the region. Vegetation consists of dwarf
birch, juniper, heath, grass and moss.
Bedrock of the study area consists of Precambrian metavolcanites and
granite intruded by aplite and pegmatite veins. Scheelite mineralization
with some molybdenite occurs in tuffaceous skarn close to the northern rim
of a granite dome.
The overburden varies in thickness up to a maximum of 10 m, disappear-
ing in the western and northwestern parts of the area where rocks outcrop
on the top of fells. Glacial till with a podsolic soil profile on top is covered in
the southeastern corner of the area by peat. A 1 km long scheelite-bearing
garnet-skarn boulder fan curves round the flanks of the Allebuouda fell
towards the southeast at the apex about 320°, at the tail 350°. Glacial striae
are oriented between 320° and 335°, but there is also an older direction
from 300°.
The purpose of the study at this fairly well-known prospect was to
compare the results of a chemical analysis of humus with the results of an
analysis of the heavy fraction of till to assess the suitability of humus as an
exploration material for tungsten. Counting of scheelite (and wolframite)
grains under ultraviolet light from a panned heavy mineral concentrate of
till has long been considered a sensitive and reliable exploration method for
tungsten, whereas to determine tungsten chemically has been tedious.
Samples were collected from the C-horizon of till at about 50 cm depth
on a 100 x 200 m grid. After drying and sieving of the samples, the heavy
metal concentrate of the 0.5-5 mm fraction was analyzed for selected
metals by X-ray fluorescence. Humus, which in the area is 2-5 cm thick,
Allebuouda: W in humus of humid mountainous environment 353

Fig. 11-65. Distribution of tungsten in the heavy mineral concentrate of the 0.5-5 mm fraction
of till (ppm) at Allebuouda, Swedish Lapland.

was sampled from the same grid, ashed at 450°C and analyzed for metals
by ICP spectrometry
Tungsten concentrations of the till samples are not high even though
the material analyzed was enriched by panning; neither is the anomaly/
background contrast strong. The anomaly in heavies of till has clearly been
derived from the ore suboutcrop in the bedrock. Its failure to coincide
with the ore boulder fan, though not usual, requires an explanation.
Separation due to solifluctional creeping of surface till rich in boulders
seems improbable. In fact, there are two anomalies in till, separated
by a 200 m wide gap (Fig. 11-65). If these represent the same source,
the material lying to the southeast must have been transported by an
older ice movement and then may be pushed a little towards the southeast
354 Geochemical exploration examples

Fig. 11-66. Distribution of tungsten in ashed humus (ppm) at Allebuouda, Swedish Lapland.

by the younger movement.


The tungsten contents in humus are given as concentrations in ash;
though the anomaly itself is not strong, the anomaly /background contrast
is clear. The tungsten anomaly in humus almost totally coincides with the
anomaly in till and like it is divided into two parts (Fig. 11-66). The humus
anomaly is bent along contour lines towards the south. There is no anomaly
in humus directly above the ore suboutcrop.
Humus samples were also analyzed for molybdenum and tin. The tin
anomalies coincide rather well with the tungsten anomalies. Molybdenum
contents show a totally different distribution pattern, but there are two
anomalous points directly above the ore suboutcrop. Two almost economic
molybdenum mineralizations have been identified in the area and it is a
pity the sampling grid was not extended over them.
Hannukainen: Cu in humus in humid hilly environment 355

The report does not tell about the overburden above the ore suboutcrop,
nor what results were obtained in the analysis of the - 0 . 1 mm till fraction.
It would have been important to investigate the occurrence of tungsten
deeper down in the till, at the apex of the anomaly. What is the relation
between the anomaly and the suboutcrop of the mineralization? Deeper
sampling of till at the gap between anomalies might have revealed the
reason for the gap.
There is in the report a mention of a mineralogical study of the till and
the identification of scheelite and wolframite by XRD. Presenting of the
results of the scheelite grain count of the heavy fraction of till would have
assisted the reader in the evaluation of the suitability of humus analysis
for tungsten exploration.
As an assessment of the usefulness of humus analysis in tungsten
exploration, the investigation perhaps was good enough. In the case of
real exploration, the sampling network should have been larger and, in
the second stage, denser in the central area to verify the preliminary
interpretation. Some deeper pits should have been dug northwest of the
apex of the till anomaly and sampled at different depths to follow the
anomaly in vertical direction.
Geochemical study of humus is not accurate enough to allow location
of diamond drill holes that will hit an ore and allow inventory of the
mineralization.

HANNUKAINEN: Cu IN HUMUS IN HUMID HILLY ENVIRONMENT

From: Äyräs, M., 1979. Eräiden hivenmetallien esiitymisestä moreenissaja humuksessa


Hannukaisen alueella Kolarissa. Geol. Surv. Finl., Rep. Inv., 44, 34 pp.

The Hannukainen prospecting area lies in northern Finland close to


the Swedish border, about 100 km north of the Arctic Circle. The area is
gently rolling and freely drained, the highest hilltop being 300 m and the
lowest stream valley 167 m a.s.l. Climate is cold and humid. Vegetation
consists of northern-type spruce and some pine, with birch and willow
in the valleys. Surface vegetation is mostly blueberries, lingonberries and
heath on hillsides and sedge and moss in moist depressions.
Bedrock from east to west (from the lowest unit upwards) is quartzite,
quartz feldspar schist, amphibolitic skarn with iron ores and monzonite
with a dioritic east rim dipping gently westwards (Fig. 11-67). Younger
granites and pegmatites intersect all these formations. Four skarn iron
orebodies are encountered in the area, the southernmost and northernmost
being "blind", while the two in the middle, Laurinoja and Kuervaara, have
a suboutcrop under the overburden. The middle two orebodies contain up
0.4% copper.
356 Geochemical exploration examples

Cu Cu/C

ppm
15 0.015

— i i 1 1 1
500 1000 1300r

FeyFe/C

2.5 10

0-L0
1300 m
SP

mV
300H

150

OH
1300m
-75

200m ^Μο/}ζοη]}9/

Fig. 11-67. Copper and iron contents of humus and their ratio to carbon contents as well as self
potential along a line crossing the Laurinoja ore deposit at Hannukainen, North Finland.

Overburden consists of till interbedded with sand with a total thickness


up to 30 m. The uppermost brownish till, which was transported from the
southwest, is underlain by grey till transported from the west. Beneath
these an older till is sometimes encountered, itself overlying a thin layer of
in situ weathered regolith. The transport of till was short because the area
lies in the zone of the ancient ice divide. Soil is podsolic.
The investigation was done to determine the suitability of humus
prospecting in an area of complex, rather thick overburden. Samples of
Hannukainen: Cu in humus in humid hilly environment 357

the 1-5 cm thick humus were taken at 50 m intervals along E-W directed
lines 200 m apart. In depressions where there was a thin peat layer, this
was sampled instead of humus. By way of comparison, the suitability of till
geochemistry was investigated as well.
Samples of upper till were taken from a depth of 2 m, at 100 m intervals
along lines (NW-SE, perpendicular to glacial transport) 1 km apart. Some
of the drill holes were driven down to the bottom of the overburden and
sampled every metre.
Till samples were dried and sieved. The 0.06-0.25 mm fraction was
reserved for mineralogical studies while the —0.06 mm fraction was divided
into three parts. Total contents of Cu, Ni, Co, Zn, Mo and Fe were
determined in one part by optical emission spectrography, and the other
two parts were digested, one in a hot HC1 + H2O2 mixture and the other in
diluted cold HC1 before analysis by AAS for the same heavy metals.
Humus and peat samples were dried, ashed at 500°C, digested in hot
(1:1) HC1 and analyzed by AAS. Results were calculated back to dry mass.
The mean concentrations of the metals in about 450 till, 500 humus and 90
peat samples are presented in Table 11-XII.
Amounts of magnetite determined by susceptibility from the 0.06-0.25
mm fraction of mineral materials were highest in the tills, whereas the
sand bedded between them was poor in magnetite. The same pattern was
mostly true for the heavy metals too. Copper showed highest average
concentrations in the lower till.
Immediately east of the largest ore outcrop (Kuervaara) the copper
content in the upper till is highly anomalous: hot leach copper concentration
350 ppm, compared with a concentration of 1000 ppm in the ore below. The
distribution pattern of the heavy metals in the upper till is very peculiar:
namely, the anomalies partly lie to the southwest of the suboutcrops of
the ore bodies, on the "wrong" side to what one would expect from the
glacial transport. Anomalies in humus show the same feature near the ore
suboutcrops (Fig. 11-68). The most distinctive anomalous areas of the cold

TABLE 11-XII
Average metal concentrations in till, humus and peat in the Hannukainen area

Till Humus Peat


Cu (ppm) 25 6 8
Ni(ppm) 15 4 4
Co (ppm) 9 2 3
Zn (ppm) 16 37 5
Pb (ppm) 8 22 6
Mo (ppm) 1 0.2 0.7
Fe(%) 1.7 0.3 0.6
358 Geochemical exploration examples

&-'·?

rtJ \ Λ fV«'

Jfifi'''

Cu concentrations

rySj in dry humus


jy^ll >7ppm

f
in fines of till
acid soluble
> 28 ppm

Ore outcrop and


^ orebody surface
v—-^ projection

V ; \»"i

Fig. 11-68. Distribution of copper concentrations in humus and till at Hannukainen, North
Finland.

leach copper concentrations in till are to the west of the second outcropping
ore body (Laurinoja).
The maximum total concentrations of copper in till are found on the
northern side of the long "blind" orebody (Vuopio). The maximum total
concentration of Cu in the upper till determined by optical emission
spectrography is 1050 ppm, compared with 600 ppm in hot leach and 180
ppm in cold leach. The maximum total copper concentration in the lower
till is 1770 ppm. There are very high levels of copper in the sand between
the tills too: maximum total concentration is 1450 ppm.
Petalax: heavy metals in humus of humid temperate climate, plain terrain 359

It is a great pity that only the "anomalies" — the concentrations lying in


the upper 20% of the cumulative distribution curve — are presented on the
maps, and not the original concentrations, so that there is little possibility
for the reader to make his/her own interpretation (for copper the threshold
values were hot leach 28 ppm and cold leach 7 ppm). The total picture of
the Hannukainen target therefore remains unclear.
A differentiation between the clastic and hydromorphous transport of
till was attempted (1) by comparing the concentration of magnetite in
the 0.06-0.25 mm fraction with the concentration of iron in the —0.06
mm fraction, and (2) by comparing the results of hot and cold chemical
extractions. The latter comparison showed that 10-30% of the metals of the
fine fraction of the upper till was readily extractable and likely transported
in water solution.
There is thus some evidence that the material from ore outcrops was
originally transported clastically and that later hydromorphic transport
was responsible for the anomalies on the "wrong" side of the suboutcrops.
A microscopic study of heavy metal minerals would have helped to clarify
the true situation; the sequential chemical analyses were not sufficient
in this case. The relation between the upper and lower tills was not
established; stone counts could have helped in deducing transport distances
and directions, but for this more samples should have been taken from the
lower till.
The main purpose, evaluating the usefulness of humus studies in
prospecting, was not fully achieved. The distribution of heavy metals in till
and the suitability of till geochemistry for exploration in the conditions of
this case remain obscure. The explanation of the location of geochemical
anomalies in humus preferred by the author may be approximately correct
— i.e. cations may have been lifted up to humus by electrochemical poten-
tial: (1) from the ore bodies through monzonite and overburden, (2) from
the hydromorphic anomaly in till, or (3) from both. But there are other
possible explanations, equally as good and the reader remains confused.
It seems that neither the surface till method nor the humus method can
easily be applied in the case of multiple till with sand layers in between,
and the better approach would be to use deep till assaying.

PETALAX: HEAVY METALS IN HUMUS OF HUMID TEMPERATE CLIMATE,


PLAIN TERRAIN
From: Kauranne, L.K., 1976. Petolahti: copper and nickel in till and humus. J. Geochem.
Explor., 5: 292-296.

The Petalax study area is located in the low-lying western coastal zone
of Finland. The area appears to be fairly freely drained, although the
topographical variation is small. The target lies on a low crest between two
360 Geochemical exploration examples

peat bogs. Climate is temperate and humid. Forests consist of deciduous


trees — birch, alder and rowan — with coniferous species — spruce, pine
and juniper — intermixed. Heath, blueberries, lingonberries and grass
cover the forest floor.
Bedrock of the area is Proterozoic mica gneiss oriented in N20°W25°E
direction and sharply cut by a pentlandite-chalcopyrite bearing diabase in
N65°W direction (Ervamaa, 1962).
The overburden consists of sandy lodgement till 1-6 m thick, covered in
places by a thin silt layer or peat. There are five Ni-Cu-rich ore boulders
south of the ore suboutcrop, approximately in the direction (towards 170°)
of the younger ice movement, the older direction being towards southeast.
The purpose of the study was to see whether determination of cold-
extractable metals in humus could be used for geochemical location of ore
(Kauranne, 1967).
Humus was sampled at 20 m intervals along E-W lines 50 m apart in
1960. Till was sampled with a percussion drill and an Irish-type sampler
along every second line at 20 m intervals and at 1 m depth in 1975.
Humus was extracted with cold ammonium citrate buffer (pH 8.5) and
shaken with 0.03% dithizone xylene solution (Bloom, 1955). Xylene rises to
the surface of the water and the green dithizone turns red, the brighter the
red the greater the concentration of Cu and other base metals. Till samples
were dried, sieved and the —0.06 mm fraction was analyzed for total Ni,
Cu, Zn and Pb by optical emission spectrography.
Metals of sulphides are concentrated in the glacial process into the fines
of till, but only a small part is in easily soluble form (see Table 11-XIII).
The situation is different in Lapland where a large amount of preglacially
weathered material is mixed in till.
The areal distribution of citrate extractable heavy metals (ex HM) in
humus and bulk base metal (Ni, Cu, Zn, Pb) concentration in till are
presented in Fig. 11-69, A and B. The blank areas on the humus map are
cultivated fields.

TABLE 11-XIII

The average total, mineralic acid-soluble and exchangeable Ni and Cu of till in Petalax

Total content HN0 3 -leach Citrate extraction


Ni < 0.06 mm 46.8 28.8 5.0
0.06-0.25 mm 39.8
Cu < 0.06 mm 44.0 28.4 4.3
0.06-0.25 mm 29.3
a
The average bulk heavy metal concentration of humus in the entire study area converted to
copper equivalents is 42 ppm.
Petalax: heavy metals in humus of humid temperate climate, plain terrain 361

Fig. 11-69. A. Distribution of cold citrate extractable heavy metals (ex Hm) (Ni + Cu + Zn
+ Pb) in humus (ml of dithizone consumed). B. Distribution of the Ni + Cu + Zn + Pb
concentration (ppm) in the fines of till at 1 m depth, at Petalax, western Finland.

The humus study revealed an anomaly pointing directly towards the


ore suboutcrop. This suggests that analysis of cold extractable bulk heavy
metals in humus can suitably be employed for target studies in geochemical
exploration in this type of situation.
By contrast, neither the total concentrations nor the HN0 3 -soluble
or cold extractable base metal concentrations in till formed coherent
anomalies. Their sum amounts formed a narrow anomaly, but one running
in a direction of about 150°, whereas the long axis of the rather clear
ex HM anomaly in humus is about 180°. For some unknown reason, the
orientations of the anomalies deviate from one another.
The area studied by humus analysis was very narrow and therefore the
area of the comparative till study became much too small. The main mass
(sand size and finer fractions of till) usually is not transported far; very
often the first signs of anomaly at the till surface are seen at 200-300 m
distance and the last detectable traces of ore material 500-2000 m from the
362 Geochemical exploration examples

ore suboutcrop. In the present case the higher concentration in surficial till
in the southern part of the study area, 300 m south from the deposit, may
be the first real expression of the ore in the down-ice direction and the till
anomaly just begins to show up.
Maybe the ore-bearing tillmaterial at the apex of the anomaly is situated
close to the bedrock surface and was not reached by the 1 m sampling.
The anomaly in humus may reflect that material and maybe the weak sum
anomaly in till above it has hygroscopically climbed up from that deeply
situated material. Analysis of the hydromorphically transported part of
the heavy metal concentrations in till should have been performed. This
case was not studied properly and even the classification of the metal
concentrations in till and their presentation with class symbols on the map
prevents readers from drawing their own conclusions.
363

Chapter 12

FOCAL ASPECTS OF GEOCHEMISTRY APPLIED IN ARCTIC AND


TEMPERATE REGIONS

THE PLACE OF SOIL GEOCHEMISTRY IN EXPLORATION

The aim of regional geological research is to clarify the lithology,


stratigraphy, tectonics and possibly the technical character of the bedrock
and drift. When mineralized or altered horizons of bedrock, mineralized
erratics or gossan are discovered in the course of geological mapping,
follow-up studies are appropriate.
Airborne magnetic, electromagnetic and radiometric surveys and regional
gravimetry provide useful information on geological structures. Areas
that differ strikingly from the environment (anomalies) may host ore
mineralization and need to be studied in greater detail.
Soil geochemistry is an essential link in the chain of explorational
operations, whether used before, after or at the same time as geological and
geophysical methods.
Geochemical mapping, involving the assessment of abundant and evenly
distributed materials such as stream and lake sediments, till, humus and
peat, reveals geological structures, and as an element-selective method
it can also give direct indication of hidden mineralization. Even a pair
of samples, especially if taken from transported overburden, is at the
same time a source of regional, local and site information. The elemental
associations in these samples reflect the trends in the geochemical province,
the nature of the local formations and the character (fingerprint) of the
lithology at the site (Saarnisto and Taipale, 1985).
Bedrock is usually considered to be the primary geological material
in geochemistry, although it contains formations of different age, even
fossilified geochemical secondary anomalies. Secondary geological materials,
such as overburden, waters and vegetation, contain element populations
derived from the bedrock lithologies and mineral deposits transformed by
the geochemical events the material has experienced. These populations
can be discriminated if a statistically significant number of samples are
analyzed over a large enough area (Sinclair, 1976). Sometimes the "extra
noise" of larger, sparsely mineralized formations such as black schists,
364 Focal aspects of geochemistry applied in arctic and temperate regions

or the "obscuring veil" of pollution caused by present or past mining


activities, makes identification of the signal from an ore mineralization
difficult. In such a case a multivariate statistical analysis or a closer
geochemical study with sequential analysis and/or mineralogical study is
needed to discriminate the anomalies from the superimposed element popu-
lations.
It is always wise to begin both regional and local geochemical studies with
a pilot study to support the planning of the study proper. In geochemistry
the study is often executed in different stages and scales: country-wide,
regional, areal, local and detailed target study. In preparatory stage several
materials should be studied by different methods as a means of assessing
local factors and conditions.
In the final stage of detailed prospecting, if not before, a variety of
supporting geological and geophysical methods should be used to ensure
that the geochemical results are interpreted as comprehensively as possible.
This kind of saturation prospecting is strongly recommended, even today
when geochemical surveys are "one of the principal methods, if not the
principal method, for discovering blind deposits of nearly all types of
minerals" (Boyle, 1982). The last stage in target study is deep drilling
to determine the form, size and grade of the deposit. Semi-industrial
pilot feasibility and refining studies may follow if the size and grade are
encouraging.

PRINCIPLES OF GEOCHEMISTRY

An ore mineralization is a concentration of elements in bedrock or


overburden, while an ore deposit is an ore mineralization rich enough to
be economically exploited. Geochemical exploration for ores involves the
analysis of anomalous concentrations of ore elements themselves or of
indicator (paragenetic) elements for ores or their host rocks (Boyle, 1974).
The success of the method rests on the tendency of aureoles of greater than
average concentrations of elements to form around ore bodies.
An aureole, or halo, in the bedrock is called primary regardless of
whether it was formed in connection with the genesis of the ore itself or
by later processes. An anomalous aureole in the overburden, groundwater,
soil, air, surface waters or atmosphere is called secondary and is classified
according to the process by which it was formed, as a residual, clastic or
chemical anomaly. Higher than average concentrations of elements close
to ores in vegetation and animals are also called secondary anomalies,
although they might better be defined as tertiary. Vegetation and animals,
at death, are returned to the overburden, which itself through diagenetic
and metamorphic processes turns back to bedrock. Eventually the various
anomaly types become "fossilified" in the bedrock.
Principles of geochemistry 365

The prospector wishing to use geochemical methods may choose the


material to be assessed quite freely, as there are indications of bedrock
lithologies in all of them. Some materials are nevertheless better than
others. Because the results are comparable only if one and same material is
used throughout the area, the material should cover the area as completely
as possible. It should be easily obtainable, be in intimate or at least close
contact with the bedrock and be formed by simple processes.
If the bedrock is widely exposed, rock chemistry can be used (see the
Volume 2 of the Handbook, Rock Geochemistry in Mineral Exploration, by
Govett, 1983), with the caution, however, that the great differences in
chemistry between lithologies may hide any weaker geochemical variations
of elements (the halos) in the rock due to a deeper lying ore. In areas of in
situ weathered regolith rock geochemistry can be recommended.
In warm climate in situ weathered rock makes up the main part of the
overburden. The use of soils of tropical regions is described in another
volume of the Handbook edited by Butt. Weathering tends to smoothen
the chemical differences between lithologies, while at the same time the
soil forming processes discriminate the elements into horizontal layers.
The transformations of the original compositions into new associations and
internal relations can usually be traced back with the aid of mathematical
methods.
In transported formations of overburden the internal relations of el-
ements differ still more from the original ones depending of course on
the mode of transport — whether the elements were moved as mineral
grains clastically, in water solution as separate ions or in successive stages
as a combination of both. The use of different overburden formations for
exploration has been called pedogeochemistry in the west (Rose et al.,
1978) and lithogeochemistry in the east (Beus and Grigorian, 1977). The
use of recent water-laid sediments in exploration is described in detail in a
separate volume of the Handbook edited by Plant.
Sometimes the transporting medium itself can usefully be used as
investigation material: hydrogeochemistry is widely practiced and atmogeo-
chemistry is increasing in popularity. Plants obtain their elements either by
water through roots or by air through leaves and needles. Animals in turn
may obtain their elements by eating plants, but a part also by drinking
water and inhaling air.
The materials found in a region and useful for geochemical exploration
depend to a very large extent on climate. In the high ice-covered Arctic,
geochemical exploration based on sediments of meltwater streams on the
ice surface has led to the location of mineralization on the flanks of
nunataks (Lawson, 1974). However, the analysis of dust layers in ice has
also revealed the ash of distant, but identifiable volcanic eruptions (Körner
et al., 1989). Such airborne dust complicates exploration, though improves
perhaps the construction of the local tephrachronological time-scale.
366 Focal aspects ofgeochemistry applied in arctic and temperate regions

On barren, permanently frozen tundra the same normal, geochemical


materials as in the temperate zone — till, stream and lake sediments
(Batterson, 1989), humus, etc. — can be investigated. More exotic is
the assaying of mudboils (frostboils) (Shilts, 1978; Riese et al., 1968). In
fossilized form, mudboils or "ice wedges" are also found outside permafrost
areas. Preglacially weathered rock, deposits of gossan and other chemically
discriminated formations are found also in glaciated terrain in places and
can be assayed for exploration purposes (as described for other regions by
Laville-Timsit and Wilhelm, 1979; Matheis and Pearson, 1982).
In contrast with arctic areas where glacial erosion was weak, in the zone
approximately down to latitude 50°-55° nearly all preglacially weathered
crust has been abraded away and replaced by glacial or postglacial sed-
iments. The most suitable materials for geochemical exploration in this
temperate region are till, watercourse sediments, humus and peat.
At the maximal extension of the continental ice where ice erosion again
was weak, several partially sorted layers of till were deposited above each
other making the use of till in geochemical exploration of questionable
value. Higher up in mountains it may be acceptable, but for lowlands and
valleys watercourse sediments would be better.
Farther to the south (or north in the southern hemisphere), in the
nonglaciated zone, the in situ weathered regolith, if exposed, will be better
than the transported materials as the target of analysis for geochemical
exploration.
For geochemical mapping of whole continents, an appropriate material
must be selected for each climatic zone, but the zones and materials should
overlap to allow the results for one material to correlate with the results
for the other when presented on the map.
A slightly different situation in which materials depend on climate
is encountered in mountainous regions, where soil formation at higher
elevations resembles that of colder latitudes. The increased precipitation,
better drainage and landslides may all affect the formation of geochemical
anomalies on mountains.
It must be remembered as well that climate changes through time.
The tree limit has travelled back and forth and living species have varied
(Edlund, 1989; Anderson, 1989). For example, 14C determinations have
shown the growth of peat in a Finnish bog to have been about 0.3 mm/
year, 5000-8000 years ago, but only 0.15 mm/year during the last 2000
years (T Kankainen, personal communication, 1989). This has affected the
anomalies in peat and humus.
Finally in this connection a word of caution: some researchers (e.g.,
Legkova et al., 1980; Nikitichev, 1980; Proskuryakov et al., 1980) in
selecting materials for analysis and in interpreting the obtained results
have made use of the concept of landscape. However, if one connects
together the quality and depth of overburden, topography soil profile and
Anomalies 367

type of vegetation, one is mixing causes and consequences. Already complex


geochemical/geological interrelations are made still more complex through
detailed landscape classifications, and it is only too easy to make the fatal
error of mixing together the analytical results of different materials.
If a single type of sample material is used and this is in close relation to
bedrock, the interpretation of anomalies should not be overly difficult. In
addition to multi-element chemical analysis for complete characterization
of the source rocks, also selective dissolution may be used to determine the
mode of occurrence of elements in the anomaly and the type of transport.
Statistical methods are used to discover faint anomalies and the typical
paragenic associations and relations of elements — the fingerprints of the
source mineralizations.
Perhaps there is no suitable secondary material covering the entire
area. The difference between the geochemistry of mineralization and
the host rock may be slight and weathering, erosion, transport and
sedimentation processes may have confused the original element relations.
If materials not closely related to the source are investigated, even the
most experienced scientific detective may find the anomalies impossible to
interpret without mineralogical studies, because of the overlapping effects
of several processes.
The amount of precipitation, the mean temperature and the vegetation
have varied during geological time, changing the weathering and soil
formation and causing different products at different times (Luckman,
1989; Ovenden, 1989). Volcanoes have complicated matters by spewing
their contents over large parts of the world (Körner et al., 1989). Through
their agriculture, industry and transportation systems human cultures
have been impacting on the natural distribution of elements (Shuts and
Kettles, 1989). In geochemical exploration, all these potential problems
need to be kept in mind, and their effects minimized by well-conceived
sampling and taken into account at the interpretation stage.

ANOMALIES

Every element has a certain average concentration (clarke) with its spe-
cific standard deviation in any particular material. A standard composition
for every material can thus be defined and used in identification of that ma-
terial from a group of samples. Deviations from the standard composition
are called anomalies.
In exploration geochemistry it is usual to consider concentrations greater
than the areal average plus 2 standard deviations to be anomalous. This
simple definition of anomalies is suitable for small area/high density
sampling studies. Sometimes anomalies are true, sometimes "false" (Ra-
manamurthy, 1983), sometimes strong, and sometimes so weak that only
368 Focal aspects of geochemistry applied in arctic and temperate regions
[-100%

Cr NiCuPb 80
f 60

40

h20
+ %
Averages
Cr 232 ppm
Ni 102 - -
Cu 90 - -
Pb 31 - -

Fig. 12-1. Residual anomalies showing copper and lead, chromium and nickel and lead-rich
bedrock horizons on the northeast side of Levi fell, Finnish Lapland.

a summing up of the concentrations of paragenetic elements reveals ex-


istence. Average concentrations of a characteristic set of common major,
minor and trace elements calculated from samples taken from a limited
area can be considered the geochemical spectrum of the assayed material.
Deviations of concentrations of the whole set from this spectrum — the
residuals — likewise constitute an anomaly (Fig. 12-1). Over larger areas
where the chemical composition of overburden materials varies according
to the bedrock beneath, a moving average of concentrations will need to
be calculated to represent the background, and the concentrations above or
below this undulating "surface" are then anomalous.
In one and the same region there may be areas of high or areas of low
concentrations, as well as areas of high variation of concentrations and
very stable concentrations. The boundaries of anomalies are gradual and
a certain flexibility should be allowed for the prospector to use his own
geological knowledge and imagination in delineating them.
The significance of an anomaly can be checked by comparing the
deviation of the elemental association from the bulk chemistry of the
lithologies of the underlying bedrock (Aucott, 1987; Gustavsson, 1983). The
Scale 369

effects of sampling density and the homogeneity of the sampled material,


the reliability of chemical analyses and methods of statistical mathematics
available for the interpretation of single-element and, more usefully, multi-
element data have been widely discussed by Gustavsson in Chapter 10. The
methods recommended for mathematical treatment and visualization of the
results are mainly suited for modern personal desk-top computers.
Discrimination of weak anomalies from background variations for pur-
poses of follow-up studies, and a classification of these anomalies, can also
be done by checking the mineralogical characteristics of the assayed mate-
rial, e.g., fluid inclusions (Kinnunen, 1979) or minor elements in magnetite
(Granath, 1983), or by investigating the isotopic composition, as in the case
of lead (Gulson and Vaasjoki, 1987).
The reliability of chemical analysis depends not only on the concentration
of elements and the homogeneity of the material but also on the sensitivity
and accuracy of the method. These can vary between laboratories or even
within one laboratory (Xie and Zheng, 1983) complicating the setting
of fixed limits to an anomaly. The suitability of different preparation
and analytical methods for a specific element, as well as the standards
needed for different materials and different purposes of study, are discussed
by Noras in Chapter 9. Recommendations for small and medium-sized
laboratories are given.
The amount of metal in a unit volume of assayed material in an anomaly
(concentration) reflects the source deposit (large low grade or small high
grade) and the total amount of metal in the secondary anomaly gives a
measure of the erosion and transport and of course also the magnitude
(grade and the area) of the ore exposure. The ratio of the anomaly to
the source nevertheless depends on so many geological, topographical and
climatological factors that no universal factor can be calculated, even
though such are used in regional ore potential assessments (Solovov, 1959;
Rundqvist, 1979).

SCALE

A reconnaissance survey with, say, one composite sample per 100-500


km 2 is recommended over the whole region as the first step in a geochemical
study; it is especially important that this be done in terrains for which
geological or geophysical maps are not available. To permit the delineation
of anomalous zones of 2000-10,000 km 2 the survey should cover the region
of a geochemical province, say 20,000-100,000 km 2 .
From the geochemical map or from the existing geological or geophysical
maps, promising zones for areal geochemical mapping with a sampling
density of one composite sample per 1-10 km 2 can then be selected.
Anomalous formations 100-1000 km 2 in size can be delineated from areal
370 Focal aspects of geochemistry applied in arctic and temperate regions

geochemical maps. Several materials and analytical procedures and a large


selection of elements should be tested at this stage. Stepwise study is
the most cost-effective exploration method if the variations caused by
mineralized rocks can be (see Chapter 10) detected from the variation of
concentrations caused by barren lithologies with sparse sampling.
The most promising formations revealed in the areal mapping can be
studied through local sampling on a grid at least ten times denser (5-50
samples/km 2 ). Using the best secondary material available (preferably soil),
anomalies of 0.2-1 km 2 can be revealed.
The local scale studies are completed with the final stage of geochemical
study, detailed target exploration, in which the most promising anomalies
are assayed on a 5-50 m sampling grid. Since target exploration is aimed at
pinpointing sites for diamond drilling, samples should be taken from a vari-
ety of depths at selected points, and various grain size fractions of selected
samples should be analyzed by different methods. The criteria choosing
of materials, sampling densities and sampling methods are discussed by
Salminen in Chapters 7 and 8. Recommendations for different scales of
study are given. A good example of a geochemical study proceeding step by
step is given by Salminen and Hartikainen (1986).
The geochemical mapping of Finland (338,000 km 2 ) was begun in the
1970's using till with a sampling distance of 100 m along lines 1000 m
apart, perpendicular to the main direction of till transport (Kauranne,
1975). According to Savinskii (1965), with 80% probability this system will
reveal anomalies 150 x 1000 m in size with two anomalous points. Mapping
was proceeding slowly when a new, more rapid system was adopted in
1980's: one composite till sample/300 km 2 or corresponding ground water
sample from different types of wells over the entire land area of the country,
with each sample representing a catchment area.
Such sparse sampling was encouraged by the successful experiments with
stream sediments (1 sample/200 km2) in Sierra Leone (Garrett and Nichol,
1967) and Zambia (Armour-Brown and Nichol, 1970), with heavy minerals
of till in Sweden (1 sample/300 km 2 , Brundin and Nairis, 1972) and with six
different materials in northern Fennoscandia (1 point/30 km 2 , Bölviken et
al., 1986). The 1/300 km 2 density study initiated by Björklund in the 1980s,
and where the results are visualized as weighted average concentrations of a
moving window (Gustavsson and Kontio, 1988), reveals the regional trends
and lithological zones with diagnostic element association. The system has
allowed quick preparation of national geochemical atlases (groundwater,
Lahermo, 1990; till, Koljonen, 1992).
The "national" geochemical mapping at 1 sample/300 km 2 is being con-
tinued with "regional scale" mapping where 1 random sample (consisting
of 9 subsamples) is taken per 4 km 2 . The "regional scale" density reveals all
significant features of bedrock lithology and most geochemical anomalies
of 150 x 1000 m size (Koljonen, 1992). The follow-up exploration is done
Sampling material 371

in two or three successive phases. The same final procedure has been used
by Horsnail (1975), who recommends for Ireland rectangular sampling on
250 x 125 m and 50 x 50 m nets.
In work at scales greater than 10 samples/km 2 , maximal results will
be achieved with minimum cost if the following factors are taken into
account in planning the sampling grid: strike of bedrock formations (see,
e.g., Savinskii, 1965; De Geoffroy and Wu, 1970), the stratigraphy of
sampled material and its transport directions (e.g., DiLabio, 1989; Hirvas
and Nenonen, 1987; Shuts and Smith, 1986; Stea et al., 1986), grain size
and inhomogeneity of material (Gleeson et al., 1988; Nichol, 1986; Shuts,
1973; Xie and Zheng, 1983) and pollution (Qvarfort, 1979).

SAMPLING MATERIAL

Sampling is the most important of all stages of geochemical exploration.


Samples must be representative, relationships of their material to the
target must be known, they must be easily obtainable and results of their
analysis must be comparable (see Chapter 10, "Quality control").
Bedrock with its ores consists of lithologies which are composed of stable
chemical compounds, the minerals. Overburden formations consisting of
mineral materials chemically resemble the bedrock, the more so the closer
it, the less sorted and the less weathered they are. Some overburden
formations are on the site of their source formation, while some have been
mechanically moved and some chemically transported. In the last case a
part of the chemical constituents will have been transported to sea, the
other captured from ground-water solution into organic and mineral soil
deposits. Organic overburden formations differ sharply both chemically and
physically from the bedrock, but their heavy metal composition continues
to reflect the mineralization of the bedrock.
Materials having an intimate relation to the source are more suitable
for geochemical exploration than those that have been chemically altered
or far transported. The importance for geochemical exploration of a thor-
ough understanding of the composition, stratigraphy, transportation and
provenance of the material to be assayed cannot be overstressed (DiLabio
and Coker, 1989). To this end, additional geological and related studies of
materials in every field of geochemical exploration can be recommended.
The anomalous concentrations found by geochemical target exploration
are a product of (1) chemical processes associated with weathering in which
released ions are moved in gas or water phase or as "dry" ionic diffusion
(dispersion aureoles, Polikarpochkin et al., 1983), or (2) clastic processes in
which mineral grains are moved by gravity, frost, ice or water streams (dis-
persion flow; Polikarpochkin et al., 1983). Clastic and chemical transport,
two-vector dispersion, can take place either simultaneously or successively,
372 Focal aspects of geochemistry applied in arctic and temperate regions

complicating the differentiation and correspondingly the interpretation.


Examples of this kind of multiple transport and the interpretation are
given Chapter 3 for Paskolahti (Fig. 3-7) and in Chapter 11 for Maaselkä
(Figs. 11-20 to 11-26) for Porte aux Moines (Figs. 11-56 to 11-58) and for
Petalax (Fig. 11-69). Similar forces and media of transport act differently
on different materials, so that sample type must be selected and sampling
done with the greatest of care to ensure comparable results.
Whatever the sample density or scale of study, usually a meaningful
pattern of element distributions will be found (Björklund and Gustavsson,
1987) in almost any material assayed. The suitability of study material
depends not only on climate and overburden but also on the scale of
study. For country-wide, regional and areal geochemical mapping and for
delineating of geochemical provinces, zones or formations, a material well
homogenized and possibly far transported may be good. Stream and lake
sediments, bog rim peat and B-horizon soil may be used. For local and
target exploration and especially for drilling point location, a material
closely related to bedrock must be chosen. The right material for the
latter purpose is C-horizon mineral soil, preferentially with depth profile
sampling, or in some cases humus. The results obtained with sorted clastic
sediments have so far not been encouraging, with the exception of the
present watercourse bottom sediments.

Residual mineral formations


Residual formations are deposits developed in situ through mechanical
or chemical disintegration of rocks. Usually both processes go on simul-
taneously, making it impossible to draw absolute boundaries between the
soils developed in one or another of these two ways. On top of the in situ
weathered regolith there may be layers of organic deposits or transported
mineral drift usually with some sort of soil profile. Deeper down the regolith
is tougher and denser, gradually transforming into solid bedrock.
On the surface of rock exposures in the Arctic, and on mountain tops
in temperate regions, stony material is mechanically loosened from the
bedrock. Mechanical disaggregation is most intense under dry conditions
where daily temperature changes are great and rapid; crumbling is acceler-
ated by freezing and swelling of the moisture in cracks. Gravel developed
from readily disintegratable coarse lithologies (e.g., from rapakivi granite)
has been shown to be chemically almost identical with the parent material.
Any anomalies in such materials lie directly on top of the mother lode.
In temperate and more humid climate, rocks are predominantly chem-
ically weathered; some minerals (sulphides, micas, carbonates) are easily
leached and their constituents are either transported away by groundwater
or become trapped in secondary minerals (oxyhydrates, clay minerals).
The weathering of Quaternary deposits, e.g., till depends largely on their
Sampling material 373

petrological composition (Hall, 1981). The mobility of elements depends on


the pH and Eh of pore water, as described by Salminen in Chapter 5 and in
many textbooks (e.g., Garrels and Christ, 1965; see also Blain and Andrew,
1977). The leached elements are separated from their original association
and some of them may be enriched into certain layers or minerals (e.g.,
copper in mica, Puustinen, 1977).
Chemically weathered rock differs from the original in both physical
appearance and chemical composition. Acid-type weathering of arctic and
temperate regions produces silica-rich resistates or quartz sand. In warmer
climate the less acid weathering leads to aluminium-rich resistates and the
formation of true clays. The humus layer binds considerable amounts of
metal, while the pale A-horizon underneath is poor in almost all cations
except silica, aluminium and potassium, and sometimes lead in sulphate,
as described in Chapter 4. The smallest mineral particles like finely ground
mica, newly formed clay minerals and organic or mineral colloids scavenge
heavy metal ions migrating downwards with infiltrating ground water.
Oxidation is often quickly followed by dissolution. Sulphides, for example,
are easily oxidized and the released heavy metals are picked up by water,
but iron is soon precipitated as limonite coating on grain surfaces or as
filling in voids. Other metals, cobalt for example, are coprecipitated, while
still others continue their journey. This leads to chemical zonation of the
weathered regolith, where some zones are poorer and some richer than the
original rock in a particular element.
Anomalies in weathered bedrock are mainly hydromorphic and usually
metals in such anomalies occur in easily soluble form. The chemically
resistant minerals and their elements are enriched in the residuum. In size
the hydromorphic anomalies tend to be somewhat larger than their original
source. The decreased concentration of mobile elements can be called for
negative anomalies. Residual and negative anomalies have not been moved,
while positive ones are elongated in the direction of water movement.
Chemical alterations are often easily recognized from their "unnatural"
colours, as bleached or brownish red horizons on a valley side.
A typical anomaly in residual formations, but occurring also in trans-
ported drift, is the "iron cap", or gossan, on the surface of soil above
or near by a sulphide mineralization. Sometimes "iron pan" is found at
the level of the ground-water surface, where limonite cements the loose
mineral particles into a hard crust. More often hydromorphic anomalies
are hidden in the regolith, right under one's feet but detectable only by
chemical analysis.
Weathered rock is a good material for geochemical local and target scale
exploration where dense sampling laterally and some sampling vertically
needs to be done. Sequential analysis of selected samples is required
for study of the mode of occurrence of the metals, without which the
interpretation of results may be difficult. In contrast, the geochemical
374 Focal aspects of geochemistry applied in arctic and temperate regions

inhomogeneity of weathered bedrock makes its use in regional geochemical


studies of questionable value; its analysis for purposes of regional geological
mapping has been examined by Shepherd et al. (1987).
In nonglaciated temperate zone the in situ weathered regolith is the
most common type of overburden, although often covered by transported
mineral sediments or organic deposits. The main part of the literature on
soil geochemistry deals with just this type of deposit (e.g., Rose et al., 1979;
Levinson, 1974; Ginzburg, 1960; Polikarpochkin et al., 1983). In glaciated
terrain residual deposits are met with mainly in depressions, in the lee of
the glacial abrasion and in areas of gentle abrasion at the former ice divide
or the marginal zone of the glaciation.
Mixing of preglacially weathered material with till usually increases the
heavy metal concentrations of till (Kauranne et al., 1977). Pure weathered
rock material, if sampled simultaneously with till, should be presented on
separate maps and the results evaluated individually.

Transported drift
The geochemical character of transported overburden formations de-
pends on the type of transport and mode of occurrence of their constituents.
Disintegration of rocks and the separation of minerals and chemical ele-
ments from each other depend on the transporting medium, whether this
be gravity, pushing glacier, flowing water or air, as described in Chapter
3. To a large extent transport occurs mechanically as grains, but in water
also as ions or molecules. To understand fully the geochemical character of
a material and the provenance of anomalies in it, the way in which it has
been formed needs to be known. Thus Chapters 2, 3 and 4 of this volume
have presented a broad description of the formation of glacigenic and other
types of overburden deposits and soils.
Talus represents the original bedrock, which has simply been loosened
mechanically and transported freely by gravity down along the mountain
flanks. Only some separation of rock types according to form and size of
the fragments occurs during the rolling and sliding of the material.
Most glacial till is likewise chemically fresh, although preglacial materials
— chemically altered crust, sorted sediments and even organic deposits
— may be intermixed. During glacial transport minerals are comminuted
and elements differentiated into different grain sizes and lithologies are
described by Eriksson in Chapter 2 and by Salminen in Chapter 5 (see also
Andrews et al., 1983; DiLabio, 1982; Hyvärinen et al., 1973; Minell, 1980;
Peuraniemi, 1989; Saarnisto and Taipale, 1985; Salminen, 1980; Shuts,
1975; Shuts and Smith, 1989; Toverud, 1984).
The main mass of till has typically been transported only some hundred
metres (Kauranne, 1976b), but an identifiable part may have been trans-
ported tens (Batterson, 1989; Perttunen, 1977) or even hundreds (Shuts,
Sampling material 375

1972; Schuddebeurs, 1981) of kilometres. Stones and boulders form disper-


sal trains of variable length and width as described by Salonen in Chapter
6. Particle direction measurements have also been described, e.g., by Glen
et al. (1957), Kauranne (1960a), Boulton (1971), Puranen (1977) and East-
erbrook (1981) and the stone counts, e.g., by Okko (1941), Dreimanis and
Vagners (1971) and Perttunen (1977). Chemically detectable anomalies in
the finer grain size fractions of till are usually only some hundreds of
metres long — the main part of the till material obviously has not been
transported far (Chapter 5, see also Kinnunen, 1979).
Sometimes till has been transported in several successive stages (Geddes,
1982; Hirvas and Nenonen, 1987; Klassen and Thompson, 1989; Stea et al.,
1989), and anomalies in different till sheets are then partially overlapping.
14
C-age determinations of the organic material found between different
till sheets have shown that the ice-free intervals between glaciations in
some cases have lasted thousands of years (Korpela, 1969). Weathering and
soil formation have thus had time to alter the chemistry of the already
deposited formations. If the material of the older sheet is then mixed with
younger drift during a new glacial advance and new direction of transport,
the anomalies in the new till bed will be patchy and broad. In regional and
areal geochemical mapping, when samples are taken from shallow depths
of the surface till bed, the anomalies of deeper till sheets may be missed. In
local and detailed geochemical exploration, several depths and even several
fractions of material should always be sampled (Boyle and Troup, 1975).
A good summary of experiences in explorational till geochemistry can be
found in Gleeson and Nichol (1987).
Glacial ice is highly viscose and, unlike water and wind, does not
separate the particles it transports according to grain size and density. In
the "glacial mill" minerals are instead comminuted according to strength,
with the weakest ones concentrated into the finest fractions (Kauranne,
1967; Dreimanis and Vagners, 1971; Salminen, 1980;).
In glaciated areas the clastic sorted sediments (cobbles, stones, grains),
transported and sorted by water and wind have obtained a considerable
part of their material from till, but also contain large amounts of pre-
and interglacial sorted materials as evident in the differences of grain
size fractions. In nonglaciated areas their source is the weathered crust.
Streaming, standing and waving water is an effective separator of mineral
grains. Wind easily retransports the finer mineral fractions, further sorting
the mineral grains according to their grain form and weight (Chapter 3).
Direction of transport is easily determined if transport has occurred in
just one stage but in the case of multiple till it has occurred in several
directions and in sorted sediments the material has moved in several
successive stages. The length of the transport is always difficult to estimate,
and still more difficult for sorted sediments than for till. Roundness of the
particles and their lithology (Hellaakoski, 1930; Okko, 1941; Matisto, 1961;
376 Focal aspects of geochemistry applied in arctic and temperate regions

Virkkala, 1971; Perttunen, 1977) or chemical composition (Granath, 1983)


may indicate the provenance. In contrast to till, the finer grained the sorted
materials the farther they have been transported (Kauranne, 1976b; Miller,
1984). Particles of coarse sorted sediments consist mainly of different
lithologies: sand size sediments consist of monomineralic grains, with
quartz and feldspar dominating, while in finer grained sediments micas are
most abundant. In stream and beach sands and even in wind blown dunes
there may be enrichments of certain heavy minerals (placers), but the very
fine airborne dust is mineralogically poorly sorted.
The total grain surface in a unit volume (specific surface) of fine
sediments is much larger than in a unit volume of coarse fractions, with
the result that the fines are much more sensitive to chemical attack by
pore water. Sulphide and carbonate grains are easily dissolved and parts
of their constituents soon again precipitated. Heavy metals like cobalt and
zinc tend to coprecipitate with iron and manganese, oxyhydrated either as
a thin film on the silicate grain surfaces or in the voids. In silt, loess and
clay a hygroscopic rise of metal-rich water solution may bring anomalies to
the surface (Lukashev and Lukashev, 1977; Smee and Sinha, 1979). This
means that elements transported and sedimented as mineral grains may
be hydromorphically redistributed postdepositionally (e.g., Maceahern and
Stea, 1987).
Materials transported in ionic form in water are separated according to
their affinity to other elements through reduction and oxidation reactions
and precipitation, all depending on variations in pH and Eh of the solution
(see Chapter 5). Voluminous deposits of iron oxyhydrate (gossan) are often
encountered above or in the vicinity of sulphide mineralizations or sulphide
rich schists. Gossan has been widely used in geochemical exploration
(Blain and Andrew, 1977; Laville-Timsit and Wilhelm, 1979). Comparison
of the chemistry of gossan with that of sulphide mineralizations, e.g.,
by lead isotope studies (Gulson and Vaasjoki, 1987), helps in selecting
the most promising gossan deposits for follow-up studies, as described in
Chapter 3.
The chemical migration of elements in overburden occurs downwards
with percolating rain water, laterally with ground water, and upwards with
hygroscopic water, with evaporating gases and by electrochemical potential
(Goldschmidt, 1934; Sveshnikov and Ryss, 1964; Govett, 1973; Bölviken
and Logn, 1975). Heavy metals migrating upwards to the surface of sorted
sediments may be captured by organic compounds of humus or peat.
Although till is a clastically transported material it may also contain
hydromorphically transported elements. These can be discriminated from
the clastically transported ones by sequential analysis and do not hinder
the use of till for geochemical exploration. Nurmi (1976), for example
calculated the hydromorphic portion of the total anomaly in till close to the
Talluskylä nickeliferous deposit to be Cu 10%, Co 45% and Ni 10%.
Sampling material 377

Purely chemically transported anomalies are sometimes fairly close to


their source like gossan and can easily be used for geochemical exploration.
Certain very mobile elements (e.g., uranium) may be transported much
longer distances and tracing of their source has been difficult (see, e.g., the
case history from Masugnsbyn, U in peat; also Armands, 1967; Michie et
al., 1972; Lett and Fletcher, 1978).
Due to upwards migration of elements it should be possible to use
clay as a sampling medium, as investigated experimentally by Smee and
Sinha (1979); loess, too, has been used for geochemical studies — both
explorational and environmental — with the help of a special ion trap
system (Lukashev, 1983). Sorted coarser glacifluvial mineral sediments
have been found mainly to represent the average lithology and chemistry
of the tills from which they were derived.
At the same sampling points of some 100 km 2 map sheet areas in
central Finland the Pearson correlation coefficient between the elemental
concentrations in the fines of till and in the fines of sorted glacial sediments
was 0.5-0.7 for major elements, 0.3-0.6 for minor elements and 0.3-0.9 for
trace elements (Kauranne et al., 1977). The average chemical composition
of glacilacustrine silt and clay may be the same as that of the fine fractions
of till of a given area (DiLabio, 1989), but in larger clay deposits may also
be totally different (Kaszycki and DiLabio, 1986).
The grains of possible clastic anomalies in sorted glacial sediments will
have been transported along different paths, first in the ice and then in
the water, which makes it very difficult to interpret their provenance. For
this reason, varved sediments and sands have been shunned as a sample
medium (DiLabio, 1989).

Organic material
Typically a thin layer of organic material is found in dry places and
a thicker layer in moist places at the surface of the overburden. After
decaying of the organic tissue such material turns into humus or peat as
described in Chapter 3. The C/N ratio in plants is over 15, whereas it is in
humus under 12, showing the trend of the alteration.
Humus consists of fresh and decomposed organic litter, mainly plant
remnants, but also contains resistant relics of insects or other animals
and living micro-organisms like microbes, fungi mychorriza and chrysales.
The totally humified part of humus consists of humic and fulvic acids
and their salts and chelates in which the heavy metals are fixed and
enriched. Humic acids and their salts are high molecular weight, base-
soluble compounds of large ion exchange capacity. Fulvic acids are smaller
molecular weight water-soluble compounds, which have a clay clumping
and water retention capacity important for agriculture. Antropova (1975)
claims that humic acids are the main collectors of heavy metals and that
378 Focal aspects of geochemistry applied in arctic and temperate regions

only they should be analyzed in exploration. Analyzing this fraction she


reports (N.T. Antropova, pers. commun., 1982) obtaining the signal of an
ore deposit through 150 m clay and 350 m limestone (altogether 500 m). The
metal fixation capacity of humus depends on the type of vegetation, e.g.,
whether the forest is coniferous or deciduous (Kauranne, 1967). Humus is
the first object of attack of acid rain and metals bound in it may become
soluble in toxic amounts.
Humus is an easy material to sample and is easily analyzed on site
for bulk heavy metals (Bloom, 1955). It has been widely used as a
sampling medium in exploration geochemistry, although its inhomogeneity
and instability somewhat impair its usability (e.g., Vogt and Bergh, 1943;
Govett, 1973; Kokkola, 1977; Nuutilainen and Peuraniemi, 1977). The
levels of heavy metals in humus have been found to vary with the season
(Salminen and Kokkola, 1984) and in peat from year to year (Gleeson et
al., 1989), but the anomalies persist.
In areas of poor drainage, humified organic material builds up into
thick layers of peat. In arctic and subarctic regions where bogs are widely
occurring, peat layers are typically 0.5 to 2 m thick. In the more localized
bogs of temperate regions they range from 0.5 to 10 m, and in bogs in the
tropics are up to tens of metres thick.
The texture, composition and degree of humification of peat depend
on the original vegetation. Tanskanen (1976) did not find any significant
variation in the heavy metal concentrations with type of peat, but the
sampling depth and ash content were relevant.
Peat has been used for both regional and target exploration. Eriksson
(1976) found peat samples from the outer rims of bogs to be very useful in
regional (first stage) prospecting, which then was followed up by boulder
tracing and till geochemistry. Salmi (1967) observed that heavy metal
anomalies in peat have their source in ore bodies directly below, some
metals being enriched in the bottom part of the mire, some in surficial
parts and some located indifferently.

Soil profile
During weathering and soil formation, minerals are decomposed and
part of their elements are dissolved and transported down by percolating
rain water (or up by hygroscopic water). Some of the metals are almost
immediately precipitated, while others get caught in resistant primary
or secondary minerals. The various layers in the soil profile are thus
impoverished or enriched in some elements (Chapter 4).
The humus in podsol and certain other soil types is usually directly
underlain by a bleached mineral horizon A, which is enriched in silica and,
in warmer climates, in aluminium. Lead appears to be enriched as well
(Kauranne, 1967). No enrichment of Pb or Zn seems to occur (Horsnail,
Sampling material 379

TABLE 12-1

Average exchangeable (BaC^-soluble) element concentrations of cultivated fields in southern


Finland (Tamminen and Starr, 1990), podsolic surface (0-30 cm), layering destroyed

Layer pH aq pH ba Al Fe Mn Cu Zn Cd Pb Cr Ni
(%) (%) (%) (ppm) (ppm) (ppm) (ppm) (ppm) (ppm)
Humus 4.17 3.15 18.9 4.4 10.1 7.4 54.7 0.47 33.8 9.4 5.1
0- 5 cm 4.35 3.45 22.7 2.0 0.30 3.3 22.9 0.24 16.7 12.9 0.04
5-20 cm 4.94 3.76 9.3 0.49 0.18 4.0 28.1 0.32 20.6 19.9 0.13
20-40 cm 5.25 3.76 3.9 0.10 0.05 5.9 24.0 0.32 18.7 21.5 0.70
60-70 cm 5.56 3.76 1.7 0.05 0.01 7.8 19.0 0.28 14.2 18.3 0.72

1975) directly under humus in brown earth soil profile. The A-horizon
has been used experimentally for geochemical exploration, but evidently
without much success since it has not become popular in spite of easy
and inexpensive sampling. The great deviation in results probably depends
on the varying content of organic compounds and the thinness of the
A-horizon.
The boundary of the A-horizon towards the rusty B-horizon is usually
well developed. The B-horizon can be divided into subhorizons too, though
this is not necessary for purposes of geochemical prospecting. The B-horizon
is clearly enriched in Fe, Mn, Co, Zn and Al with minor enrichment of
many other elements. In temperate zones it has become a very popular
material for geochemical exploration since it is easily sampled from about
0.5 m depth and easily recognized by its colour. Thickness varies according
to grain size: in clay it may be a few centimetres, in sand as much as a
couple of metres thick. The border towards the C-horizon is transitional.
No real C-horizon is found in residual soils because the material is
weathered down to bedrock. A virgin C-horizon occurs in transported
mineral soils, where the minerals, even sulphide grains, are fresh in
the ground-water regime. An iron pad layer cemented with limonite is
sometimes encountered at the level of the ground-water table. Above this,
sulphides are weathered, metals soluted and precipitated, clastically and
hydromorphically transported anomalies are superimposed.
Table 12-1 shows the variation of concentrations of the easily BaCl2-
leachable metals with depth in the soils of southern Finland (Tamminen
and Starr, 1990).
Although sampling of the C-horizon is more tedious than that of the
B, it is nevertheless the most popular sampling medium for geochemical
exploration in glaciated terrain. In detailed target exploration it is advisable
to analyze several soil horizons and to take some vertical sample series
through the whole overburden down to the surface of the bedrock.
380 Focal aspects ofgeochemistry applied in arctic and temperate regions

SAMPLING DENSITY AND DEPTH

Density
A sparse shallow sampling of till or other widely distributed overburden
materials is sufficient in regional and areal scale geochemical mapping, but
denser and deeper sampling is required in local and especially in target
exploration (see Chapter 7). The smaller the target sought, the denser
the sampling must be and the weaker the signal (anomaly) given by the
target, the larger (or more homogeneous) the samples should be. The
representativeness of the samples can be improved by combining several
subsamples, and the interpretation improved by analyzing a whole set of
related elements (Björklund and Gustavsson, 1987).
As a rule of thumb, in regional (over 100,000 km 2 ) reconnaissance
work one composite sample (consisting of five subsamples from an area of
perhaps 400 m 2 ) per 100-500 km 2 is sufficient. A widely distributed mixture
of bedrock materials — stream sediment or till (e.g., the B-horizon) — can
be recommended as material. In areal (1000-10,000 km 2 ) geochemistry one
composite sample of the above materials or peat or humus per 1-10 km 2
can be recommended.
For local scale (10-100 km2) exploration a denser grid of samples, say
5-50 samples per 1 km 2 , is needed. A short-transported nonsorted sediment
or residual mineral material or humus may be recommended. If till is used
it should be sampled from the C-horizon — from a depth of about 1 m
if there is only one till bed in the area. If there are several till beds, the
stratigraphy, the transport direction and length may decide which is the
optimal bed to be sampled.
Still denser sampling is needed for detailed target exploration (<1 km 2 ).
Both till, bedrock and weathered rock samples can be recommended. Some
samples of every target should be taken as a series through the whole
overburden to permit explanation of the shape and mode of origin of the
anomalies and pinpointing of the drilling target.

Depth
The chemical composition of the surficial layer of residual soils, even if
adjusted for any differences between the possible soil profile layers, does not
necessarily resemble that of the unweathered rock beneath. If the region
happens to be anthropogenically altered by acidic precipitation, fertilizers
or other pollutants, the chemistry of the surface layer may be very different
from that of the underlying bedrock. But the deeper one goes, the closer is
the chemical connection with the fresh rock.
The same is true of relatively short-transported formations like till.
The deeper one goes, the closer one is to the bedrock both physically and
Sampling density and depth 381

t pt
• VIHANTI
MINE
·

' ■ · ·,·
Y
f ►
\ Zn in
ORGANIC MATTER

· · · . under 43 ppm
0 ^ 5km 1 i ^ 43-83 ppm
y ES3 over 83 ppm

Fig. 12-2. High zinc concentrations in stream sediments in the Vihanti river due to contamination
from a flotation waste pond. Redrawn with permission from L-M. Kauranne (1979).

chemically until at the till/rock interface the material is the same, if one
doesn't count the thin layer of quartz-rich "rock flour" often encountered
at the interface (Chapter 2).
Contamination transported by wind as sulphide mineral dust has been
traced down to a depth of 0.5 m in till (Qvarfort, 1977) in villages
where mining and smeltering was done 200-400 years ago. In stream
and lake sediments, traces may be found many kilometres away (30 km,
L-M. Kauranne, 1977; 10 km, Salminen, 1981; 5 km, Qvarfort, 1977). For
example, in the Vihanti river the contamination from a flotation waste
basin could be traced over 30 km in analysis for zinc; the other heavy
metals became fixed more quickly (Fig. 12-2).
Some element concentrations appear to vary with sampling depth even if
the overburden is formed of only one layer of till. In central Finland, in the
overburden above mica schist bedrock, there is a slight increase of average
Mn, \ζ Cu and Zn contents downwards, while the concentrations of Cr, Co
and Ni seem to have no correlation with depth (Kauranne et al., 1977).
By contrast, Pb tends to be highest close to the surface. These patterns
may indicate that the upper horizons are somewhat diluted with granitic
material. In northern Finland concentrations of all these elements increase
strongly downwards in till, not only above basic but above granitic rocks
as well. Possibly in northern Finland there is more preglacial weathered
material intermixed in the till deeper down. See the following Table 12-IIA.
Approximately the same thing applies to sorted sediments (Table 12-IIB).
In central Finland the variation of element concentrations with depth is
very slight, whereas in northern Finland it is marked. In more detailed
geochemical surveys it is important to study the relationship between
concentration and depth to be able to choose the optimal sampling depth
and to interpret the obtained results.
382 Focal aspects of geochemistry applied in arctic and temperate regions

TABLE 12-11

Average total concentrations (ppm) of geochemical samples taken over different lithologies

Depth (m) N Mn V Cr Co Ni Cu Zn
A In till at different depths
Central Finland
Mica schist <1 327 676 372 184 25 106 115 24
1-2 430 781 392 191 30 132 128 36
2-3 19 902 427 197 35 151 132 46
>3 10 847 349 186 28 114 135 72
Northern Finland
Granite <1 238 553 284 371 22 141 49 9
1-2 183 821 374 560 31 238 86 20
2-3 65 882 442 691 37 296 107 32
>3 76 783 478 744 38 320 119 33
Gabbro <1 281 548 301 424 26 177 55 4
1-2 292 625 282 654 37 319 119 10
2-3 101 639 300 740 40 360 179 15
>3 99 690 300 791 42 364 123 18
B. In sorted sediments at different depths
Central Finland
Mica schist <1 18 846 299 153 20 78 73 19
1-2 66 833 333 171 26 97 121 27
2-3 23 728 338 179 20 81 110 32
>3 19 681 336 183 22 111 87 53
Northern Finland
Granite <1 116 551 254 394 22 142 48 11
1-2 51 654 293 485 26 184 58 16
2-3 5 602 470 724 31 249 102 37
>3 4 786 447 826 35 353 105 26
Gabbro <1 11 634 301 364 23 150 64 7
1-2 26 805 300 458 29 186 98 17
2-3 15 594 298 514 32 220 264 29
>3 33 650 349 580 38 237 176 33
C. At the surface of weathered rock under varying thickness of overburden
Central Finland
Granodiorite <1 2 943 285 147 75 83 91 33
1-2 3 947 247 147 43 102 62 39
2-3 1 1300 162 642 28 279 12 106
Northern Finland
Granite <1 18 510 134 126 15 54 24 9
1-2 6 814 145 253 24 96 66 16
2-3 7 1046 269 373 32 145 59 19
>3 5 1035 116 127 21 60 34 14
Gabbro <1 42 638 268 772 35 323 119 4
1-2 109 596 217 634 34 293 132 6
2-3 55 715 292 602 34 229 170 10
>3 64 725 223 607 38 274 174 11
Sample preparation and analysis 383

The commonly used Irish-type sampler does not penetrate (weathered)


bedrock, so that the depth given in Table 12-IIC does not mean totally
the same thing as depth in the A and B tables. It refers often only to the
thickness of overburden to the surface of weathered bedrock. The values
shown for weathered bedrock in Table 12-IIC do not really show a relation
between element concentrations and depth.

SAMPLE PREPARATION AND ANALYSIS

As described in Chapter 9, sample preparation and analysis are almost as


vital for successful exploration as the sampling. Through ashing, sieving,
grinding, fusion and solution operations the desired form of occurrence
of the elements is obtained for analysis. Preparation procedures affect
the results directly. Therefore, both pretreatment and analysis must be
strictly standardized for obtaining comparable results and "maximation of
information recovery for the dollar spent" (Closs, 1987).

Sample preparation
The elements of interest may occur in the assayed material in many
forms. To achieve the highest concentration together with the greatest
anomaly/background contrast and most easily interpretable results, one
must consider which mode of occurrence of the sought metal is best for
analysis and on that basis choose the most suitable way of preparing
samples. The most common modes of occurrence of elements are in their
own primary mineral grains, in their own secondary minerals, as coating on
other mineral grains, absorbed in colloids, in organic tissue or in pore water
solution. It is also important to study how the elements in the material
occur to know how it has been transported from the mineralization to
the point of sampling (Nikkarinen et al., 1984). Depending on the method
of preparation, the modes are either separated or mixed homogeneously
together before chemical analysis.
Usually samples are dried by warming before weighing them for analysis.
Elements in solution then become fixed onto the surfaces of grains or
absorbed into colloids. In dry matter from which water has been evaporated,
the elements are rather uniformly distributed. Drying can also be done by
other methods, e.g., centrifuging, which removes the main part of water
along with the elements in solution.
Materials rich in organic material are ashed. Most elements become
oxidized and water soluble, though some of the lighter elements are partly
lost to the air. Hard mineral grains are weakened or broken. Degradation
of organic material can also be achieved by wet oxidation, but the process
tends to be tedious and time-consuming.
384 Focal aspects of geochemistry applied in arctic and temperate regions

To increase the solubility and homogeneity, samples are sometimes


crushed and ground, causing the physically weak minerals to be concen-
trated into fine fractions. In fact this has already happened in nature, and
the greatest amounts of heavy, chalcophile metals will be found in the finest
fractions.
More often samples are simply sieved, to separate the finest fractions for
analysis. Higher concentrations and greater anomaly/background contrast
are obtained simultaneously. Elongated and flaky minerals can even be
separated from the main mass by sieving (e.g., asbestos from till, Halonen,
1967) and visually "analyzed". The very finest grain sizes can be separated
by mixing the sample with water, allowing the coarser fraction to sedi-
mentate (or accelerating sedimentation by centrifuging) and decanting the
suspension to be ultrafiltered. In the clay fraction of till, metals from far
away mineralizations are evenly distributed and can be traced by compar-
atively sparse sampling (Shilts, 1973). It is important to know in which
fractions the element of interest occurs in order to select the right ones
for analysis. Gold, which usually occurs as nuggets, may behave peculiarly
in sampling and analysis (Boyle, 1979; Nichol, 1986; Sopuck et al. 1986)
and its mode of occurrence in sample material should be studied before
embarking on larger geochemical operations.
Note that some further separation of heavy particles may occur during
sieving and shaking of the sample. It has also been observed that heavy
minerals (e.g., gold) tend to settle to the bottom of the sample container
during transport of oven dry or oversaturated samples.
Although drying increases the clumping of materials rich in clay or
colloids, this must usually be done. Ashing causes losses of volatile elements
and is better avoided. Floccules need to be gently crushed by hand
after drying, but heavier crushing breaks the grains and causes noise
and dust. Sieving is preferable to heavy crushing, and is also more
suitable for exploration of sulphides, which occur in the fines. Some hard
oxide ore minerals, in contrast, occur in the sand size fraction and will
escape detection if grinding does not precede analysis. Centrifuging and
ultrafiltering will be necessary to separate the clay fraction, because simple
decanting takes too much time.
Some elements occur in such small concentrations that no method of
analysis is able to detect them without pre-enrichment. Heavy minerals
can be panned out or separated by heavy liquids or shaking table. Magnetic
minerals can be separated first by hand magnet and then by magnetic
separator. Other, more sophisticated separation methods suitable for field
use are also available. Separation increases the sensitivity of analysis, or
allows the amount of concentrate to be used directly as a measure of the
ore material in the sample, e.g., in gold or tungsten exploration.
Sample preparation and analysis 385

Analysis

It is most important in exploration geochemistry that the results be


comparable. This has been stressed above in the context of selection of
the sample material and the sampling and sample preparation procedures.
Elements of interest usually occur in the sampled materials in very
low concentrations. Moreover, successful geochemical exploration requires
the collection of many samples over a large area. All these demands
presuppose sensitive, selective, simple (cheap), rapid and reliable methods of
analysis.
Heavy mineral separation by panning is a suitable method of analysis
only for elements like Au, W and U, which occur as very heavy minerals,
are easily identifiable, and are present in silt size or coarser fraction. Gold
in till is often concentrated in the <63 μηι fraction and is lost both in
panning and in reverse circulation sampling (Nichol, 1986; see also Sopuck
et al. 1986 and Huhta, 1988). A thorough examination of the mode of
occurrence of gold and platinum group elements showed (DiLabio, 1982)
the grains in till to be somewhat smaller in size than those in the parent
rock.
Sometimes gold is coated with a rusty film, and sometimes it contains
large amounts of copper that darkens the appearance, but mostly it is
relatively pure, or contains silver and is shimmering yellow. Its shiny
grains are easy to recognize even in silt size, though grains of that
size are otherwise hardly visible to the naked eye. The same applies
to minerals like scheelite and wolframite that are strongly luminescent
under ultraviolet light. All the identifiable grains in a certain volume of
sample can be visually counted, their sizes evaluated and the concentration
roughly calculated (Sopuck et al., 1986). Cassiterite, chromite, haematite
and garnets are also easily separated by density-based methods, but their
identification requires experience. As a rough estimate of other minerals
present, one can measure the percentage of the heavy fraction, the bulk
susceptibility of the sample, the weight per cent of the hand-separated
magnetic fraction (for magnetite), or the bulk radioactivity for uraninite
(although this last can be misleading where there is an imbalance in U and
its daughter elements due to leaching by water).
A bulk analysis of base metals in humus with dithizone was very popular
in the 1950s (Rose et al., 1979). This gives a measure of the bulk Zn, Cu, Pb
and Ni in the sample, but is subject to interfering factors (Kauranne, 1967).
Other semiquantitative methods still being used include neocuproine for
copper, dimethylglyoxime for nickel, diphenyl carbatzime for chromium,
fluorimetry for uranium, the Gutzeit test for arsenic and many others.
Better of course than these rough methods is real quantitative chemical
analysis, as discussed in detail in Chapter 9 by Noras. The greatest
invention for exploration geochemistry was atomic absorption spectroscopy
386 Focal aspects ofgeochemistry applied in arctic and temperate regions

(AAS), which established its reputation in the 1960's. With its various
modifications — flame, graphite furnace, hydride generation — most
elements can be determined in a wide range of concentrations, in any
material, accurately, rapidly and cheaply. Samples must be in solution.
For easier operation, instruments can be equipped with automatic sample
changers, lamp exchangers and readout. AAS is universal and it can be
warmly recommended for any geochemical laboratory.
Another good analytical method for many elements is optical emission
spectrometry (OES). For this technique, buffer powder and internal stan-
dard are added usually to the finely ground sample to avoid any errors
due to the inhomogeneity of samples and interference of other elements.
The fastest and cheapest version of the optical emission spectrometer has
fixed channels with semiautomatic sample feed and allows the use of fine
mineral powder as such. It is only a pity that this otherwise fine method
does not always give good quality results because of the great variation in
sample matrices.
X-ray fluorescence is also popular, and the latest instrumentation and
improved methods of correction for matrix and interelement effects allow
the analysis of minor and even trace elements. The best results are obtained
using samples that have been finely ground and then homogenized by
fluxing.
Neutron activation analysis (NAA) is a good method for the determina-
tion of uranium, the noble metals and especially the rare earths. Studies
of the rare earths have become popular as a means of investigating the
origin of geological formations, and could perhaps be used in identifying
the source of geochemical anomalies in the same way that determinations
of sulphur and lead isotopes are used.
Small single-mineral grains can be identified and analyzed by scanning
electron microscopy (SEM) with a coupled energy dispersive X-ray analyzer
(EDX). The latest development for isotope analysis of single grains is the
secondary ion mass spectrometer (SHRIMP) — very sophisticated and very
expensive.
The most versatile instrument for today's high level geochemical labo-
ratory is the inductively coupled plasma excitation optical emission quan-
tometer (ICP). Although the requirement for dissolution of the sample
makes it a slower method than semiquantitative OES, it has the invaluable
feature of combining the sensitivity and repeatability of flameless AAS with
the versatility of NAA.
The latest development in the ICP area is the coupled ICP-MS, which
is even capable of separating some isotopes. The technique is a good tool
for identifying the source of an anomaly and can be used for elements like
lanthanides, tantalum, tungsten and the whole group of platinum metals,
which are otherwise difficult to measure. It is useful both for geochemical
exploration and for environmental research.
Interpretation of results 387

Total concentrations are suitable for geochemical and geological map-


ping, whereas partial concentrations provide more useful information for
environmental and exploration purposes. The purpose of the geochemical
study thus determines the method of extraction.
The salts of fulvic acids can be removed from humus by water leaching.
Water leach is seldom used in mineral soils. Exchangeable elements that
are weakly bound in colloids and clay minerals are usually extracted with
neutral ammonium acetate or slightly acidic EDTA (ethylenediamine tetra-
acetic acid) (Chao, 1984), or slightly alkaline hydroxylamine hydrochloride,
which attacks manganese oxyhydrate and the metals scavenged by it (de
Walque and Martin, 1983). Metals fixed in carbonates and organic material
may be leached with acetic acid or hydrogen peroxide. For sulphide-bound
metals, weak mineral acids — hydrochloric or nitric acids or a mixture
(aqua regia) of them — are usual. Metals occurring in oxides and more
resistant silicates require stronger attack with hydrofluoric acid or fusion
with a flux. Cole and Rose (1984) describe the modes of occurrence and
leaching of zinc and lead from oxides and silicates in glacial deposits,
Sandström (1984) in stream sediments, Mäkitie and Virri (1965) in clays
and Sillanpää and Lakanen (1966) in different soils. Chao (1984) gives
several methods suitable for sequential multiple extraction, and Bradshaw
et al. (1974) describe methods for differential extraction of soil profile
layers.

INTERPRETATION OF RESULTS

Geochemical research mainly deals with weak and highly varying con-
centrations of trace elements; sets of these elements are determined in
every sample and many samples are collected from every target. Thus
enormous amounts of data must be evaluated and interpreted. As described
in Chapter 10 by Gustavsson, methods of statistical analysis are indispens-
able in the task. The leap forward in computer technology has allowed
the automation of analytical equipment and, together with improved data
processing methods, has opened the way for large scale use of geochemistry.
Despite the remarkable advances in geological theory and skills and
improved geophysical techniques, geochemistry fills many specific needs
that other methods cannot. Not only is it suited for the mapping of large
unexposed regions and for delineating of ore potential areas, but it can
reveal the site and chemical nature of hidden mineralization. Geochemistry
has been successfully applied in both trackless wilderness and industrialized
countries and vast amounts of data have been gathered.
High concentrations of a single element do not necessarily point to an
ore; and sometimes a weak anomaly is more indicative than a strong one,
but unfortunately it can be interpreted as natural variation, due to the
388 Focal aspects of geochemistry applied in arctic and temperate regions

heterogeneity of material or the analytical standard deviation. One way of


detecting weak anomalies, to improve the anomaly-background contrast, is
to sum up the concentrations of several related elements. In assessing the
significance of geochemical results this multi-element approach has been
growing in importance.
The mode of occurrence of elements is often more important than mere
concentration. Chemical methods used for study of the mode of occurrence,
and via it the way of transport and possible provenance, have been
described above. Elemental associations extracted by different solutions
and revealed by statistical analysis help in understanding the different
phases of migration and fixation of the metals in the anomaly (Wilhelm
and Kosakevitch, 1979; Stea et al., 1986). For example, gossans originating
from laterite are rich in Cr, X Ti, Mn and 1^ while those from sulphides
are rich in Cu, Zn and Pb. The paragenetic associations known from
bedrock deposits deserve to be used more in interpreting and evaluating
anomalies.
To discriminate the part of the total lead in soils due to pollution from
natural lead (Erviö and Lakanen, 1973), it is not sufficient to determine the
mode of occurrence and the dispersion pattern, but the isotopic composition
needs to be determined as well. The isotopic composition can also be used
in classifying the gossans into promising and unpromising follow-up targets
(Vaasjoki and Gulson, 1985).
Computer processing of data provides a powerful means of searching
for natural elemental associations in a set of data, e.g., by principal
component and factor analysis. Trends can be calculated, and trend surfaces
constructed for single elements or different combinations of elements. The
program can be allowed to search freely for sets of correlating elements
or it can be "taught" to search for predetermined combinations (see
Chapter 10). Such clustering procedures have been used when applying
geochemical results to bedrock mapping (Nichol et al., 1969; Gustavsson
and Björklund, 1976; Shepherd et al., 1987) in poorly exposed regions.
Statistically enhanced results have led in many cases to findings that,
without data manipulation, would not have been possible (Riese et al.,
1986; Rogers et al., 1986).

FOLLOW-UP

Geochemical study of overburden over larger regions is done to aid


bedrock mapping in poorly exposed regions, study of the origin of overbur-
den formations and delineation of areas of mineral potential or pollution.
Over smaller areas it is done for the study of overburden stratigraphy and
transport of materials, for the purpose of locating anomalies originating
from hidden mineralizations and pollutional sources. Locally, it is done
Follow-up 389

to pinpoint targets for geophysical and geological follow-up studies and


diamond drilling.
Exploration geochemistry tells what elements are present, how they
occur and where they are situated. It does not tell how large and rich the
source of an anomaly is, nor if it is in a form that would permit mining
and extraction. Even the indication of location is usually so vague that
drilling sites cannot be selected without the help of other methods. The
role of other methods as a support to geochemistry has been discussed by
Stone and Gallagher (1984). Geophysical methods are often good indicators
of place. However, if the physical character of a mineralization differs
from the country rock only slightly, or if the overburden is thick, then
geophysics fails as well. In that case, the geological imagination is left
to work — guided by tectonic and stratigraphic field measurements and
analogies from known deposits; in many cases it is almost the only tool for
assessing the location of deeper ore mineralizations.
The next stage is the diamond drilling. In the search for an ore body of
20 x 10 x 100 m at a dark depth of 500 m, one has to be extremely lucky
to make a hit with a furiously rotating 0 40 mm drill rod. Geophysical
measurements in the first holes and geological observations on rock cores
will draw one closer to the ore. But rock geochemistry of the core —
calculation of the ratios of lighter and heavier elements and trends in
element concentrations with depth — have also proved to be extremely
informative about location.
Sound exploration strategy demands that all possible methods should be
available, and applied flexibly and in optimum combination as demanded
by the case.
391

REFERENCES

Aario, R., 1977. Classification and terminology of morainic landforms in Finland. Boreas, 6:
87-100.
Agha, M. and Ibrahim, M.T., 1984. Maximum likelihood estimation of mixtures of distributions.
Appl. Stat., J. R. Stat. Soc, Ser. C, 33(3): 327-332.
Agricola, G., 1556. De Re Metallica. Translated from the first Latin edition of 1556, with
biography, introduction, annotations and appendix upon the development of mining methods,
etc., from the earliest times. Published by the Mining Magazine, London, 1912.
Agterberg, F.P., 1974. Geomathematics, Mathematical Background and Geo-science Applications.
Elsevier, Amsterdam, 596 pp.
Aho, L., 1980. On the Sb-mineralization in the Seinäjoki-Nurmo area. Rep., Geol. Surv. Finland
(unpublished).
Ahrens, L.H., 1953. A fundamental law of geochemistry. Nature, 172: 1148.
Ahrens, L.H. and Taylor, S.R., 1961. Spectrochemical Analysis. Addison-Wesley, London, 454 pp.
Alhonen, P., 1988. Suomen kasvillisuuden kehitysvaiheet jääkauden jälkeen. Lounais-Hämeen
luonto. Vuosikirja, 75: 56-61.
Alley, R.B., Blankenship, D.D., Bentley, C.R. and Rooney, S.T., 1986. Deformation of till beneath
ice stream B, West Antarctica. Nature, 322(3): 57-59.
Alviola, R., 1974. Selostus pegmatiittitutkimuksista Kiteen-Tohmajärven alueella vuosina 1971-
1973. Rep., Geol. Surv. Finland (unpublished).
Anderberg, M.R., 1973. Cluster Analysis for Applications. Academic Press, New York, NY, 359
pp.
Anderson, J , Van der Walt, T.N. and Strelow, F.W, 1985. Determination of gallium at the
ppb level in South African primary and secondary reference samples by ion exchange
chromatography and electrothermal atomic absorption spectrophotometry. Geostand. Newsl.,
9: 17-18.
Anderson, T.W, 1989. Vegetation changes over 12 000 years. Geos, 38: 39-47.
Andrews, D.F., 1973. Graphical techniques for high dimensional data. In: T. Cacoullos (Editor),
Discriminant Analysis and Application. Academic Press, New York, N.Y, pp. 37-59.
Andrews, J.T., Shilts, W.W and Miller, G.H., 1983. Multiple deglaciations of the Hudson Bay
lowlands, Canada, since deposition of the Missinaibi (last interglacial?) Formation. Quat.
Res, 19: 18-37.
Andrews, M.J., Bibby, J.M, Fuge, R. and Johnson, C.C., 1984. The distribution of iodine and
chlorine in soils over lead-zinc mineralisation, east of Glogfawr, Mid-Wales. J. Geochem.
Explor, 20: 19-32.
Antropova, N T , 1975. Forms of Occurrence of Elements in Dispersion Haloes of Ore Deposits.
Nedra, Leningrad, 144 pp.
Armands, G, 1967. Geochemical prospecting of a uraniferous bog deposit at Masugnsbyn,
392 References

northern Sweden. In: A. Kvalheim (Editor), Geochemical Prospecting in Fennoscandia.


Interscience, London, pp. 127-154.
Armour-Brown, A. and Nichol, I., 1970. Regional geochemical reconnaissance and the location
of the metallogenic provinces. Econ. Geol., 65: 312-320.
Aucott, J.W, 1987. Workshop 5: Geochemical anomaly recognition. J. Geochem. Explor., 29:
375-376.
Aurola, E., 1955. Über die Geschiebeverfrachtung in Nordkarelien. Acta Geogr., 14(4): 32-52.
Äyräs, M.,1979. Eräiden hivenmetallien esiintymisestä moreenissa ja humuksessa Hannukaisen
alueella Kolarissa. Summary: Some trace metals in till and humus in the Hannukainen area,
Kolari, northern Finland. Geol. Surv. Finl., Rep. Inv., 44, 34 pp.

Barnes, R.M. and Diallo, A., 1985. Application of inductively coupled plasma atomic emission
spectrometry to the determination of platinum group metals, silver, tantalum, thorium and
uranium in geological materials with poly(dithiocarbamate) resin separation. In: L.R.P. Butler
(Editor), Analytical Chemistry in the Exploration, Mining and Processing of Materials, 2nd
International Symposium under IUPAC, Pretoria, 15-19 April 1985. Blackwell, Oxford, pp.
3-13.
Barnett, V and Lewis, T., 1978. Outliers in Statistical Data. John Wiley and Sons, New York,
NY, 365 pp.
Batterson, M.J., 1989. Glacial dispersal from the Strange Lake alkalic complex, northern
Labrador. In: R.N.W DiLabio and WB. Coker (Editors), Drift Prospecting. Energy, Mines
and Resources Canada, 1989, pp. 31-40.
Belyaev, Yu. and Koveshnikova, T.A., 1975. On the possibility of estimating the information
capacity of the methods of analysis of the composition of matter. In: A.I. Tugarinov (Editor),
Recent Contributions to Geochemistry and Analytical Chemistry, Keter Publ. (distributed by
Wiley), Jerusalem, pp. 661-664.
Beus, A.A., and Grigorian, S.V, 1977. Geochemical Exploration Methods for Mineral Deposits.
Applied Publishing Company, Wilmette, 111., 287 pp.
Bezdek, J.C., 1981. Pattern Recognition with Fuzzy Objective Function Algorithms. Plenum
Press, New York, NY, 256 pp.
Billings, G.K., 1969. Abundance in common igneous rock types and terrestrial abundance.
Lithium. In: K.J. Wedepohl (Editor), Handbook of Geochemistry, Springer-Verlag, New York,
NY, II-l, 1028 pp.
Björklund, A. (Editor), 1984. Geochemical Exploration 1983. Proceedings of the 10th Inter-
national Geochemical Exploration Symposium, Espoo. J. Geochem. Explor., 21(1-3): 501
pp.
Björklund, A., 1985. Moränen innehaller enorma svavelmängder. Finl. Natur, 1: 4-7.
Björklund, A. and Lummaa, M., 1983. Representation of regional, local and residual variability
of geochemical data by means of filtering techniques. Proc. 2nd Int. Symp. on Methods of
Prospecting Geochemistry. Irkutsk, U.S.S.R., pp. 25-34 (in Russian).
Björklund, A. and Gustavsson, N., 1987. Visualization of geochemical data on maps: New
options. J. Geochem. Explor., 29: 89-103.
Blain, C.F. and Andrew, R.L., 1977. Sulphide weathering and the evaluation of gossans in
mineral exploration. Minerals Sei. Eng., 9(3): 119-150.
Bloom, H., 1955. A field method for the determination of ammonium citrate-soluble heavy
metals in soils and alluvium. Econ. Geol., 50: 535-541.
Bock, R., 1972. Aufschlussmethoden der Anorganischen und Organischen Chemie. Verlag
Chemie, Weinheim, 232 pp.
References 393

Bouchard, M.A., 1986. Geologie des depots meubles de la region de Temiscamie (Territoire-du-
Nouveau-Quebec). MM 83-03. Gouvernement du Quebec, 80 pp.
Bouchard, M.A., Cadieux, B. and Goutier, F., 1984. L'origine et les charasteristiques des
lithofacies du till dans le secteur nord du Lac Albanel, Quebec: une etude de la dispersion
glaciaire clastique. In: J. Guha and E.H. Chown (Editors), Chibougamou — Stratigraphy
and Mineralization, Can. Inst. Min. Metall., Spec. Vol., 34: 244-260.
Bouchard, M.A. and Martinau, G., 1984. Les aspects regionaux de la dispersion glaciaire,
Chibougamou, Quebec. In: J. Guha and E.H. Chown (Editors), Chibougamou — Stratigraphy
and Mineralization, Can. Inst. Min. Metall., Spec. Vol., 34: 431-440.
Bouchard, M.A. and Salonen, V-R, 1990. Boulder transport in shield areas. In: R.R. Kujansuu
and M. Saarnisto (Editors), Glacial Indicator Tracing. Balkema, Rotterdam, pp. 87-107.
Boulton, G.S., 1971. Till genesis and fabric in Svalbard, Spitsbergen. In: R.P Goldthwait
(Editor), Till: A Symposium. Ohio State Univ. Press, Columbus, Ohio, pp. 41-72.
Boulton, G.S., 1974. Processes and patterns in glacial erosion. In: D.R. Coates (Editor), Glacial
Geomorphology: Proceedings, 5th Geomorphological Symposium. State University of New
York, Binghampton, N.Y., pp. 41-87.
Boulton, G.S., 1975. Processes and pattern of sub glacial sedimentation, a theoretical approach.
In: A.E. Wright and F. Moseley (Editors), Ice Ages: Ancient and Modern. Shell House,
Liverpool, pp. 7-42.
Bouma, J., 1987. Soil resources, In: A.A. Archer, G.W Lüttig and I.I. Snezhko (Editors)
Man's Dependence on the Earth. UNEP/UNESCO Schweizerbartsche Verlagsbuchhandlung,
Stuttgart, pp. 117-135.
Box, G.E.P and Cox, D.R., 1964. An analysis of transformations. J. R. Stat. Soc, Ser. B, 26:
211-252.
Boyle, R.W, 1967. Geochemical prospecting-retrospect and prospect. Geol. Surv. Can., Paper,
66-54: 30-43.
Boyle, R.W, 1974. Elemental associations in mineral deposits and indicator elements of interest
in geochemical prospecting. Geol. Surv. Can., Paper 74-45, 40 pp.
Boyle, R.W, 1979. The geochemistry of gold and its deposits. Bull. Geol. Surv. Can. 280, 584 pp.
Boyle, R.W, 1982. Geochemical methods for the discovery of blind mineral deposits. Can. Inst.
Min. Metall. Bull., 75(845): 113-132.
Boyle, R.W. and Cragg, C.B., 1957. Soil analysis as a method of geochemical prospecting in Keno
Hill-Galena Hill area, Yukon Territory. Geol. Surv. Can. Bull. 39, 27 pp.
Boyle, D.R. and Troup, A.G., 1975. Copper-molybdenum porphyry mineralization in central
British Columbia, Canada: an assessment of geochemical sampling media useful in areas
of glaciated terrain. In: Prospecting in Areas of Glaciated Terrain. Institute of Mining and
Metallurgy, London, pp. 6-15.
Bradshaw, PM.D. (Editor), 1975. Conceptual models in exploration geochemistry. The Canadian
Cordillera and Canadian Shield. J. Geochem. Explor., 4: 1-213.
Bradshaw, PM.D. and Koksoy, M., 1968. Primary dispersion of mercury from cinnabar and
stibnite deposits. In: W Turkey (Editor), 23rd IGC, 7: 341-355.
Bradshaw, PM.D., Clews, D.R. and Walker, J.L., 1972. Exploration Geochemistry: A Series of
seven articles. Reprinted from Mining in Canada and Canadian Mining Journal. Barringer
Reseach Ltd., Toronto, Ont., 49 pp.
Bradshaw, PM.D., Thomson, J., Smee, B.W and Larsson, J.O., 1974. The application of
different analytical extractions and soil profile sampling in exploration geochemistry. J.
Geochem.Explor. 3: 204-225.
Breiman, L., Friedman, J.H., Ohlsen, R.A. and Stone, C.J., 1984. Classification and Regression
394 References

Trees. Waldsworth, Belmont, 358 pp.


Brent, E.E., 1989. Statistical Navigator™, Users' Guide and Reference Manual, Version 1.1,
Idea Works, Columbia, Mo.
Brown, B.W and Hilchey, G.R., 1975. Sampling and analysis of geochemical materials of gold.
In: I.L. Elliott and WK. Fletcher (Editors), Geochemical Exploration. Elsevier, Amsterdam,
pp. 683-690.
Brundin, N.H. and Nairis, B., 1972. Alternative sample types in regional prospecting. J.
Geochem. Explor., 1: 7-46.
Brune, D., Forkman, B. and Persson, B., 1984. Nuclear Analytical Chemistry. Verlag Chemie,
Weinheim, 557 pp.
Brummer, G., Tille, KG., Herms, U. and Clayton, P.M., 1983. Adsorption-desorption and/or
precipitation-dissolution processes of zinc in soils. Geoderma, 31: 337-354.
Bölviken, B., 1971. A Statistical approach to the problem of interpretation in geochemical
prospecting. In: R.W Boyle and J.I. McGerrigle (Editors), Geochemical Exploration. Can.
Inst. Min. Metall., Spec. Vol., 11: 564-567.
Bölviken, B., 1979. The redox potential field of the earth. In: L.H. Ahrens (Editor), Origin and
distribution of the elements. Phys. Chem. Earth, 11: 649-665.
Bölviken, B. and Logn, 0., 1975. An electrochemical model for element distribution around
sulphide bodies. In: I.L. Elliot and WK. Fletcher (Editors), Geochemical Exploration 1974.
Elsevier, Amsterdam, pp. 631-648.
Bölviken, B., Bergström, J., Björklund, A., Kontio, M.,Ottesen, R.T., Steenfelt, A. and Volden, T,
1986. Geochemical Atlas of Northern Fennoscandia. Scale 1:4 000 000. Geological Surveys
of Finland, Greenland, Norway and Sweden, Espoo, 19 pp., 155 map sheets.

Cameron, E.M., 1977. Geochemical dispersion in mineralized soils of a permafrost environment.


J. Geochem. Explor., 7: 301-326.
Carpenter, R.H., Pope, T.A. and Smith, R.L., 1975. Fe-Mn oxide coatings in stream sediment
geochemical surveys. J. Geochem. Explor., 4: 349-363.
Cazalet, P.C.D., 1973. Notes on the interpretation of geochemical data in glacial areas. In:
Prospecting in Areas of Glaciated Terrain. Institution of Mining and Metallurgy, London,
pp. 25-29.
Cazalet, PC.D., 1982. A review of geochemical exploration techniques 1971-1981. In: A.G.
Brown (Editor), Mineral Exploration in Ireland, Progress and Developments 1971-1981.
Irish Association for Economic Geology, 179 pp.
Chao, T.T., 1972. Selective dissolution of manganese oxides from soils and sediments with
acidified hydroxylamine hydrochloride. Soil. Sei. Soc. Am. Proc, 36: 764-768.
Chao, T.T., 1984. Use of partial dissolution techniques in geochemical exploration. J. Geochem.
Explor, 20: 101-135.
Chao, T.T, Sanzalone, R.F. and Hubert, A E., 1978. Flame and nameless atomic-absorption
determination of tellurium in geological materials. Anal. China. Acta, 96: 251-257.
Chisholm, E.O., 1950. A simple chemical method of tracing mineralization through light
non-residual overburden. Can. Inst. Min. Metall. Bull, 43(454): 38-39.
Chork, C.Y, 1990. Unmasking multivariate anomalous observations in exploration geochemical
data from sheeted-vein tin mineralization near Emmaville, N.S.W, Australia. J. Geochem.
Explor, 37(2): 205-223.
Closs, L.G, 1987. Exploration geochemistry. Geotimes, 32(2): 23-24.
Cogger, N, 1974. An absorptiometric method for the determination of small amounts of tin in
rocks and ores with toluene-3,4-dithiol. Inst. Geol. Sei. Inf. Ser, Rep, 13, 8 pp.
References 395

Coker, W.B. and DiLabio, R.N.W, 1989. Geochemical exploration in glaciated terrain: geochemical
responses. In: G.D. Garland (Editor), Proceedings of Exploration 87. Ontario Geol. Surv.,
Spec. Vol., 3: 336-383.
Cole, D.R. and Rose, A.W., 1984. Distribution and mode of occurrence of zinc and lead in glacial
soils. J. Geochem. Explor., 20: 137-16.
Crow, E.L. and Shimizu, K., 1988. Lognormal Distributions — Theory and Applications. Marcel
Dekker, New York, N.Y., 387 pp.
Currie, L.A., 1968. Limits for qualitative detection and quantitative determination. Anal. Chem.,
40: 586-591.

Danielsson, A., Lundgren, F. and Sundkvist, G., 1959. The tape machine. A new tool for
spectrochemical analysis, Part I. Spectrochim. Acta, 15: 122-125.
Date, A.R. and Gray, A.L., 1983. Development progress in plasma source mass spectrometry.
Analyst, 108: 159-165.
David, P.P. and Bedard, P, 1986. Stratigraphy of the McGerricle Mountains granite trains of
Gaspesie. Geol. Surv. Can., Paper, 86-IB: 319-328.
Davis, J.C., 1973. Statistics and Data Analysis in Geology. John Wiley and Sons, New York, NYJ
550 pp.
De Geer, G., 1940. Geochronologia Suecica. Principles, Text and Atlas. K. Svenska Vetenskap.
Handl, 3, Sei. Bd. 18, 6.
De Geoffroy, J. and Wu, S.M., 1970. Design of a sampling plan for regional geochemical surveys.
Econ. Geol, 65: 340-347.
Delecour, F., van Praag, H. and Cherduville, G, 1974. Pedologie, Ghent, XXIV(3): 216-237.
Denoyer, E , Ediger, R. and Hager, J , 1989. The determination of precious metals in geological
samples by ICP-mass spectrometry. Atomic Spectrosc, 10: 97-102.
De Vos, W and Viaene, W, 1980. Geochemical study of soils and metallogenetic implications at
Hiendelaencina, Guadalajara, Spain. Mineral. Deposita, 15: 87-99.
De Walque, L. and Martin, H , 1983. Geochemical dispersion of lead and zinc around a gossan
in a carbonate environment at Heure (Belgium). Mineral. Deposita, 18: 27-38.
DiLabio, R.N.W, 1979. Drift prospecting in uranium and base-metal mineralization sites,
District of Keewatin, Northwest Territories, Canada. In: Prospecting of Areas in Glaciated
Terrain. Institution of Mining and Metallurgy, London, pp. 91-100.
DiLabio, R.N.W, 1981. Glacial dispersion of rocks and minerals in the south end of Lac
Mistassini, Quebec, with special reference to the Icon dispersal train. Geol. Surv. Can, Bull,
323, 46 pp.
DiLabio, R.N.W, 1982. Gold and tungsten abundance vs. grain size in till at Waverley, Nova
Scotia. Current Reseach, Part B. Geol. Surv. Can, Paper, 82-1B: 57-62.
DiLabio, R.N.W, 1989. Terrain geochemistry in Canada. In: R.J. Fulton (Editor) Quaternary
Geology of Canada and Greenland. Geol. Surv. Can, 1: 647-663.
DiLabio, R.N.W and Shuts, WW, 1978. Compositional variation of debris in glaciers, By lot
Island, District of Franklin. Current Research, Part B. Geol. Surv. Can, Paper, 78-1B: 91-94.
DiLabio, R.N.W and Shuts, WW, 1979. Composition and dispersal of debris by modern glaciers,
Bylot Island, Canada, In: Ch. Schlüchter (Editor), Moraines and Varves; Origin, Genesis,
Classification. Balkema, Rotterdam, pp. 145-155.
DiLabio, R.N.W and Coker, WB, 1982. Geochemistry of peat associated with uraniferous
bedrock in the Kashmere lake area, Manitoba. In: Prospecting in Areas of Glaciated Terrain.
Institution of Mining and Metallurgy, London, pp. 179-194.
DiLabio, R.N.W and Coker, W.B. (Editors), 1989. Drift prospecting. Geol. Surv. Can, Paper,
89-20: 170 pp.
396 References

Donner, J.J., 1986. Weichselian indicator erratics in the Hyvinkää area, southern Finland. Ann.
Acad. Sei. Fenn., A-III, 140 pp.
Drake, L.D, 1983. Ore plumes in tiU. J. Geol., 91(6): 707-713.
Dreimanis, A., 1956. Tracing of ore boulders as a prospecting method in Canada. Can. Min.
Metall., Bull., 51(550): 73-80.
Dreimanis, A., 1960. Geochemical prospecting for Cu, Pb, Zn in glaciated areas, eastern Canada.
Int. Geol. Congr., XXI, Copenhagen, Part II, pp. 7-19.
Dreimanis, A., 1982. Genetic varieties of tills. Int. Congr. Sedimentol., 111(11), 77 pp.
Dreimanis, A., 1983. Quaternary geological glacial deposits: Implications for the interpretation
of Proterozoic glacial deposits. Geol. Soc. Am., Mem., 161: 299-307.
Dreimanis, A. and Vagners, VJ., 1971. Bimodal distribution of rock and mineral fragments in
basal tills. In: R.P. Goldthwait (Editor), Till, A Symposium. Ohio State University Press,
Columbus, Ohio, pp. 237-250.
Dreimanis, A., Reaveley, G.H., Cook, R.J.B., Knox, K.S. and Moretti, F.J., 1957. Heavy mineral
studies in tills of Ontario and adjacent areas. J. Sediment. Petrol., 27: 148-161.
Drewry, D., 1986. Glacial Geologic Processes. Edward Arnold, London, 276 pp.
Duchaufour, P., 1982. Pedology. George Allen and Unwin, London, 448 pp.
Duda, R. and Hart, P., 1973. Pattern Classification and Scene Analysis. John Wiley and Sons,
New York, N.Y., 482 pp.
Dyck, W, 1969a. Field and laboratory methods used by the Geological Survey of Canada in
geochemical surveys, No. 10. Radon determination apparatus for geochemical prospecting
for uranium. Geol. Surv. Can., Paper, 68-21, 30 pp.
Dyck, W, 1969b. Development of uranium exploration methods using radon. Geol. Surv. Can.,
Paper, 68-46, 26 pp.

Easterbrook, D.J., 1981. Remanent magnetism in glacial tills and related diamictons. In:
E.B. Evenson, Ch. Schlüchter and J. Rabassa (Editors), Till and Related Deposits. INQUA
Symposium U.S.A.-Argentina 1982, pp. 303-313.
Edlund, S., 1989. Vegetation indicates potentially unstable Arctic terrain. Geos, 18(3): 9-13.
Ehlers. J., 1982. Different till types in North Germany and their origin. In: E.B. Evenson, Ch.
Schlüchter and J.Rabassa (Editors), Till and Related Deposits. INQUA Symposium U.S.A.
1981/Argentina 1982, pp. 61-80.
Ek, J., Ohlsson, S.A. and Selinus, O., 1988. Bly kadmium, seien- heia Sverige kartläggs. Forsk.
Framsteg, 2, 7 pp.
Eriksson, K., 1973. The distribution of some metals in different till fractions. In: Prospecting
in Areas of Glaciated Terrain. Institution of Mining and Metallurgy, London, pp. 83-86.
Eriksson, K, 1976. Regional prospecting by the use of peat sampling. J. Geochem. Explor., 5(3):
387-388.
Eriksson, K., 1983. Till investigations and mineral prospecting. In: J. Ehlers (Editor), Glacial
Deposits of North-West Europe. Balkema, Rotterdam, pp. 107-113.
Eriksson, K. and Eriksson, G., 1976. Strömbo: heavy metals in peat close to a boulder train. J.
Geochem. Explor., 5(3): 342-344.
Ermengen, S.V, 1957. Geochemical prospecting in Chibougamau. Can. Min. J., 78(1): 99-104.
Ervamaa, P, 1962. The Petolahti diabase and associated nickel-copper-pyrrhotite ore, Finland.
Bull. Comm. G6ol. Finl., 199, 80 pp.
Erviö, R. and Lakanen, E., 1973. Maan lyijysaastuminen sulattamon ympäristössä Tikkurilassa.
English summary: Lead contamination of soil in the environment of a smeltery in South
Finland. Ann. Agric. Fenn., 12: 200-206.
Everitt, B.S., 1978. Graphical Techniques for Multivariate Data. Heineman, London, 117 pp.
References 397

Eyles, N. and Menzies, J., 1983. The sub glacial landsystem. In: N. Eyles (Editor), Glacial
Geology. An Introduction for Engineers and Earth Scientists. Pergamon Press, Oxford, pp.
19-70.

Fletcher, W.K., 1981. Analytical Methods in Geochemical Prospecting. Handbook of Exploration


Geochemistry, Vol. 1. Elsevier, Amsterdam, 255 pp.
Flint, R.F., 1947. Glacial Geology and the Pleistocene Epoch. John Wiley and Sons, New York,
N.Y, 589 pp.
Flint, R.F., 1967. Glacial and Pleistocene Geology. John Wiley and Sons, New York, N.Y, 553 pp.
Forsskal, J., 1736. Underrättelse for Almogen och Gemene Man i Storfurstendömet Finland,
om Malmers och andre nyttige Tings Kännande och Efterletande. Aufbewart in Archiv des
Bergkolleqium in Reichsarkiv, Stockholm.
Foster, J.R., 1974. The efficiency of various digestion procedures in the extraction of metals
from rocks and rock-forming minerals. Can. Inst. Min. Metall. Bull., 66: 85-92.
Friedman, G.M. and Sanders, J.E., 1978. Principles of Sedimentology. John Wiley and Sons,
New York, N.Y, 792 pp.
Friedrich, G.H., 1974. Occurrences of heavy mineral sands in the coastal areas of Australia.
Erzmetall, 27(7/8): 350-353 (in German).
Fuge, R. and Andrews, M.J., 1985. The automated photometric determination of total fluorine
in mineral exploration. J. Geochem. Explor., 23: 293-297.
Fulton, R.B., 1950. Pospecting for zinc using semiquantitative analysis of soils. Econ. Geol., 45:
654-670.
Fyfe, WS., 1982. Recent sediments: The front line of enviromnental protection. Geosci. Can.,
9(1): 71-74.

Gabriel, K.R., 1981. Biplot display of multivariate analysis for inspection of data and diagnosis.
In: V Barnett (Editor), Interpretating Multivariate Data. John Wiley and Sons, Chichester,
pp. 147-173.
Gallagher, M.J., Smith, R.T., Fortey, N.J. and Parker, M.E., 1986. Lead-zinc exploration in
the lower carboniferous of south Scotland. In: Prospecting in Areas of Glaciated Terrain.
Institution of Mining and Metallurgy, London, pp. 39-47.
Garreis, R.M. and Christ, C.L., 1965. Solutions, Minerals and Equilibria. Harper and Row, New
York, NY, 450 pp.
Garrett, R.G., 1969. The determination of sampling and analytical errors in exploration
geochemistry. Econ. Geol., 64: 568-569.
Garrett, R.G., 1971. The dispersion of copper and zinc in glacial overburden at the Louvem
deposit, Val d' Or, Quebec. Toronto Symp. Vol., pp. 157-158.
Garrett, R.G., 1979. Sampling considerations for regional geochemical surveys. Geol. Surv. Can.,
Paper, 79-1A: 197-204.
Garrett, R.G., 1989. The chi-square plot: a tool for multivariate outlier detection. J. Geochem.
Explor., 32: 319-341.
Garrett, R.G. and Nichol, I., 1967. Regional geochemical reconnaissance in eastern Sierra Leone.
Inst. Min. Metall., Trans., Sect. B, 76: 97-112.
Garrett, R.G. and Lynch, J.J., 1976. A comparison of neutron activation delayed neutron
counting vs. fluorimetric analysis in large scale geochemical exploration for uranium. In:
Exploration for Uranium Deposits. Proceedings, Symposium on Exploration of Uranium Ore
Deposits, Vienna, 1976. IAEA, Vienna, pp. 321-334.
398 References

Garrett, R.G. and Sinding-Larsen, R., 1983. Optimal Composite Sample Size Selection,
Applications in Geochemistry and Remote Sensing. A poster session presented at the 10th
Int. Geochem. Explor. Symp., Helsinki, August 29-September 1983 (abstract).
Garrett, R.G. and Leymarie, E, 1989. Report on Workshop, 2: Data processing. J. Geochem.
Explor., 32: 477-483.
Gatehouse, S., Russell, D.W and van Moort, J.C., 1977. Sequential soil analysis in exploration
geochemistry. J. Geochem. Explor., 8: 483-494.
Geddes, R.S., 1982. The 'Vixen lake indicator train, northern Saskatchewan. In: Prospecting in
Areas of Glaciated Terrain. Institution of Mining and Metallurgy, London, pp. 264-283.
Gillberg, G., 1965. Till distribution and ice movements of the northern slope of the South
Swedish Highlans. Geol. Foren. Stockholm Förh., 86: 433-484.
Gillberg, G., 1967. Further discussion of the lithological homogeneity of till. Geol. Foren.
Stockholm Förh., 89: 29-49.
Ginzburg, I.I., 1960. Principles of Geochemical Prospecting (translation from Russian). Pergamon
Press, Oxford, 311 pp.
Gleeson, C.F. and Cormier, R., 1971. Evaluation by geochemistry of geophysical anomalies and
geological targets using overburden sampling at depth. Can. Inst. Min. Metall., Spec. Vol., 1:
159-165.
Gleeson, C.F. and Nichol, L, 1987. Workshop 2: Till Geochemistry. J. Geochem. Explor., 29:
359-373.
Gleeson, C.F., Rampton, VN., Thomas, R.D., Paradis, S. and Corden, S.G., 1989. Peatlands and
gold exploration. In: Prospecting in Areas of Glaciated Terrain. Institution of Mining and
Metallurgy, London, pp. 403-431.
Glen, J.W, Donner, J.J. and West, R.G., 1957. On the mechanism by which stones in till become
oriented. Am. J. Sei., 255: 194-205.
Gluckert, G., 1976. The Kuhmo drumlin field. Publ. Dep. Soils Geol., Univ. Turku, 11 pp.
Goldschmidt, VM., 1934. Drei Vorträge über Geochemie. Geol. Foren. Stockholm Förh., 56,
385 pp.
Goldschmidt, VM., 1937. The principles of distribution of chemical elements in minerals and
rocks. J. Chem. Soc., pp. 655-673.
Goldthwait, R.P, 1971. Introduction to till, today. In: R.P Goldthwait (Editor), Till, A Symposium.
Ohio State University Press, Columbus, Ohio, pp. 3-26.
Goldthwait, R.P, 1982. The growth of glacial geology and glaciology. Geosci. Can., 9(1): 36-37.
Gonzales, R.C. and Wintz, P., 1987: Digital Image Analysis. Addison-Wesley, Reading, Mass.,
503 pp.
Govett, G.J.S., 1960. Geochemical prospecting for copper in Northern Rhodesia. Proc. Int. Geol.
Congr., XXI, Copenhagen, 1960, Part II, pp. 44-56.
Govett, G.J.S., 1973. Differential secondary dispersion in transported soils and post-minerali-
zation rocks: an electrochemical interpretation. Int. Geochem. Explor. Symp., London,
pp. 81-91.
Govett, G.J.S., 1975. Soil conductivities: assessment of an electrogeochemical exploration
technique. Int. Geochem. Explor., Symposium, Vancouver, B.C., pp. 101-118.
Govett, G.J.S., 1976. Detection of deeply buried and blind sulphide deposits by measurement of
H+and conductivity of closely-spaced surface samples. J. Geochem. Explor., 6: 359-382.
Govett, G.J.S., 1983. Rock Geochemistry in Mineral Exploration. Handbook of Exploration
Geochemistry, Vol. 3. Elsevier, Amsterdam, 461 pp.
Govett, G.J.S. and Chork, C.Y., 1977. Detection of deeply buried sulphide deposits by measure-
ments of organic carbon, hydrogene-ion and conductance in surface soils. In: Prospecting in
References 399

Areas of Glaciated Terrain. Institution of Mining and Metallurgy, London, pp. 49-55.
Granath, G., 1983. Trace elements in magnetites as pathfinders for base-metal deposits in
Dalecarlia, Sweden. Bull. Geol. Inst. Univ. Uppsala, N.S., 9: 153-159.
Granath, G., 1984. Application of fuzzy clustering and fuzzy classification to evaluate the
provenance of glacial till. Math. Geol., 16(3).
Graybill, F.A., 1961. An Introduction to Linear Statistical Models, Vol. 1. McGraw-Hill, New
York, N.Y., 463 pp.
Grip, E., 1953. Tracing glacial boulders as an aid to ore prospecting in Sweden. Econ. Geol.,
48(8): 715-725.
Gubac, J., 1986. On the character of distribution of chemical elements in nature. Math. Geol.,
18(4): 429-432.
Gulson, B.L. and Vaasjoki, M., 1987. Lead isotope data from Thalanga, Dry River and MT
Chalmers base metal deposits and their bearing on exploration and ore genesis in eastern
Australia. Aust. J. Earth Sei., 34: 159-173.
Guo, X., Zhang, J., Yang, M and Fan, F., 1989. The application of hydride-generation atomic
fluorescence spectrometry in geochemical sample analysis. In: X. Xie, and S.E. Jennes
(Editors), J. Geochem. Explor., Special Issue, 33: 237-246.
Gustavsson, B. and Minell, H., 1977. Case history of discovery and exploration of Pleutajokk
uranium deposit, northern Sweden. In: Prospecting in Areas of Glaciated Terrain. Institution
of Mining and Metallurgy, London, pp. 72-79.
Gustavsson, N , 1983. Use of pattern recognition methods in till geochemistry. In: R.J. Howarth
(Editor), Statistics and Data Analysis in Geochemical Prospecting. Handbook of Exploration
Geochemistry, Vol. 2. Elsevier, Amsterdam, pp. 303-309.
Gustavsson, N. and Björklund, A., 1976. Lithological classification of tills by discriminant
analysis. J. Geochem. Explor., 5: 393-395.
Gustavsson, N. and Kontio, M., 1990. Statistical classification of regional geochemical samples
based on local characteristic models and data of the geochemical Atlas of Finland and from
the Nordkalott Project. In: D.F. Merriam and G. Gall (Editors), Computer Applications in
Resource Exploration: Prediction and Assessment for Petroleum, Metals and Nonmetals.
Comput. Geol., 7: 23-41.
Gustavsson, N , Noras, P. and Tanskanen, H., 1979. Seloste geokemiallisen kartoituksen
tutkimusmenetelmistä. Summary: Report on geochemical mapping methods. Geol. Surv.
Finl., Rep. Inv., 39, 20 pp.
Gy, P.M., 1982. Sampling of Particulate Materials — Theory and Practice. Developments in
Geomathematics, 4. Elsevier, Amsterdam, 431 pp.

Haldorsen, S., 1983. Mineralogy and geochemistry of basal till and their relationship to till
forming processes. Nors. Geol. Tidsskr., 63: 15-25.
Hale, M., Thompson, M. and Wheatley, M.R., 1984. Laser ablation of stream sediment pebble
countings for simultaneous multi-element analysis in geochemical exploration. J. Geochem.
Explor., 21(1/3): 361-371.
Halicz, L. and Russell, G.M., 1986. Simultaneous determination, by hydride generation and
inductively coupled plasma atomic emission spectrometry, of arsenic, antimony, selenium
and tellurium in silicate rocks containing the noble metals and suphide ores. Analyst, 111:
15-18.
Hall, R.D., 1981. The influence of till petrology on weathering of moraines in southwestern
Montana and northwestern Wyoming. In: E.B. Evenson, Ch. Schlüchter and J. Rabassa
(Editors), Tills and Related Deposits. INQUA Symposium on the Genesis and Lithology of
Quarternary Deposits, U.S.A. 1981. pp. 127-139.
400 References

Hall, G.E.M., Park, C.J. and Pelchat, J.C., 1987. Determination of tungsten and molybdenum at
low levels in geological materials by inductively coupled plasma mass spectrometry. J. Anal.
Atomic Spectrom., 2: 189-196.
Halonen, 0., 1967. Prospecting for asbestos. In: A. Kvalheim (Editor), Geochemical Prospecting
in Fennoscandia. Interscience, London, pp. 279-317.
Harbaugh, J.W., 1953. Geochemical prospecting abstracts through June, 1952. U.S. Geol. Surv.,
Bull., 1000-A: 1-50.
Hartikainen, A. and Salminen, R., 1982. Tohmajärven karttalehtialueen geokemiallisen kar-
toituksen tulokset. Summary: The results of the geochemical survey in the Tohmajärvi map
sheet area. Geol. Surv. Finl., Explanatory Notes to Geochemical Map, Sheet 4232, 58 pp.
Hausen, H., 1912. Studier öfver de sydfinska ledblockens spridning i Ryssland, jämte en öfversikt
af is-recessionens fbrlopp i Ostbaltikum. Deutsches Referat: Studier über die Ausbreitung
der südfinnischen Leitblöcke in Russland, nebst einer Übersicht der letzten Eisrezession im
Ostbaltikum. Fennia, 32, 32 pp.
Hawkes, H.E., 1952. Geochemical prospecting in the Blackbird cobalt district, Idaho. Econ.
Geol., 47(7): 771-772.
Hawkes, H.E., 1954. Geochemical prospecting investigations in the Nyeba lead-zinc district,
Nigeria. U.S. Geol. Surv., Bull, 1000-F: 225-355.
Hawkes, H.E, 1957. Principles of geochemical prospecting. U.S. Geol. Surv., Bull, 1000-F:
225-355.
Hawkes, H.E. and Lakin, H.W, 1949. Vestigial zinc in surface residuum associated with primary
zinc ore in east Tennessee. Econ. Geol, 44(4): 286-295.
Hawkes, H.E. and Webb, J.S, 1962. Geochemistry in Mineral Exploration. Harper and Row,
New York, N.Y, 415 pp.
Hedström, H , 1894. Studier over bergarter frän morän vid Visby. Sver. Geol. Undersök, Sen C,
139.
Hellaakoski, A, 1930. On the transportation of materials in the esker of Laitila. Fennia, 52(7):
1-41.
Helmersen, G.V, 1882. Geologische und physico-geographische Beobachtungen im Olonezer
Bergrevier. Beitr. Kennt, des russ. Reiches 2 Folge, Bd. 5, St. Petersburg, pp. 1-411.
Hirvas, H , 1977. Glacial transport in Finnish Lapland. In: Prospecting in Areas of Glaciated
Terrain. Institution of Mining and Metallurgy, London, pp. 128-137.
Hirvas, H , 1980a. Soil studies supporting mineral exploration at Kalliosalo, Nurmo. Rep, Geol.
Surv. Finland (unpublished).
Hirvas, H , 1980b. Moreenistratigrafiasta ja sen merkityksestä malminetsinnässä. Geologi, 32:
33-37.
Hirvas, H , Alfthan, A, Pulkkinen, E , Puranen, R. and Tynni, R, 1977. Raportti malminetsintää
palvelevasta maaperätutkimuksesta Pohjois-Suomessa vuosina 1972-1976. Summary: A
report on glacial drift investigations for ore prospecting purposes in northern Finland
1972-1976. Geol. Surv. Finl, Rep. Inv, 19, 54 pp.
Hirvas, H. and Nenonen, K, 1987. Till stratigraphy in Finland. Geol. Surv. Finl, Spec. Paper,
3: 49-63.
Hjulström, F, 1939. Transportation of detritus by moving water. In: P.D. Trask (Editor) Recent
Marine Sediments, A Symposium. American Association of Petroleum Geologists, Tulsa,
Okla, pp. 5-31.
Hoffman, S.J. and Fletcher, WK, 1979. Selective sequential extraction of Cu, Zn, Fe, Mn and
Mo from soils and sediments. Int. Geochem. Explor. Symp, Denver, Colo, pp. 289-299.
Holman, R.H. and Webb, J.S, 1957. Exploratory geochemical soil survey at Ruhiza ferberite
References 401

mine, Uganda. In: S.F. Kelly and E.W Westrick (Chairmen), Methods and Case Histories in
Mining Geophysics. 8th Commonw. Mining Congr., Montreal, Ont., pp. 353-357.
Holmes, A., 1972. Principles of Physical Geology. William Cloves and Sons, London, 1288 pp.
Holmes, C.W, 1952. Drift dispersion in West-Central New York. Bull. Geol. Soc. Am., 63:
993-1010.
Horsnail, R.E, 1975. Strategy and tactical geochemical exploration in glaciated terrain:
illustrations from Northern Ireland. In: Prospecting in Areas of Glaciated Terrain. Institution
of Mining and Metallurgy, London, pp. 16-31.
Horsnail, R.F. and Elliott, I.L., 1971. Some environmental influences on the secondary dispersion
of molybdenum and copper in Western Canada. Can. Inst. Min. Metall., Spec. Vol., 11: 166-
175.
Hosking, K.F.G., 1957. Identification of lithium minerals. Min. Mag., Vol. XCVI(5): 271-275.
Hosmer, D.W, 1973. A Comparison of iterative maximum likelihood estimates of the parameters
of a mixture of two normal distributions under different types of sample. Biometrics, 29:
761-770.
Howarth, R.J., 1973. FORTRAN IV programs for empirical discriminant classification of spatial
data. Geocom Bull., 6: 1-31.
Howarth, R.J. and Sinding-Larsen, R., 1983. Multivariate Analysis. In: R.J. Howarth (Edi-
tor), Statistics and Data Analysis in Geochemical Prospecting. Handbook of Exploration
Geochemistry, Vol. 2. Elsevier, Amsterdam, pp. 207-289.
Huber, PJ., 1981. Robust Statistics. John Wiley and Sons. New York, NY, 308 pp.
Hubert, A.E. and Chao, T.T., 1985. Determination of gold, indium, tellurium and thallium in the
same sample digest of geological materials by atomic-absorption spectroscopy and two-step
solvent extraction. Talanta 32: 568-570.
Huff, L.C., 1951. A sensitive field test for detecting heavy metals in soil or sediment. Econ.
Geol, 46(5): 517-542.
Huff, L.C., 1952. Abnormal copper, lead and zinc content of soil near metalliferous veins. Econ.
Geol., 47: 517-542.
Huhta, P, 1988. Studies of till and heavy minerals for gold prospecting at Ilomantsi, Eastern
Finland. In: Prospecting in Areas of Glaciated Terrain. Institution of Mining and Metallurgy,
London, pp. 285-292.
Hyvärinen, L., 1969. On the geology of the copper ore field in the Virtasalmi area, Eastern
Finland. Bull. Comm. Geol. Finl., 240: 1-82.
Hyvärinen, L., Kauranne, L.K. and Yletyinen, V, 1973. Modern boulder tracing in prospecting.
In: Prospecting in Areas of Glaciated Terrain. Institution of Mining and Metallurgy, London,
pp. 87-95.
Hyyppä, E., 1966. I Salpausselän rakenne Lahden seudulla. Summary: On the structure of I
Salpausselkä at Lahti. Geologi, 18: 73-76.
Högbom, A., 1931. Nya iaktelser inom Norr- och Västerbottens urberg. Geol. Foren. Stockholm
Förh., 53, 415 pp.

James, C.H., 1967. The use of the terms "primary" and "secondary" dispersion in geochemical
prospecting. Econ. Geol., 62(7): 977-999.
Jarvis, K.E., 1988. Inductively coupled plasma mass spectrometry: a new technique for the
rapid or ultra-trace level determination of the rare-earth elements in geological materials.
Chem. Geol., 68: 31-39.
Jenkins, R., Gould, R.W and Gedcke, D., 1981. Quantitative X-ray Spectrometry. Marcel Dekker,
New York, NY, 586 pp.
402 References

Jenny, E.A., 1968. Controls on Mn, Fe, Co, Ni, Cu and Zn concentrations in soils and water: The
significant role of hydrous Mn and Fe oxides. Am. Chem. Soc, Adv. Chem. Sen, 73: 337-387.
Johnson, WM. and Maxwell, J.A., 1981. Rock and Mineral Analysis. Wiley, New York, N.Y, 489
pp.
Jöreskog, K.G., Klovan, J.E., Reyment, R.A., 1976. Geological Factor Analysis. Elsevier,
Amsterdam, 178 pp.

Kabata-Pendias, A. and Pendias, H., 1984. Trace Elements in Soils and Plants. CRC Press, Boca
Raton, Fla., 315 pp.
Kahma, A., 1965. Trained dog as tracer of sulphide bearing glacial boulders. Sedimentology,
Vol. 5; Atlas Vol. 1, 4.
Kahma, A., 1973a. The main metallogenic features in Finland. Geol. Surv. Finl., Bull., 265, 28
pp.
Kahma, A., 1973b. Metallogenic map of Finland. 1:2000000. Maanmittaushallituksen kartta-
paino, Helsinki.
Kahma, A., Nurmi, A. and Mattson, P., 1975. On the composition of the gases generated by
sulphide-bearing boulders during weathering, and on the ability of prospecting dogs to detect
samples treated with these gases in the terrain. Geol. Surv. Finl., Rep. Inv., 6, 4 pp.
Kaszycki, C.A. and DiLabio, R.N.W, 1986. Surficial geology and till geochemistry, Lynn Lake-
Leaf rapids region, Manitoba. Current Research, Part B. Geol. Surv. Can., Paper, 86-IB:
245-256.
Kauranne, L.K., 1951. Outokummun lohkarevastan moreenin mineraalikoostumuksesta. Manu-
script, Archives of Helsinki Univ., Dep. Geol., 20 pp.
Kauranne, L.K., 1957. Maaperägeologisista malminetsintämenetelmistä-kokeilu Makolassa.
Rep., Archives of Helsinki Univ., Dep. Geol., 84 pp. (unpublished).
Kauranne, L.K., 1958. On prospecting for molybdenum on the basis of its dispersion in glacial
tül. Bull. Comm. Geol. Finl., 180: 31-34.
Kauranne, L.K., 1959. Pedogeochemical prospecting in glaciated terrain. Bull. Comm. Geol.
Finl., 184, 10 pp.
Kauranne, L.K., 1960a. A statistical study of stone orientation in glacial till. C.R. Soc. Geol.
Finl., 32: 87-97.
Kauranne, L.K., 1960b. Moreeni-murske. Summary: Glacial till — a product of crushing.
Geologi, 12(6): 72-74.
Kauranne, L.K., 1967. Aspects of geochemical humus investigation in glaciated terrain. In:
A. Kvalheim (Editor), Geochemical Prospecting in Fennoscandia. Interscience, London, pp.
261-272.
Kauranne, L.K., 1970. On the abrasion and impact strength of gravel and rocks in Finland.
BuU. Comm. Geol. Finl., 243: 1-61.
Kauranne, L.K., 1975. Regional geochemical mapping in Finland. In: Prospecting in Areas of
Glaciated Terrain. Institution of Mining and Metallurgy, London, pp. 71-81.
Kauranne, L.K. (Editor), 1976a. Conceptual models in exploration geochemistry — Norden
1975. J. Geochem. Explor., 5: 173-420.
Kauranne, L.K., 1976b. Petolahti: Copper and nickel in till and humus. J. Geochem. Explor.
5(3): 292-296.
Kauranne, L.K., 1980. On regional differences in trace element contents. In: H. Tanskanen
(Editor), Methods of the Geochemical Mapping and Boulder Prospecting in the Eastern Part
of the Baltic Shield. Proc. Finn.-Sov. Seminar, Leningrad, 3-5 Sepember, 1979. Geological
Survey of Finland, Espoo, pp. 39-64.
References 403

Kauranne, L.K., Salminen, R. and Äyräs, M., 1977. Problems of geochemical contrast in Finnish
soils. In: Prospecting in Areas of Glaciated Terrain. Institution of Mining and Metallurgy,
London, pp. 34-44.
Kauranne, L-M., 1979. Vihannin karttalehtialueen geokemiallisen kartoituksen tulokset. Ex-
planatory Notes to Geochemical Maps, Sheet 2434. Geolological Survey of Finland, Espoo,
55 pp.
Kinealy, K. and Eadington, P.J., 1984. The effect of temperature of reaction on the extraction of
tin from calc-silicate rocks by NH4I volatilisation. J. Geochem. Explor., 20: 9-17.
Kinnunen, K., 1979. Ore mineral inclusions in detrital quartz contained in basal till and the
glacial transport from Ylöjärvi copper tungsten deposit, Southwestern Finland. Bull. Geol.
Surv. Finl., 298, 55 pp.
Kivekäs, E.J., 1946. Zur Kenntnis der mechanischen, chemischen und mineralogischen Zusam-
mensetzung der finnischen moränen. Acta Agralia Fennica, 60(2): 1-122.
Klassen, R.A. and Thompson, F.J., 1989. Ice flow history and glacial dispersal patterns, Labrador.
In: R.N.W DiLabio and WB. Coker (Editors), Drift Prospecting. Geol. Surv. Can., Paper,
89-90: 21-29.
Kobayashi, J., 1972. Air and water pollution by Cd, Pb and Zn attributed to the largest Zn
refinery in Japan. In: D.D. Hemphill (Editor), Trace Substances in Environmental Health,
V University of Missouri, Columbia, Mo., pp. 117-128
Koivisto, T. and Manner, R., 1981. Koivun karttalehtialueen esitutkimukset. Rep., Geol. Surv.
Finland (unpublished).
Kokko, J., 1981. Eräiden metallien esiintymisestä moreenissa, humuksessa ja turpeessa
Lutsokurun alueella Kittilässä. Rep., Archives of Helsinki Univ, Dep. Geol., 94 pp.
(unpublished).
Kokkola, M., 1977. Application of humus to exploration. In: Prospecting in Areas of Glaciated
Terrain. Institution of Mining and Metallurgy, London, pp. 104-110.
Kokkola, M. and Penttilä, VJ., 1976. Aittojärvi: molybdenum in humus, drift and bedrock. J.
Geochem. Explor., 5(3): 198-208.
Koljonen, T. (Editor), 1992. Geochemical Atlas of Finland, Part 2. Till. Geological Survey of
Finland, Espoo (in press).
Koljonen, T, Gustavsson, N., Noras, P., Tanskanen, H., 1989. Geochemical Atlas of Finland:
preliminary aspects. J. Geochem. Explor. 32(1/3): 231-242.
Kontas, E., 1981. Rapid determination of gold by nameless atomic absorption spectrometry
in the ppm and ppb ranges without organic solvent extraction. Atomic Spectrosc, 2: 59-
61.
Kontas, E., 1991. Gold contamination of the fine fraction of till during sampling and sample
preparation. In: A.J. Björklund (Editor), Gold Geochemistry of Finland. J. Geochem. Explor.,
39: 289-294.
Korpela, K., 1969. Die Weichsel-Eiszeit und ihr Interstadial in Peräpohjola (nördliches Nordfinn-
land) im Licht von submoränen Sedimenten. Ann. Acad. Sei. Fennicae, Ser. A, III, 99: 1-
108.
Kovalevskii, A.L., 1984. The cadmium informativeness of various plant species. Geochem. Int.,
21(2): 148-157.
Krumbein, W.C., 1937. Sediments and exponential curves. J. Geol., 45: 577-601.
Krylova, L.Ya, 1977. Exploration significance of iodine distribution in the bedrock and weathering
crust of the Maykain deposit. Int. Geol. Rev., 20: 357-361.
Kujansuu, R. and Niemelä, J., 1984. The Quaternary deposits in Finland. 1:1000 000 map.
Geological Survey of Finland, Espoo.
404 References

Kvalheim, A. (Editor), 1967. Geochemical Prospecting in Fennoscandia. Interscience, New York,


N.Y., 350 pp.
Königsson, L.-K, 1992. Economic Quaternary geology in Sweden — some comments on the
contemporary situation. In: Kalevi Kauranne (Editor), Stratigraphy, Engineering Geology
and Construction. Geol. Surv. Finl., Spec. Paper, 15: 75-87.
Körner, R., Dubey, R. and Parnandi, M., 1989. Scientists monitor climate and pollution from ice
caps and glaciers. Geos, 18(3): 33-38.

Lahermo, E, Ilmasti, M., Juntunen, R. and Taka, M., 1990. The Geochemical Atlas of Finland,
Part 1. Ground water. Geological Survey of Finland, Espoo, 66 pp.
Lahti, H.R., 1971. Factors contributing to secondary dispersion of trace elements in glacial soils,
St. Stephen area, Brunswick. Rep., University of New Brunswick (unpublished).
Lalonde, J.P. and Beaumier, M., 1984. Pedogeochemical prospecting in a clay environment. Can.
Inst. Min. Metall., Bull., 77(862): 56-76.
Lavikainen, S., 1973. Map sheet 4244, Ilomantsi. Geological Map of Finland 1:100000,
Pre-Quaternay rocks. Geol. Survey Finland, Espoo.
Laville-Timsit, L., 1986. Neoformation de suliures de fer et de zinc en milieu reducteur en aval
du gisement Pb, Zn, Cu, Ag. de Porte-aux Moines (Cötes-du-Nord): L'anomalie Geochemique
Pb-Zn des sols de Kerouaran. Sei. Geol. Bull., 39(3): 263-275.
Laville-Timsit, L. and Wilhelm, E., 1979. Comportement supergene des metaux autour du gite
sulfure de Porte-aux-Moines (Cotes-du-Nord). Application ä la prospection geochimique. Bull.
B.R.G.M., Sect., II, 2-3: 195-228.
Lawley, D.N. and Maxwell, A.E., 1971. Factor Analysis as a Statistical Method. Butterworths,
London, 153 pp.
Lawson, A., 1974. A comparison of the pebble orientations in ice and deposits of the Matanuska
glacier, Alaska. J. Geol., 87: 629-645.
Lee, H.A., 1965. Investigation of eskers for mineral exploration. Geol. Surv. Can., Paper, 65014,
17 pp.
Lee, H.A., 1971. Mineral discovery of the Canadian Shield using the physical aspect of
overburden. Can. Inst. Min. Metall., Bull., 64, pp. 32-36.
Legkova, VG., Bonbenkov, VN, Plisov, A.A., Bezukladnov, A.A. and Legkova, O.E., 1980.
Distribution of chemical elements in Quaternary rocks of the Kola-Karelian region. In: H.
Tanskanen (Editor), Methods of the Geochemical Mapping and Boulder Prospecting in the
Eastern Part of the Baltic Shield. Proc. Finn.-Sov. Seminar, Leningrad, 3-5 September, 1979.
Geological Survey of Finland, Espoo, pp. 93-114.
Lehmuspelto, P, 1987. Some case histories of the till transport distances recognized in
geochemical studies in northern Finland. In: R. Kujansuu and M. Saarnisto (Editors),
INQUA Till Symposium, Finland, 1985. Geol. Surv. Finl., Spec. Paper, 3: 163-168.
Lepeltier, C, 1969. A simplified statistical treatment of geochemical data by graphical represen-
tation. Econ. Geol., 64: 538-550.
Lett, R.E.W and Fletcher, W.K., 1978. The secondary dispersion of transition metals through a
copper-rich hillslope bog in the Cascade mountains, British Columbia. Proc. 7th Geochem.
Explor. Symp., Golden, Colo., pp. 103-115.
Levinson, A.A., 1974. Introduction to Exploration Geochemistry. Applied Publishing, Calgary,
Alta., 612 pp.
Li, T. and Yio, C-L., 1966. The abundance of chemical elements in the earth's crust and its
major tectonic units. Sei. Sinica, 15: 258-272.
Linden, A., 1975. Till petrographical studies in an Archaean bedrock area in southern central
Sweden. Striae, 1, 57 pp.
References 405

Longerich, H.E, Fryer, B.J. and Strong, D.F., 1987. Determination of lead isotope ratios by
inductively coupled plasma mass spectrometry (ICP-MS). Spectrochim. Acta, 42B: 39-48.
Lounamaa, J., 1967. Trace elements in trees and shrubs growing on different rocks in Finland.
In: A. Kvalheim (Editor), Geochemical Prospecting in Fennoscandia. Interscience Publishers,
London, pp. 287-317.
Luckman, B.H., 1989. Global change and the record of the past. Geos, 18(3): 1-8.
Lukashev, K.I. and Lukashev, VK., 1977. Distribution of heavy metals in populated areas:
problems of contamination and prospecting methods. English translation by V Lukashev from
Russian. Eksperimentaljnye issledovanija form i prozessov gipergennoi migrazii elementov.
ANBSSR Institut Geohimii i Geofisiki, Minsk, pp. 5-17.
Lukashev, VK., 1983. Mode of occurence of elements in secondary environments. J. Geochem.
Explor., 21(1/3): 73-87.
Lundbohm, Hj., 1887. Beskrifning till kartbladet Halmstad. Sver. Geol. Unders., Aa, 12.
Lundqvist, G., 1948. Blockens orientering i olika jordarter. Sver. Geol. Unders., C, 497: 1-29.
Lundqvist, J., 1980. New data concerning the Quaternary glaciation in Sweden. In: V Sibrava
(Editor), Quaternary glaciations in the northern hemisphere. IGCP Proj. 24, Rap., 5: 176-
182.
Lüttig, G., 1990. Quaternary research in view of modern requirements of applied geology. In:
L.K. Kauranne and L.-K. Königsson (Editors), Geology in the Nordic Countries. Striae, 29:
15-29.

Maceachern, I.J. and Stea, R.R., 1987. Geochemical studies on gold in till in Nova Scotia. J.
Geochem. Explor., 29: 425-426.
Magnusson, N.H., Granlund, E., Lundqvist, G., 1949. Sveriges Geologi. Nordstedts, Stockholm,
556 pp.
Malmqvist, L. and Kristiansson, K., 1983. Evidence for fast migration of metals through the
bedrock: a potential new exploration technique. In: A. Björklund and T. Koljonen (Editors),
10th International Geochemical Exploration Symposium. Abstracts, 50 pp.
Malmqvist, L. and Kristiansson, K, 1984. Experimental evidence for an ascending microflow of
geogas in the ground. Earth Planet. Sei. Lett., 79: 407-416.
Marcussen, Ib., 1973. Stones in Danish tills as a stratigraphical tool: A review. Bull. Geol. Inst.
Univ. Uppsala, N.S., 5: 177-181.
Martin, H., Wilhelm, E., Laville-Timsit, L., Lecomte, P, Sondag, F. and Warnant, P., 1984.
Enhancement of stream-sediment geochemical anomalies in Belgium and France by selective
extraction and mineral separations. J. Geochem. Explor., 20: 179-203.
Maritz, J.S., 1981. Distribution-free statistical methods. Monographs on applied probability and
statistics. Chapman and Hall, Cambridge, 256 pp.
Markowicz, A.A. and Van Grieken, R.E., 1986. X-ray spectrometry. Anal. Chem., 58: 279R-294R.
Mason, B., 1966. Principles of Geochemistry. John Wiley and Sons, New York, N.Y., 329 pp.
Massart, D.L., Dijkstra, A. and Kaufman, L., 1978. Evaluation and Optimization of Laboratory
Methods and Analytical Procedures. Elsevier, Amsterdam, 596 pp.
Matheis, G. and Pearson, M.J., 1982. Mineralogy and geochemical dispersion in lateritic soil
profiles of northern Nigeria. Chem. Geol., 35: 129-145.
Matisto, A., 1961. On the relation between stones of the eskers and the local bedrock in the
area northwest of Tampere, southwestern Finland. Bull. Comm. Geol. Finl., 193, 53 pp.
Matisto, A., 1976. Kallioperäkartan selitykset. Lehti 2132 Valkeakoski. Summary: Precambrian
rocks of the Valkeakoski map-sheet area. Suomen geologinen kartta 1:100000. Geological
Survey of Finland, Espoo, 34 pp.
406 References

McCarthy. J.H., Jr., Lambe, R.N. and Dietrich, J.A., 1986. A case study of soil gases as an
exploration quide in glaciated terrain-Crandon massive sulphide deposit, Wisconsin. Econ.
Geol, 81: 408-420.
McCorkell, R.H., Porritt, J.WM. and Brameld, M.P., 1981. A comparison of uranium exploration
methods at the South March uranium-copper occurrence. Can. Inst. Min. Metall., Bull.,
74(828): 93-98.
McLaughlin, R.J.M., 1955. Geochemical changes due to the weathering under varying climatic
conditions. Geochim. Cosmochim. Acta 5: 109-130.
Meier, A.L., 1980. Flameless atomic-absorption determination of gold in geological materials. J.
Geochem. Explor., 13: 77-85.
Mertie, J.B., Jr., 1954. The gold pan: a neglected geolocical tool. Econ. Geol., 49(6): 639-651.
Michelson, D.M., Clayton, L., Fullertone, D.S. and Borns, H.W, Jr., 1983. The late Wisconsin
glacial record of the Laurentide ice sheet in the United States. In: H.E. Wright (Editor), Late-
Quaternary Environments of the United States. University of Minnesota Press, Minneapolis,
Minn., pp. 3-37.
Michie, U.M., Gallagher, M.J. and Simpson, A., 1972. Detection of concealed mineralization in
northern Scotland. Int. Geochem. Explor. Symposium, London, pp. 117-130.
Miesch, A.T., 1981. Estimation of the geochemical threshold and its statistical significance. J.
Geochem. Explor., 16: 49-76.
Miller, J.K., 1979. Geochemical dispersion over massive sulphides within the continuous
permafrost zone, Bathurst Norsemines, Canada. In: Prospecting in Areas of Glaciated
Terrain. Institution of Mining and Metallurgy, London, pp. 101-109.
Miller, J.K., 1984. Model for clastic indicator trains in till. In: Prospecting in Areas of Glaciated
Terrain. Institution of Mining and Metallurgy, London, pp. 69-77.
Miller, J.K. and Cazalet, P.C.D., 1979. Barium geochemistry in Irish mineral exploration. In:
Prospecting in Areas of Glaciated Terrain. Institution of Mining and Metallurgy, London,
pp. 16-21.
Minell, H., 1980. The distribution of local bedrock material in some moraine forms from the
inner part of northern Sweden. Boreas, 9: 275-281.
Minczewski, J., Chwastowska, J. and Dybczynski, R., 1982. Separation and Preconcentration
Methods in Inorganic Trace Analysis. John Wiley and Sons, New York, N.Y., 543 pp.
Minkkinen, P., 1987. Evaluation of the fundamental sampling error of particulate solids. Anal.
Chim. Acta, 196: 237-245.
Minkkinen, P, 1989. SAMPEX — A computer program for solving sampling problems.
Chemometrics and Intelligent Lab. Systems, 7: 189-194.
Mitchell, R.L., 1964a. The Spectrochemical Analysis of Soil, Plants and Related Materials.
Commonw. Bur. Soils, Harpenden, Tech. Comm., No. 44A, 225 pp.
Mitchell, R.L., 1964b. Trace elements in soils. In: F.E. Bear (Editor), Chemistry of the Soil.
Reinhold, New York, N.Y, 515 pp.
Morris, J.C. and Stumm, W, 1967. Redox equilibria and measurements of potentials in the
aquatic environment. In: F.R. Gould (Editor), Equilibrum Concepts in Natural Water
Systems. Am. Chem. Soc, Adv. Chem. Ser., 65: 270-285.
Mutanen, T, 1976. Komatiites and komatiite provinces in Finland. Geologi, 4/5: 49-56.
Myers, D.E., 1988. Kriging hydrogeochemical data. In: D.F. Merriam (Editor), Current Trends
in Geomathematics. Plenum Press, New York, NY, pp. 117-142.
Mäkinen, J. and Lestinen, P., 1990. Sinkki-, lyijy-, kupari-, hopea— ja uraaniaiheiden moreeni-
geokemiallinen etsintä Kiihtelysvaaran itäpuoleisella alueella. Rep., Geol. Surv. Finland
(unpublished).
References 407

Mäkitie, 0. and Virri, K., 1965. On the exchange characteristics of some clay soils in the Middle
Uusimaa. Ann. Agric. Fenn., 4: 277-289.
Molder, K, and Salmi, M., 1954. General Geological Map of Finland 1:4 00000, Sheet B3, Vaasa.
Explanation of the map of the superficial deposits, Geol. Survey Finland, Espoo, 109 pp.

Neuvonen, K.J., 1961. Geological Map of Finland 1:100000, Sheet 2222, Seinäjoki. Geological
Survey of Finland, Espoo.
Nichol, I., 1987. Geochemical exploration for gold deposits in areas of glaciated overburden:
problems and new developments. In: Prospecting in Areas of Glaciated Terrain. Institution
of Mining and Metallurgy, London, pp. 7-16.
Nichol, I. and Björklund, A., 1973. Glacial geology as a key to geochemical exploration in areas
of glacial overburden with particular reference to Canada. J. Geochem. Explor., 2: 133-170.
Nichol, I., Garrett, R.G. and Webb, J.S., 1969. The role of some statistical and mathematical
methods in interpretation of regional geochemical data. Econ. Geol., 64: 204-220.
Nieminen, K. and Yliruokanen, I., 1976. Kitee: The copper-nickel anomalies of Lietsonsuo peat
pog. J. Geochem. Explor. 5(3): 248-253.
Nikitichev, A.P, 1980. Principles of mapping landscape geochemical areas at a scale of
1:1000000. In: H. Tanskanen (Editor), Methods of the Geochemical Mapping and Boulder
Prospecting in the Eastern Part of the Baltic Shield. Proc. Finnish-Soviet Seminar, Leningrad,
pp. 161-176.
Nikkarinen, M. and Björklund, A., 1976. Emmes: The use of till in spodume exploration. J.
Geochem. Explor., 5(3): 212-213.
Nikkarinen, M., Kallio, E., Lestinen, P and Äyräs, M., 1984. Mode of occurrence of Cu and Zn
in till over three mineralized areas in Finland. J. Geochem. Explor., 21: 239-247.
Niskavaara, H. and Kontas, E., 1990. Reductive coprecipitation as a separation method for
the determination of gold, palladium, platinum, rhodium, silver, selenium and tellurium in
geological samples by graphite furnace atomic absorption spectrometry. Anal. Chim. Acta,
231: 273-282.
Noras, P., 1992. Determination of gold by aqua regia-potassium bromate digestion, methyl
isobutyl-ketone extraction and flame atomic absorption. In: E. Kontas (Editor), Analytical
Methods for Determining Gold in Geological Samples. Geol. Surv. Finl., Rep. Invest, (in
press).
Noras, P and Kontas, E., 1989. Extractabilities of elements — a new approach to interpretation
of regional geochemical data. XII Int. Geochem. Explor. Symp.-II Brazilian Geochem. Congr.,
Rio de Janeiro, Abstracts, pp. 114.
Nordenskiöld, N , 1863. Beitrag zur Kenntnis der Schrammen in Finland. Acta Soc. Sei.
Fennicae, 7: 505-543.
Nuotio, T, Hyyppä, J. and Kämäri, J., 1985. Buffering properties of soils in 53 forested
catchments in southern Finland. Aqua Fennica, 15(1): 35-40.
Nurmi, A., 1976. Geochemistry of the the till blanket at the Talluskanava Ni-Cu ore deposit,
Tervo, Central Finland. Geol. Surv. Finl., Rep. Inv, 15, 84 pp.
Nurmi, A., 1977. Explorational applications of pedogeochemistry. Acad. Diss. Synopsis, Univ.
Helsinki, 10 pp.
Nurmi, P and Isohanni, M., 1984. Rock, till and stream sediment geochemistry in the search
for porphyry-type Mo-Cu-Au deposits in the proterozoic Rautio batolith, western Finland. J.
Geochem. Explor., 20: 209-228.
Nurmi, PA., Kukkonen, I.T. and Lahermo, P.W, 1988. Geochemistry and origin of saline
ground waters in the Fennoscandian shield. Appl. Geochem,. 3(2): 185-203.
408 References

Nuutilainen, J. and Peuraniemi, V, 1977. Application of humus analysis to geochemical


prospecting: some case histories. In: Prospecting in Areas of Glaciated Terrain. Institute of
Mining and Metallurgy, London, pp. 1-5.
Nykänen, O., 1968. Kallioperäkartan selitys. Lehti-Sheet 4232-424. Tohmajärvi. Summary:
Explanation to the map of rocks. Geological map of Finland, 1:100 000. Geological Survey
of Finland, Espoo, 66 pp.
Nykänen, O., 1971a. Kallioperäkartta-Pre-Quaternary rocks. Lehti-Sheet 4241, Kiihtelysvaara.
Geological map of Finland, 1:100 000. Geological Survey of Finland, Espoo.
Nykänen, 0., 1971b. Kallioperäkarttojen selitykset-Explanation to the maps of Pre-Quaternary
rocks. Lehti-Sheet 4241, Kiihtelysvaara. Summary: Explanation to the map of rocks.
Geological map of Finland, 1:100000. Geological Survey of Finland, Espoo, 68 pp.

O'Leary, R.M. and Viets, J.G., 1986. Determination of antimony, arsenic, bismuth, cadmium,
copper, lead, molybdenum, silver and zinc in geologic materials by atomic absorption
spectrometry using a hydrochloric acid-hydrogen peroxide digestion. Atomic Spectrosc, 7:
4-8.
Oivanen, P., 1982. Antimony ore studies in the Seinäjoki-Nurmo area in the years 1975-1982.
Rep., Geol. Surv. Finland (unpublished).
Okko, V, 1941. Über das Verhältnis der Gesteinzusammensetzung der Moräne zum Felsgrund
in den Gebieten der Kartenblätter von Ylitornio und Rovaniemi in nördlichen Finnland.
Geol. Rundsch., 32: 627-643.
Ollier, C., 1969. Weatherig. Ollier and Boyd Ltd., Edinburgh, 304 pp.
Ollila, J.T., 1976. On the geology of the Mäkärärova gold occurrence and its surroundings. Rep.,
Helsinki Univ, 62 pp. (unpublished).
Osterlund, S.E., 1982. Micro boulder tracing with a portable gamma spectrometer. In:
Prospecting in Areas of Glaciated Terrain. Institution of Mining and Metallurgy, London,
pp. 240-248.
Ovenden, L., 1989. Peatlands: a leaky sink in the global carbon cycle. Geos, 18: 14-24.

Pekkarinen, L., 1976. Selostus malmitutkimuksista Kiihtelysvaaran Hyypiän ja Karsikkojärven


alueilla vuosina 1972-1974. Rep., Geol. Surv. Finland, 4 pp. (unpublished).
Peltoniemi, H., 1985. Till lithology and glacial transport in Kuhmo, eastern Finland. Boreas,
14: 67-74.
Penttilä, S., 1963. The glaciation of the Laanila area. Bull. Comm. geol. Finl., 203, 71 pp.
Perttunen, M., 1977. The lithological relation between till and bedrock in the region of
Hämeenlinna, southern Finland. Geol. Surv. Finl., Bull., 291, 68 pp.
Perttunen, M. (Editor), 1989. Transport of glacial drift in Finland. Geol. Surv. Finl., Spec.
Paper, 7, 74 pp.
Peuraniemi, V, 1982. Geochemistry of till and mode of occurrence of metals in some moraine
types in Finland. Geol. Surv. Finl., Bull., 322, 75 pp.
Peuraniemi, V, 1989. Till stratigraphy and ice movement directions in the Kittilä area, Finnish
Lapland. Boreas, 18: 145-157.
Pirttisalo, K. and Räisänen, M-L., 1988. Raskasmetallit Kuopion alueen rantasedimenteissä
vuosina 1974-1977 ja 1985-1986 (Heavy metals in Kuopio city area in central Finland in
1974-1977 and 1985-1986.) Kuopion kaupunki, ER 1988: 3, 30 pp.
Pitulko, VM., 1968. Features of geochemical searches for rare metal deposits in permafrost
area. Int. Geol. Rev., 11: 1239-1246.
References 409

Plant, J.A., Smith, R.T., Stevenson, A.G., Forrest, M.D. and Hodgson, J.F., 1984. Regional
geochemical mapping for mineral exploration in northern Scotland. In: Prospecting in Areas
of Glaciated Terrain. Institution of Mining and Metallurgy, London, pp. 103-120.
Polikarpochkin, W , Lomonosov, I.S., Kitaev, N.A., Filippova, L.A., Gapon?, A.E., Konstantinova,
I.M., Sarapulova, VN. and Belogolova, G.A., 1983. Theory and practice of geochemical
methods of prospecting according to secondary aureoles and dispersal flows. Sov. Geol.
Geophys., 24(1): 36-46.
Pollard, J., 1977. A Handbook of Numerical and Statistical Techniques with Examples Mainly
from the Life Sciences. Cambridge University Press, Cambridge, 349 pp.
Potts, P.J., 1987. A Handbook of Silicate Rock Analysis. Blackie, Glasgow, 622 pp.
Potts, P.J., Webb, PC. and Watson, J.S., 1985. Energy-dispersive X-ray fluorescence analysis of
silicate rocks: comparisons with wavelength dispersive performance. Analyst, 110: 507-513.
Prest, VK.,1968. Nomenclature of moraines and ice-flow features as applied to the glacial map
of Canada. Geol. Surv. Can., Paper, 67-57, 32 pp.
Prest, VK. and Nielsen E., 1987. The Laurentide ice sheet and long-distance transport. Geol.
Surv. Finl., Spec. Paper, 3: 91-101.
Proskuriakov, V, Katskov, A., Plisov, A., Chekuskin, V, Bogdanov, L. and Kuznetsov, V,
1980. Technique and procedure of integrated geochemical and geocphysical prospecting
for ore deposits in the Kola-Karelian region. In: H. Tanskanen (Editor), Methods of the
Geochemical Mapping and Boulder Prospecting in the Eastern Part of the Baltic Shield.
Proc. Finnish-Soviet Seminar, Leningrad, pp. 191-201.
Pulkkinen, E., Puranen, R. and Lehmuspelto, P., 1980. Interpretation of geochemical anomalies
in glacial drift of Finnish Lapland with the aid of magnetic susceptibility data. Geol. Surv.
Finl., Rep. Inv., 47, 39 pp.
Pulkkinen, E. and Rossi, S. 1984. Gradually proceeding geochemical exploration, a case history
of the discovery on Maaselkä copper-cobalt-showing in Finnish Lapland. In: Prospecting in
Areas of Glaciated Terrain. Institution of Mining and Metallurgy, London, pp. 1-5.
Punkari, M., 1979. Skandinavian jäätikön deglasiaatiovaiheen kielekevirrat Etelä-Suomessa.
English summary: The ice lobes of the Scandinavian ice sheet during the deglaciation in
South Finland. Geologi, 2: 22-28.
Punkari, M., 1984. The ice lobes of the Scandinavian ice sheet during the deglaciation in
Finland. Boreas, 9: 307-310.
Puranen, M. and Puranen, R., 1977. Apparatus for the measurement of magnetic susceptibility
and its anisotropy. Geol. Surv. Finl., Rep. Inv., 28, 46 pp.
Puranen, R., 1977. Magnetic susceptibility and its anisotropy in the study of glacial transport
in northern Finland. In: Prospecting in Areas of Glaciated Terrain. Institution of Mining
and Metallurgy, London, pp. 111-119.
Puranen, R., 1988. Modelling of glacial transport of basal tills in Finland. Geol. Surv. Finl., Rep.
Inv., 81, 36 pp.
Puranen, R. and Kivekäs, L., 1979. Grain density of till: a potential parameter for the study of
glacial transport. Geol. Surv. Finl., Rep. Inv., 40, 32 pp.
Puustinen, K., 1977. Exploration in the northwest region of the Koitelainen gabbro complex,
Sodankylä, Finnish Lapland. In: Prospecting in Areas of Glaciated Terrain. Institute of
Mining and Metallurgy, London, pp. 6-13.
Pääkkönen, V, 1966. The geology and mineralogy of the occurrence of native antimony at
Seinäjoki, Finland. Bull. Comm. Geol. Finl., 225, 70 pp.

Qvarfort, U., 1977. Some problems asssociated with exploration geochemistry in mining areas
in Sweden. Striae, 7, 77 pp.
410 References

Qvarfort, U., 1979. Sulfidmalmuppslag som miljöproblem. Rep,, Statens Naturvärdsverk (un-
published).

Raeburn, C. and Millner, H.B., 1927. Alluvial Prospecting. Thos Murby and Sons, London, 487
pp.
Rahkola, P, 1982. Report on studies in the Mäkärärova and Kiviaapa areas. Rep., Geol. Surv.
Finland (unpublished).
Ramanamurthy, M.V, 1983. Efficacy of some chemical extractions in the recognition of true and
false anomalies in geochemical exploration. J. Geol. Soc. India, 24(1): 43-53.
Rankama, K., 1940. On the use of trace elements in some problems of practical geology. C.R.
Soc. Geol. Finl., 14: 90-106.
Rankama, K. and Sahama, Th. G., 1950. Geochemistry. University of Chigaco Press, Chicago,
111., 612 pp.
Raukas, A. and Hyvärinen, H. (Editors), 1992. Geology of the Gulf of Finland. Estonian Academy
of Sciences/Academy of Finland, Tallinn, 422 pp. (in Russian with English summary).
Repo, R., 1957. Untersuchungen über die Bewegungen des Inlandeises in Nordkarelien. Bull.
Comm. Geol. Finl., 179, 178 pp.
Riddle, C., Van der Voet, A. and Doherty, W, 1988. Rock analysis using inductively coupled
plasma mass spectrometry: a review. Geostand. Newsl., 12: 203-234.
Riese, WC., Herald, C.J. and Flammang, J.A., 1986. Application of advanced statistical analysis
to interpretation of geochemical data from glaciated terrains, northern territories, Canada.
In: Prospecting in Areas of Glaciated Terrain. Institute of Mining and Metallurgy, London,
pp. 151-162.
Robbins, J.C., 1973. Zeeman spectrometer for measurement of atmospheric mercury vapour.
Int. Geochem. Explor. Symp., London, pp. 315-323.
Rock, N.M., 1987. ROBUST: An interactive FORTRAN-77 package for exploratory data
analysis using parametric, robust and nonparametric location and scale estimates, data
transformations, normality tests, and outlier assessment. Comput. Geosci., 13(5): 463-494.
Rodionov, D.A., 1971. Distribution Functions of the Elements and Mineral Contents of Igneous
Rocks (translated from Russian). Consultants Bureau, New York, N.Y., 80 pp.
Rogers, PJ. (Editor), 1988. Prospecting in areas of glaciated terrain, Halifax, Nova Scotia.
The Canadian Institute of Mining and Metallurgy and Institute of Mining and Metallurgy,
London, 644 pp.
Rogers, PJ., Bonham-Carter, G.F. and Ell wood, D.J., 1986. Anomaly enchancement by use of
catchment basin analysis of surfical geochemical data from the Cobequid Highlands, Nova
Scotia. In: Prospecting in Areas of Glaciated Terrain. Institution of Mining and Metallurgy,
London, pp. 163-174.
Rogerson, R.J., 1982. The glaciation of Newfoundland and Labrador. In: Prospecting in Areas
of Glaciated Terrain. Institute of Mining and Metallurgy, London, pp. 37-56.
Roos, S., 1981. Ilomantsin Jerusalemin molybdeeniesiintymästä. Rep., Turku Univ., 93 pp.
(unpublished).
Rose, A., Hawkes, H.E. and Webb, J.S., 1979. Geochemistry in Mineral Exploration. Academic
Press, London, 657 pp.
Rosenberg, R.J., Pitkänen, V and Sorsa, A., 1977. An automated uranium analyzer based on
delayed neutron counting. J. Radioanal. Chem., 37: 167-179.
Rouhunkoski, P. and Isokangas, P, 1974. The copper gold vein deposits of Kivimaa at Tervola,
N-Finland. Bull. Geol. Soc. Finl., 46: 29-35.
Rundqvist, D.H. (Editor), 1979. Kalitäestvennoe Prognosirovanie pri Regionalnyh Metallogenit-
seskih Issledovanijah. VSEGEI, Leningrad, 88 pp.
References All

Rosier, H.J. and Lange, H., 1972. Geochemical Tables. Elsevier, Amsterdam, 468 pp.

Saarnisto, M. and Taipale, K., 1985. Lithology and trace-metal content in till in the Kuhmo
granite-greenstone terrain, eastern Finland. J. Geochem. Explor. 24(3): 317-336.
Salmi, M., 1955. Prospecting for bog-covered ore by means of peat sampling. Bull. Comm. Geol.
Finl., 169: 5-34.
Salmi, M., 1967. Peat prospecting: applications in Finland. In: A. Kvalheim (Editor), Geochemical
Prospecting in Fennoscandia. Interscience, London, pp. 113-126.
Salminen, R., 1976. Identification of weathered bedrock in till derived from it. J. Geochem.
Explor. 5(3): 409-412.
Salminen, R., 1980. On the geochemistry of copper in the Quaternary deposits in the
Kiihtelysvaara area, north Karelia, Finland. Geol. Surv. Finl., Bull., 309, 48 pp.
Salminen, R., 1981. The effect of mining on the metal contents of streamsediments. In: K.
Puustinen (Editor), Geological, Geochemical and Geophysical Investigations in the Eastern
Part of the Baltic Shield. The Committee for Scientitic and Tecnical Co-operation between
Finland and Soviet Union, pp. 169-181.
Salminen, R. and Kokkola, M., 1984. Annual variation in metal contents of humus. In:
Prospecting in Areas of Glaciated Terrain. Institute of Mining and Metallurgy, London, pp.
171-177.
Salminen, R. and Hartikainen, A., 1985. Glacial transport of till and its influence on
interpretation of geochemical results in North Karelia, Finland. Geol. Surv. Finl., Bull., 335,
48 pp.
Salminen, R. and Hartikainen, A., 1986. Tracing of gold, molybdenum and tungsten mineraliza-
tion by use of a step by step geochemical study in Ilomantsi, eastern Finland. In: Prospecting
in Areas of Glaciated Terrain. Institute of Mining and Metallurgy, London, pp. 210-210.
Salo, A. and Voipio, A., 1973. Transport of radionuclides in lake and river systems flowing
through areas characterized by Precambrian bedrock and peat-bogs. IAEA-SM-158/12: 195-
217.
Salonen, V-P, 1986. Glacial transport distance distributions of surface boulders in Finland.
Geol. Surv. Finl., Bull., 338, 57 pp.
Salonen, V-P, 1987. Observations on boulder transport in Finland. Geol. Surv. Finl., Spec.
Paper, 3: 103-110.
Salonen, V-P and Palmu, J-P, 1989. On measuring the length of glacial transport. Geol. Surv.
Finl., Spec. Paper, 7: 25-32.
Sandström, H., 1984. Selective sequential dissolution of organic-rich stream sediments from
Talvivaara, Finland. J. Geochem. Explor., 21: 341-353.
Sanzalone, R.F. and Chao, T.T., 1981. Determination of submicrogram amounts of selenium
in geological materials by electrothermal atomic-absorption spectrophotometry after solvent
extraction. Analyst, 106: 647-652.
Sauramo, M., 1924. Tracing of glacial boulders and its application in prospecting. Bull. Comm.
Geol. Finl, 67: 1-37.
Savinskii, I.D., 1965. Probability Tables for Locating Elliptical Underground Masses with
Rectangular Grid. Consultants Bureau, New York, N.Y., 110 pp.
Schroll, E , 1975. Analytische Geochemie, Band I: Methodik. Enke Verlag, Stuttgart, 292 pp.
Schuddebeurs, A.P, 1981. Results of counts of Fennoscandinavian erratics in the Netherlands.
Meded. Rijks Geol. Dienst, 34(3): 10-14.
Seifert, G, 1954. Das Mikroskopische Korngefüge des Geschiebemergels als Abbild der Eisbe-
wegung, zugleich Geschichte des Eisabbaues in Fehmarn, Ostwagrien und der Dänischen
Wohld. Meyniana, 2: 124-154.
412 References

Sergeev, Ye.A., 1941. Geochemical method for prospecting of ore deposits (in Russian).
Translation in: Selected Russian Papers on Geochemical Prospecting for Ores. U.S. Geological
Survey, pp. 15-87.
Sergeev, Ye.A. and Solovov, A.P., 1937. The ionic method of geophysical prospecting (in Russian).
Translation in: Selected Russian Papers on Geochemical Prospecting for Ores. U.S. Geological
Survey, pp. 89-96.
Shen, R., 1989. Simultaneous determination of 37 elements in geochemical samples by
spectrographic method using a specially designed covered electrode. In: X. Xie and S.E.
Jennes (Editors), J. Geochem. Explor., Spec. Issue, 33: 215-236.
Shepherd, A., Harvey, P.K. and Leake, R.C., 1987. The geochemistry of residual soils as an aid
to geological mapping: a statistical approach. J. Geochem. Explor., 29: 317-331.
Shetsen, L, 1984. Application of till pebble lithology to the differentiation of glacial lobes in
southern Alberta. Can. J. Earth Sei, 21: 920-933.
Shuts, WW, 1971. Till studies and their application to regional drift prospecting. Can. Min. J ,
92: 45-50.
Shuts, WW, 1972. Drift prospecting in the Kaminak Lake area, District of Keewatin. Geol.
Surv. Can., Paper, 72-A1: 182-189.
Shuts, WW, 1973. Glacial dispersal of rocks, minerals and trace elements in Wisconsinan till,
southeastern Quebec, Canada. Geol. Soc. Am., Mem., 136: 189-219.
Shuts, WW, 1975. Principles of geochemical exploration for sulphide deposits using shallow
samples of glacial drift. Can. Inst. Min. Metall., Bull., 68: 73-80.
Shuts, WW, 1976. Glacial till and mineral exploration. In: R.F. Legget (Editor), Glacial Till. R.
Soc. Can, Spec. Publ, 12: 205-225.
Shuts, WW, 1978. Nature and genesis of mudboils, central Keewatin, Canada. Can. J. Earth
Sei, 15(7): 1053-1068.
Shuts, WW, 1982. Glacial dispersal; principles and practical applications. Geosci. Can, 9(1):
42-47.
Shuts, WW, 1984a. Important principles and recent research in drift prospecting. In: Till
Tomorrow '84, Kirkland Lake, Ont. Canadian Institute of Mining and Metallurgy, pp. 1-3.
Shuts, WW, 1984b. Till geochemistry in Finland and Canada. J. Geochem. Explor, 21: 95-117.
Shuts, WW and Smith, S.L, 1986. Stratigraphy of placer gold deposits, overburden drilling in
Chaudiere Walley, Quebec. Geol. Surv. Can, Paper, 86- 1A: 703-712.
Shuts, WW, Aylsworth, J.M, Kaszycki, C. and Klassen, R.A, 1987. Canadian Shield. In: WL.
Graf (Editor), Geomorphic System of North America. Geol. Soc. Am, Spec. Vol., 2: 119-161.
Shilts, WW. and Smith, S.L, 1989. Drift prospecting in the Appalachians of Estrie-Beauce,
Quebec. In: R.N.W DiLabio and WB. Coker (Editors), Drift Prospecting. Geol. Surv. Can,
Paper, 89-20: 41-60.
Shilts, WW and Kettles, I.M, 1989. Geology and acid rain in eastern Canada. Geos, 18(3):
25-32.
Sibrava, V and Billard, A, 1986. Correlation Chart of European Glaciations. Quaternary Science
Reviews, 5. Pergamon, Oxford.
Siegel, R, 1974. Applied Geochemistry. John Wiley and Sons, New York, N.Y, 535 pp.
Sillanpää, M. and Lakanen, E , 1966. Readily soluble trace elements in Finnish soils. Ann.
Agric. Fenn, 5: 298-304.
Silvennoinen, A, Honkamo, M, Juopperi, H , Lehtonen, M, Mielikäinen, P, Perttunen, V,
Rastas, P, Räsänen, J. and Väänänen, J, 1980. Main features of the stratigraphy of northern
Finland. In: A. Silvennoinen (Editor), Jatulian Geology in the Eastern Part of the Baltic
Shield. Proc. Finnish-Soviet Symp, Finland, 21-26 August 1979, pp. 153-162.
References 413

Simonson, R.V 1957, What soils are. 1957 Yearbook of Agriculture, U.S. Department of
Agriculture, pp. 17-31.
Sinclair, A.J., 1976. Applications of probability graphs in mineral exploration. Assoc. Explor.
Geochem., Spec. Vol., 4, 95 pp.
Sinclair, A.J., 1983. Univariate analysis. In: R.J. Howarth (Editor), Statistics and Data Analysis
in Geochemical Prospecting. Handbook of Exploration Geochemistry, Vol. 2. Elsevier,
Amsterdam, pp. 59-81.
Sinding-Larsen, R., Berner, H., Ekström, T, Gustavsson, N, Larkin, S., Nilsson, G., Saltikoff,
B., Strand, G., Taipale, K. and Tontti, M., 1986. The Mineral Resource Assessment Map
of Northern Fennoscandia. Nordkalott Project. Geological Surveys of Finland, Norway and
Sweden, Espoo.
Sippola, J. and Tares, T, 1978. The soluble content of mineral elements in cultivated Finnish
soils. Acta Agr. Scand., Suppl., 20: 11-25.
Slavin, W, 1984. Graphite Furnace AAS: A Source Book. Perkin-Elmer Corp., Ridgefield, 230
pp.
Smee, B.W,1983. Laboratory and field evidence in support of the electrogeochemically enhanced
ionic diffusion through glaciolacustrine sediment. Int. Geochem. Explor. Symp., Saskatoon,
pp. 277-304.
Smee, B.W 1987. Soil sampling strategies in an area with alpine glaciation, British Columbia,
Canada. In: A.A. Levinson, PM.D. Bradshaw and I. Thomson (Editors), Practical Problems
in Exploration Geochemistry. Applied Publishing Company, 269 pp.
Smee, B.W. and Sinha A.K., 1979. Geological, geophysical and geochemical considerations for
exploration in clay-covered areas: a reviev. Can. Inst. Min. Metall., Bull., 72(804): 67-82.
Smith, R.T and Gallagher, M.J., 1975. Geochemical dispersion trough till and peat from
metalliferous mineralization in Sutherland, Scotland. In: Prospecting in Areas of Glaciated
Terrain. Institute of Mining and Metallurgy, London, pp. 134-149.
Solovov, A.P, 1959. The Theory and Practice in Metallometric Surveys. Akad. Nauk Kazakhskoy
S.S.R., Alma-Ata (in Russian).
Solovov, A.P, 1987. Geochemical Prospecting for Mineral Deposits. Mir, Moscow, 288 pp.
Sondag, F., 1981. Selective extraction procedures applied to geochemical prospecting in an area
contaminated by old mine workings. J. Geochem. Explor., 15: 645-652.
Sopuck, V, Schreiner, B.T. and Averill, S., 1986. Drift prospecting for gold in the southeastern
shield of Saskatchewan, Canada. In: Prospecting in Areas of Glaciated Terrain. Institute of
Mining and Metallurgy, London, pp. 217-240.
Soveri, J., 1985. Influence of meltwater on the amount and composition of ground water in
Quaternary deposits in Finland. Publ. Water Res. Inst., 63, 92 pp.
Späth, H., 1980. Cluster Analysis Algorithms for Data Reduction and Classification of Objects.
Ellis Horwood Ltd., 226 pp.
Stanley, C.R. and Sinclair, A.J., 1989. Comparison of probability plots and the gap statistic
in selection of thresholds for exploration geochemistry data. J. Geochem. Explor., 32(1/3):
355-357.
Stanton, R.E., 1966. Rapid Methods of Trace Analysis for Geochemical Applications. Edward
Arnold, London, 103 pp.
Stanton, R.E., 1976. Analytical Methods Used in Geochemical Exploration. Edward Arnold,
London, 54 pp.
Stary, J., 1964. The Solvent Extraction of Metal Chelates. MacMillan, New York, NY, 240
pp.
Stea, R.R., Day, T.E. and Ryan, R.J., 1986. Till and bedrock Cu-Pb-Zn geochemistry in northern
414 References

Nova Scotia and its metallogenic implications. In: Prospecting in Areas of Glaciated Terrain.
Institute of Mining and Metallurgy, London, pp. 241-260.
Stea, R.R., Turner, R.G., Fink, P.W and Groves, R.M., 1989. Glacial dispersal in Nova Scotia: a
zonal concept. Geol. Surv. Can., Paper, 89-20: 155-169.
Stepanova, M.D., 1976. Microelements in Organic Matter of Soils. Izd. Nauka, Novosibirsk, 105
pp.
Stewart, R.A. and van Hees, E.H., 1982. Evaluation of part-producing gold mine properties
by drift prospecting: an example from Matachewan, Ontario, Canada. In: E.B. Evenson,
Ch.Schlüchter and J. Rabassa (Editors), Tills and Related Deposits. INQUA Symp. U.S.A.
1981/Argentina 1982, pp. 179-193.
Stoch, H. and Dixon, K., 1983. Manual of analytical methods used in MINTEK. Council for
Mineral Technology, Spec. Publ., No. 4, Randburg (RSA), 314 pp.
Stone, P. and Gallagher, M.J., 1984. Mineral exploration in lower paleozoic turbidites of south
Scotland. In: Prospecting in Areas of Glaciated Terrain. Institute of Mining and Metallurgy,
London, pp. 201-211.
Strauss, M.F., Story, S.L. and Mehlhorn, N.E., 1989. Applications of dual-wall reverse-circulation
drilling in ground water exploration and monitoring. Ground Water Monitoring Rev., 9:
63-71.
Ström, K.M., 1937. Land-locked waters. Hydrography and bottom deposits in badly-ventilated
(sic) Norwegian fjords, with remark upon sedimentation under anaerobic conditions. Nors.
Vidensk.-Akad., Oslo, Math. Naturw. KL, Skr., Vol. 1, 85 pp.
Sturhahn, H.H. and Otto, H.J., 1974. Erosion measurements and comminution in centrifugal-
type ball mills. Chem.-Anlagen and Verfahren, 12: 37-42.
Sulcek, Z. and Povondra, P., 1989. Methods of Decomposition in Inorganic Analysis. CRC Press,
Boca Raton, Fla., 325 pp.
Sundborg, Ä., 1956. The river Klarälven, a study of fluvial process. Geogr. Ann., 38: 127-316.
Sveshnikov, G.B. and Ryss, Yu.S., 1964. Electrochemical processes on sulphide deposits and
their geochemical significance. Geochem. Int., 1964: 198-204.
Szabo, N.L., Govett, G.J.S. and Lajtai, E.Z., 1975. Dispersion trends of elements and indicator
pebbles in glacial till around M. Pleasant, New Brunswick, Canada. Can. J. Earth Sei., 12:
1534-1556.
Szalay, A., 1958. The significance of humus in the geochemical enrichment of uranium. Proc.
2nd U.N. Int. Conf. Peaceful Uses Atom Energy, Geneva, 1731.
Säynäjärvi, K, 1958. Unpublished report on studies in the Mäkärärova area. Archives of
Suomen Malmi Oy.

Tamminen, P and Starr, M.R., 1990. A survey of forest soil properties related to soil acidification
in southern Finland. In: P. Kauppi, K. Kenttämies and P. Anttila (Editors), Acidification in
Finland. Springer-Verlag, Berlin, 1237 pp.
Tanner, V, 1915. Studier öfver kvartärsystemet i Fennoscandias nordliga delar. III. Om
landisens rörelser och afsmältning i finska Lapland och gränsande trakter. Resume: Etudes
sur le Systeme quaternaire dans les parties septentrionales de la Fennoscandia. III. Sur la
progression et le cours de la recession du glacier continental dans la Laponie finlandaise et
les regions environnantes. Bull. Comm. G6ol. Finl., 38, 815 pp.
Tanskanen, H., 1976. Rookkijärvi: the nickel/copper ratio in peat in gabbro-peridotite area. J.
Geochem. Explor., 5(3): 309-310.
Tavela, M., 1957. The study of sulphide minerals of till in prospecting. Rep., Helsinki Univ., 104
pp. (unpublished).
References 415

Tenhola, M., 1988. Regional geochemical mapping based on lake sediments in eastern Finland.
In: Prospecting in Areas of Glaciated Terrain. Institute of Mining and Metallurgy, London,
pp. 305-331.
Theobald, PK, Jr., 1957. The gold pan as a quantitative geological tool. U.S. Geol. Surv., Bull.,
1071A: 1-54.
Theobald, P.K., Jr. and Thompson, C.E., 1959. Geochemical prospecting with heavy mineral
concentrates used to locate a tungsten deposit. U.S. Geol. Surv., Circ, 411, 13 pp.
Thompson, M., 1983. Control Procedures in Geochemical Analysis. In: R.J. Howarth (Edi-
tor), Statistics and Data Analysis in Geochemical Prospecting. Handbook of Exploration
Geochemistry, Vol. 2. Elsevier, Amsterdam, pp. 39-58.
Thompson, M. and Howarth, R.J., 1978. A new approach to the estimation of analytical
precision. J. Geochem. Explor., 9: 23-30.
Thompson, M. and Walsh, J.N., 1983. A Handbook of Inductively Coupled Plasma Spectrometry.
Blackie, Glasgow, 273 pp.
Thompson, M., Pahlavanpour, B., Walton, S.J. and Kirkbright, G.F., 1978. Simultaneous
determination of trace concentrations of arsenic, antimony, bismuth, selenium and tellurium
in aqueous solution by introduction of gaseous hydrides into an inductively coupled plasma
source for emission spectrometry. Analyst, 106: 568-579 and 705-713.
Thornton, I., 1988. Soil features and human health. In: G. Grube and B. Herrman (Editors),
Trace Elements in Environmental History. Springer-Verlag, Heidelberg, pp. 135-144.
Tilas, D., 1743. Tankar om malmletande, i anledning af löse stenar. K. Sver. Vet. Akad. Handl.,
Vol. I, 1939-1940. 2. Aulf., Stockholm 1974, pp. 190-193.
Toverud, Ö., 1979. Humus: a new sampling medium in geochemical prospecting for tungsten in
Sweden. In: Prospecting of Areas in Glaciated Terrain. Institute of Mining and Metallurgy,
London, pp. 74-79.
Toverud, Ö., 1984. Dispersal of tungsten in glacial drift and humus in Bergslagen, southern
central Sweden. J. Geochem Explor., 21(1/3): 261-272.
Tukey, J.W, 1977. Exploratory Data Analysis. Addison-Wesley, Reading, Mass., 273 pp.
Tukey, PA. and Tukey, J.W, 1981. Summarization; smoothing; supplemented views. In: V
Barnett (Editor), Interpreting Multivariate Data. John Wiley and Sons, Chichester, pp.
245-275.
Tuominen, H.V and Mikkonen, A.A., 1955. Unpublished report on the studies in the Mäkärärova
area, Archives of Suomen Malmi Oy.

Ure, A.M., 1975. The determination of mercury by non-flame atomic absorption and fluorescence
spectrometry — a review. Anal. Chim. Acta, 76: 1-27.

Vaasjoki, M. and Gulson B.L., 1985. Evaluation of drilling priorities using lead isotopes at the
Lady Loretta lead-zinc-silver deposit, Australia. J. Geochem. Explor., 24(3): 305-316.
Vaasjoki, M. and Gulson, B.L., 1986. Carbonate-hosted base metal deposits: lead isotope data
bearing on their genesis and exploration. Econ. Geol., 81: 156-172
Van Loon, J.C. and Barefoot, R.R., 1989. Analytical Methods for Geochemical Exploration.
Academic Press, San Diego, Calif., 344 pp.
Virkkala, K, 1958. Stone counts in the esker of Hämeenlinna, southern Finland. Bull. Comm.
Geol. Finl., 180: 87-103.
Virkkala, K., 1969. Maaperäkartan selitys. Lehti 2131 Hämeenlinna. Summary: Explanatory
text to the map of Quaternary deposits. Suomen geologinen kartta 1:100000, 69 pp.
Virkkala, K., 1971. On the lithology and provenance of the till of a gabbro area in Finland. VIII
Int. Congr., INQUA, Paris 1969, pp. 711-714.
416 References

Virtanen, K., 1991. Effect of bedrock on trace element concentration of fuel peat. In: E.
Pulkkinen (Editor), Environmental Geochemistry in Northern Europe. Geol. Surv. Finl.,
Spec. Paper, 9: 241-246.
Vogt, T., 1939. Kjemisk og botanisk malmletning vid Röros. Summary: Chemical and botanical
ore prospecting in the Röros area. K. Nor. Vidensk. Förh., 12: 81-84.
Vogt, T. and Bergh, H., 1943. Geokjemisk and geobotanisk malmletning,IX. Bestemmelse Cu,
Zn, Pb, Mn og Fe i planter fra Röros-feltet. K. Nor. Vidensk. Förh., 16: 55-58.
Vuorinen, A. and Lahermo, P, 1986. Enviromental factors affecting the acid neutralization
capacity of Finnish glacial till deposits. Toxicol. Environ. Chem., 11: 61-78.
Vuorinen, A., Vuorela, I. and Welinder, S., 1988. Pollution of lakes in a former mining and
smeltering area: Evidence from successive extraction and pollen analysis of lake sediments,
Part 1. Lake Lissjön; Part 2. Lake Dammsjön. Int. Environ. Anal. Chem., 34: 265-297.
Väänänen, P, 1976. Eronlampi: uranium and radiation determinations from till samples. J.
Geochem. Explor., 5(3): 214.

Waldron, T., 1988. The heavy metal burden in ancient sediments. In: G. Grube and B. Herrman
(Editors), Trace Elements in Environmental History. Springer-Verlag, Heidelberg, pp. 125-
133.
Walsh, J.N., Buckley, F. and Barker, J., 1981. The simultaneous determination of the rare earth
elements in rocks using inductively coupled plasma source spectrometry. Chem. Geol., 33:
141-153.
Ward, F.N., Lakin, H.W., Canney EC, 1963. Analytical methods used in geochemical exploration
by the U.S. Geological Survey. U.S. Geol. Surv., Bull., 1152, 100 pp.
Warren, H.V and Delavault, R.E., 1956. Soils in geochemical prospecting. Min. Eng., 8(10):
992-996.
Webb, J.S., 1958. Observations on geochemical exploration in tropical terrains. Int. Geol. Congr.,
Mexico, 1, pp. 143-173.
Welsh, E.P, 1979. Determination of arsenic in geologic materials using silver diethyldithiocar-
bamate. U.S. Geol. Surv., Open-File Rep., 79-1442, 10 pp.
Welsh, E.P, 1983. A rapid geochemical spectrophotometric determination of tungsten with
dithiol. Talanta, 30: 876-878.
Welsh, E.P. and Chao, T.T., 1976. Determination of trace amounts of tin in geological materials
by atomic absorption spectrometry. Anal. Chim. Acta, 82: 337-342.
Wennervirta, H., 1968. Application of geochemical methods to regional prospecting in Finland.
Bull. Comm. Geol. Finl., 234, 41 pp.
Wennervirta, H., Bölviken, B. and Nilsson, C.A., 1971. Summary of research and development
in geochemical exploration in Scandinavian countries. Can. Inst. Min. Metall., Spec. Vol., 11:
11-14.
White, WH. and Allen, T.M., 1954. Copper soil anomalies in the Boundary District of British
Columbia. Min. Eng., 6: 49-52.
Wilhelm, E. and Kosakevitch, A., 1979. Utilisation des chapeux de fer comme guides de
prospection. Bull. B.R.G.M., II, 2-3: 109-140.
Wilhelm, E., Laville-Timsit, L., Leleu, M., Cachau-Herreillat, F. and Capdecomme, H., 1979.
Behaviour of base metals around ore deposits: application to geochemical prospecting in
temperate climates. In: J.R. Watterson and PK. Theobald (Editors), Geochemical Exploration
1978. Association of Exploration Geochemists, pp. 185-199.
Willie, S.N., Sturgeon, R.E. and Berman, S.S. 1986. Hydride generation atomic absorption deter-
mination of selenium in marine sediments, tissues, and seawater with in situ concentration
in a graphite furnace. Anal. Chem., 58: 1141-1143.
References All

Willis, J.P., 1985. Applications of X-ray fluorescence spectrometry and electron microprobe in
the exploration, mining and processing of materials. In: L.R.P. Butler (Editor), Analytical
Chemistry in the Exploration, Mining and Processing of Materials, 2nd Int. Symp., IUPAC,
Pretoria, 15-19 April 1985. Blackwell, Oxford, pp. 45-56.

Xie, X., Sun, H. and Li, S., 1981. Geochemical exploration in China. J. Geochem. Explor., 15:
489-506.
Xie, X. and Zheng, K., 1983. Recent advances in geochemical exploration in China. J. Geochem.
Explor., 19: 423-443.

Yfantis, E.A., Flatman, G.T. and Behar, J.V, 1987. Efficiency of kriging estimation for square,
triangular, and hexagonal grids. Math. Geol., 19(3): 183-205.

Zheng, J., 1984. Tentative geochemical rock survey in an impregnation type Nb-Ta granite
terrain of South China. Geophys. Geochem. Explor., 8(4): 212-222.
Zhou, Di., 1987. Robust statistics and geochemical data analysis. Math. Geol., 19(3): 207-218.
Zhou, L., Chao, T.T. and Meier, A.L., 1984. Determination of indium in geological materials
by electrothermal-atomization atomic absorption spectrometry with a tungsten-impregnated
graphite furnace. Anal. Chim. Acta, 161: 369-373.
419

REFERENCES INDEX

Aario, R., 31, 277, 391 Belyaev, Yu., 190, 392


Agha, M., 235, 391 Bentley, C.R., 391
Agricola, G., 391 Bergh, H., 61, 378, 416
Agterberg, F.P., 262, 391 Bergström, J , 394
Aho, L, 318, 391 Berman, S.S., 416
Ahrens, L.H., 208, 237, 391 Berner, H., 413
Alfthan, A., 400 Bernhardi, R., 4
Alhonen, P., 81, 391 Beus, A.A., 5, 93, 365, 392
Allen, T.M., 5, 416 Bezdek, J.C, 257, 392
Alley, R.B., 24, 391 Bezukladnov, A.A., 404
Alviola, R., 322, 391 Bibby, J.M., 332, 391
Anderberg, M.R., 255, 391 Billard, A, 412
Anderson, J., 206, 391 Billings, G.K., 325, 392
Anderson, T.W, 366, 391 Björklund, A., 5, 31, 153, 242, 245, 259, 322,
Andrew, R.L., 55, 373, 376, 392 370, 372, 379, 380, 388, 392, 394, 399,
Andrews, D.F., 249, 391 407
Andrews, J.T., 374, 391 Blain, C.F., 55, 373, 376, 392
Andrews, M.J, 203, 332, 333, 391, 397 Blankenship, D.D., 391
Antropova, N.T., 60, 61, 150, 377, 378, 391 Bloom, H , 5, 360, 378, 392
Armands, G, 67, 348, 377, 391 Bock, R, 195, 392
Armour-Brown, A., 5, 370, 392 Bogdanov, L., 409
Aucott, J.W, 368, 392 Bonbenkov, VN., 404
Aurola, E., 4, 392 Bonham-Carter, G.F., 410
Averill, S, 413 Borns, H.W, Jr., 406
Aylsworth, J.M., 412 Bouchard, M.A, 128, 129, 131, 132, 134, 137,
Äyräs, M., 60, 310, 355, 392, 403, 407 393
Boulton, G.S, 25, 130, 133, 375, 393
Barbier, J, 5 Bouma, J., 52, 393
Barefoot, R.R., 186, 415 Box, G.E.P, 232, 393
Barker, J., 416 Boyle, D.R, 375, 393
Barnes, R.M., 209, 392 Boyle, R.W, 3, 5, 364, 384, 393
Barnett, V, 239, 392 Bradshaw, P.M.D, 5, 31, 102, 103, 122, 127,
Batterson, M.J., 366, 374, 392 144, 387, 393
Beaumier, M., 53, 404 Brameld, M.P, 406
Bedard, P, 136, 395 Breiman, L, 257, 393
Behar, J.V, 417 Brent, E.E, 247, 394
Belogolova, G.A., 409 Brown, B.W, 319, 394
420 References index

Brummer, G., 51, 52, 394 Denoyer, E., 211,595


Brundin, N.H., 370, 394 Diallo, A, 209, 392
Brune, D, 214, 394 Dietrich, J.A., 406
Buckley, F, 416 Dijkstra, A., 405
Bölviken, B., 8, 63, 119, 144, 150, 221, 237, DiLabio, R.N.W., 5, 31, 37, 44, 66, 67, 76, 117,
370, 376, 394, 416 127, 279, 371, 374, 377, 385, 395, 402
Dixon, K, 206, 414
Cachau-Herreillat, F., 416 Doherty, W, 410
Cadieux, B., 393 Dormer, J.J., 141, 396, 398
Cameron, E.M., 5, 394 Drake, L.D., 40, 63, 66, 102, 129, 147, 396
Canney F.C., 416 Dreimanis, A., 5, 21, 22, 23, 27, 29, 30, 31,
Capdecomme, H., 416 97, 105, 127, 128, 130, 375, 396
Carpenter, R.H., 117, 394 Drewry, D., 132, 396
Cazalet, P.C.D., 63, 90, 91, 120, 150, 394, 406 Dubey, R., 404
Chao, T.T., 124, 205, 206, 387, 394, 401, 411, Duchaufour, P., 62, 79, 81, 83, 86, 396
416, 417 Duda, R., 255, 396
Chekuskin, V, 409 Dybczynski, R., 406
Cherduville, G., 395 Dyck, W, 122, 181, 396
Chisholm, E.O., 4, 394
Chork, C.Y., 119, 120, 240, 394, 398 Eadington, PJ., 279, 403
Christ, C.L., 116, 373, 397 Easterbrook, D.J., 375, 396
Chwastowska, J., 406 Ediger, R., 395
Clayton, L., 406 Edlund, S., 366, 396
Clayton, PM., 394 Ehlers. J., 13, 31, 396
Clews, D.R., 393 Ek, J., 72, 396
Closs, L.G., 383, 394 Ekström, T., 413
Cogger, N , 203, 394 Elliott, I.L, 279, 401
Ellwood, D.J., 410
Coker, W.B., 5, 31, 37, 66, 67, 76, 371, 395
Eriksson, G., 67, 68, 396
Cole, D.R., 387, 395
Eriksson, K, 24, 31, 67, 68, 76, 374, 378, 396
Cook, R.J.B., 396
Ermengen, S.V, 5, 396
Corden, S.G., 398
Ervamaa, P, 360, 396
Cormier, R., 5, 398
Erviö, R., 388, 396
Cox, D.R., 232, 393
Everitt, B.S., 238, 257, 396
Cragg, C.B., 5, 393
Eyles, N., 67, 130, 397
Crow, E.L, 238, 395
Currie, L.A., 189, 395
Fan, F, 399
Filippova, L.A., 409
Danielsson, A., 208, 395 Fink, PW, 414
Date, A.R., 210, 395 Flammang, J.A., 410
David, P.P., 136, 395 Flatman, G.T., 417
Davis, J.C., 255, 395 Fletcher, W.K., 124, 186, 197, 377, 397, 400,
Day, T.E., 413 404
De Geer, G., 32, 395 Flint, R.F., 4, 13, 100, 110, 397
De Geoffroy, J., 371, 395 Forkman, B., 394
De Vos, W., 335, 395 Forrest, M.D., 409
De Walque, L., 387, 395 Forsskäl, J., 3, 397
Delavault, R.E., 5, 416 Fortey, N.J., 397
Delecour, F., 86, 395 Foster, J.R., 199, 397
References index 421

Frick, C, 196 Hager, J , 395


Friedman, G.M., 44, 46, 48, 52, 53, 54, 397 Hakala, P, 25
Friedman, J.H, 393 Haldorsen, S, 25, 105, 399
Friedrich, G.H., 53, 397 Hale, M, 51, 399
Fryer, B.J., 405 Halicz, L, 209, 399
Fuge, R., 203, 332, 333, 391, 397 Hall, G.E.M, 211,400
FuUertone, D.S., 406 Rail, R.D, 373, 399
Fulton, R.B., 5, 397 Halonen, O, 7, 145, 384, 400
Fyfe, WS, 397 Harbaugh, J.W, 5, 400
Hart, P, 255, 396
Gabriel, K.R., 397 Hartikainen, A, 101, 104, 107, 108, 127, 130,
Gallagher, M.J, 69, 343, 389, 397, 406, 413, 131, 134, 135, 141, 144, 157, 285, 322,
414 370, 400, 411
Gapon, A.E, 409 Harvey, P.K., 412
Garreis, R.M, 116, 373, 397 Hausen, H , 4, 400
Garrett, R.G, 5, 203, 219, 222, 228, 232, 258, Hawkes, H.E, 4, 5, 82, 83, 84, 85, 93, 400,
370, 397, 398, 407 410
Gatehouse, S, 124, 398 Hedström, H , 4, 400
Gedcke, D, 401 Hellaakoski, A, 375, 400
Geddes, R.S., 31, 375, 398 Helmersen, G.V, 4, 400
Gillberg, G, 129, 130, 131, 132, 134, 135, Herald, C.J, 410
141, 398 Herms, U, 394
Ginzburg, I.I, 5, 144, 374, 398 Hilchey, G.R, 319, 394
Gleeson, C.F, 5, 31, 371, 375, 378, 398 Hirvas, H , 26, 58, 134, 135, 139, 297, 318,
Glen, J.W, 25, 398 371, 375, 400
Gluckert, G, 32, 398 Hjulström, F, 46, 400
Goldschmidt, VM, 4, 60, 376, 398 Hodgson, J.F., 409
Goldthwait, R.E, 4, 97, 398 Hoffman, S.J, 124, 400
Gonzales, R.C., 246, 398 Holman, R.H, 5, 400
Gould, R.W, 401 Holmes, A, 110, 401
Goutier, F, 393 Holmes, C.W, 4, 131, 401
Govett, G.J.S, 5, 60, 63, 119, 120, 365, 376, Honkamo, M, 412
378, 398, 414 Horsnail, R.F, 279, 371, 378, 401
Granath, G, 40, 52, 53, 257, 369, 376, 399 Hosking, K.F.G, 322, 401
Granlund, E , 405 Hosmer, D.W., 235, 401
Gray, A.L, 210, 395 Howarth, R.J, 189, 242, 252, 255, 401, 415
Graybill, FA, 232, 399 Huber, P.J., 240, 401
Grigorian, S.V, 5, 93, 365, 392 Hubert, A E , 205, 394, 401
Grip, E , 4, 399 Huff, L.C, 4, 5, 401
Groves, R.M, 413 Huhta, P, 385, 401
Gubaö, J , 238, 399 Hyvärinen, H , lb, 410
Gulson, B.L, 40, 55, 56, 369, 376, 388, 399, Hyvärinen, L, 27, 29, 52, 127, 136, 139, 374,
415 401
Guo, X, 205, 399 Hyyppä, E , 34, 401
Gustavsson, B , 137, 399 Hyyppä, J , 407
Gustavsson, N , 145, 153, 242, 256, 259, 298, Högbom, A, 4, 401
368, 369, 370, 372, 380, 387, 388, 392,
399, 403, 413 Ibrahim, M.T, 235, 391
Gy, P.M., 222, 399 Ilmasti, M, 404
422 References index

Ilvonen, E., 310 Kontas, E., 193, 199, 205, 206, 270, 271, 306,
Isohanni, M., 6, 407 318, 319, 403, 407
Isokangas, P., 304, 410 Kontio, M., 256, 370, 394, 399
Korpela, K., 15, 58, 305, 375, 403
James, C.H., 93, 401 Kosakevitch, A., 55, 388, 416
Jarvis, K.E, 210, 401 Kovalevskii, A.L., 65, 403
Jenkins, R., 213, 401 Koveshnikova, T.A., 190, 392
Jenny, E.A., 279, 402 Kristiansson, K, 63, 120, 150, 405
Joensuu, 0., 5 Krumbein, W.C., 127, 128, 130, 403
Johansson, P., 2, 168 Krylova, L.Ya., 332, 403
Johnson, C.C., 332, 391 Kujansuu, R., 139, 403
Johnson, WM, 186, 402 Kukkonen, LT., 407
Juntunen, R., 404 Kuznetsov, V, 409
Juopperi, H., 412 Kvalheim, A., 5, 31, 37, 348, 404
Jöreskog, K.G., 250, 402 Kämäri, J., 407
Königsson, L.-K., 37, 404
Körner, R., 365, 367, 404
Kabata-Pendias, A., 65, 85, 87, 402
Kahma, A., 122, 152, 402 Lagerbäck, R., 28
Kallio, E., 407 Lahermo, P, 80, 370, 404, 407, 416
Kalm, P, 4 Lahti, H.R., 91, 404
Kankainen, T., 25, 366 Lajtai, E.Z., 414
Kaszycki, CA., 377, 402, 412 Lakanen, E., 387, 388, 396, 412
Katskov, A., 409 Lakin, H.W, 4, 400, 416
Kaufman, L., 405 Lalonde, J.P, 53, 404
Kauranne, L.K., 4, 5, 10, 25, 31, 37, 40, 42, Lambe, R.N., 406
59, 62, 74, 76, 80, 88, 94, 102, 103, 127, Lange, H., 115, 411
131, 134, 145, 150, 279, 298, 318, 344, Larkin, S., 413
359, 360, 370, 374, 375, 377, 378, 381, Larsson, J.O., 393
385, 401, 402, 403 Lavikainen, S., 157, 404
Kauranne, L-M., 5, 73, 381, 403 Laville-Timsit, L., 338, 339, 366, 376, 404,
Kettles, I.M., 367, 412 405, 416
Kinealy, K., 279, 403 Lawley, D.N, 252, 404
Kinnunen, K, 53, 369, 375, 403 Lawson, A., 365, 404
Kirkbright, G.R, 415 Leake, R.C., 412
Kitaev, N.A., 409 Lecomte, P., 405
Kivekäs, E.J., 4, 403 Lee, H.A., 4, 37, 129, 404
Kivekäs, L., 129, 409 Legkova, O.E., 404
Klassen, R.A., 375, 403, 412 Legkova, VG., 366, 404
Klovan, J.E., 402 Lehmuspelto, P., 40, 325, 404, 409
Knox, K.S., 396 Lehtonen, M., 412
Kobayashi, J., 73, 74, 403 Leleu, M., 416
Koivisto, T, 302, 305, 403 Lepeltier, C, 238, 404
Kokko, J., 263, 403 Lestinen, P., 57, 283, 406, 407
Kokkola, M., 60, 69, 88, 170, 173, 175-177, Lett, R.E.W., 377, 404
272, 274, 290, 293, 378, 403, 411 Levinson, A.A., 5, 279, 374, 404
Koksoy, M., 122, 393 Lewis, T, 239, 392
Koljonen, T, 245, 370, 403 Leymarie, P, 219, 398
Konstantinova, I.M., 409 Li, S., 417
References index 423

Li, T., 319, 322, 404 Minczewski, J., 206, 406


Linden, A., 131, 132, 404 MineU, H., 131, 134, 137, 374, 399, 406
Logn, O., 63, 119, 376,594 Minkkinen, E, 222, 406
Lomonosov, I.S., 408 MitcheU, R.L., 208, 406
Longerich, H.P., 211, 405 Moretti, F.J., 396
Lounamaa, J., 64, 405 Morris, J.C., 114, 406
Luckman, B.H., 367, 405 Mutanen, T., 296, 406
Lukashev, K.I., 376, 405 Myers, D.E., 243, 406
Lukashev, VK, 55, 60, 61, 89, 376, 377, 405 Mäkinen, J., 57, 406
Lummaa, M., 245, 392 Mäkitie, 0., 387, 407
Lundbohm, Hj., 4, 405 Mölder, K, 318, 407
Lundgren, R, 395
Lundqvist, G., 4, 25, 405 Nairis, B., 370, 394
Lundqvist, J., 13, 31, 405 Nenomen, J., 179
Lynch, J.J., 203, 397 Nenonen, K., 58, 139, 371, 375, 400
Lüttig, G., 37, 405 Neuvonen, K.J., 318, 407
Nichol, I., 5, 7, 31, 370, 371, 375, 384, 388,
Maceachern, I.J., 376, 405 392, 397, 398, 407
Magnusson, N.H., 20, 405 Nielsen E., 141,409
Malmqvist, L., 63, 120, 150, 405 Niemelä, J., 139, 403
Manner, R., 305, 403
Nieminen, K, 346, 407
Marcussen, Ib., 132, 405
Nikitichev, A.P., 366, 407
Maritz, J.S., 233, 249, 405
Nikkarinen, M., 117, 118, 320, 322, 383, 407
Markowicz, A.A., 213, 405
Nilsson, C.A, 416
Martin, H., 387, 395, 405
Nilsson, G., 413
Martinau, G., 132, 393
Niskavaara, H., 206, 407
Mason, B., 95, 405
Noras, E, 199, 205, 369, 385, 399, 403, 407
Massart, D.L., 186, 405
Nordenskiöld, N., 4, 407
Matheis, G., 366, 405
Nuotio, T., 74, 80, 407
Matisto, A., 137, 277, 375, 405
Nurmi, A, 124, 145, 376, 402, 407
Mattson, E, 402
Nurmi, P.A., 6, 51, 407
MaxweU, A.E, 252, 404
Nuutilainen, J., 60, 119, 121, 378, 408
MaxweU, J.A., 186, 402
Nykänen, 0 , 283, 284, 285, 290, 321, 408
McCarthy, J.H., Jr., 122, 150, 406
McCorkeU, R.H., 122, 181, 406
McLaughlin, R.J.M., 325, 406 O'Leary, R.M., 196, 204, 408
Mehlhorn, N.E, 414 Ohlsen, R.A., 393
Meier, A.L., 205, 406, 417 Ohlsson, S.A., 396
Menzies, J., 130, 397 Oivanen, E, 318, 408
Mertie, J.B., Jr., 4, 406 Okko, V, 31, 375, 408
Michelson, D.M., 13, 406 OUier, C, 94, 408
Michie, U.M., 377, 406 OUila, J.T., 270, 408
MictheU, R.L., 325 Osterlund, S.E., 7, 408
Mielikäinen, E, 412 Ottesen, R.T., 394
Miesch, A.T, 237, 406 Otto, H.J., 193, 414
Mikkonen, A.A., 270, 415 Ovenden, L., 66, 367, 408
MiUer, G.H., 391
Miller, J.K., 90, 91, 328, 376, 406 Eahlavanpour, B., 415
Millner, H.B., 4, 410 Ealmu, J-E, 129, 411
424 References index

Paradis, S., 398 Raukas, A., 15, 410


Park, C.J., 400 Reaveley, G.H., 396
Parker, M.E., 397 Repo, R., 127, 130, 141, 410
Parnandi, M., 404 Reyment, R.A., 402
Pearson, M.J., 366, 405 Riddle, C, 211, 410
Pekkarinen, L., 285, 408 Riese, WC, 366, 388, 410
Pelchat, J.C., 400 Robbins, J.C., 122, 410
Peltoniemi, H., 132, 408 Rock, N.M., 240, 410
Pendias, H., 65, 85, 87, 402 Rodionov, D.A., 238, 410
Penttüä, S., 297, 408 Rogers, PJ., 31, 388, 410
Penttüä, VJ., 69, 403 Rogerson, R.J., 13, 410
Perelman, A.I., 5 Rooney, S.T., 391
Persson, B., 394 Roos, S., 285, 410
Perttunen, M., 27, 31, 127, 130, 131, 132, Rose, A.W, 5, 93, 108, 112, 117, 279, 365,
134, 303, 374, 375, 376, 408 374, 385, 387, 395, 410
Perttunen, V, 412 Rosenberg, R.J., 214, 410
Peuraniemi, V, 60, 119, 121, 277, 303, 374, Rossi, S, 296, 409
378, 408 Rouhunkoski, P, 304, 410
Pirttisalo, K, 73, 408 Rundqvist, D.H., 8, 369, 410
Pitkänen, V, 410 Russell, D.W., 398
Pitulko, VM., 5, 408 Russell, G.M, 209, 399
Plant, J.A., 143, 409 Ryan, R.J., 413
Play fair, J., 4 Ryss, Yu.S, 93, 376, 414
Plisov, A.A., 404, 409 Räisänen, M-L., 73, 408
Polikarpochkin, VV, 5, 371, 374, 408 Rosier, H.J., 115, 411
Pollard, J., 233, 409
Pope, T.A., 394 Saarnisto, M., 127, 363, 374, 411
Porritt, J.W.M., 406 Sahama, Th. G., 43, 55, 112, 410
Potts, P.J., 186, 200, 212, 409 Salmi, M, 67, 70, 318, 348, 378, 407, 411
Povondra, P, 195, 414 Salminen, R., 42, 60, 73, 88, 101, 104, 107,
Prest, VK, 31, 141, 409 108, 117, 127, 130, 131, 134, 135, 141,
Proskuriakov, V, 366, 409 144, 157, 285, 290, 322, 370, 373, 374,
Pulkkinen, E., 25, 296, 325, 400, 409 375, 378, 381, 400, 403, 411
Punkari, M., 139, 277, 409 Salo, A., 51, 411
Puranen, M., 326, 409 Salonen, V-P, 128, 129, 131, 132, 134, 135,
Puranen, R, 25, 127, 129, 137, 325, 326, 375, 137, 138, 139, 141, 375, 393, 411
400, 409 Saltikoff, B., 413
Puustinen, K, 373, 409 Sanders, J.E., 44, 46, 48, 52, 53, 54, 397
Pääkkönen, V, 318, 409 Sandström, H., 124, 197, 387, 411
Sanzalone, R.F., 206, 394, 411
Qvarfort, U., 72, 74, 371, 381, 409, 410 Sarapulova, VN., 409
Sauramo, M., 4, 127, 411
Raeburn, C, 4, 410 Savinskii, I.D., 145, 221, 370, 371, 411
Rahkola, P, 270, 410 Schreiner, B.T., 413
Ramanamurthy, M.V, 367, 410 Schroll, E., 186, 411
Rampton, VN., 398 Schuddebeurs, A.P, 40, 375, 411
Rankama, K, 4, 43, 55, 112, 410 Seifert, G., 25, 411
Räsänen, J., 73, 412 Selinus, O., 396
Rastas, P, 412 Sergeev, Ye.A., 4, 5, 412
References index 425

Shen, R., 208, 412 Sturgeon, R.E., 416


Shepherd, A, 374, 388, 412 Sturhahn, H.H., 193, 414
Shetsen, I., 141, 412 Sulcek, Z., 195, 414
Shuts, WW, 5, 17, 21, 24, 31, 37, 44, 117, Sun, H., 417
127, 132, 136, 141, 150, 366, 367, 371, Sundborg, Ä., 46, 414
374, 384, 391, 395, 412 Sundkvist, G., 395
Shimizu, K, 238, 395 Sveshnikov, G.B., 376, 414
Sidorenko, A.V, 5 Szabo, N.L, 130, 414
Siegel, R, 5, 412 Szalay, A., 349, 414
Sillanpää, M., 387, 412 Säynäjärvi, K., 270, 414
Silvennoinen, A., 304, 412 Sibrava, V, 17, 412
Simonson, R.V, 83, 84, 413
Simpson, A., 406 Taipale, K, 127, 363, 374, 411, 413
Sinclair, A.J., 138, 237, 363, 413 Taipale, T., 179
Sinding-Larsen, R., 222, 252, 258, 398, 401, Taka, M, 404
413 Tamminen, P., 379, 414
Sinha A.K., 376, 413 Tanner, V, 136, 414
Sippola, J , 74, 413 Tanskanen, H., 71, 378, 399, 403, 414
Slavin, W, 206, 413 Tares, T, 74, 413
Smee, B.W, 37, 53, 63, 376, 393, 413 Tavela, M., 150, 414
Smith, R.L., 394 Taylor, S.R., 208, 391
Smith, R.T., 69, 343, 397, 409, 413 Tenhola, M., 59, 415
Smith, S.L., 371, 374, 412 Theobald, P.K., Jr., 4, 415
Solovov, A.P, 4, 5, 6, 8, 49, 369, 412, 413 Thomas, R.D., 398
Sondag, F., 74, 405, 413 Thompson, C.E., 4, 415
Sopuck, V, 384, 385, 413 Thompson, F.J., 375, 403
Sorsa, A., 410 Thompson, M., 189, 209, 225, 399, 415
Soveri, J., 55, 413 Thomson, J., 393
Späth, H., 255, 413 Thornton, I., 72, 415
Stanley, C.R., 237, 413 Tilas, D, 4, 415
Stanton, R.E.., 203, 413 Tille, K.G., 394
Starr, M.R., 379, 414 Tontti, M., 413
Stary, J., 206, 413 Toverud, Ö, 60, 63, 352, 374, 415
Stea, R.R., 371, 375, 376, 388, 405, 413, 414 Troup, A.G., 375, 393
Steenfelt, A, 394 Tukey, J.W, 240, 259, 415
Stepanova, M.D., 87, 414 Tukey, P.A., 415
Stevenson, A.G., 409 Tuominen, H.V, 270, 415
Stewart, R.A., 74, 414 Turner, R.G., 414
Stoch, H , 206, 414 Tynni, R., 400
Stone, C.J., 393
Stone, P., 389, 414 Ure, A.M., 205, 415
Story, S.L, 414
Strand, G, 413 Vaasjoki, M., 40, 55, 56, 369, 376, 388, 399,
Strauss, M.R, 172, 414 415
Strelow, RW, 391 Vagners, V.J., 22, 23, 105, 127, 128, 130, 374,
Strong, D.R, 405 375, 396
Ström, K.M., 53, 414 Vallius, H., 178
Stumm, W, 114, 406 Van der Voet, A., 410
426 References index

Van der Walt, T.N., 391 Webb, J.S., 5, 82, 83, 84, 85, 93, 400, 407,
Van Grieken, R.E., 213, 405 410, 416
van Hees, E.H., 74, 414 Webb, PC, 409
Van Loon, J.C., 186, 415 Welinder, S., 416
van Moort, J.C., 398 Welsh, E.P, 203, 205, 416
van Praag, H., 395 Wennervirta, H., 51, 143, 416
Viaene, W, 335, 395 West, R.G., 398
Viets, J.G., 196, 204, 408 Wheatley, M.R., 399
Virkkala, K, 25, 277, 376, 415 White, W.H., 5, 416
Virri, K, 387, 407 Wühelm, E., 5, 55, 279, 338, 339, 366, 376,
Virtanen, K., 75, 416 388, 404, 405, 416
Virtanen, P., 167, 169, 171, 172, 180 Willie, S.N, 205, 416
Vogt, T, 4, 61, 378, 416 Willis, J.P, 213, 417
Voipio, A., 51, 411 Wintz, P, 246, 398
Volden, T, 394 Wu, S.M., 371, 395
Vuorela, I., 416
Vuorinen, A., 72, 76, 80, 416 Xie, X., 332, 369, 371,427
Väänänen, J., 412
Väänänen, P., 56, 416 Yang, M, 399
Yfantis, E.A., 222,417
Waldron, T., 73, 416 Yio, C-L., 319, 404
Walker, J.L., 393 Yletyinen, V, 401
Walsh, J.N., 209, 415, 416 Yliruokanen, I., 346, 407
Walton, S.J., 415
Ward, EN., 203, 416 Zhang, J., 399, 417
Warnant, P., 405 Zheng, K, 369, 371, 417
Warren, H.V, 5, 416 Zhou, Di, 240, 417
Watson, J.S., 409 Zhou, L, 206, 417
427

SUBJECT INDEX

A-horizon, 55, 58, 77, 79, 83, 86-88, 90, 373, alpine, 85
379 alteration zone, 338
AAS, 64, 195, 196, 200, 203, 204, 206-210, amblygonite, 322
216, 264, 264, 270, 276, 278, 293, 298, ammonia, 112
306, 312, 322, 329, 336, 357, 386 ammonium iodide, 205
ablation moraine, 35 amphibole, 43, 279, 280
ablation tiU, 29, 97, 97, 349 amphiboHte, 132, 263, 269, 272, 291, 310, 321
abrasion, 39, 41, 42, 99, 104, 105, 131, 132, amphibolitic, 355
270, 374 anaerobic bacteria, 43
abrasion resistance, 22, 25, 36, 109, 128, 130 analysis of variance, 187, 225, 228, 231-233
absorbance, 202, 205 analytical method, 100, 124, 143, 153, 160,
absorption, 70, 114, 188, 189, 195, 201, 202, 163, 186, 187, 189, 195, 200, 208, 215,
204-207, 224, 228, 312, 334, 385 218, 223, 225, 231, 369, 386
accuracy, 99, 166, 186-188, 212, 224, 369 analytical performance, 185, 187
acid digestion, 199 analytical precision, 187
acid soil, 77 analytical result, 270, 343, 367
acidity, 96, 111, 112, 350, 351; see also pH analyzing crystal, 211, 212
adsorption, 51, 52, 111, 112, 116, 117, 125 Andrews' curve, 249, 250, 258
aeolian, 46, 54; see also loess; dune angle of climb, 137, 141
Ag, 3, 189, 196, 199, 200, 204, 206-210, 214, anglesite, 340
271, 335-339, 343, 385 annual moraine, 20, 33
age determination, 66, 366, 375 annual precipitation, 338
aggressive reagent, 197 anomaly, 5-8, 11, 12, 20, 31, 33, 40, 42, 49,
air-acetylene flame, 204 50-53, 55-58, 60-62, 63, 66-69, 74-77,
Aittojärvi, 69, 273 82, 83, 85, 90, 100, 102-111, 119, 122-
Al, 43, 62, 77, 79, 80, 86, 98, 193, 196, 197, 125, 127, 129, 131, 135, 138, 141, 143-
199, 204, 209, 339, 340, 379 145, 147-150, 152, 153, 156, 160, 185,
alkali, 77, 113 187, 197, 215-218, 220-222, 224, 226-
alkali metal, 196, 204 228, 230, 232-248, 250, 252-254, 256,
alkaline, 77, 113, 115, 195, 198, 387 258-260, 262, 265, 269-271, 275-277,
alkaline earth element, 113 280-284, 286-288, 290, 293, 295, 297,
alkaline soil, 77 298, 301, 302, 306, 308, 310, 316, 319,
Allebuouda, 352-354 320, 323, 32^-329, 332-334, 336, 337,
allo-till, 21, 29 339, 341-346, 348, 349, 353-355, 359,
aluminium 43, 49, 56, 85, 88, 90, 174, 373, 361-364, 366-370, 372-380, 383, 384,
378;, see also Al 386-389
almandite, 280 anomaly pattern, 102, 103, 105-109, 119,
428 Subject index

135, 152, 156, 187, 197, 218, 220, 222, Ba, 43, 98, 193, 196, 198-200, 204, 209, 336,
224, 226, 228, 230, 232, 234, 236, 238, 340
240-242, 244-246, 248, 250, 252, 254, background shift, 208
256, 258, 260, 262, 288, 308, 310, 316 ball mill, 194
anomaly-background contrast, 242, 388 Baltic Sea, 15, 18, 39, 53, 56, 88, 277
antimony, 153, 320, see also Sb barium, 146, 153, 199; see also Ba
apatite, 280 baryte, 335
apex, 40, 352, 355, 362 basal till, 29, 33, 97, 137, 163, 312, 316, 318
apical boulder, 135, 137 base-exchange capacity, 77
aquaregia, 98, 114, 146, 191, 195, 196, 199, base metal, 51, 56, 88, 90, 165, 269, 334, 338,
204, 205, 209, 387 348, 360, 361, 385
arc nebulization, 211 Bayes' decision rule, 256
Ar, 44 Be, 200, 207, 325
Archaean basement, 284, 285, 290 beach sand, 20, 53, 109, 376
arctic region, 11 beaded esker, 38
Arctic Circle, 263, 269, 325, 348, 352, 355 Beer's law, 202, 204
areal average, 367 Bersbo, 74
argillic, 87 beryllium, see Be
argon, see Ar beta decay, 213
arithmetic mean, 135, 264, 278, 279, 298, 302 Bi, 200, 204, 205, 207, 208, 210, 271
arsenic, 115, 153, 336, 337, 385; see also As bias, 186, 224, 227, 232, 233, 248
As, 196, 199, 204, 205, 207-209, 212, 214, bimodal, 236, 241, 280
271, 333, 335, 336, 338, 343 binding capacity, 68
ash layer, 67 binocular microscope, 285
assaying method, 103 biogeochemistry, 9, 116
atlantic, 81, 85, 88 biplots, 253
atmogeochemistry, 9, 365 bismuth, 153; see also Bi
atmosphere, 113, 114, 364 black cotton soil, 87
atomic absorption, 385 black schist, 105, 296, 321, 338, 363
atomic absorption spectrometer, 201 bog, 56, 59, 66-68, 76, 88, 263, 264, 269, 272,
atomic absorption spectrometry (AAS), 195, 293, 310, 321, 325, 332, 338, 339, 343,
202, 203, 205, 319 346-351, 360, 366, 372, 378
atomic emission spectrometry (AES), 200, bogs on slope, 67
203, 206, 210, 212 bones of animals, 59
Au, 187, 189, 196, 200, 205, 206, 211, 214, boreal, 81, 85
269, 271, 302, 305, 306, 308, 310, 318, boric acid, 196, 204
385 boron, 153; see also B
auger bit, 170, 174, 176 bottom sediment, 51, 52, 67, 196, 372
auger driU, 170, 344 boulder, 4, 5, 11, 21-25, 32, 33, 36, 44, 46,
aureole, 6, 93, 94, 96, 124, 156, 364, 371 51, 54, 67, 102, 108, 109, 122, 127-133,
Australia, 56 135-141, 170, 270, 276, 291, 292, 305,
automatic locating satellite system, 167 318, 343, 349, 353, 360, 375
autosampler, 191 boulder circle, 328
boulder fan, 108, 128, 136, 291, 292, 352, 353
B, 207, 209-211, 214 boulder tracing, 4, 63, 122, 127, 139, 141,
B-horizon, 55, 59, 72, 77, 79, 80, 83, 85, 87, 291, 378
88, 90, 91, 117, 264, 332, 333, 372, 379, Box-Cox power transformation, 232
380 Br, 196, 200, 212, 214
Subject index 429

breccia, 277, 328, 332, 334 China, 46, 54, 55, 208, 332
bromium, 206; see also Br chitinozoan, 59
bromoform, 322 chloride, 56, 88, 112, 202, 203
brown soil, 88 chlorine, see Cl
brunined, 88 chromium, 193, 266-268, 312, 314, 327, 368,
brunisol, 85, 328 385; see also Cr
buffer capacity, 80 Cl, 212, 332-334
buffering agent, 202 classification criterion, 254
bulk analysis, 64, 182, 385 classification tree, 257
bullet shape, 25 clast, 23-25, 28, 30, 31, 33, 36, 40, 52, 127,
burosem, 87 130, 139
14
clastic, 8, 53, 57, 66, 75, 96, 100, 102, 123-
C-age determination, 375 125, 127, 141, 247, 280, 320, 344, 359,
C-horizon, 77, 79, 83, 85, 87, 88, 90, 91, 285, 364, 371, 377
329, 352, 372, 379, 380 clastic anomaly, 57, 58, 93, 102, 123, 124, 377
Ca, 4, 43, 56, 193, 196, 197, 199, 204, 209, clastic dispersion, 96, 97, 110-112, 125, 128,
326, 327, 339 145, 326, 332
cadmium, 65; see also Cd
clastic material, 50, 133
calc-silicate mineral, 279
clastic sediment, 372
calcimagnesian soil, 88
clastic transport, 11, 51, 58, 96, 123
calcium, 43, 56, 71, 85, 87, 88, 327; see aho
clay, 3, 15, 20-22, 35-40, 42, 45, 49, 51-53,
Ca
55, 56, 72, 74, 76, 77, 79, 81, 85, 86,
calibration, 188, 189, 205, 224, 227, 230
109, 110, 116-118, 265, 274, 277-280,
Camp Lake, 328-331
335, 336, 346, 348, 373, 376-379, 384,
Canada, 4, 66, 91, 102, 103, 130, 136, 150,
387
205, 328, 330, 331
clay mineral, 43, 76, 77, 88, 90, 96, 117, 264,
Canadian Shield, 15
capillary rise, 11, 55, 60, 63, 77, 82 279, 323, 334, 372, 373, 387
capture, 49, 77, 226 closed tube sampler, 174, 175
carbon dioxide, 43, 44, 94, 112 C/N ratio, 81
carbonaceous material, 114 Co, 69, 83, 114, 117, 193, 196, 199, 200, 204,
carbonation, 82 207, 209, 212, 264, 265, 269, 276, 285,
carrier gase, 123 293, 296, 298, 302, 310, 312, 317, 336,
cassiterite, 40, 53, 109, 196, 385 357, 376, 379, 381, 382
Cd, 74, 193, 199, 200, 204, 207-210, 214, 379 coarse fraction, 23, 27, 35, 40, 165, 279, 322,
central Lapland peneplain, 296 376
cesium, see Cs coating, 24, 51, 117, 373, 383
chalcopyrite, 269, 280, 289, 290, 293, 304 cobalt, 49, 115, 293, 297-301, 302, 373, 376;
characteristic optical wavelength, 203 also see Co
chelate, 59, 79, 377 Cobra percussion drill, 272, 275
chemical alteration, 49, 96, 373 coefficient of variation, 278
chemical anomaly, 8, 55, 123, 123, 364 cohesive soil, 50
chemical dispersion, 96, 111, 112, 121, 123- cold base glacier, 19
125, 145, 148, 150, 152, 181, 326 cold extraction heavy metals (ex HM), 360,
chemical transport, 11, 371 361
chemical weathering, 43, 94-96, 106, 112, colloid, 7, 45, 51, 58, 77, 90, 91, 96, 112, 116,
279, 330 117, 270, 373, 383, 384, 387
chernosem, 11, 79, 83, 86-88 colloidal, 45, 51, 67, 111, 116, 117
cherty quartzite, 263, 267 colorimetric, 90, 91, 202, 203, 216
430 Subject index

colorimetry, 64, 202, 203 crystal lattice, 49, 116


coloured surface map, 218, 245, 259 Cs, 200, 212, 214
comminution, 19, 22, 39, 94, 128-131, 290 Cu, 67, 69, 74, 80, 83, 98, 104-106, 114, 118,
communality, 251, 253 195, 196, 199, 200, 204, 207, 209, 212,
complementary colour, 202 214, 233, 263-265, 267, 269, 271, 276-
complexes, 51, 67, 68, 88, 90, 115, 198 280, 283-290, 293, 296, 298, 302, 305,
composite sample, 76, 150, 152, 162, 222, 369, 306, 308, 310, 312, 317, 326, 327, 329,
380 331-334, 336, 338, 339, 343, 346-348,
compressed-air flush, 172 355, 357, 358, 360, 361, 376, 379, 381,
compressive flow, 19 382, 385, 388
computer processing, 160, 223, 388 cumulative frequency curve, 137, 237-239
concentrate, 7, 10, 59, 67, 90, 347, 348, 352, cutting mill, 194
353, 384 Czechoslovakia, 109
concentration map, 153, 306
conifer, 62, 64, 81, 343, 346 data compression, 257, 258
"consistency" of data, 187 De Geer moraine, 32
contamination, 170, 175, 192-194, 196, 279, dead ice, 33, 35, 38
336, 381 deciduous tree, 64, 81, 85, 346, 360
continental ice, 53, 81, 96, 109, 110, 131, 137, decision limit, 189
291, 305, 308, 366 decomposition, 66, 112, 192, 194-197, 202,
control chart, 188 206, 209, 235, 239
coordinate, 149, 166, 183 deformation, 21, 24, 30, 130
copper, 3, 10, 61, 65, 67, 91, 104, 278, 279, deglaciation, 20, 96, 99, 102, 131, 134, 139,
280, 285, 286, 288, 289, 290, 293, 297, 147, 277
298-302, 305, 307, 309, 327, 329, 330, delayed neutron counting (DNC), 202, 214
331, 347, 348, 355, 357, 358, 359, 360, delta, 39, 51, 52
368, 373, 385; see also Cu delta formation, 51, 110
coprecipitation, 205, 319 deposition, 20, 21, 23, 29, 31-33, 38, 54, 99,
cost, 102, 144, 149, 152, 153, 155, 159, 160, 106, 133, 134, 139, 291, 305
163, 165, 167, 168, 171, 180, 181, 185, Derbyshire, 72
186, 190, 191, 195, 202, 204-207, 213, detailed study, 8, 155, 162, 302, 306
215, 222, 223, 225, 226, 256, 371 detailed study phase, 144, 163
cost-benefit analyses, 190 detection limit, 188, 189, 191, 200, 201, 204-
cost-effectiveness, 195 206, 210-213, 215, 216, 224, 225, 229,
counters, 181 235, 238, 239, 242, 248, 270, 319, 338
counting interval, 213 determination limit, 189
country rock, 7, 8, 60, 100, 118, 119, 312, determination range, 208
316, 389 diabase, 132, 302-304, 306, 310, 338, 360
cover moraine, 139, 277 diamicton, 10, 27
Cr, 83, 98, 193, 194, 196, 199, 200, 204, 207, diamond drilling, 55, 156, 163, 164, 171, 296,
209, 212, 263-265, 267, 269, 276, 285, 302, 317, 370, 388, 389
310, 312, 316, 317, 326, 327, 379, 381, diatom, 59, 70
382, 388 digestion, 195, 196, 205, 206, 336
crawler-based vehicle, 178 dilution, 103, 129, 131, 191, 195, 196, 211,
crevasse filling, 35 223, 235
crushing, 41, 42, 48, 130, 132, 192, 194, 384 dimethylglyoxime, 181, 385
cryoturbation, 328 diorite, 285, 310
cryptocrystalline, 340 diphenyl carbatzime, 385
Subject index 431

direct current plasma atomic emission spec- electrochemical force, 111, 119
trometer (DCP-AES), 206 electrochemical potential, 94, 359, 376
direct reading spectrometer, 206 electron microprobe, 202, 211
direct sample insertion, 211 electrostatic attraction, 117
direction of ice movement, 18, 25, 32, 99-104, electrothermal atomic absorption spectrome-
145, 270, 322, 344 try, 205
discriminant analysis, 155, 160, 291 electrothermal furnace atomizer, 205
dispersal train, 128, 135-137, 141, 375 electrothermal vaporization, 211
dispersion parameter, 132, 135 Elessiinlampi, 284, 286, 287, 290
dispersion type, 123 Elster, 15
dissolution, 51, 55, 69, 76, 82, 124, 198, 219, emission spectra, 207
322, 367, 373, 386 emission spectrometry, 189, 190, 194, 195,
dissolved material, 50, 51 200, 201
distal side, 33, 107, 134, 145, 305, 318 emission wavelength, 207
distance measure, 254, 256, 257 empirical discriminant analysis, 255, 256
dithiol, 203, 272 end moraine, 33-35
dogs, 7, 122 energy dispersive X-ray analyzer (EDX), 386
dot map, 243-245, 259
energy-dispersive X-ray fluorescence spec-
drift, 1, 3, 7, 11, 41, 54, 55, 60, 66, 68, 77, 96,
trometer (ED-XRF), 212
97, 108-110, 113, 123, 130, 131, 153,
englacial, 137
159, 160, 188, 208, 224, 266, 297, 310,
engorged esker, 38
311, 322, 326-328, 333, 334, 336, 337,
enrichment, 90, 91, 98, 117, 204, 214, 323,
346, 363, 372, 375
376, 378, 379
drift formation, 108
epithermal neutron activation analysis
drill, 162, 165, 168-172, 174-176, 179, 269,
(ENAA), 214
297, 305, 306, 310, 318, 339, 340, 342-
epithermal neutron flux, 214
344, 346, 352, 355, 357, 389
epofix, 278
drilling equipment, 160, 171, 173, 178, 179,
182 equilibrium, 94, 95, 113, 114, 116-118, 349
drilling machine, 155, 160, 162, 172, 182 erosion, 21, 31, 35, 38, 43, 45, 46, 54, 58, 85,
dropstone, 109 106, 108, 129, 133, 322, 366, 367, 369
drumlin, 18, 30-32, 134, 137, 139, 293, 318, erratic, 4, 55, 141, 155, 328, 343, 349, 363
328 esker, 20, 21, 34, 35, 37, 38, 74, 110, 270, 328
drumlinoid, 31, 32 esker chain, 38
dry ashing, 196 Estonia, 15
drying, 192-194, 264, 293, 312, 318, 352, 383, Europe, 3, 15, 17, 81, 88, 149
384 evaporates, 43, 55
dune, 41, 46, 54, 55, 62, 83, 110, 376 evaporation, 82, 88, 329, 336
duplicate sample, 166, 230, 244 excavator, 134, 161, 162, 168, 169
dynamic equilibrium, 59, 68 excavator pitting, 291
experimental mean, 187
economic viability, 186 expert system, 219, 247
Eh, 10, 11, 44, 86, 113-116, 125, 182, 346, exploration, 3-7, 9, 12, 21, 31, 32, 37, 39,
351, 373, 376 53, 56, 63, 67, 68, 74, 76, 82, 87, 91,
eigenvalue, 251, 253 110, 122, 124, 125, 127-129, 138, 144,
electrical charge, 117 145, 155, 161, 165, 169, 174, 180, 182,
electrical conductivity, 118 186, 188, 190-192, 194, 197, 202, 203,
electrical potential field, 118 205, 208, 209, 212-217, 221, 240, 242,
electrochemical dispersion, 119, 125 263-272, 274, 276, 278, 280, 282, 284,
432 Subject index

286, 288, 290-292, 294, 296, 298, 300, "fixed channel", 212, 386
302, 304, 306, 308, 310, 312, 314, 316, fixed charge, 117
318, 320, 322, 324, 326, 328, 330, 332, fixed flow cell, 202
334-336, 338, 340-342, 344, 346, 348- fixed surface potential, 117
350, 352, 354-356, 358-362, 364-367, flame atomic absorption, 199
370-373, 375-380, 383-389 flame atomic absorption spectrometry (FAAS),
exploratory data analysis (EDA), 240 204-207, 285
extending flow, 19 flameless atomic absorption spectrometry, 200
extraction agent, 185, 198 flame-mode AAS, 203
floating bog, 66, 67
F, 200, 333 flood deposits, 39, 51
fabric, 24, 28, 30, 275, 305 flow-injection analysis (FIA), 203
factor analysis, 155, 160, 190, 249-251, 255, flow-till, 22, 31
257, 258, 262, 267, 338, 388 fluoride, 202
factor matrix, 252 fluorine, 203; see also F
factor scores, 252 fluorimetry, 385
"false" anomaly, 49, 367 fluorite, 343
Falun, 3, 74 flushing, 168, 170-174
faulting, 24, 37, 38 fluting, 31, 32
Fe, 43, 62, 67, 77, 79, 86, 98, 113, 114, 117, fluvial, 45, 50-52, 99, 137
190, 193, 196-199, 204, 209, 271, 326, flux, 133, 137, 195, 214, 387
327, 336, 339, 340, 342-344, 346, 357, folding, 24, 37
379 fossil secondary pattern, 93
Fe-Mn coating, 117 fracture zones, 122, 156
feldspar, 10, 22, 37, 43, 45, 54, 74, 76, 279, friction soil, 50
355, 376 frost heave, 100, 110, 328
Fennoscandia, 15, 81, 150, 348, 370 frost wedge, 24
Fennoscandian Shield, 15 fulvic acid, 59, 79, 88, 377, 387
fermentation, 328 fulvic salt, 85
ferric iron, 43, 79, 85 fusion, 195, 196, 198, 344, 383, 387
ferri-oxyhydrate, 47 Fuzzy-set theory, 257
ferrous iron, 43
field laboratory, 181 Ga, 200, 206, 207, 210
Filon Rico, 335-337 gabbro, 131, 296, 310, 382
fine fraction, 42, 97, 104, 129, 131, 135, 141, gallium, see Ga
165, 193, 194, 208, 222, 264, 265, 270, galvanic cell, 118
279, 280, 285, 290, 308, 312, 319, 320, gamma emitter, 214
322, 325, 344, 359, 377, 384 gamma ray, 213
fingerprint, 6, 53, 258, 259, 348, 363, 367 gamma spectrometry, 7, 213
Finland, 4, 5, 24, 26, 29, 33, 34, 42, 53, 56, 67, gangue, 338
74, 80, 81, 98, 101, 102, 104-106, 108, gangue minerals, 124, 334
121, 122, 128, 134, 137, 139, 140, 144, garnet, 10, 40, 45, 52, 53, 109, 269, 385
146, 150-152, 157, 198, 272, 277, 279, gas anomaly, 122
280, 283, 290, 293, 296, 302, 310-312, gas flame, 204
318, 321, 346, 347, 355, 356, 358, 359, gas geochemistry, 9, 122, 123
361, 370, 377, 379, 381 gaseous dispersion, 111, 122
fissility, 24, 30 gaseous mercury, 181
fission neutron energy spectrum, 213 gaseous phase, 121, 125, 183
fissure, 94, 122, 348, 352 Gaussian scale, 238
Subject index 433

Ge, 196, 200, 205, 210 glacilacustrine, 20, 36, 39, 40, 377
genesis of till, 21, 29, 97 glacimarine, 20, 36, 39
geochemical cycle, 93 gley, 79, 88
geochemical data, 144, 155, 161, 181, 186, gleysol, 328
187, 190, 219, 223, 224, 233, 237, 238, gliding mean, 160
243, 245, 246, 253, 255, 260 global zoning, 77
geochemical dispersion, 1, 93, 124, 183, 328, Glogfawr, 332-335
343 gneissic granite, 285, 288, 289
geochemical prospecting, 33, 49, 51, 58, 63, goethite, 55, 269, 270, 280, 340
74, 96, 105, 109, 112, 114, 115, 163, 164, gold, 4, 7, 40, 60, 76, 109, 153, 165, 182,
348, 352, 379 193, 204, 210, 222, 269-271, 307-310,
geochemical province, 163, 164, 363, 369, 372 318-320, 376, 384, 385; see also Au
geochemical relief, 187, 218, 220, 221, 232, gold ring, 193
233 goodness-of-fit, 253
Geochemical Atlas of Finland, 144, 146, 151 good laboratory practice (GLP), 215, 227
Geological Survey of Finland, 185, 193, 217, gossan, 49, 55, 56, 216, 338-340, 342, 363,
227, 285 366, 373, 376, 377, 388
germanium, see Ge grain size, 8, 10, 15, 19, 21-23, 27, 30-32, 35,
Germany, 4, 109 36, 40, 42, 45, 46, 52, 54, 74, 78, 105,
glacial drift, 113, 131, 187, 296, 297, 325-327 117, 125, 134, 141, 165, 166, 168, 182,
glacial erosion, 97-99, 108, 147, 160, 297, 193, 194, 197, 222, 264, 278-280, 328,
326, 366 330, 341, 344, 370, 371, 374, 375, 379,
glacial lobe, 99, 141 384
"glacial miü", 375 granite, 4, 7, 42, 43, 131, 132, 136, 151, 263,
glacial overburden, 5, 96, 102, 113, 193, 200 267, 269, 272, 284, 285, 288-291, 310,
glacial till, 1, 8-10, 39, 96, 114, 115, 117-119, 321, 325, 326, 343, 346, 348, 349, 352,
263, 319, 352 355, 372, 382
glacial transport, 23, 57, 58, 96, 100, 102, graphite, 204, 205, 207, 270, 276, 277, 306,
127-132, 134, 135, 137-139, 141, 145, 319, 386
149, 155, 159, 272, 326, 332, 334, 349, graphite furnace atomic absorption (GFAAS),
357, 374 205
glaciated terrain, 2, 5, 21, 41, 58, 96, 99, grass, 59, 66, 81, 263, 269, 328, 332, 343, 352,
108-110, 122, 124, 125, 141, 144, 145, 360
149, 150, 162, 296, 328, 343, 352, 366, gravity, 10, 43, 44, 50, 52, 94, 96, 110, 111,
374, 379 123, 124, 129, 371, 374
glaciation, 13, 15-17, 20, 21, 24, 26, 33, 41, gravity differentiation, 93
42, 46, 54, 55, 81, 86, 96-99, 108, 114, graywacke, 296
115, 123, 128, 145, 297, 310, 326, 328, Great Britain, 72
374, 375 Great Lakes, 15
glacier, 4, 15, 19-21, 28-30, 33-35, 37-39, greenstone, 132, 263, 296, 302, 304
41, 44-46, 54, 79, 83, 96, 97, 99, 100, grinding, 10, 32, 191-194, 383, 384
102-104, 107, 109, 110, 123, 127, 131, groove, 19, 54, 99
133, 134, 136, 145, 148, 156, 163, 279, ground-penetrating radar, 181
322, 343, 374 ground water, 43-45, 51, 52, 55-58, 66-68,
glacifluvial formation, 1, 35, 37, 277, 346 72, 79, 88, 112, 115, 117, 118, 123, 326,
glacifluvial process, 27 328, 332, 340, 349, 351, 352, 370, 373,
glacigenic anomaly, 49, 110, 148, 295, 326 376
glacigenic dispersion, 124, 127, 141, 320 ground-water table, 44, 56, 83, 91, 113, 152,
434 Subject index

166, 379 hydration, 43, 82


Grudie Burn, 343, 345, 346 hydride generation cell, 204, 205
Gulf of Bothnia, 152 hydride generator, 336
Gulf of Finland, 15 hydrocarbon, 121, 122, 150, 181
Günz, 15 hydrofluoric acid, 195, 198, 206, 278, 387
Gutzeit test, 203, 385 hydrogen peroxide, 198, 387
gyttja, 52, 58, 70 hydrogeochemistry, 9, 365
hydrolysate, 49
half-distance method, 127, 137, 138, 141 hydrolysis, 43, 82, 95
half-distance value, 129, 130, 132-135 hydromorphic, 11, 23, 47, 53, 55-58, 63, 66,
hafnium, see Hf 76, 85, 88, 247, 328, 332, 344, 359, 373,
halo, 6, 66, 153, 364, 365 376
Hannukainen, 355-359 hydromorphous, 359
Harjunpää, 274 hygroscopic water, 54, 376, 378
He, 44, 120 Hämeenlinna, 277
health risk, 192
heating cabinet, 193
I, 200, 210, 333, 334
heating identification method, 322
ice-dammed lake, 39
heavy drilling equipment, 161
ice divide, 9, 41, 98, 99, 136, 139, 264, 270,
heavy liquid, 384
326, 349, 356, 374
heavy metal spectrum, 74, 348
ice flow, 19, 99, 100, 104, 106-108, 136, 138,
heavy mineral separation method, 182
heavy mineral study, 156 275, 277, 312, 322
heavy nuclide, 214 ice margin, 99, 107, 109
helium, 44, 181; see also He ice movement, 31, 32, 34, 41, 99, 100, 103,
hematite, 269, 270 105, 108, 133, 149, 322, 323, 353, 360
Hercynian orogeny, 335 ice transport, 149
heterogeneity, 166, 195, 220-222, 225, 228, ice wedge, 24, 366
388 iceberg, 36, 109
Hf, 193, 200, 210, 214 Illinoian, 15
Hg, 122, 196, 200, 270, 319 illite, 77
Hiendelaencina, 335-337 Ilomantsi, 101, 157, 285
Hieronmäki, 277 image processing, 218, 245, 246
highest coast line, 36 immature, 79, 81, 85
histogram, 235-237, 241 impact mill, 194
Hjulström diagram, 45 impeded drainage, 79
Hudson Bay, 15 In, 200, 205, 206, 210
human inpact, see pollution in situ, 41, 49
humates, 60, 61 indicator element, 320
humic acid, 52, 79, 88, 377 indium, see In
humic compounds, 112 inductively coupled plasma-mass spectrome-
humification, 59, 66, 80, 81, 328, 351, 378 ter (ICP-MS), 201, 202, 210, 211, 214,
hummocky, 33, 131, 134, 349 216, 386
hummocky moraine, 134, 137, 138 information capacity, 186, 190, 191, 208
humus, 8, 11, 39, 52, 55, 56, 58-64, 66, 67, insect shards, 59
69, 76, 77, 80, 81, 83, 85, 88, 89, 91, insoluble residue, 199, 209
119, 121, 150, 196, 198, 216, 265, 264, instrumental neutron activation analysis
268, 269, 328, 331, 349, 352, 354-363, (INAA), 198, 201, 210, 213, 216
366, 372, 373, 376-380, 385, 387 integrated computer control, 208
Subject index 435

interface, 19, 25, 40, 133, 270, 309, 318, 345, Kraft paper, 193
381 Kriging interpolation, 243
interference, 190, 202, 204, 205, 207, 208, Kylmäkoski, 291, 292
210, 212, 214, 386
interglacial, 13, 15, 17, 31, 42, 375 La, 199, 207, 209, 212, 214
international quality system standard, 215 laboratory information management system,
interstadial, 15 191
iodine, 332-334; see also I labour-intensive, 192, 195
ion exchange, 90, 111, 116, 125, 210, 377 Lady Loretta, 56
ion probe (SIMS), 202 lake sediment, 1, 8, 9, 72, 73, 110, 215, 328,
ion-selective electrode, 202 329, 331, 332, 363, 366, 372, 381
ion size, 45 laminae, 24, 71
ionic diffusion, 371 laminar movement, 25
ionic dispersion, 98 lamination, 30
ionic potential, 95 land uplift, 15, 38
ionization, 117, 210, 211 landscape, 366
Ir, 211 landslide, 85, 110, 123, 366
iridium, see Ir lanthanides, 386; see also La; rare earth ele-
Ireland, 89-91, 371 ments
iron, 3, 10, 43, 44, 49, 51, 53, 54, 55, 56, Lapland, 71, 98, 136, 263, 266-269, 271, 296-
67, 70, 79, 85, 86, 88, 90, 115, 198, 264, 298, 325, 350, 351, 353, 354, 360, 368
289, 326, 327, 338, 339, 348, 355, 356, laser ablation, 208, 211
359, 373, 376, 379; see Fe laser probe, 202
iron cap, 49, 373 lateral diffusion, 67
iron pan, 47, 56, 85, 373 lateral moraine, 33
Iron Age, 73 laterite, 388
irradiation, 213, 214 Laukunkangas, 293, 295
lead, 3, 49, 55-58, 65, 67, 71-73, 76, 83, 90,
jarosite, 340 91, 122, 329-331, 334, 336, 337, 339,
Jatulian, 302, 304 340, 342, 347, 348, 368, 369, 378, 387,
jaw crusher, 42, 194 388; see also Pb
leaves of trees, 59
K, 43, 56, 95, 158, 193, 196, 198, 199, 204, lee, 4, 30, 32, 37, 54, 129, 374
209, 339, 340 lee-side cone, 31, 32
Kajaani, 272 lepidocrocite, 280
Kalahari Desert, 109 lepidolite, 322, 323
käme, 38, 39, 328 Li, 193, 196, 199, 204, 209, 320-325
käme delta, 39 lichen, 59, 76, 269, 328
käme moraine, 33 Lietsonsuo, 346, 347
käme plateau, 38 lime, 45, 85, 131, 132, 138, 378
käme terrace, 38 limit of detection, 186, 188, 189, 212
Kansan, 15 limonite, 43, 45, 47, 49, 53, 55, 265, 270, 335,
kaolinite, 77, 79 373, 379
Karelian schist belt, 321 linear discriminant analysis, 255
key element, 153 lithium, 195, 322; see also Li
key element paragenese, 153 lithogeochemistry, 5, 9, 222, 365
Kissanloso, 283, 284, 289, 290 lithological composition, 22, 102
knob and kettle, 33 lithology, 6, 10, 19, 21, 23, 25, 30, 31, 40,
Kokka, 103 45, 71, 87, 88, 94, 127-132, 135-137,
436 Subject index

141, 150, 155-158, 160, 194, 265, 269, methane, 44


283, 285, 288, 289, 312, 363, 365, 368, methyl isobutyl ketone (MIBK), 204-206
370-372, 374-377 Metsola, 303, 304, 306
Utter, 81, 89, 328, 377 Mg, 43, 56, 95, 193, 196, 199, 204, 209, 310,
living organism, 58, 59 312, 316, 326, 327, 339
local phase, 156, 159, 160 mica, 10, 22, 37, 40, 42, 43, 54, 76, 274, 276,
lodgement, 19, 21, 24, 25, 30-33, 35, 137, 280, 291, 310, 322, 323, 360, 372, 373,
139, 263, 285, 304, 343, 360 376
loess, 41, 46, 54-56, 376, 377 mica schist, 104, 131, 199, 263, 277, 321-323,
lognormal distribution, 237 338, 343, 346, 381, 382
Losomäki, 103, 105 microdensitometer, 207
Lutsokuru, 263, 266-268, 310 microwave oven, 196
luvisol, 88 Mindel, 15
mineral acid, 7, 124, 197, 204, 387
Maaselkä, 296, 297, 301, 302, 372 mineral anomaly, 109
Madrid, 335
mineral material, 21, 24, 27, 30, 35, 51, 58,
magmatic crystallization, 93
67, 77, 79, 82, 85, 94, 109, 328, 332,
magnesium, 313, 327; see also Mg
344, 351, 357, 371, 380
magnetic anomaly, 274-277, 326
mineralization, 6-8, 11, 49, 55-57, 60, 64, 75,
magnetic susceptibility, 25, 181, 325, 327
82, 100, 102, 103, 121, 123, 127, 131,
magnetite, 10, 40, 45, 52, 53, 109, 181, 269,
143, 144, 149, 152, 153, 216, 217, 265,
277, 304, 326, 328, 349, 357, 359, 369,
267, 272-274, 277, 280, 281, 291, 293,
385
295, 296, 302, 306, 308, 310, 311, 316,
Mahalanobis' distance, 257-259
318, 320, 328, 329, 331-334, 338, 339,
main element, 160
341, 343, 344, 346, 348, 352, 354, 355,
major element, 160, 377
363-365, 367, 371, 373, 376, 383, 384,
manganese, 51, 86, 88, 115, 116, 198, 327,
349, 351, 376, 387; see aho Mn 387-389
manual processing, 160 mineralogical study, 355, 364
marine mud, 88 minerogenic, 143, 168
masking agent, 202 minor element, 160, 369, 377
mass movement, 30, 31, 37, 110 Miocene, 335
mass spectra, 210 mixing, 9, 19, 37, 130, 134, 140, 173, 195,
Masugnsbyn, 348-351, 377 197, 257, 329, 367, 374, 384
matrix modification, 190, 205 Mn, 43, 67, 77, 79, 86, 87, 98, 114, 117, 193,
mature, 59, 66, 79, 89, 90, 130 196, 199, 204, 209, 276, 326, 327, 336,
mean temperature, 367 338, 339, 342-344, 346, 351, 379, 381,
mechanical disintegration, 42 382, 388
mechanical dispersion, 185 Mo, 67, 69, 87, 189, 193, 196, 199, 200, 204,
mechanical transport, 11, 50, 96 207-210, 212, 264, 265, 267, 269, 272,
median, 135, 137, 141, 155, 237, 243, 245, 273, 285, 287, 288, 290, 343, 344, 346,
278, 290, 302 357
melt, 50, 195 mobility of element, 114, 373
melt out till, 22 mobility of trace element, 116
melt water, 97 mode of occurrence, 6, 11, 52, 123, 320, 337,
mercury, 181; see also Hg 340, 344, 367, 373, 374, 383-385, 388
metal ratio, 97, 102, 161 mode of transport, 19, 20, 53, 130, 365
metamict material, 39 moisture, 10, 58, 88, 152, 166, 372
metamorphism, 10, 93 molecular absorption band, 202
Subject index 437

molybdenite, 263, 289, 343, 344, 352 noble gases, 120


molybdenum, 69, 272, 273, 288, 343-346, non-destructive technique, 198
348, 354; see also Mo non-singular, 252
monochromator, 204, 207 nonparametric method, 232, 233, 240
montebrasite, 322, 323 Norway, 4, 51, 143, 150
montmorillonite, 49, 77, 79 Nordkalott project, 8, 144, 150, 258, 305
monzonite, 355, 359 nuclear reactor, 200, 213
mor, 85
morphology, 166, 322 O, see oxygen
mortar, 194 olivine-tholeiite series, 293
motor sledge, 178, 179 optical emission spectrometry (OES), 206-
moulin käme, 38 208, 216, 298, 306, 339, 386
mountainous terrain, 110 optical resolution, 202
moving average, 155, 243, 280, 368 optimal grid, 220-222
mud, 24, 39, 52, 53, 56, 58, 66, 67, 81, 328, ore element, 6, 52, 153, 183, 364
332, 344, 348 ore outcrop, 104-106, 127, 164, 291, 357, 359
"multichannel" instrument, 191 organic acid, 7, 11, 80
multimict lithology, 40 organic compound, 59, 83, 115, 119, 339, 376,
multimodal, 22, 237 379
multiple transport, 8, 107, 136, 372 organic deposit, 1, 41, 58, 74, 76, 372, 374
multistage transport, 139 organic material, 1, 11, 45, 51-53, 58, 59, 67,
multivariate statistical method, 163 68, 79, 80, 87, 90, 114, 117, 195, 214,
mussel shell, 59 269, 335, 349, 351, 375, 377, 378, 383,
Mäkärärova, 269-272 387
orientation, 19, 25, 30, 31, 33, 50, 100, 136,
N, 44, 81, 204 243, 272, 291-293, 304, 312, 318, 320,
Na, 43, 56, 95, 193, 196, 199, 204, 209, 339 361
Nb, 200, 207, 210, 212, 325 Oriselkä, 321-323, 325
Nebrascan, 15 ortho-till, 21, 22
nebulized solution, 208 orthogonal VARIMAX criterion, 252
nebulizer, 204, 209 Ostrobothnia, 279
nebulizing, 205, 210 Outokumpu, 73, 102-106, 259
neocuproine, 385 overbank sediment, 39, 45, 51
Neolithic, 73 overburden, 1-9, 11, 15, 19, 35, 40-44, 49, 50,
neutron activation, 201, 202, 212 55-60, 77-79, 81, 83, 86, 88, 91, 93, 110,
neutron activation analysis (NAA), 195, 200, 112, 119, 120, 122, 124, 127, 143-145,
213, 214, 386 147-150, 152, 155, 156, 160-166, 168,
neutron emitter, 214 171, 172, 176, 181-183, 185, 198, 203,
Ni, 60, 64, 80, 83, 103-105, 114, 117, 158, 204, 221, 222, 263, 264, 270, 271, 277,
193, 196, 199, 200, 204, 207, 209, 212, 285, 293, 295, 304, 318, 328-330, 332,
263-265, 267, 269, 276, 291, 293, 310, 333, 335, 337-339, 342-344, 346, 348,
312, 316, 317, 336, 346-348, 357, 360, 351, 352, 355-357, 359, 360, 363-366,
361, 376, 379, 381, 382, 385 368, 371, 372, 374, 376, 377, 379-381,
nickel, 42, 60, 62, 64, 67, 74, 91, 104, 115, 383, 388, 389
158, 181, 210, 266-268, 293, 315, 316, oxidate, 49, 55
326, 347, 348, 359, 368, 385; see also Ni oxidation, 43, 59, 65, 79, 82, 118, 121, 339,
niobium, see Nb 373, 376, 383
nitric acid, 195, 312, 387 oxidation-reduction potential, 96; see also Eh
nitrogen, see N oxidation-reduction condition, 106
438 Subject index

oxidation state, 113 pH, 10, 42-44, 51, 80, 86, 95, 113-117, 125,
oxide, 7, 10, 43, 49, 51, 79, 95, 113, 114, 116, 182, 202, 333, 346, 349, 351, 360, 373,
117, 185, 190, 193, 197, 198, 210, 269, 376, 379
279, 289, 338, 384, 387 phosphates, 196, 340
oxidizing acid, 195, 196 phosphorus, 153; see also P
oxidizing agent, 113, 196 photographic film, 181
oxy ions, 115 photographic recording, 206
oxygen, 11, 44, 58, 72, 94, 112, 113 photomultiplier, 204, 206, 208
oxyhydrates, 47, 338, 372 phyUite, 131, 321, 348
physical methods, 181
ί 193, 199, 207, 209, 340, 388 physical weathering, 94, 95
palladium, see Pd pilot study, 225, 364
Pallastunturi, 310, 311 placer deposit, 40, 53, 76, 109
panning, 4, 7, 63, 353, 385 plagioclase, 280
paragenetic element, 7, 246, 368 plant pollen, 59
partial extraction, 195, 197, 199, 203, 216 plasma emission, 188, 189, 191, 195, 199
partial leach, 6, 215 plasma-mass spectrometry (ICP-MS), 190,
particle induced X-ray emission (PIXE), 202 200, 201, 208
passive ice, 102 plasma source, 206
pathfinder element, 153, 185, 189, 271 plasma source spectrometry, 195
Pb, 57, 74, 80, 114, 189, 193, 196, 199, 200,
platinum, 195, 385, 386; see also Pt
204, 205, 207-336, 338-340, 342, 343,
platinum crucible, 195
357, 360, 361, 378, 379, 381, 385, 388
platinum group of elements (PGE), 189, 200,
Pd, 205,206,211
209, 210, 214
Pearson coefficient of correlation, 247, 377
Pleistocene, 9, 13, 16, 46, 81, 86, 128
Pearson product moment correlation coeffi-
plumbogummite, 340
cient, 247
pneumatic drill, 297
peat, 8, 58, 59, 61, 65-71, 76, 81, 88, 89, 93,
podsol, 11, 55, 79, 80, 82, 83, 85-88, 117, 264,
196, 263-265, 269, 272, 274, 293, 304,
332, 335, 338, 339, 343-352, 357, 360, 343, 378
363, 366, 372, 376-378, 380 polarizing microscope, 278, 322
peat geochemistry, 9 pollution, 49, 55, 67, 72-74, 76, 90, 364, 371,
pebble, 51, 54, 100, 117, 128, 130, 131, 141, 388
157, 285, 290 polychromator, 207-209
pedalfer, 85 polyethylene bottel, 196
pedocal, 85 pore water, 50, 373, 376, 383
pedogeochemistry, 4, 9, 31, 365 Pori, 274
pedology, 82 porosity, 55, 77, 87, 122
pentlandite, 263, 293, 317 portable Hg determinator, 205
perchloric acid, 329 portable XRF analyzer, 182
percussion drill, 160, 170, 171, 264, 270, 272, Porte aux Moines, 338, 340, 342, 372
275, 278, 285, 293, 297, 305, 306, 312, postglacial, 9, 15, 18, 21, 39, 41, 53, 56, 58,
318, 322, 360 99, 100, 150, 152, 304, 326, 366
perennial, 110 postglacial clay, 38
periglacial, 24, 110 potassium, 49, 198, 373; see also K
permafrost, 9, 24, 67, 79, 328, 329, 331, 332, powder, 195, 196, 199, 208, 211, 212, 214,
366 264, 386
permeability, 52, 55, 76, 81, 86 Precambrian, 335, 343, 352
Petalax, 359-361, 372 precipitate, 43, 44, 55, 95, 113, 117, 123, 319
Subject index 439

precipitation process, 55 Re, 200


precision, 186-188, 190, 204, 211, 212, 223, reconnaissance scale, 144
224, 236, 256 redox potential, 44, 95, 111-113, 351; see also
preconcentrate, 209 Eh
preconcentration, 189, 191, 201, 206, 209, reducing, 47, 55, 56, 59, 88, 113, 114, 167,
210 191, 192, 257, 319, 339
preglacial, 9, 22, 25, 31, 37, 41, 42, 49, 98, reduction, 82, 194, 205, 240, 376
296, 297, 318, 320, 326, 374, 381 redusates, 55, 339
presentation of data, 217, 223 refractory mineral, 196, 201
pressed pellet, 211, 212 regional phase, 148, 153, 155
pretreatment, 185, 190-193, 195, 210, 211, regolith, 6, 9-11, 41, 85, 112, 263, 270, 356,
215, 232, 383 365, 366, 372-374
primary dispersion pattern, 93 regosol, 328
primary halo, 93 relative variation, 229
primary mineral, 95, 111, 124, 197, 383 reliability, 186-188, 369
primary mineralogy, 185 rendsina, 11, 88
primary pattern, 93 repeatability, 186, 386
principal component analysis, 155 replicate analysis, 225, 232
probability plot, 236, 237, 258 replicate sampling, 225, 231, 232
prompt gamma-ray NAA (PGNAA), 214 representativeness, 162, 165, 166, 169, 192,
prospector, 3, 52, 72, 114, 124, 145, 234, 235, 193, 212, 213, 221-223, 225, 239, 380
365, 368 reproducibility, 186, 208, 224-226, 228
proximal, 32, 130, 134, 137 residual anomaly, 155, 160, 368
proximal side, 107, 134 residual soil, 155, 338, 379, 380
pseudomorphs, 270 residuum, 1, 11, 373
Pt, 206, 211 resistates, 149, 373
pyrite, 104, 269, 270, 280, 304, 343 reverse circulation drilling, 172
pyrrhotite, 263, 280, 293, 317 Rh, 206, 211
rhenium, see Re
quality control, 188, 215, 217, 221, 224, 225, rhodium, see Rh
231, 232, 248, 252, 371 ribbed moraine, 32, 33
quartz, 18, 22, 37, 42, 49, 53, 54, 74, 85, 194, riffle sizer, 194
279, 280, 285, 335, 343, 355, 373, 376 Riss, 15
quartzite, 104, 131, 132, 138, 263, 269, 296, Rn, see radon
302-304, 306, 310, 338, 348, 355 road cutting, 99
robust estimate, 240
radiation detector, 122 roches moutonnees, 15, 18, 19
radioactive decay, 120 rock geochemistry, 5, 365, 389
radiochemical, 214 rod, 170, 172-174, 275, 389
radiometric survey, 363 Rogen moraine, 32, 131, 291, 304
radon, 44, 122, 181, 349 roUing, 44, 96, 110, 328, 332, 355, 374
raised bogs, 66, 67 Rovaniemi, 296, 302
Rana, 335-337 rubidium, see Rb
random error, 186, 241
randomization, 226, 227 Saale, 15
rapakivi, 109, 131, 150, 151, 198, 199, 372 Sahara, 54, 109
rare earth elements (REE), 200, 201, 209, saline soil, 88
210, 214 Salpausselkä, 33, 34, 134, 277
Rb, 196, 200, 204, 207, 210, 212 salt solution, 7, 124
440 Subject index

sample contamination, 172 sediment, 1, 4, 8-11, 15, 20-22, 24, 30, 31,
sample decomposition, 191 35-42, 45, 46, 48, 50-53, 55, 56, 58, 67,
sample logistics, 185, 190, 215 72-74, 76, 79, 87, 97, 98, 109, 110, 114,
sample size, 165, 222 116, 127, 133, 150, 168, 169, 193, 196,
sampler, 38, 162, 170, 174-176, 264, 270, 272, 215, 270, 278, 280, 285, 304, 310, 312,
275-278, 285, 360, 383 320, 328, 329, 331, 332, 335, 341, 363,
sampling density, 143-145,148-152, 155,158, 365, 366, 372, 374-377, 382
163, 164, 166, 221, 222, 297-300, 302, sedimentation, 10, 20, 23, 30, 37-39, 42, 45,
305, 307, 369 46, 55, 93, 94, 96, 97, 133, 137, 367, 384
sampling depth, 143, 149, 159, 163, 165, 166, Seinäjoki, 318, 319
168-170, 277, 306, 378, 381 seismic sounding, 152, 181
sampling equipment, 143, 165, 168, 169, 180- selective extraction, 197, 215, 340
182, 192 selectivity, 186, 188, 190, 197
sampling error, 187, 222 selenium, see Se
sampling grid, 102, 145, 147-149, 152, 159, self potential, 356
161, 165, 219, 220, 222, 234, 280, 302, semiquantitative analysis, 207
354, 370, 371 semiquantitative method, 160, 385
sampling interval, 156, 160, 293, 312, 336 sensitivity, 7, 100, 103, 186, 188, 369, 384,
sampling method, 165, 166, 168, 218, 370 386
sampling variance, 187 separation, 10, 32, 37, 39, 42, 54, 95, 189,
sand, 3, 20-22, 24, 25, 28, 30, 33, 35-42, 45, 190, 201, 203-206, 209, 210, 214, 231,
49, 51-55, 60, 62, 74, 76, 81, 86, 93, 353, 374, 384, 385
110, 117, 124, 131, 197, 270, 274, 275, sequential analysis, 364, 373, 376
277-280, 293, 318, 335, 336, 356-359, sequential extraction, 197
361, 373, 376, 377, 379, 384, 387 sequential XRF, 212
sandstone, 131, 132, 343 serpentinite, 103-105, 263, 265, 267, 269,
sandur, 39, 83 310, 316, 317
saturation, 96, 114, 326, 364 sesquioxides, 83, 88
Sb, 83, 193, 196, 200, 204, 205, 207-210, 212, shifting, 136
214, 271, 318, 319, 336, 338 Si, 43, 79, 193, 196, 204, 209, 211
Sc, 199, 200, 207, 209, 214 Sierra Leone, 370
Scandinavia, 4, 5, 102, 103, 128, 144, 160 sieve shaker, 191
scandium, see Sc sieving, 23, 192-194, 265, 276, 352, 383, 384
scanning electron microscopy (SEM), 386 silica, 43, 116, 373, 378; see also Si
scatter diagram, 228-230, 233, 243, 244, 247, silicate, 7, 36, 42, 49, 95, 185, 195, 196, 198,
249, 252 199, 376, 387
scavenge, 77, 96, 117, 373 silicate-forming element, 204
scheelite, 63, 277-281, 352, 355, 385 silt, 20-22, 30, 33, 35-41, 45, 52, 53, 56, 74,
Scotland, 143 76, 80, 109, 110, 130, 277-280, 332, 348,
Scottish Highlands, 345, 346 360, 376, 377, 385
Se, 196, 200, 205, 206 Silurian, 332, 335
secondary dispersion, 77, 78, 93, 94, 96, 339, silver, 3, 335, 337; see aho Ag
341 silver halogenide, 335
secondary halo, 93 simultaneous spectrometer, 208, 213
secondary mineral, 43, 95, 162, 216, 372, 378, simultaneous XRF, 212
383 single element anomaly, 234
secondary radiation, 211 skarn, 104, 263, 277, 352, 355
secondary till, 29, 30 slip, 19, 110, 175,227
Subject index 441

slumping, 38, 44, 45, 50 statistical classification method, 253


Sn, 187, 190, 193, 200, 205, 207, 210, 212, statistical dependence, 246, 247, 250
277-279, 325 statistical method, 74, 76, 153, 155, 161, 217-
snowmobile, 170 219, 223, 224, 231, 233, 240, 262, 338,
sodium, 88, 198, 205; see also Na 367
soil, 1, 3, 4, 7, 11, 24, 27, 43-46, 49-51, 55, stems of coral, 59
57, 58, 64, 65, 72, 74, 77-88, 90, 91, 114, step-by-step procedure, 144, 146, 163, 164
117, 119, 120, 123, 155, 156, 161, 165, stone, 4, 23-25, 28, 30, 31, 44, 50, 54, 100,
166, 168-174, 176, 180-182, 198, 263, 107, 109, 264, 276, 277, 291, 305, 310,
275, 276, 320, 321, 328-340, 343, 346, 312, 332, 375, 389
352, 356, 364-367, 370-375, 378-380, stone count, 100, 129, 130, 156, 291, 318,
387, 388 359, 375
soil creep, 96, 110, 123, 335 stratigraphy, 139, 149, 152, 155, 166, 169,
soil-forming process, 77 181, 310, 311, 363, 371, 380, 388
soil gas, 44 stream sediment, 1, 8, 73, 117, 143, 144,
soil geochemistry, 1, 3, 9, 363, 374 149, 150, 155, 197, 220, 221, 283, 284,
soil horizon, 77, 82, 379 296-298, 302, 316, 336, 339, 344, 370,
solar energy, 43 380, 381, 387
solid mode, 195 stream sediment geochemistry, 9, 143
solid phase, 112 striae, 4, 15, 18, 19, 32, 99, 270, 272, 275,
solifluction, 21, 50, 100, 110, 328 291, 293, 312, 318, 349, 352
solubility, 96, 111, 113-116, 125, 199, 351, striation, 25, 100
384 strontium, see Sr
soluble, 47, 49, 55, 76, 77, 86, 95, 113, 114, structure, 1, 5, 21, 23, 24, 28, 30, 31, 38, 39,
146, 191, 199, 205, 209, 333, 360, 373, 45, 51, 66, 85, 87, 156, 222, 246, 249,
378, 383 250, 253, 254, 269, 270, 291, 310, 337,
solution cell, 202 363
solution mode, 185, 195, 213 subalpine, 85
solvent extraction, 204 subaquatic, 30
sorted sediment, 10, 11, 15, 24, 31, 35-37, subglacial, 21, 30, 31, 38
40, 41, 42, 74, 270, 278, 280, 285, 310, sublimation, 20, 21, 30
312, 320, 335, 374-376, 381, 382 suboutcrop, 4, 40, 42, 57, 64, 93, 100, 102-
source area, 100, 103, 104, 110, 129-131, 133, 104, 122, 144, 145, 148, 160, 161, 269,
145, 152, 153, 163, 327 272, 333, 353-355, 357, 359-362
Spain, 335-337 subsample, 152, 165, 166, 192, 195, 231, 232,
specific gravity, 109 370, 380
spectrometer, 181, 191, 192, 201, 206-212, sulphate, 56, 88, 121, 340, 373
326, 386 sulphide, 7, 10, 22, 43, 49, 56, 95, 104, 109,
spectrophotometry, 195, 201, 202 113, 114, 119-122, 152, 153, 185, 196-
sphalerite, 263, 280, 334 199, 210, 263, 265, 269, 277, 279, 280,
spodic, 85 289, 306, 326, 328, 331, 339, 349, 360,
spodumene, 322, 323, 325 372, 373, 376, 379, 381, 384, 388
squeeze-up esker, 38 sulphide-forming element, 204
Sr, 196, 199, 200, 204, 209, 211, 212 sulphuric acid, 112-114
stability of excitation, 208 supervised method, 253
standard deviation, 40, 128, 132, 138, 187, suppa, 2, 35, 38
189, 232, 249, 251, 264, 265, 278, 298, supracrustal rock, 289
302, 367, 388 supraglacial, 19, 21, 30, 33
442 Subject index

surface area, 42, 51, 102, 104, 106, 113, 192 Tohmajärvi, 320-322
surface erosion, 110 toluene, 206
surface water, 56, 112, 113, 123, 339, 364 topography, 42, 58, 99, 107, 130, 141, 143,
survey strategy, 165, 166, 168 152, 166, 178, 246, 277, 293, 318, 346,
susceptibility, 181, 357, 385 366
suspension, 44, 384 total analysis, 124, 291
Svecokarelian orogenic belt, 277 total decomposition, 197, 204
Sweden, 3, 72, 150, 152, 348, 352, 370 tourmaline, 280
swing mill, 194 trace element, 6, 9, 12, 71, 72, 74, 114, 116,
symbol map, 242, 243 153, 160, 195, 196, 198, 199, 200, 201,
212, 215, 368, 377, 386, 387
Ta, 193, 200, 209, 210, 212, 214, 325 training data, 253-256
talus cone, 41, 50, 110 transition element, 113
Tampere, 290 translocation, 21
tantalum, 386; see also Ta transport, 6, 10, 19, 21, 25, 27, 29, 32, 33,
Te, 196, 200, 205, 206, 210 35-37, 39, 42-44, 46, 47, 50, 51, 54,
tectonics, 94, 363 56-58, 63, 69, 94-96, 99, 100, 102, 103,
tellurium, see Te 106-111, 123, 124, 128, 130, 131, 133,
tephra, 46 135, 137-180, 270, 272, 320, 326, 328,
terminal grade, 22, 128, 130 329, 332, 337, 338, 344-346, 356, 359,
terminal moraine, 33 367, 369, 370, 372, 374, 375, 384, 388
Th, 199, 200, 209, 210, 212, 214, 325, 351 transport direction, 11, 99, 100, 108, 110,
thallium, see Tl 123, 145, 148, 297, 305, 308, 328, 346,
thermal, 133, 134, 211, 213, 214 348, 371, 380
thermal ionization mass spectrometry (TIMS), transport distance, 100-102, 105, 106, 109,
211 123, 127-132, 134, 135, 137-141, 145,
thorium, see Th 148, 277, 291, 308, 312, 323, 359
throughput, 186, 191, 192, 196, 201-203, 206, transport distance (TDD) method, 128, 137,
209, 213, 215 138, 141
Ti, 43, 193, 196, 199, 204, 388 transported drift, 373
till, 5, 10, 11, 15, 19-25, 27-33, 35-42, 49, 53, transport mechanism, 1, 109, 123
55, 58, 62, 68, 69, 72, 74-76, 80, 81, 85, trimming, 240
90, 91, 93, 96-110, 113-115, 117-119, triphylite, 322, 323
127-132, 134-141, 143, 144, 146, 148- "true value", 242
153, 155, 157-163, 165, 166, 168, 169, tungsten, 7, 63, 193, 278, 279, 280, 352, 353,
171-173, 188, 190, 193, 194, 197, 198, 354, 355, 384, 386; see also W
204, 206, 208, 215, 216, 220, 222, 227, turnaround, 186, 192
228, 244, 253, 257, 264-267, 269-280, two-peaked "rabbit ear" anomaly, 119
283, 285-299, 301, 302, 304-306, 308- two-vector dispersion, 371
310, 312, 313, 316, 318-326, 328, 329, twofold unbalanced design, 225, 228
332, 335, 343-346, 348, 349, 352-363,
366, 370, 372, 374-378, 380-382, 384, U, 67, 98, 200, 202, 207-212, 214, 346, 348,
385 349, 351, 377, 385
till blanket, 100, 101, 117, 131, 152, 285, 287, ultramafite, 310, 316, 317
322 ultrasonic sieving, 194
time-dependent shift, 188 ultraviolet (UV), 7, 63, 352, 385
tin, 3, 203, 277-280, 282, 354; see also Sn unconsolidated, 1, 11, 41, 77, 98, 155
titanium, 67, 348; see also Ti United Kingdom, 149
Tl, 200, 205, 210, 211 unsupervised method, 253
Subject index 443

uranium, 51, 56, 58, 122, 137, 202, 203, 347, weathering, 1, 7, 10, 11, 19, 27, 31, 41, 43,
348, 349, 350, 351, 352, 377, 385, 386; 49, 58, 77, 79-83, 85, 93-99, 104-106,
see also U 110-113, 123, 160, 216, 280, 290, 318,
UV radiation, 202 323, 326, 329, 332, 337, 338, 365, 367,
371-373, 375, 378
y 196, 199, 200, 207, 209, 212, 381, 382, 388 weathering crust, 42, 296, 297, 326
vacuum path, 209 Weichsel, 15
valence, 113 weighted median, 245
vanadium, 115, 348; see also V wet ashing, 196
vaporization, 205 white spirit, 194
Visasaari, 325-327 winsorized average, 240
void, 44, 49, 55, 81, 373, 376 winsorizing, 240
volatile, 123, 205, 384 Wisconsin, 15
volcanite, 277, 279, 301, 306, 318 wolframite, 63, 352, 355, 385
volcano, 46, 348, 367 Wurm, 15
Vuonelonoja, 325-327
X-ray fluorescence, 187, 190, 194, 201, 211,
W, 187, 193, 200, 207-210, 212, 277-280, 352, 352, 386
385 X-ray fluorescence spectrometry (XRF), 188,
Wales, 333, 334 190, 195, 198, 200, 201, 211-213, 264,
warm base glacier, 19 276, 291, 312, 347
washboard moraine, 32 Y, 199, 200, 207, 209, 210, 212
water, 1, 3, 4, 8, 10, 11, 15, 20, 21, 23, 27, Yb, 207, 209
30-32, 35-46, 49-53, 55, 56, 58, 65-68, ytterbium, see Yb
72, 73, 77, 79, 81, 82, 85, 88, 91, 94-97, yttrium, 151; see also Y
99, 108, 110-114, 117, 123, 124, 156,
168, 170-173, 182, 194, 195, 198, 199, Zambia, 370
202, 205, 216, 222, 265, 296, 328, 329, zinc, 61, 64, 65, 75, 101, 122, 267, 278, 279,
331-378, 383-385, 387 280, 329, 330, 331, 334, 337-340, 342,
water table, 79, 88, 113, 114, 117, 161 347, 348, 376, 381, 387; see also Zn
wave action, 27, 35, 36, 52, 109 zirconium, see Zr
weak acid, 124 Zn, 51, 64, 69, 74, 80, 83, 114, 117, 118, 193,
weak agent, 124 196, 199, 204, 207, 209, 211, 212, 244,
weathered bedrock, 1, 93, 96-98, 104, 113, 264, 265, 267, 271, 276-280, 310, 312,
116, 117, 122, 143, 160, 162, 216, 264- 317, 329, 331-333, 335, 336, 338, 339,
267, 269-273, 275, 285, 296-298, 300- 342, 343, 357, 360, 361, 378, 379, 381,
302, 306, 312, 318-320, 322-326, 336, 382, 385, 388
339, 373, 374, 383 Zr, 193, 196, 200, 209, 212

Potrebbero piacerti anche