Download as pdf or txt
Download as pdf or txt
You are on page 1of 255

HANDBOOK OF EXPLORATION GEOCHEMISTRY

G.J.S. GOVETT (Editor)

1. ANALYTICAL METHODS IN GEOCHEMICAL PROSPECTING


2. STATISTICS AND DATA ANALYSIS IN GEOCHEMICAL PROSPECTING
3. ROCK GEOCHEMISTRY IN MINERAL EXPLORATION
4. DRAINAGE GEOCHEMISTRY IN MINERAL EXPLORATION
5. SOIL GEOCHEMISTRY IN MINERAL EXPLORATION
6. BIOGEOCHEMISTRY AND GEOBOTANY IN MINERAL EXPLORATION
7. VOLATILE ELEMENTS IN MINERAL EXPLORATION
Handbook of Exploration Geochemistry

VOLUME I
Analytical Methods in
Geochemical Prospecting

by

W.K.FLETCHER
Associate Professor
Department of Geological Sciences
University of British Columbia, Vancouver, B.C., Canada

ELSEVIER SCIENTIFIC PUBLISHING COMPANY


Amsterdam — Oxford — New York 1981
ELSEVIER SCIENTIFIC PUBLISHING COMPANY
1, Molenwerf, 1014 AG Amsterdam
P.O. Box 211,1000 AE Amsterdam, The Netherlands

Distributors for the United States and Canada:

ELSEVIER/NORTH-HOLLAND INC.
52, Vanderbilt Avenue
New York, N.Y. 10017

Library of Congress Cataloging in Publication Data


F l e t c h e r , William K
A n a l y t i c a l methods i n geochemical prospecting.
(Handbook of e x p l o r a t i o n geochemistry ; v. 1)
Bibliography: p.
Includes index.
1. Geochemical prospecting. I . Title.
TT S e r i e s
TN270.F55 622'.13 8θ-398θ6
ISBN 0-khk-kl930-6 (v. l )

ISBN 0-444-41930-6 (Vol. 1)


ISBN 0-444-41932-2 (Series)

© Elsevier Scientific Publishing Company, 1981


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 Scientific Publishing
Company, P.O. Box 330, 1000 AH Amsterdam, The Netherlands

Printed in The Netherlands


EDITOR'S FOREWORD

During the past 25 years exploration geochemistry has developed from an


esoteric subject pursued by a few academics and a novelty practised by a few
companies to a standard mineral exploration technique. In the 1950s region-
al scale stream sediment surveys were first being tested, atomic absorption
spectrophotometry barely existed as an analytical technique, and the height
of sophistication in data interpretation was to calculate threshold as the
mean plus two standard deviations of a presumed background population.
Today most exploration programmes use soil and drainage geochemical
surveys as a matter of routine, and rock and volatile element surveys are
increasingly becoming standard techniques. Multi-element analyses and com-
puter data processing of geochemical exploration data are commonplace.
In the course of various assignments around the world (often in places
remote from comprehensive libraries) I became aware of the difficulties of
following and bringing together developments in exploration geochemistry —
widely scattered in the literature as they are — in a sufficiently detailed form
for use by the average company or government geologist. Even in places
where detailed information is available, the geologist (who may have had
little formal training in exploration geochemistry) has difficulty in deter-
mining what information is relevant to a particular exploration problem. It
is the geologist, not the professional exploration geochemist, who is still
responsible for most of the applications of exploration geochemistry today.
The scope of exploration geochemistry is now too large to encompass
in a single text which can do no more than present a broad survey of the sub-
ject. The Handbook of Exploration Geochemistry, of which this is the first
volume, was conceived as a solution to the problem of the geologist who
needs to use or interpret exploration geochemical results. It is also designed
to serve the professional exploration geochemist working in the field who
does not have access to the large number of separate papers required to solve
a particular problem, as well as a reference text for those engaged in explora-
tion geochemistry research.
The Handbook will be published over the next few years in a series of
individual volumes. In addition to this book the following volumes are being
prepared: Statistics and Data Analysis in Geochemical Prospecting, Rock
VI

Geochemistry in Mineral Exploration, Drainage Geochemistry in Mineral


Exploration, Soil Geochemistry in Mineral Exploration, Biogeochemistry
and Geobotany in Mineral Exploration, and Volatile Elements in Mineral
Exploration. Each of these volumes will have different authors (or editors in
the case of multi-author books), and the approaches to the subject will
necessarily be different.
An attempt has been made to conform to a common set of objectives.
These are:
(1) The theme of the series is ore-finding; the broad test for inclusion of
material is that it should be relevant to the discovery of mineral deposits.
(2) The information is presented in sufficient detail and in a form to be
immediately understood — and applied — by non-specialist practising
geologists.
(3) Notwithstanding to the "how-to" practical exploration geochemical
approach to ore-finding, the dedication of a volume to each of the
various aspects of exploration geochemistry allows adequate space for
a philosophical examination of geochemical problems and applications
and the presentation of the theoretical bases of exploration geochemistry.
It is appropriate that the first volume in this series is Analytical Methods
in Geochemical Prospecting by W.K. Fletcher. The rapid growth in explora-
tion geochemistry alluded to above — apart from the pressure from mining
companies for increased exploration capacity — is due to two main factors:
the development of rapid, cheap multi-element analytical techniques; and
the simultaneous development and increasing availability of computer tech-
nology that allows the vast increase in analytical data to be handled
efficiently.
Analytical techniques are fundamental to the practice of exploration geo-
chemistry. Outside of the U.S.S.R. the use of geochemistry as a significant
mineral exploration method dates from the development of simple, rapid
and sensitive colorimetric techniques in the 1950s. The availability of atomic
absorption spectrophotometry in the following decade placed an incompara-
ble tool in the hands of the exploration geochemist — the range of elements
that can be determined is greatly increased, the instrumentation is relatively
simple and inexpensive, its sensitivity is excellent, and productivity is high.
Most geologists practising exploration geochemistry never have to per-
form an actual analysis; however, many are called upon to recommend ana-
lytical methods for their samples and some have to supervise the installation
of laboratories. It is therefore vital that they have an understanding of the
principles, capabilities, and constraints of the various analytical methods.
This volume provides information on a range of techniques from simple
colorimetric tests to the advanced inductively coupled plasma emission
spectrometry. It should enable the geologist to communicate intelligently
with the chemical analyst to obtain the maximum information from geo-
chemical samples. The geochemical implications of various sample diges-
VII

tion techniques and the effect they have on precision and accuracy (as well
as the precision and accuracy of the actual analytical method) are of funda-
mental importance to the correct interpretation of analytical data. These
topics are emphasized and discussed in detail.
In conformity with the general objective of the Handbook series, Dr.
Fletcher also provides the theoretical background to the analytical tech-
niques he describes. This volume is therefore also a comprehensive reference
text on analytical methods in exploration geochemistry for the professional
geochemist and university student.
As the editor of the Handbook I am particularly pleased with this, the
first volume. An important facet of the volume — and one that distinguishes
it from other treatises on analytical methods — is the liberal illustration of
the analytical techniques by results from actual exploration geochemical
surveys. The constant reference to a large number of such surveys and the
careful interpretation of the results is an excellent example of the approach
which the Handbook hopes to achieve in all of its volumes.

G.J.S. GOVETT
Sydney, N.S.W., Australia
December, 1980.
PREFACE

Books on geochemical analysis are usually general texts emphasizing aquisi-


tion of data for mineralogical and petrological purposes, specialized texts
dealing with specific methods of analysis, or manuals providing step-by-step
instructions. This book fits none of these catagories: it is intended as an
introduction to the choice of analytical methods that might be available to a
geologist submitting samples to a typical prospecting laboratory, the advan-
tages and disadvantages of those methods and the problems that can arise in
their use. Quality control and the importance of sample decomposition are
emphasized, and only sufficient theoretical background has been included as
is necessary to provide the non-specialist with a sound basis for discussion of
each method's scope and limitations.
It is hoped that this book will bridge the communication gap that is some-
times observed between the geologist in the field and the analyst in the
laboratory. To this end it is directed principally to mineral exploration
geologists who must organize geochemical surveys and submit samples for
analysis, and geochemists,with perhaps little or no formal analytical training,
who find themselves involved in analysis or responsible for a laboratory.
Students of geology and those in other disciplines, who determine trace ele-
ments in earth materials, may also find the book useful. For those interested
in specific details about particular methods of analysis, an attempt has been
made to cite the more significant original papers from the abundant geo-
chemical and analytical literature.
Many friends and colleagues have assisted the author but special mention
should be made of Gerry Govett who first suggested the book. Criticism of
the manuscript and helpful suggestions were forthcoming from D. Brabec, J.
Davidson, I. Elliott, and G. Holmes, S. Horsky, A.J. Sinclair and SJVI. Flet-
cher. J. Homenuk, I. Mclntyre and B. Robillard were responsible for typing
and most of the figures were drafted by G. Hodge. To each of these and
many others, the author is indebted.

W.K. FLETCHER
Vancouver, B.C.
Chapter 1

INTRODUCTION

In their Introduction to Geochemistry in Mineral Exploration, Hawkes


and Webb (1962), page 1) defined geochemical prospecting as "any method
of mineral exploration based on systematic measurement of one or more
chemical properties of a naturally occurring material". Commonly, this
involves the determination of trace metal abundances in systematically
collected samples, many (or most) of which may contain only normal (back-
ground) concentrations. Results must then be interpreted to provide thresh-
old values which, taking into account the origins of the materials and factors
influencing their composition, most effectively separate background con-
centrations from those to be considered anomalous and hence worthy of
further investigation.
Once the threshold value has been obtained the magnitude of the anomaly,
or anomaly contrast, can be expressed as the ratio of peak to threshold con-
centrations. The greater the anomaly contrast, the smaller the chance of
missing significant geochemical patterns: sampling, analytical and interpre-
tive techniques should therefore be chosen, on the basis of either previous
experience or an orientation survey, to optimize contrast. Because of the
importance of anomaly contrast, exploration geochemists generally attach
greater importance to relative, rather than absolute, metal abundances. This,
as we shall see, has had a considerable influence on the approach taken to
analysis and on the choice of analytical methods.
A great diversity of natural materials can and have been utilized for geo-
chemical prospecting: these include rock samples, drainage sediments, soils
and overburden, surface and ground waters, mineral separates, plants, atmo-
spheric particulates, and soil and atmospheric gases. Even trout have been
shown to give a geochemical response (Warren et al., 1971). Collection and
analysis of many of these media requires specialized equipment, considera-
tion of which is beyond the scope of this text. The vast majority of explora-
tion analyses, however, involve rocks, soils and sediments (Table 1-1). Con-
centrations of trace elements in these media can range from their typical
crustal abundances to ore grades so that there is considerable variability in
both chemical and physical characteristics of the samples (Fig. 1-1, Table 1-
II). Suitable analytical methods must therefore combine adequate sensitivity
2

TABLE 1-1
Census of geochemical samples collected in North America (from A.E.G. Analysis Com-
mittee, 1971, unless otherwise indicated)

Material sampled Percentage of total samples (1970-1971)

Canada United States

Soils 56.7 19.0


Rocks 14.6 44.0
Stream sediments 23.4 23.0
Vegetation 1.9 4.3
Water 2.7 8.3
Air 0.2 0.9
Other 0.4 0.5
Total number of samples
June 1, 1970 to May 30, 1971: 812,456 337,370
l 2
1975: 1,100,000 374,020
1
Boyle (1976).
2
Canney and Post (1977).

(Figs. 1-9, 1-10 and 1-11) with freedom from interferences over a wide range
of compositional variability. The need to handle large numbers of samples
(100—1000 per day), with short turnaround times and at minimum cost,
imposes further constraints on choice of methods and also requires careful
organization of laboratory operations.

10 000

z 1000
o

ioo H
cc
LU
Q_

CC

Fig. 1-1. Abundance of some trace elements in soils. Unshaded sections indicate more un-
usual values; abnormally high values found in proximity to ore deposits have been ignored.
(From Mitchell, (1964a, in: F.E. Bear (Editor), Chemistry of the Soil, 2nd ed. © 1964 by
Litton Educational Publishing, Inc. Reprinted by permission of Van Nostrand Reinhold.)
TABLE l-II
Chemical composition (%) of some igneous and sedimentary rocks (based on Ahrens, 1965, unless otherwise indicated)
3
Con- Ultrabasic Basalts Inter- Granites Shales Sand- Lime- Laterite
! 2
stituent rocks mediate stones stones
rocks

Si0 2 43.5 48.5 54.5 69.1 58.1 93.2 5.19 17.02


Ti02 0.8 1.8 1.5 0.5 0.7 0.03 0.06 1.96
A1 2 0 3 2.0 15.5 16.4 14.5 15.4 1.28 0.81 32.93
Fe203 2.5 2.8 3.3 1.7 4.0) 0.54
0.43 46.51
FeO 9.9 8.1 5.2 2.2 2.5 j
MnO 0.2 0.17 0.15 0.07 — — 0.05 0.75
MgO 37.0 8.6 3.8 1.1 2.4 0.07 7.90 0.24
CaO 3.0 10.7 6.5 2.6 3.1 3.12 42.61 0.34
Na20 0.4 2.3 4.2 3.9 1.3 0.05 0.06
0.39
K20 0.1 0.7 3.2 3.8 3.2 0.33 0.18

Orthoquartzitic sandstone (Pettijohn, 1957).


Composite of 345 limestones (Pettijohn, 1957).
Nichol and Henderson-Hamilton (1965): total Fe as F e 2 0 3 .
4

Progress in exploration geochemistry has been closely linked to develop-


ments in analytical chemistry; the rapid and almost universal replacement of
colorimetric methods by atomic absorption spectrophotometry, within only
ten years of Walsh (1955) describing a simple atomic absorption spectro-
photometer is a good example. This freed the geochemist from considering
only elements of immediate interest by making it possible to provide quan-
titative data on several additional elements at little or no extra cost. However,
the significance of analysis in geochemical prospecting goes beyond the mere
provision of abundant systematic data; it must be regarded as a part of the
interpretive process insofar as choice of appropriate analytical methods can
provide a filter against much of the normal compositional variability of
natural materials. Ideally this would be achieved with analytical methods
that only determined those trace constituents related to or derived from
mineralization. Although this ideal is not attainable, it is possible to achieve
considerable selectivity for particular components of a sample. Used to

HCIO4CU \ *

i
N
- Vx V"*" x

UMW-.\\)^
^ w </·» H I >S-M

HCl0 4Cu EDTA Cu


ppm ppm
X BACKGROUND 0-70 0-10
THRESHOLD 70 10
rd
0 3 ORDER ANOMALOUS 71-140 11-20
nd
Θ 2 ORDER ANOMALOUS 140 - 280 21-40
st
• I ORDER ANOMALOUS > 280 > 40
SOIL ANOMALY WITH KNOWN MINERALIZATION
-CHAL€OPYRITE AND PYRITE (PORPHYRY TYPE)

Fig. 1-2. Perchloric acid- and EDTA-extractable Cu in stream sediments, Fiji. Both extrac-
tions outline the area of anomalous soils over known mineralization. However, with per-
chloric acid the limit of the third-order anomaly, defined by a distinct break at 70 ppm in
the histogram and indicated on the map by a dashed line, extends upstream of the soil
anomaly and is much more extensive than the EDTA-extractable Cu anomaly. This broad
anomaly is apparently geologically controlled and related to the distribution of un-
mineralized intrusive rocks with slightly raised Cu contents. (From Bradshaw et al., 1974.)
5

advantage such partial, rather than total, analyses of the sample can enhance
geochemical patterns related to mineralization (Fig. 1-2) and also enable the
geochemical dispersion of an element to be studied. Conversely, inappropri-
ate analytical methods can mask geochemical patterns of interest.
The benefits of close liaison between the field and laboratory should be
obvious, ideally the analyst-geochemist should be one person. However, as
noted in a 1968 discussion of "What is a Geochemical Analysis?" (Hansuld
et al., 1969), communication between geologist and analyst is often lacking
to the overall detriment of many geochemical programmes. It is hoped that
this text will provide field-orientated geologists and geochemists with useful
insights into the methods available for analysis of their samples, the princi-
ples behind those methods, and their relative merits and limitations. No
specific analytical methods are described because many excellent publications
already fulfill that purpose. Convenient compilations of analytical methods
used in geochemical prospecting are given by Stanton (1966, 1976), Ward
(1975) and Ward et al. (1963, 1969).

ANALYSIS OF EXPLORATION SAMPLES: CHOICE OF METHODS

Analysis of most natural materials involves three major stages — sample


preparation, digestion or extraction, and the final determination (Fig. 1-3).
For soils or sediments the first stage would typically include drying, dis-
aggregation, and sieving. This might be followed by strong or partial extrac-
tions to release trace elements into solution for their final determination by
colorimetry, atomic absorption spectrophotometry (AAS) or perhaps by
inductively coupled plasma emission spectroscopy (ICP-ES). Alternatively,
the sieved sample might be analyzed directly by either X-ray fluorescence
(XRF) or DC-arc emission spectroscopy (DC-ES). Because of the considera-
ble variation to be found in the aims of geochemical prospecting programmes
and the problems encountered, no single combination of sample preparation,
decomposition and determinitive methods can be universally applicable.
Instead, the geochemist and analyst must select the most appropriate meth-
ods, from the many available, for their particular problem. Although many
geochemical, analytical and administrative considerations can influence the
choice of an analytical method, sample preparation and decomposition are
the principal analytical factors influencing anomaly contrast.

Geochemical considerations

Geochemical considerations arise from the behaviour of the trace elements


during their dispersion and the development of geochemical anomalies.
Dispersion processes are considered in detail by Rose et al. (1979) and
Levinson (1974): very briefly geochemical anomalies in soils or sediments
SAMPLE DISSOLUTION
MATERIAL PREPARATION OF SAMPLE ANALYTICAL METHOD

Atomic absorption
Plasma emission
Colorimetry
υ
O

DC-arc emission
XRF

Atomic absorption
Plasma emission
Colorimetry

DC-arc emission
XRF

"4
Atomic absorption
Dry at 110°C Acid Attack
Plasma emission
Colorimetry
DC-arc emission
XRF

Plasma emission
Atomic absorption

Colorimetry

Precipitation

Fig. 1-3. Some of the pathways for preparation, dissolution and analysis of exploration
samples.

can result from either mechanical (clastic) transport of metal-rich particles


or by hydromorphic (saline) dispersion of dissolved ions. As would be
expected relatively more mobile elements (Fig. 1-4), released from unstable
minerals during weathering, commonly form hydromorphic anomalies
whereas mechanical dispersion predominates for those elements that have
limited solubilities or are present in resistate minerals. Clastic and saline
anomalies are often associated with a single source but, because of their
different origins, can become separated on the geochemical landscape
(Fig. 1-5). Where this separation exists the exploration target can shift from
one anomaly to another as exploration proceeds. For example, in explora-
tion for lead-zinc ores relatively long down drainage dispersion patterns of
Zn, as the more mobile element, might provide the ideal reconnaissance
target whereas distribution of Pb in soils would probably be a better guide
for follow-up studies.
7

Relative pH of Environment
Mobility Acidic Alkaline

Very Mobile
Θ {ä}
S, Ca, Mg, Na

Ag, As, Cd, Co | Mo, Se, U,V 1


Cu, Ni, Zn

Low | Si, P, K | | Si, P, K |


Mobility | Fe, Mn | | S, Ca, Mg,Na |
I Mo, Se, U, V

Immobile AI, Ti, Sn, W, Nb ΑΙ,ΤΊ, Sn, W, Nb


Ta, Cr, Zr, Th Ta,Cr, Zr,Th
Ag, As, Cd, Co
Cu, Ni, Zn
Fe, Mn

Fig. 1-4. Relative mobilities of selected elements.

Taking into account the dispersion characteristics of the element, sample


preparation and analytical methods should be chosen to enhance anomaly
contrast and reduce the likelihood of false anomalies being pursued. The op-
tions available include use of various size fractions, mineral separates, or one
or more of the many decomposition techniques described in Chapter 4.
Choice of methods may be based either on previous experience or, if suitable
mineral occurrences exist, on an orientation survey. With respect to labora-
tory procedures, an orientation survey should at least establish the optimum
size fraction and whether a total analysis, strong decomposition or partial
extraction will provide the best anomaly contrast.

E
Ω. Immobile element Mobile element (hydromorphic)
a anomaly anomaly

fr—7-^^--Soii

Bedrock J&r ^^^^I3^^bf>^^ Bog - Seepage


zo
'^e**^^^^^^^^ , ne

W
Fig. 1-5. Idealized separation of mobile and immobile elements in the geochemical land-
scape.
/
1 ^V
LEGEND
Niobium ppm

BACKGROUND VALUES
Formation Threshold Samples

Tokio 7 11
Trinity :
Siltstone 3.5 8
Limestone 3.5 9

Fig. 1-6. Dispersion of Nb in (A) the minus 10/plus 20-, and (B) minus 80- and minus 300-mesh fractions of stream sediments asso-
ciated with kimberlite pipes, Arkansas. (From Gregory and Tooms, 1969.)
9

Routinely, analysis of minus 80-mesh (177 μηι) material has become a


generally accepted procedure. Sometimes, however, better contrast and more
extensive anomalies can be obtained with other size fractions (Fig. 1-6).
Intuitively it might be anticipated that the coarser size fractions would give
the best contrast for elements present in resistate minerals, especially near
their source, or in rugged terrain with rapid mechanical erosion. Conversely,
the finer fractions would be expected to play a more important role in devel-
opment of hydromorphic anomalies by adsorption. In either case the objec-
tive of separating a particular size fraction is to reject any diluting material
not involved in the dispersion of metal-rich material from the source of the
anomaly. From a practical viewpoint, fractions coarser than 80 mesh may
require grinding prior to analysis whereas fractions finer than 270 mesh
(50 μηι) can only be separated by relatively time consuming wet sieving or
sedimentation procedures. Mineral separates can improve anomaly contrast
(Fig. 1-7) but their preparation is very time-consuming compared to analysis
of the whole sample.
Decomposition of the sample provides a bewildering array of procedural
choices: these vary from dissolutions with either strong acids or by fusion to
release most of the sample's trace element content, to extraction and
leaching with comparatively mild reagents. The latter can be used either to
liberate weakly adsorbed metals in a non-selective fashion or to selectively
release metals associated with ion-exchange sites, sulphides, organic matter,
or amorphous and crystalline hydroxides of Fe and Mn. Thus, although the
decomposition stage can introduce several additional steps into the analysis,
it provides the geochemist with very powerful methods of investigating geo-
chemical dispersion processes, determining the origins of anomalies and
enhancing anomaly contrast (Figs. 1-2 and 1-8). This is particularly valuable
in unravelling the behaviour of hydromorphologically transported elements in
soils and sediments, or in distinguishing metals associated with different
mineral phases, especially sulphides, in bedrock.
A considerable proportion (depending on mineralogy) of the total trace
element content of a sample may be extracted by a strong acid decomposi-
tion; such determinations, however, cannot be regarded as total analyses
compared to results from XRF or DC-ES. This should not necessarily be
regarded as a disadvantage because, as noted by Ward et al. (1969, p. 8) in
relation to their use of boiling 7 M nitric acid, ". . . the significant trace
amounts of metal in exploration have quite likely been introduced into the
rocks by hydrothermal or other genetic processes, and such metals are easily
solubilized by boiling nitric acid. Background amounts of metals such as
copper and zinc in crystal lattices of silicates are less significant in explora-
tion, and the need to solubilize them is not as important in exploration as in
abundance and distribution studies." Whatever decompositon procedure is
used the final solution should be compatible with the determination of as
many elements as possible.
10

Fig. 1-7. Distribution of Sn in (A) minus 10/plus 35-mesh till and (B) the minus 10/plus
35-mesh heavy mineral fraction of the till, Mount Pleasant, New Brunswick. (From Szabo
et al., 1975.)

Widespread use of partial extractions in exploration geochemistry has much


in common with analysis of agricultural and environmental samples to esti-
mate plant-available or biologically active metal concentrations in soils and
sediments. Much useful information is to be found in the literature of these
disciplines.

Analytical and organizational considerations

Assuming that the choice of an analytical method has not been severely
curtailed by geochemical considerations, the principal analytical factors that
determine if a particular method is worthy of further consideration are its
sensitivity, freedom from interferences, and its reliability and reproducibil-
ity. Final selection of a method might then be based on organizational and
administrative factors such as the numbers of samples to be analyzed and the
number of elements to be determined; location of the laboratory; capital
and operating costs; availability of supplies; and the skills of the staff.

Analytical sensitivity
The detection limits attainable with AAS, DC-ES and XRF are considered
in relation to average crustal abundances of the elements in Figs. 1-9, 1-10
and 1-11. None of the methods provides adequate sensitivity for all elements
11

Fig. 1-8. Extraction of Cu from stream sediments with (A) cold 6 M hydrochloric acid,
and (B) hot 4 M nitric acid, Coppermine Basalt Belt, Northwest Territories. The cold
extraction gives the best contrast. (From Allan and Hornbrook, 1971.)

that might reasonably be expected to be of interest in a large exploration


laboratory and some elements, for example F, must be determined by other
methods. Furthermore, although perhaps AAS comes closest to the ideal if
flameless determination of Hg and hydride generation of As, Sb, Se and Te
are included, it is notably inadequate for the estimation of Bi, Nb, U and W,
and only marginally adequate for Ag and perhaps Pb and Mo. Although not
necessarily required for exploration purposes, reliable estimation of these
elements at their background concentrations requires their preliminary con-
centration by solvent extraction or ion exchange. The time involved in pre-
concentration must then be balanced against the merits of employing some
other, more sensitive, analytical method. For some elements with very low
crustal abundances, for example Au, preconcentration will almost invariably
be needed whichever of the three methods of analysis is employed. This is
also true for the determination of almost all trace elements in natural waters.
Unless special precautions are taken DC-ES generally gives poorer detec-
tion limits for volatile elements, such as As, Hg, Se, Zn and Cd, than atomic
12

KEY
0 - can be determined by flame emission
hydride generation
H n - nitrous oxide - acetylene flaine He

p Be
s
*
- solvent extraction
- cold vapour
B C N 0 F Ne

11Mg mh
ίΑΠ ::,?-.U
P S Cl Ar

11 Sc
1 m\n
jjCrJ s K sH s B Ga Ge am Br Kr

pRbj m Y Zr Nb ijMo; Tc Ru Rh ""1


PdIs
::Cd| In m i l jij 1 Xe

Cs mi
• Sr;:;
La Hf Ta Si Re Os Ir Pt •:Au: Tl W£ Bi Po At Rn
•:Ba:; w
Fr Ra Ac

Ce Pr Nd Pm Sm Eu Gd Tb Dy Ho Er Tm Yb Lu

Th Pa U Np Pu Am Cm Bk Cf Es Fm Md Lw

Fig. 1-9. Analysis of exploration samples by AAS. Elements most suitable for determina-
tion by AAS shown in stipple; bold face letters indicate that the concentration of the ele-
ment can normally be estimated without difficulty after sample decomposition with
strong acids. Small letters indicate that special operating conditions are required or
recommended as shown in the key.

absorption but is superior for some refractory elements (Nb and W) and also
for Bi, Sb and Sn. Comparable and possibly better detection limits than
those attainable with AAS appear to be possible with ICP-ES, although
published data on analyses of exploration samples by this method are still
limited. Detection limits with XRF are generally inferior to those with AAS
or ES. Nevertheless, XRF is capable of determining concentrations at least as
low as 10 ppm for most elements heavier than Ca: it is particularly well
suited to the rapid determination of major elements and the more abundant
trace elements, and to the analysis of heavy mineral separates.
For special determinations the analyst may have to resort to analytical
techniques of exceptional sensitivity. These include neutron activation,
mass spectroscopy and graphite furnace atomic absorption. However, with
the exception of delayed neutron counting in the determination of U
(Garrett and Lynch, 1976), these techniques are normally too time con-
suming and expensive for routine analysis.
13

H He

01ill 111 C N 0 F Ne

Na 11 ÖH Z'pi s Cl Ar

IIIli II W\\Els Mn K Co s ;cu jjzrii JGa Ge iAs:·: Se Br Kr

Rb
11 IYI Wtä ΖΓ| gNbi; Tc Ru Rh Pd
3 Cd In | s n | £sP Te 1 Xe

Cs
ill pjLaJlJ Hf Ta nyy* Re Os Ir Pt Au Hg Tl fpbi; 110 Po At Rn

Fr Ra Ac

Ce Pr Nd Pm Sm Eu Gd Tb Dy Ho Er Tm Yb Lu

Th Pa U Np Pu Am Cm Bk Cf Es Fm Md Lw

Fig. 1-10. Analysis of exploration samples by DC-ES. Emission lines of elements in bold
face and stipple are almost invariably present in the DC-arc spectra of geochemical
samples: elements in stipple only are often visible and those in half-stipple are occa-
sionally visible. Remaining elements either cannot be determined by ES or are only very
rarely seen in emission spectra of rocks, soils and sediments.

H He

Li Be B C N 0 F Ne

Na m ΪΜί il igiPj::: &§S; :::;Cl::: Ar

pi ;Ca i| 81| y | B Mn H iicol B B SI Ga EiGej SAsj: Se jIBrj Kr

Rb Ir IYI !2rimy Mo Tc Ru Rh Pd Ag Cd In i;Sn;: Sb Te 1 Xe

bcsi Ba La: Hf Ta isws Re Os Ir Pt Au Tl !JPb! Bi Po At Rn

Fr Ra Ac

be SPrS iiiNdi Pm •Sm! Eu Gd Tb Dy Ho Er Tm Yb Lu

iThi: Pa s-us Np Pu Am Cm Bk Cf Es Fm Md Lw

Fig. 1-11. Analysis of exploration samples by XRF. Concentrations of elements in bold


face and stipple can be estimated by XRF in most samples. Concentrations of elements in
stipple only are usually close to or below their detection limits, they will only be mea-
surable in samples with above average contents.
14

Interferences
Materials collected in geochemical prospecting programmes vary consider-
ably and often unpredictably in their trace element and bulk chemical com-
position. Moreover, their bulk composition is seldom either known or deter-
mined. Consequently, it is usually impossible to match the composition of
standards and samples. This gives rise to a variety of matrix interferences
which often tend to reflect or follow changes in bedrock geology (and are
therefore not always a disadvantage). In addition both ES and X R F are
especially subject to spectral interferences caused by coincident or over-
lapping emission lines.
No analytical method is free of interferences. Nevertheless, compared to
DC-ES and X R F , AAS has considerable advantages in this respect as a result
of the comparative simplicity of absorption compared to emission spectra. In
addition, with AAS the few matrix interferences of sufficient severity to be a
problem in analysis of exploration samples, for example the suppression of
Mo by Ca, are relatively easily avoided by simple modifications to the com-
position of the sample solution. Correction of interferences therefore
involves no additional calculations or data manipulation.
Modification of the sample matrix is also used to reduce interferences in
DC-ES by mixing the sample powder with graphite and spectroscopic buffers
— usually salts of the alkali or alkali earth elements. Even with this precau-
tion interferences, attributable to matrix variability, could still amount t o as
much as ±30 to ±100% for samples of unusual composition unless further
corrections are applied. This is not practical in rapid analysis when the emis-
sion spectra are to be recorded photographically and their intensities esti-
mated visually: such methods are therefore best regarded as semi-quantita-
tive. If a multi-channel direct-reading spectrometer is used for the analysis,
interferences can be corrected by use of internal standards and by simulta-
neous estimation of the bulk composition of the sample. Access to a com-
puter is, however, necessary to calculate the corrections for matrix and spec-
tral interferences with reasonable efficiency.
Compared to flames and DC-arcs, the ICP has several advantageous
characteristics which make its results less susceptible to matrix effects
although spectral interferences are still present, and background emission
and stray light in the spectrometer can be a problem. Because of the
enormous analytical capacity of the ICP-ES, data handling and interference
corrections require at least a small dedicated computer.
In XRF analysis there is a relatively simple relationship between the bulk
composition of a sample, expressed as its mass absorption coefficient, and
the intensity of the characteristic X-rays it emits. Consequently, to correct
for variations in bulk chemical composition, it is necessary to obtain the
relative mass absorption coefficients of samples and standards. A simple,
elegant method of doing this with sufficient reliability for most exploration
analyses has been described by Feather and Willis (1976). If more than one
15

or two elements are to be determined manual calculation of the corrections


becomes laborious.

Reliability and reproducibility


Reliability of an analytical procedure involves two related factors: first, its
overall performance in day-to-day analysis, and secondly, the actual estima-
tion of analytical reproducibility as part of a quality control programme.
The overall performance of a method is related to its robustness or ability to
withstand departures from the optimum conditions. These departures may
reflect changes in local conditions, for example in temperature or reagent
purity, or result from the inexperience of the analyst; some methods (the
determination of Au with Brilliant Green is an example) are notoriously
difficult or require especial care and skill. Although most (but not all)
modern analytical instruments are reasonably robust and designed for
operator convenience and safety, the author's experience is that downtime
can be expected to increase with their increasing sophistication.
Despite the general reliability of the analytical methods and instruments
used in exploration laboratories, it is extremely important to measure any
short- or long-term fluctuations in analytical performance. This reflects the
special significance of relative metal contents and anomaly contrast in
exploration geochemistry. Quality control is therefore considered in detail
in Chapter 2. An example of the spurious geochemical trends that can arise
from systematic analytical errors is shown in Fig. 2-11: typically these fol-
low the sequence of sample analysis. Although random errors introduced
into the data during analysis can be an important component of its total
variability, errors arise at all stages of sample collection, preparation and
analysis. There is therefore little merit in improving the reproducibility of
the laboratory procedures much beyond the point where sampling errors
become a major source of variability. This, of course, will differ from ele-
ment to element depending on their dispersion characteristics and on the
way in which the samples are collected. A well designed prospecting pro-
gramme should therefore enable each of the principal sources of variation to
be assessed as shown in Table l-III.
Taking the existence of sampling errors into account a precision of ±10—
15% at the 95% confidence level is generally regarded as acceptable for
laboratory variability in most exploration programmes. Except for DC-ES,
which is best regarded as semi-quantitative, all of the instrumental methods
of analysis considered are capable of providing acceptable precision pro-
viding the elements are at concentrations several times higher than the detec-
tion limit of the method (the detection limit is defined as the concentration
at which precision becomes equal to ±100%, Fig. 2-5).

Sample throughput
If samples are digested in test tubes approximately one hundred can be
16

TABLE l-III
The ratio (R) of overall variance to analytical variance and analytical precision (P) for
analyses of minus 80-mesh stream sediments by emission spectroscopy and atomic ab-
sorption. Overall variance was estimated for 20 duplicate pairs of sediments collected
from a single stream draining sandstones, and P by 10 replicate analyses of one of these
samples. Results indicate that sampling errors are significantly greater than analytical
errors when atomic absorption is used. With semi-quantitative emission spectroscopy
analytical errors (except for Sn) are greater than sampling errors (data from Howarth and
Lowenstein, 1971)

Element Atomic absorption Emission spectroscopy

Ä1 P2 R P

Cu 6.8 5 - 3 2
Mn 92 14 (3.9) 42
Ni 7.1 8 - 8 8
Pb not detected — 68
Sn not determined 65 641
Zn 37 6 no data

R = ratio of population mean variance to analytical variance: — = R is less than 1; 3 =


R when within-site random error component differs from zero at 0.05 level of signifi-
cance; (3) = R when within-site random error component is not different from zero at
the 0.05 level of significance.
P = analytical precision: ±P percent.

handled on a 50 X 30 cm hotplate. Most of the colorimetric procedures


developed for geochemical prospecting enable about this number of deter-
minations per man day, whereas a technician operating an atomic absorption
spectrophotometer should be able to analyze 600—1000 solutions per day
(depending on the number of additional dilutions required by anomalous
samples). Alternatively one hundred solutions might be analyzed for six ele-
ments. With a second technician responsible for weighing and digestion of
samples this would amount to three hundred determinations per man day.
Atomic absorption is, however, a single-element method and if more than six
elements per hundred samples per day are required it becomes necessary to
consider either the use of more than one atomic absorption unit or to take
advantage of the simultaneous multi-element capability of analytical meth-
ods based on emission spectra.
Direct-reading emission spectrometers and X-ray emission (fluorescence)
spectrometry can both provide rapid sample throughput with simultaneous
multi-element determinations, the number of elements determined ranging
from perhaps as few as ten to more than forty depending on the number of
analytical channels provided. The multi-element capability of ES has always
been valued in geochemical prospecting, whereas the greater cost and poorer
detection limits of multi-channel (wavelength-dispersive) X-ray spectrometers
17

have discouraged their installation in exploration laboratories, although there


are some notable exceptions. For example, the Swedish Geological Survey
uses XRF extensively for analysis of heavy minerals separated from till
(Brundin and Bergström, 1977), and the Research Laboratory of Anglo
American has used XRF to analyze an average of 300 prospecting samples
per month, peaking to more than 600 per day. One notable advantage of
XRF, compared to other methods of analysis, is that under favourable condi-
tions the loose sample powder can simply be transferred to a sample container
and analyzed directly.
Direct-reading spectrometers and multi-channel XRF units have an
enormous analytical capacity: they are also relatively sophisticated and
correspondingly less robust than simpler instruments. Their use is therefore
only practical or justified in the largest laboratories. If multi-element analy-
ses are required and semi-quantitative data are acceptable, photographic
recording of DC-arc spectra should be considered; using the procedure
development by Nichol and Henderson-Hamilton (1965) up to 20 samples
can be analyzed for 15 elements per man day.
Irrespective of the method of analysis, sample throughput can be
increased and errors reduced by organizing an efficient flow of samples
through the laboratory. A worksheet to accompany the samples and to be
completed at each stage of the analysis, enables their progress to be fol-
lowed and delays remedied. Transfer and manipulation of samples should be
kept to a minimum and maximum use made of the many laboratory aids
now available for aliquoting and diluting solutions. (New equipment designs
should always be checked as sources of contamination — the colourful pig-
ments used in plastic ware often contain large quantities of the trace ele-
ments!)

Location of the laboratory and other factors


Samples are commonly transferred, perhaps after drying and sieving in
field camps, to a central in-house or commercial laboratory. This provides
greater flexibility and control, at the sacrifice of the time lost in transit of
samples and results, than analysis on-site or in camp. When minimum delay
is essential, for example in anomaly follow-up or on an overburden drilling
programme, a field laboratory can save time and avoids the need to disrupt
normal laboratory operations with requests for "High Priority — Rush"
analyses. Many colorimetric procedures are readily adapted to simple field
kit or field laboratory procedures and portable XRF analyzers are also well
suited to many such situations — particularly if ore or sub-ore grades of
finely disseminated or weathered mineralization are to be estimated. More
sophisticated field laboratories equipped with atomic absorption have also
been described (Horton and Lynch, 1975) and the U.S. Geological Survey
has used trailers equipped with emission spectrographs for many years
(Canney et al., 1957). Mobile laboratories have also been described by Smith
00

TABLE 1-IV
Evaluation of performance of analytical methods commonly used in exploration geochemistry (see Figs. 1-9, 1-10 and 1-11 for
analytical sensitivities)

Method Cost Precision Freedom Multi- Deter- Solid Comments


($) from inter- element minations sampl es
ferences capa- per man
bility day

1. Colorimetry 1 X 10 3 poor-good good no 20-100 no very simple: adaptable


) to field use; special
reagents needed for
each element
2. Atomic absorption 2 X 10 4 good very good no 500 no easy to set-up and
operate: several ele-
ments can be deter-
mined on same solu-
tion; special methods
for Hg, As, Te, Se;
dilutions required for
high concentrations
3. Emission spectroscopy
visual comparison 4 X 10 4 very poor very poor yes 500 yes simple robust equip-
ment: requires com-
parator and darkroom;
results semi-quantita-
tive
5
direct reader 1—2X10 poor poor yes >1000 yes sophisticated equip-
ment: requires ex-
perience^ analyst to
set-up and supervise
ICP-direct reader 1- 2 X 10 5 good good yes >1000 no operations; requires
dedicated computer
or access to computer
4. X-ray fluorescence
wavelength dispersive 3 X 10 5 good good yes >1000 yes sophisticated equip-
ment: requires super-
energy dispersive 1 X 10 5 good (?) good (?) yes > 500 yes vision by a skilled
analyst; dedicated com-
puter or access to
computer; very simple
sample preparation
20

TABLE 1-V
Methods of estimation 1970—1971 (from A.E.G. Analysis Committee, 1 9 7 1 )

Method Percent of total number of samples

Canada United States

Atomic absorption 80.6 65


Emission spectroscopy 4.3 14
Colorimetry 14.3 17
Cold extraction colorimetry 3.3 4
X-ray fluorescence 4.4 0.1
Paper chromatography 0.7 0.2
Selective ion electrode 0.5 1.0
Other 3.5 4.0
Total number of samples 8 1 2 , 4 56
56 337,370

and Washington (1962), Holman and Durham (1967), Kvalheim (1967), and
Kinson and Belcher (1970).
In making the final choice of an analytical method other factors which
must, of course, be considered include (1) the capital and operating costs of
equipment and the facilities available for servicing and repairing it; (2) costs
and availability of reagents or suitable local substitutes; and (3) the extent to

TABLE 1-VI
Approximate costs of sample preparation and geochemical analysis by commercial labo-
ratories, Vancouver, B.C., Canada, 1979—1980

Procedure Cost per sample


($)
Soil or sediment: dry and sieve $ 0.45
Rock: crush and pulverize $ 2.00
Hot acid extraction/atomic absorption 1st element $ 1.50; 2nd element $ 0.65
Cu, Mo, Pb, Zn, Ag, Cd, Ni, Co, Fe, Mn then $ 0.55 for additional elements
Twenty to thirty elements by semi- $ 20.00—25.00
quantitative emission spectroscopy

Special analyses
Sb: solvent extraction/atomic absorption $ 3.00
As: hydride generation/atomic absorption $ 3.00
Au: solvent extraction/atomic absorption $ 3.00
Sn: solvent extraction/atomic absorption $ 3.00
U: fluorimetric $ 3.00
21

which professional staff are essential to set up and supervise the operation of
the equipment (in general, maximum use should be made of non-professional
staff for all routine analysis). A qualitative attempt has been made to
evaluate the principal analytical methods with respect to these and other fac-
tors in Table 1-IV; the impact of these factors on usage of these methods is
reflected in Table 1-V and on analytical cost in Table 1-VI. It should be
noted that the adequacy of equipment maintenance and reagent supply
varies greatly throughout the world: with respect to instrumentation it is
often easier to repair older units locally than state-of-the-art microprocessors.
As an exploration technique, geochemical prospecting has the outstanding
virtue that systematic use of even the simplest equipment can successfully
lead to the discovery of both exposed and concealed orebodies.

REPORTING RESULTS

From the foregoing discussion of the choice of analytical methods it is


apparent that the same sample could be analyzed by many different proce-
dures depending on the objectives of the analysis and the facilities available.
Furthermore, although many laboratories might use the same general method,
detailed procedural steps (for example, solution to sample ratios and extrac-
tion temperatures) will often differ substantially. As a result, very different
metal contents can be reported for the same sample without any errors in
the analytical technique being involved. An example of this is shown in the
analysis of a stream sediment in Table 1-VII, for which the reported Ni
concentrations range from less than 20 ppm to more than 1000 ppm. Results
of this kind emphasize the need for the geologist or geochemist to clearly
specify the analysis required (and understand its implications) and for the
laboratory to provide a full description of the procedure actually used.
Neither a request for "AA copper" nor the description "Acid extractable AA
copper" can be regarded as adequate. The analysis required is usually
specified on a "Request for Analysis" form — an example is shown in
Fig. 1-12 which should be completed as fully as possible.
In addition to reporting metal concentrations the analyst should also
monitor analytical reproducibility and long term drift. Providing results are
within acceptable limits (and the geologist is aware of what those limits are)
there may be no need to report them. They should, however, be readily
available on demand for any particular batch or day's work. Preparation of
standards and plotting of control graphs does require some additional effort
but will avoid the pursuit of spurious geochemical trends resulting from
analytical errors.
It is not unusual for results to be required again many years after the
original analysis. The laboratory files may then become the last resort. A
22

TABLE 1-VII
Comparison of six stream sediment analyses for nickel (from Hansuld et al., 1969)

Preparation Digestion Nickel content


(ppm)
crushing fraction used
(mesh)

Lab A no minus 80 70% HC10 4 20


LabB no minus 80 1 : 3 HNO3 60
LabC no minus 80 HNO3/HCI mix*id acid 150
LabD no minus 80 1 : 1 HC1 320
LabE yes minus 100 1 : 1 HC1 14
LabF no minus 100 1 : 1 HC1 1120

Sample description
stream sediment containing 0.5% magnetite
magnetite contains average of 0.28% Ni
80% of magnetite is minus 100 mesh
99% of sample is plus 100 mesh and 96% is plus 80 mesh

From: Results to:


Number of sampies =__

Material· Rock = Soi I jnt


= Sedim« Other:
Sample prep. : 1Dry Sieve - 8 0 Retain+80 Pulverize ■
Other:
:
Extraction: HNO3/HCIO4 = 0.05M EDTA 1.0M HCI
Other:
Elements:

:
Emission Spec. : 10 elements = 20 elements Other
Special instructions:
Date:
Sample No. Location Sample No. Location

Fig. 1-12. A typical "Request for Analysis" form.


23

complete and well-ordered analytical archive should therefore be maintained


of all analyses.

SAFETY

Although the decline in use of colorimetric techniques has greatly reduced


the number of reagents routinely required in the exploration laboratory a
considerable number of potential hazards remain. These are summarized in
Table 1-VIII together with an indication of the appropriate precautions:
Hazards in the Chemical Laboratory (Muir, 1977) or some other laboratory
safety manual should be consulted for detailed information.
Apart from the obvious dangers in working with strong acids, areas of
particular concern include (1) control of dust during sample preparation;
(2) the hazard of explosion from hot perchloric acid in contact with organic
material; and (3) the insidious effects of inhalation of toxic vapours — this is
particularly a problem in hot climates with inadequate ventilation of the
laboratory. A first aid kit, eye wash and fire fighting equipment should all be
rapidly available and everyone should be familiar with their use. Eating,
drinking and smoking should be absolutely forbidden in working areas and
non-professional staff made aware of any potential dangers involved in
mixing reagents, transferring solutions or operating equipment. Because of
their use in remote regions, safety is of particular concern in mobile labora-
tories.

TABLE 1-VIII

Some hazards and safety precautions in the laboratory

Hazard Precautions

Dust during rock crushing and work in well ventilated area with efficient dust
grinding or sieving extraction system (see p. 48)
Flammable liquids and vapours store minimum quantities in approved cabinets;
(MIBK, benzene, acetone, etc.) work in fume hood; fire fighting equipment
readily available
Toxic vapours (benzene, carbon avoid use if possible; use only in well ventilated
tetrachloride, chloroform, toluene, areas and in fume hoods
etc.)
Radiation: X-rays check equipment for X-rays leakage with a
scintillometer; regular operators wear
dosimeters
radon gas avoid working on high-grade uranium samples
in poorly ventilated areas
1
Guidelines only — consult Hazards in the Chemical Laboratory (Muir, 1977) or similar
manual for detailed information. See p. 63 for use of perchloric acid.
Chapter 2

QUALITY CONTROL IN THE LABORATORY

INTRODUCTION

To make the most efficient use of geochemical analyses their reliability


must be known. Quality control throughout sample preparation and analysis
is therefore as important in the geochemical laboratory as in any other
production facility and should be regarded as an essential aspect of any
geochemical prospecting programme. Unfortunately, it is an aspect of analy-
sis that all too often is neglected by both the laboratory and its clients.
Two types of errors contribute to unreliability of an analytical result:

>- _ _
u C. p x D. p x

' JA CONCENTRATION
" ►
Fig. 2-1. Random and systematic errors. The variation in concentration caused by random
errors is represented by the normal curve with an average value X: μ is the true concentra-
tion of the analyte. A. The dispersion (width) of the normal curve is narrow and sym-
metrical around μ — i.e. X = μ and results are both accurate and precise. B. The disper-
sion is greater but still symmetrical about μ — precision is therefore relatively poor and
although the average value (X) is accurate, this is not necessarily true of individual analy-
ses. C. Dispersion is narrow but a systematic positive error has been introduced (X > μ) —
results are precise but inaccurate. D. Systematic error and poor precision.
26

random errors arising from the variations inherent to any sampling or mea-
surement process, and non-random errors causing systematic negative or
positive deviations from the true result. Accuracy, which is the closeness of a
result to its true value, is dependent on both random and systematic errors,
whereas precision, which is the ability to obtain the same result repeatedly,
is a measure of random errors alone. It follows that results can be precise
without being accurate (Fig. 2-1).
Exploration geochemistry utilizes geochemical patterns, that is natural
variation of element abundances, as a guide to the presence or absence of ore
mineralization. Relative abundances of elements are therefore more impor-
tant than their absolute concentrations and strictly, it is only necessary that
any variability introduced during sample collection and analysis be signifi-
cantly less than the variability sought (Fig. 2-2). Consequently, quality con-
trol in the exploration laboratory, unlike the assay office, usually emphasizes
relative variability (i.e. precision) over accuracy. Development of exploration
geochemistry has therefore followed, and encouraged, development of
simple, rapid analytical methods capable of achieving adequate precision.

2>
<
cc

ill
o
z
o
o

VEIN
Fig. 2-2. Influence of analytical precision on anomaly contrast. Noisy data (A) arising
from random errors in sampling or analysis obscures the anomaly. Although absolute con-
centrations are lower in (B) and (C), the data are less noisy and the anomaly contrast
improved.
27

RANDOM ERRORS

Precision

Random errors are assumed to follow a normal Gaussian distribution


about their mean concentration (c). Analytical precision is then specified as
the percent relative variation at the two standard deviation (95%) confidence
level:

Pc = — c -x 100%
c

where Pc is the precision in percent at concentration c, and Sc is an estimate


of the standard deviation (oc) at that concentration.
A value of Pc = ±20% for 70 ppm Cu indicates that, on the average, 95 out
of 100 analyses of the sample will be within the range 56—84 ppm: the five
remaining results will be outside this range.This is equivalent to a precision
of ±10% at the one standard deviation (68%) confidence level or a coeffi-
cient of variation of 0.10. Commonly, a laboratory precision of ±10 to ±15%
is regarded as adequate for most prospecting purposes. However, if anomaly
contrast is low a relatively poor precision increases the probability of
anomalous samples going unrecognized (Howarth and Martin, 1979).
Calculation of precision is often based on replicate analysis of selected
samples to determine Sc, or on analyses of a statistical series prepared by
mixing samples with high and low metal concentrations in definite propor-
tions (Craven, 1954; Stanton, 1966; James, 1970). However, as discussed by
Thompson and Howarth (1973, 1976, 1978), these approaches have several
disadvantages:
(1) Control samples are not necessarily representative, either physically or
chemically, of the samples as a whole. Furthermore, with a statistical series
there is the inherent difficulty of producing homogeneous mixtures of
powders.
(2) Precision is either obtained at specific concentrations or an average
precision value is obtained for a range of concentrations.
To overcome these failings alternative ways of estimating precision using
actual samples, randomly selected and analyzed in duplicate, have been con-
sidered in detail by Thompson and Howarth (1973, 1976, 1978) and
Howarth and Thompson (1976). Briefly, their method employs the absolute
difference \Xl — X21 between pairs of duplicate analyses (Xu X2) as an esti-
mator of the standard deviation (a c ), and the mean value (Xx + X2)/2 as an
estimator of average concentration. Ι^Ί — X21 is normally distributed and
relates to the parent population, with a standard deviation a c , such that:

od = y/2oc
28

where od is the standard deviation of the difference \Xl—X2\\

d = 1.128ac

where d is the mean value for the difference; and:

Md = 0.954a c

where Md is the median value for the difference. oc can be obtained from
each of these relationships but the median (Md) is the most convenient
estimator because it is (1) relatively little affected by wild values; (2) readily
estimated graphically; and (3) corresponds very closely to oc without further
calculation. The following rapid procedure is suggested for estimation of
precision from a minimum of 50 pairs of duplicates (Thompson and
Howarth, 1976).
(1) From the duplicate analyses obtain a list of the means (Xx + X2)/2 and
absolute differences Ι-ΧΊ — X21.
(2) Arrange the list in increasing order of concentration means.
(3) From the first 11 results obtain the mean concentration and median
difference for that group.
(4) Repeat this for each successive group of 11 results, ignoring any
remainder less than 11.
(5) Calculate, or obtain graphically, the linear regression of the median
differences on the means and multiply the intercept and coefficient by 1.048
(i.e. 1/0.954) to obtain σ0 and k, respectively. As an example the variation of
oc with c for the determination of copper in soil samples, based on 50 dupli-
cate analyses, is shown in Fig. 2-3. From the regression oc = σ0 + kc so that
Pc is given by:

This method of estimating Pc needs a minimum of 50 duplicate analyses.


With fewer pairs it is still possible to judge if they conform to a particular
precision by graphically comparing \Xi~X2\ versus (Χχ +X2)/2 with the
distribution percentiles for Md at any specified precision. Thus, in Fig. 2-4,
all but one of thirteen duplicate Cu analyses fall below the 90th percentile
on a Pc = 10% control graph. Reference to tables of the binomial probability
of M out of AT points falling above the 90th percentile, tabulated in Table 2-1
from Thompson and Howarth (1978), indicates that chance alone gives an
approximately 75% (p = 0.745813) probability of such an event. Results are
therefore judged to be consistent with a precision of ±10%. In contrast four
or more points above the 90th percentile (with p = 0.034161 for N = 13 and
M = 4) would have suggested that precision was almost certainly worse than
±10%.
29

KEY
* Individual points
O Interval median

50
-*i_
100 +
150

(X!+X2) / 2

Fig. 2-3. Regression of the median differences of \Χχ — X2 I against average concentration
(Xi + X2)/2 for duplicate pairs of analyses. Vertical dashed lines separate intervals con-
taining eleven pairs of analyses and their median value.

To construct similar control graphs at different levels of precision it is


necessary to understand the origin of the distribution percentiles. The
50th percentile for Md at concentration c is given by 0.6745\A2ac, that is
0.954a c , where 0.6745 is the normal deviate for the 50th percentile of the
half normal distribution. Similarly the 90th and 99th percentiles with nor-
mal deviates of 1.6449 and 2.5758, are 2.3262a c and 3.6427a c respectively.
Hence, to construct a Pc = 20% control graph: at a concentration of 100 ppm
and precision of ±20%, oc is 10 ppm and the control points for the 90th and
99th percentiles of \XX — X21 will be at 23.3 and 36.4 ppm, respectively.
In determining precision from duplicate pairs it should be noted if the
duplicates are all from the same analytical batch, only within batch precision
is estimated. A more realistic estimate of overall precision is obtained by
analyzing duplicates randomly distributed throughout many batches. Diffi-
culties in the use of duplicates to estimate precision arise when differences
between them follow a non-gaussian distribution. This can happen if: (1) the
sample is heterogeneous and sampling errors are skewed; (2) concentrations
are close to the resolution of the analytical method and results are reported
as discrete values giving a discontinuous distribution; (3) concentrations are
30

/ s*

— —



»
U 2
X ·*.

1
10 20 50 100 200 500 1000
(Xi+X2)/2
Fig. 2-4. A precision control chart for P = ±10% at the 95% confidence level. See text for
explanation of data points.

close to the detection limit and values below the limit are set to zero,
recorded as "less than" or are set to the detection limit; and (4) if systematic
differences (bias) arise between batches (Thompson and Howarth, 1976).

Detection limits

The detection limit for an analytical System is the minimum concentra-


tion than can be detected with a specified degree of confidence. It follows
that it is closely dependent on the noise (σ0) associated with the zero con-
centration (blank) measurement. This, from the foregoing discussion of
precision, is readily estimated graphically, as in Fig. 2-3, or by calculation of
the linear regression of IX! — X21 on (Xi +X 2 )/2 to give Md at c 0 , and σ0
from Md = 0.954a c . Usually two standard deviations are taken as the con-
fidence level and the detection limit is then defined as the concentration that
gives rise to a signal equal to twice the standard deviation of blank fluctua-
tions: thus at the detection limit Pc = 100% and c = 2σ0 (Fig. 2-5). At con-
centrations higher than the detection limit precision falls asymptotically
towards the value of 2k in the expression:

Some implications of this are discussed by Thompson and Howarth (1976).


Instrument manufacturers generally quote 2σ0 detection limits for instru-
TABLE 2-1
Tables of the probability that M or more points out of N will fall above the 90th and
99th percentiles of a precision control chart (from Thompson and Howarth, 1978)

A. 90th percentile (single event probability = 0.100000)

100000
190000 .010000
271000 .028000 .001000
3-43900 .052300 .003700 .000100
4095,10 .081460 .008560 .000460 .Ü00010

468659 .114265 .015850 .001270 .000066 .000001


521703 .149694 .025692 .002726 .00017; .000006 .000000
569633 .186895 .038092 .005024 .000432 .000023 .000001 .CÜÜCÜC
612580 .225159 .052972 .008331 .000891 .000064 .00000 5 .31X000 . ..\;JL\. C - " ■ —
651322 .263901 .070191 .012795 .001636 .000147 .000009 .000000 .CCC03C

686189 .302643 .089562 .018535 .002761 .000296 .000023 .G00001 .JCC3ÜC :.:ccccc
717570 .340998 .110870 .025637 .004329 .000641 .000060 .000003 .200300 CC02C2
7-46813 .378655 .133883 .034161 .006460 .000920 .000099 .000008 . JCOCGC 3CCCCC
771232 .415371 .158360 .044133 .009230 .001474 .000181 .00001 7 .2000c: 22-0222
79-4109 .450957 .184061 .055556 .012720 .002250 .000311 .000034 ..'J000C3 222222

81-4698 .485272 .210751 .068406 .017004 .003297 .000606 .000061 . ICCCCb 222300
833228 .518215 .238203 .082641 .022144 .00466 7 .000 7^4 .000106 -ocoo:: 222C21
8-49905 .549716 .266204 .098197 .028194 .006416 .001172 .000173 .JUÜ021 3C0022
864915 .579735 .294555 .114998 .035194 .008693 .001696 .000273 .000036 000004
878423 .608253 .323073 .132953 .043174 .011253 .002386 .000416 ,000060 CCC2C?

890581 .635270 .351591 .151965 .062162 .014446 .0032 73 .000613 .000096 oooc: 2
901523 .660801 .3 79959 .171928 .062134 .018216 .004390 .0008 79 .00014/ 0CCC21
911371 .684873 .408043 .192731 .073113 .022608 .006773 .001230 .000220 000033
920234 .707523 .435726 .214262 .085075 .027658 .007466 .001684 .000 321 000062
928210 .728794 .462906 .236409 .097994 .033400 .009476 .002261 .000468 OCCC79

935389 .748736 .489495 .259058 . 111835 .039859 .011869 •0D2983 .000638 CCC11 ?
941850 .767401 .515419 .282102 .12655 7 .04706 7 .014668 .0038 7 1 .0008 72 000169
947665 .784846 .540617 -.305434 .142112 .055007 .017907 .004961 .001172 C00239 2C0C4?
952899 .801128 .565040 .328952 .158444 .063717 .021617 .00624 7 .001660 3CQ333 200062
957609 .816305 .588649 .352561 .175495 .073190 .025827 .007784 .-.702020 0OC464 202089

961848 .830435 .611414 .376170 .193201 .083421 .030563 .009688 .002696 000611
965663 .843577 .633316 .399694 .211498 .094399 .035849 .011686 .003296 0C0809
969097 .855785 .654342 .422066 .230318 . 106109 .041704 .014102 .004134 001068
972187 .867116 .674487 .446185 .249592 .118530 .048144 .016862 .006131 001366
974968 .877624 .693750 .469015 .269251 .131636 .066183 .019990 .006304 001740

977472 .887358 .712137 .491489 .289227 .146397 .062828 .023509 .00 7673 3C2198 3666
979724 .896369 .729659 .613553 .309454 . 159780 .071086 .027441 .009256 CC2 7->6
981752 .904705 .746330 .635164 .329864 .174748 .079955 .031806 .011076 0C3397 192 3
983577 .912410 .762168 .566281 .350394 .190259 .089434 .036621 .C1314Ö 20-Ί66
985219 .919526 .777192 .576869 .370982 .206273 .099616 .041302 .016496 226263

B. 99th percentile (single event probability = 0.010000)

N M=l M=2 f-U3 M=4 M-6 ri=6 M- / M-6 r-i-


9 r-t-1
0 Mil '-.- u
.010000
2 .019900 .000100
3 .029701 .000298 .000001
4 .039404 .000692 .000004 .000000
6 .049010 .000980 .000010 .000000 .000000

6 .058520 .001460 .000020 .000000 .000000 .000000


7 .067936 .002031 .000034 .000000 .000000 .000000 .000000
8 .077255 .002690 .000054 .000001 .000000 .000000 .000000 .000000
9 .086483 .003436 .000080 .000001 .000000 .000000 .000000 .000000 .000000
10 .095618 .004266 .000114 .000002 .000000 .000000 .000000 .000000 .000000 .000000

11 .104662 .005180 .000165 .000003 .000000 .000000 .000000 .000000 .000000 .000000 .000000
12 .113615 .006175 .000206 .000006 .000000 .000000 .000000 .000000 .000000 .000000 .000000 .oooooo
13 .122479 .007249 .000266 .000007 .000000 .000000 .000000 .000000 .000000 .000000 .000000 .oooooo
14 131254 .008401 .000336 .000009 .000000 .000000 .000000 .D00000 .000000 .000000 .000000 .oooooo
15 .139942 .009630 .000416 .000012 .000000 .000000 .000000 .000000 .000000 .000000 .000000 .oooooo
16 .148642 .010933 .000508 .000017 .000000 .000000 .000000 .000000 .000000 .000000 .000000 .000000
17 . 157057 .012309 .000612 .000021 .000001 .000000 .000000 .000000 .000000 .000000 .000000 .000000
18 .165486 .013756 .000729 .000027 .000001 .000000 .000000 .000000 .000000 .000000 .000000 .oooooo
19 .173831 .015274 .000859 .000034 .000001 .000000 .000000 .000000 .000000 .000000 .000000 .oooooo
20 .182093 .016859 .001004 .000043 .000001 .000000 .000000 .000000 .000000 .000000 .000000 .oooooo
21 .190272 .018512 .001162 .000052 .000002 .000000 .000000 .000000 .000000 .000000 .000000 .oooooo
22 .198369 .020229 .001336 .000063 .000002 .000000 .000000 .000000 .000000 .000000 .000000 .000000
23 .206386 .022011 .001525 .000076 .000003 .000000 .000000 .000000 .000000 .000000 .000000 .oooooo
24 .214322 .023854 .001729 .000091 .000004 .000000 .000000 .000000 .000000 .000000 .oooooo .oooooo
25 .222179 .025759 .001951 .000107 .000004 .000000 .000000 .000000 .000000 .000000 .oooooo .oooooo
26 .229957 .027723 .002189 .000125 .000006 .000000 .000000 .000000 .000000 .000000 .oooooo .oooooo
27 .237657 .029746 .002444 .000146 .000007 .000000 .000000 .000000 .000000 .000000 .oooooo .oooooo
28 .245281 .031825 .002717 .000169 .000008 .000000 .000000 .000000 .000000 .000000 .oooooo .oooooo
29 .252828 .033959 .003008 .000194 .000010 .000000 .000000 .000000 .000000 .000000 .oooooo .000000
30 .260300 .036148 .003318 .000223 .000012 .000000 .000000 .000000 .000000 .000000 .oooooo .ocuooo
31 .267697 .038390 .003646 .000254 .000014 .000001 .000000 .000000 .000000 .000000 .oooooo .oooooo
32 .275020 .040683 .003993 .000287 .000016 .000001 .000000 .000000 .000000 .000000 .000000 .oooooo
33 .282269 .043026 .004360 .000325 .000019 .000001 .000000 .000000 .000000 .000000 .oooooo .oooooo
34 .289447 .045418 .004747 .000365 .000022 .000001 .000000 .000000 .000000 .000000 .oooooo .oooooo
35 .296552 .047859 .005154 .000409 .000025 .000001 .000000 .000000 .000000 .000000 .oooooo .000000

36 .303587 .050346 .005581 .000456 .000029 .000002 .000000 .000000 .000000 .000000 .oooooo .oooooo
37 .310551 .052878 .006028 .000507 .000033 .000002 .000000 .000000 .000000 .000000 .000000 .oooooo
38 .317445 .055455 .006497 .000563 .000038 .000002 .000000 .000000 .000000 .000000 .oooooo .oooooo
39 .324271 .058075 .006986 .000622 .000043 .000002 .000000 .000000 .000000 .000000 .oooooo .000000
40 .331028 .060737 .007497 .000686 .000049 .000003 .000000 .000000 .000000 .000000 .000000 .000000
32

Pc%=(^+2k)l00

σ0 =5
z k =0.05
o
CO
Ü
111 \
cc
Q.

20

100 k U·-

CONCENTRATION, c
Fig. 2-5. Variation of precision with concentration for OQ = 5 and k = 0.05. The detection
limit (d = 11) is the concentration corresponding to P = ±100%.

ment performance obtained in an ideal matrix and under ideal conditions


with respect to stability so that in some cases a higher multiple of σ0 is
proposed as a more realistic "working" limit. However, the most meaning-
ful estimate of a laboratory detection limit can be obtained, as described
here, as part of the routine monitoring of laboratory precision. With dupli-
cates carried through the complete analytical procedure this will take into
account random errors introduced during sample decomposition as well as
those generated during the final measurement.

Sources of and reduction of random errors

So far only the overall analytical precision, arising from random errors
throughout sample preparation, decomposition and measurement, has been
considered. Occasionally it may be necessary to determine the relative
magnitude of errors generated at each step of the analysis. By including
duplicates or replicates at appropriate stages in the analysis the necessary
estimates can be made from an analysis of variance.
As a simple example Table 2-II tabulates results of duplicate (; = 2) analy-
ses of a sample in each of six (i = 6) batches. We wish to evaluate the
possibility that errors between the batches are significantly greater than
those within batches. First, variance between batches (i>|) and within
batches (S^) is estimated from the corresponding sums of squares, SSB and
SSW, given by the relationships:

SSB = n Σ (C, - C)2 and SSW = Σ (Cy - Ct)2


33

TABLE 2-II
Analysis of variance for duplicate analyses from six batches of analyses

Analysis Batch (i) (ior analyst or decomposition)


U)
1 2 3 4 5 6 -k

1 41.0 39.4 39.6 43.3 42.3 36.9


2 42.4 39.8 40.2 43.1 42.9 36.3

n
Means
Q 41.7 39.6 39.9 43.2 42.6 36.6
c 40.6

Source Sum of Degrees of Variance F*


of error squares freedom ratio

Between SSB= 58.92 k —1 = 5 si = SSB/5 = 11.78


11.78/0.27 = 43.6
Within SSW= 1.62 N-k = 6 = SSW/6 = 0.27
Total SSC = 60.54 N—l = 11 sh = SSC/11 = 5.50

* From tables 95F(5>6) = 4.39 and 99F(95>6) = 20.81.

then:

SSB SSW
and Sw =
«-s N-k

where k is the number of batches (/), n the number of replications (;) and N
the total number of analyses: C is the overall average concentration and C,
the average of the n replications on the ith batch. The F ratio (F) is then
computed as:

Ei _ larger variance
smaller variance

and compared with values of F, from tables, at k — 1 and N — k degrees of


freedom.
For the data in Table 2-II the value of F(43.6) is much greater than
tabulated values at either p = 0.05 (F = 4.39) or p = 0.01 (F = 20.81). It is,
therefore, concluded that errors arising between batches make a significantly
greater contribution to overall laboratory error than within batch errors.
Reasons for this might be sought in variations of reagent quality, variation of
34

analytical conditions from day to day, or in the relative skill and care of
different analysts.
Analysis of variance, even in its simplest forms, is very versatile. Batches in
the worked example might equally well represent different laboratories or
analysts. Alternatively, by digesting the same sample several times and
analyzing each of the resulting solutions in duplicate, conclusions could have
been drawn about the relative contribution to overall errors of subsampling
and decomposition compared to the measurement step. Having identified the
relative magnitude of errors arising at each stage, efforts to improve overall
precision should focus on the sources of the greatest error — obviously in
Table 2-II considerable reduction of between batch error is needed to reduce
it to the level associated with within batch variations. By extending duplica-
tion to field sampling, total data variance can be partitioned between
regional variability, local (sampling) variability and analytical (laboratory)
variability. The probable significance of the regional patterns can then be
estimated (Plant, 1971; Plant et al., 1975; Chork, 1977; Howarth and
Lowenstein, 1971; Garrett, 1969; Garrett and Goss, 1979). As noted in
Chapter 1, there is little point expending effort to improve laboratory preci-
sion if errors generated during sample collection dominate the data.
Random errors can often be reduced, albeit at the expense of time and
cost, by more careful analysis: for example, by use of volumetric glassware
rather than calibrated test tubes, by more rigorous control of operating con-
ditions or by closer bracketing of unknowns with standards. Sometimes,
particularly in emission spectroscopy, rationing the analyte signal to that of
an internal standard, having the same characteristics in the analytical system
as the analyte, enables the effects of fluctuating conditions to be com-
pensated (Fig. 2-6).
Errors can also be reduced statistically. If the standard deviation for a
single measurement is S, the standard deviation Sa for the average of a series
of n measurements is:

Sa = S/y/n

Reduction of errors is therefore possible by replication at any step of the


analysis. At the measurement step this is achieved automatically in many
instruments by increasing the counting time or by rapid sampling and
averaging of the analyte signal.

Laboratory sampling

In a thoroughly mixed single-component sample, such as a monomineralic


mineral separate, sampling reproducibility improves as the number of grains
(n) increases. This can be achieved by taking a larger sample or by grinding
to reduce grain size. The sampling standard deviation is proportional to ll\fn
35

Si 252.8

Be 249.4

ISi 252.8
cc I Be 249.4

, o
z
UJ
I-
—I 1 1 1 1 1 1 —
i 1 1 1 —
i 1 1 1 1
5 10 15 20

NUMBER OF ARCINGS
Fig. 2-6. Use of an internal standard to correct for- poor reproducibility in the DC-arc.
(Reproduced from Ahrens and Taylor, 1961, Spectrochemical Analysis, 2nd ed., with
permission of Addison-Wesley, Advanced Book Program.)

so that the standard deviation, £,·, for weight Wj can be related to standard
deviation, Sh at weight Wi by:

s,-*· y/Wj/Wi
3

For multi-component samples in which the element of interest is more or


less uniformly distributed throughout the components, the single-component
model is a reasonable approximation of the effects of sample size on repro-
ducibility. At this point it should be noted that grinding to achieve further
reduction of grain size can seldom be justified in routine analysis of soils and
sediments. Unground minus 80-mesh material is, therefore, often taken for
analysis because, even with small sub-samples ( ^ 1 0 0 mg), adequate sampling
reproducibility can be obtained for many (but not all) samples (Fig. 3-1).
Furthermore, as well as increasing the risks of contamination, grinding will
not invariably reduce sampling errors. For example, grinding to pass 200 mesh
(75 μηι) would not effect any improvement in sampling error for gold if it
was already present in the unground material as minus 200-mesh particles.
The effect of grinding on sub-sampling reproducibility is therefore complex,
depending on the distribution of the element of interest between size frac-
36

tions. Some examples are described by Howarth and Lowenstein (1971) and
Plant (1971).
With multi-component mixtures in which the abundance of the analyte
differs appreciably between components, the effects of sampling are more
complex. In the simplest case with two-component mixtures the binomial
distribution can be used to estimate the relative standard deviation (R).
Using the nomenclature of Ingamells and Switzer (1973) and Ingamells
(1974):

D _100y/pq B-H
y/n K

where p and q are the volume proportions of constituents with H and B


content of the element of interest, respectively, and K is the overall content
of the element. Providing the densities of the mineral grains are the same
their proportions by weight or volume will also be the same and the number
of grains (n) can be estimated by:

w
n
" dour V)3
where w is sample weight in grams, d is density and μ the effective mesh size
in microns. If densities of the mineral species differ n can be obtained more
accurately from:

= WB
+ WH

3
dß(io-V) dH(io-V)3
although this correction will be negligible if B is only present as rare grains
of an accessory mineral. The relationship between volume proportions (p, q)
and weight proportions (p w , qw) with minerals of different densities is given
by:

_ dBpw _ dpq^
dßPw + dHqw dBpw + dHqw

These relationships have been used to derive R for different mesh sizes
and different proportions (q) of an accessory mineral (B) containing 80% of
the element of interest compared to 1 ppm in the gangue (H) (Fig. 2-7).
Densities dB and dH have been taken as 7.0 and 3.0, respectively, and the
model would correspond approximately to a soil or sediment containing rare
grains of cassiterite. The same assumptions and values have been used to esti-
mate R at different sub-sample weights in Table 2-III. It is apparent that
sampling errors increase to a maximum at the concentration at which both
37

100

80

60
o\
cc

20

/'.'
υ
/^
P ' 100 ' ' 200
mesh 200 100 80
Grain Size
Fig. 2-7. The relative error (R%) for an element (X) present as a major constituent (80%)
of a trace mineral. Concentration of the element in the gangue is 1 ppm, sub-sample size
is 1 g, and the densities of the gangue and special mineral are 3 and 7, respectively.
Dashed curves indicate the relative proportions of grains of the special mineral (q) and
gangue (p). Curve A: q = 4 X 10" 4 ^ 7 5 0 p p m X ; B : q = 4 X 10" 5 ^ 7 5 p p m X ; C : q = 2 X
10" 5 ^ 40 ppm X; D: q = 4 X 10" 8 ^ 1 ppm X; E: q = 4 x 10" 6 ^ 8 ppm X; F: q = 2 X
10~ 6 ^ 5 ppm X; G: q = 4 X 10~ 7 ^ 2 ppm X. Solid lines separate fields containing, on
the average, (/) more than five grains of the special mineral; (//) between one and five
grains; and (///) less than one grain.

the accessory mineral and the gangue contribute equally to the overall
content. Under the circumstances described this occurs at 2 ppm, when q ^
5.36 XI0" 7 . At concentrations above and below this the sampling error
decreases. The latter trend, however, would be of little practical significance
if errors from other sources were rising rapidly towards the detection limit of
the analytical method.
Increasing sample weight (or grinding) reduces R, but in the circumstances
described even 1-g sub-samples are inadequate (R > 50%) for materials
coarser than 100 mesh containing low, but possibly anomalous tin values, up
to about 20 ppm. Using emission spectroscopy, consuming a few tens of
milligrams of sample, a single analysis or even the average of several analyses
is unlikely to indicate the overall content of an element present as a major
constituent of a few rare mineral grains. One practical solution to this prob-
lem if the rare mineral has a distinctly higher density than the gangue., is to
pan, or otherwise separate, a heavy mineral fraction thereby greatly
38

TABLE 2-III
Effect of sub-sample weight and grain size on sampling error (R) for tin present as rare
grains of cassiterite in a gangue containing 1 ppm

Sn Mesh Sub- Number Average number R


(ppm) sample of grains * of cassiterite (%)
(g) grains
2
8.5 80 0.2 12,022 0.05 173
(177 Mm) 0.5 30,056 0.12 109
1.0 60,112 0.24 77
200 0.2 158,025 0.63 48
(75 Mm) 0.5 395,062 1.58 30
1.0 790,124 3.16 21
75 80 0.2 12,022 0.63 61
0.5 30,056 1.20 39
1.0 60,112 2.40 27
200 0.2 158,025 6.32 17
0.5 395,062 15.80 11
1.0 790,124 31.61 7.5
2
n = u>/<2(10~4 M)3) with d = 3.0 and assuming error introduced by neglecting density of
cassiterite to be negligible.
2
For 8.5 ppm Sn volume proportion of cassiterite (q) = 0.000004; for 75 ppm q -
0.00004.

increasing the effective sample weight and reducing R accordingly.


So far it has been assumed that random errors, estimated as the relative
standard deviation (i?), will follow a Gaussian distribution with a sym-
metrical distribution around their mean. When sampling for trace elements
present as major constituents of rare mineral grains, this is not necessarily so
and normality becomes unlikely or impossible when JR exceeds 50%. In some
areas of Fig. 2-7 and Table 2-III only a fraction of a single whole grain of
cassiterite is present. However, real grains are not divisible in this manner and
under these circumstances the outcome of laboratory sub-sampling becomes
very dependent on the size of the sub-sample and scarcity of special grains.
Three possibilities can be recognized (Ingamels and Switzer, 1973).
(1) Sub-samples are so small, as in emission spectroscopy, and grains so
scarce that the chances of a sub-sample containing a grain are remote.
Sampling errors will be approximately Gaussian and the background con-
centration of the element in the matrix or gangue will be estimated.
(2) The average sub-sample contains fewer than 5 special grains. Single
analytical values follow a strongly skewed Poisson distribution with a pre-
ponderance of low values and occasional wild highs. The average of a few
replicates, or of larger sets if wild results are ignored, seriously underesti-
mates the true trace element content.
39

(3) The average sub-sample contains more than 5 special grains. The Poisson
distribution becomes approximately Gaussian and replicate analyses provide
an estimate of the true mean. Relative standard deviation (R) depends on the
size of the sample and number of grains taken for analysis as described previ-
ously.
Three zones corresponding approximately to these conditions, with the
average number of grains (z) fewer than one, between one and five, and
greater than five, are indicated in Fig. 2-7. The probability (Pn) that a certain
number of grains (n) will be found in a particular sub-sample is estimated
from the Poisson distribution:

f
so that the probability (P0) of finding no special grains is e and probabil-
ities for 1, 2, ..., n grains are calculated from:

P =- i - ·P

A single analytical result will have the value (Xi):

Xt=H+ Cizi

when H is the content of the element of interest in the gangue, Q the con-
tribution of a single special grain to the overall concentration, and z{ the
number of special grains. Poisson distributions with z = 0.4, 1.0, 3.0 and
10.0 are shown in Fig. 2-8.
Returning to our example of rare grains of cassiterite in a soil or sediment,
one 100-mesh (150 μπι) grain weighs 23.6 μg with a Sn content (C) of
18.9 μg. If the average number of grains in a 1-g sub-sample is z = 1 the
overall content, with 1 ppm in the gangue, is approximately 20 ppm. How-
ever, with z = 1, P0 = 0.3679 and there is therefore a 37% chance of such a
sub-sample analyzing as only 1 ppm Sn (Fig. 2-8). Similarly, at an overall Sn
content of 8.5 ppm (q = 4 X 10" 6 , z = 0.4) the sampling distribution is
strongly skewed towards low values with approximately two thirds of the
samples reporting the background content (1 ppm): no single analysis will
correspond to the overall Sn content but one in twenty analyses will be as
high as 40 ppm. Skewed Poisson distributions are often reported for Au
(Jones and Beaven, 1971) and extreme examples have been reported for
diamonds (Phillips, 1971). The problems of sample size in estimation of Au
content are throughly discussed by Clifton et al. (1969).
Sampling errors can also create problems when data obtained by different
analytical methods, using different size sub-samples, are compared. Ingamells
40

z = 10.0
~c= 190 ppm

z = 3.0
c= 58 ppm

>-
o
z z = 1.0
LU

D 50 ~c= 19 ppm
σ
LU
CO L3^
z = 0.4
c^=8.5ppm Sn

' 7 ' 8 ' 9 Ί θ ' 1 1 '12'13' 1 4 ' 1 5 ' 1 6 ' 1 7 ' 1 8 ' 1 9 '
,
|o>|o|»|Hcol i|2?!S!|:r|o
|i-|co|tf)|^|o)|^|$2|i2|^|q> o N k © oo o N k <£

Sn ppm = H + z
Fig. 2-8. Histograms of the Poisson distribution with different values for the average
number of special grains (z) in a sub-sample. Corresponding concentration of Sn have
been calculated assuming that a 100-mesh grain of cassiterite contains 18.9 μg of Sn and
that there is 1 ppm Sn in the gangue.

and Switzer (1973) illustrate this by considering determination of Cr in


USGS granite G-l. The overall Cr content of G-l is about 20 ppm, roughly
half of which is present as individual chromite grains so that, on average, one
in twenty 10-mg samples of 170-mesh material contains a single grain. A
spectroscopist analyzing 10-mg samples will report 10 ppm, with good preci-
sion, and occasional wild highs of around 200 ppm. In contrast, a colorimetric
analysis on 1-g samples would indicate an overall content of 20 ppm but
precision would be poor due to the variable number of chromite grains.
These examples clearly illustrate the extent to which concentrations and
distributions of trace elements can become an artefact of laboratory
sampling procedures. This is much more likely to be a major problem in an
assay laboratory, but should also be taken into account in evaluating results
whenever a trace element is likely to be present as a rare accessory mineral.
A very thorough discussion of the problems and the benefits to be gained
from a statistical approach to them is given by Ingamells and Switzer (1973).
41

SYSTEMATIC ERRORS

Systematic analytical errors, giving rise to spurious geochemical trends,


result from contamination, drift, or physical and chemical interferences.
Between different analysts or laboratories systematic errors frequently arise
as a result of slight differences in the analysts routine or operating condi-
tions — an example of this is shown in Fig. 2-9. Only an outline of their
origins and the means of monitoring them is given here, specific causes and
their elimination being considered in greater detail at appropriate points
throughout the book.

SET IE Range 3 . 3 6 - 4 . 7 8 g/t


Mean 4.16 g/t
St. dev. 0.33 g/t
C.V. 7.9

oi
3.0 4.0
GOLD ( g / t )

12
SET I R □ nge 3 . 0 5 - 4 . 0 6 g/t
Mean 3.4 3 g/t

§
II
St. dev. 0.24 g/t
10 C.V. 7.0
9

>-
LU
3 6
7Z
σ \
a: °
////////////
4

3 -

w,fe
2

/%;
o 1 1
3 0 5 5 4 0 4.5 5.
GOLD (g/t)

Fig. 2-9. Comparison of duplicate analysis for Au by two analysts using different instru-
ments. (From Allcott and Lakin, 1975.)
42

Contamination

Perhaps the most common form of contamination arises if anomalous and


background samples are prepared and analyzed together. Sample carryover,
of metal-rich dusts during sieving and grinding or metal-rich solutions during
decomposition and analysis, results in spuriously high values for background
samples. At the instrumental measurement stage such carryover between
samples is sometimes described as memory. Contamination also arises during
preparation and analysis from the addition of extraneous material to the
sample, for example from wear of crusher plates, use of insufficiently puri-
fied reagents, or corrosion of instrument components.
It is often worthwhile to have a rough idea of the approximate extent of
contamination likely to occur. For example, in grinding 50-g samples with a
carryover of about 0.1 g in a ring mill, a sample containing 10,000 ppm of an
element will contaminate the succeeding sample, containing 100 ppm, by
about 20 ppm. Samples containing the same concentration of an element
cannot contaminate each other with respect to that element! In order to
detect the effects of contamination or carryover the sequence of sample
preparation and analysis must be known.

Drift

Drift is a source of systematic errors if, after calibration, there is some


change in the response of the analytical system to the analyte. With colorim-
etry, AAS and other techniques where calibration is simple and frequent,
this should not be a problem unless the calibration standards themselves
deteriorate. However, if photographic methods of recording signal intensity
are used, as in ES, preparation of standard films may be very infrequent and
drift, both long and short term, can occur for a variety of reasons.

Physical and chemical interferences

Physical and chemical characteristics of a sample influence the liberation


of trace elements during sample decomposition, and constituents taken into
solution along with the analyte can be a source of interference during analy-
sis. To the extent that such interferences reflect the mineralogy or bulk
chemistry of the sample an incidental effect can be to selectively exaggerate
or suppress the effects of geological (lithological) changes on geochemical
patterns.

Control and monitoring of systematic errors

Chemical interferences usually can be eliminated or minimized by meth-


ods to be described in subsequent chapters. Systematic errors arising from
43

drift or external contamination of the samples can be monitored by random


inclusion of standard control samples and reagent blanks in all analytical
batches. A typical control graph that detected a systematic drift in Cu values
is illustrated in Fig. 2-10. When samples are analyzed in the same sequence as
their collection (as is commonly the case), the presence of systematic drift
causes spurious geochemical trends that follow this sequence: a spectacular
example is shown in Fig. 2-11. This effect can be avoided by analyzing the
samples in a random order so that drift in the analytical system is ran-
domized and becomes part of the random laboratory component to the
overall data variance. The sequence of analysis must, of course, still be
known so that the presence of any sample carryover and contamination can
be recognized.
It is the author's experience that, although most analysts include control
standards, it is only in a few government and university laboratories that
incoming samples are randomized. Obviously in many circumstances, where
there is considerable pressure to complete geochemical programmes as
quickly as possible, any suggestion of delaying analysis until all samples are
received at the laboratory will be strongly contested. A more realistic
approach is, therefore, to only randomize the samples within each batch
shipped for analysis. This will not prevent systematic errors arising between
batches and if control standards indicate that this might have occurred, the
limits of each analytical batch should be indicated when results are plotted.
When the benefits of randomizing samples are considered to justify the
additional costs, the process of re-organizing a batch of sequentially num-
bered samples into a random sequence is simplified if the field envelopes are
initially numbered in a random sequence (Plant, 1973; Plant et al., 1975).
After the samples have been collected the envelopes are sorted into numeric
sequence. Plots of concentration against sample number then reveal any
trends introduced in the laboratory (Fig. 2-12). Other approaches to the

I I I I I I l_l I I I I

BATCH NUMBER
Fig. 2-10. Control graph, based on the analysis of three laboratory standards, indicating
that systematic between batch variations are present for Cu.
44

\to Woodstock
4.8 km

Fig. 2-11. Spurious Cu anomalies introduced into a soil survey by analytical errors. The
east-west orientation of the anomalies parallels the soil traverse lines and has resulted
from the analysis of the samples in the same order as their collection. (Modified from
Lockhart, 1976.)

randomized submittal of samples have been described by Howarth (1977).


Implementation of these quality control schemes is facilitated by the
increasing use of computerized data management systems.
Control samples for monitoring drift must be carefully homogenized and
available in large quantities. The steps involved in the preparation of six
exploration geochemical standards by the U.S. Geological Survey are infor-

Original values
Redetermined values

70

60

10 15 20 25 30 35 40 45 50 55 60 65 70
Sample number

Fig. 2-12. A randomized numbering scheme for geochemical sampling and analysis. The
randomized field numbers have been plotted sequentially to reveal the presence of
systematic analytical errors caused by a leaking standard. (From Plant et al., 1975.)
45

mative (Allcott and Lakin, 1975):


(1) Rock samples of about 450 kg were crushed to pass a 5-mesh sieve.
(2) Each bulk sample was then ground for 12 hours in an aluminium oxide
ceramic-lined mill with ceramic balls untill 97% of every sample would pass a
325-mesh (44 μπι) screen. During grinding frictional heating to 80—90°C
may have caused partial loss of volatile elements and contamination by
alumina resulted from wear of the ball mill.
(3) Bulk samples were then divided between 45-kg drums and 80-g bottles
filled from each drum, each bottle being assigned a random number. To test
sample homogeneity an analysis of variance model, described in detail by
Allcott and Lakin (1975), was used to establish variance between drums,
within layers within drums, between bottles within layers and between
determinations within drums.
In accepting and using a standard it should be remembered that, as
described on p. 40, apparent concentrations and homogeneity can be
artefacts of sub-sample size. As a Cr standard G-l, with a Cr content likely to
be reported anywhere between 7 and 200 ppm depending on sub-sample
size, is obviously not very satisfactory.

Accuracy

Although absolute accuracy is less important in analysis of exploration


geochemical samples than in most other forms of analysis, it would be
inappropriate to conclude this discussion without considering it briefly. The
true content of a constituent in a sample is probably impossible to determine
and from a practical viewpoint adequate accuracy can be considered to have
been achieved when, taking into account their reproducibility and the repro-
ducibility required of the analysis, different analytical methods give essen-
tially identical results. Simple statistical tests such as Student's t can be used
to quantify the comparison of two sets of results. In planning such com-
parisons the effects of sub-sample size, as previously discussed for Cr in G-l,
used in different methods of analysis must be considered carefully.
The wide discrepancies reported for trace element determinations in com-
pilations for standard rocks (e.g. Flanagan, 1969), for lunar samples
(Table 2-1V) and in inter-laboratory assay checks (Lister and Gallagher,
1970), all confirm the difficulty of obtaining accurate trace element deter-
minations for geological materials. Under these circumstances the problems
of assigning accepted or recommended values by consensus is only too
apparent. Nevertheless, despite their inherent problems, carefully prepared
reference materials are invaluable for comparison of analytical methods and
in developing new procedures. Each laboratory should therefore have as
many as possible. Abbey (1977) has compiled usable values for geological
standards and Geostandards Newsletter (from Geostandards, 15, rue Notre-
Dame-des-Pauvres, Case Officielle No. 1, 54500 Vandoeuvre-les Nancy,
46

TABLE 2-IV
Relative deviations between laboratory determinations of trace elements in lunar samples
(based on data in Morrison, 1971)

Concentration Number of Average relative Range of relative


range elements standard deviation standard deviation
(ppm) (%) (%)
1-10 15 21 7.9- -61
10-50 13 14 6.7- -31
>50 10 15 3.4- -35

France) regularly contains information on new reference standards. Results


obtained on the six exploration geochemical standards prepared by the U.S.
Geological Survey are summarized by Allcott and Lakin (1978).
Most certified standards are only available in small quantities and only the
values for their total trace element contents are certified — publication of
provisional values for ascorbic acid/hydrogen peroxide soluble Cu, Co and Ni
in the Canadian ultramafic rocks UM-1, UM-2 and UM-4 is a notable excep-
tion (Cameron, 1972; Faye, 1975). Because of their scarcity primary stan-
dards are best used to calibrate secondary standards of similar bulk composi-
tion.

A QUALITY CONTROL PROGRAMME

Every geochemical laboratory should have a routine procedure for evalua-


tion of precision and detection of drift or bias between batches. When sam-
ples are submitted to external laboratories duplicates and standard samples
can be inserted in each batch to provide an independent check on their per-
formance. No reputable laboratory or analyst should object to this pro-
viding it is done in a responsible fashion and any problems detected are dis-
cussed with them. The records provided by such a programme are partic-
ularly valuable if it becomes necessary to change laboratories (Hill, 1975).
The following minimum scheme is proposed for a laboratory in which
samples are analyzed in batches of 100:
(1) Duplicate, preferably random, analysis of approximately 10% of all sam-
ples in different batches to the original analysis. Results are plotted as Ι-ΧΊ —
X21 versus (A^ + X2)/2 control graphs (Fig. 2-4).
(2) Inclusion of at least one reagent blank and two control standards per
batch. Control standards should be coded and not recognizable by the
analyst. Results are plotted as in Fig. 2-10.
This scheme provides a basic quality control programme. A more compre-
hensive scheme, including randomized submission of samples and com-
puterized evaluation of the quality control programme, has been described
by Plant (1973) and Plant et al. (1975).
Chapter 3

SAMPLE PREPARATION

INTRODUCTION

Routine preparation of sediments and soils for analysis involves drying


and sieving, and for rocks size reduction by crushing and grinding. Step-by-
step procedures are described by Lavergne (1965). Water samples present
quite different problems in that the main concern is to ensure that their
composition is not too altered by the changes which begin from the moment
a sample is collected.
The entire field sample is seldom used for the final analysis and an impor-
tant part of sample preparation is sub-sampling to obtain a representative
portion of the particular fraction chosen for analysis. Because of the exten-
sive handling that samples usually receive during preparation, contamination
must continually be guarded against. A summary of potential contaminants
from common laboratory materials appears in Table 3-1.

TABLE 3-1

Potential contaminants from laboratory materials

Material Potential contaminants

Grinding equipment
Steel and iron grinding plates Fe, Co, Cr, Cu, Mo, Mn, Ni, V
Alumina ceramic plates l Al, Cu, Fe, Ga, Li, Ti, B, Ba, Co, Mn, Zn, Zr
Tungsten carbide Co, Ti, W
Lubricants Mo
Packaging materials
Polythene Ti, Ba, Zn, Cd
Polypropylene Ti
PVC Ti, Zn, Na, Cd
Brown paper Si
Rubber Zn
1
Thompson and Bankston (1970).
2
Scott and Ure (1972).
48

PREPARATION OF ROCK SAMPLES

Reduction of rock fragments to powder involves crushing in a laboratory


jaw crusher, which should be capable of handling about 50 kg per hour,
followed by grinding in a disc mill, swing mill or ball mill (Fig. 3-1). Disc
mills, in which the sample is fed between one rotating and one stationary
vertical plate, are able to handle relatively large volumes of material and
samples weighing several kilograms can be reduced rapidly to minus
100 mesh, which is sufficiently fine for many purposes. Disc mills are not,
however, very effective for size reduction much below 100 mesh and if a
very fine powder is needed, further reduction must continue in a ball or
swing mill. The author has found swing mills, of the type manufactured by
Angstrom * or Rocklabs **, to be very effective. Loss of fine powder is pre-
vented by the tightly fitted lid of the grinding vessel. A simple ball mill
shaker capable of grinding several samples simultaneously is readily con-
structed from a paint shaker (Myers and Wood, 1960; Lavergne, 1965).
Laboratory swing mills and ball mills have the disadvantage of only being
able to handle relatively small quantities (20—200 g) of material and sample
bulk must sometimes be reduced accordingly by coning and quartering or
splitting.
Laboratory crushing and grinding equipment is made from a variety of
materials (Table 3-1) and wear causing contamination of samples is inevitable.
However, this is probably less of a problem than the contamination resulting
from sample carryover, particularly if anomalous and background samples
must be processed with the same equipment: preparation of assay samples
on the same equipment is, of course, to be discouraged! The health hazard
from silicate dusts and avoidance of cross-contamination requires a reasona-
bly clean, well ventilated work area. If possible, dust escaping from heavy
equipment, such as jaw crushers and disc mills, should be extracted down-
wards and to the rear along conduits in the equipment pedestal. Fine mate-
rial lodging in equipment can be dislodged with blasts of compressed air.
Apart from contamination, comminution of the sample can cause loss of
volatiles and oxidation of ferrous to ferric iron. Behavior of Hg is of partic-
ular interest to exploration geochemists. Watling et al. (1973) and Koksoy
et al. (1967), using a modified Tema swing mill and a ceramic ball mill,
respectively, found losses of Hg during grinding to be insignificant. However,
in both cases increased grinding increased the amount of Hg released subse-
quently by pyrolysis.
Usually the sample powder taken for analysis will be intended to repre-
sent the whole of the original rock sample. Sometimes, however, analysis of

* Angstrom, Inc., Belleville, MI 48111, U.S.A.


** Rocklabs, Auckland, New Zealand.
49

Fig. 3-1. Rock crushing equipment: in sequence from the foreground are (1) a disc mill
open for cleaning; (2) a jaw crusher; (3) a closed disc mill; and (4) a second jaw crusher.
The equipment is mounted on a hollow pedestal constructed by covering a heavy welded
steel frame with sheet metal: within the pedestal ducts, connecting each piece of equip-
ment to an air exhaust system, extract any escaping dust downwards. The compressor in
the background provides air for blowing equipment clean.

a particular mineral fraction might be needed and it will be necessary to


prepare a mineral separate. For this purpose the rock should only be ground
sufficiently to release monomineralic grains without excessive production of
fine material. The next step is to screen the sample through nested sieves so
that separations can be made on uniformly sized fractions — minus 40/plus
60-, minus 60/plus 80-, and minus 80/plus 100-mesh material is usually
easiest to process. Mineral fractions with different physical properties can
then be separated using heavy liquids (for example, bromoform, specific
gravity 2.89; or methylene iodide, specific gravity 3.31), vibrating tables to
separate platy minerals, or magnetic susceptibility. Rotating-drum magnetic
separators, such as the Carpco *, are able to handle several kilograms of
material in a few minutes and are more satisfactory for handling large
samples than the more familiar Frantz magnetic separators. Both bromo-
form and methylene iodide are toxic and should only be used in a well
ventilated hood. They are also expensive, but can be recovered from acetone

* Carpco, Jacksonville, Florida.


50

washings of the mineral separates by vigorous shaking with an excess of


water.

Soils and sediments

Preparation of soils and sediments involves drying, then disaggregation and


sieving to separate the size fractions chosen for analysis. An 80-mesh
(177 Mm) screen is usually used because the minus 80-mesh fraction is com-
monly found to give good anomaly contrast and is sufficiently abundant in
most soils and sediments that obtaining the few grams needed for analysis
presents no difficulty. It is also sufficiently fine that additional size reduc-
tion by grinding seldom produces a worthwhile reduction in sub-sampling
errors (Fig. 3-2). These attributes should never, however, be taken for
granted and there is no substitute for a detailed orientation survey on the
influence of particle size, sieving and grinding on anomaly contrast. The
study by Plant (1971) of metal dispersion in Scottish streams is a good
example.

100°/

T3
e
o
J-l
be
Ö
3

Ü- 50

50 100%
P (ground)
Fig. 3-2. Comparison of analytical precision (P%) for trace element concentrations in
ground and unground minus 80-mesh stream sediments. Open symbols: no significant
difference, at the 95% confidence level, between the two populations on the basis of the
Kolmogorov-Smirnov test; solid symbols indicate that populations are significantly
different. (From Howarth and Lowenstein, 1971.)
51

Soils and sediments should be collected in high wet strength paper bags in
which they can be dried without opening the packet. In hot, non-humid
climates it is sufficient to place the bags outside in the sun. Where this is not
practical a simple drying tent can be improvised with four uprights, plastic
sheeting and a kerosene heater. Electric drying cabinets are used in the
laboratory. To avoid baking clay-rich samples, which makes subsequent dis-
aggregation difficult, temperatures should be no greater than 65° C. Koksoy
et al. (1967) reported losses of up to 42% of the Hg content of soils dried at
80° versus 20° C, and other volatiles, such as Se, could probably also be lost.
Contamination of samples which originally had low Hg contents was ob-
served within thirty days of their storage in the same container as an
anomalous sample.
Once dry the sample can be disaggregated and sieved. The object of dis-
aggregation is not to grind the material but to break down aggregates and
clumps to obtain the natural grain size distribution. This is best achieved by
a light pounding with a porcelain pestle and mortar. The sample is then
screened on a silk or nylon sieve mounted in a cardboad or plastic frame —
4|" (11.4 cm) diameter PVC drainpipe is ideal. Stainless steel sieves can also
be used but are expensive and must be treated with caution because silver-
solder is sometimes used in their construction: only sieves with a friction-
flange fitting should be used. Between samples, equipment can be cleaned
with a stiff paint brush or compressed air. Normally only the fines are
retained for analysis. Hoffman (1974), however, has suggested collection of a
coarse fraction (>2 mm) on bitumen coated pebble cards as a guide to sedi-
ment lithologies and provenance.
Mesh sieve sizes are summarized in Table 3-II. In some systems of size
classification the silt-sand boundary is arbitrarily placed at 20 μηι whereas in
others it is at 50 μιη. The latter, at 270 mesh, is about the lower limit of
rapid dry screening and separation of finer fractions requires methods based
on their sedimentation rates. Because dry disaggregation and sieving are not
very efficient, estimates of size fraction distributions so obtained will seldom
agree closely with results from more reliable wet dispersion methods and
wet-sieving or sedimentation.
Detailed schemes for separation of mineral fractions from overburden or
sediment are beyond the scope of this text. However, general remarks on this
topic in the preceeding section apply equally here. The value of panning bulk
samples to obtain reasonably representative sub-samples of rare grains of
heavy minerals, such as cassiterite, was noted in Chapter 2. The scheme used
by the Geological Survey of Sweden for processing tills is summarized in
Fig. 3-3.
Separation of mineral fractions requires considerable effort and the little
additional time required for a rapid visual examination under a binocular
microscope can be very rewarding. In Zambia, the author was able to trace
the sub-outcrop of weathered zinc mineralization by washing the plus 2-mm
52

TABLE 3-II
Approximate equivalent mesh sizes in the British Standard (B.S.), American Society for
Testing Material (ASTM) and Tyler sieve series

Approximate aperture B.S. ASTM Tyler


(Mm)

2000 8 10 9
1000 16 18 16
850 18 20 20
420 36 40 35
355 44 45 42
300 52 50 48
250 60 60 60
212 72 70 65
180 85 80 80
150 100 100 100
125 120 120 115
90 170 170 170
75 200 200 200
63 240 230 250
53 300 270 270
45 350 325 325
38 — 400 400

(10-mesh) fraction of overburden drill samples and staining with ferricyanide/


oxalic acid/diethylaniline solution (Reid, 1969) to identify the presence of
secondary zinc minerals.

Vegetation

Vegetation samples are best collected in brown paper grocery bags. They
should not be packed tightly or placed in closed plastic bags because this will
cause rotting and ultimately produce silage! If the material is to be ashed it
should be remembered that ash content usually amounts to only 3—10% of
the dry weight and an appropriate amount of material collected.
Samples can be dried by hanging the bags in a warm dry room or by using
an oven with forced ventilation. Temperatures above 40° C should not be
used. Dried material can then be ground in a hammer mill equipped with
steel blades and a non-contaminating screen through which the fine, ground
material falls.

Waters

Natural waters may have to be analyzed for trace metals and pH, and less
often for anions (particularly Cl~, SO*", HCOJ, CO3" and F"), temperature,
53

Sampling

Glacial till
| 10 liters

S ieving 5 m m ^ Material >5mm Waste

Material < 5mm

Suction dredge _
and sluice box —H Overflow Waste

Retained on
sluice box

4
Drying

i
Sieving 0 5 mm

[Material <0.5 mm]

Separation with
tetrabromoethane (296 g/cm 3 ) HMatenal<2.96 Waste

I Material >2.96|

Separation with
hand magnet

H Weakly and non-


magnetic material

Weakly and non - Separation with Weakly and non-


magnetic material - dimethyliodide magnetic material
>3 31 ( 3.31 g / c m 3 ) > 2 9 6 < 3 31

Separation with Separation with


Franz magnet 07 A Franz magnet 13 A

i
Non-magnetic Weakly magnetic Weakly magnetic
i
Non-magnetic
Magnetite material material material magnetic
material >0.5mm < 5 m n
>3.31 >331 >2.96<3.31 >2.96<3.31
Sub- Kept for further
fraction 1 investigations

Y
Chemical and mineralogical analysis

Fig. 3-3. Flow chart of the stages in heavy mineral separation and processing used by the
Geological Survey of Sweden. (From Brundin and Bergström, 1977.)
The average weights of the subfractions according to this flow sheet are:

Sub fraction Average weight from


3872 samples (g)

1 (d = 2.96—3.31 g/cm 3 ; isodynamic separation at 1.3 A) 9.0


2 (d = 2.96—3.31 g/cm 3 ; "non-magnetic" residue) 8.5
3 (d 3.31 g/cm 3 ; isodynamic separation at 0.7 A) 3.8
4 (d 3.31 g/cm 3 ; "non-magnetic" residue) 9.4
5 (magnetite) 9.5
54

conductivity and dissolved oxygen. pH and the anions are usually best deter-
mined as soon as possible at the sample site or shortly after collection. How-
ever, unless colorimetric field procedures are used, determination of trace
metals will usually require storage and shipment of waters to a laboratory.
During storage the very low concentrations of trace metals found in natural
waters are susceptible to changes resulting from bacterial activity, from con-
tamination by their container, and from losses due to adsorption of dis-
solved metals on suspended sediments, on precipitates and on the container
walls.
For field use polyethylene or polypropylene bottles are preferred to glass.
However, several important trace metals are incorporated in plastics during
their manufacture (Table 3-1) and are often present in a leachable form
(Scott and Ure, 1972). Zinc is especially troublesome. To prevent con-
tamination a thorough cleaning of bottles (and caps), first by protracted
soaking in 50% hydrochloric or nitric acids and then by rinsing with distilled
water and aliquots of the water to be collected, is recommended. Marchant
and Klopper (1978) reported that the levels of contamination found after
storing ultrapure water for four years in acid-washed polyethylene, were too
low to be of concern in exploration geochemistry.
Losses by adsorption on container walls have been studied by many
workers and a variety of special procedures have been suggested for individual
elements. Acidification to pH 1 with hydrochloric acid and addition of
50 mg/1 Br is effective for Au (Chao et al., 1968) and preservation by addi-
tion of an acidified solution of potassium permanganate (Jonasson et al.,

Iron
00

/
Acidified

00 Γ

+ / +
+
10 +^ +
+
- +
+ + +-H-
-y
1n ■ i » ■ ....I L _ J _ i, iml , i , , MM
1.0 10 100 1000
Unfiltered Fe m g / l

Fig. 3-4. Influence of acidification on Fe content of filtered and unfiltered waters:


organic carbon, Li, Ba, As, B, Mg, Ra, Mn, Pb, Ca, K, Al and U behave similarly. (From
Wenrich-Verbeek, 1977.)
55

1973) or dichromate (Feldman, 1974) prevents loss of Hg. The only


preservation procedure of general applicability for several trace elements is
acidification to ^ p H 1 with nitric or hydrochloric acid which in practice
means addition of several millilitres of concentrated acid per litre of sample.
However, unless the sample is first filtered, the increased acidity causes
leaching of elements from suspended sediment thereby increasing apparent
dissolved metal content (Fig. 3-4). (In waters with a high content of dis-
solved organics, polymerization and precipitation of organic material at low
pH values can actually cause losses of metals during acidification; Sholkovitz,
1976.) Ideally, therefore, filtration soon after collection should precede
acidification. Unfortunately filtration, particularly through the 0.45-μηι
filters used in standard water analysis, can be time consuming and unsuited
to rapid surveys even if a portable vacuum or pressure filtration unit
(Herbert and Young, 1977) is available. Consequently, bearing in mind that
the turbidity of lakes and streams varies with time, the relative merits of
collecting unfiltered-acidified, unfiltered-unacidified, and filtered-acidified
samples should be carefully considered.
Wenrich-Verbeek (1976, 1977) has discussed the problem of preservation
of water samples in relation to exploration for U: storage without filtration,
and acidification without filtration both increased dissolved U concentra-
tions. Field-filtration and acidification were therefore recommended. How-
ever, in a study of Canadian lakes and streams, Hall (1979) found that dis-
solved U remained stable for at least five months without addition of preser-
vatives. Filtration had no effect on dissolved U concentrations. The reasons
for the differences between the two studies are not known although it seems
likely that they result from differences in the suspended sediment load.
Parslow and Dwairi (1977) avoided the potential problems associated with
water samples by concentrating U on an ion exchange resin loaded into a tea-
bag. The teabag was left overnight in the water sample and then transferred
to a smaller bottle for shipment to the laboratory where U was recovered by
elution.
Chapter 4

SAMPLE DECOMPOSITION-SOLUTION TECHNIQUES

INTRODUCTION
During decomposition the sample is opened up or decomposed releasing
the elements to be determined. This can be brought about by digestion and
leaching with concentrated or dilute acids; fusion with acid or alkaline
fluxes; liberation of volatile constituents by pyrolysis; and by partial or
selective extraction of constituents with a wide variety of reagents. Strictly,
the almost simultaneous decomposition and excitation of the sample in DC-
ES should also be considered as sample decomposition. Here, however, only
those procedures producing solutions, suitable for analysis by AAS, ICP-ES
or colorimetry, will be considered.
Although decomposition introduces an additional step between sample
preparation and the final analysis, it provides the geochemist with consider-
able freedom to liberate and determine either all or only a particular fraction
of the trace constituents. As discussed in Chapter 1, selection of the appropri-
ate decomposition technique can be a major factor in enhancing anomaly
contrast. Choice of a method will depend largely on the dispersion character-
istics of the element to be determined and on its mode of occurrence within
the sample. Other factors to be considered include: (1) the final method of
analysis and the need to avoid interferences; (2) the desirability of deter-
mining several elements in the same solution; (3) ease of handling, rapidity
and adaptability to processing large numbers of samples; and (4) availability
and costs of reagents and equipment.
The decomposition procedures to be described are summarized in Table
4-1. They fall into two broad groups: strong decompositions capable of
releasing a large proportion of the trace constituents from mineral lattices,
and weak or partial decompositions intended to remove only weakly bonded
elements or those associated with a particular fraction of the sample. Strong
and weak decompositions are often referred to casually as total and cold
extractions, respectively. Neither term can be recommended because the
former are seldom true total attacks (i.e. attacks capable of releasing all of
the metal inside and outside the lattice in whatever form (Lapointe, 1968),
and the efficiency of the latter is very susceptible to even small changes of
ambient temperatures (Ellis et al., 1967).
58

TABLES 4-1
Classification of some decomposition techniques useful in exploration geochemistry

Decomposition Reagents

Strong decompositions
(1) Digestions with hot, usually concen- H N 0 3 , HCl, HCIO4, HF
trated, mineral acids
(2) Fusions:
acid fusions KHSO4, K 2 S 2 0 7
ammonium halide sublimations NH4I, NH4CI
alkaline fusions Na 2 C0 3 , NaOH, L1BO3
oxidative-alkali fusions Na 2 C0 3 or NaOH with ΚΝΟ3 or N a 2 0 2

Partial decompositions
(1) Non-selective decompositions:
cold dilute mineral acids: e.g. 0.1-1.0 N HCl
buffers: e.g. NH 4 -citrate/NH 2 OH · HCl, pH 2-
chelating agents: e.g. 0.05-0.25 M EDTA, pH 4 - 7
(1) Selective decompositions:
(a) removal of exchangeable metals NH 4 -acetate; MgCl2
(b) removal of organic matter H 2 0 2 ;NaOCl
(c) secondary iron and manganese Na-dithionite;
oxides hydrazine;
hydroxylamine hydrochloride;
NH4-oxalate
(d) sulphides KCIO3/HCI; ascorbic acid/H 2 0 2 ;
bromine

STRONG DECOMPOSITIONS

Strong decompositions usually involve either digestion with concentrated


acids or fusion with fluxes. The acid is then diluted with water or the residue
leached with a dilute acid. Where possible acid digestions are generally to be
preferred because they introduce less extraneous material into the final solu-
tion, thereby reducing chances of contamination and interferences, and are
more suited to handling of large numbers of samples. When the final deter-
mination is to be by AAS, fusions are particularly undesirable because the
high dissolved solids content of the fusion leachate can cause rapid clogging
of the burner slot by deposition of evaporated salts. The problem is even
more severe with ICP-ES nebulizers. However, under field conditions where
acids may be hazardous, fusion provides a convenient alternative for many
elements.
59

Acid digestion

Physical and chemical properties of the important acids are summarized


in Table 4-II. Mixtures of nitric acid with either hydrochloric or perchloric
acids are probably most widely used in routine work. However, procedural
details with respect to temperature and length of the extraction period vary
greatly between laboratories. Except for mixtures including hydrofluoric
acid, all the digestions are readily carried out by heating the sample (0.2—
0.5 g) with a few millilitres of acid in a hard glass test tube. Air baths, water
baths, sand baths and drilled aluminium blocks (Fig. 4-1) all make suitable
receptacles for test tubes and provide reasonably uniform heating for decom-
position of large batches of samples on a single hot plate.
Depending on the heating arrangement, samples can either be allowed to
reflux for a few hours or be evaporated to dryness. The latter gives the most
complete extraction, but has the disadvantage that the residue must then be
leached with a dilute acid (usually hydrochloric acid) whereas the refluxed
acids can simply be diluted to volume with water, mixed and any residue
allowed to settle. In a recent innovation Lovell and Hale (1980) have over-
come this problem and increased the efficiency of extraction with a mixture
(5 : 1) of nitric and hydrochloric acids, by decomposing the sample in a pres-
sure leach tube with a crimped-on teflon seal. An additional advantage is that
volatile elements are retained.
Foster (1971, 1973) has compared the efficiency of various acid diges-
tions on some common minerals and igneous rock. His results show consider-
able variability in the susceptibility of silicate lattices to decomposition
(Figs. 4-2 and 4-3). With mixtures of concentrated hydrochloric, nitric and
perchloric acids extraction of Co, Cu, Mn, Ni and Zn was greater than 70%
for olivine, biotite and limonite compared to total decomposition with

TABLE 4-II
Physical and chemical properties of mineral acids used in sample decomposition

Property Acid

HN03 HC1 HCIO4 HF

Molecular weight: 63.01 36.46 100.46 20.01


AR grade (%) 1 70 36 60 or 72 40 or 48
Specific gravity 1.42 1.18 1.54 or 1.70 1.13 or 1.15
Boiling point (°C) 120.5 110 203 120
Azeotrope (%) 68 20 72.5 35
2
Cost per litre (C.$) $ 15.00 $ 13.00 $ 40.00 $ 30.00
1
Percent acid (w/w) in analytical reagent grade.
2
Reagent grade.
60

Fig. 4-1. Sample decomposition equipment. The hot air bath (background), drilled alumi-
nium block (left) and sand bath (right) are all convenient ways of decomposing large
numbers of samples on a single hot plate.

hydrofluoric acid. Lower recoveries were obtained from some pyroxenes and
amphiboles and from most of the rocks studied. Grain size was also found to
influence extraction efficiency.
A similar study of the release of Cr, Co, Cu, Fe, Mn, Ni, V, Zn and Pb
from the products of lateritic weathering, after decomposition with hydro-
fluoric acid, nitric and perchloric acid, and perchloric acid alone, was
reported by Gedeon et al. (1977). Sample mineralogy was examined by
X-ray diffraction, both before and after decomposition, and extraction effi-
ciency compared to total concentrations estimated by XRF. Hydrofluoric
acid left little or no residue and was satisfactory for all elements except Cr
and V. In contrast, residues from the nitric-perchloric decomposition con-
tained quartz, talc, muscovite, feldspars, amphiboles and kaolinite;
depending on the type of material from 50 to 100% of the trace metals were
extracted. Perchloric acid used alone was least satisfactory with regard to
both extraction efficiency and reproducibility.
Clearly, even with strong acid decompositions extraction of lattice-bound
trace elements, from both fresh silicates and the residues of intensive
weathering, is often far from complete. The resulting geochemical patterns
will therefore reflect, in part, mineralogical and lithological differences
among samples rather than absolute differences in their trace element
61

120 -

100 - t^r/o »\ - / " Feldspar 9


Vi \\Ä// ° A m P h i b o ' e 260
\ Υ \ ♦ Biotite 265
80 - \* * \ \ · Pyroxene 105
\ η \ \ Δ Limonite 285
ω
Ο 60-
<

^ \ \

Vu
I-
X
LU
40-

20- ^^^>Τ*
* Ε
α
η _
α
CO
^ ι^ ° <"> o o
ζ ζ z
ο
X ° ^ I z
g
o o
* Ö

Fig. 4-2. Variation in the extraction of Zn from some common rock-forming minerals
with acid decompositions. (From Foster, 1971.)

100-,

>■ 80H Rhyodacite 168


Diabase 140
60 Peridotite 96
O
< ♦ Diopside
cc Gneiss 64
X
LU

E
aa
co i- co Ü
O

0
<N

CO
CO

O
O
C ix P z
X
O z z o
z 1 X •5 <
o
o X
o X

Fig. 4-3. Variation in the extraction of Zn from dacite, diopside gneiss, diabase and
peridotite with acid decompositions. (From Foster, 1973.)
62

content. As noted in Chapter 1 (p. 9) this is often more of an advantage


than disadvantage. Dolezal et al. (1968) is invaluable as a source of informa-
tion on the action of acids and acid mixtures on specific minerals.

Nitric acid
Nitric acid, dilute or concentrated, is a strong oxidizing agent and an
excellent solvent for most sulphides, except cinnabar, which are oxidized to
sulphates. Lead can be subsequently lost by precipitation of insoluble
lead sulphate. Phosphates and most primary and secondary uranium minerals
are also readily decomposed. Silicates are only partly attacked and organic
matter is not completely destroyed. Organic-rich samples should therefore be
ignited prior to treatment with nitric acid.
Despite incomplete breakdown of organic matter, boiling concentrated
nitric acid in a 5 : 1 (v/w) ratio has been recommended as a general decom-
position procedure for rocks and soils (Ward et al., 1969). The resulting solu-
tion, after dilution, is suitable for the determination of Bi, Cd, Pb, Cu, Zn,
Co and Ni by AAS. Reasons given for the choice of nitric acid, in preference
to mixtures of nitric-perchloric or nitric-hydrochloric acids, were greater
safety and reduced nebulizer corrosion, respectively. A hot dilute nitric acid
leach (4 M at 90° C) has been used to extract Cu and Zn from lake sediments
(Timperley and Allen, 1974); the solution is also suitable for the fluorimetric
determination of U (Smith and Lynch, 1969).

Nitric acid/hydrochloric acid


Because of its ability to dissolve Au, the 3 : 1 mixture of hydrochloric
and nitric acids is known as aqua regia. The active agents are nitrosyl chlo-
ride and chlorine:

3 HC1 + H N 0 3 -* 2 H 2 0 + NOC1 + Cl 2

As well as dissolving the noble metals its oxidizing properties make it a


powerful solvent for sulphides. Iron oxides, apatite, uranium oxides and
oxygen containing molybdenum minerals (wulfenite, molybdite and
powellite) are also readily decomposed. Silicates are only partly decomposed
and aqua regia has therefore been suggested as a sulphide-selective attack for
bedrock geochemistry (Warren and Delavault, 1959a, b; Olade and Fletcher,
1974). Aqua regia is unsuited to decomposition of organic-rich sample unless
they are first ignited.
Decomposition with aqua regia is often used in the determination of Au.
In the method described by Tindall (1965) a 10-g sample is first treated with
hydrochloric acid (25 ml). Nitric acid (15 ml) is then added and the solu-
tion evaporated to dryness and the residue baked before taking the salts into
solution with 50 ml of 50% hydrochloric acid. Initial addition of hydro-
chloric acid alone ensures release of sulphur from sulphides as hydrogen
63

sulphide, thereby precluding precipitation of sulphates. Baking of the residue,


until the odour of nitric acid is no longer discernible, is required to minimize
formation of insoluble nitrates and to ensure their subsequent decomposi-
tion.
Lynch et al. (1973) have described a procedure for determination of both
trace element and organic carbon content of lake sediments. The sediments
are leached with a mixture of 4 M HNO 3 /0.1 M HC1 at 90° C and trace ele-
ments determined by AAS. Organic carbon content is estimated by mea-
suring optical density of the solution at 500 nm. The method was used by
Garrett and Hornbrook (1976) to study the relationship between Zn and
organic content of sediments.

Perchloric acid, and mixtures of nitric and perchloric acid


Hot perchloric acid is a powerful oxidizing and dehydrating agent, and all
common perchlorates, except potassium perchlorate, are readily soluble. Its
main disadvantage is the danger of an explosive reaction with organic mate-
rial. Mixtures with an excess of nitric acid, which is evaporated off as the
decomposition proceeds, are therefore usually used to moderate the initial
reaction with readily oxidizable compounds. Perchloric acid alone is, how-
ever, used in some laboratories (Bradshaw et al., 1974; Stanton, 1970a).
Ideally, it should only be used in a fume hood constructed from non-reactive
impermeable materials, and equipped with a venturi-exhaust and built-in
wash down system to prevent accumulation of perchlorate residues. Other
precautions to be taken are described by the Analytical Methods Committee
(1959).
Mixed nitric acid/perchloric acid is excellent for decomposition of soils
and sediments and is also used for wet ashing of vegetation in biogeochem-
istry. Silicates are partly decomposed leaving a residue of dehydrated silica
and insoluble potassium perchlorate; organic matter and sulphides are
oxidized. Decomposition is most complete when taken to dryness and, after
taking the residue up in dilute acid, the resulting solution is suitable for
determination of many elements (Stanton, 1966). However, prolonged
evaporation results in loss of As and Se which, because cold perchloric acid is
not an oxidizing agent, could otherwise be determined as their hydrides
(pp. 132 and 165). Perchloric acid digestions can, therefore, only be used for
generation of arsine and other hydrides providing evaporation is sufficient to
remove all traces of nitric acid but not so prolonged as to cause analyte
losses by volatilization. Charring of organic materials during digestion can
also cause loss of Se and As but can be prevented by further additions of
nitric acid (Gorsuch, 1970).
In the determination of Mo, Stanton (1970a) found that digestion of
inorganic samples with perchloric acid was more rapid and gave comparable
results to pyrosulphate fusion. Subsequently, because of low recoveries of
Mo from organic-rich samples with pyrosulphate fusion, digestion with
64

nitric acid/perchloric acid was recommended (Stanton et al., 1973). To avoid


precipitation of potassium perchlorate, sodium iodide rather than potassium
iodide must be used to suppress copper interferences when Mo is estimated
as its dithiol complex on perchloric acid leachates (Stanton, 1970a; Hoffman
and Waskett-Myers, 1974). Low recoveries for Pb in perchloric acid diges-
tions of Ba-rich materials have been reported (Peachey and Vickers, 1978).

Hydrofluoric acid
Hydrofluoric acid is unique in its ability to decompose silicates, Si being
lost as the volatile tetrafluoride. All but the most refractory silicates are dis-
solved and hydrofluoric acid, either alone or with other mineral acids, has
therefore found wide use in many analytical schemes for determination of
major constituents of rocks (Maxwell, 1968; Langmhyr and Paus, 1968).
However, in the exploration laboratory, where rigorous estimation of
absolute metal concentrations is seldom needed, the precautions necessary
in handling this acid make its general use inconvenient except when total
concentrations are required or it is necessary to achieve complete decomposi-
tion of a particular mineral. At temperature below 120°C inexpensive
plastic beakers or bottles, heated on an air- or water-bath, make satisfactory
decomposition vessels. Polytetrafluorethylene (teflon) can be used up to
250° C but at higher temperatures platinum ware is needed.
Hydrofluoric acid is sometimes used both to decompose the sample and
complex the analyte. For example, B can be determined colorimetrically as
its fluorborate complex (BF4) after the sample has been decomposed with
hydrofluoric and sulphuric acids (Stanton and McDonald, 1966). Similarly,
in the estimation of Nb, pyrochlore and other Nb-bearing minerals are
decomposed with hydrofluoric acid; the soluble niobium fluoride complex is
then isolated and estimated by paper chromatography (Hunt et al., 1955).

Other acid decompositions


In determination of Hg by cold vapour AAS (p. 134) it is necessary to
bring Hg into solution in a form that can be reduced to elemental Hg (Hatch
and Ott, 1968; Ure and Shand, 1974). The hot acid decomposition proce-
dures already described can result in appreciable losses of Hg by volatiliza-
tion. Satisfactory recovery can be obtained by digestion with mixtures of
nitric acid/sulphuric acid/potassium permanganate at temperatures between
60 and 90°C (Barakso and Tornocai, 1970; Huffman et al., 1972; Ure and
Shand, 1974).
Another specialized procedure is the dissolution of Au by hydrobromic
acid containing free bromine. In a rapid method described by Thompson et
al. (1968), 10 g of ignited pulverized sample is shaken for 15 minutes with
20 ml hydrobromic acid containing 0.1 ml Br. After centrifuging and
washing with 0.1 N hydrobromic acid, the aurobromate complex is extracted
65

into methyl isobutylketone (MIBK) and Au estimated by AAS. Results are


comparable to those obtained by fire assay and the procedure is simpler and
more rapid than digestion with aqua regia. Pd, Pt and Te can also be deter-
mined (Thompson, 1967; Nakagawa and Thompson, 1967; Ward et al.,
1969; Stanton, 1975, 1976).
An unusual decomposition for F is based on formation of the soluble
beryllium fluoride complex when fluorite dissolves in an acidified solution of
beryllium nitrate (Pliiger and Friedrich, 1973). Dissolved F is then measured
with a specific ion electrode; results are lower than those obtained with
sodium hydroxide or potassium nitrate fusions (Fig. 4-4). Solutions of
aluminium chloride also dissolve fluorite (Dolezal et al., 1968) and were used
by Carlson and Manus (1979), prospecting for fluorite, in a field procedure

900

■ 10 N HCI soluble F"


* 10~ 3 N NaOH soluble F"
o TISAB soluble F"

E
aa

z
o
o
LJJ
Q * Smelting with NaOH
CO

o ■ Sintering with K N 0 3 - N a 2 C 0 3
o Complexing with Be(N0 3 ) 2

5000

+" ~+ + + + φΜ + + + + + + + + +~^r-v
+ V + % Vw £ ^ + V ^ Granite + V + V V l
+ + + + ( J ^ + + + + + + + + + + + -(■

Fig. 4-4. Fluoride values in soils estimated by six different extractions. (From Pliiger and
Friedrich, 1973.)
66

for analysis of stream sediments. Between 1 and 5 g of sediment was


extracted with 20 ml of a solution containing 241.44 g A1C13 · 6 H 2 0 per
litre. A sodium acetate buffer controls pH and addition of sodium citrate
prevents aluminium interference in estimation of F with the specific ion
electrode (p. 208).

Fusions

Fusion or sintering with a flux is an effective method of decomposing


many minerals. Acid fusions with either potassium pyrosulphate or bisul-
phate, and alkaline fusions, with or without an oxidizing agent, are partic-
ularly useful in exploration geochemistry. Fire assay for Au is an example of
a reductive fusion and fusion with ammonium iodide is used to separate Sb
and Sn, which form volatile iodides, from the bulk of the sample.

Fusion with potassium bisulphate and pyrosulphate


Acid fusions with potassium pyrosulphate (K 2 S 2 0 7 ) or bisulphate
(KHS0 4 ) are conveniently made in borosilicate test tubes held in either a
simple fusion rack on a Primus stove (Stanton, 1966) or a more elaborate
apparatus (Marranzino and Wood, 1956). Sulphur trioxide is the effective
agent:
2KHS04->K2S207 + H20
K2S207 - K2S04 + S 0 3
Release of water during fusion with bisulphate causes frothing and requires
more careful control than the pyrosulphate.
After cooling and leaching the fused mass with dilute acid, the resulting
solution is suitable for determination of many elements (Table 4-III). Better
recoveries are obtained for Mo and W than with alkali fusions (Stanton and
Hardwick, 1967; Stanton, 1970b; Quin and Brooks, 1972). The pyrosul-
phate fusion includes no oxidizing agent and is therefore not suitable for
decomposition of sulphide- or organic-rich samples.
Harden and Tooms (1964) reported the effects of sample mineralogy on
bisulphate fusion for determination of Cu, Zn, Ni and Co. Their results show
considerable variability, with extraction efficiencies ranging from less than
10% for Co, Ni and Zn in amphiboles and pyroxenes, to more than 90%
from quartz and feldspars. Extraction of Cu was always relatively high (60—
100%). Results for minerals are reflected in related lithologies, and in the
soils and sediments derived from them.

Fusion with ammonium salts

Ammonium iodide (and chloride). Fusion with ammonium iodide provides


a method of decomposing samples and separating Sn, Sb and Hg from poten-
67

TABLE 4-III
Applications of bisulphate and pyrosulphate fusions

Element 1
Flux: sample Leach acid Reference

As 4 :1 HC1 1
Bi 3 :1 HN03 1, 2
Bi(col./XRF) 3 :1 HNO3 3
Co 4—5 : 1 HC1 1, 2
Cu 5 :1 HC1 1, 2
Mn 5 :1 H 2 S 0 4 or ΗΝΟ3 1, 2
Mo 4 :1 HC1 4
Nb 20 : 1 tartaric acid 2
Ni 4-5 : 1 HC1 1, 2
Pb 5 :1 HC1 2
Sb 5-7.5 : 1 HC1 or tartaric acid 2, 5, 6
Sb(AAS) 4 :1 HC1 7
V 3-5 : 1 HNO3 1, 2, 8 3 , 9 3

w 4.5 : 1 HC1 10, 11


Zn 5 :1 HC1 1, 2

Final determination by colorimetry unless otherwise indicated: col. = colorimetry,


XRF = X-ray fluorescence, AAS = atomic absorption spectrometry.
Many earlier references are given in Stanton (1966) and Ward et al. (1963). References:
I = Stanton (1966); 2 = Ward et al. (1963); 3 = Stanton (1971a); 4 = Stanton and
Hardwick (1967); 5 = Jardine (1963); 6 = Schnepfe (1973); 7 = McHugh and Welsch
(1975); 8 = Roberts (1971); 9 = Stanton and Hardwick (1971); 10 = Stanton (1970b);
I I = Quinn and Brooks (1972).
N a 2 S 2 0 7 used rather than K 2 S 2 0 7 to avoid precipitation of potassium phosphotung-
state.

tial interferents in their determination by colorimetry (Wood, 1959; Ward et


al., 1963; Stanton, 1966). More recently the method has been used in the
determination of Sn by ICP-ES (Pahlavanpour et al., 1979) and both Sn and
Sb by AAS (Heffernan et al., 1967; Nicolas, 1 9 7 1 ; Schweinsberg and
Heffernan, 1972; Welsh and Chao, 1975, 1976).
When heated ammonium iodide releases anhydrous hydriodic acid:

NH4I - NH 3 + HI

which decomposes the sample and converts Sb, Sn and Hg to volatile halides
which sublimate, together with excess ammonium iodide at the cooler, upper
end of the fusion tube. The sublimate is then dissolved with dilute hydro-
chloric acid. In the determination of Sn, cassiterite is readily decomposed;
however, Sn occluded within silicate lattices is n o t released unless silicates are
destroyed by pre-treatment with hydrofluoric acid (Agterdenbos and
Vlogtman, 1972; Welsh and Chao, 1976). Ammonium chloride can be used
in place of ammonium iodide in the determination of Sb, volatile antimony
68

trichloride being formed (Stanton and McDonald, 1961).


According to Welsh and Chao (1975, 1976) correct fusion conditions are
critical to obtaining consistent results. For Sb they recommend drying the
4 : 1 flux/sample mixture overnight at 105° C and then fusing for 10 minutes
at 350°C in 25 X 200-ml culture tubes: for Sn a temperature of 500°C is
recommended. The tubes are heated in a Pyropot (Heffernan et al., 1967)
designed to ensure uniform heating at the lower end of the tube while the
upper end remains cool. Other heating devices for ammonium iodide fusions
have been described by Smith (1967) and Sierra and Leon (1967).
Pahlavanpour et al. (1979) describe a simple detachable condenser to collect
the sublimate (Fig. 4-5).

Ammonium fluoride. On heating to 145°C ammonium fluoride releases


anhydrous hydrogen fluoride which is very effective in decomposing such
resistant silicates as beryl, tourmaline, topaz and kyanite. It has been used in
the decomposition of beryl for estimation of Be (Hunt et al., 1959—60) and
could probably find wider application.

Fig. 4-5. Sample decomposition by sublimation with ammonium iodide using a detacha-
ble condenser (A); and (B) dissolution of the sublimate. (From Pahlavanpour et al.,
1979.)
69

Alkaline fusions
Alkaline fusions, usually with sodium carbonate or lithium metaborate,
appear in many schemes for whole rock analysis (Abbey, 1970; Ingamells,
1970; Maxwell, 1968). In the exploration laboratory their use has generally
been limited to determination of: (1) amphoteric elements of Groups V and
VI of the Periodic Table — these form soluble anions in alkaline media, e.g.
chromate, molybdate and tungstate; (2) uranium; and (3) chloride and
fluoride.
North (1956) first applied alkaline fusion to the colorimetric determina-
tion of W and Mo in geochemical prospecting. The sample is fused, in a
nickel crucible or disposable culture tube, with a flux consisting of sodium
carbonate/potassium nitrate/sodium chloride in a 5 : 1 : 4 ratio (Stanton,
1966; Ward et al., 1963). Potassium nitrate is added as an oxidant and
sodium chloride to decrease the viscosity of the melt. The solidified melt is
leached with water to dissolve molybdate and tungstate together with
aluminate and sodium silicate. Interfering elements, notably Fe, remain in
the residue as their insoluble carbonates. Low recoveries obtained by this
method, compared to pyrosulphate fusion, have been attributed to the diffi-
culty of fully leaching the fused mass with a small volume of water (Stanton
and Hardwick, 1967; Stanton, 1966; Quin and Brooks, 1972).
Chromium is determined colorimetrically, as chromate or with diphenyl-
carbazide, after fusion with a sodium hydroxide/sodium peroxide flux
(Wood and Stanton, 1956—57). Peroxide is used as an oxidant, rather than
potassium nitrate, to avoid interference from nitrate. Manganate, which
would also dissolve in the water leachate and interfere, is reduced to insoluble
manganese dioxide by addition of ethyl alcohol. Most other elements remain
in the residue as insoluble hydroxides.
For the determination of F and Cl samples can be sintered with a 2 : 1
mixture of sodium carbonate and potassium nitrate (Ficklin, 1970; Haynes
and Clark, 1972; Crenshaw and Ward, 1975), or fused with either sodium
carbonate/zinc oxide (Ingram, 1970) or sodium and potassium carbonate
(Hopkins, 1977). The sinter or fused mass is then leached with water and the
leachate analyzed with a specific ion electrode. Sodium hydroxide fluxes
(Kesler et al., 1973; Josephson et al., 1977; Hopkins, 1977) and lithium
metaborate have also been used (Bodkin, 1977). Results obtained with
several of these methods are compared in Fig. 4-4 and examples of their
applications to prospecting are to be found in Plüger and Friedrich (1973),
Farrell (1974), Lalonde (1974, 1976) and Phuphatana et al. (1976). Ficklin
(1975) used a sodium carbonate/potassium carbonate/magnesium oxide flux
in the determination of I.
Fluorimetric determination of U takes advantage of the intense fluores-
cence of uranyl compounds in the presence of fluoride. A powdered sample
(Grimaldi et al., 1954), or a suitable aliquot of a solution containing U
(Smith and Lynch, 1969), is fused in a small platinum crucible or dish with a
70

1 : 1 mixture of sodium and potassium carbonates containing 10% sodium


fluoride. As little as 2 ppm U can be determined on a 5-mg sample (Stanton,
1966), the small sample and large excess of flux (3 g) minimizing the
quenching effect of other elements, especially Mn and Fe, on the fluores-
cence. Fusion conditions affect the sensitivity (Fig. 4-6) and precision of the
method to the extent that Parslow and Dwairi (1976) concluded that the
care needed to achieve accurate results was incompatible with rapid analysis.
Subsequently, Parslow (1979) demonstrated that the effects of fusion condi-
tions and quenching could be related to absorption of infra-red (IR) by the
fused pellet. A modification to the fluorimeter to enable IR absorption to be
measured was described and an accuracy and precision of ±5% on a through-
put of one hundred samples per day was claimed. Some laboratories prefer
to avoid problems associated with quenching by digesting the sample
(p. 114) and then separating U by solvent extraction prior to its fluorimetric
estimation (p. 106).
Other fluxes, based on mixtures of sodium fluoride, lithium fluoride and
alkali carbonates, can also be used (Price et al., 1953; Centanni et al., 1956).

Dry ashing of organic-rich materials


For vegetation and other organic-rich samples, dry ashing or ignition is a

S £
o «> 620° C •
6 w 30 - ! 650° C
Q CM W~
z o ~** · ·
< d
*
z
c
o
JIT
<
-I W
CQ £
o
Z .2 20 -
LU > ^^e^700°C
g 5
m 2
Q
z<
< CO
Ui 3
£ u. i o - \7500C
</> O
>- 5
H <
α
>
t O v 800° C
(/> DC
z So
C/) 0 "I I I I I I I I
5 10 15 20 25 30 35 40
TOTAL TIME IN FURNACE, minutes
Fig. 4-6. Influence of fusion temperature, with a sodium fluoride/sodium carbonate/
potassium carbonate (9 : 45.5 : 45.5) flux, on sensitivity for the fluorimetric determina-
tion of U. The vertical axis shows the difference in readings between the blank and
0.005 Mg of U. (From Fletcher, 1954.)
71

convenient preliminary step in their decomposition providing such volatile


elements as As, Hg and Se are not to be determined. It is an essential pre-
treatment in the decomposition of organic-rich samples by fusion with pyro-
sulphate or digestion with aqua regia, and is also required to destroy sulphides
before extraction of Au with hydrobromic acid/bromine. The ground mate-
rial, dried at 60—100°C, is weighed into a platinum, silica or glazed porcelain
crucible and ignited for several hours in a muffle furnace at 450—550°C. If
large numbers of samples weighing 1 g or less have to be processed, short
wide-mouthed borosilicate glass test tubes can be used. The glass, however, is
more susceptible to attack by alkaline plant ash than either silica or glazed
porcelain.
When ignition is completed the fluffy white or grey-white ash is dissolved
in dilute acid (usually 2—6 M HCl) and an aliquot taken for analysis. Results
can be expressed on either a dry weight or ash weight basis: whichever is
chosen should be clearly stated in presenting the data. Loss (of weight) on
ignition (LOI) provides a reasonable estimate of soil organic matter content
(Davies, 1974).
Low recovery of an element after dry ashing of vegetation can result from
losses due to volatilization or by its retention on the walls of the crucible
and the insoluble constituents of the ash. Except for the volatile elements
already referred to, losses of metals are unlikely to be large enough to be
significant in routine exploration analyses. The author has, however, ob-
served rapid loss of Cu from solutions allowed to remain in contact with car-
bon particles that resulted from an incomplete ashing. Peachey (1976)
attributed low values obtained for Cu, in soils that were ignited prior to their
decomposition with nitric acid/perchloric acid, to its occlusion in dehydrated
iron oxides.

PARTIAL EXTRACTIONS

Partial extractions can enhance anomaly contrast when the anomalous


element is concentrated, compared to its distribution in non-anomalous sam-
ples, in a particular component (or components) of the sample that can be
selectively dissolved with a suitable reagent (Figs. 1-2, 1-8 and 4-13). Their
most general application is extraction of soluble or weakly bonded metals in
the familiar cold extraction tests for hydromorphic anomalies. They can also
be used t o distinguish metals distributed between silicate and sulphide phases
in bedrock, and to selectively liberate trace elements associated with a partic-
ular phase of a soil or sediment.

Application to bedrock samples

In the application of bedrock geochemistry to prospecting use of partial


extractions appears to have been limited t o dissolution of sulphides, leaching
72

of water-soluble constituents, and the estimation of non-silicate U. Dolezal


et al. (1968) and a review of phase analysis by Steger (1976) provide
stimulating insights into other possibilities.

Dissolution of sulphides
Several procedures have been suggested for selective liberation of metals
associated with sulphides: these include aqua regia (Warren and Delavault,
1959a, b; Brabec, 1971); a mixture of ascorbic acid and hydrogen peroxide
(Lynch, 1971); potassium chlorate and hydrochloric acid (Olade and
Fletcher, 1974); and various solutions containing bromine (Hausen et al.,
1973; Davis, 1972; Czamanske and Ingamells, 1970; Peachey et al., 1978).
When ascorbic acid/hydrogen peroxide is used sulphides are preferentially
dissolved by their oxidation with hydrogen peroxide, the ascorbic acid pro-
viding a reducing, acidic medium to prevent precipitation of hydrous ferric
oxides. Galena, arsenopyrite, chalcopyrite, pyrite, pyrrhotite, pentlandite,
tetrahedrite, sphalerite and awaruite are all dissolved in the method
described by Lynch (1971). With potassium chlorate/hydrochloric acid the
active oxidizing agent is the nascent chlorine generated: most sulphides are
rapidly dissolved (Dolezal et al., 1968).
Olade and Fletcher (1974) have compared the selectivity of aqua regia,
ascorbic acid/hydrogen peroxide and potassium chlorate/hydrochloric acid
for copper sulphides associated with porphyry copper deposits. Concentra-
tions of copper (Cu x ) and zinc (Zn x ) liberated with these reagents were com-
pared to total contents (Cu t and Z n t ) determined after decomposition with
hydrofluoric acid/nitric acid/perchloric acid. Two criteria based on geo-
chemical behaviour of Cu and Zn were used to evaluate the results: Cu, a
strongly chalcophile element, is probably largely present as sulphide inclu-
sions in most igneous rocks, whereas the more oxyphile Zn occurs predomi-
nantly within silicate lattices. Consequently:
Criterion 1: an efficient copper sulphide selective leach will give a high
Cu x /Zn x ratio in samples containing background Zn values.
Criterion 2: Cu x /Cu t will increase with Cu t as copper sulphide content
increases, until in strongly mineralized samples Cu x equals Cu t within the
limits of analytical error.
Results (Fig. 4-7) show that all three reagents partly meet these criteria.
However, the Cu x /Zn x ratio is greatest with the potassium chlorate/hydro-
chloric acid leach and lowest with aqua regia. Similarly, in samples with low
values of Cu t , Cu x /Cu t is lowest with the potassium chlorate/hydrochloric
acid. On this basis it was concluded that this procedure was least damaging
to silicates and most selective for sulphides. In a later study, Chao and
Sanzolone (1977) investigated the efficiency of potassium chlorate/hydro-
chloric acid for dissolution of individual sulphide minerals. Following the
procedure of Olade and Fletcher (1974), dissolution of galena, cinnabar,
orpiment, stibnite, sphalerite and tetrahedrite was essentially complete,
73

C Aqua regia

o ·

° °
O o O

o ° °° o ° ,
οθ o o °,

2
I-
X ° *<* o a o 0 8 °° °
o o o
0 o o o

B H ?O o - A s c .

l0
K
O
°l • • • · o g

• 1
·· • *· ·
• ••
• • • o 8 °
• ··· • • 6
ÜC 60|

UJ
0. •
• o o

• · o
o
o
o
o
o
o •
o
o cS> o o
°8 o o
i
1000 10,000 10 10,000

TOTAL COPPER CONTENT (ppm)

Fig. 4-7. Sulphide-selective decompositions: relationships between extractable Cu and Zn,


and total Cu content with potassium chlorate/hydrochloric acid, hydrogen peroxide/
ascorbic acid, aqua regia, and nitric acid/perchloric acid decompositions on bedrock sam-
ples associated with porphyry copper deposits, Guichon Batholith, British Columbia. See
text for discussion. (From Olade and Fletcher, 1974.)

whereas chalcopyrite and pyrite were only moderately decomposed (40—


70%) and molybdenite was scarcely affected. Potassium chlorate/hydro-
chloric acid followed by boiling for 20 minutes with 4 N nitric acid was
effective for all the sulphides studied. With the exceptions of orpiment and
stibnite, which were only weakly attacked, all the sulphides were moderately
decomposed with ascorbic acid/hydrogen peroxide.
Olade and Fletcher (1976) used the potassium chlorate/hydrochloric acid
leach to study the distribution of sulphides associated with porphyry copper
deposits. It was found that the leach enhanced bedrock geochemical patterns
74

related to distribution of sulphides and hence provides a better guide to


mineral zoning than can be obtained with a stronger, less selective, decom-
position.
Cameron et al. (1971) tested the efficiency of the ascorbic acid/hydrogen
peroxide leach on monomineralic concentrates of olivines and pyroxenes.
Their results (Table 4-IV) indicate that only minor quantities of Ni and Co
are released whereas Cu, which is present as sulphide inclusions in the ortho-
pyroxene, is almost fully liberated. The extraction was used by them
(Cameron et al., 1971) to evaluate the ore potential of ultramafic rocks in
the Canadian Shield and by Garrett (1975) as a guide to copper and zinc
sulphides in Proterozoic volcanic rocks of the same region. Subsequently,
Peachey and Allen (1977) investigated the application of the ascorbic acid/
hydrogen peroxide leach to weathered materials. It was concluded that,
although not quantitative, the method could be useful in identifying the
presence of sulphides. Cameron (1972) reported the extraction of Cu, Ni and
Co from the ultramafic rock standards UM-1, UM-2 and UM-4 with this
leach.
Davis (1972) reported considerable damage to silicates and carbonates
when ascorbic acid/hydrogen peroxide was used on Kambalda nickel ores
and preferred to oxidize sulphides with cold bromine followed by 0.01 M
hydrochloric acid. A similar method is given by Stanton (1976) and a
bromine/hydrochloric acid mixture (30 ml Br/1 6 M HC1) was also used by
Peachey et al. (1978) in a field test for determination of Cu in drill-sludge
samples. The Cu released was estimated colorimetrically and results were
judged to be better than those obtained with a portable XRF. The potassium
chlorate/hydrochloric acid extraction is also satisfactory for this purpose.

Water-soluble constituents
Determination of water-soluble constituents in bedrock has included esti-
mation of chloride released from fluid inclusions in freshly crushed plutonic

TABLE 4-IV
Extraction of copper, nickel and cobalt from some silicate minerals with an ascorbic acid/
hydrogen peroxide leach (data from Cameron et al., 1971)

Mineral Percent extraction *

Cu Ni Co

Olivine 67 4 4
Olivine 56 3 4
Clinopyroxene 43 5 3
Orthopyroxene 91 2 0
1
100 X (ascorbic acid/H 2 0 2 )/HF.
75

rocks (Van Loon et al., 1973), and leachable Ca, Mg, Na, K, F and Cl in
mafic and felsic volcanic rocks at the Brunswick No. 12 and Heath Steele
massive sulphide deposits (Goodfellow and Wahl, 1976). Although the
extraction procedure used by Goodfellow and Wahl was extremely simple
(1.0 g of rock powder was stirred with 10 ml of water for one minute) it
successfully outlined anomalous halos extending up to 800 m around the
deposits.

Non-silicate uranium
Under oxidizing, alkaline conditions in carbonate-rich waters U 4 + in
uraninite and pitchblende is oxidized to U 6 + and dissolves as a carbonate
complex — U0 2 (C0 3 )3~ (Langmuir, 1978). Solutions of sodium or ammoni-
um carbonates and hydrogen peroxide are therefore employed as lixiviants in
solution mining of uranium ores. The same principle can be applied to
extraction of U in the laboratory. For example, Ward and Bondar (1979)
were able to distinguish U anomalies attributable t o ore minerals, from
equally anomalous concentrations of lattice held U, by extraction with a 2%
sodium carbonate/5% hydrogen peroxide leach. In addition to its direct
applications to geochemical prospecting, this approach has potential value in
the identification of source rocks, with relatively high content of leachable
(labile) U, favourable to the development of sandstone-type uranium depos-
its.

Application to soils and sediments

During weathering primary minerals of igneous and metamorphic rocks


decompose at varying rates and new products, especially clays and hydrous
oxides or iron and aluminium, form. Trace elements in primary minerals are
dispersed mechanically until their host is sufficiently decomposed to release
them into solution. Their mobility is then determined by the chemical stabil-
ity of the dissolved species. Changes in solution chemistry, especially Eh and
pH; exchange reactions on clays and colloids; and fixation by organic matter
are among the many processes whereby dissolved metals become reassociated
with the solid phases of soils and sediments. As a result of these processes
distribution of trace elements in these media is extremely complex. The
principal associations likely to be present in most background soils or sedi-
ments are:
(1) Trace elements in lattices of undecomposed primary minerals. Varia-
tions in mafic and heavy mineral content, which are enriched in many trace
elements relative to quartz and feldspars, can be an important factor in the
contribution of this component to the overall trace element content of the
soil or sediment.
(2) Trace elements in the lattices of secondary minerals or occluded in
amorphous compounds; for example, the occurrence of trace elements in
76

lattices of clay minerals or in amorphous and crystalline sesquioxides of iron


and manganese.
(3) Trace elements adsorbed on surfaces of clays, iron and manganese
sesquioxides, and organic compounds, or on surface and interlayer exchange
sites of clay minerals. Ions held in exchange sites are very sensitive to
changes in solution chemistry and the extent to which surface adsorption
occurs is strongly dependent on pH (Fig. 4-8). With continued aging of
amorphous compounds, increasing crystallinity can result in increased fixa-
tion of the trace elements.
(4) Trace elements associated with organic matter either as a result of up-
take by the living organism or by complexation and chelation with organic
matter. The ability of peat bogs to scavenge and concentrate up to several
percent Cu from Cu-rich ground waters is an extreme example of this asso-
ciation.
In soils and sediments in proximity to a mineral deposit, a significant
proportion of the associated trace elements may also be present as:
(5) Major constituents of surviving ore minerals. This will apply partic-
ularly to such resistant minerals as cassiterite, scheelite, beryl and gold, but
other less stable minerals will also persist where physical erosion proceeds
faster than their chemical decomposition. Primary sulphides are only likely
to persist in environments with low oxidation potentials; for example below
the ground water table and in impermeable tills.
(6) Major constituents of secondary minerals, resulting from alteration of
primary ore minerals or by precipitation from metal-rich solutions. The
nature and distribution of the secondary minerals (for example malachite
and smithsonite) will reflect the chemistry and hydrology of groundwaters
derived from the mineral deposit.
Trace elements in lattices of relatively resistant primary minerals are

Fig. 4-8. Influence of pH on the adsorption of cations (Ag+) and anions (ΜοΟ^ ) on
hydrous oxides, clays and related materials. Based on results from Dyck (1971) and Jones
(1957).
77

usually much more difficult to liberate than those adsorbed on or occluded


by clays, iron and manganese oxides, and organic matter. Similarly, the ease
with which the trace elements occluded in secondary minerals and amorphous
phases can be released depends on the stability and reactivity of their host.
This has been considered by Rose (1975) in relation to the Eh and pH of the
chemical treatments commonly used as partial extractants.
Manganese oxides are relatively soluble, except under strongly oxidizing
neutral to basic conditions, whereas iron oxides can only be dissolved by
strongly reducing or acidic solutions. Acidic solutions also cause partial
dissolution of clay minerals, with concomitant release of lattice held trace
elements (Fig. 4-9), and displace metals associated with organic matter
(Fig. 4-10). Trace elements associated with organic matter can also be
released by strong chelating agents. However, breakdown of peat and similar
materials only approaches completion with strong oxidizing agents such as
hydrogen peroxide or hypochlorite. Cations loosely adsorbed or associated
with exchange sites on clay mineral surfaces are relatively readily removed
by competition with hydrogen ions (Figs. 4-8 and 4-9), or an excess of some
other cation, by ion exchange. Taking advantage of these differences in the
reactivity of the components of soils and sediments, it is possible to either
non-selectively remove trace elements associated with several phases; or
selectively dissolve a specific phase, releasing the associated elements with
only minimum damage to the other phases present.
Apart from enhancing contrast, non-selective partial extractions can pro-
vide insights into element behaviour during dispersion. Ratio maps, com-
paring results of partial and strong extractions, are particularly useful in this
respect. Low partial extraction ratios are likely to occur if physical

200 -i
Freely drained soil, UL1(A), extracted for 15 min. at 25°C

0 1 2 3 4 5
PH
Fig. 4-9. Influence of pH and solvent on extraction of Cu from a freely drained soil.
(From Ellis et al., 1967.)
78

^ & tit
Ξ — Ξ Zn
CO 50 H • · Ni
< A A Co
UJ
-1
<·> <·> Pb
UJ

♦ ♦ Cu

5 4 3 2 1 0
pH
Fig. 4-10. Release of metals at different pH values from humic materials. (From
Chowdury and Bose, 1971.)

weathering is proceeding much faster than chemical decomposition and ele-


ments are still locked-up within silicate lattices; or if chemical weathering
and leaching have depleted the pool of readily solubilized elements. Extreme
examples of this has been reported for acidic arctic soils associated with
weathering massive sulphides (Cameron, 1977; Miller, 1979): in these soils
leaching of Zn from the strongly acidic zones over the suboutcrop of the sul-
phides is so complete that negative Zn anomalies have developed in which
virtually none of the remaining Zn can be released by partial extractants.
Conversely, high partial extraction ratios can indicate chemical decom-
position without concomitant leaching or, more importantly, development
of transported hydromorphic anomalies by fixation of dissolved metals at
suitable sites. Some partial extractions, for example dilute hydrochloric acid,
will dissolve sulphides (or secondary oxides and carbonate minerals) and will,
therefore, also give high extraction ratios and accentuate anomaly contrast
when sulphide-bearing rock fragments or minerals are present in soils or sedi-
ments.
Selective dissolution of a specific phase is usually too specialized and too
time-consuming to be routinely used on exploration samples. It is, neverthe-
less, of great value, particularly when used sequentially, in providing infor-
mation on the distribution of trace elements within samples and hence on
factors influencing their dispersion. This is especially useful during orienta-
tion surveys, in designing analytical schemes, and in investigating unusual
anomalies.
Partial extractions of soils and sediments are much more sensitive to the
bulk physical and chemical composition of the sample than are strong
decompositions. Consequently, results obtained on different sample types
may not be strictly comparable. Furthermore, changes caused by sample
preparation, particularly dehydration and oxidation during drying, probably
influence subsequent extraction of the trace elements. Results obtained with
partial extractions are also greatly influenced by the sample to solution ratio,
the extraction period and temperature — all of which must be maintained as
79

constant as possible. Ellis et al. (1967) found, for example, that differences
of 10°C in room temperature could change the amount of Cu released with a
citrate/hydroxylamine buffer by as much as 50%. Response surfaces (Fig. 4-
11) from Sorensen et al. (1971) illustrate the influence of extraction time
and sample to solution ratio on extraction of Zn with 0.1 N hydrochloric
acid. The long extraction period before equilibrium is approached is typical
of many partial extractions, other than those involving ion exchange which
proceeds relatively rapidly. Reaction kinetics, i.e. changes in the rate of
release of elements as extraction proceeds (Fig. 4-12), were used by Ellis et
al. (1967) and Warnant et al. (1980) to characterize the distribution of Cu in
soils and sediments.

Non-selective partial extractions


Non-selective partial extractions, usually involving treating the sample
with a buffer solution, dilute acid or complexing agent, have been used in
the determination of loosely bonded heavy metals and halogens. Sample and
extractant, usually in a 1 : 10 to 1 : 50 ratio, are shaken vigorously in a test
tube. Because it is not the intention with these extractions to obtain detailed
information on distribution of trace elements within the sample, there is

E
a
a

<
en
h-
X
UJ

υ
z
N

300 100
SHAKING PERIOD, min
Fig. 4-11. Response surface showing the effect of different shaking periods and solution
to soil ratios on the amount of Zn extracted with 0.1 N hydrochloric acid from nine
alkaline soils. (From Sorensen et al., 1971.)
80

100000 -d A. Anomalous freely drained soil, UL1(A)

10000
E
α
α

1000
Ο

5
Ο
100

<
CC
I-
X 100000-J B. Background freely drained soil, UL2(A)
LU

UJ
CO
UJ
ζ 10000 -d
<
ο
ζ
<
2

CL
α.
Ο
ο

TIME, minutes

Fig. 4-12. Comparison of the rate of extraction, at 100°C with 0.2 M hydrochloric acid,
of Cu, Fe and Mn from (A) anomalous and (B) background, freely drained A horizon soil
samples, Kilembe, Uganda. Similarities between Cu and Fe curves suggest an association
between Cu and Fe in both sets of samples. (From Ellis et al., 1967.)

little advantage to prolonging the extraction in an attempt to achieve equilib-


rium and shaking for a few minutes is usually adequate to release sufficient
metal for analysis. After extraction the residue is allowed to settle, or
separated by centrifuging, prior to taking an aliquot of the supernatant solu-
tion for analysis. Stanton (1976) summarizes some partial extractions for As,
Cu, Pb, Ni, Zn and heavy metals: with the exception of Ni, colorimetry is
used for the final determination.
81

Determination of loosely bonded heavy metals. Before the development of


AAS, determination of loosely bonded heavy metals usually involved their
liberation with a buffer solution that provided the appropriate chemical
environment'for their reaction with a colorimetric reagent. This continues to
be very useful in field tests. For example, with dithizone field tests the
buffer is an acidic or alkaline solution of ammonium citrate/hydroxylamine
hydrochloride, depending on whether extractable Cu or total heavy metals
are to be determined, respectively. This buffer will release metals on
exchange sites and, depending on the length of the extraction period and pH,
partly dissolve manganese oxides by the reducing action of the hydroxyl-
amine (Table 4-V). Results obtained by Ellis et al. (1967) show a marked
increase in the amount of Cu released at pH values below 3 (Fig. 4-9) sug-

TABLE 4-V
Influence of hydroxylamine hydrochloride on the extraction of heavy metals from
stream sediments in relation to their manganese content l . Heavy metals determined by
the dithizone field test (from Canney and Nowlan, 1964)

Sample No. Citrate-soluble heavy metals Manganese

with NH 2 OH · HC1 without NH 2 OH · HC1

A-1046 900 17 150,000


738 550 17 80,000
1834 350 14 60,000
1914 450 25 60,000
638 225 11 45,000
653 140 27 45,000
1034 250 14 45,000
1037 180 17 45,000
736 140 22 30,000
1837 130 32 30,000
600 60 17 15,000
602 90 14 15,000
1839 70 22 15,000
1917 32 9 12,000
598 70 17 11,250
1918 32 9 7500
1915 27 20 4000
1932 70 40 4000
1916 9 4 1500
1929 14 9 1000
1934 11 5 1000
1927 14 9 750
1931 9 7 500
1935 5 4 500
1936 4 3 250

Values in parts per million (Zn equivalent).


82

gesting that under the acidic conditions used in determination of Cu, there
may also be some dissolution of clay minerals and iron oxides.
Cold dilute hydrochloric acid provides an excellent medium for AAS and
is therefore often used as a weak extractant. To avoid erratic results caused
by soil minerals neutralizing the acid, solutions stronger than 0.5 M should
be used (Ellis et al., 1967). Exchangeable metals will be released together
with some of the metals associated with clay minerals, manganese oxides and
organic matter. Sulphides and carbonates are also partly decomposed and the
almost complete release of Pb from Pb-rich soils near sulphide mineraliza-
tion has been attributed to the solubility of galena or the secondary mineral
plumbojarosite (Cameron, 1977; Miller, 1979). Dissolution of iron oxides is
probably limited by low solubility of Fe 3+ .
In geochemical prospecting for U the use of relatively strong decomposi-
tions appears to have been favoured. However, as described on p. 75 for
bedrock geochemistry, there appears to be some merit in using partial
extractants to selectively leach uranium minerals. The observations of Rose
and Keith (1976) corroborate this. Investigating geochemical dispersion in
stream sediments associated with uranium deposits in sandstones of north-
eastern Pennsylvania, they found that an acetic acid/hydrogen peroxide
leach was considerably better than total U (by neutron activation), and
slightly better than extraction with nitric acid, in distinguishing anomalous
sediments (Fig. 4-13). Similar conclusions were reached by Olade and
Goodfellow (1979), using a sodium carbonate/hydrogen peroxide leach to
study distribution of U in sediments of streams draining the Toombstone
Batholith. One advantage of the alkali-oxidizing leach is that the elements,
especially Mn and Fe, that quench the fluorescence of U in its fluorimetric
estimation are insoluble in the leachate.

Determination of fluorine. Pliiger and Friedrich (1973) and Schwartz and


Friedrich (1973) have compared the action of strong and weak decomposi-
tions on removal of F from soils and sediments in the vicinity of fluorite
deposits. Extracted F was measured with a specific ion electrode.
For partial extractants both groups of authors used the total ionic
strength adjustment buffer (TISAB), which is formulated to avoid inter-
ferences in the measurement of fluoride activity with the electrode, and 0.01
hydrochloric acid. In addition 0.003 N sodium hydroxide and 0.037 N ferric
chloride solutions were used by Pliiger and Friedrich (1973), and Schwartz
and Friedrich (1973), respectively. Results (Fig. 4-4; Table 4-VI) indicated
that TISAB (which has a pH of 5—6) removes least F and ferric chloride
removes most F. With anomalous samples ferric chloride removed almost all
the F, compared to 5—30% for background samples, and gave the best
anomaly contrast. Its action is attributed to the strong complexing ability of
Fe 3+ for fluoride, most of which is presumably present in the anomalous
samples as fluorite. Alternative extraction procedures for fluorite, with
83

k • Near prospects
o Background

[-
Ε Γ 3/
07
- / γ • ° /
- #/
O CO <k>/
I
03 o o^4

01 ι/ι i i i i IX 1 1 1 1 1 1 1 1
03 10
Total U (ppm)
Fig. 4-13. Total U and acetic acid/hydrogen peroxide-extractable U in stream sediments,
northeastern Pennsylvania. Sediments related to uranium prospects contain an increased
proportion of readily extractable U. (From Rose and Keith, 1976.)

TABLE 4-VI
Extraction of fluoride from soils with TISAB 0.01 M HCl and 0.037 M FeCl 3 (from
Schwartz and Friedrich, 1973)

Soil sample No. Fluoride content (ppm)

TISAB HCl FeCU

26/2a 160 1400 7900


26,12a 85 90 5300
51/3a 210 240 6400
45/4a <5 11 60
45/15a 70 660 3100
45/28a 100 750 3900
35/31a 120 1100 2000
45, 55a <5 <6 10
1
TISAB: 1 M NaCl, 0.25 M acetic acid, 0.75 M sodium acetate, and 0.001 M sodium
citrate; pH 5—6.
84

acidified beryllium nitrate or aluminium chloride solutions, were described


on p . 6 5 .

Selective partial extractions


In this section release of exchangeable ions, and the selective dissolution
of organic matter and iron and manganese oxides is considered. The signifi-
cance of selective must be emphasized: the decomposition procedures
involved are certainly not specific for particular phases and, strictly, it would
perhaps be better to use only operational definitions (e.g. metal extracted by
sodium hypochlorite . . .) rather than associating the elements released to a
specified phase. However, interpretation of geochemical patterns must often
take into account the abundance of organic matter and presence of hydrous
oxide precipitates. Consideration of extraction data in these terms is there-
fore both convenient and reasonable, providing it is remembered that the
action of the extractants can vary considerably depending on sample type.
Sampling should therefore be as consistent as possible and especial care
taken in interpreting the results when the presence of atypical phases is
suspected — this is often the case in regions with a previous history of
mining or smelting.

Exchangeable ions. Ions held in exchange sites on surfaces of clays and


colloids are very sensitive to solution chemistry and respond rapidly to
changes in pH (Figs. 4-8 and 4-9) or competition from other ions. They are
liberated by displacement with another ion:

TE e x + M a q = TE a q + M e x

where M is usually an excess of NH4, K + , Na + , Ba 2+ or Mg 2+ added in a


neutral solution to minimize dissolution of solid phases. Although divalent
ions should be more effective in removing exchangeable ions, univalent ions
are, nevertheless, used in many procedures. Use of K+ of NH4 results in
irreversible replacement of ions in interlayer exchange sites of vermiculite.
Providing the sample is fully dispersed and exchange sites are not
protected by a coating of iron and manganese oxides or organic colloids, ion
exchange proceeds rapidly. For example, exchange of Na from the clay frac-
tion of stream sediments was found to be essentially complete within sec-
onds (Kennedy and Brown, 1966). In most geochemical samples the propor-
tion of trace elements retained at exchange sites is generally low and will
often be below the detection limit of conventional AAS unless large samples
and pre-concentration techniques are used. Such extractions are therefore
seldom used in geochemical prospecting.

Organic matter. Most soils contain from less than 1 to 20% organic matter
(equivalent to up to about 10% organic carbon). Associated trace metals can
85

be released with a suitable chelating agent or by destruction of organic


matter with a strong oxidant such as hydrogen peroxide or sodium hypo-
chlorite. In peats and other organic-rich samples, successive leaches with
dilute acids of increasing strength can provide information on the fixation of
trace metals. In the case of Cu it has been found that the binding of the
remaining Cu to soil organic matter increases as Cu is removed, so that suc-
cessively stronger solutions of acids are needed to continue the extraction
(Lewis and Broadbent, 1961; Davies et al., 1969). Chowdury and Bose
(1971) reported that at pH values below 1 humates released almost all of the
associated trace metals (Fig. 4-10).
Hydrogen peroxide, which decomposes to water and does not introduce
extraneous ions into the analysis, is often used to destroy organic matter
(Jackson, 1958; Rose, 1975). However, it is less effective than alkaline
sodium hypochlorite solutions (pH 9.5), does more damage to amorphous
constituents and clay minerals (Anderson, 1963; Lavkulich and Wiens,
1970), and, by reaction with Ca, can produce calcium oxalate (Farmer and
Mitchell, 1963). Manganese oxides are reduced and solubilized by hydrogen
peroxide whereas hypochlorite oxidizes some Mn to permanganate
(Anderson and O'Conner, 1972). Both reagents oxidize sulphides.
A disadvantage of sodium hypochlorite, compared to treatment with
acidified hydrogen peroxide, is that in the alkaline solution many of the
trace metals released will be precipitated as their basic hydroxides which
must then be dissolved with a weak acid leach (Table 4-VII). However,
Hoffman and Fletcher (1980) found that when a peat containing 2.6% Cu
was treated with sodium hypochlorite, the resulting brown solution con-
tained 20,000 μg Cu. This unexpected result was attributed to soluble
organic compounds, produced by incomplete oxidation of the peat, com-
plexing Cu and preventing its precipitation. The solubilizing action of humic
and fulvic acids on metal compounds is well known: Rashid and Leonard

TABLE 4-VII
Comparison of the action of sodium hypochlorite on release of copper from granodiorite
and peat samples (from Hoffman and Fletcher, 1980)

Extraction Peat Granodiorite

Cu extracted Cu extracted
(ppm) (%) (ppm) (%)

H N 0 3 * HCIO4 (total) 26,100 100 14,000 100


Sodium hypochlorite 20,100 77 500 3.6
Acidified water * 2090 8 10,000 71

pH 3.0: redissolves basic copper hydroxides precipitated in the alkaline hypochlorite


solution.
86

(1973) have shown, for example, that alkaline solutions containing humic
acids can retain approximately five times the amount of Cu in solution than
the same solutions without humic acids.
Extracts of organic-rich samples and natural waters draining peat lands are
often strongly coloured by organic colloids and soluble organic compounds.
These organic compounds, which are capable of sequestering heavy metals,
must be broken down prior to colorimetric analysis. Oxidation with
hydrogen peroxide or bromine (Mitchell and Smith, 1974) can be used.
Alternatively an aliquot can be evaporated to dryness with nitric acid/
perchloric acid. In the case of natural waters, the difference between total
dissolved metal and the fraction capable of reacting directly with a colori-
metric reagent such as dithizone, provides a rough estimate of the extent of
organometallic complexing.
Providing it is not necessary to remove organic matter as part of a sequen-
tial scheme, chelating agents can be used to extract associated heavy metals.
Ethylenediaminetetracetic acid (EDTA), a strong chelating agent for heavy
metals, is commonly used as 0.05- to 0.25-M solutions of its disodium salt
for this purpose. The potential value of this extraction is illustrated in a com-
parison of the action of 0.1 M EDTA and hot 4 M nitric acid on gytta
(organic-rich) lake sediments (Timperley and Allan, 1974): EDTA was found
to give the more consistent background results for Zn (Fig. 4-14). This was
attributed to the selective action of EDTA on the organic matter whereas the
hot acid also partly decomposed silicates and ferromanganese nodules which

100]

80

601

·%·.
10 20 30 £0 50 60 70 80 20 40 60 80 100 120 U0 160

Fig. 4-14. Analysis of lake sediments. A. EDTA-soluble Cu versus nitric acid-soluble Cu.
B. EDTA-soluble Zn versus nitric acid-soluble Zn. ♦ = samples containing ferromanganese
nodules given abnormally high Zn concentrations with the acid decomposition. (From
Timperley and Allan, 1974.)
87

were present in variable quantities in the sediments. In a study of Cu in peat


bogs (Maynard and Fletcher, 1973), 0.05 M EDTA and 0.5 M hydrochloric
acid were found to extract similar amounts of Cu ranging from 13 to 100%
of the total content. No significant difference in the amount of Cu extracted
was observed between anomalous and background samples (Table 4-VIII).

Removal of iron and manganese oxides. The association of trace elements


with precipitated iron and manganese oxides in soils and sediments ranges
from very loosely adsorbed (exchangeable) forms, through moderately fixed
forms associated with amorphous oxides, to relatively strongly fixed forms
occluded in goethite, lepidocrocite and other oxide minerals. Strongly acidic
or strongly reducing conditions will dissolve all these phases. Alternatively,
by controlling the Eh and pH of the decomposition some selectivity can be
obtained.
Manganese oxides. Manganese oxides are stable under a narrower range of
conditions than iron oxides so that they can be separated from them by their
relatively rapid dissolution under moderately reducing, neutral to acidic
conditions. Hydrogen peroxide (pH 3.5) can be used for this purpose
providing organic matter has already been destroyed (Rose, 1975; Jackson,
1958). Alternatively, acidified solutions of hydroxylamine hydrochloride
can be used — dissolution of manganese oxide by this reagent as a con-
stituent of buffer solutions has already been noted (Table 4-V). Chao (1972)
investigated optimum conditions for removal of manganese oxides with
minimum dissolution of iron oxides: these were found to be 0.1 M hydroxyl-
amine hydrochloride in 0.1 M nitric acid (pH 2) with a 30-minute extraction.
This released an average of 85% of the total Mn but only 5% of total Fe
content. Increased extraction time, lower pH and increased concentrations
of hydroxylamine hydrochloride all increased solution of iron oxides.
Horsnail et al. (1969), following a procedure described by Sherman et al.
(1942), extracted easily reducible Mn with a 0.2% solution of hydroquinone
in N ammonium acetate (pH 7) and distinguished between exchangeable,
readily reducible, and relatively inert forms of manganese oxides in stream
sediments. Readily reducible Mn was much more abundant in sediments of
streams draining poorly drained moorland soils than in those draining freely
drained soils (Fig. 4-15) and was probably responsible for accumulation of
As, Co, Ni and Zn in the sediments of the moorland streams.
Removal of iron oxides. Many methods have been described for the reduc-
tion and removal of iron oxides from soils prior to studies of clay mineralogy.
For geochemical purposes extractions with acid ammonium oxalate
(LeRiche and Weir, 1963; McKeague and Day, 1966); oxalic acid (Alminas
and Mosier, 1975; Carpenter and Hayes, 1979); sodium dithionite (Mehra
and Jackson, 1960; Mackenzie, 1954; Mitchell et al., 1971; Deb, 1950); or
hydrazine (Gatehouse et al., 1977) are probably most useful.
Removal of iron oxides with acid ammonium oxalate (pH 3.3) was first
00
00

TABLE 4-VIII
Extraction of copper from anomalous and background peat samples with nitric acid/perchloric acid, 0.5 M hydrochloric acid,
0.05 M EDTA and 1 M ammonium acetate (from Maynard and Fletcher, 1973)

HNO3/HCIO4 0.5MHC1 0.05 M EDTA CH3COONH4

A B A B A B A B

Analytical precision; 2σ level (±%) 3.2 6.8 18.6 5.7 7.7 22.0 n.d.
Cu extracted (ppm) GM 116 19 84 11 73 15 3 n.d
R 26-565 9-32 24-444 2-23 19-307 5-26 2-15 n.d,
Cu extracted (% of total) Av — — 73 67 68 82
n d
R 49-92 13-100 18-100 30-100 <l-8
Threshold (ppm) (GM B + 2σ) 36 38 36 n.d.
Average contrast (GM A /threshold) 3.2 2.2 2.1 n.d.

A = anomalous samples; B = background samples; GM = geometric mean; R = range; Av = average; n.d. = no data.
89

25000 POORLY FREELY 500


DRAINED DRAINED
Ηοο7„ι

h 400

TOTAL Mn

z
O 15000

CC
LU UJ
O. Q.

10000 200
CC CC

_ EASILY
REDUCIBLE
5000 hlOO

EXCHANGEABLE
Mn

Fig. 4-15. Total, exchangeable and easily reducible Mn in drainage sediments from poorly
drained and freshly drained environments. Note the difference in scale. (From Horsnail et
al., 1969.)

described by Tamm (1922). The reaction is light sensitive and, unless carried
out in strong sunlight or ultra-violet light, dissolves only amorphous and
poorly crystallized oxides (Deb, 1950; McKeague and Day, 1966). According
to LeRiche and Weir (1963), gibbsite and boehmite are only very slightly
decomposed whereas appreciable quantities of organic matter are co-ex-
tracted with the amorphous oxides. This is consistent with the observation
of Hoffman and Fletcher (1980) that ammonium oxalate extracts are
strongly coloured by organic compounds and contain greater concentrations
of the trace elements if organic matter is not first destroyed by a preliminary
treatment with hypochlorite. In carbonate terrains extraction with either
ammonium oxalate or oxalic acid can precipitate calcium oxalate causing
low recoveries for Pb (Alminas and Mosier, 1975).
Reduction with sodium dithionite (hydrosulphite) is more vigorous than
with ammonium oxalate and both amorphous and crystalline iron oxides
dissolve. Many procedures have been described: in that of Mehra and
Jackson (1960) sodium bicarbonate is used to buffer the reaction at the
optimum pH of 7.3 (Fig. 4-16) and sodium citrate (0.3 N) is added to com-
90

-100

-80

o—o After 15 min.


After 2 min. Q
co 0.6 H

O
UJ
-40 ü-o

-20

0
5 6 7 8 9 10 11 12 13
PH

Fig. 4-16. Curve of pH against the oxidation potential of a sodium dithionite/citrate


systems buffered with sodium bicarbonate, and associated solubility of iron oxides.
(From Mehra and Jackson, 1960.)

plex released Fe and. minimize precipitation of sulphides. Solid sodium


dithionite is added to the solution and the temperature is maintained at
80° C for the 15-minute extraction period. Any precipitated sulphides can be
dissolved, after decanting the supernatant extract, by treating the residue
with acidified hydrogen peroxide. The high salt content of dithionite solu-
tions can make AAS difficult and Gatehouse et al. (1977) therefore sub-
stituted hydrazine in hydrochloric acid as a reducing agent. No significant
damage to silicate minerals was detected.
The differences between the action of ammonium oxalate, in the absence
of sunlight, and dithionite were used by Hoffman and Fletcher (1979) to
study partitioning of trace elements between amorphous and well crystallized
iron oxide phases in soils and sediments. Some of their results for lake sedi-
ments are shown in Fig. 4-17.

Sequential extractions. After removal of organic matter and oxide phases the
bleached residue, consisting principally of primary and secondary silicate
minerals, can be mineralogically examined and the size fractions separated.
Strong decompositions can then be used to determine their trace element
content. Four sequential extractions schemes, in which organic matter and
hydrous oxide associated trace elements are estimated prior to analysis of
91

500
200

120 180 240 300 360 420 480 120 180 240 300 360 420 480

500
200

k<^^__*

120 180 240 300 360 420 480 60 120 180 240 300 360 420 480
MOLYBDENUM ZINC

z
O
υ

LEGEND

1 ORGANIC FRACTIONS

1A DILUTE HCI
2 AMORPHOUS OXIDES
3 AMORPHOUS Fe OXIDES

4 CRYSTALLINE Fe OXIDES

5 — SILICATE MINERALS

6 TOTAL CONTENT

180 240 3θΟ 360 420 4ΘΟ

MANGANESE

Fig. 4-17. Sequential extraction of Mn, Mo and Zn from lake sediments, on a transect
across Capoose Lake, British Columbia, using the extraction scheme of Hoffman and
Fletcher (1979). Details of the extraction appear in Table 4-IX. (Modified from Hoffman
and Fletcher, 1981.)
CO
to

TABLE 4-IX
Four sequential extraction procedures (intermediate extraction steps, to redissolve any precipitates formed, are not included)

Step Extraction procedure

Hoffman and Fletcher Gatehouse et al. Rose and Suhr Chao and Theobald
(1979) (1977) (1971) (1976)

II sodium hypochlorite *, H.O sodium acetate/acetic 0.1 M hydroxylamine hy-


pH9.5 acid, pH 5 drochloride in 0.01 M H N 0 3
[OM, S] 2 [exch., carb.] [Mnox.]
II hydroxylamine hydro- ammonium acetate/acetic hydrogen peroxide 0.25 M hydroxylamine
chloride, pH 2.5 acid, pH 4.5 hydrochloride in 0.25 M
HCL
l
[Mnox.] [exch.] [OM,S, Mnox.] [amor. Feox.]
III ammonium oxalate, pH 3.5 hydroxylamine hydro- sodium dithionite- sodium dithionite
chloride, pH 4.5 citrate-bicarbonate
[amor. Feox] [Mnox.] [Feox.] * [xst. Feox.]
IV sodium dithionite/citrate/ hydrogen peroxide size fraction and spectro- potassium chlorate/
bicarbonate, pH 7.0 graphic (DC-arc) analysis hydrochloric acid
of residues
[xst. Feox.] * [OM, S] [S]
V size fraction and H N 0 3 / hydrazine chloride silicate residue with
HCIO4 acid decomposition [Feox.] HF/HNO3
of residues
VI size fractions separated:
HCIO4 acid digestion
1 Precipitates formed with these extractions must be redissolved (see original references).
2 [Fractions decomposed]: OM = organic matter; S = sulphides; Mnox. = manganese oxides; amor. Feox. = amorphous iron oxides;
xst. Feox = crystalline iron oxides; Feox. = iron oxides; exch. = exchangeable; carb. = carbonates.
93

the silicate residue, are summarized in Table 4-IX with some typical results
in Figs. 4-17 and 4-18. The sequence of extraction is, of course, critical —
the least aggresive reagents being applied first and organic matter removed
before dissolution of the oxides. This prevents partial release of humic sub-
stances solubilizing metals and influencing the results of subsequent stages in
the procedure (Hoffman and Fletcher, 1980).
The application of selective and sequential extractions to the special prob-
lems of geochemical prospecting in former mining districts has been dis-
cussed by Dijkstraet al. (1979). In these circumstances, with unusual mineral
phases present, results of sequential extractions can be misleading and the
distribution of ore or smelter contaminants in soils and sediments can some-
times be more readily identified as the difference between two extractions,
only one of which is capable of dissolving the contaminant phases.

depth Fe Mn Zn Pb

Fig. 4-18. Distribution of the total amounts of Fe, Mn, Zn, Pb and Cu available in each
10-cm interval of South Comet soils, western Tasmania, using the extraction scheme of
Gatehouse et al. (1977). See Table 4-IX for details of the extraction. (From Gatehouse et
al., 1977.)
94

SUMMARY

Comparing the dispersion of trace elements and the ability of different


extractions to delineate soils and sediment anomalies related to mineral
occurrences, Bradshaw et al. (1974, p. 223)'concluded that:
(1) "While surface geomorphological conditions may differ drastically in
tropical residual, Alpine glaciated, and continentally glaciated terrain, the
mechanisms of geochemical dispersion are essentially the same, although the
surface expression may be different.
(2) That the choice of chemical extractant used on sediment and soil
samples is very important to the character and significance of the results ob-
tained. While few extractions can be regarded as giving results of no value to
exploration, some provide far more meaningful information than others.
The type of extraction used may well vary between reconnaissance and fol-
low-up stages depending on the information required.
(3) The stronger chemical attacks, when used on sediments, tend to
accentuate the effects of changes in rock composition and consequently
relatively reduce the effect of mineralization. In some cases mineralization
will be clearly seen by the use of a total attack, even in areas of varying
geology. In other instances, however, the mineralization is not easily seen by
the use of total attack on stream sediments because the variations in metal
content of different rock types are large with respect to anomalies from
mineralization.
(4) A particular weak extraction, for example 0.5 N HC1 or EDTA, will
not produce the same effect when used to analyze sediments in all areas of
the world.
In some instances, the metal is relatively tightly bonded even after hydro-
morphic movement for reasons that are not clearly understood at this time.
In this case, EDTA-extractable metal will not detect the strongest anomaly
with the best contrast and the use of a stronger, but still less than total
attack should be considered.
(5) The percentage of cold-ex tractable metal is increased in mineral soils
where the anomalies are a result of hydromorphic accumulation and increases
very dramatically in organic soils under the same conditions.
(6) It is very important, even in relatively well explored areas, to under-
take orientation surveys, in order to determine the variations in the param-
eters discussed above."
On the basis of the foregoing comments it is suggested that under most
circumstances, where anomalous concentrations of the metals can be intro-
duced into the sediments or soils by both mechanical and hydromorphic dis-
persion, the strong decompositions are of most general applicability. Their
use is unlikely to result in many, if any, anomalies being entirely missed.
However, if the anomalies sought are largely of hydromorphic origin (or
sulphides are present) an appropriate partial extraction can significantly
95

enhance anomaly contrast. It should be noted that the more specific the
action of the partial extractant, the more demanding becomes the task of
sampling similar materials at each site. Consequently, the non-selective
partial extractions are of greatest use in routine geochemical prospecting.
They are also well suited, in conjunction with colorimetry, to field tests.
From an anlytical viewpoint the decomposition procedure selected should
lend itself to rapid sample throughput, be compatible with the determina-
tion of a large number of elements on the same solution, and minimize inter-
ferences.
If the elements of interest are present in resistate minerals, decomposi-
tion can often be avoided and the solid sample analyzed directly by DC-ES
or XRF.
Chapter 5

COLORIMETRY AND RELATED TECHNIQUES

INTRODUCTION

In this chapter three analytical methods — colorimetry, fluorimetry and


turbidimetry — relating analyte concentration to intensity or absorption of
visible light are considered. The principles involved are illustrated in Fig. 5-1
and can be summarized as follows:
(1) Colorimetry. Estimation of intensity or hue of colour in a coloured
solution, either by visual comparison or instrumentally by measurement of
the absorption of the complementary colour (spectrophotometry) as light is
transmitted through the solution.
(2) Fluorimetry. Estimation of the intensity of fluorescence of a solu-
tion or solid excited by exposure to ultra-violet (UV) light.
(3) Turbidimetry. Estimation of turbidity of a suspension either by visu-
ally estimating light extinction or instrumental measurement of absorption
of a light beam passing through the suspension.
Of the three methods colorimetry is by far the most important and the
development of sufficiently sensitive, rapid colorimetric techniques,
requiring only simple equipment, was a major impetus in development of
exploration geochemistry in the 1950's. Excellent texts on colorimetric tech-
niques are available ranging from the comprehensive treatment of Sandell
(1959) to manuals providing detailed instructions for analysis of exploration
geochemical samples (Stanton, 1966, 1976; Ward et al., 1963). Most of the
procedures developed specifically for analysis of exploration samples enable
about a hundred determinations per man day. These procedures will not be
redescribed here and this chapter provides only a general review of the topic.
Although supplanted by AAS for determination of many elements in the
laboratory, the simplicity and suitability of colorimetric methods to field
usage should be emphasized. Furthermore, because of widespread concern
over the environmental significance of many elements, simple inexpensive
water testing kits are available for many cations and anions. Field determina-
tion of stream pH, using liquid indicators or test strips, is an excellent exam-
ple requiring no further elaboration. Reid (1969) has prepared a useful com-
pilation of mineral staining tests.
98

Transmittance

B. lo-»t
Ύ
i m ^

solution-ce

lo-»t
Scattering
Ύ

solution-cell

Ύ
E? κ :

Fig. 5-1. Idealized representations of (A) colorimetry, (B) spectrophotometry, (C)


turbidimetry, and (D) fluorimetry. I0 and J t = intensity of incident and transmitted light,
respectively.

COLORIMETRY

When white light passes through a coloured solution some wavelengths are
selectively absorbed: non-absorbed radiation transmitted by the solution
causes its colour. For example, the purple colour of potassium permanganate
solutions is due to their strong absorption of green, and transmission of
violet and red wavelengths (Fig. 5-2). Solutions of coloured substances can
therefore be characterized by their transmission or absorption curves for visi-
ble light.
99

400 450 500 550 600


WAVELENGTH, nm
Fig. 5-2. Transmittance curve for a solution of potassium permanganate.

Solutions of most metallic salts are not usually strongly coloured except
at high concentration (the blue colour of a copper sulphate solution becomes
apparent at about 100 Mg/ml) and are therefore not directly amenable to
trace analysis. Exceptions are determination of trace levels of Mn as
permanganate or Ti as the yellow colour produced with hydrogen peroxide
in an acidic solution. Almost invariably, however, it is necessary to combine
the analyte with an appropriate colorimetric reagent, usually an organic com-
pound, with which it develops a stable coloured complex at very low con-
centrations. An additional step, usually required, is the extraction and con-
centration of the coloured complex from the aqueous phase into an immisci-
ble organic phase. Providing the solution of the coloured substance is suffi-
ciently dilute the relationship between its concentration and absorption is
given by Beer's Law:

absorbance (A) = log(I0/It) = <*bc

where I0 and It = intensity of the incident and transmitted light, respectively,


a = a constant (absorbtivity), b = length of the light path, and c = concentra-
tion. The reciprocal ratio It/I0 is the transmittance of the solution. In practice
deviations from Beer's Law are not a problem because analyses are almost
invariably made by comparison to standard solutions, prepared in an identical
manner to the unknowns, of known concentration.
Concentration of the coloured substance can be estimated visually by
matching unknowns against a series of standards or instrumentally by mea-
suring the absorption of the complementary colour with a spectrophotom-
eter. The former, though less quantitative, has the advantage of being
100

simple, inexpensive and suited to field use. Many methods based on visual
colour comparison have therefore been developed for analysis of exploration
samples. Usually the coloured solution in a test tube is matched, against a
uniform white background, with a series of identical test tubes containing
standards of known concentration (see Fig. A-l, Appendix 1). A typical
series might, for example, cover the range 0—1 Mg/ml in 0.2^g increments,
1—10 μg in l.O^g increments and 10—20 μg in steps of 2.0^g. With such
a series an experienced analyst can rapidly match unknowns and standards
and interpolate between them giving results which are quite adequate for
most routine work. Regular preparation of standard series, using the same
batches of reagents used for unknowns, avoids the problem of deterioration
of standards with time and provides a check on methodology and reagent
quality. Any standard series that cannot be arranged in the correct sequence
without reference to its known content should be discarded. Use of artificial
standards, such as coloured plastics, is not recommended unless the real
standards deteriorate very rapidly.
In spectrophotometric analysis concentrations are estimated by measuring
the intensity of those wavelengths most strongly absorbed by the coloured
substance. The instrument is set at zero absorption against a solution blank
and absorption by standards and unknowns determined. Differentiation of
Beer's Law with respect to concentration shows that, ignoring other sources
of error, optimum concentrations for measurements are in the range 0.2—
0.7 absorbance (40—80% absorption): measurements close to 0 or 100%
absorption should be avoided if possible.
A wide variety of spectrophotometers are available. These vary from
simple instruments, using filters to isolate the absorption wavelength, to
more sophisticated instruments with monochromators. Solutions are usually
placed in rectangular glass or quartz cells which have been carefully matched
to have the same absorption characteristics. These cells must be handled with
great care and their optical surfaces kept scrupulously clean. To avoid
repeated handling and changing of cells some spectrophotometers employ
flow cells through which the solutions can be pumped. At least one manufac-
turer produces an instrument with a glass fibre optic cell which is lowered
into the solution to be tested. In general the simplest spectrophotometers are
fully adequate for routine exploration analysis.
Very few colorimetric reagents, 2,2' biquinoline for Cu being one, are
specific for a particular element. Most react to a group of elements and must
be made more selective by controlling reaction conditions with respect to
pH, Eh, time and temperature, and by use of masking agents which prevent
reaction with unwanted elements. Conditions for determination of most of
the elements of interest are treated very thoroughly by Sandell (1959) and
colorimetric reagents that have been used for analysis of exploration mate-
rials are summarized in Table 5-1. As an example of the applications and
problems of colorimetry the uses of dithizone (diphenylthiocarbazone), a
101

very versatile colorimetric reagent, will be considered briefly.


Dithizone, an indio-black crystalline solid, is used in colorimetric analysis
as its solution in hydrocarbons or chlorinated hydrocarbons, especially
chloroform and carbon tetrachloride in which its solubility is 2.0 and 0.5 g/
100 ml, respectively (Irving, 1977; Sandell, 1959). Dilute solutions as used
in analysis are green with an absorption minimum at about 510 nm; it reacts
with nineteen metals to form orange or red dithizonates. When a dithizone
solution in an organic solvent is shaken (under appropriate conditions) with
an aqueous phase containing any of these nineteen elements the final colour
of the organic phase, ranging from green through grey to blue, purple or red,
reflects the relative proportions of unreacted dithizone and metal dithizonates
(see Fig. A - l , Appendix 1). Metal concentrations are estimated either by
visually comparing the hue of the organic phase against a similarly prepared
series of standards, or by measuring absorption due to either the dithizonate
or remaining dithizone.
Obviously dithizone is far from specific for any one element. Fortunately
its selectivity can be improved by choosing appropriate reaction conditions.
An important factor is pH (Fig. 5-3) and it is usually adjusted with a suita-
ble buffer containing a complexing agent, such as citrate or tartrate, to pre-
vent loss of metals by precipitation under weakly acidic or alkaline condi-
tions. Because of the sensitivity of colorimetric reactions buffer solutions
must be purified by successive extractions with the colorimetric reagent.
In some cases control of pH alone does not provide adequate selectivity.
For example, in an ammonium citrate buffer at pH 8.5 dithizone reacts with
both Pb and Zn and several other metals. This is used to advantage in the
Bloom Total Heavy Metals test (Bloom, 1955; Stanton, 1976). If, however,
greater selectivity is required under these conditions a complexing agent,
which masks the reaction of one or more elements with dithizone, must be
added. Thus, addition of cyanide to the citrate buffer at pH 8.5 complexes
Cu, Co, Ni and Zn and the test becomes virtually specific for Pb (assuming
that Sn, Bi or Th are unlikely to be present in appreciable amounts in most
samples). A procedure for determination of Pb is summarized in Table 5-II.
Apart from its lack of specificity, problems can also arise with dithizone
as a result of its susceptibility to oxidation. This is particularly likely in
chlorinated hydrocarbons which can contain phosgene as a decomposition
product. In this respect non-halogenated solvents such as toluene and
benzene are preferable providing impurities are removed by treatment with
potassium dichromate (Stanton, 1966, 1976). Oxidizing agents such as ferric
iron in the sample solution can also decompose dithizone but this can be
prevented by addition of a reducing agent, such as hydroxylamine hydro-
chloride, to the buffer. Even with purified solvents dilute solutions of
dithizone slowly deteriorate. Working solutions should therefore be prepared
fresh each day from a much stronger (0.01%) stock solution in chloroform.
This stock solution should be stored in a cool, dark place.
102

TABLE 5-1
Some colorimetric reagents used in analysis of geochemical samples

Element Reagent Reference

Ag dithizone 56
catalytic oxidation Mn 2+ 55
As HgCl2 1-3,9-11
silver diethyldithiocarbamate 4,12,75
Molybdenum Blue 4
Cassius gold purple 84
Au Brilliant Green 2,27
4,4'-bisthiobenzophenone (TMK) 28
crystal violet 29
B methylene blue 1,17
Ba sulphate 2,3,13
Be beryllon II 2,14
Bi diethyldithiocarbamate 1 - 3 , 15, 16
Co 2-nitroso-l -napthol 3,19
tri-rc-butylamine 2,20
Cr as chromate 2,3,18
diphenylcarbazide 2,18
Cu 2,2 , -biquinoline 2 , 3 , 21, 77
dithizone 1, 2 2 , 2 3
rubeanic acid 24,76
Neocuproine 25
Ge phenylfluorone 2, 3, 26
Heavy metals dithizone 1,3,72-74
Hg dithizone 2,3,32
I as iodine 30
Mn as permanganate 2 , 3 , 31
Mo dithiol 1,2, 3 3 - 4 1 , 78, 85
thiocyanate 3, 4 2 - 4 4
Nb thiocyanate 3,79
Ni a-furildioxime 2, 3, 4 5 - 4 7
dimethylglyoxime 48
P as phosphomolybdovanadic acid 2,3,49
Pa reduction of phosphomolybdic acid 1, 50,80
Pb dithizone 1-3
Pt reduction of phosphomolybdic acid 1, 50,80
Rhodamine 6G 51
Rare earths (several methods) 52,81
Re thiocyanate 53
103

TABLE 5-1 (continued

Element Reagent Reference

Sb Brillant Green 2,5,82


Rhodamine B 3,6-8
Se 3-3'-diamino-benzidine 2,54
Sn Gallein 1 - 3 , 58, 59, 60, 67
dithiol 61
Th Arsenazo III 1, 5 7 , 8 3
Ti Tiron 2,3
V as phosphotungstovanadic acid 1 - 3 , 68, 69
W dithiol 1,2,33,62-64
thiocyanate 3,65,66
Zn dithizone 1 - 3 , 70, 71

References: 1 = Stanton (1976); 2 = Stanton (1966); 3 = Ward et al. (1963); 4 = Ward


(1975); 5 = Stanton and McDonald (1962); 6 = Ward and Lakin (1954); 7 = Jardine
(1963); 8 = Schnepfe (1973); 9 = Almond (1953a); 10 = Stanton (1964); 11 = Lynch and
Mihailov (1963); 12 = Rundle et al. (undated); 13 = Brobst and Ward (1965); 14 = Hunt
et al. (1959—60); 15 = Ward and Crowe (1956); 16 = Stanton (1971a); 17 = Stanton and
McDonald (1966); 18 = Wood and Stanton (1956—57); 19 = Almond (1953b); 20 =
Stanton and McDonald (1961-62a); 21 = Almond (1955); 22 = Holman (1963); 23 =
Holman (1956—57); 24 = Warren and Delavault (1958); 25 = Kauranne and Nurmi
(1967); 26 = Almond et al. (1955); 27 = Stanton and McDonald (1964); 28 = Lakin and
Nakagawa (1965); 29 = Kothny (1969); 30 - Grimaldi and Schnepfe (1971); 31 =
Almond (1953c); 32 = Ward and Bailey (1960); 33 = North (1956); 34 = Stubbs (1968);
35 = Stanton and Hardwick (1967); 36 = Stanton (1970a); 37 = Marshall (1968); 38 =
Marshall (1964); 39 = Stanton et al. (1973); 40 = Hoffman and Wasket-Myers (1974);
41 = Baker (1965); 42 = Ward (1951a); 43 = Lapointe (1968); 44 = Barakso (1967); 45 =
Stanton and Coope (1958-59); 46 = Nowlan (1970); 47 = Lynch (1967); 48 = Bloom
(1962); 49 = Peachey et al. (1973); 50 = Thompson (1967); 51 = Kothny (1974); 52 =
Rose (1969); 53 = Mahaffey (1974); 54 = Stanton and McDonald (1965); 55 = Nakagawa
and Lakin (1965); 56 = Bloom (1966); 57 = Stanton (1971b); 58 = Wood (1959); 59 =
Stanton and McDonald ( 1 9 6 1 - 6 2 b ) ; 60 = Smith (1967); 61 = Cogger (1974); 62 =
Stanton (1970b); 63 = Bowden (1964); 64 = Quin and Brooks (1972); 65 = Ward
(1951b); 66 = Cogger (1976); 67 = McDonald and Stanton (1962); 68 = Stanton and
Hardwick (1971); 69 = Roberts (1971); 70 = Lakin et al. (1949); 71 = Stanton (1962);
72 = Bloom (1955); 73 = Hawkes (1963); 74 = Smith (1964); 75 = Marshall (1978); 76 =
Delavault (1977); 77 = Peachy et al. (1978); 78 = Grifitts et al. (1976); 79 = Ward and
Marranzino (1955); 80 = Stanton (1975); 81 = Rose (1976); 82 = Stanton and McDonald
(1961-62c); 83 = Aly et al. (1977); 84 = Krings et al. (1976); 85 = Clark and Axley
(1955).

Most colorimetric procedures involve similar steps to those with dithizone:


decomposition of the sample; transfer of an aliquot of sample solution; addi-
tion of a buffer and, if necessary, a masking agent; addition of the colori-
metric reagent and finally extraction of the coloured complex into a suitable
104

pH
ION
1 2 3 4 5 6 7 8 9 10
+
Ag
H9 2 +
2+

.2+

Pb2

Fig. 5-3. Optimum pH ranges for formation of metal-dithizone complexes.

organic solvent. A few determinations deviate from this sequence. For exam-
ple, in the determination of As, arsine gas is produced by reduction of As
with nascent hydrogen: the arsine is then allowed to react with filter paper
soaked in mercurous chloride giving a coloured spot or, alternatively, is
bubbled through a solution of silver diethyldithiocarbamate (Fig. 5-4).
Chromatographie techniques, especially paper chromatography, can also
be used to separate and estimate concentration of coloured complexes of
trace elements. Theory and application of chromatography in geology have
been described by Ritchie (1964, 1969) and Hunt et al. (1955). Ritchie

TABLE 5-II
Procedure for the colorimetric determination of lead with dithizone (based on Stanton,
1966)

1. Weigh 0.1 g of sample into a test tube calibrated at 10 ml.


2. Add 2 ml 4 M nitric acid and digest for 1 hour.
3. Dilute to 10 ml with water and mix well.
4. Transfer 2 ml to a test tube containing 10 ml of buffer solution (100 g citric acid,
20 g hydroxylamine hydrochloride and 18 g potassium cyanide in 1 litre water: the
pH is adjusted to 9.5—9.8 with ammonia solution before addition of 100 ml of 18%
w/v potassium cyanide solution).
5. Add 5 ml 0.0008% dithizone in benzene (prepared by dilution of a 0.01% dithizone
in benzene solution).
6. Stopper the tube and shake vigorously for 15 seconds.
7. Compare with standard containing 0, 0.5, 1.0, 1.5, 2.0, 2.5, 3.0, 4.0, 5.0, 6.0 and
7.0 ßg Pb. The primary standard, containing 100 μg Pb/ml is prepared by dissolving
80 mg of lead nitrate in 1.2 M nitric acid and diluting to 500 ml with the acid.
105

Mercuric chloride
paper
ABSORPTION
0 ring
HEAD

glass wool soaked


in lead acetate

Absorption tube

ARSINE
GENERATOR
Silver
Diethyldithiocarbamate
solution

- Zinc pellets

Fig. 5-4. Determination of As by generation of arsine and its reaction with (A) mercuric
chloride paper, or (B) silver diethyldithiocarbamate dissolved in pyridine or quinoline.

(1964, p . 30) summarizes the essentials of paper chromatography as fol-


lows: "A drop or less of solute (for example, solution of a mineral) is
'chromatographed' by allowing solvents to move across the chromatography
paper past the point of application of the solute. The metal ions move across
the paper usually according to their ability to form complexes. The move-
ment of the ion proportional to that of the solvent is called the RF value of
the ion on that solvent:

distance moved by spot


Rv -
distance moved by solvent front

Some ions in some solvents form coloured spots which can be marked on
the chromatogram as soon as it is taken out of the vessel. Other ions, being
colourless on the paper, need to be treated with reagents (by spraying or
dipping) to produce either coloured or fluorescent spots." The solvent is
usually a mixture of a miscible organic solvent and a mineral acid.
Ritchie (1964) gives detailed Chromatographie methods for many trace
metals. However, the only elements routinely determined by chromatog-
raphy in the exploration laboratory are Nb (and Ta) and perhaps U. In the
determination of Nb (Hunt et al., 1955; Stanton, 1966) the sample is decom-
posed with HF and a 0.05-ml aliquot applied to the chromatography paper.
After drying for an hour the paper is stood in a solvent mixture of ethyl
methyl ketone and hydrofluoric acid until the solvent has diffused to the
106

top. The paper is removed, exposed to ammonia vapour for three minutes
and sprayed with tannic acid. Niobium then appears as an orange-yellow
band with a detection limit of about 4 ppm. Rapid Chromatographie methods
for determination of U have been described using potassium ferrocyanide
(Thompson and Lakin, 1957; Hunt et al., 1955; Ward et al., 1963) or l-(2-
pyridylazo)-2-naphthol (Plamondon, 1968); detection limits are 4 ppm and
1 ppm, respectively.

FLUORIMETRY

When a fluorescent substance is exposed to UV radiation it emits visible


light at a longer wavelength than the excitation source (Fig. 5-1). Intensity
of the fluorescence is proportional to concentration of the fluorescor and
can be estimated visually, by comparison to standards, or measured in a
fluorimeter.
In the exploration laboratory fluorescence of the uranyl ion, which is
particularly intense in the presence of fluoride, provides a sensitive method
for estimation of U in discs prepared by fusing the sample with a K 2 C0 3 /
Na 2 C0 3 /NaF flux (p. 69). Either less than 5 mg of sample powder is taken
for direct fusion (Standon, 1966; Grimaldi et al., 1954) or a suitable aliquot
of an acid digestion (p. 62) is evaporated, heated to destroy organics and
then fused with the flux (Smith and Lynch, 1969). Quenching agents, espe-
cially Mn (Table 5-III) but also Cr, Co, Ni, Fe, Ag, Pb, Pt and others, sup-
press the fluorescence. However, providing the sample to flux ratio is kept
low and fusion conditions are carefully controlled (Fig. 4-5), the method is
adequate for many purposes and has the advantage of simplicity (A.Y. Smith,
in Garrett and Lynch, 1976). Furthermore, the presence of excessive
amounts of Mn is readily apparent from the blue colour it imparts the fused
disc, and can be corrected for by taking either a smaller sample weight or
by precipitating manganese dioxide from the sample solution by addition of
potassium bromate (Ward and Bondar, 1979). Alternatively, by measure-
ment of the IR absorption of the disc, it is possible to correct for both
quenching effects and variations caused by fusion conditions (Parslow,
1979).
Despite the additional steps involved many laboratories prefer to avoid
potential quenching problems by separating U from other elements by
solvent extraction (with tributyl phosphate (Whitehead and Brooks, 1969),
ethyl acetate (Grimaldi et al., 1954) or tetrapropylammonium (Sandell and
Onishi, 1978)) from nitrate solutions salted with aluminium nitrate. A com-
parison of the fluorimetric determination of U, with and without solvent
extraction, in the presence of variable amounts of Mn is given in Table 5-III.
Recently, direct measurement of U in waters has been described using
laser-stimulated uranyl ion fluorescence (Robbins, 1978). The UV laser
107

TABLE 5-III
Comparison of uranium content of sediments determined by fluorimetry, with and
without solvent extraction, after digestion with 4 N nitric acid (data provided by S.J.
Hoffman)

Uranium content (ppm) Mn


(ppm)
direct fluorimetry with solvent extraction

58 282 238
1 3 118
2 5.5 226
5 16 954
0.2 1.0 5200
0.5 1.1 440
0.2 3.9 4300
16.8 62.5 240
0.7 6.8 3850
1.3 11.4 300
8.0 18.8 720
161.0 435.0 665
0.1 10.7 1640
31.2 105.0 680
0.8 370.0 5400
0.1 160.0 405,000

excites fluorescence by both natural dissolved organics and uranyl ions (Fig.
5-5). However, the fluorescence of the latter persists very slightly longer than
that of the organics when the laser is switched off. Consequently, by modu-
lating the laser and tuning the detector, it is possible to measure only the

"ORGANIC"
FLUORESCENCE
URANIUM
FLUORESCENCE

300 400 500 600


WAVELENGTH, nm
Blue Green
Fig. 5-5. Direct determination of U in solution by laser-induced fluorescence. As well as
being at a longer wavelength than the fluorescence due to organic matter, the fluores-
cence of the uranyl ion persists longer after the laser is turned off. (From Robins, 1978.)
108

uranyl ion fluorescence. A proprietary solution is added to the sample to


adjust pH and ensure U is present in the uranyl form; similar results can be
obtained by addition of a sodium hexametaphosphate solution. This nor-
mally ensures that U is present as a soluble uranyl phosphate complex
(υθ 2 (ΗΡ0 4 )2~). However, with very acid solutions (for example, if acid has
been added in the field as a preservative) problems can be encountered in
adjusting the pH to the optimum range (5—7) for formation of the complex.
If possible, therefore, waters should probably be analyzed, as soon as possi-
ble after collection, without acid preservation. The instrument described by
Robbins is portable and designed for field use.

TURBIDIMETRIC METHODS

Turbidimetric methods are seldom used except in the rapid determina-


tion of S, as sulphate, in natural waters and sample extracts. The sulphate is
precipitated from an acidic solution, by addition of barium chloride, as a
fine suspension of barium sulphate. Transmission of light through the
suspension is then compared to standards either visually, against a black
background, or by using a spectrophotometer to measure absorption of
either white light or any wavelength between 380 and 420 nm. The standard
suspensions must be stabilized by addition of glycerol or gum arabic
(Golterman, 1970; A.G.R.G., 1962).
Chapter 6

ATOMIC ABSORPTION SPECTROPHOTOMETRY

INTRODUCTION

Atomic absorption spectrophotometry involves the conversion of com-


pounds to their constituent atoms and then excitation of the atoms by
absorption of radiant energy. The amount of energy absorbed is measurable
and is proportional to the concentration of absorbing atoms. Although
atomic absorption was described by Kirchoff and Bunsen in 1860, and then
used to identify the elements present in the solar atmosphere, its potential
applications to analytical chemistry, with the notable exception of the deter-
mination of Hg, went unfulfilled for almost one hundred years. Then, in
1955, Walsh described both the theoretical applications of AAS to analysis
and the design of an atomic absorption spectrophotometer utilizing a hollow
cathode light source (Walsh, 1955). Subsequently, AAS expanded rapidly
into many fields of analysis, including exploration laboratories where it soon
replaced colorimetric procedures as the principal method for many trace
metal determinations (Table 1-V; Fig. 1-9). Although already being partly
replaced by ICP-ES in some large laboratories, it seems likely that AAS will
continue to be the principal analytical method in most exploration laborato-
ries over the next decade.

THEORY

Atomic absorption and atomic emission are related phenomena arising


from changes in the energy levels of an atom's outer electrons (Fig. 6-1). If
these are excited, for example in a flame or electric arc, they are raised from
their neutral, or ground state, to higher energy levels, the transitions involved
depending on the atomic configuration of the particular element and the
selection rules of quantum mechanics. In collapsing back from higher to
lower energy levels electrons loose energy by emission of photons whose
energy, and hence wavelength, corresponds exactly to the difference
between the initial and final energy levels of the transition. Each transition
therefore gives rise to a narrow emission line at a fixed wavelength and each
110

High energy level, E2

absorption

Low energy level, E-j

Fig. 6-1. Excitation of an electron from a lower (£1) to a higher (#2) energy level by
absorption of energy (E^) where E^ = E2 — E\. On collapsing back from E2 to E\ the elec-
tron releases a photon with an energy equal to E± and hence, from Planck's Law, a wave-
length in nanometres (nm) of 1.24/ü^, if E± is in kilovolts.

atom (or ion) has its own unique emission spectrum (Fig. 6-2). Conversely,
excitation of an electron to a higher energy level can result from absorption
of radiant energy of the same wavelength as the emission line for that partic-
ular transition. This is the basis of atomic absorption spectrophotometry.
The ratio of excited (N{) to ground state (N0) atoms at different tempera-
tures is given by:

i Pi (-ΕΛ
N,

where P{ and P0 are statistical weights for the ground and excited state, Ei is
the energy difference between the states, k is the Boltzmann constant, and T
is temperature in °K. As the value of Ei increases (i.e. as wavelength
decreases according to Planck's Law) the smaller will be the fraction of
excited atoms for a given temperature. As temperature increases the number
of excited atoms increases. However, even at high flame temperatures the
proportion of excited atoms remains relatively low (Table 6-1). Atomic
absorption measurements are usually made in flames with temperatures
between 2000 and 3000°C so that virtually all the atoms are in the ground
state (or energy levels very close to the ground state) and N0 is relatively
insensitive to small fluctuations of flame temperature. Also, because only
absorption lines originating in the ground state (or close to the ground state
e.g. Sn 235.48 nm and Zr 360.12 nm) will be sufficiently abundant to cause
measurable absorption, the absorption spectrum is much simpler than the
emission spectrum (Fig. 6-2).
The relationship between the intensity of an incident (I0) and transmitted
(It) light beam, of wavelength λ, passing through an absorbing cloud of
atoms follows Beer's law:

It = I0 exp — (Kxcl)
Ill

10,000

20,000

30,000

40 000 E
υ

50,000

μβο,οοο

^70,000

Fig. 6-2. Grotrian energy level diagram for the Au atom. The wiggly arrows indicate a
transition that is forbidden by the selection rules, a and b are metastable levels. Only
those wavelengths associated with the ground state (i.e. 164.6 nm, 242.8 nm and
267.6 nm) are involved in atomic absorption but the line at 164.6 nm is too low in the
UV to be useful and the 242.8-nm line is usually used for AAS determination of Au.
(Slightly modified from Ahrens and Taylor, 1961, Spectrochemical Analysis, 2nd ed.,
with permission of Addison-Wesley, Advanced Book Program.)

where Κλ is the absorption coefficient for the particular wavelength (λ), c is


the concentration of atoms, and / is the length of the absorbing path. There-
fore:

logioCo/Λ) = K\cl = absorbance (A)


With J0, as percent transmission, set to 100% (J0 — h) gives percent absorp-
tion, and:
100
absorbance {A) = log
100-(/0-/t)
112

TABLE 6-1
The relationship between temperature, wavelength and the number of excited (Nj) to
neutral (N 0 ) atoms

Absorption line Pi IPo


1
Ni/N 0
(nm)
2000°K 3000°K 5000°K

Na 589.0 2 9.89 X 10~6 5.81 X 10" 4 1.51 X 10" 2


Ca 422.7 3 1.21 X 10" 7 3.54 X 10" 5 3.31 X 10" 3
Cu 324.8 2 4.78 X 10" 1 0 7.70 X 10" 7 2.83 X 10" 4
Zn 213.8 3 7.25 X 10" 1 5 5.40 X 10" 1 0 4.27 X 10" 6

I\ and P0 are statistical weights derived from the quantum number J such that P =
2 J + 1: e.g. Zn 213.8 arises in the transition 'S0 -+ 'Px so that P0 = 1 and P{ = 3.

so that 1% absorption = 0.0044 absorbance units. In practice, the atomic


absorption instrument is calibrated with standards of known concentration
and it is therefore not necessary to know the values of either Κλ or /. The
sensitivity for the determination of an element by AAS is defined as the con-
centration required to give a 1% absorption signal. Typical values are
summarized in Table 6-II.
The natural width of an absorption line is about 10" 5 nm. However, as a
result of the Doppler effect and pressure broadening, caused by collisions
between atoms, this is increased to 0.001—0.003 nm. Additional line
broadening occurs in strong magnetic (Zeeman effect) or electrical fields
(Stark effect). In some specialized applications, described further on p. 135,
advantage is taken of pressure broadening or Zeeman broadening to measure
and correct absorbance results for unwanted, non-atomic background
absorption.

INSTRUMENTATION

Schematic diagrams of two atomic absorption spectrophotometers are


shown in Fig. 6-3. In the simpler, single-beam design, the instrument consists
of: (1) a light source emitting the sharp line spectrum of the element to be
determined, (2) a means of generating a cloud of atomic vapour, (3) a mono-
chromator and slit to select the desired absorption line, and (4) a detector,
amplifier and readout system. Components of the double-beam system are
identical but the light from the source is split into two beams, one of which
bypasses the atomic cloud and provides a reference against which instability
in the output of the source can be corrected.
113

TABLE 6-II
Crustal abundance of some trace elements compared to their analytical sensitivity by
atomic absorption spectrophotometry

Element * Crustal 2 Sensitivity concentration for 1% absorption


abundance
(ppm) flame furnace other
(Mg/ml) 3
(Mg/ml per 5μ1) 4 methods 5

Ag 0.07 0.029 0.00004


As 1.80 0.92 0.02 0.002 (c)
Au 0.004 0.11 0.002
B 10 8.4 (a)
Ba 425 0.20 (a)
Be 2.8 0.016(a) 0.00018
Bi 0.17 0.20 0.0014
Cd 0.20 0.011 0.00002
Co 25 0.053 0.0012
Cr 100 0.055(b) 0.001
Cu 55 0.040 0.0014
Hg 0.08 2.2 0.02 0.0001 (d)
Mn 950 0.021 0.0001
Mo 1.5 0.28 (a) 0.008
Nb 20 19 (a)
Ni 75 0.06 0.002
Pb 12.5 0.11 0.001
Rb 90 0.03 0.0012
Sb 0.2 0.29 0.006 0.08 (c)
Se 0.05 0.38 0.02 0.002 (c)
Sn 2 0.39 (a) 0.012
Sr 375 0.041 (a) 0.001
Ta 2 11 (a)
Te 0.026
U 2.7 113 (a)
V 135 0.75 (a) 0.02
w 1.5 5.8 (a)
Zn 70 0.009 0.000016
Zr 165 9.1 (a)
1
Elements in italics detectable by flame AAS at normal crustal abundances with a dilu-
tion factor of 50.
2
From Levinson (1974).
3
Lean air/acetylene flame unless otherwise stated: data based on manufacturer's litera-
ture for Techtron VI: a = nitrous oxide/acetylene flame, b = reducing air/acetylene
flame.
4
Manufacturer's data for Model 63 (Varian Techtron) carbon rod atomizer with 5-μ1
injections.
5
Other methods: c = hydride generation; d = flameless cold vapour method.
114

electronic modulation

"o-'t
*γ«
nm i · i n \

[ Ί
I i' ·m
^
>t lo

LIGHT SOURCE ABSORPTION DETECTION

B.

«o""t
~γ*

i m

Half-silvered
Rotating mirror
mirror
sectors

Fig. 6-3. Schematic diagram of atomic absorption spectrophotometers: (A) single-beam,


and (B) a double-beam instrument, h = hollow cathode light source; m = monochromator;
and p = photomultiplier. IQ is the intensity of the signal from the light source and It its
intensity after absorption by analyte atoms in the flame.

The light source

With a width of 0.001—0.003 nm, atomic absorption lines are too narrow
to be isolated by small monochromators, but if an attainable bandpass of
0.1 nm is used complete absorption of a continuum source by a line
0.002 nm wide only results in a 2% change in signal strength — too small a
difference to be measured accurately. Walsh (1955) realized that this
dilemma could be overcome by having the light source emit only the spec-
trum of the element of interest. Ideally the emitted lines should be sharper
(narrower) than the absorption line of the ground state atoms.
For a few volatile elements, such as Hg, As and Se, electrodeless discharge
lamps are ideal sources of an intense, stable emission. With most elements,
however, the best source is the sealed hollow cathode lamp. This consists of
115

a tubular cathode, containing the element(s) of interest, and an anode


mounted in a sealed envelope containing an inert gas, usually Ar or Ne, at
pressures around 2 mm Hg. For work in the ultraviolet, where the strongest
absorption lines of many metals occur, the lamps have a silica window. A
stabilized DC power supply, providing up to 800 V and a variable current
from 2 to 50 mA, is used to operate the lamp. Inert gas ions, produced at the
anode, are accelerated to the cathode where they sputter off and excite atoms
of the cathode causing them to emit the desired sharp line spectrum together
with that of the inert gas. The lamp is designed to enhance emission of
ground state transition lines relative to non-absorption lines, and, because of
the low pressure in the lamp, line widths of the emitted spectrum are less
than widths of absorption lines, at atmospheric pressure, in the atomic
absorption flame.
Apart from aligning the beam from the lamp through the atomic cloud,
the only variable in operating the hollow cathode lamp is the current. Above
and optimum current, which varies for each element, absorption decreases
with increasing current and the calibration line shows increasing curvature
(Fig. 6-4). This arises from several causes: (1) as current increases the num-
ber of atoms sputtered from the cathode increases, causing resonance
broadening — a special case of pressure broadening — so that the wings of the
broadened emission line may exceed the width of the absorption line in the
flame and therefore not be absorbable; (2) all the atoms sputtered from the
cathode are not excited; as the number of ground state atoms increases, self-
absorption of the central portion of the absorption lines becomes signifi-
cant; and (3) as cathode temperature increases, the number and intensity of
non-ground state lines in the emitted spectrum increases; if these lines are
not resolved from the absorption line, curvature of the calibration line
results.
Despite the decrease in absorption associated with increasing lamp current
above the optimum value, this may still be a justifiable procedure when mea-

Ni, j j g / m l
Fig. 6-4. Influence of varying lamp current (mA) and slit width (nm) on the sensitivity for
Niat 232.0 nm.
116

suring very small absorption signals. At low concentrations calibration


curvature is not a problem and the reduced signal amplification (gain)
required if lamp current is increased may result in a worthwhile improve-
ment of the signal to noise ratio and hence of the detection limit (Table 6-
III). Increasing the lamp current will also sometimes improve the signal to
noise ratio if the analyte must be measured in a strongly luminous flame.

Production of atomic vapour

With the exception of Hg, which is sufficiently volatile to produce ground


state atomic vapour at room temperature, some heat source is needed to
volatilize the sample, dissociate chemical compounds and produce a cloud of
absorbing atoms. The temperature must not, however, be too high or too
many atoms will be thermally excited to energy levels above their ground
state. Furthermore, it is in the establishment of complex chemical and
ionization equilibria during production of the atomic vapour, that many of
the interference problems associated with AAS arise.
Many methods of producing ground state atoms have been investigated
but only flames and, to a lesser extent, electrically heated furnaces have
found general application.

Flames
Usually the sample solution is aspirated, by a venturi driven by the pres-
sure of the oxidant gas, into a chamber where the resulting aerosol mixes
with fuel gas before entering the flame through a long, narrow slit. Burners
of this type, known as laminar-flow or pre-mix burners, produce very steady,
stable flames. Aspiration rates with aqueous solutions are usually 3—5 ml/
minute. However, only a small fraction of the solution reaches the flame; the
remainder, consisting of the larger droplets, is trapped by baffles in the

TABLE 6-III
Modifications to routine operating conditions

Situation Action

A. High analyte concentration — to decrease (a) use less sensitive absorption line
sensitivity (b) rotate burner
B. Very low analyte concentrations — to (a) increase slit
improve signal to noise ratio (b) increase lamp current
(c) both (a) and (b)
C. Strongly luminous flames — to improve (a) decrease slit
signal to noise ratio (b) increase lamp current
(c) both (a) and (b)
117

mixing chamber and drains to waste. This low efficiency in transfering


sample to the flame is one of the principal limitations on analytical sensitiv-
ity.
A variety of gas mixtures have been used but the lean, oxidizing air/
acetylene flame has the most general application. Its temperature (^2200°C)
is sufficiently high to ensure breakdown of most compounds but is not so
high as to cause significant ionization for elements other than the alkalies. As
would be expected the cooler (1800—1900° C) air/propane and air/coal gas
flames are more prone to chemical interferences and are therefore seldom
used. The hotter (2955°C), fuel-rich nitrous oxide/acetylene flame provides
a reducing environment for determination of those elements — especially Mo
in exploration samples — which tend to form refractory oxides (Fig. 1-9). Its
relatively high temperature causes significant ionization of both alkali and
alkali earth elements, and for Ba the ion absorption line at 455.4 nm pro-
vides better sensitivity than the atomic line (553.6 nm) unless ionization is
suppressed (p. 128, Fig. 6-10).
Wavelengths below 200.0 nm are strongly absorbed by the air/acetylene
flame. Hydrogen/argon and hydrogen/nitrogen flames are more transparent
in this region and are therefore recommended for measurements on the As
and Se absorption lines at 193.7 nm and 196.0 nm, respectively.
Whichever gas mixture is chosen, the heat of the flame is used to complete
the process of evaporation which begins in the mixing chamber, and then to
volatilize the solids and dissociate the chemical compounds producing a
reservoir of atoms. The number of ground-state atoms produced depends on
the aspiration efficiency and on the equilibrium established in the flame
between the dissociation of compounds, formation of new compounds and
the extent of ionization. The equilibrium condition is determined by flame
temperature and composition, the solvent used, and on the other elements
introduced into the flame by the sample. Consequently, production of
ground state atoms will vary from sample to sample, causing chemical and
ionization interferences, and, since flames are strongly zoned, in different
parts of the flame.
The distribution of absorbing atoms in a laminar air/acetylene flame is
shown in Fig. 6-5 from a study by Rann and Hambly (1965). In the lean
flame the maximum density of ground-state atoms, and hence greatest
analytical sensitivity, is found about 0.5 cm above the burner where the
flame is hottest. Higher in the flame the number of atoms decreases due to
dilution by flame gases and recombination of atoms. Molybdenum and other
refractory elements show a particularly rapid recombination rate, which can
be somewhat reduced by use of fuel-rich air/acetylene flame. Better sensitiv-
ity for these elements can, however, be obtained in nitrous oxide/acetylene
flames (Fig. 1-9).
For optimum sensitivity the burner height and fuel to oxidant ratios
and pressures should be determined separately for each element. In practice,
118

Rich Lean Rich Lean


COPPER MOLYBDENUM

Fig. 6-5. Distribution of Cu and Mo atoms in lean and fuel rich air-acetylene flames. Con-
tours are drawn at intervals of 0.1 absorbance units with maximum absorbance in the
centre: height is measured above the burner slot. (Reprinted with permission from Rann
and Hambly (1965), Distribution of atoms in an atomic absorption flame. Analytical
Chemistry, 37: 879—884. © 1965 American Chemical Society.)

however, these parameters are not too critical in the lean air/acetylene flame
and the best conditions for any one of the elements which are probably most
frequently requested in the exploration laboratory (Cu, Co, Ni, Pb, and Zn)
will usually be satisfactory for the remainder of the group. Considerably
greater care is needed to optimize flame conditions and the position of the
light beam in the red cone or feather of the fuel-rich nitrous oxide/acetylene
flame.

Furnaces
Resistance heated furnaces offer several theoretical advantages over flames
as a means of producing ground-state atoms, notably more efficient utiliza-
tion of the sample solution and a longer residence time for atoms in the fur-
nace than in the flame. This gives sensitivities appreciably better than those
attainable in flames (Table 6-II) and enables very small quantities of material
to be analyzed. Major disadvantages of furnaces, compared to flames, are
more severe interferences and their much lower analytical productivity.
The simplest furnaces are the silica tubes, heated either by resistance wire
(Chu et al., 1972) or by mounting them above an air/acetylene flame
(Thompson and Thomerson, 1974), used for atomization of elements, such
as As, that form gaseous hydrides. The hydride generating system and fur-
nace used by Aslin (1976) to determine As and Sb in geochemical samples is
shown in Fig. 6-6.
Less specialized furnaces for general analytical applications are usually
resistance-heated graphite tubes (Fuller, 1977). Alternatively, a strip (fila-
ment) of an unreactive metal, for example tantalum, enclosed in a chamber
119

Nitrogen

Dry Ash Atomize

Nitrogen B
Tygon tubing HOLLOW CATHODE LAMP
Plastic
stopper (tight fit)

CONTINUUM LAMP
ΊΓ
Eppendorf tip

Sodium NET SIGNAL


borohydnde
solution
time -
Fig. 6-6. Apparatus for AAS determination of elements forming gaseous hydrides. Cell A
is a 18 mm X 16 mm plastic test tube with a hole, 30 mm from the bottom, to take the
tip of a micropipet. A l-μΐ volume of sample is injected into the borohydride solution in
the cell. Gaseous hydrides are then swept by a flow of nitrogen into the silica absorption
cell (D) heated by the air/acetylene flame of a triple-slot burner. (From Aslin, 1976.)

Fig. 6-7. Typical absorption curves for the dry, charr and atomize stages of electro-
thermal atomizers. Subtraction of non-atomic absorption, measured with a continuum
source, from total absorption, measured with the hollow cathode lamp, gives the signal
due to the analyte. During drying and charring, smoke and vapour can give large non-
atomic absorption signals.

can be used. The power supply to such furnaces must provide three cycles,
each with variable time and current settings, for drying, charring, and finally
volatilization and atomization of the analyte. At the high temperatures (up to
3000°C) used for atomization, the furnace material would be rapidly con-
sumed by oxidation in air. It is, therefore, arranged for the furnace to be
sheathed with an inert gas (argon or nitrogen) or, to provide a more reducing
environment, hydrogen.
Sample solutions are usually introduced into the furnace with a micro-
pipet (1—100 μΐ) and the solvent evaporated during the drying cycle. The
current is then increased sufficiently to char and burn off organic matter
and, if possible, other matrix constituents. Ideally this step, which produces
smoke, should be completed with no volatilization of the analyte before the
atomization cycle begins. If charring overlaps with atomization, light losses
caused by the smoke must be subtracted from the true absorption signal by
simultaneous background correction. Some typical drying, charring and
atomizing, cycles are shown in Fig. 6-7.
120

Wavelength selection

Hollow cathode lamps or electrodeless discharge lamps emit a virtually


pure spectrum of the element to be determined. Consequently, requirements
for wavelength selection are relatively simple, it only being necessary to iso-
late the atomic absorption line from any neighbouring non-absorbing lines
emitted by the light source. For many elements, with simple spectra, a 1-nm
or wider bandpass is adequate. However, for elements with more complex
spectra, for example Mn and Ni, resolution of at least 0.1 nm is necessary.
These requirements can be met by small, relatively inexpensive, grating
monochromators.
When low absorbance values are to be measured curvature of the calibra-
tion line, due to inclusion of non-absorbing lines in the bandpass, is not
usually significant and it may be worthwhile to increase the slit width, there-
by letting more light into the monochromator and allowing the gain, and
hence noise, to be reduced (Table 6-III). It should be noted that doubling
the slit width will increase the light intensity for a sharp line source twofold,
whereas the intensity of a continuum source will be squared.

Detection and readout systems

Light from the monochromator enters a photomultiplier where it is con-


verted to an electrical signal and fed to an amplifier whose output goes to the
readout system. Photomultipliers supplied with atomic absorption spectro-
photometers usually have their optimum sensitivity in the 200.0- to 700.0-
nm region. Photomultipliers giving better response in the UV below 200.0 nm
are available to improve sensitivity for determination of As and Se.
In order to detect only the absorption signal and to reject flame emission,
particularly in the visible light region, the output of the light source must be
modulated and the detector tuned to accept only the modulated signal. This
can be achieved by either mechanically chopping the light between the
source and flame with a revolving mirror (Fig. 6-3), or by using alternating
current or modulated direct current as the lamp power supply. With strongly
emitting flames, as when large amounts of Ca are present, increased noise
and apparent interferences can result from the inability of the electronics to
fully reject the DC signal from the flame (Rooney and Woolley, 1978).
Increasing the brightness of the light source reduces the problem.
Readout systems with early atomic absorption spectrophotometers were
simple galvanometers calibrated in percent absorption or absorbance units.
Digital readouts, providing calibration in concentration units and features
such as curve correction, peak signal retrieval and signal integration, are now
standard with many instruments. Rapid detector response and peak signal
retrieval are particularly valuable with furnace methods that give very fast,
transient signals.
121

APPLICATIONS

Furnace techniques are well suited to direct analysis of waters but cannot
compete, either in productivity or freedom from interferences, with flame
AAS for the rapid routine determination of most of the elements of interest
in soils and sediments. Consequently, even although concentration by
solvent extraction may be required (Table 6-IV), most AAS analyses in the
prospecting laboratory involve aspiration of solutions into laminar flow
flames. Notable exceptions are the flameless determination of Hg and the
electrothermal decomposition of gaseous hydrides.
Willis (1975) has described an interesting technique for avoiding the
sample decomposition stage, whereby a 5% suspension of the powdered
sample in water is aspirated directly into the flame: Cu, Ni, Co, Zn and Pb
were determined in stream sediments.

TABLE 6-IV
Application of some solvent extraction schemes to geochemical samples
l
Element Extraction scheme Reference

Ag TOTP-MIBK C h a o e t a l . (1971)
Bratzel et al. (1972)
Au Cl-MIBK Bratzel et al. (1972)
Br-MIBK Tindall (1965)
Thompson et al. (1968)
DBS-toluene Rubeskaet al. (1977)
Be acetylacetone Terashima (1973)
Bi APDC-MIBK Ficklin and Ward (1976)
Mo SCN-MIBK Kim et al. (1974)
Aliquat 336-MIBK Rao (1971)
Pd DBS-toluene Rubeskaet al. (1977)
Pt Sn n -MIBK Stanton and Ramankutti (1977)
Sb Cl-MIBK McHugh and Welsch (1975)
TOPO-MIBK Welsch and Chao (1976)
Sn TOPO-MIBK Welsch and Chao (1976)
Te Br-MIBK Nakagawa and Thompson (1968)
C h a o e t a l . (1978)
Tl and In Br-MIBK Hubert and Lakin (1973)
w SCN-Alamine 336-MIBK K i m e t a l . (1976)
1
Abbreviations: TOTP = triisoctyl thiophosphate; MIBK = methyl isobutylketone, 4-
methyl-2-pentanone; Cl = chloro; Br = bromo; DBS = dibutyl sulphide; APDC = ammo-
nium pyrrolidinedithiocarbamate;SCN = thiocyanate; TOPO = trioctylphosphine oxide.
122

Flame atomic absorption

When solids are to be analyzed, the sample solution is generally either a


dilute acid or some other aqueous extract prepared as described in Chapter 4.
If analytical sensitivity is too low for direct determination of the analyte in
the original solution it may be concentrated into an organic phase by solvent
extraction.

Operation and calibration


Operation of atomic absorption spectrophotometers is usually straightfor-
ward, routine operating conditions being adequately described in most
manufacturer's manuals. At low concentrations, requiring scale expansion, or
when working with intensely luminous flames some deviation from these
conditions may improve signal to noise ratios (Table 6-III). High-precision
methods, for example standard additions or closely bracketing unknowns
with standards, are too time-consuming for routine use in the exploration
laboratory and the calibration curve is therefore usually based on a reagent
blank and two or three standard solutions. Also, to avoid further calculations,
the sample dilution factor (i.e. volume (ml)/weight (g)) is taken into account
and the instrument calibrated directly in concentration units. On instru-
ments not equipped with digital readout this can be done by inserting a
calibration scale (one for each element) over the face of the meter readout
and using scale expansion to make any necessary small adjustments for daily
changes of sensitivity, so that concentrations and the corresponding readings
for standard solutions coincide.
Ideally, standard solutions should have the same composition as sample
solutions. However, differences in absorption between dilute aqueous solu-
tions are often not too great and standards in dilute hydrochloric or nitric
acid will often suffice for analysis of samples in other dilute aqueous media.
This must, of course, be established for the particular solutions to be ana-
lyzed. Preparation of standards is described further in Appendix 2.
Sample solutions with element concentrations above the concentration
range of the instrument can be brought on-scale by use of a less sensitive
absorption line, by shortening the absorption light path by rotating the
burner, or by further dilution. When a constituent with a wide concentration
range, or several constituents at very different concentrations Eire to be deter-
mined from the same original solution, several dilutions may be needed.
Under these circumstances an automatic dilutor, of the type used in clinical
laboratories, providing a ten- to twenty-fold dilution saves considerable time.

Sensitivity
Of the trace elements in Table 6-II, crustal abundance of twelve (Ba, Be,
Co, Cr, Cu, Mn, Ni, Pb, Rb, Sr, V and Zn) is such that they should be
detectable with routine operating conditions and dilution factors in the
123

range 20—50 after complete, or almost complete, sample dissolution. For-


tunately, this group includes many of the elements most often sought in the
exploration laboratory. Sensitivity for several other important elements,
notably Ag, Cd and Mo, is borderline and these will be reported as not
detected in many background samples.
The remaining elements will usually only be detectable after their pre-con-
centration or by using special techniques to enhance sensitivity. Because of
its simplicity and rapidity, solvent extraction, whereby an uncharged ion
association or metal chelate is extracted from an aqueous phase into an
immiscible organic solvent, has found many analytical applications in the
concentration and separation of trace elements from interferences (Morrison
and Freiser, 1957; Stary, 1964). When used in conjunction with AAS the
organic solvent, in addition to having a low solubility in, and providing a
clean separation from, the aqueous phase, must have good burning character-
istics in the air/acetylene or nitrous oxide/acetylene flames. Despite a rather
high solubility in water (22 ml/1) methyl isobutylketone (MIBK: 4-methyl-2-
pentanone) fulfils most of these requirements and has become the preferred
solvent in many extraction schemes. Compared to aqueous solutions, use of
MIBK also results in a two- to five-fold increase in sensitivity, apparently due
to an increase in the aspiration rate and more efficient transfer of the analyte
to the flame (Allan, 1961; Chakrabarti and Singhal, 1969).
Although many solvent extraction schemes are described in the analytical
literature only a few have been applied to routine analysis of exploration
samples (Table 6-IV). Extraction of the chloro- or bromo-aurate ion into
MIBK in the determination of Au is probably the most frequently used.
Prior to solvent extraction Au can be solubilized with either aqua regia diges-
tion (Tindall, 1965) or, in the rapid method described by Thompson et al.
(1968), a hydrobromic acid/bromine mixture. Campbell (1980) found that
interferences from Mn at the 242.8-nm Au line could be overcome by boiling
the solution to remove unreacted bromine. With little modification solvent
extraction from hydrobromic acid/bromine can also be used for pre-con-
centration of In and Tl (Fratta, 1974; Hubert and Lakin, 1973), Te
(Nakagawa and Thompson, 1968) and Pt (Stanton and Ramankutti, 1977).
Some separation schemes, for example extraction of the red Mo-SCN com-
plex or the Pt IV compound formed in the presence of SnCl2, have evolved
directly from the colorimetric methods employed for determination of these
elements.
In addition to solvent extraction schemes for individual elements, multi-
element extraction schemes have also been described for determination of
trace elements in geological material. Hannaker and Hughes (1977, 1978),
for example, developed an extraction procedure to reduce interferences and
increase sensitivity in the determination of Cu, Ni, Co, Cr, Ag, Pb, Bi, Cd,
Zn, Mn, Au, Tl, Sb, Ga and Mo. Fortunately, providing background absorp-
tion is corrected for (p. 128), interferences in flame AAS are seldom suffi-
124

ciently troublesome to warrant use of solvent extraction for those elements


with adequate sensitivity for their direct determination in acid extracts.
Viets (1978) has described a solvent extraction scheme using tricaprylyl-
methylammonium chloride (Aliquat-336) for rapid determination of Ag, Bi,
Cd, Cu, Pb, and Zn after their extraction with the sulphide selective potassi-
um chlorate/hydrochloric acid leach.
Analysis of waters by AAS has been reviewed by Ediger (1973). Trace
metals at parts per billion (ppb) concentrations are not usually directly
detectable in the flame and must therefore be concentrated. The simplest
methods of concentration are evaporation or solvent extraction. Extraction
with APDC*-MIBK is suitable for many elements (Mulford, 1966) and has
been used by Brown et al. (1970) to concentrate Ag, Cd, Co, Cr, Cu, Ni, and
Pb from samples acidified to pH 2.8. However, Kinrade and Van Loon
(1974) found extraction with APDC-DDDC **-MIBK to be more effective
over a wider pH range (Fig. 6-8). With a citrate buffer at pH 5.0 extraction
of Ag, Cd, Co, Fe, Ni, Pb and Zn is quantitative. References to many other
solvent extraction schemes are to be found in Ediger (1973) and the Annual
Reports of Analytical Atomic Spectroscopy. Careful cleaning of storage
bottles and reagents is necessary to free them from trace metal contaminants
at ppb concentrations.
Without resorting to solvent extraction, sensitivity for some elements can
be substantially increased by adjusting bulk composition of the solution to
ensure that formation of non-absorbing species, for example ions and refrac-
tory oxides, is minimized. At the same time the amount of solution reaching
the flame can be increased. For example, Sen Gupta (1976) and Ooghe and
Verbeek (1974) increased sensitivity for rare earth elements by suppressing
their ionization in the nitrous oxide/acetylene flame with an alkali metal
buffer. A further increase in sensitivity was obtained by aspiration of an
ethanolic, rather than an aqueous solution, to improve overall atomization
efficiency.
A number of elements, notably Zr, Hf, Ta and Ti, have very poor sensitiv-
ities even in the nitrous oxide/acetylene flame because of the formation of
their refractory oxides. All these elements also form strong fluoride com-
plexes, with much lower melting points than the oxides, and Bond (1970)
reported up to a tenfold increase of their sensitivity in the presence of
ammonium fluoride. This is presumably due to the formation, volatilization
and direct atomization of the fluoride complexes without formation of the
refractory oxide. Similarly, Husler (1972) obtained a marked increase in
sensitivity for Nb in the presence of hydrofluoric acid and an alkali-metal
buffer.

* Ammonium pyrrolidine dithiocarbamate


** Diethylammonium diethyldithiocarbamate
125

Cd scale = 2.5
Α
2.0 η —Α Α
9

1 · — · Cd
I D—D CO
1.6 A
j o o Cu

0 2 4 6 8 10 12 14
pH

Zn scale = 2.5

♦ — ♦ Ni
1.6 H
0—0 Pb
LU Δ—Δ Zn

0 2 4 6 8 10 12 14
PH

Fig. 6-8. Influence of pH on the extraction of metals into MIBK with 1% w/v APDC +
DDDC. (Reprinted with permission from Kinrade and Van Loon, 1974, Solvent extrac-
tion for use with flame atomic absorption. Analytical Chemistry, 46: 1894—1898. ©
1974 American Chemical Society.)

Interferences
Freedom from interferences is often cited as one of the principal advan-
tages of AAS. Nevertheless, if samples and standard solutions differ in bulk
composition a variety of interferences can cause enhancement or depression
of the analyte signal. Principal sources of interferences and some remedial
methods are summarized in Table 6-V.
Most interferences could be overcome either by careful matching of
sample and standard solutions, or by use of the method of additions. Alter-
natively, the analyte could be isolated from interferents by solvent extrac-
tion or some other separatory technique. For example, prior to analysis of
126

TABLE 6-V
Some interferences in the determination of trace elements in geological matrices by flame
atomic absorption spectrophotometry

Element Inter- Comments Reference


ference

Ag B background correction Minkkinen (1975)


Ba* C, I suppression by Si, Al and Marutaetal. (1972)
P — add oxine or NH4CI;
enhancement by alkalies and Cioni et al. (1976a)
alkali earths — add
K radiation buffer Miller and Cazalet (1979)
Be* C suppression by Al Terashima (1973)
Cd, Co B background correction Fletcher (1970)
Cr C suppression by Fe, Na and K;
enhancement by AI, Mg, Ca —
interference and sensitivity
reduced in Ν 2 0 / ϋ 2 Η 2 flame
Mo* C suppression by alkalies, Ca, Van Loon (1972)
Fe — add up to 1000 Mg/ml Al Sutcliffe (1976)
Ni, Pb B background correction Fletcher (1970)
Rb I add K radiation buffer
Sr* C, I add K radiation buffer and La Carter et al. (1975)
releasing agent
v* C suppression by Fe and Ti — Terashima (1973)
add 1000Mg/ml Al
Zr* C add NH 4 F Bond (1970)
1
Determined in lean air/acetylene flame except (*) in nitrous oxide/acetylene flame.
2
B = background absorption; C = chemical (suppression); I = ionization (enhancement).

Fe-rich solutions, produced by acid decomposition of magnetite or pyrite,


ferric iron can be extracted into MIBK from 6 M HCl leaving many trace
elements in the aqueous phase (Nakagawa, 1975; Sinclair et al., 1977). How-
ever, such procedures are generally too time-consuming for routine analysis
and it is, therefore, fortunate that for prospecting purposes analytical inter-
ferences are normally within tolerable limits.

Spectral interferences. Spectral interferences arise when absorbing lines of


other elements coincide or overlap with the absorption line of the analyte.
Because of the simplicity of absorption spectra and the purity of the spec-
trum emitted by a hollow cathode lamp, such interferences are seldom
encountered. One of the few cases that might be encountered in routine
127

analysis of geological matrices is the overlap of Fe 213.859 nm on Zn


213.856 nm causing high absorption readings for Zn in Fe-rich materials.
Kelly and Moore (1973) found that 0.2-2.0% Fe in solution gave apparent
Zn concentrations ranging from 0.031 to 0.26 Mg/g.

Chemical interferences. As the solvent evaporates in passing from the spray


chamber into the lower part of the flame a condensed phase consisting of
solid clotlets is formed. If the composition of the sample results in formation
of refractory compounds which do not dissociate (or only dissociate above
the zone illuminated by the light source) the population of absorbing atoms
and hence the absorption signal will be correspondingly reduced. This is
particularly significant when silicates are analyzed for Ca or Mg in the air/
acetylene flame (Fig. 7-10) and is also well illustrated by the severe depres-
sion of Mo readings in the presence of Ca in the nitrous oxide/acetylene
flame, probably due to formation of calcium molybdate (Fig. 6-9). Addition
of Al which preferentially combines with the Ca and acts as a releasing agent
for the Mo, minimizes the interference. Van Loon (1972) recommends addi-
tion of sufficient Al to give a final concentration in solution of 1000 Mg/ml
for a 0.3-g sample.

lonization interferences. lonization interferences are associated with the


determination of the alkali metals and, in the hotter nitrous oxide/acetylene
flame, the alkali earths. The alkalies have relatively low ionization potentials
and are appreciably ionized even in cool flames. If another alkali is added to
a solution already containing an alkali, its ionization contributes to the total

0.5Ί

g ,
< /
er °· 3 Ί 7

°
CO /
/
CD ' 10 uq/ml Mo
< 02J /
1 / 10^ig/ml Mo +
/ 1 0 0 p g / m l Ca
/

1 1 1 1 1
200 400 600 800 1000
ALUMINIUM, ^g/ml

Fig. 6-9. Interference of Ca in the determination of Mo in the nitrous oxide/acetylene


flame, and the influence of varying concentrations of Al. At approximately 1000 Mg/ml
Al the suppression due to Ca is almost eliminated.
128

electron population in the flame. This results in partial suppression of ioniza-


tion for the first alkali and an increase in the number of its ground-state
atoms causing an increase in the absorption signal. The effect of varying con-
centrations of alkalies in sample solutions can be overcome and sensitivity
increased by addition of an alkali metal ionization-suppression buffer, for
example, a large excess of Cs in the determination of Rb or Sr. Enhanced
absorption by the Ba 553.6-nm atom line due to ionization suppression by
increasing concentrations of K and Ca is shown in Fig. 6-10: there is a corre-
sponding reduction of absorption at the 455.4-nm ion line. However, in the
case of Ba the interactions between chemical and ionization interferences are
so complex that effective correction with geological matrices may be
impossible (Cioni et al., 1976a). Despite this, Miller and Cazalet (1979)
reported obtaining satisfactory results for prospecting purposes using the
455.4-nm line and addition of K (1000 Mg/ml) to standards. Use of this line
also avoids increased signal noise associated with the coincidence of a Ca
bandhead with the atom line (Rooney and Woolley, 1978).

Background interference. Background absorption is a special case of spectral


interference resulting from absorption by molecular species in the flame and
possibly also by light scattering by solid particles. It has been described in
detail by Billings (1965) and Koirtyohann and Pickett (1966). Unlike the
sharp absorption lines resulting from atomic absorption, background absorp-

0.15 0.15
A 40ppm Ba + 10,000 ppm K
• 40 ppm Ba + 2,000 ppm Ca

o
z
<
CO

o
c/>
CO
< 0.05 -4 0.05 H

1000 2000 3000


CONCENTRATION OF INTERFERENT, }jg/ml CONCENTRATION
OF ALUMINIUM, jig/ml x103

Fig. 6-10. Determination of Ba in the nitrous oxide/acetylene flame. A. Effects of various


elements on the absorbance of a solution containing 40 ppm Ba at the 553.4-nm atom
line: the dashed line shows the effect due to addition of Ca at the 455.4 nm ion line. B.
Effect of Al additions on absorption by Ba-K and Ba-Ca mixtures at the 553.4-nm line
(From Cioni et al., 1976a.)
129

tion is a broad band phenomenon the absorption being additive on any


atomic absorption due to the analyte.
Fletcher (1970) and Foster (1971) have shown that background absorp-
tion can give large errors when low concentrations of an element are to be
measured in geochemical samples. For example, Fletcher (1970) found that
a Pb-free synthetic granodiorite gave an absorption equivalent to 25 ppm Pb
(Table 6-VI). On real samples background interference was particularly
severe for Pb, Co and Ni, but not Cu and Zn. Ag is also affected and, since
Ag values of only a few parts per million may be very significant, spurious
geochemical patterns and anomalies can easily be created. Minkkinen (1975)
found that Ag-free solutions containing 20,000 ppm Ca gave readings equiva-
lent to 4 ppm Ag.
Background absorption is best corrected by its measurement with a con-
tinuum light source, either a hydrogen or deuterium lamp in the UV region,
and then subtracting it from total absorption, measured with the hollow
cathode lamp, to give a corrected value (Fig. 6-11). This can be done
automatically on many instruments and is recommended for routine deter-
mination of Ag and Pb. If a continuum lamp is unavailable background
absorption can still be measured if the light source emits a suitable non-
absorbing line close to the absorption line. For example, background absorp-
tion at Pb 216.99 nm can be estimated using Pb 220.35 nm.
Background absorption is additive and the error, on a percentage basis, is
therefore greatest at low concentrations. At high concentrations it may be
insignificant. Furthermore, if background absorption and a suppressive inter-
ference are both present, background correction can give the best results at
low concentrations but cause over correction at high values (Fig. 6-12).
Fortunately, in exploration geochemistry negative errors at high concentra-
tions are usually less serious than large positive errors at background levels.

TABLE 6-VI
Correction for background absorption in lead analysis of a synthetic rock solution (from
Fletcher, 1970)

Pb added Absorbance Pb found (ppm) 1

(ppm) *
cathode lamp H 2 lamp corrected uncorrected corrected

0.0 0.015 0.015 0.000 24.8 0.0


20.0 0.025 0.015 0.010 47.7 21.0
50.0 0.041 0.015 0.026 82.1 53.5
100.0 0.063 0.016 0.047 122.2 95.5
200.0 0.111 0.015 0.096 217.7 190.9

Dilution factor = 50
130

SOURCE FLAME DETECTOR READOUT

L.JbJ
100 100 100
A% l% A%
Fig. 6-11. Background correction using a continuum source. The profile of the atomic
absorption line is shown in solid black and the continuum is represented, over a slit width
of Δλ, as the area within the dashed line. At the light source the emission intensity of the
atomic line and continuum are balanced at 100%. In the flame atomic absorption is super-
imposed on background absorption so that the intensities reaching the detector are Ia and
Ib) respectively. The readout displays these intensities as a net signal equal to (100 — Ia) —
(100 -Ib).

100 —Φ— U n c o r r e c t e d , Pb line 2170.5 Ä

--O— Corrected on H-famp


90 ---a—- Corrected on Pb non-absorbing
line 2 2 0 4 Ä

80

70

60

50

40

30

20

10

0\- O^

-10
4 6
■r-ψ-^-ψ: 12
Θ 10
ΡΡΦ Pb in solution

Fig. 6-12. Errors in uncorrected and background-corrected measurement of Pb in the


presence of 4000 ppm Fe. (From Govett and Whitehead, 1973.)
131

Electrothermal atomization

For those elements readily determined in the flame use of electrothermal


furnaces offers no advantages. Potential applications of electrothermal meth-
ods in the exploration laboratory are therefore largely confined to deter-
mination of elements with inadequate sensitivity in the flame and to deter-
mination of ppb concentrations in weak partial extracts of soils and sedi-
ments or in natural waters. Use of electrothermal atomization in these situa-
tions should be carefully weighed against the merits of solvent extraction
and conventional flame atomization — particularly if sampling criteria
require decomposition of relatively large sub-samples anyway.
Geochemical matrices are usually analyzed by injection of an aliquot of
sample solution, although direct atomization from a few milligrams of finely
ground powder is also possible (Langmhyr et al., 1974; Langmhyr, 1977)
and Friedrich et al. (1973) injected an ethanolic suspension to determine Cu
in soils and sediments. When acid decomposition is employed, nitric acid is
preferable to a halogen acid to avoid losses by direct volatilization of the un-
dissociated metal halide: with oxy-acids atomization proceeds via the rela-
tively less volatile metal oxide. Loss of analyte can also occur, especially
with the more volatile elements, if temperatures are too high during sample
charring (Fig. 6-7). Ediger (1975) reduced such losses by matrix modifica-
tion. For example, addition of Ni retards loss of Se and Te up to 1200° C
(probably due to formation of nickel selenides and tellurides) and has been
applied to determination of Se in sediments (Martin et al., 1975).
Compared to flames, non-specific absorption, due to smoke and high atom
densities, is more severe with electrothermal devices and requirements for
correction of background absorption are correspondingly more stringent.
Other types of interference also occur in analysis of geochemical matrices.
However, the processes and factors involved are not well understood. In a
study of determination of Pb in carbonates, Campbell and Ottaway (1974,
1975) found that Ca depressed the analyte signal in hydrochloric but not in
nitric acid. A similar phenomenon has been reported by Sighinolfi (1972) in
the determination of Be, and Cruz and Van Loon (1974) attributed the
depressant effect of Ca on Cd, Co, Cu, Ni, Pb and Zn to their occlusion
within a refractory matrix formed in the furnace. Hutton et al. (1977)
suppressed ionization interferences in the determination of Ba in limestones
by addition of Cs. Interference problems could also be avoided by solvent
extraction. This, however, would seem to negate the advantages of the great
sensitivity of electrothermal atomization except for a few elements such as
Au (Bratzel et al., 1972; Meier, 1980), with such low concentrations that
adequate sensitivity can sometimes only be achieved by combining furnace
AAS with solvent extraction.
Examples of the applications of flameless AAS to geochemical analysis, in
addition to those already mentioned, include: Ba (Cioni et al., 1976b): Bi
132

(Ficklin and Ward, 1976); Cd (Gong and Suhr, 1976); Te (Corbett and
Godbeer, 1977; Chao et al., 1978); Mo (Kontas, 1976); and precious metals
(Fryer and Kerrich, 1978). Other references and a discussion of the deter-
mination of individual elements are to be found in Fuller (1977).
In marked contrast to analysis of solids, a rapidly expanding literature on
application of flameless AAS to water samples has developed in response to
environmental problems. Natural fresh waters contain a wide range of dis-
solved ions and constitute a complex, non-uniform analytical matrix. Conse-
quently, background absorption and inter-element interferences, although
less severe than with solids, are still present. Nevertheless, multi-element
studies by Edmunds et al. (1973) and Barnard and Fishman (1973) have
shown that for rapid screening and anomaly detection, direct injection and
comparison with standard calibration curves could be adequate if accom-
panied by simultaneous background correction. Fordham (1978) obtained
more accurate results by the method of additions: however, only sixty deter-
minations per day were possible.

Hydride generation

Hydride generation combined with AAS (or ICP-ES, p. 164) provides a


rapid and very sensitive method for determination of elements forming
gaseous hydrides. It has been applied to analysis of geochemical samples,
including waters, for As, Sb, Se, Sn, and Te and may also be applicable to
estimation of Ge and Bi — all elements difficult to determine by conven-
tional flame methods.
Initial studies of hydride generation-AAS involved reduction of As 3+ to
arsine (AsH 3 ) in zinc-stannous chloride/hydrochloric acid reduction cells
(Fernandez and Manning, 1 9 7 1 ; Dalton and Malanoski, 1971). The arsine
was then swept, either directly or after collection in a reservoir, into the
transparent argon/hydrogen or nitrogen/hydrogen flames. Subsequently,
hydrides of As, Bi, Sb, Se, Sn and Te were all generated more conveniently
from acidic sample solutions using sodium borohydride, either as pellets or a
stable alkaline solution, as a reductant (Fernandez, 1973; Thompson and
Thomerson, 1974). Precise control of the acidity of the sample solution is
not critical except in the determination of Sn.
Rapid quantitative evolution of the hydrides only proceeds from the
lowest valency state of the elements. Se 6+ and Te 4 + can be reduced to their
trivalent states by warming in hydrochloric acid or directly by the action of
the borohydride. If, however, arsine and stibine are to be evolved from
weakly acidic solutions, sodium or potassium iodide must be added as pre-
reductants to ensure quantitative reduction.
Sensitivities attainable by decomposing the hydrides, both in flames and
in resistance or flame heated cells, are summarized in Table 6-VII. These
values are very much dependent on generator design, reagent strengths, solu-
Application of hydride generation to analysis of geochemical samples by atomic absorption spectrophotometry
1
Element Decomposition Generator Atomizer Detection limit Comments Reference

As aqua regia 2% NaBH 4 flame heated 0.16 ppm on interference from Ni 1


HF/HCIO4/HNO3 solution quartz cell 0.5-g sample prevented by addition of
6 M HC1 0.01 M EDTA
As HNO3/HCIO4 1% NaBH 4 resistance 0.25 ppm on semi-automated Technicon 2
solution heated 0.1 g sample system for analysis of
quartz cell soil and vegetation
As 5% NaBH 4 N2/H2/air modification for rapid 3
flame analysis of exploration
samples

As HC1 1% KI/ resistance 10 ng/ml in KI added as a pre-reduc- 4


1% NaBH 4 heated solution tant; rapid technique for
HNO3/HCIO4 quartz cell analysis of soils, waters
and vegetation
As HF/HCIO4/ KI/SnCl 2 /Zn Ar/H 2 flame 0.04 ppm for K M n 0 4 added to avoid 5
HN03/KMn04 1 g sample volatilization losses;
interference from Pb
minimized by addition
Fe3+
As HN03/H2S04 2% NaBH 4 resistance 0.001 ^g/ml no interferences from Cd, 7
heated silica Cu, Pb, Zn, Mg, Ca, Fe
tube and Al at concentrations
found in soil samples
Sb Aqua regia 2% NaBH 4 flame heated 0.08 ppm on interference from Ni and 1
HF/HCIO4/HNO3 quartz cell 0.05-g sample Ag reduced by addition of
6MHC1 0.01 M EDTA
Te HF/HCIO4/HNO3 5% NaBH 4 resistance 5 ppb on 0.25-g no serious interferences 6
heated quartz sample for most samples: up to
cell 2 0 0 ppm Cu can be
tolerated 00
GO
1 :
References: 1 = Aslin (1976); 2 = Vijan et al. (1976); 3 = Kokot (1976); 4 Wauchope (1976); 5 = Terashima (1976); 6 = Green-
land and Campbell (1976); 7 : = Thompson and Thoresby (1977).
134

tion volumes and gas flow rates. Greenland and Campbell (1976), deter-
mining Te in rocks, found it difficult to maintain maximum sensitivity
without constant attention to the equipment. They, therefore, opted to use
less sensitive conditions for routine analysis: even so, as little as 5 ppb Te
could be detected with a 0.25-g sample.
As usual, early reports on the absence of interferences proved illusory.
Interferences reported by various workers do, however, differ somewhat,
apparently depending on reagent strengths and volumes and order of mixing.
In a detailed study, Smith (1975) generated the hydrides of As, Bi, Ge, Sb,
Se and Te from dilute hydrochloric acid with sodium borohydride. Using an
argon/hydrogen flame he found: Cu, Ag, Au, Pt, Pd, Rh, Ru, Ni and Co
always suppressed hydride generation — probably due to their preferential
reduction and precipitation, resulting in coprecipitation or adsorption of the
analyte, and consumption of reductant; and the hydride-forming elements
mutually interfere. Belcher et al. (1975) found that suppression of the evolu-
tion of arsine and stibine, from 0.1 M hydrochloric acid, by Co, Ni, Zn, Fe,
Bi, Cd and Cu could be almost eliminated by addition of EDTA. In more
strongly acidic media (5 M HC1), Aggett and Aspell (1976) found no inter-
ference from up to a 10,000-fold excess of Cu on the evolution of arsine.
Problems with high reagent blanks for As, Sn and Sb in sodium borohydride
have been reported.
Rapid methods of analysis of geochemical samples for As, Sb, Se and Te
have been described (Table 6-VII). With hot acid decompositions it is neces-
sary to prevent volatilization of As and Se by maintaining them in their
highest valency state, which must then be reduced to the trivalent state for
quantitative hydride generation. Large amounts of nitric acid remaining after
decomposition cannot be tolerated and should be removed by evaporation
with perchloric acid. Interference studies suggest that Fe, Cu, Ni, Mn and Co
are most likely to be troublesome in geological materials. Aslin (1976) found
that Ni suppressed evolution of arsine from 1.5 M hydrochloric acid and
both Ni and Ag suppressed stibine; the interferences could, however, be con-
trolled by addition of EDTA and it then became possible to determine
2 ppm Sb in the sulphide standard SU-1. Thompson and Thoresby (1977)
and Thompson et al. (1978b) found no significant interferences in the evolu-
tion of arsine from 5 M hydrochloric acid. However, the latter authors did
find it necessary to avoid interference from Cu in the hydride generation of
Bi, Se and Te by their coprecipitation and separation on lanthanum
hydroxide. Greenland and Campbell (1976) also reported interference from
Cu, in excess of 200 ppm, in the determination of Te.

Determination of mercury by flameless AAS

Estimation of Hg by flameless AAS predates conventional AAS by many


years and concern over environmental concentrations of Hg and its use as a
135

pathfinder for sulphide deposits have produced an extensive literature on its


determination in natural media. Ure (1975), for example, cites more than
400 papers on determination of Hg by flameless AAS after its release from
samples either by pyrolysis or acid digestion followed by reduction in solu-
tion to elemental Hg.
Pyrolytic methods can be used to determine either total Hg or, by con-
trolling the heating rate, to determine different forms of Hg and its parti-
tioning between different minerals (Koksoy et al., 1967; Watling et al.,
1973). Pyrolysis, however, has the disadvantage that geochemical samples,
especially those rich in organic matter or sulphides, produce smoke or S0 2
causing non-specific absorption at the 253.7-nm Hg line. With conventional
atomic absorption instruments, background correction with a continuum
source can be used as already described. Alternative correction methods
include:
(1) Measurement of non-specific absorption on the edges of the 253.7-nm
line using either (a) the wings of a pressure broadened, self-absorbed line
produced in a Hg-saturated cell as shown in Fig. 6-13 (Barringer, 1966); or
(b) the Zeeman effect to split the absorption line (Hadeishi and McLaughlin,
1966;Robbins, 1973).
(2) Dividing the gas flow from the heated sample into two streams (Fig. 6-
14), one of which passes through a PdCl2 filter to remove Hg, before entering
the reference cell of a double-beam instrument (James and Webb, 1964). A
similar method has been applied to recirculating single-beam systems
(Windham, 1972).
(3) Collection of Hg liberated during pyrolysis by amalgamation on Au or
Ag, from which it is subsequently re-released by heating (Vaughn and
McCarthy, 1964; Vaughn, 1976; Marinenko et al., 1972). Organic com-
pounds condensing on the Au trap can be troublesome, either due to their
occlusion of Hg or their re-release on heating causing background absorption.
Azzaria and Webber (1969) and Weissberg (1971) prevented condensation of
organics by maintaining the trap at 150—170° C. As an alternative to
amalgamation, Hg released by pyrolysis can be collected in a potassium
permanganate solution and subsequently determined by the cold vapour
method described next.
The cold vapour method is readily adaptable for use on any conventional
atomic absorption spectrometer (Fig. 6-15). In the original procedure,
described by Hatch and Ott (1968), Hg was liberated by the action of sul-
phuric acid, hydrogen peroxide and potassium permanganate on powdered
rock samples. After cooling, excess permanganate and precipitated manganese
oxides were reduced with hydroxylamine sulphate. Hg is then reduced to the
elemental form with stannous sulphate and flushed into an absorption cell
by a stream of air. Subsequently, many variations of the cold vapour method
and a wide variety of decomposition mixtures have been described (Ure,
1975). Some workers have combined cold vapour and amalgamation meth-
136

A it

I Differential
I Amplifier

Fig. 6-13. Design of a spectrometer using pressure broadening of the 253.7-nm Hg line to
correct for non-atomic absorption in the determination of Hg. A pressure-broadened Hg
spectrum, emitted by a low-pressure mercury lamp, passes through an absorption
chamber and is then split into two beams, one of which falls directly on a photomultiplier
(PMA). The other beam passes through a cell saturated with mercury vapour so that the
central portion of the 253.7-nm Hg line is completely absorbed leaving only the wings of
the line to reach the reference photomultiplier (PMR). When a vapour causing non-atomic
absorption is pumped into the absorption chamber, output from both PMA and PMR is
proportionally reduced: atomic absorption due to Hg only reduces the signal fromPM A .
(Modified from Barringer, 1966.)

ods (Head and Nicholson, 1973; Huffman et al., 1972).


Providing potassium permanganate is not used in the digestion solution
Hg can be reduced directly with Sn11 without addition of the hydroxyl-
amine. Loss of Hg in the absence of permanganate has, however, been
reported (Iskander et al., 1972). Interferences arise, particularly if the Hg is
recycled through the absorption cell and aeration vessel, from elements
which are also reduced to elemental form (Au, Pt, Pd, Te) and then
amalgamate with Hg (Jonasson et al., 1973). Interferences from Cu and Ag,
reported by Band and Wilkinson (1972) in a chloride-rich system using
hydroxylamine hydrochloride, are avoided by reduction with stannous sul-
phate (Jonasson, 1974).

Indirect determinations of elements by AAS

A large number of elements are either not determinate by AAS or are


only determinable at sensitivities too low to be of practical value. In some
137

Ultra-violet
(?fi Lamp
/\
Photocell Photocell

hm. ^m. E&.


Suction

4 pump

Glass w o o l '
mm m W^A
Palladium chloride
on glass wool

Silica gel

tm
Fig. 6-14. A double-beam mercury vapour meter. The sample stream is split into two
halves, one of which passes through glass wool loaded with palladium chloride, to remove
Hg but not constituents causing non-atomic absorption, before entering the reference
absorption cell on the right side of the instrument. Output from the two photomultipliers
is compared. (From James and Webb, 1964.)

Infra-Red
Heater Lamp
" I I I i \ | | | I i I \ \ V

Hg°
(U Absorption Cell

Quartz Window

Air

Gas Washing

Bottle

: Ä ■

ym
Fig. 6-15. Cold vapour generation and determination of Hg by AAS. In the gas washing
bottle Hg 2+ is reduced to Hg° with Sn 2+ and then swept into the absorption cell by a
stream of air. The IR heater lamp prevents condensation of water vapour in the cell.
138

cases these elements can be determined by indirect methods based on either


changes in the absorption signal of an element after its reaction with the
analyte; or the formation of complexes containing the analyte and an ele-
ment readily determined by atomic absorption.
Kirkbright and Johnson (1973) have reviewed indirect methods for some
16 elements. One of the few geochemical applications is the determination
of sulphate in soils, by measurement of excess Ba after precipitation of
barium sulphate (Varley and Chin, 1970). The author, following Kirkbright
et al. (1967), has determined phosphate by extraction of the phosphomolyb-
dovanadic acid complex, used for colorimetric estimation of phosphate
(Peachey et al., 1973), into MIBK and measurement of the Mo content of
the organic phase by AAS. The 11 : 1 ratio of Mo to P in the acid complex
makes the method extremely sensitive. Similar amplification schemes, based
on formation of molybdenum-heteropoly compounds for As, Ge, Si, V, Nb,
Th, and Ti, are summarized by Kirkbright and Johnson (1973).
Chapter 7

EMISSION SPECTROSCOPY

INTRODUCTION

Rapid, virtually simultaneous multi-element analysis with minimal sample


pre-treatment offers obvious advantages, particularly for regional recon-
naissance programmes in which it is desirable to determine as wide a range of
elements as possible, and has always had considerable appeal to exploration
geochemists. Hence, despite greater capital cost and complexity of equip-
ment compared to single element methods, there has been a continuing inter-
est in emission techniques based on sample excitation in DC-arc and, more
recently, inductively coupled plasma (ICP) sources. The spectra emitted by
the excited atoms can either be recorded photographically, or measured
instantaneously with photomultipliers in direct reading spectrometers.
To obtain quantitative results with photographically recorded spectral
lines their intensities must be measured using a densitometer and corrected
for emulsion characteristics. This is generally too time-consuming for routine
analysis. Concentrations are therefore usually estimated by visual com-
parison to standard films prepared under identical conditions. This, together
with the inherent noisiness of the DC-arc and the sampling problems asso-
ciated with the small quantities of material (10—50 mg) taken for analysis,
makes it difficult to obtain results that are better than semi-quantitative.
Nevertheless, such results can still be of considerable value, particularly for
those elements that are either associated with resistant minerals or have
inadequate sensitivity for determination by flame AAS — for example Be,
Nb, Zr, W, Sn, Sb (Figs. 1-9 and 1-10).
With suitable internal standards and matrix corrections based on simulta-
neous determination of the bulk chemical composition of samples, the DC-
arc direct reading spectrometer can provide more reliable results than semi-
quantitative photographic procedures. Providing there is no objection to the
additional effort required to decompose samples and analyze solutions, the
ICP holds considerable promise for further improvements, particularly with
respect to reproducibility and freedom from interferences. However, the
direct-reading spectrometer, with either a DC-arc or ICP source, is a sophisti-
cated instrument. Where ready access to a serviceman and spare parts cannot
140

be assured, the rugged character and simplicity of photographic spectrographs


can have considerable advantages.
Anyone about t o embark on emission spectroscopy of geochemical sam-
ples for t h e first time should consult the excellent texts by Ahrens and
Taylor (1961) and Mitchell (1964b).

THEORY

The theoretical basis of atomic emission and absorption was briefly


reviewed in Chapter 6 and the relationships between excitation states and
atomic spectra illustrated for the gold atom (Fig. 6-2).
Temperature in the DC-arc is closely related t o the ionization potentials
of the elements present in the arc. With a pure carbon arc in air, tempera-
tures are as high as 7000—8000° K. However, the presence of only a few per-
cent of a readily ionized element, such as K or Na, can lower the tempera-
ture by as much as 4000° K. Nevertheless, even in the presence of alkali ele-
ments, arc temperatures are still significantly higher than those encountered
in the flames used for AAS. The proportion of atoms excited to energy levels
above the groundstate is correspondingly increased (Table 6-1). The arc spec-
trum of an element is therefore much richer in lines than the flame spec-
trum, the difference being most notable with elements, such as Fe and Mn,
for which a very large number of transitions are possible between energy
levels above the groundstate. As a result of the almost inevitable presence of
several percent Fe, geochemical samples emit DC-arc spectra of great com-
plexity in which several thousand Fe lines can be present. Analysis of such
spectra requires a spectrograph with good resolving power and careful choice
of analytical lines to avoid spectral interferences.
Apart from increasing the proportion of atoms excited t o energy levels
above their groundstate, arc temperatures also increase the extent of ioniza-
tion compared t o flames. The relationship between temperature (T) in °K,
the apparent ionization potential (V) of an element, and t h e degree of
ionization (x) being expressed by the Saha equation:

x 5040 _
log = -logp e + 2.5 log T - — - - V - 6 . 1 8
1 —x 1
where log pe, the electron pressure, has a value of 10" 4 t o 10~ 3 atmospheres
in the carbon arc (Boumans, 1966).
The difference between t h e ionization potential (V) and the apparent
ionization potential (V) is small and need not concern us for the purpose of
illustrating the relative effects of temperature and ionization potential on the
degree of ionization (Table 7-1). It is apparent from the Saha equation that
for many elements both atom and ion lines will be present in their arc spec-
141

TABLE 7-1
Percent ionization of Na, Ca, Cu and Zn at different temperatures in the DC carbon arc.
Calculated using the Saha equation assuming an electron pressure of 4.1 X 10~4 atm

Element V1 V2 Percent ionization

4000°K 5000°K 6000°K

Na 5.14 5.49 16 89 99
Ca 6.11 5.82 7 79 99
Cu 7.72 8.13 <1 2 40
Zn 9.39 9.06 <1 <1 10

V = ionization potential.
V = apparent ionization potential from Boumans (1966).

trum and for those elements that are relatively readily ionized there will be
a pronounced weakening of their atomic spectrum. In the ICP, which oper-
ates at temperatures above 5000° K, ion lines can assume considerable
analytical importance (see Fig. A-2, Appendix 1). It should be noted that
when an atom is ionized by loss of one electron, the resulting ion spectrum
will resemble the atomic emission spectrum of the element whose atomic
number is one less.
In the DC-arc the sample powder is volatilized directly. Sample composi-
tion, particularly the content of readily ionized elements, therefore has a
considerable influence on arc temperatures, on the rates of volatilization of
individual elements, and on the distribution of excitation states in the
resulting atom and ion populations. Furthermore, conditions in the arc
column will progressively change as the relatively more volatile elements are
lost first, by selective volatilization, and arc temperatures rises (Fig. 7-3).
The proportions of emitting species can therefore vary considerably, both
between samples and at different stages in the arcing process. In quantitative
analysis these variations must be controlled, with spectroscopic buffers, and
monitored and corrected for using internal standards (Fig. 2-6).

FLAME EMISSION SPECTROSCOPY

Flame emission spectroscopy is simply the application of the familiar


qualitative flame tests, used in mineral identification, to quantitative analysis.
The sample, in solution, is sprayed into a flame and the emission intensity of
a suitable line measured with a monochromator and photomultiplier.
Because flames are less energetic excitation sources than DC-arcs the emis-
sion spectra are much simpler, requiring a spectrograph of only moderate
resolving power, but the technique is largely limited to elements that are
142

readily excited — notably the alkalies and alkali-earths. Chemical and ioniza-
tion interferences are similar to those described for AAS.
Analysis by flame emission spectroscopy has almost entirely been replaced
by AAS and is seldom used in the exploration laboratory. However, it should
be remembered that most atomic absorption spectrophotometers are readily
converted to operate in the flame emission mode. Consequently, if no
hollow cathode lamp is available, quantitative determination of Li and other
alkalies is still possible providing precautions are taken to prevent chemical
and ionization interferences.

SEMI-QUANTITATIVE DC-ARC SPECTROSCOPY

Equipment
The basic components of a spectrographic system (Fig. 7-1) consist of an
excitation source, usually a DC-arc, and its power supply; the spectrograph
itself, with either a prism or diffraction grating; and a camera to record spec-

photomultipliers
behind exit slits
on focal circle

lens source

DIRECT READING SPECTROGRAPH WITH GRATING

source
Y

slit A

F7 \
—— . v
«^^OOnm
prism
— ■ — ^ 0 0 nm
camera or
plateholder

PHOTOGRAPHIC SPECTROGRAPH WITH PRISM

Fig. 7 - 1 . Schematic diagrams o f direct-reading and photographic spectrographs.


143

tral intensities on a photographic plate or film. Essential ancillary equip-


ment is a darkroom for film processing and some means of measuring or
comparing the line intensities recorded on the films. Holman and Durham
(1967) and Canney et al. (1957) have described mobile spectrographic
laboratories.
A grating or quartz prism spectrograph with a reciprocal linear dispersion
of 0.5—0.7 n m / m m provides adequate resolution and at the same time allows
a wavelength range of 220—500 nm to be covered with one exposure on a
25-cm plate or film strip. For visual comparison of line intensities against
standard films, choice of a grating or prism instrument is not critical. How-
ever, the constant linear dispersion obtained with gratings makes identifica-
tion of unknown lines somewhat easier.
The usual excitation source for analysis of geochemical samples is the DC-
arc with the lower, sample holding electrode, as the anode. Arc currents of
10—15 A are needed so that the power source should be capable of pro-
viding a minimum of 12 and preferably up to 30 A. A built-in high-voltage
spark circuit to initiate the arc by ionizing the air gap between electrodes is
a useful convenience. Alternatively, the arc can be struck by drawing an
insulated graphite rod between the electrodes. Cathode-layer excitation, in
which the lower electrode is the cathode, is discussed by Mitchell (1964b).
The spectrograph and source can be arranged optically so that the image
of the source is focussed either at the entrance slit of the spectrograph or at
the prism or grating (Fig. 7-2). The latter give uniform illumination of the
slit and uniform line intensities on the film, whereas with the former arrange-
ment variations of emission intensity in the arc column are reproduced by
variations in the line intensity at the camera. Either arrangement is suitable
for semi-quantitative analysis but uniform illumination of the slit is essential
for quantitative work and is also required if a step-sector or step-filter is to
be used to extend the usable concentration range of analytical lines — as in
the method of Nichol and Henderson-Hamilton (1965).
Uniform illumination of the slit can be achieved in several ways, the sim-
plest being to place a lens at the slit to focus the image of the source on the
internal collimating lens of the spectrograph (Fig. 7-2). With this arrange-
ment a diaphragm at the collimator can be used to select the segment of the
arc column desired for analysis or mask out the incandescent tips of the elec-
trodes. Alternatively, two spherical lens can be employed, one giving an
image of the source on an adjustable diaphragm and the other focussing the
portion of the image selected at the diaphragm on to the collimator.
For rapidity, and to economize on film, it should be possible to expose
several spectra across the width of the film. Although a Hartman diaphragm
mounted at the slit can be used for this purpose it is usually better to rack
the camera across the exit slit of the spectrograph. By adjusting the width of
the exit slit as many as 24 spectra can be obtained across the width of a
35-mm film.
144

Internal Spherical Screen and Source


Collimating Lens Adjustable
Lens Diaphragm

Fig. 7-2. External optical arrangements to give uniform illumination of the slit by
focussing the image of the source at the grating. In (A) a lens immediately in front of the
slit focusses the image of the source at the spectrograph's internal collimating lens; in (B)
focussing the image of the source on an adjustable diaphragm allows any portion of the
arc to be selected for analysis.

With practice visual comparison of line intensities becomes very rapid and
is entirely adequate for semi-quantitative analysis. Manual densitometric
measurement is both considerably slower and unjustified when the magni-
tude of errors arising from other sources is considered. The additional
expense of recording densitometers or microphotometers is therefore seldom
warranted and a simple projector comparator will suffice for estimating line
intensities. The comparator should provide a magnification of about 20X, to
avoid operator eyestrain, and be designed to permit rapid alignment of the
sample and reference film or plate.
Heiz (1973) and Walthall (1973) have automated plate reading using a
sophisticated high-speed scanning microphotometer capable of measuring
the intensity of up to 400 lines in only 70 seconds. In a semi-automatic
system, Tait and Coats (1976) transferred digitized densitometric readings
directly from a manually operated microphotometer to punch tape for com-
puter calculation of element concentrations and correction of interferences.
Analytical throughput was comparable to that with visual estimation but
results were more reliable. It seems probable tht attempts to automate plate
reading, where semi-quantitative visual estimation is regarded as inadequate,
will be superceded by the use of ICP direct-reading spectrometers.
145

Operating conditions

Apart from the optical arrangement of the spectrograph the main variables
to be considered are excitation conditions and choice of analytical lines. Five
semi-quantitative procedures used in analysis of exploration samples are
compared in Table 7-II. All five employ uniform illumination of the slit with
anode excitation and the electrodes are of similar shapes and sizes, except
that a hemispherical or conical cathode is used in methods 1 and 4 in an
attempt to reduce arc wandering. Other workers have apparently found flat
cathodes equally satisfactory.
Principal differences between the methods are the duration of excitation,
ranging from 5 seconds to total consumption of the sample at 120 seconds,
and the use or absence of spectroscopic buffers. Total consumption of the
sample has the advantage of eliminating variations in line intensity caused by
different rates of element volatilization in samples of different bulk composi-
tion. On the other hand, background is increased and the line to background
ratio becomes less favourable for volatile elements that are released at an
early stage (Fig. 7-3).
Sample admixtures with graphite and buffers are used to improve and
maintain arc conditions. Addition of graphite alone, usually in a 1 : 1 to 1 :
4 ratio, results in a smoother burn and higher arc temperatures, thereby
reducing differences between volatilization rates of different elements and
increasing the intensity of lines of involatile elements. Avni et al. (1972)
found matrix effects, in standard rocks ranging from granite to peridotite,
were appreciably reduced when samples were mixed 1 : 5 with graphite and
only the central 2 mm of a 6-mm arc gap recorded. A disadvantage is greatly
increased cyanogen (CN) band emission caused by the combination of car-
bon and atmospheric nitrogen in the arc.
Addition of an alkali salt (L1CO3, NaCl, Na 2 C0 3 , NaF) as a spectroscopic
buffer, together with graphite, lowers arc temperature, enhances selective
volatilization and reduces CN emission. Still more important, differences in
bulk composition between samples are reduced and background variations
and matrix effects on line intensities are significantly diminished. For exam-
ple, with a lithium carbonate—carbon buffer Nichol and Henderson-Hamilton
(1965) found that matrix effects on analyte concentration were less than
twofold even between such compositional extremes as a siliceous soil and a
ferruginous laterite. In contrast, method 5, with no addition of graphite or
buffer, is subject to considerable matrix effects — though results are still
acceptable for many purposes. Tennant (1967) has described an A1203/
CaC0 3 /K 2 C0 3 ( 7 : 3 : 1 ) buffer which increases selective volatilization,
lowers general background and greatly increases sensitivity for volatile ele-
ments. Sensitivity for Mo is also improved (Delavault and Marshall, 1973).
To determine concentrations ranging from less than 10 ppm up to 1%
several analyte lines of differing sensitivities may have to be used for each
146

TABLE 7-II
Operating conditions for five DC-arc semi-quantitative methods of analysis of geochemi-
cal samples (all dimensions in millimetres unless otherwise indicated) 1
2
Reference Spectro- D Wave- Film/ N ζ Anode
graph (nm/mm) length plate
(nm) a b c
Myers et al. Wadsworth 225.0
0.524 4"X10" 38 68 4" ^ m w a ^
(1961) grating 475.0 United Carbon
3170
Nichol and Hilger and 280.0 6.15 2.91 5.00
2"X10" 12 14
Henderson- Watts E492 495.0 Ringsdorff
Hamilton RW402 Graphite
(1965)
Dorrzapf Ebert 230.0 y thin wall
0.500 4"X20" 68
(1973) 3.4 m 470.0 Ultra No. 3170
(Mark III)
Tait and Hilger 225.0 3.2 1.6 6.0
0.500 24 20
Coats 3.4 m 350.0 Johnson-Matthey
(1976) Ebert Spec-pure
ARL 232.0 35-mm
— 0.695 22 24 6.15 3.0 5.0
1.5 m 474.0 film
1
D = dispersion; N - number of spectra per film; Z = number of elements determined;
T = exposure time.
2
a: anode diameter, b: diameter of crater, c: depth of crater.

element. Alternatively, the upper concentration range of a sensitive line can


be extended by use of a step-sector or step-filter, mounted at the slit, to
reduce line intensity. The most sensitive lines suitable for analysis of geologi-
cal materials are summarized in Table 7-III, other compilations of useful
lines appear in Ahrens and Taylor (1961), Myers et al. (1961) and Heiz
(1965).
Choice of suitable lines can be restricted by their proximity to interfering
lines, the problem being most severe when the analytical line is absent or
weak and the interfering element is abundant. In some instances the theoret-
ically most sensitive line is unsuited to geochemical analysis because it is
invariably interfered with. For example, proximity to Si 252.85 precludes
use of Sb 252.85. In other cases the interference may only become signifi-
cant in samples of a particular composition. The interference of Mn 324.754
on Cu 324.754 in basic rocks is an example. If interference is suspected
other analytical lines should be checked and the presence of sufficiently high
concentrations of potential interferents confirmed by examination of several
of their analytical lines. Tables of spectral interferences are available
(Kroonen and Vader, 1963).
147

3
tode Sample Arc T Arc Slit Comments
charge gap (s) current (μιη)
e (mm) (A)

hemisphere S/C central 2—3 m m of arc


1 : 2 5-6 120 12-13 ? gap r e c o r d e d , with mask
at collimator
> S/C/L1CO3 3-step n e u t r a l density
3 20 12.5 15
gsdorff 2:1:1 filter at slit
4 0 3 Graphite

TM C-6) S/C 20 5 70% A r / 3 0 % 0 2


1 : 2 4 then 25 a t m o s p h e r e ; mask at
130 15 collimator
j conical S/C/SrF2 10 8 2-step r o t a t i n g sector;
nson-Matthey 1:1:1 2.5 th<en 8 arc image m a s k e d at
c-pure 70 15 diaphragm
sample only 3 5 15 20 n o buffer

c a t h o d e d i a m e t e r , e: shape,
imple: buffer r a t i o s .

A further restriction on choice of lines is CN band emission between


350.0 and 421.6 nm. The most sensitive lines of several elements lie in this
region and are only usable if CN emission is quenched by surrounding the
arc with a nitrogen-free atmosphere: a mixture of 20% oxygen and 80%
argon is often used for this purpose. Jets which surround the arc with a
sheath of oxygen-argon without the inconvenience of an enclosing chamber
have been described by Heltz (1964), Dorrzapf (1973), Timperley (1974)
and El-Kholy et al. (1975). Use of a jet also tends to stabilize the arc and
reduce self-absorption caused by relatively high densities of ground state
atoms in the cooler outer fringes of the arc. Despite these advantages, use
of a jet is probably not justified in routine semi-quantitative analysis.
The penultimate step in the analysis is the development of the photo-
graphic emulsion. Manufacturers of photographic materials produce special
emulsions, both as films and plates, for spectral analysis in the range 250—
450 nm. Their technical data should be consulted for choice of suitable
emulsions and development procedures. In order to obtain consistent results
over a long period, considerable attention must be given to standardizing and
maintaining all darkroom conditions: developer strength and temperature
148

Fig. 7-3. Volatilization curves for various metals, and variations of voltage, current and
electrode gap with a DC-arc. (From Timperley, 1974.)

are especially important.


Detailed discussion of photographic emulsion characteristics and their
calibration is not relevant to semi-quantitative analysis employing visual
comparison. It should, however, be noted that in its optimum working range
the photographic emulsion has a logarithmic response to changes in light
intensity. Consequently, as described in the next section, standard films for
semi-quantitative analysis are prepared with equal logarithmic, rather than
arithmetic, increments of analyte concentration.

Standards for semi-quantitative DC-arc spectroscopy

Preparation of standard films involves first the preparation of standard


trace element mixes in a suitable base and then exposure and development of
the films under identical operating conditions to those used for samples.
Standard mixes for semi-quantitative analysis are conveniently prepared
by mixing trace elements with a base to cover the range 1—10,000 ppm in
logarithmic intervals equal to the cube root of ten (i.e. 1, 2.2, 4.6, 10, 22, ...,
ppm) or approximately 1, 2, 5, 10, ..., 10,000 ppm. If the materials to be
analyzed all have similar bulk chemical compositions a trace element-free
base of the same composition can be prepared. Since, however, bulk com-
position of exploration samples varies considerably, a single general purpose
TABLE 7-III
Some suitable lines for semi-quantitative analysis of geochemical samples in the DC-arc

Element Wave- I2 Inter- Element Wave- I2 Inter-


1
length ference length * ference
(nm) (nm)

Ag 338.29 2 Mo 319.40 1 V
328.07 1 Mn 317.04 1 Fe
281.62 2
As 278.02 2
234.98 1 Cu Nb 313.08 1
228.81 1 309.44 1
295.09 2
Au 267.60 2 SiO
242.80 1 SiO Ni 349.30 1
341.48 1 Zr
B 249.77 1 Fe : SiO
249.68 2 SiO P 255.49 2 Fe
255.33 1 Fe
Ba 455.40 1
307.16 2 Pb 287.33 3
283.31 1
Be 332.13 2 266.32 2
313.04 3
234.86 1 Zr Sb 287.79 1 Cr
259.81 2
Bi 306.77 1 Fe
289.80 2 Mn Sn 326.23 2
317.50 1
Cd 346.62 2 Fe 284.00 3 Mn
326.11 3
228.80 1 Sr 460.73 1 Fe, Mn
Co 345.35 1 Ti 398.98 1
340.51 2 Ti 337.76 2
Cr 428.97 3 V 437.92 1
427.48 2 Ti 318.54 1
425.44 1 Ho
W 294.70
Cu 327.40 2
324.75 1 Mn Zn 334.50 Ti
282.43 3
Zr 339.20 Fe
Ga 294.36 1 Fe
2874.2
Li 323.26 1 Ti
Mn 403.45 2
280.11

3 Most frequently used wavelength in italics.


2
Relative intensities 1 > 2 > 3.
3
Interferences from Heiz (1965) and Ahrens and Taylor (1961).
150

base, with the composition of an intermediate igneous rock, is often pre-


pared.
In preparing the base it is possible to use either spectrographically pure
compounds, or minerals, rocks and soils with very low trace element con-
tents. If the latter are used their trace element content can generally be
reduced further by repeated leaching in hot hydrochloric or nitric acids and
washing with distilled water. Mitchell (1964b) prepared a synthetic soil base
by mixing together purified Si0 2 (63%), A1203 (20%), Fe 2 0 3 (5%), CaO (2%),
MgO (2%), Na 2 C0 3 (3.5%), K 2 S0 4 (3.5%), and Ti0 2 (1%), and sintering over-
night at 1250°C. Once prepared the base should be spectrographed and
approximate levels of any trace element impurities noted. Impurities in the
electrodes, particularly Cu and B, will also appear on the base film and fix
their lowest determinable concentrations.
Trace element mixes for dilution with the base can be prepared by com-
bining appropriate compounds, usually spec-pure oxides, with sufficient base
to give an initial concentration of 1% of each element. Two such mixtures
are given in Table 7-1V: other combinations and compounds can, of course,
be used. Standards are then prepared by successive dilutions of each mixture
with the base. If -%/TÖ logarithmic intervals are to be used standards and
base are successively mixed with 0.4642 as the dilution factor: i.e. at each
step 0.4642 g of the preceding standard is mixed with 0.5358 g of base. Very
thorough mixing is obviously essential. This is most likely to be achieved if
the base is added as a coarse powder and then ground to an impalpable
powder with the trace element mixture.
As an alternative to formulating trace element mixes, commercially avail-
able "cocktails" containing a large number of elements at a specified con-
centration can be used. Jarrel-Ash SQ powder for example contains 1.30% of
each of 45 elements. When diluted with a base, standards with concentra-
tions exceeding 1000 ppm will contain a large proportion of the trace ele-
ment mix. Resulting spectra can therefore be subject to unusual matrix and
interference effects and cannot be regarded as typical of geological materials.
Standard reference films are prepared by firing the standard mixes under
identical operating conditions to those used for unknowns. If possible the
spectra for the 13 standards, from 1 to 10,000 ppm, plus the base should be
arranged in sequence across the width of a film. Because of the poor repro-
ducibility of semi-quantitative methods it is not unusual for at least one of
the standards to give a poor spectrum and spoil the film. Method 5 of
Table 7-II is particularly prone to this problem. Under these circumstances a
better series of gradational intensities can often be obtained if reproducibil-
ity is improved by superimposing three spectra for each standard (i.e. loading
and firing three electrodes per standard without moving the film) with a
neutral density filter to reduce line intensities by 1/3 in the optical path. Un-
knowns are fired in the normal fashion without the filter.
Once a satisfactory reference film has been obtained it is worthwhile to
151

TABLE 7-IV
Composition of two standard mixtures to give 1% of each element

Mix A Mix B

compound g compound g

CuO 0.1252 NiO 0.1273


PbO 0.1077 C03O4 0.1362
ZnO 0.1245 Cr203 0.1461
CdO 0.1142 Mn203 0.1437
As203 0.1320 wo 3 0.1261
Sb203 0.1197 LiF 0.3738
Bi203 0.1115 Sn02 0.1270
Sum 0.8348 1.1802
Base 9.1652 8.8198

identify and mark all analytical lines. Alternatively line identification can be
simplified by construction of a set of cardboard or plastic wavelength guides,
showing both analytical lines and prominent major element lines, to fit the
comparator screen. Films are positioned against the wavelength guide by
aligning the most readily identifiable major element lines. Prominent lines
and groups of lines suitable for this purpose are shown in Table 7-V.
Occasionally unknown lines will have to be identified. This task is greatly
facilitated if films are available in which the pure spectra of individual ele-
ments alternate with the spectrum of a typical geochemical sample.

Analysis of unknowns

Preliminary sample preparation usually only involves sieving or grinding to


minus 80 mesh or finer, ensuring the powder is dry, and perhaps mixing with

TABLE 7-V
Some prominent lines in geochemical spectra

Element Wavelength (nm)

Mg 1 line 285.21
Si 1 line 288.16
Fe 3 lines 309.99, 310.03 310.07
CN bandhead 359.04
CN bandhead 388.34
Ca 2 lines 393.37, 396.85
Al 2 lines 394.40, 396.15
152

graphite or buffer. In some laboratories it is also the practice to ignite sam-


ples at 450—500° C in a muffle furnace. This dehydrates clay minerals,
destroys organic matter and helps to prevent evolved gases from ejecting the
sample from the electrode when the arc is struck. Sample and buffer can be
mixed in a small mortar but it is usually more convenient to use a high-speed
laboratory mixer mill adapted to hold several plastic vials, of 2 ml capacity,
into which the powders are weighed. At least one plastic ball is added to the
vial to promote mixing and grinding.
Electrode loading requires practice to obtain a uniform packing of the
powder. This can be achieved by carefully pouring the sample in small incre-
ments into the electrode cavity and either tamping the powder or leaving it
loosely packed between increments. Alternatively, the electrode can be
inverted and pressed into a small heap of the powder. Any tendency for sam-
ples to be ejected from the arc during excitation can be reduced by adding a
few drops of sucrose solution to the loaded electrodes and then drying them
on a hot plate.
Having excited the samples and recorded their spectral intensities these are
then compared with standards on a comparator. The operator can either
extrapolate concentrations between standards or simply report results at the
logarithmic midpoints between standards, so that a line with an intensity
between the 20- and 50-ppm standards would be reported as 30 ppm (Myers
et al., 1961). Inevitably some concentrations will exceed the intensity of the
highest standard, whereas for others the analytical line is absent or too weak
to evaluate. For these situations Berenice et al. (1975) suggests the following
alphabetic qualifiers to the reporting limits: N — line not detectable; L — a
weak line, concentration too low to evaluate; and G — line too intense to
evaluate. H — preceding a numeric value, indicates a possible interference.
Finally in organizing a semi-quantitative method it must be emphasized
that, for comparison with standard films to be valid, consistent attention
must be given to ensuring that excitation and operating conditions remain
constant over long periods of time. Hence, in order to monitor systematic
errors resulting from equipment drift, changes in film batches or deterioration
of old stock, operator proficiency etcetera, it is important to include at least
one of a series of standard control samples on each film. Because of the small
quantity of material analyzed it is particularly important to achieve a uni-
form distribution of the trace elements throughout the matrix of the control
standards.

DIRECT-READING DC-ARC SPECTROMETRY

In the direct-reading spectrometer the photographic plate is replaced by


photomultipliers mounted around the focal curve of the instrument so that,
after dispersion, each of the chosen analytical lines falls on a photomultiplier
153

(Fig. 7-1). Output from the photo multiplier, which is dependent on incident
light intensity and hence on element concentration, is fed to a voltmeter or
other electronic readout system.
With only a few photomultipliers it would be feasible to convert volts to
concentration units manually. However, for efficient utilization of a direct
reader, with perhaps as many as 60 analytical channels, concentrations must
be calculated automatically and most direct readers are therefore interfaced
to small dedicated computers. Ideally the computer should also be pro-
grammed to correct for background, matrix effects and spectral interferences
thus improving precision and accuracy of the data. Various aspects of the
application of direct reading spectrometers to determination of trace ele-
ments in geochemical matrices have been discussed by Cruft and Giles
(1967), Tennant and Sewell (1969), Scott et al. (1969), Foster (1970),
Timperley (1974) and De Pablo Galan (1975).
Initial standardization of the direct-reading spectrometer involves firing of
synthetic or natural standards of known major and trace element composi-
tion. Calibration curves, either of line intensity versus concentration or
intensity ratio to an internal standard versus concentration, are then stored
in the computer. Depending on the concentration range of interest and the
sensitivity of the analytical lines, the calibration curves can be either
approximately linear, or exhibit a pronounced flattening at low concentra-
tions because of background or the presence of an interfering line (Fig. 7-4).

100

without correction
corrected

1 10 100 1000
CONCENTRATION, ppm

Fig. 7-4. On a log-log plot the presence of background or an interfering line can cause the
slope of the calibration line to flatten at low concentrations (dashed line). Subtraction of
the intensity obtained with the same matrix, in the absence of the analyte, extends the
lower limit of the linear range (solid line).
154

Non-linear calibrations can either be broken into straight line segments or


represented by low-order polynomials.
Once calibration curves have been established daily checks involve firing
of high and low standards. Because of the inherent instability of the DC-arc
replicate firings are necessary and, even with replication, mean values may lie
some distance from accepted values without the instrument being out of
calibration. Foster (1970) has suggested that with nine replicates, calibration
is acceptable if their mean lies within half a standard deviation of the
accepted value.
The great rapidity of the direct reader and its use in conjunction with a
computer justifies far more attention to methods of improving the quality of
data than is possible with visual estimates of spectral densities on film. Such
improvements largely relate to correction of background, matrix effects and
spectral interferences.

Matrix effects

Matrix effects can be reduced in several ways most of which are normally
used, either singly or in combination, in quantitative spectroscopy and are
fully described by Ahrens and Taylor (1961). Use of spectroscopic buffers
has already been mentioned on p. 145. Other important methods of
reducing matrix effects include the use of internal standards and calculation
of matrix correction factors from the bulk chemical composition of the
sample.
Introduction of an internal standard into the sample allows variations in
line intensities, due to matrix effects or arc instability, to be monitored and
corrected providing behaviour of the internal standard is similar to that of
the analyte (Fig. 2-6). Their respective volatilization rates are particularly
important in this respect. Indium is generally suitable as an internal standard
for volatile elements (Ga, Pb, Ag, Cu, Tl, Ge, Sn, Zn) and Pd for involatiles
(Zr, Sc, Y, Nd, La, Sr, Ba, V, Ni, B, Co, Cr, Mo, Mn, Ti). Suitability of an
internal standard should, however, be checked by comparing volatilization
curves under actual operating conditions. This can be done by programming
the direct reader to present results for each channel At short intervals of time
throughout the duration of the burn. Timperley (1974) has presented a
series of curves prepared in this way (Fig. 7-3).
By analyzing natural and synthetic standards of known major and trace
element composition it is possible to calculate correction factors for matrix
effects of major elements on each other and on trace elements. For example,
De Pablo Galan (1975) prepared 13 synthetic matrices, with variable Si0 2 ,
A1203, Fe 2 0 3 , CaO, MgO, Na 2 0 and K 2 0 contents representing a wide range
of geological materials, containing trace element concentrations from 1 to
4600 ppm. By analyzing these matrices and solving a series of simultaneous
equations, correction factors (x) were obtained for the effect of the con-
155

centrations of major elements on apparent concentrations of trace elements.


For unknowns preliminary measurements of intensity ratios, with 500 ppm
Pd in a graphite-lithium borate flux as internal standard, could then be
refined by equations of the type:
=
(Je/'pd)c (Vpd)M +*sio 2 (Si0 2 ) + χ Αΐ2θ3 (Α1 2 0 3 ) ...χΚ2θ(Κ20)

where (7 e // P d ) c = corrected intensity ratio, (/ e /J Pd ) M = measured intensity


ratio, and x S io 2 = correction factor for the concentration of silica (Si0 2 ).
Corrections on a small computer enabled analyses to be completed in less
than 1 minute. If the concentration of a single major element accounts for
most of the error in the estimation of the concentration of a trace element,
a simple linear correction can be calculated as described for spectral inter-
ferences (Fig. 7-4).
A very different approach to avoiding matrix effects has been taken by
Danielsson et al. (1959), Danielsson and Sundkvist (1959) and Danielsson
(1967), for which Danielsson introduced the term isoformation. Soils and
sediments are sieved or ground to minus 100 mesh (isoformation). The fine
material is then automatically fed onto an adhesive tape which passes
through a spark discharge in an argon atmosphere (Fig. 7-5). Each discharge
strikes a new portion of the sample thereby eliminating effects of variations
in volatilization rates resulting from compositional differences between
matrices. The method is popular in Scandinavia — see, for example, Brundin
and Nairis (1972) and Kauranne (1976).

Fig. 7-5. The tape machine. A = stock roll; B = folding device; C = safety device; D = per-
forating device; E = sample hopper; F = pressure-suction device for removal of excess
sample powder; G = spark gap; H = air blast; K = pulling wheels; and L = waste bin. Sam-
ple powder is continually added to the adhesive tape via the hopper (E) and then passes
through the spark gap where a new portion of the sample is continually excited.
156

Background

Spectral background causes asymptotic flattening of the calibration curve


and at concentrations close to the detection limit line intensities as little as
one tenth the background intensity must be measured (Fig. 7-4). Under
these circumstances an error of only 5% in estimating background would
cause a 50% error in estimation of line intensity. Reduction and reliable esti-
mation of background is therefore an important consideration in direct
reading spectrometry.
For elements that are fully volatilized early in the burn the most favourable
line to background ratios are obtained by recording only the early part of
the burn, optimum integration times being estimated from volatilization
curves such as Fig. 7-3. When this is not feasible background can be reduced
by using alkali buffers to lower arc temperature or by sheathing the arc with
an inert gas to reduce CN band emission as described on p. 147.
Background intensity, which varies with wavelength (Fig. 7-6), can be
measured by locating photo-multipliers at one or more line free portions of
the spectrum and assuming relative intensities of background at different
points in the spectrum remain constant despite matrix variations. The
intensity ratio for analytical to background line positions must initially be
estimated using a trace element-free base. Alternatively, as described by
Timperley (1974), by shifting the spectrograph slightly off the line a profile
of the intensities on either side can be obtained. If similar background inten-
sities are found on both sides of the line the average value is taken as an esti-
mate of background. Large discrepancies between the readings indicates a

l.b - .
1 4- //
z
LU
1.2 -

1.0-
1 //
/ ^ / /
(0 0.8-
>^^~\ Diabase W-1 ^ /y/Synthetic //
3 < ^ — \ ^>^ * ^ / ^ - b a s i c rock //
0.6-
o f V ^ ^ - - ^ " G i r a n i t e G-1^/^-->^ /
cc
O
(J
0.4 -<
\/y\^<zzzri^_^Z^
^ Synthetic acidic rock
^^
0.2-
<
0 1 1 1 1 1 1 1 1 1—i—
250 260 270 280 290 300 320 340
WAVELENGH, nm
Fig. 7-6. Relationship between background intensity and wavelength in a DC-arc for
several different rock types. (From Tennant and Sewell, 1969.)
157

line (spectral) interference and only the background estimate from the inter-
ference-free side of the line is used. With some direct-reading spectrometers
the entrance slit of the spectrograph can be automatically oscillated to pro-
vide a display of the line profile and adjacent spectrum.

Spectral interferences

Spectral interferences, caused by coincident and overlapping lines or


molecular band emissions, can be reduced by careful line selection (Table 7-
III) or, if sufficient analytical channels are available, by use of alternative
lines in the presence of significant levels of an interferent. If this is not
feasible correction factors can be calculated from synthetic standards con-
taining known concentrations of the interferent. For example, Scott et al.
(1969) found that, measuring the Sn 284.0 nm line, 1% Cr gave a reading
equivalent to 100 ppm Sn. A similar interference of Fe on Mo 317.035 nm
was described by Tait and Coats (1976). Simple linear corrections are
usually adequate to remove this type of error (Fig. 7-4).
Matrix effects, causing changes in arc conditions, background and spectral
interferences have been described separately. In practice they usually exhibit
complex interrelationships and cannot always be readily distinguished. For
example, an abundance of alkalies lowers arc temperatures, increases selec-
tive volatilization, reduces background and might either increase or decrease
the relative intensity of an interfering line or band. Both background and
spectral interferences cause flattening of the calibration curve at low analyte
concentrations (Fig. 7-4). Careful choice of analytical lines, and use of spec-
troscopic buffers together with calculation of corrections for matrix and
spectral interferences, can overcome these problems sufficiently to provide
more reliable results, with a faster analytical throughput than with semi-
quantitative photographic methods. Precision and accuracy is not, however,
comparable to that obtained with AAS.

PLASMA SOURCES

Greenfield et al. (1975a) list the desirable properties for spectroscopic


sources as: (1) the ability to excite the lines of a large number of elements;
(2) high sensitivity; (3) good stability; (4) freedom from interferences; (5)
reproducibility in the introduction of samples; and (6) convenience of opera-
tion. Particularly with regard to (3), (4), and (5) the DC-arc is not an ideal
source: the inductively coupled plasma (ICP) is now attracting considerable
attention as an alternative source.
Plasmas are masses of gas sufficiently ionized for this to have a significant
effect on their properties. Most important for their use as emission sources,
large quantities of electrical energy can be pumped into them giving tempera-
158

tures much higher than those attainable in flames or arcs. Microwave plasmas
have been used to analyze the volatiles released by inductive heating of soils
and sediments (Meyer and Lam Shang Leen, 1973). However, the greatest
interest in plasmas has focussed on the general analytical capabilities of the
inductively coupled (radio-frequency) plasma in conjunction with the direct-
reading spectrometer. In contrast to DC-arc spectroscopy the sample is
usually introduced into the plasma in solution: the technique therefore has
all the advantages and disadvantages associated with dissolution of geochemi-
cal samples.

The inductively coupled plasma

The first reports on the use of the ICP as an excitation source for trace ele-
ment analysis were published by Greenfield et al. (1964) and Wendt and
Fassel (1965). Theory and analytical applications have been reviewed by
Greenfield et al. (1975b, c), Fassel and Kniseley (1974a, b), Fassel (1978)
and Boumans (1979). State of the art information regularly appears in the
ICP Information Newsletter edited by R.M. Barnes (Department of Chem-
istry, University of Massachusetts, Amherst, MA 01002, U.S.A.).
Various designs of plasma cells have been used but modifications of that
first described by Greenfield et al. (1964) are generally regarded as most
satisfactory. These consist of two concentric quartz tubes surrounding a glass
sample injector (Fig. 7-7). Argon to form the plasma passes through the con-
centric quartz tubes, flow in the outer tube being tangential to produce a
vortex which both stabilizes the plasma and keeps it away from the tube
walls. (Consumption of argon, at 10—20 1/minute, is the major operating
cost of the ICP.) The upper part of the plasma cell is surrounded by a few
turns of an induction coil powered by a high-frequency generator: crystal-
controlled generators operating at 27 MHz and providing up to 2 kW of
power appear to be becoming standard. Greenfield et al. (1975b, c), how-
ever, prefer a more powerful (up to 15 kW) free-running generator operating
at lower frequencies (7 MHz).
High-frequency currents in the induction coil generate a magnetic field
around the plasma cell (Fig. 7-7). If the plasma gas is then seeded with elec-
trons using a Telsa coil, the electrons and ions move in closed annular paths
within the cell. Heating occurs from collisions between charged particles to
give temperatures of the order of 7000—10,000°K. After the plasma has
formed the sample aerosol is injected into its centre, creating a doughnut-
shaped plasma in which the sample passes through the hole without pene-
trating the skin of the plasma. This impenetrability of the plasma results in it
being relatively insensitive to injection of aerosols and hence able to tolerate
high sample loadings without loss of stability or quenching.
Although temperatures in the axial channel of the plasma fireball are
lower than those within the plasma, they are still much higher than tempera-
159

60 -
A. > - 3 5 7 . 8 6 9 nm
Excitation energy =3.46 eV /^^Nv
Q 20 mm / ^^
Z
3
O
GC
O
40 H
A-*"' >
/ y 15mm \

*
o
/ ' ■

<
CO
\UJ 2 0 - /
z
/^ #' -"♦~~ ^N. 10 mm

" V
mr—^^^f''
ψ:°—-~
^* >-♦
n

AEROSOL Ar FLOW RATE, I min"1

Hottest zone 25-

v'-'-Magnetic
field
/ ^ 15 mm
20-
'/ \ \
Q 'i
{
N
\l0mm
\
Z
•D
i* / \♦ \
o J" \
QC
O
15-
\ \
o* \\ \•
<
oo \\
X
UJ
10- 20^Tun ^
\
z
-1
s' \\ \

5- B. /> =267.716 nm \\ \ \\ \
Argon
cooling
Excitation energy \ v
12.6 eV V >
gas \ Λ
Sample 0
Plasma aerosol
Argon AEROSOL Ar FLOW RATE, I min -

Fig. 7-7. The inductively coupled plasma.

Fig. 7-8. Influence of the aerosol argon flow rate on emission of (A) the atomic Cr line at
357.869 nm; and (B) the ionic Cr line at 267.716 nm, at three observation heights in an
inductively coupled plasma. (From Berman and McLaren, 1978a.)

tures attainable in flames or arcs and are more than adequate to vaporize the
solute and dissociate molecular species. However, the background continuum
in the plasma fireball is intense and for analytical purposes it is necessary to
view the spectrum at some point in the tail flame where background is low
but temperatures are still sufficiently high to prevent recombination of
atoms. Because the axial channel is cooler than the surrounding plasma the
problems of self-absorption and self-reversal — encountered with flames and
arcs in which temperatures decrease towards the outer fringes — are absent
and the plasma behaves as an optically thin source.
160

Temperature variations along the tail flame of the plasma can be dramati-
cally illustrated by aspirating a solution containing Ba (see Fig. A-2, Appen-
dix 1). Emission from the yellow (Ba 585.37) and blue (Ba 493.4) ion lines
is well separated from the continuum whereas the most intense emission of
the green (Ba 553.5) atom line occurs immediately above the continuum and
again, though somewhat weaker, higher in the plasma. As discussed below,
the response of ion and atoms lines to operating conditions in the plasma
can differ considerably. Boumans (1979) has listed the most sensitive lines
for seventy elements; approximately half of them are ion lines.
Principal variables controlled by the analyst are: (1) the argon and aerosol
flow rate, (2) the generator power fed to the plasma, and (3) the height in
the tail flame at which the emission is viewed. Generally the objective will be
to establish optimum compromise conditions for determination of as many
elements as possible, rather than to optimize for a single element. The prob-
lem of establishing compromise conditions has been discussed by Berman
and McLaren (1978a) who present a series of line to background ratios under
varying operating conditions for atom and ion lines. For atomic lines of Pb,
Mn and Cr the line to background ratio increased with viewing height and
aerosol flow rate up to about 1.0 1/minute (Fig.7-8). In contrast, with ion
lines the optimum line to background ratio occurs lower in the plasma and
with lower aerosol flow rates. Increasing power input to the plasma increases
line intensity but also causes a rapid rise in background so that line to back-
ground ratios are less favourable at maximum power (Fig.7-9). On the basis

-- A - 357.869 nm
Aerosol Ar
0.95 I min"1
— A = 267.716 nm fi
Aerosol Ar /
>* 150 0.80 I min"1 '

n
<

z
UJ
I-
Z 50

1.0 1.2 1.4 1.6


INCIDENT POWER, kW
Fig. 7-9. Effect of power on background intensity at two Cr lines in an inductively
coupled plasma. (From Berman and McLaren, 1978a.)
161

of these results, it was concluded that the best compromise conditions would
be obtained with relatively low argon flow rates and observation of ion lines
at a height of about 15 mm. These conditions should also minimize non-
spectroscöpic interferences.
Kirkbright and Ward (1974), comparing ICPs and flames as emission
sources, have noted that in addition to its higher temperature and lower self
absorption, the plasma is also characterized by both longer particle residence
times and freedom from the intense band spectra of CH, OH, C and CO
found in flames (or CN found in DC-arcs). Also, because the free electron
density is much greater than sample particle density, the degree of ionization
is essentially unaffected by sample introduction.
In practical terms the longer residence times and higher temperatures
should ensure complete volatilization of the solute and dissociation of
molecular species, thereby increasing sensitivity and eliminating band spectra
and interferences caused by formation of refractory compounds. This is
borne out by the virtual absence of the Ca-Al and Ca-P interferences which
are so troublesome in flames (Fig. 7-10; see also p. 127). Ionization inter-
ferences also appear to be less than with other sources presumably due to the
high electron densities (Larson et al., 1975). As a further advantage, the
combination of increased sensitivity and, at high analyte concentrations,
lower self-absorption gives the plasma a linear working range extending over
five orders of magnitude compared to three for flames. It is therefore possi-
ble to determine both trace and major elements on the same solution
without dilution.
In the absence of significant chemical and ionization interferences the
most serious analytical problems, particularly near an elements detection
limit, have been associated with spectral interferences from overlapping and
coincident emission lines or from background. Coincident or overlapping
spectral lines are not, of course, restricted to plasma sources: assuming
spectrometer resolution is adequate they are best avoided by careful selec-
tion of lines. If this is not possible, appropriate correction factors can be
computed by measuring the signal produced in the analyte channel by a
known concentration of interferent (Marciello and Ward, 1978). Tables of
line interferences in the ICP have been published by Parsons et al. (1980).
Correction of background is particularly important if low signal to back-
ground intensities are to be measured. Under these circumstances small
changes in background emission can severely degrade analytical performance.
Variations in intensity of background emission between samples and stan-
dards can result either from changes in the proportions of molecular species
emitting band spectra or from light scattering, causing stray light, in the
spectrometer. Stray light is most troublesome when strong emitters, such as
Ca, are abundant in the sample. Apart from minimizing stray light with
improved spectrographs and gratings, the most promising approach to this
problem appears to be measurement of background intensity adjacent to the
162

1.0 -e
£--*■-—r- ICP
» -*
=> (Ca 393.4 nm)
0.8 -
s.
<
0.6 -
AAS
or Air Acetyl ene <§L
h-
0.4 -
Z (Ca 422.7 nm) ^s.
LU
CC
<
a. 0.2 - v§)
o. A,
< —,
0 L-// .— 1
0 100 500 1000 □ 3.5 7o HCI0 4
PHOSPHORUS/CALCIUM RATIO
• 500 ppm Ca
CO
2.0 n z A 100 ppm Sr
AAS LU
♦ 100 ppm Ti
N 2 0-Acetylene I-
z
(Ca 422.7 nm) ($-
< LU

υ >
LAJ
ICP
z (Ca 393.4 nm)
LU 0.5 H
cc
<
Q.
Q.
< -//- —\ 1 —r~ ~v
0 10 30 100 300 1000 0.91 0.90 0.89 0.88 | 0.87 0.86
POTASSIUM/CALCIUM RATIO 400.875
WAVELENGTH, nm

Fig. 7-10. Comparison of chemical and ionization interferences in flames and the induc-
tively coupled plasma. A. Chemical interference of P on Ca. B. Ionization interference of
K on Ca. (From Ediger and Wilson, 1979.)

Fig. 7-11. Some wavelength profiles showing interelement effects on emission at the W
400.875 nm line in an inductively coupled plasma, a = 3.5% HCIO4 background; plus, · =
500 ppm Ca — broadband scatter; * - 100 ppm Sr — line reflection or scatter; and ♦ =
100 ppm Ti — spectral overlap from 400.893 nm. (Reprinted with permission from
McQuaker et al., 1979, Calibration of an inductively coupled plasma-atomic emission
spectrometer for analysis of environmental matrices, Anal. Chem.y 5 1 : 888—895. © 1979
American Chemical Society.)

analyte line (Fassel, 1978). This can be done by oscillating the spectrometer's
entrance slit to display a scan of the line profile (Fig. 7-11).
Detection limits reported by various workers using compromise conditions
and pneumatic nebulizers are summarized in Table 7-VL Appreciably lower
limits are obtained with ultrasonic nebulizers; these, however, are unsuited
to rapid sample throughput. The stability of the plasma enables relative stan-
dard deviations of about 5% to be achieved.
163

TABLE 7-VI
Detection limits in solution ^ g / m l ) with the ICP *

Element 1 2 3
As 0.039
B 0.1 0.001
Ba 0.0001 0.001 0.0003
Be 0.0002
Cd 0.1 0.002
Co 0.004 0.02 0.029
Cr 0.001 0.02 0.004
Cu 0.01 0.003
Fe 0.04 0.004
Ga 0.022
Mn 0.003 0.0006
Mo 0.02 0.004
Nb 0.001
Ni 0.003 0.01 0.017
P 0.17
Pb 0.008 0.07 0.037
Sb 0.050
Sn 0.1 0.029
Sr 0.02 0.0001
Te 0.19
Ti 0.005 0.03 0.002
V 0.002 0.009 0.002
w
Zn 0.01 0.1
0.15
0.003
1
Reference: 1 = Scott et al. (1974): detection limit twice standard deviation of back-
ground signal; 2 = Greenfield et al. (1975c): compromise conditions; 3 = McQuaker et
al. (1979): means over a one-year period.

Analysis of geochemical samples

Analysis of fresh waters with the ICP appear to present few problems and
has rapidly gained acceptance. Considerably less information is available on
application of the ICP to trace element determinations in soils, sediments
and geological materials, although the topic has been reviewed by Berman
and McLaren (1978b).
In one of the few published studies Scott and Kokot (1975) compared
determination of Cu, Pb, Zn, Co and Ni with the ICP and AAS. One-gram
soil samples were digested with nitric-perchloric acids and diluted to 50 ml
with water for aspiration into a 27-MHz, 1-kW argon plasma at 1 ml/minute.
When viewed 20 cm above the load coil no problems were encountered in
determination of Cu (324.7 nm) and Zn (213.8 nm) which had linear calibra-
tions over four orders of magnitude. Linear calibrations were also obtained
164

for Ni (351.5 nm) and Co (345.3 nm) but results were lower than those by
AAS: this appeared to be due to interferences in the flame rather than the
plasma. In the case of Pb, high Ca content of the soils caused spectral back-
ground in the vicinity of the most sensitive line (405.8 nm) and it was neces-
sary to apply a correction. Scott et al. (1976) obtained a detection limit of
0.1 ppm U in solution with the line at 378.28 nm. Motooka et al. (1979)
eliminated major element interferences and enhanced sensitivity for trace
elements (Ag, Au, Bi, Cd, Cu, Pb and Zn), after their extraction from geologi-
cal materials using potassium chlorate/hydrochloric acid (see p. 72), by sol-
vent extraction into MIBK with Aliquat 336. Although the plasma is some-
what less stable with MIBK than with aqueous solutions, results on standard
samples were well within the limits necessary for exploration purposes.
Dahlquist and Knoll (1978), McQuaker et al. (1979) and Floyd et al.
(1980) have discussed problems encountered in analysis of soils and similar
materials. Spectral wavelength scans indicated that broad band scatter and
spectral overlaps, caused by varying amounts of major concomitants, give
rise to significant interferences for most elements at low concentrations
(Fig. 7-11). Contribution of these interferences to the analyte signal can be
corrected using linear or second-order polynomial regressions. As reported
by other workers, the effects of chemical and ionization interferences in the

105-
ί
95
-
°/0 U p t a k e R a t e Γ

<
g
υ
100 -
<KJ ►
L
^90 UJ
UJ
CC
i H
► '
Q.
3
95- i

- |
% Recovery

90-

1 1 1 80
7.0 14.0

% HCI0 4

Fig. 7-12. Effect of perchloric acid content on uptake rate and recovery (when calibration
is carried out in 3.5% perchloric acid) in the inductively coupled plasma. Recoveries are
mean effects found for Cd, Cr, Cu, Fe, Mg, Mn, Mo, Pb and Zn. Bars indicate spread.
(Reprinted with permission from McQuaker et al., 1 9 7 9 , Calibration of an inductively
coupled plasma-atomic emission spectrometer for analysis of environmental matrices,
Anal. Chem., 5 1 : 888—895. © 1 9 7 9 American Chemical Society.)
165

plasma were generally found to be insignificant. Acid concentration in sample


solutions and standards should, however, be reasonably matched to avoid
apparent interferences from the effects of changing viscosity on nebulization
rates (Fig. 7-12).
Rather than aspirating solutions into the ICP, Thompson et al. (1978a, b)
determined As, Sb, Bi, Se and Te by generation of their gaseous hydrides,
using 5 M hydrochloric acid — 1% sodium borohydride as the reductant
(p. 132), and direct introduction of the hydride into the ICP. Detection
limits of 1 ppb or better in solution were obtained but it was necessary to
control the interference of Cu on Bi, Se and Te by their coprecipitation and
separation on lanthanum hydroxide. No other serious interferences were
found with typical geochemical matrices. Using the same ICP system,
Pahlavanpour et al. (1979) separated Sn from interfering elements by sample
decomposition with ammonium iodide fusion (p. 66, Fig. 4-4). The
sublimate was leached with a 1% solution of tartaric acid; sodium boro-
hydride was then added to generate stannane (SnH4). Direct introduction of
stannane into the ICP gave a working range of 0.02—50 ppm compared to
10 ppm to 10% for aspiration of the solution into the plasma.
As yet it is too early to assess the impact of the ICP on analysis in the
exploration laboratory. It seems probable, however, that whenever several
hundred samples per day have to be analyzed for more than about eight ele-
ments the ICP will become competitive with AAS. Lower detection limits
and the need for fewer dilutions, combined with comparable precision, are
additional benefits. However, the potential advantages of the ICP to analysis
of geochemical samples will not be fully realized without parallel develop-
ments in rapid decomposition procedures suitable for a wide range of ele-
ments and matrices.
Chapter 8

X-RAY FLUORESCENCE

INTRODUCTION

X-rays, which occupy the electromagnetic spectrum between γ-rays and


the vacuum ultra-violet (0.01—10 nm), are the most energetic form of non-
nuclear radiation. When a sample is bombarded by electrons, X-rays or 7-rays
of sufficient energy, secondary X-rays are emitted at wavelengths (or
energies) and intensities that are determined by the elemental composition
of the sample. Measurement of the intensity of this secondary, characteristic
radiation is the basis of X-ray emission spectroscopy or X-ray fluorescence.
In conventional X-ray spectrometers the sample is the target for primary
X-rays generated in an X-ray tube requiring a stable, high voltage power
supply. Where this is impractical, as in portable field instruments, radio-
isotopes can be used as a primary radiation source. The characteristic radia-
tion of the element to be determined is isolated from other X-rays on the
basis of its wavelength or energy, and its intensity measured. For elements
lighter than Fe (atomic number Z = 26) the characteristic radiation is
strongly absorbed by air and it becomes necessary to mount the sample and
detector in a vacuum chamber.
XRF, which has found wide acceptance for determination of the major
constituents of rocks, is also capable of providing excellent data for many
trace elements in geochemical matrices. Moreover, sample preparation can
be extremely simple. However, compared to ES, colorimetry and AAS,
application of XRF in exploration laboratories has been relatively limited.
This perhaps reflects the higher capital cost of the conventional wavelength-
dispersive X-ray spectrometer, its somewhat poorer detection limits for many
of the elements of interest, and the need to correct data for matrix effects.
Recent advances in energy-dispersive (or non-dispersive) X-ray spectrometry
have reduced capital costs for simultaneous multi-element determinations.
This should promote a re-evaluation of the potential applications of X-rays
in geochemical exploration.
For analysis outside the laboratory there has been considerable interest
in development of small portable XRF units using radio-isotope sources, to
excite sample fluorescence, and filters to screen out unwanted radiation. At
168

present these instruments are only able to measure trace element concentra-
tions in excess of several hundred parts per million. They are therefore most
useful for determination of elements with background values in this range,
for example Ba and Sr (Ball et al., 1979; Grout and Gallagher, 1980) or for
on-site inspection of drill-core, pit faces or outcrops where the presence of
"invisible" mineralization or anomalous values is suspected. Friedrich et al.
(1974) have described a sled mounted system for remote analysis of manga-
nese nodules on the ocean floor. Development of portable instruments, their
use and associated problems have been described by Bowie et al. (1965),
Bowie (1968), Gallagher (1967, 1969), Garson and Bateson (1967); and Von
Alfthanetal. (1980).
Other specialized geochemical applications of XRF, beyond the scope of
this general text, include its use to determine trace elements after their pre-
concentration from solution by ion exchange. With natural waters this
involves collection of particulate and dissolved metals on ion-exchange
loaded filters (Blasius et al., 1972). Alternatively, trace elements can be
co-precipitated from natural waters or analytical digests, the precipitate
collected as a thin film on a filter paper and analyzed by XRF (Campbell et
al., 1966; Green et al., 1971; Watanabe et al., 1972). Stanton (1976), fol-
lowing procedures described by Luke (1968), gives rapid methods for deter-
mination of Bi; As, Se and Te; and Au, Pa and Pt by co-precipitation and
XRF.

THEORY

Numerous texts have described the properties of X-rays and the nature of
their interactions with matter — the reader is especially referred to Jenkins
and De Vries (1970). In this section the fundamental principles underlying
X-ray spectrometry are only briefly described.

Excitation of X-rays

When an element is bombarded by sufficiently energetic X-rays, electrons


are ejected from their orbits leaving holes which are filled by transfer of elec-
trons from the outer electron shells (Fig. 8-1). The difference between the
energy of the original and final state of the electron is released as an X-ray
photon of the same energy. Consequently, the energy of the photon is
characteristic of the particular transition involved and the target element.
The relationship between energy (V) and wavelength (λ) of the characteristic
radiation is given by:

\ = hc/V=1.24/V
169

Fig. 8-1. Idealized representation of the production and nomenclature of characteristic


X-rays. Only a few of the many possible transitions for the L- and M-shells are shown.

where V is in kilovolts, h is Planck's constant, c the velocity of light, and λ is


in nanometres.
Nomenclature of X-ray spectra is determined by the site of origin of the
ejected electron: thus, transitions terminating in holes resulting from ejec-
tion of electrons from the K-shell give rise to the K spectrum, holes in the
L-shell the L spectrum and so on. Each of these spectra comprises several
characteristic lines, with the K spectrum being the simplest. Because the
energy of the emitted X-ray photon is equal to the difference in energy
arising from the electron transition, K^ radiation, resulting from M -► K shell
transitions, is more energetic and at a slightly shorter wavelength than the
corresponding Ka radiation from the L -> K transition. Similarly, K spectra
are more energetic than L spectra (Fig. 8-1).
In 1913 Moseley established the relationship, now known as Moseley's
Law, between the wavelength of the characteristic radiation associated with
a particular transition and the atomic number (Z) of the target:

l/X = k(Z-o)2
where k and σ are constants. A corollary to Moseley's Law is that more
energetic primary radiation is necessary to produce X-rays from the heavier
elements. For example, the minimum potentials needed to excite Ka spectra
for Cu (Z = 29), Sn (Z = 50) and U (Z = 92) are 8.98, 29.19, and 115.9 kV,
170

respectively (Table 8-1). The L spectra of all elements up to and including U


are excited if the energy of the primary radiation exceeds 22 kV.
Relative intensities of X-ray lines depend on the probabilities associated
with the corresponding electron transitions. Ka lines are most intense and

TABLE 8-1
K spectrum excitation potentials and wavelengths for selected elements (based on
Jenkins and De Vries, 1970)

Element Atomic Excitation Wavelength (iim)


number, potential
Z (kV) *ß Ka

Na 11 1.1 1.1617 1.1909


Mg 12 1.3 0.9889 0.9889
Al 13 1.6 0.7981 0.8338
Si 14 1.8 0.6769 0.7125
P 15 2.1 0.5804 0.6155
S 16 2.5 0.5032 0.5372
Cl 17 2.8 0.4403 0.4728
K 19 3.6 0.3454 0.3742
Ca 20 4.0 0.3089 0.3359
Sc 21 4.5 0.2780 0.3031
Ti 22 5.0 0.2514 0.2749
V 23 5.5 0.2285 0.2503
Cr 24 6.0 0.2085 0.2290
Mn 25 6.5 0.1910 0.2102
Fe 26 7.1 0.1757 0.1936
Co 27 7.7 0.1621 0.1789
Ni 28 8.3 0.1500 0.1658
Cu 29 9.0 0.1392 0.1540
Zn 30 9.7 0.1296 0.1435
Ga 31 10.4 0.1207 0.1340
Ge 32 11.1 0.1129 0.1255
As 33 11.9 0.1057 0.1175
Br 35 13.5 0.0933 0.1040
Rb 37 15.2 0.0829 0.0926
Sr 38 16.1 0.0783 0.0875
Y 39 17.1 0.0740 0.0825
Zr 40 18.0 0.0701 0.0786
Nb 41 19.0 0.0665 0.0747
Sn 50 29.2 0.0435 0.0491
Ba 56 37.4 0.0341 0.0385
La 57 38.9 0.0328 0.0371
Ce 58 40.4 0.0316 0.0357
W 74 69.5 0.0184 0.0209
Pb 82 88.0 0.0146 0.0165
Th 90 109.8 0.0117 0.0132
U 92 115.0 0.0111 0.0126

Based on the wavelength of the absorption edge.


171

are therefore generally favoured for analysis. However, for heavy elements
the excitation potential of the K spectra may exceed the instrument power
and safety rating. One of the more intense lines in the L spectrum is then
chosen. M spectrum lines are very weak.
The actual measured intensity or fluorescent yield for a particular transi-
tion is always less than predicted. This results from the Auger effect, which
involves the internal re-absorption of the secondary X-ray photon and causes
serious reduction of X-ray intensity for light elements.

Interaction of X-rays with matter

Emission of secondary fluorescence involves photo-electric absorption of


primary X-rays striking the target. A proportion of the primary beam is also
lost by scatter or is transmitted through the target. If a monochromatic
beam of X-rays (λ), with intensity J0, is incident on a homogeneous target
the beam will be attenuated by (I0 — I) where I is the intensity of. the trans-
mitted beam. The extent of attenuation by absorption and scatter is related
to the mass (M) of the absorber by:

dl0 = —Jo · μλάΜ

where μ λ , the mass absorption coefficient, depends only on the wavelength


(λ) of the X-rays and the atomic number (Z) of the target. Tables of μλ for
different elements and wavelength are available. For composite materials the
average mass absorption coefficient (μ λ ) is given by:

μλ = Σ (μΖλ · Wz)
where μΖχ and Wz are the mass absorption coefficient and weight fraction
for each element (Z) in the material.
The mass absorption coefficient includes both photo-electric absorption
(r) and scatter (σ) so that μ = r + o. However, photo-electric absorption is
quantitatively far greater than scatter, so μ ^ r. A plot of μ for an element as
a function of wavelength shows several major breaks, or absorption edges,
each one corresponding to the minimum potential required to excite the
spectrum associated with a particular shell. For example, in Fig. 8-2 the K
absorption edge for Ni at 0.149 nm corresponds to the energy (8.33 kV) just
necessary to eject an electron from the K-shell of a Ni atom. Wavelengths
slightly longer than the K absorption edge have lower energies, cannot eject a
K-shell electron, and are, therefore, only subject to the photo-electric
absorption associated with emission of the L and M spectra. There is there-
fore an abrupt decrease of μ at the absorption edge. Maximum absorption of
incident X-rays, and therefore most efficient excitation of secondary X-rays,
occurs with wavelengths just shorter than the targets absorption edge: thus,
172

Fig. 8-2. Relationship between wavelength and the mass absorption coefficients of Ni and
Co in the vicinity of their K-shell absorption edges. Mass absorption coefficients of Ni and
Co for CU-KQ, are approximately 50 and 350, respectively.

in Fig. 8-2, Cu-Ka is strongly absorbed by Co (but not Ni) and would be an
ideal primary radiation for excitation of Co-Ka. Between absorption edges μ
is proportional to λ 3 and Z 4 .
Scattering of X-rays by the target is usually much less important than
photo-electric absorption. Nevertheless, scattering cannot be ignored as it
contributes to background radiation reaching the detector and also has some
very useful analytical applications. Total scatter is a combination of coherent
(Rayleigh) and incoherent (Compton) scatter. Rayleigh scatter, which is the
phenomenon involved in X-ray diffraction, results from elastic collisions
between the incident X-rays and target electrons. The primary radiation is
scattered without a change in wavelength. In contrast, Compton scatter
involves inelastic collisions in which the primary X-ray photons loose some
of their energy to the electrons. As a result the scattered X-rays are at a
slightly longer wavelengths than the primary beam. The change in wave-
length (Xd) is given by:

Xd = (1 — cos φ)
mc
where h = Planck's constant, m = rest mass of an electron, c = velocity of
light, and φ = angle between the primary beam and the scattered beam. For
most spectrometers φ is approximately 90° so that the wavelength shift
between coherently and incoherently scattered lines is about 0.0024 nm
(Fig. 8-3). The intensity of the scattered radiation is proportional to Z" 3 t o " 2 .
Consequently, as shown in Fig. 8-3 for Mo-Ka radiation scattered from
teflon, silica and lead discs, scattering of radiation is most intense from
light matrices and can be used as an estimate of μ (Fig. 8-11).
173

KocCompton

26 24 22 20 18 16 12
DEGREES 2Θ
Fig. 8-3. Relationship between the intensity of coherent (Rayleigh) and incoherent
(Compton) scattering of Mo tube lines with sample discs having low (teflon), medium
(silica) and high (lead) mass absorption coefficients. Note that the Compton peaks are too
small to be detected with the lead disc and that the general intensity of background also
decreases as the mass absorption coefficient increases. See also Fig. 8-11.

INSTRUMENTATION

All X-ray emission spectrometers provide some means of exciting sec-


ondary fluorescence from a sample, selecting the desired characteristic line
and measuring its intensity. These goals can, however, be achieved in several
ways as shown by the sketches (Fig. 8-4) of a conventional wavelength-dis-
persive spectrometer, an energy-dispersive spectrometer, and a portable field
instrument.
In the conventional spectrometer characteristic radiation, excited by a
primary X-ray beam produced in a high-voltage X-ray tube, is dispersed by
crystal diffraction and its intensity measured at the appropriate wavelength
by a proportional or scintillation detector on the movable arm of a gonio-
meter. The geometry of the energy-dispersive spectrometer is much simpler
in that incident X-ray photons are sorted solely on the basis of their energies.
174

A. WAVELENGTH DISPERSIVE

(sample
Goniometer arm

Detector

Detector HV, Amplifier,


Electronics, PHA, Readout
B. ENERGY DISPERSIVE

(sample

°o°
Electronics,
o w Amplifier,
o
Multi-Channel
Analyzer
C. PORTABLE UNIT WITH

ISOTOPE SOURCE
Nal Crystal Filter pair on
rotating mount

Electronics m
<s~ Photomultiplier fc^l
Safety
shutter

Cable to power supply, Window


electronics and readout
Fig. 8-4. X-ray spectrometers. A. Wavelength dispersive. B. Energy dispersive. C. A porta-
ble instrument with an isotope source and filters.

This is achieved with a semi-conductor detector which both resolves and


measures the intensity of the secondary fluorescence. When the resulting
information is fed to a multi-channel analyzer, simultaneous multi-element
175

analysis becomes possible. Jenkins and De Vries (1970) and Woldseth (1973)
should be consulted for more detailed information on wavelength- and
energy-dispersive X-ray spectrometers, respectively.
Primary radiation sources
In commercially available X-ray fluorescence units sample fluorescence is
invariably excited in one of three ways: with an X-ray continuum generated
in a X-ray tube; with monochromatic X-rays produced by exposing a suitable
target to an X-ray continuum; or with λ-rays emitted by a radio-isotope.
X-ray tubes
When sufficiently energetic electrons strike a target they loose their
energy in an irregular step wise fashion and an X-ray continuum (or Brems-
strahlung) is emitted. The hardest radiation emitted, i.e. the short-wave-
length end of the continuum (X min ), is related to the energy of the electron
beam by Planck's Law:

Xmin = 1.24/V

where V is the voltage in kilovolts acrsss the X-ray tube. Peak intensity of
the continuum is at approximately 1.5 X min . The overall intensity of the con-
tinuum increases with tube voltage and filament current, and with the
atomic number (Z) of the target. If the energy of the electron beam exceeds
the excitation potential for the target's X-ray spectrum, the target element's
characteristic lines will be superimposed on the continuum.
In practice, primary X-rays are generally produced in a sealed vacuum
tube containing a W filament (the cathode) and a target which forms the
anode. A high-voltage generator, stabilized to better than 0.5%, is required to
operate the tube. A current through the filament produces electrons which
are accelerated by the tube potential towards the anode. On striking the tar-
get the electrons lose their energy producing the X-ray continuum and a
large amount of heat. The target must therefore be cooled and fabricated
from materials with good thermal characteristics. Suitable elements are Au,
W, Ag, Mo, Cu and Cr. X-rays leave the tube through a side window made
from a thin Be sheet.
Targets of heavy elements such as W produce the most intense X-ray con-
tinuum and are, therefore, often favoured for routine trace element analysis.
However, a characteristic line is best excited by wavelengths just shorter than
the corresponding absorption edge. The shortest wavelengths of the con-
tinuum are therefore not necessarily the most desirable, and the peak inten-
sity of the continuum and the presence of characteristic target lines may be
important. For example, a Mo target emitting Mo-Ktt at 0.071 nm is very
effective in exciting Rb-Ka and Sr-Ka with absorption edges at 0.0816 and
0.0770 nm, respectively.
176

X-ray tubes are normally operated close to their maximum power rating in
order to produce both the most intense continuum and t o excite the K spec-
tra of heavy elements. Occasionally, to prevent line interferences or reduce
background, it may be helpful to operate below the minimum excitation
potential for the K spectrum of a particular element.
Finally, the use of primary X-rays to excite secondary monochromatic
fluorescence, which is then used to excite characteristic radiation from the
sample, should be mentioned. The secondary fluorescence is, of course,
much less intense than the primary continuum source. However, provided
the target is chosen so that the monochromatic radiation lies on the short
wavelength side of and close to the absorption edge associated with the
characteristic line of interest, adequate intensities can be obtained. The tech-
nique is particularly useful in energy dispersive spectrometry for which the
high background scatter produced by a continuum source is unacceptable.

Isotope sources
Radio-isotopes can provide suitable primary sources when either back-
ground radiation is a limiting factor (as in energy-dispersive methods) or the
instrument must be compact and/or portable. Source strength is generally a
few millicuries. The relevant properties and analytical applications of several
useful isotopes are summarized by Gallagher (1969).

Dispersion of X-rays

Before its intensity can be measured the characteristic line must be sepa-
rated from other secondary X-rays and from scattered primary radiation.
Three methods are employed: (1) wavelength dispersion by X-ray diffrac-
tion; (2) filters with absorption edges bracketing the analytes characteristic
line; and (3) energy discrimination with semi-conductor detectors. Descrip-
tion of the last method is deferred to p. 1 8 1 .
Wavelength dispersion
In the wavelength-dispersive spectrometer the principal components of the
dispersive system are the primary and secondary collimators, the analyzing
crystal, and the movable goniometer arm carrying the detector (Fig. 8-4).
The collimators are a series of parallel metal plates arranged, rather like
Venetian blinds, to ensure that an essentially parallel beam of X-rays reaches
the crystal and detector and thereby exclude unwanted wavelengths.
Increasing the collimation, by decreasing the gap between the plates or
increasing their length, improves resolution by decreasing the range of wave-
lengths reaching the detector but also decreases line intensity.
The collimated beam of X-rays falling on the crystal is diffracted according
to Bragg's Law so that, if λ is in nanometres:

ηλ = 2d - sin Θ
177

where d, in nm is the distance between the reflecting planes of the crystal; Θ


the angle of incidence, equals the angle of reflection; and n is an integer. For
first-order reflections n is unity: high-order (n = 2, 3, ...) reflections are
much weaker but can sometimes cause spectral interferences by harmonic
overlap with analyte lines.
A variety of analyzing crystals are available and a number of factors must
be considered in selecting the most appropriate (Table 8-II). These include
mechanical and thermal stability and reflection efficiency, as well as the
d-spacing. The principal restriction on the choice of a crystal is imposed by
the upper angular limit through which the goniometer can traverse. This is
usually about 145° 2Θ so that the longest wavelength (X m a x ) reaching the
detector is:

Xmax = 2d · sin 72.50° = 1.91 d

or about 2d. Thus a LiF crystal, reflecting on the (200) crystal plane with
d = 0.2014 nm, has a X m a x of 0.4028 nm and potassium is the lightest ele-
ment that can be determined.
The ability of the spectrometer to resolve adjoining lines depends on its
dispersive power and the line widths. Dispersive power, the angular separa-
tion of adjoining lines, is given by:

d0/,dX = n / 2 d - cos Θ

Dispersion and resolution are therefore poorest when determining the

TABLE 8-II
d-spacings of some diffraction crystals used in X-ray fluorescence (based on Jenkins and
De Vries, 1970, and manufacturer's literature)

Crystal Reflection 2d Lowest atomic


plane (nm) number detectable

K series L series

Topaz (303) 0.2712 V Ce


Lithium fluoride (220) 0.2848 V Ce
Lithium fluoride (200) 0.4028 K In
Penta-erythritol (PET) (002) 0.8742 Al Rb
Ethylenediamine tartrate (020) 0.8808 Al Br
Ammonium dihydrogen phosphate (ADP) (110) 1.065 Mg As
Gypsum (020) 1.519 Na Cu
Thallium acid phthalate (TLAP) (100) 2.575 O V
Rubidium acid phthalate (RAP) (100) 2.612 O V
Potassium acid phthalate (KAP) (100) 2.663 o V
178

heaviest elements towards the shorter-wavelength end (low 20) of the spec-
trometer. Since λ is proportional 1/Z2 this is also a region where character-
istic lines are relatively closely spaced thereby increasing the possibility of
spectral interference. Consequently, providing some loss in line intensity can
be tolerated, it may be desirable to improve resolution by using increased
(finer) coUimation. Conversely, light elements with low fluorescent yields are
best determined using coarse coUimation in conjunction with crystals dif-
fracting their characteristic lines at large values of 20.

Filters
The simplest methods of isolating a particular characteristic line is with
balanced filters having absorption edges immediately adjacent to and on
either side of the analyte's characteristic line. A typical example would be
the use of a Ni-Co filter pair to isolate Cu-Ka which is transmitted by the Ni,
but not the Co, filter (Fig.8-2). The filters must be carefully matched, by
adjusting their thickness, so that they transmit equal intensities except for
wavelengths lying between their absorption edges. Readings are taken first
with one filter in position and then the other, the difference in readings
giving the intensity of the characteristic line. If the element to be determined
is a major constituent, for example Fe, giving rise to almost all the short
wavelength radiated emitted by the sample, it may only be necessary to iso-
late the characteristic line using the long-wavelength filter.
Use of filters is restricted to compact, portable XRF units, a purpose for
which their simplicity is ideally suited (Fig. 8-4). Generally such instruments
can be supplied with sets of interchangable filter discs, each containing
several filter pairs chosen by the analyst.

Detectors
The detector has the dual role of both measuring X-ray intensities and
sorting them on the basis of their energies. Measurement of X-ray intensities
is dependent on the ability of X-rays to ionize matter, releasing electrons and
creating a pulse which can be amplified and measured. Sorting them on the
basis of their energy is possible because the voltage (v) of the pulse produced
by an X-ray photon entering the detector is proportional to the energy of
the incident X-ray photon (e0), that is:

v °c e0/ei

where ex is the ionization potential for the detector. Three types of detectors
are in general use: gas-filled proportional counters, scintillation counters, and
semi-conductor detectors.

Gas-filled proportional counters


These detectors contain an anode wire along the radial axis of an earthed
179

chamber, containing an inert gas, which X-rays enter through a thin mylar
window. The thinness (less than 6 μιη) of the window makes it impractical
to exclude air or prevent gas leakage. The inert gas, usually argon mixed with
methane, is therefore allowed to flow continuously through the chamber.
On entering the detector an X-ray photon ionizes inert gas atoms (e{ ^
20 eV) to produce primary ion pairs:

Ar° -> Ar+ + e'

the average number of primary ion pairs depending on the energy (e0) of
the photon. The potential applied to the anode causes the electrons to
accelerate towards it, resulting in collisions between accelerated electrons
and inert gas atoms that produce secondary ion pairs and give a cascade
effect known as gas amplification. In the region of proportional response the
voltage of the amplified pulse produced by the counter remains proportional
to e0. At higher anode voltages this relationship breaks down and the detector
behaves as a geiger counter.

Scintillation counters
In a typical scintillation counter X-rays penetrating a Be window are
absorbed by a thallium-activated sodium iodide crystal causing it to phos-
phoresce in the blue region of the visible spectrum. Some of the emitted
light photons strike a photomultiplier, mounted in optical continuity with
the crystal, to produce a voltage pulse. The amplitude of the pulse depends
on the efficiency of the detector, for which the effective e{ is about 50 eV,
and on counter voltage.

Comparison of gas proportional and scintillation counters


Most wavelength-dispersive X-ray spectrometers permit a choice of either
a gas proportional or scintillation counter to measure secondary X-rays. In
either case the relationship between the energy of the incident X-rays (e0)
and amplitude of the pulse produced by the detector can be used to reject
any unwanted X-rays entering the detector (pulse height analysis — PHA). It
is, therefore, informative to compare detector performance with respect to
(1) efficiency at different wavelengths; and (2) ability to resolve X-ray
photons of different energies.
Variations in detector response at different wavelengths are largely a func-
tion of the X-ray absorption characteristics of the detector and its window.
For the best response at a particular wavelength the incident X-rays should
all penetrate the window and be absorbed, causing ionization, in the detector.
However, some photons may fail to penetrate the window whereas harder
radiation may travel right through the detector without causing ionization.
Attenuation of X-rays in the air-path of the spectrometer and by absorption
at the detector window are particularly critical in measurement of soft radia-
180

tion. Consequently, to measure Ka lines of elements lighter than Fe it is best


to use a gas flow proportional counter, with a thin window ( < 6 μπι)
mounted in the spectrometer's vacuum chamber. A very thin window
( < 3 μπι) is needed if Na-K a or F-KQ, are to be measured. Harder radiation is
measured with either a sealed proportional counter or scintillation counter.
In some spectrometers a scintillation counter can be mounted in tandem
behind a gas flow proportional counter thereby ensuring optimum counting
efficiency over a wide range of wavelengths.
Detector resolution is determined by the energy (e0) of the incident X-ray
photon and the effective ionization potential (e^ of the detector in the rela-
tionship:

n = e0/ei

where n is the number of electrons initially produced when the X-ray photon
enters the detector. Like radioactive decay, the phenomenon is random and
n has a Poisson distribution so that the standard deviation on = \fn. With
typical values of n the Poisson distribution approximates a normal, Gaussian
distribution and counter resolution (r) is defined as:

2.35σ„Χ100 w X 100
r(%) =

where w is the full-width-at-half-maximum-height (FWHM) and v is the pulse


amplitude or voltage (Fig. 8-5). Because n changes with e 0 , counter resolu-

Natural Peak

VOLTAGE
Fig. 8-5. X-ray pulse amplitude distribution of a detector. Detector resolution (%) is
defined as 100 X w/υ By setting the detector electronics to a lower threshold voltage (Vi)
together with a window (V2 — Vi), pulses lying below V\ and above V2 can be rejected.
181

tion varies with wavelengths and a particular wavelength must be chosen to


compare detector resolution: the Mn-Ka or Fe-K^ lines are often used. The
effective ionization potentials of gas proportional counters and scintillation
counters are approximately 25 and 50 eV, respectively.

Semi-conductor detectors
The development of semi-conductor detectors is a relatively new field
which has advanced rapidly and several energy-dispersive X-ray analyzing
systems are now available. Their outstanding advantage, compared to wave-
length-dispersive spectrometers, is that true simultaneous multi-element
determinations become practical with a relatively inexpensive instrument. A
serious disadvantage is that the detector must be maintained at low tempera-
tures, using liquid nitrogen, to prevent its deterioration.
In X-ray spectrometry the semi-conductor detector is a lithium-drifted
silicon crystal between two metal electrodes. A high voltage applied across
the electrodes produces a large electric field in the detector. When an X-ray
photon enters the detector ionization occurs and the resulting ions migrate
to the electrodes. The effective ionization potential is much lower (about
3 eV) for semi-conductor detectors than for gas proportional or scintillation
counters. Consequently, their resolution is much greater. For example,
resolution of Cu-K^ with a scintillation counter, a proportional counter and
with a semi-conductor detector, would be about 65%, 15% and 5%, respec-
tively. Thus, with a semi-conductor detector it is possible to divide the X-ray
spectrum emitted by a sample into a number of relatively narrow energy
channels; the X-ray intensity within each channel is stored, processed and
displayed by a multi-channel analyzer.
Energy-dispersive spectrometry permits a much simpler sample-detector
geometry than can be used with wavelength dispersion (Fig. 8-4). This con-
siderably improves counting efficiency, permitting use of less intense
primary sources. However, rather than measuring only a narrow band of
X-rays dispersed by an analyzing crystal, the semi-conductor must handle all
incoming X-ray photons emitted by the sample. Consequently, a particular
characteristic line may account for only a small proportion of the total num-
ber of counts — perhaps a few hundred or a few thousand out of up to
50,000 cps — registered by the detector. Under these conditions it is desira-
ble to reduce the contribution of background radiation as much as possible.
The largest single source of background is normally the back-scattered con-
tinuum and characteristic radiation from the X-ray tube. Background can
therefore be greatly reduced if an isotope source or monochromatic radia-
tion, generated by bombarding a secondary target with primary X-rays, is
used to excite the sample.
182

OPERATION OF THE X-RAY FLUORESCENCE SPECTROMETER

Instrument conditions

Having described the components of the X-ray spectrometer, it is appro-


priate to consider operation of the complete instrument. This, however, will
be confined to general remarks on wavelength dispersive instruments because
these are the type most often available in geological laboratories. As might
be expected, operation of the portable units intended for field analysis is
relatively straightforward.
The main operating variables with a wavelength dispersive spectrometer
are: (1) the primary X-ray tube target and the intensity of the primary radia-
tion; (2) selection of a dispersing crystal; (3) use of a flow proportional or
scintillation detector; and (4) the counting time. Because the response of
either type of detector is related to the energy of the incident X-ray
photons, pulse height analysis can be used to discriminate against unwanted
X-rays by appropriate adjustment of the detector electronics. In practice,
choice of operating conditions is largely fixed by the elements (and hence
wavelengths) to be measured (Fig. 8-6). Providing only elements heavier than
Cr are of interest a W target X-ray tube, a LiF (200) crystal and a scintilla-
tion detector would be standard. No vacuum would be required and with a
100 kV generator it would be possible to excite Ka radiation up to Bi
(excitation potential 90.44 kV; Table 8-1). The same conditions could also
be used to measure Th-L^ (20.42 kV) and U-L ai (21.72 kV). For elements
lighter than Cr, as in whole rock analysis, a Cr target X-ray tube and flow

o o o o o o o q o r- N n n ^ m Φ N o <j> o
do odd d ö ό ö ό o o d d ό ό ό ό ό^
Mill I I 1I I I I I I I I I I I I I J L
ZnNi
w Dy
__|
Ba
I
Sn Ag
I I
Mo
I
Sr
i
As GalCul Fe Cr Ti
I I I ιΠ ii I I I I
CaK Cl S P Si AI Mg Na
I I I I I
Th PbAuW Dy Ba Sn Ag Mo

LIF(220)
- TOPAZ ~
-LIF(200)-
_E.D.D.;r_
CRYSTAL
PE
--+ GYPSUM* rt
* KAP-
S.C.-
F.C.
DETECTOR

COLLIMATOR COARSE-

- AIR
MEDIUM -VACUUM -

Fig. 8-6. General operating conditions with a wavelength-dispersive X-ray spectrometer.


(From Jenkins and De Vries, 1970.)
183

proportional counter, in conjuction with LiF (for Fe, Ti and Ca), PET (Al
to K) and TLAP (Na and Mg) crystals, provide optimum conditions.
The principal remaining operating conditions relate to the detector, in
particular to the detector potential and the settings for pulse height analysis.
Optimum detector voltage should be selected for each element following the
manufacturer's instructions — generally this involves setting the goniometer
20 position for the peak to be measured and with a sample in place gradually
increasing the detector voltage. The count rate should initially increase
rapidly until a plateau is reached. Optimum voltage is near the middle of this
plateau. Hutchison (1974) gives an excellent description of optimizing
counter voltage and pulse height analysis with the Philips 1540 spectrometer.
Pulse height analysis often seems to be a subject of some confusion and,
since its misapplication is probably worse than no use at all, it warrants a
rather detailed description. It will be remembered from p. 178 that the
voltage of the pulse produced by both proportional and scintillation detectors
is proportional to the energy (e0) of the X-ray photon. Pulses corresponding
to X-ray photons having different energies can therefore be sorted and
rejected by setting the detector electronics to a lower threshold value and a
window (Fig. 8-5). Only pulses having energies lying above the threshold and
within the window are recorded.
It remains to account for the unwanted pulses registered by the detector.
These usually arise as a result of either harmonic repetition of the X-ray
spectrum or from fluorescence of the components of the spectrometer.
From the Bragg equation higher-order (n > 1) lines of heavy elements can
coincide with or overlap a first-order analytical line. However, their energies
will be very different. For example, first-order Mn-Ka (0.2103 nm) is partly
overlapped by the second-order reflection of Au-L^2; their respective energies
are 5.88 kV and 11.57 kV. Because of the large energy differences involved,
pulse height analysis with a flow proportional counter can completely
separate the lines. With the poorer resolution of a scintillation counter some
overlap remains — the calculations to demonstrate this are given in a worked
example by Jenkins and De Vries (1970). The intensity of second-order and
higher lines is relatively weak so the harmonic overlap is unlikely to be a
problem unless light elements are to be determined in heavy matrices.
Instrument fluorescence can arise if the sample emits or scatters X-rays at
wavelengths shorter than the absorption edges associated with the con-
stituents of the dispersing crystal or some other component of the spectrom-
eter: these can then fluoresce and become a source of extraneous pulses.
Crystal fluorescence is not a problem with the LiF crystal. It might, how-
ever, for example be encountered in the measurement of P-Ka with a
gypsum crystal — P-Ka, Ca-K^ and S-Ka would all reach the detector.
Unwanted pulses can also arise within an argon-filled proportional
detector if an X-ray photon, giving rise to a natural peak (i.e. the peak to be
measured), has sufficient energy (e0) to eject an electron from the argon
184

K-shell. The ejected escape electron has an energy (ee) where:

ee = e0 — eA

when eA is the energy associated with the argon K-shell absorption edge. The
escape electron produces a pulse on the low-energy side of the natural peak,
the separation of the two peaks increasing as the energy difference between
the escape electron and eA decreases. In practice, escape peaks are encoun-
tered for elements K to Ni with argon-filled detectors.
With sodium iodide scintillation detectors escape peaks should be asso-
ciated with ejection of K- or L-shell iodine electrons with absorption edges at
0.0374 nm and 0.2389 nm, respectively. However, the scintillation detector
would not normally be used at wavelengths longer than Fe-Ka (0.1936 nm)
and only elements heavier than La will give rise to the K-shell escape elec-
trons.

Counting strategy

Having arrived at the appropriate instrumental conditions for a particular


analysis it remains to decide the counting period. This will be a function of
both the range of concentrations expected and the count rate, longer
counting periods being required to obtain reliable results at low concentra-
tions or low count rates. Obviously, however, we do not wish to count for
longer than necessary. Fortunately, though complicated by the presence of
background, there is a simple relationship between the number of counts
accumulated and the reliability of the results. This provides a rational basis
for deciding the counting strategy.
If repeated measurements were made at a constant concentration the
average number of counts (N) accumulated in a constant time and the count
rate (R, cps), would follow a Poisson distribution. Hence the standard devia-
tion (QJV) = VN. Assuming that N ^ N the relative standard deviation or
counting error (E) is then:
E-y/W/N=l/VW
but N = RT, where T is the counting time, so:

E = 1/y/RT and E% = 100 X Ι/Λ/RT


For example, the counting error with R = 10 cps and T = 5 s will be 14%,
which the analyst can reduce to 7% by increasing T to 20 s. Thus, always
remembering that errors arise at all stages of analysis, the component due to
counting can be reduced until either the benefit becomes negligible com-
pared to other sources of error or until instrument instability becomes a
limiting factor.
185

So far background has been assumed to be absent. However, for trace


analysis the background intensity, giving a count rate of Rh, is likely to be a
significant component of the total intensity. From the foregoing, the stan-
dard deviation of the count rate (oR) is:

oR=ER= y/RjT

With background present two measurements must be made: one at the


peak, giving a value (Rp) due to background rate (i? b ) plus analyte signal
(i? a ), and another at an adjoining line-free wavelength to estimate (.Rb) alone.
The standard deviations of Rh and Rp are given by:

oh = VRJT~h and σρ = y/Rp/Tp

and, since variances are additive, the standard deviation (a a ) for the analyte
signal (E a ), with Ra = Rp — Rh, is:

tfa = >/σΓ+ ol

■m+
T ' τ b ρ

and relative error (E%) becomes:

E(%)
V¥kl·'-** X 100

If counting time is the same at the background and peak positions (i.e. T p
T b = Γ/2) this can be rearranged to give:

_ 100^2 V E p + Rl
E(%)

^ρ ^b

As usual the detection limit is defined as the background signal plus two
standard deviations: in this case Rh and a b , respectively. However, as noted
above, two measurements must be made — one at the peak (Rp) and one at a
background position (Rh). The standard deviation for the analyte (a a ) is
therefore greater than the standard deviation for background ( a b ) alone, by
a factor of:
186

At the detection limit Ä p ^ Rh so that this factor is close to \/2 and the
detection limit becomes the concentration giving a net count rate (Rp ■ A b )
equal to:

2v^2 · VEbTn - 3yfllJT*


For a more detailed discussion of counting strategies Jenkins and De Vries
(1970) should be consulted.

ANALYSIS OF GEOCHEMICAL SAMPLES

Detection limits attainable with geochemical samples are summarized in


Fig. 8-7 from Leake et al. (1969). It is apparent that although the detection
limits for elements of low atomic number, notably Na, Mg, AI and Si, are
severely curtailed by the Auger effect, it should be possible to detect many
of the trace elements of interest in exploration samples (Fig. 1-11). How-
ever, because of the wide compositional variations of geochemical samples and
differential excitation of different components of individual samples as a

? 0,000 360,000

Detection limit at 95% confidence for 100 seconds

1,000

- 100

10

20 30 40 50 60 70 Θ0 90 100
Ca Zn Zr Sn Nd Lu Hg Th
Atomic number
Fig. 8-7. Detection limits (ppm) by XRF at the 95% confidence level: based on 100 sec-
onds counting time for pressed powder pellets. Broken lines join results between which
unmeasured elements occur. (From Leake et al., 1969.)
187

result of textural and mineralogical effects, it is only possible to obtain


reliable results if sample preparation is designed to minimize textural varia-
tions and if corrections are made for mass absorption differences between
samples and standards. These topics are the main subject of this section.
Some examples of the applications of XRF to analysis of exploration sam-
ples are summarized in Table 8-III. XRF is also ideally suited to the deter-
mination of major elements, as required in studies of alteration or as a means
of normalizing variations in trace element contents to bulk composition.
(Examples of major element geochemistry being used in these ways are to be
found in Davenport and Nichol, 1973; Gunton and Nichol, 1975; Wolfe,
1975; and Olade and Fletcher, 1975.) For these purposes the very accurate
whole rock analyses used in petrography, requiring fusion of the sample to
give a homogeneous glass disc in which mass absorption coefficient variations
are greatly reduced (Hutchison, 1974), are seldom necessary and quite
adequate results can be obtained, without matrix corrections, providing
pelletized samples (or powders) are analyzed against standards of approxi-

TABLE 8-III
Analysis of geochemical samples by X-ray fluorescence
A. Single-element methods

Element X-ray Analyzing Line Total Detection Reference


tube crystal counting limit
time (s) (ppm)

S Cr PET S-Ka — — Fabbi and Moore


(1970)
Zn Cr LiF(200) Zn-K a 80 — Bergseth (1975)
U W LiF(200) U-L a 100 100 Clark and Pyke
(1972)
U Mo LiF(220) U-L a 100 1.2 James (1977)
Th Mo LiF(220) Th-L a 100 1.5 James (1977)
Sn W LiF(220) Sn-K a 120 1 Goodman (1973)
B. Multi-element methods *

Elements Reference

Ni, Cu, Zn, As, Br, Pb Leake and Peachey (1973)


Si, Ca, Ti, Mn, Fe, Ni, Cu, Zn, Rb, Sr, Y, Zr, Nb,
Ag, Sn, Sb, Ba, Ce, Pb, Th and U Leake and Aucott (1973)
Ga, Rb, Sr, Th, Pb and others Feather and Willis (1976)
Nb, Ta, W, Sn, Mo, Th, V, Ba and Ti Levinson (1975)
1
Only papers relating to analysis of exploration samples are referenced: there is a con-
siderable literature on determination of trace elements in rocks for petrological studies.
188

mately similar composition. If all the major elements are to be determined,


their approximate concentrations can be refined by successive iterations of
their estimated mass absorption coefficients and corrected concentrations
(Brownetal., 1973).

Sample preparation

The simplest approach to sample preparation for rapid analysis is to pour


the sieved or ground sample powder into one of the holders, usually an
aluminium or plastic cup with a replacable mylar window, provided with the
spectrometer. The window should be kept as taut and wrinkle free as possi-
ble but does not normally need to be changed between samples. Alterna-
tively, sample powder can be mixed with a suitable binding agent (starch,
cellulose, or one or two drops of poly-vinyl alcohol) and pressed into a pellet
with a small hydraulic press. Pelletizing has the advantage of giving some-
what greater intensities, more reproducible results and enabling exactly the
same sample to be re-analyzed. However, the additional effort is certainly
not always warranted.
Irrespective of whether powder or pellets are analyzed it is important that
particle size and mineralogical effects (Fig. 8-8) be avoided, and that the
depth of sample in the holder exceeds the infinite thickness for the radiation
being measured. Providing the sample is ground sufficiently fine that the
depth from which the characteristic radiation emerges is much greater than
grain size, results should be essentially independent of particle size effects,
i.e. the intensity of the characteristic X-ray will reflect the composition of a
sample volume representative of the whole sample. For typical silicates

β-α
mjtjb.
n . v Secondary X-rays
J
Primary X-rays '
Fig. 8-8. Sample particle and mineralogical effects in XRF. In a sample consisting of
rounded grains in a finer matrix, characteristic X-rays with a penetration depth (x), less
than the diameter of the grains, will underestimate the concentrations of constituents
contributed by the grains. To avoid this effect samples should be ground until x >> grain
diameter.
189

grinding to 50 μιη or less will prevent particle size effects with wavelengths
shorter than about 0.3 nm. Where grinding is not feasible and particle size
effects are suspected, Jenkins (1970) has suggested a correction based on the
intensity ratio of an element's K and L, or L and M lines.
As a check on the amount of sample required, varying amounts of powder
(or pellets of increasing thickness) can be loaded into the holder until no
further increase in intensity is observed. Feather and Willis (1976) found
that with pellets of pure quartz, 40 mm in diameter, 8 g of material sufficed
for Mo-Ka (0.071 nm) whereas 20 g was required with Ba-Ka (0.039 nm).
Grinding of a few tens of grams of material in a ring mill will therefore
usually provide an ample quantity of sufficiently fine powder.

Matrix corrections

Count rates obtained with unknowns are normally converted to con-


centrations by comparison to count rates obtained with natural or artificially
prepared standards. However, because of the wide range of bulk composi-
tions encountered in geochemical analysis it is essential to correct for differ-
ences between background intensity and mass absorption coefficients of
samples and standards.
When bulk composition is known an average mass absorption coefficient
(μχχ) can be derived for any analyte wavelength (λ) and estimates of con-
centrations corrected accordingly: i.e., where Wx is the weight fraction of an
element or oxide with mass absorption μ1χ at wavelengths λ:

μχχ = ΐν,(μ 1λ ) + Ψ2(μ2λ) ...

and corrected concentration (Cx) is given by:

where Ι8χ and Ιχχ are intensities (cps) on the standard and sample respec-
tively, μ8λ and μχχ are the corresponding mass absorption coefficients, and
Cs is the concentration of the analyte in the standard. Brown et al. (1973)
describe an iterative procedure, employing successive approximations of the
sample's mass absorption coefficient and major element content, for accu-
rate analysis of pelletized samples without resort to fusion methods.
Unfortunately, if only trace elements are of interest the bulk composition
of samples will seldom be known and some other method of estimating
matrix effects is therefore needed. The relationships between scattered radia-
tion and mass absorption (p. 172) provide several methods of making the
necessary corrections; before describing these it is useful to consider the
X-ray absorption spectrum of silicates as presented by Hower (1959).
190

A plot of mass absorption versus wavelength for a rock shows discon-


tinuities at 0.1744 nm and 0.2498 nm representing the Fe-K and T-K absorp-
tion edges, respectively (Fig. 8-9). These discontinuities divide the curve into
three regions; region one, with wavelengths shorter than the Fe-K absorption
edge, includes most of the trace elements of exploration interest. Hower
(1959) showed that if relative absorption coefficients — rather than absolute
absorption coefficients — are plotted the discontinuities remain; however,
the relative absorption between the discontinuities is virtually constant at all
wavelengths. This led to the important conclusion, which forms a basis for
the matrix correction methods to be discussed, that a single measurement of
a specimen's relative mass absorption coefficient at any suitable line-free
wavelength shorter than the Fe-K absorption edge can be used to estimate
mass absorption at all other wavelengths shorter than the absorption edge.
Matrix correction using peak to background ratios
Between absorption edges the mass absorption coefficient is approxi-
mately proportional to Z 4 so that line intensity (Ιγ) is approximately propor-
tional to Z~4. Also intensity of background scattered radiation (I b ) is
approximately proportional to Z" 3 t o " 2 ; hence:
Z
^1 ~ ~ z-(l to 2)
α Ζ
4 ζ " ^
The peak to background ratio should therefore be less sensitive to mass
absorption changes than IY. The usefulness of this relationship was first
demonstrated by Anderman and Kemp (1958): measuring background
scatter at an arbitrary wavelength of 0.06 nm they obtained a linear (i.e.
matrix independent) calibration curve for the ratio Pb-L a // 0 .o6 versus Pb
concentration in a series of lead ores with varying Fe and Zn contents
(Fig. 8-10). They also noted that Ii/Ih is much less sensitive to instrumental
fluctuations than IY and that Ih therefore acts as an internal standard.
As used by Anderman and Kemp (1958) the ratio method neglects the
contribution of scattered background radiation to analyte line intensity. If,
however, trace elements are being determined the background intensity at
the spectral line (7lb) can be an important component of the total intensity
(Ix). Champion et al. (1966) therefore refined the ratio method by measuring
background intensity on either side of the characteristic line (in their case Sr-
K a ) and interpolating to obtain an estimate of 7 lb at the peak position. True
peak intensity due to the analyte (Ja) is then:

'a = / l - / l b

and for the ratio method:

7 7
K ib ib
191

Ca Absorption Edge
280

z z
ο 220 H Ti Absorption Edge- o
t-
Q.
OC
a.
OC

o O
</>
Fe Absorption Edge </>
03

<
03
<

O Relative Absorption
c/>
03

<

Absolute Absorption

0.20 0.30

1
Fe Absorption Edge

z
o
Q. Ca Absorption
18
O Ti Absorption 9
c/>
m
< Edge
LU
-· ♦ ♦ * *-
>
5 1.4·
_l
UJ
Diabase
Shale
Ounite
Granite

~r
0.15 0.20

WAVELENGTH, nm

Fig. 8-9. The X-ray absorption spectrum of typical silicates. A. Absolute mass absorption
coefficient of a dolerite and its absorption relative to that of alumina. B. Absorption
relative to alumina for a granite, diabase, shale and dunite. (From Hower, 1959.)

Because of its simplicity, Levinson (1975) has recommended the ratio meth-
od for analysis of exploration samples: using a 50-kV generator and 20-sec-
ond counting he obtained detection limits of 50 ppm for several elements.
Lower limits are achievable with more powerful generators and longer
counting times. In adopting the ratio method careful consideration should be
192

3- B

2-
Q. Y IFe 7„Zn
0-1 0-1
JL
/ ■
· 1 -7 0-1
1- ▲ 1 -7 2-6
+ 4-7 60

L50 100 150


—i
0.5
i
1.0
1
1.5
o 40

r~
2.0
1

PbLoc, INTENSITY PbLoc.,/-O.06nm SCATTERING

Fig. 8-10. Calibration curves for Pb in ores. A. Absolute calibration, with %Pb plotted
against the intensity of Pb-L a , showing the wide scatter caused by the effect of varying
Fe and Zn contents on the mass absorption coefficient. B. Scatter has been considerably
reduced using the ratio method and plotting %Pb against the ratio of Pb-L a to intensity of
scattered radiation at 0.06 nm. (Reprinted with permission from Anderman and Kemp,
1958, Scattered X-rays as internal standards in X-ray emission spectroscopy, Anal. Chem.,
30: 1306—1309. © 1958 American Chemical Society.)

given to the choice of the wavelength(s) for background measurement —


the following conditions are required:
(1) The wavelength should lie on the same side of the absorption edges of
major constituents of the sample as the characteristic radiations to be mea-
sured, i.e. for K^ lines of elements heavier than Ni background should be
measured at a wavelength shorter than the Fe-K absorption edge (0.1744 nm)
(because Fe is the major component with the greatest mass absorption
coefficient in most geochemical samples).
(2) The wavelength chosen should be free of any characteristic lines
emitted by the X-ray tube or sample.
The best results are obtained if background is measured at interference-
free positions adjacent to and on either side of the analyte line. However,
the additional effort is probably not warranted in routine analysis, and sim-
pler and more rapid methods of estimating background scatter from a single
measurement are described next.

Matrix correction with Compton scatter


Part of the characteristic radiation emitted by the X-ray tube is incoher-
ently scattered by the sample to give a Compton peak located, with conven-
tional spectrometer geometry, 0.0024 nm on the long-wavelength side of the
coherently scattered tube line (Fig. 8-3). Intensity of the Compton peak
(I\c), at wavelength Xc, is proportional to 1/M\C (Fig. 8-11) and from Hower
(1959) μλ will have a constant ratio to μλ provided λ and Xc lie on the same
193

F
50-
e 2 <VX

40-
Magnetites x\

ί 30
(0

y^Fe-mica
XBCR
10-
>XVGSP
0
c) 10 20 30
10 5
cps MoK Compton

Fig. 8-11. Mass absorption coefficient for Sr-K a versus Compton scattering of the Mo-K a
tube line using a LiF(200) crystal. The slight curvature, that results from the difficulty in
resolving the small Compton peak from the coherent peak in samples with a high mass
absorption coefficient (Fig. 8-3), could be reduced using a LiF(220) crystal. (Data pro-
vided by R.L. Armstrong.)

side of any major element absorption edges. Consequently, Reynolds


(1963), using a Mo target tube, was able to estimate the mass absorption
coefficients of samples, for an arbitrary wavelength of 0.09 nm, by com-
parison of the intensity of their Compton scattered Mo-K a peak to scattering
by standards of known composition and hence known mass absorption. The
mass absorption coefficients so obtained were used to correct estimates of
concentrations of elements ranging from Ni-K a to Ag-Ka with the relation-
ship:

C - λ
.^*0·09. r
8
λ ^s0.0

Using a silver-target tube the range can be extended up to Mo-K a and would
also include the La lines of Hf to U (Feather and Willis, 1976). Measurement
of Compton scatter is therefore a simple and rapid method of correcting
mass absorption differences for many elements of exploration interest.
Principal shortcomings of the original Compton scattering method, as
subsequently elaborated by Reynolds (1967), were:
(1) Samples with μ 0 .ο9 l e s s than 7 could not be handled because of the
difficulty of achieving infinite thickness in the sample cup.
(2) Samples with μ 0 .09 greater than 20 were excluded because the
Compton peak, lying on the shoulder of the more intense Mo-K a , could not
be reliably measured; and
194

(3) Mass absorption coefficients could not be extrapolated across the Fe-K
absorption edge to provide corrections for the lighter trace elements.
Solutions to each of these problems were suggested. For light, organic-rich
matrices Ryland (1964) has shown that Compton scatter can be success-
fully measured using the softer radiation of a chromium-target tube. With
high mass absorption coefficients, which would be expected in heavy mineral
concentrates, gossans or other iron-rich materials, separation of the Mo-Ka
Compton peak from the Mo-Ka peak can be greatly improved by replacing
the LiF(200) crystal with a LiF(220) crystal. This provides better resolution
with some loss of intensity. The separation can be further improved by
installing an yttrium oxide filter, for which Reynolds (1967) provides
detailed instructions, at the detector entrance slit. Mass absorption correc-
tions can then be made up to values of at least 80 for Mo.09·
The remaining problem, that of extrapolating μχ values across major ele-
ment absorption edges, was also dealt with by Reynolds (1967) with respect
to the estimation of Mn, Cr, V, Ti and Sc on the long-wavelength side of the
Fe-K absorption edge. A somewhat different approach, which also allows for
enhancement of the Ka lines of a trace element as a result of their excitation
by Fe-K radiation, has been described by Giauque et al. (1977).
It will have been recognized from the foregoing descriptions of the back-
ground ratio and Compton scatter methods that they have much in common.
In fact, the latter is a special case of the former. On this basis Feather and
Willis (1976) have proposed a simple, elegant method of background and
matrix correction based on determination of background intensity at a single
point. This can be either a Compton peak or, where this cannot be used
some other interference-free background intensity (Ih) on the short-wave-
length side of the Fe-K absorption edge. Their method is well suited to
exploration samples and therefore warrants a fuller description. The basic
premises are:
(1) Background results from scattering of the continuum emitted by the
X-ray tube and its intensity is therefore inversely proportional to the mass
absorption coefficient (Fig. 8-11).
(2) Ratios of mass absorption coefficients measured at any two wave-
lengths between major element absorption edges are constant (Hower,
1959).
(3) There is therefore a constant relationship between background mea-
sured at any two positions within the same wavelength region (Fig. 8-12).
These relationships can be used to estimate the background (7lb) and net
peak intensity (7a) for any analyte line. The first step is to use blank pellets
of pure compounds, covering a range of mass absorption values, to relate
background intensity at each line (7lb) to that (Ih) at the position chosen to
monitor background (Fig. 8-13). There is a residual component (R) to these
plots — probably arising from scatter within the spectrometer — so that true
background 7 lb , which equals mlh when m is the slope in Fig. 8-13, must be
195
'MgO
150
Al 2 0 3 >
\~

c Z MgO
D UJ
O I-
Ü z
100 A Q AI2(W^ ;,
o NaCI .y z3
O CaCOo*^
KCI« s
AC

>Ti02
V205 v ^ T i 0 2 Si
ω >^Fe 2 03 B
3 Cr290 V
3: < u0
o <^
Fe 2 0 3 _l *l
cc
O
Ag X-ray tube LiF(220) <c
o j R
H I Residual
< υ
(0 I
QQ
100 200 300 400
L. · ; : t... I
BACKGROUND AT 35°, counts/s INTERFERENCE-FREE BACKGROUND INTENSITY, l b

Fig. 8-12. The linear relationship between two backgrounds at different wavelengths,
illustrated using blank pellets having different mass absorption coefficients. (From
Feather and Willis, 1976.)

Fig. 8-13. jEstimation of the true spectral peak background intensity (I\h), and the total
background intensity (B - J l h + R), from a plot of the intensity measured at an interference-
free background position (Jb) and the spectral peak position using blank pellets of pure
compounds representing a range of values for the mass absorption coefficient. (From
Feather and Willis, 1976.)

distinguished from total background (/ lb + R), where:

J lb + R = mlh + R
It should be noted that with typical silicate matrices R will be negligibly
small if 7b is measured at a Compton peak. Net peak intensity (/a) at the
analyte line is given by:

/.=/p-(/ib+Ä)=/p-(mib+Ä)

where Ip is the measured peak intensity. For the ratio method:

r .
1
= [* _Jp-(m/b+Ä)
ratio
■»rat.ir» ~ — -
mlh

so that concentration (Cx) for an unknown is given by:

c _s[/ p -(m/ b +i?)]


196

where s is the slope of a calibration plot of IJhh versus concentration


prepared with standard samples.

Line interferences

Line interferences usually arise from two sources: contaminants in the


X-ray tube-emitting characteristic radiation and the coincidence or overlap
of lines emitted by the sample. Some line interferences encountered with
typical silicates are summarized in Table 8-1V. These can be corrected for by
measuring the contribution of the interferent at the analyte peak position in
a pellet free from the analyte. Common tube contaminants are Fe, Ni, Cr,
Cu and Zn: their characteristic lines can be filtered out of the primary radia-
tion by mounting an aluminium filter at the tube window. Filters can also be
used to remove characteristic radiation of the target element, for example a
titanium filter at the window of a Cr target tube will enable Cr to be deter-
mined.

Summary

The foregoing discussion of the application of XRF to geochemical analy-


sis has been based on the use of wavelength-dispersive spectrometers.
Because detection limits for many elements are somewhat poorer than can
be attained with ES or AAS, XRF is best suited to analysis of materials
either with relatively high concentrations of trace elements, for example,
panned concentrates (Leake and Aucott, 1973), or organic-rich samples with

TABLE 8-IV
Line interferences in X-ray fluorescence analysis of geochemical samples (based largely
on Leake et al., 1969; Webber and Newburry, 1971; and Feather and Willis, 1976)

Analysis line Interfering line Analysis line Interfering line

As-Ka Pb-L Sc-K a Ca-Kßl5


Ba-L a Ti-K U-L a Rb-K a
Ba L Ce-L V-K a Ti-K^
- 01,4
Co-K a Fe-K ßl V-K0 Cr-K a
Cr-K a V K
- 0l,3
Υ
"Και 2 Rb-K ß l 3
Mn-Ka Cr-K^ Zr-K a Sr-K0 1 3
Μο-Κα Zr-Kß Ag-Ka Compton U-L 7 , Ru-K^
Nb-K a Υ-Κ0 Mo-K a Compton U-Lp.Nb-Ka, Y-Kp
Ni-Κα W-LL (tube line) Au-L a Compton W-Ljj, Zn-Kß
Ρ-Κα Ca escape line

Lines overlaps can often be reduced by using a LiF(220) crystal rather than LiF(200).
197

low mass absorption coefficients (Leake and Peachey, 1973). For these
materials and for trace elements, such as Ba and Sr, which occur in most geo-
chemical samples well above their detection limits by XRF, the ease of sam-
ple preparation is a considerable advantage when information on the total
content of an element is required. XRF is also ideal for rapid determination
of major, rock-forming elements when information on the bulk chemical
composition of samples is required.
Feather and Willis (1976), using a multi-channel spectrometer with a 160-
position automatic sample changer, were able to analyze 25,500 samples for
three elements in 54 working days. However, a multi-channel spectrometer
represents a major capital investment (Table 1-IV) and analytical throughput
with the less expensive sequential spectrometers, in which the goniometer
arm is programmed to move the detector from one peak position to the
next, would be appreciably lower. It is, therefore, of considerable interest
that advances in energy-dispersive spectrometry have reached the state that,
using backscatter corrections for mass absorption variations, quite reasonable
multi-element results can be obtained. Hansel and Martell (1977), for exam-
ple, determined Ni, Cu, W, Pb, Bi, Nb, Ag, Cd and Sn in stream sediments
ground to pass 325 mesh. Detection limits were from 5 to 20 ppm and varia-
tions in sample mass absorption, due to varying Fe contents, were corrected
using the intensity of the Compton scattered tube peak. Sixty samples per
day could be analyzed. Other examples of the use of energy-dispersive X-ray
fluorescence in analysis of exploration samples are given by Von Alfthan et
al. (1980), Clayton and Packer (1980) and Kramar and Puchelt (1980). The
instrument described by Von Alfthan et al. uses gas proportional, rather than
semiconductor detectors, and is portable.
Chapter 9

ELECTROCHEMICAL METHODS

INTRODUCTION

Polarographic analysis was developed almost fifty years ago and applied in
the early days of exploration geochemistry to determination of Bi, Cd, Cu,
Pb and Zn (Hawkes and Webb, 1962). However, these methods were soon
superceded by faster, simpler colorimetric methods so that, with the excep-
tion of pH measurement and several non-routine determinations, they fell
into disfavour. Despite considerable improvements in polarography and
related techniques such as anodic stripping voltametry, which have found
important applications in water chemistry (Whitney and Risby, 1975), this
generally remains true. In contrast, a relatively recent development — the
specific ion electrode — has greatly simplified estimation of the halogens,
especially F, and thereby promoted investigation of their geochemical behav-
iour and application t o exploration.
Non-routine electrochemical determinations include conductivity, dis-
solved oxygen and Eh. Equipment and methods are described in most texts
on analysis of natural waters (e.g. Golterman, 1970) and, since they are only
required occasionally by the exploration geochemist, will not be discussed
further except to mention the use of conductivity measurements to estimate
concentrations of F", Cl", NOJ and SOl~ in natural waters after their separa-
tion by ion chromatography (Smee et al., 1978). Bölviken et al. (1973) have
described an instrument for in-situ measurement of pH and Eh in diamond
drill holes. Although perhaps n o t strictly electrochemical, an ingenious
method for determination of gaseous Hg by measuring changes in the
resistance caused by adsorption of Hg on a thin gold film (McNerney et al.,
1972) also deserves mention here. The instrument is portable and, with an
absolute sensitivity better than 0.05 ng Hg, has been used to determine the
Hg content of soil gases for mineral exploration purposes (McNerney and
Buseck, 1973). The remainder of this chapter is devoted to pH and specific
ion electrodes.
200

THEORY

A pH electrode is shown in Fig. 9-1: it consists of a thin-walled bulb of


pH-sensitive glass sealed to the end of a glass tube. The bulb holds dilute
hydrochloric acid or a buffered chloride solution into which dips a silver/
silver chloride electrode connected to a sheathed cable. When the electrode is
placed in a solution in which hydrogen ion activity differs from that inside
the electrode, protons accumulate at different rates on either side of the
glass membrane and a potential develops across it. In specific ion electrodes
the proton sensitive membrane is replaced by a solid state or liquid ion
exchange membrane sensitive to the ion to be determined.
The sensing electrode forms one side (a half-cell) of the measuring system,
the other side being a reference electrode providing a stable reference poten-
tial. This generally consists of a silver/silver chloride or calomel (Hg/Hg2Cl2)
electrode surrounded by a saturated solution of potassium chloride. Elec-
trical continuity between the sample solution and the filling solution of the
reference electrode is maintained by diffusion and leakage through a salt-
bridge — either a ground sleeve joint, or a wick or porous plug — forming a
liquid-liquid junction. Different diffusion rates for negative and positive ions
across the liquid junction can produce a junction potential of a few milli-
volts. The complete electrode system, with a calomel electrode, can be repre-

Γ shielded coaxial cable

fill hole

silver/silver chloride wire


filling solution

Pt wire

Mercury -filling solution

Mercurous chloride
1
Wg
(calomel)

U—pH sensitive glass


porous plug

REFERENCE ELECTRODE pH ELECTRODE


Fig. 9-1. A pH electrode and silver/silver chloride reference electrode.
201

sented as:

Ag; AgCl, 0.1 M HC1II glass II sample solution I saturated KC1, Hg2Cl2; Hg
sensing electrode reference electrode

where solutes in the same solution are separated by commas, metal-solution


boundaries by semicolons, liquid-liquid junctions by I and the sensing mem-
brane by II.
Providing it remains stable the absolute potential of the reference elec-
trode is not important because the measuring system is calibrated using stan-
dard solutions of known composition. However, for accurate work, partic-
ularly with specific ion electrodes, the junction potential must be kept as
low as possible and remain constant despite changes in sample solution com-
position that influence diffusion rates. This can be achieved with specially
formulated filling solutions, available from the manufacturers of electrodes,
or by using a double junction reference electrode having two chambers with
an internal liquid-liquid junction between them. The outer chamber is filled
with a solution giving a low junction potential whereas the inner chamber
contains a solution buffered to provide a constant level of the ion sensed
by the reference element. A double junction electrode is also required if the
filling solution of a single junction electrode contains the ion to be measured
or an ion that would interfere with the determination. Leakage from a nor-
mal reference electrode would, for example, contaminate solutions with
chloride.
Ideally the response of the electrode system to the ion to be measured is
described by the Nernst equation:
„ „ 2.3ET A
E=Ex+~^r-\ogA

where E = total potential measured; Ex = potential of the reference elec-


trode; R = the gas constant (8.316 J/deg.-mole); T = temperature, in °K; F =
the value of the Faraday (9.649 X 10 4 J/V); n = the charge on the ion
including its sign; A = the activity of the ion. The Nernst factor 2.3RT/nF
varies with temperature: at 25°C for univalent ions it is 59.16 mV and for
divalent ions 29.58 mV. There is thus a linear relationship between log ion
activity and the potential measured, a tenfold change in activity of a uni-
valent ion causing a 59.16-mV change in the measured potential (Fig.9-2). It
also follows from the logarithmic response that small errors in measurement
of potentials can result in relatively large errors in estimates of ion activity,
i.e. a 1-mV error will give relative activity errors of 4 and 8% for univalent
and divalent ions, respectively.
It will have been noted that throughout the description of electrode
202

>
E

MOO

O
Q_

4*
Q
O
cc /
/
1-300-
/
—i-
iTi 10'' 10" 10 10"' 10" J
0.0002 0.002 0.02 0.2 20

FLUORIDE CONCENTRATION
Fig. 9-2. Response of the fluoride electrode to fluoride activity. At activities above
10~ 5 M response is Nernstian: at lower concentrations the slope flattens due to con-
tamination of the reagent solution with F~ and release of F " from the electrode's
lanthanum fluoride sensing crystal.

0.001 0.005 0.05

IONIC STRENGTH, moles/litre

Fig. 9-3. Relationship between the activity coefficient (7) of an ion and the ionic strength
of a solution. Calculated from the Debey-Huckel expression:
-log Ίι = Azfy/T/il + äßVT)
where A and B are constants determined by the solvent (in this case water at 15°C), a,· is
the effective diameter of the ion, zt is its charge and / is the ionic strength of the solution.
Values used were: A = 0.5;J3 = 0.3262 X lO" 8 ;^· = 2;^· X 10 8 = 6.
203

response ion activity (A) rather than concentration (C) has been referred to.
In very dilute solutions the distinction is negligible. However, the two are
related by: A = jC where 7, the activity coefficient, decreases as the total
ionic strength of the solution increases (Fig. 9-3). Variations in ionic strength
therefore affect the results when the electrode is used to estimate concentra-
tion. In practice this problem is often overcome by preparing both sample
solutions and standards in a total ionic strength adjustment buffer (TISAB)
to eliminate such variations. For a discussion of the significance of the activ-
ity coefficient and its estimation the reader should consult Garrels and
Christ (1965).
Other factors requiring consideration in electrode systems are sensitivity,
specificity (i.e. freedom from interferences) and their speed of response.
These are considered with respect to individual electrodes in the following
pages.

DETERMINATION OF pH

The fundamental definition of pH is pH = —log An where A H is the activ-


ity of the hydrogen ion. Unfortunately this simple definition hides a host of
theoretical and practical problems and it is not, in fact, possible to measure
in any absolute sense the activity of the hydrogen ion. In practice an opera-
tional definition is used to provide a convenient working scale of acidity and
alkalinity by reference to standards that have been assigned agreed pH
values. As we shall see, even with this pragmatic definition numerous
practical problems exist in pH measurement in natural systems. Conse-
quently, it is fortunate that the exploration geochemist is usually content
with relative variations and trends in pH, as a guide to interpreting geochemi-
cal dispersion patterns, rather than absolute values.
In natural waters pH is conveniently estimated in the field using colori-
metric pH indicator solutions or papers. Similar methods can be used for
soils but under laboratory conditions potentiometric measurements, using a
pH electrode and a calomel or Ag/AgCl reference electrode, are preferred. A
saturated potassium chloride solution is usually employed as the reference
electrode filling solution, this avoids changes in the reference potential due
to changes in electrolyte strength caused by evaporation. However, an accu-
mulation of potassium chloride crystals at the salt bridge can cause unsteady
junction potentials and must therefore be avoided.
Soil pH is very sensitive to oxidation-reduction reactions (including micro-
bial activity), which affect sulphate-sulphide and nitrogen-nitrate transfor-
mations, and to the balance of carbon dioxide between soil gases and the
atmosphere. For example, drying of a waterlogged, anaerobic soil can cause
oxidation of sulphide to sulphate with a resulting drop in pH from around 7
to as low as 4. Consequently, apart from removal of stones, soils for pH
204

determination are best kept under field conditions or only gently air-dried.
Before pH can be determined dry soils must be moistened by addition of
water or a salt solution. This introduces problems in so far as soil colloids
interact with the solutions to give different pH values with different solution
to soil ratios or when salt solutions, rather than water, are used. If water is
added pH increases as the water to soil ratio increases (Table 9-1). Usually
some definite water to soil ratio between 1 : 1 and 10 : 1 is used or, in a
more time-consuming procedure, the soil is brought to its moisture satura-
tion point by careful addition of water until the surface of the wet soil
glistens (Jackson, 1958; Peech, 1965). Addition of a salt solution, for exam-
ple 0.01 M calcium chloride, has the advantage of approximating soil-water
composition for many soils. This reduces ion exchange effects and the
system is much less suceptible to changes of solution to soil ratios than when
water alone is used. Junction potentials are also reduced. A disadvantage is
that it is not possible to continue with the same solution and determine its
conductivity.
Occasionally different pH values are obtained depending on whether the
reference electrode (but not the pH electrode) is placed in the soil suspen-
sion or in the clear supernatant liquid. This phenomenon, known as the sus-
pension effect, can be avoided by consistently keeping the reference elec-
trode at the same depth in the supernatant solution.
The variations obtained in measuring the pH of a single soil (Table 9-1)
illustrates the practical difficulties in measuring any theoretically "correct"
pH. For exploration purposes a simple, rapid procedure using a definite
water to soil ratio and measuring pH to ±0.2 units will generally suffice. To
avoid breakage of electrodes or changes in their response due to scratches,
robust electrodes protected by a plastic sleeve are most useful for mea-
surements in soil sluries.

TABLE 9-1
Effect of soil to water ratio and presence of salts on estimation of pH

Treatment pH
(soil : water)
soil A soil B

1 : 0.5 - 8.1
1:1 4.8 8.1
1:2 4.9 8.2
1:5 5.1 8.5
1 : 10 5.0 8.6
1 : 2 0.01 M CaCl2 4.3 7.3
1 : 25 1.0MKC1 3.9 8.3
205

SPECIFIC ION ELECTRODES

Specific ion electrodes are available for at least sixteen ionic species and
many more can be determined using electrodes as endpoint detectors in titra-
tions. However, only a few of the available electrodes (Table 9-II) are likely
to be of interest in the exploration laboratory, and of these only the fluoride
electrode has already achieved widespread usage.
Measurements with electrodes can be made either directly or by methods
of addition. Direct measurements, which are the simplest and most rapid,
involve construction of a calibration curve on semi-log paper using standard
solutions (Fig. 9-2). Calibration should be checked regularly and a buffer
solution is usually added to both unknowns and standards to swamp out any
differences in ionic strength, adjust the pH to the optimum range and con-
trol interferences. Many specific ion meters are provided with logarithmic
concentration scales to simplify direct readout of results in concentration
units.
Methods of addition, in which a known amount of the analyte is added to
the unknown, avoid the need for frequent re-calibration of the electrode and
overcome the problems of close matching of ionic strengths and composi-
tions of standards and unknowns. However, the initial concentration of the
analyte must be known approximately so that the addition can be arranged
to give a two- to five-fold increase. Also the standard addition must not

TABLE 9-11
Characteristics of some specific ion electrodes (based on Orion Research literature)

Electrode Concentration range Interferences


(ppm) (M)

Chloride (Cl") 1.8-35,500 > 3 X 10" 6 Br;5 X 10" 1 0 I;


2X10"10CN;>3X10~7S
Cupric (Cu 2+ ) 6.4 X 10" 4 to saturated > 1 0 " 7 S 2 " , A g \ Hg 2+ ;high
levels of Cl", Br , Fe 3 + , Cd 2+
Fluoride (F") 0.02 to saturated >10~ 4 OH"
Fluoroborate (BFJ) 0.26 to saturated > 2 X 10" 2 NO3;0.2Br",
OAc", HCO3, F~> cl~> 0 H " >
SO 2 " at 10" 3 MBF4
Iodide (I") 5X 10" 3 to 127,000 >10" 7
s
(can be used to measure Hg)
Lead (Pb 2+ ) 0.2-20,700 >10" 7 A g \ Hg 2 + ,Cu 2 + ;
(can be used to measure SO^") high levels Cd 2+ , Fe 3 +
Silver/sulphide S 2 ": 0.003-32,100 >10" 7 Hg 2+
(Ag+/S2") Ag + : 0.01-107,900
206

change either the ionic strength (or else ion activities and junction potentials
would also change) or the proportion, if any, of complexed analyte ions. The
first requirement is met by addition of a small volume of a relatively con-
centrated solution of the analyte to a much larger, accurately known volume
of the unknown. Providing strong complexing agents are present in large
excess this also causes no significant change in the proportion of complexed
ions.
Calculation of results is straightforward. The original potential (Εχ) of the
system is given by:

Ε^Εχ+S log C0

where C 0 = concentration of the analyte in the unknown, S = the Nernst


factor (or an experimental value), and Ex = the reference potential corrected
for the activity coefficient of the analyte ion and for the fraction, if any, of
the analyte complexed. If a known volume (Vs) of a standard, concentration
C s , is added to a measured volume (V 0 ) of the sample the new potential (E2)
becomes:

iiC0V0 CsSVsS \
E22=EX+Sx log
8 —-5—^- + -
\V0+VS V0+Vj
However, if we make Vs negligible compared to V0 (V0/Vs> 100 : 1), by
adding a small volume of a relatively concentrated standard, the expression
simplifies t o :

E2=Ex+Slog (co + ~ )

The change in concentration AC = CSVS/VQ, so that the resulting change of


potential (E2—Ex = AE) is given by:

AE = S log(C 0 + AC)

or:

A E = S l o g ( l + AC/C 0 )

Then if we define the antilogarithm of (AE/S) as Z :

The concentration (C 0 ) in the unknown can, therefore, be calculated from a


207

single addition or, more reliably, by making several standard additions and
plotting (Z — l) versus AC to obtain an average slope. Incremental additions
to the reagent blank provide good estimates of the blank that can then be
used to obtain a linear, blank-corrected, calibration curve at lower concentra-
tions than would otherwise be possible (Smith and Manaham, 1973;
McQuaker and Gurney, 1977).
If we do not arrange for Vs to be negligible compared to V0, the calcula-
tions become relatively cumbersome and it is much simpler to plot results on
Gran's paper — a special volume corrected semi-antilog graph paper named
after one of the principal exponents of linear titrations (Orion Research,
1970). An example of a Gran's plot is shown in Fig. 9-4: the paper is availa-
ble from Orion Research, 380 Putnam Avenue, Cambridge, MA 02139 U.S.A.
Specific ion electrodes can also be used as endpoint detectors for indirect
determination of ions by titration. Many such determinations are feasible;
two of potential interest here are analysis of natural waters or leachates for
Cl" or SO*". Chloride can be determined directly with a Cl~-sensing elec-
trode, but at levels below 5 Mg/ml is best determined indirectly using a Ag/
AgS electrode to follow its titration with silver nitrate (Haynes and Clark,
1972; Orion Research, 1970). Sulphate is determined by titration of a
methanolic solution with lead perchlorate using a lead (Pb/PbS) electrode:

mLof 2X10" 3 M A g N 0 3 ADDED


Fig. 9-4. A low level titration of Cl~ on 10% volume-corrected Gran's plot paper using the
Ag/S electrode. Titrations are shown for the reagent blank and a sample containing 5 X
10~ 5 M Cl~. At these concentrations the solubility product of silver chloride is hardly
exceeded at the equivalence point so that no end-point break would be obtained with a
normal titration. The distortion of the graph paper corrects for up to a 10% volume
change during the titration. Note the vertical antilogarithmic axis. (Courtesy of Orion
Research Incorporated.)
208

lanthanum is added to prevent phosphate interference (Orion Research,


1975; Goertzen and Oster, 1972; Hulanicki et al., 1976; Scheide and Durst,
1977). Results of the titrations can be plotted conventionally but at the low
ion concentrations found in many freshwaters the endpoint becomes indeter-
minate. Linear titration on Gran's paper gives a much clearer endpoint and
extends the lower concentration of the titration (Fig. 9-4).

Fluoride

The fluoride electrode has enabled rapid, direct methods to be developed


for determination of F, freeing the analyst from the very laborious distilla-
tion procedures previously required. This has aroused greater interest in the
geochemistry of F and its possible application to prospecting.
The electrode, which has a solid state lanthanum fluoride-sensing mem-
brane, is able to measure as little as 10" 6 M F" (0.019 ppm) and is remarka-
bly specific. However, the maximum permissable level of hydroxyl ion activ-
ity [OH - ] is one tenth the fluoride activity [F"] and the pH should there-
fore be lower than 7. At pH values below 4.5 H+ interferes, due to formation
of HF and HFJ, and the optimum pH range for determination of fluoride is
therefore 5—6.5. Only free F" ions are sensed and ions forming complexes
with fluoride, for example Al3+ and Fe 3+ , should be absent.
In direct analysis of natural waters pH and ionic strength are controlled
and interferences suppressed by addition of a total ionic strength adjustment
buffer. The buffer used by Frant and Ross (1968) contained 57 ml glacial
acetic acid, 58 g sodium chloride and 0.3 g sodium citrate made up to 1 litre
with water after adjusting pH to 5.0—5.5 with sodium hydroxide. Sodium
citrate is present to preferentially complex Al 3+ . In a later TISAB formula-
tion (Orion Research, 1976), citrate is replaced by 4 g of CDTA (cyclo-
hexylene dinitrilo tetracetic acid or 1,2-diaminocyclohexane Ν,Ν,Ν',Ν'-
tetracetic acid). Both versions of the buffer are mixed 1 : 1 with sample and
standard solutions.
Fluoride in geochemical samples is often determined after an alkali fusion,
or extraction with aluminium chloride or beryllium nitrate solutions
(pp. 69 and 65). Acid leachates of the alkali fusions contain far greater
quantities of potential interferents than are found in natural waters and the
Al3+ and Be3+ are, of course, strong complexing agents for F". A buffer with
a greater complexing capacity for these potential interferents is therefore
needed. Relatively strong citrate buffers (Edmond, 1969; Ingram, 1970;
Plüger and Friedrich, 1973; Hopkins, 1977), sometimes with CDTA added
(Bodkin, 1977) are used. It is also possible to reduce the concentrations of
interferents brought into solution by an alkali fusions if the fused mass is
slowly dissolved, by addition of acid, until a pH of 8 is reached. At this point
insoluble aluminum and iron hydroxides are filtered off before addition of
the buffer and determination of F" (McQuaker and Gurney, 1977).
209

Final solutions of most geochemical samples will probably contain 0.2—


5.0 Mg/ml F" assuming a dilution factor of 100. In this range electrode
behaviour is approximately Nernstian (Fig. 9-2), with a response time of
several minutes, and direct concentration readings should be adequate
although more reliable results can be obtained by methods of addition
(Jagner and Pavlova, 1972; Kesler et al., 1973; Hopkins, 1977). At lower
concentrations (0.02—0.1 μg/ml), Ingram (1970) found that response times
in a sodium citrate/potassium nitrate buffer were up to 30 minutes and
readings taken after only 10 minutes were dependent on F" content of the
preceding sample. This problem was overcome by "standardizing" the elec-
trode for 5 minutes in a 4^g/ml F" solution between readings.
Chloride
Liquid membrane and solid state (Ag2/AgCl) Cl"-sensing electrodes are
able to measure concentrations down to 5 X 10~5 M Cl". The solid state elec-
trode, however, is subject to several interferences (Table 9-II) and can also be
influenced by temperature fluctuations and changes of light intensity
(Sekerka and Lechner, 1973). Both electrodes must, of course, be used in
conjunction with a double junction reference electrode, its outer chamber
containing a potassium nitrate solution, to avoid Cl" contamination of the
sample. Despite its potential interference problems the more robust solid
state electrode is preferable for routine geochemical analysis. Construction
of an inexpensive electrode has been described by Van Loon (1971) and Van
Loon et al. (1973).
Chloride electrodes have been used to determine total Cl" in rocks
(Haynes and Clark, 1972), and water-extractable Cl" in plutonic rocks (Van
Loon et al., 1973) and soils (Selmer-Olsen and Qien, 1973). In their deter-
mination of total Cl", Haynes and Clark (1972) fused 0.25-g samples with
1 g sodium carbonate/potassium nitrate ( 2 : 1 ) in culture tubes or with
sodium carbonate/zinc oxide (2 : 1) in platinum crucibles. Residues were
leached with distilled water and filtered into a beaker, 1.5 ml nitric acid
being added to expel carbon dioxide before bringing the volume to 100 ml.
Cl" concentrations greater than 5 Mg/ml were estimated by standard addi-
tions using 1 ml of 1000 Mg/ml Cl" standard. Below 5 Mg/ml electrode
response deviated from the Nernst slope and Cl" was estimated using a silver-
sensing electrode and titrating solutions with 0.5-ml increments of a silver
nitrate solution. The titration was plotted on Gran's paper (Fig. 9-4). With
respect to interferences (Table 9-II), I" and Br" are unlikely to be present in
significant quantities in most samples and sulphide is destroyed by the oxida-
tive fusion. Perspiration and sea-spray are potential contaminants.
Iodide
Characteristics of the solid state (Ag2S/AgI) iodide electrode are summar-
ized in Table 9-II. Ficklin (1975) used it to estimate I content of soils and
210

rocks, by a standard addition method, after decomposing the sample with a


sodium carbonate/potassium carbonate/magnesium oxide sinter. Interference
from Ag was prevented by its reduction with Zn or Sn in basic solution. Sul-
phides, which would also interfere, are destroyed in the oxidative fusion or
converted to hydrogen sulphide in the strongly acidic final solution. Results
were comparable to those obtained by neutron activation.
Copper
The cupric ion electrode, with an Ag 2 S/AgCu membrane, has a free-ion
concentration limit of 10" 8 M (0.6 Mg/1) but can measure cupric ion activities
down to a much lower level (10~ 1 7 M) provided complexed copper species
are present at concentrations of at least 10" 6 M. The electrode also responds
to S 2 ", Ag + , and Hg 2+ which must be less than 10~ 7 M, and high levels of Cl",
Br", Fe 3 + and Cd 2+ also interfere (Table 9-II). Electrode response has also
been reported to be influenced by room temperature changes (Sekerka and
Lechner, 1973).
Despite its sensitivity, interference problems have prevented geochemists
taking advantage of the cupric ion electrode for on-site field or field labora-
tory determinations of Cu. These problems appear to have been partly over-
come, however, by Smith and Manaham (1973) with their development of a
complexing antioxidant buffer. This is prepared by mixing 100.0 ml of
1.00 N acetic acid, 63.5 ml 1.00 N ultrapure potassium hydroxide, 0.84 g
ultrapure sodium fluoride and 2.0 ml of 1.0 M formaldehyde solution, and
diluting to 1 1 . The buffer, which should n o t contain more than about 1 ppb
Cu (checked by the method of additions), serves several purposes: (1) pH is
regulated to the optimum of about 5 — above this pH basic copper
hydroxides form and in very acidic solutions the electrode functions badly;
(2) the large excess of acetate forms a strong complex with cupric ions, both
keeping them in solution and ensuring that a constant fraction of Cu is
present as Cu 2 + ; (3) the electrode responds to Fe 3 + and precipitation of ferric
hydroxide can coprecipitate Cu; addition of fluoride both complexes Fe 3 +
and retains it in solution; and (4) oxidizing conditions, which cause the elec-
trode to become unstable, are prevented by addition of formaldehyde as a
reducing agent.
Using this buffer and a double junction reference electrode, equilibration
time between samples was about 10 minutes at 30 Mg/1, 60 minutes at 1 μg/l
and less than one minute at 1 mg/1. As little as 9 μg/l could be determined by
the method of additions and plots of (Z — 1) versus AVS as described on
p. 206. An alternative buffering system, giving faster response times with a
chalcocite electrode, is described by Hulanicki et al. (1977).
Schuller et al. (1975) used the cupric ion electrode to determine Cu in
cold hydrochloric acid extracts of soils. Results were reported to compare
favourably with AAS determination: unfortunately no procedural details are
given.
211

Boron

Boron can be determined either by emission spectroscopy or by several


colorimetric methods. An alternative is to convert B to tetrafluorborate
(BFi) and estimate it with the fluorborate electrode.
A procedure for analysis of soils and plants was described by Carlson and
Paul (1969). Acidic sample digests, containing 1—500 μg B, are first neutral-
ized by passage through the ammonia-form of a weak acid resin to improve
retention of B on their subsequent passage through a B specific resin
(Amberlite XE-243). Addition of hydrofluoric acid to the XE-243 column
converts B to BF4 which is then eluted with 0.3 N sodium hydroxide and
passed through a strong acid resin in the Ca-form to precipitate excess F" as
calcium fluoride. A calibration curve is prepared by passing known con-
centrations of B through the columns.
Elution of solutions through the three columns takes about an hour but
the columns are small and it would not be difficult to construct racks for
simultaneous processing of large batches. Results obtained by Carlson and
Paul (1969) compare favourably with results of the more laborious curcumin
and mannitol colorimetric methods.
APPENDIX 2 - PREPARATION OF STANDARDS

Weights of compounds to be dissolved and diluted to 1 litre to give a con-


centration of 1000 mg/1 in solution
Element Compound Weight Comments *
(g)

Al A1K(S0 4 ) 2 ' 12 H 2 0 17.582


Sb K(SbO)C 4 H 4 0 6 · \ H 2 0 2.7426
As As203 1.320 dry at 110 C: dissolve in 5 mv
water + 2 g NaOH
Ba BaCl2 · 2 H 2 0 1.7787
Be BeS0 4 11.6586 in distilled water + 2 ml nitric acid
Bi Bi(N03)3-5H20 2.321
B H3BO3 5.716
Cd Cd 1.000 warm to dissolve in 10 ml
hydrochloric acid
Ca CaC0 3 2.500 dry at 105°C; carefully dissolve by
slow addition of 50% hydro-
chloric acid; boil to expel C 0 2

cr NaCl 1.6482
Cr K2Cr2Ch 2.828
Co CoCl2 · 6 H 2 0 4.037
Cu Cu 1.000 polished Cu wire: dissolve in 10 ml
CuS0 4 • 5 H 2 0 3.929 nitric acid
F" NaF 2.210
Fe Fe 1.000 pure Fe wire: dissolve in 5 ml nitric
acid
Au Au 1.000 dissolve in aqua-regia
Pb Pb(N0 3 ) 2 1.599
Li LiCl 6.109
Mg MgO 1.658 dissolve in 10 ml nitric acid
Mn MnS0 4 2.7486 dry at 180°C; dissolves in water
Hg HgCl2 1.354
Mo Na 2 Mo0 4 · 2 H 2 0 2.522
Ni NiS0 4 ■6 H 2 0 4.477
215

APPENDIX 2 (continued)

Element Compound Weight Comments *


(g)

Nb Nb 2 O s 1.432 repeated evaporation with hydro-


fluoric acid
P KH2P04 4.393 dry at 105 C; dissolves in water
K KC1 1.9068
Se Se 1.000 dissolve in nitric acid boiling to expel
brown fumes
Si Na 2 Si0 3 • 9 H 2 0 4.73
Ag AgN0 3 1.574
Na NaCl 2.5420
Sr SrC0 3 1.685 slowly add 50% hydrochloric acid:
dilute and boil for a few minutes
to expel C 0 2
sof Na2S04 1.497
Sn Sn 1.000 dissolve in hydrochloric acid
U u3o8 1.179 dissolve in nitric acid
V v2os 1.785 ignite at 500 C then dissolve in
slight excess NaOH
W Na 2 W0 4 • 2 H 2 0 1.794
Zn Zn 1.000 dissolve in 10 ml nitric acid

* Unless otherwise indicated the compound dissolves in water. For atomic absorption
spectrophotometry, standard solution are normally prepared in dilute (^10%) hydro-
chloric or nitric acid. Hydrochloric acid should not be used for stock solutions of ele-
ments, e.g. Ag and Pb, with insoluble chlorides.
REFERENCES

Abbey, S., 1970. Analysis of rocks and minerals by atomic absorption spectroscopy, 3.
A lithium-fluorborate scheme for seven major elements. Geol. Surv. Can., Paper, 70-
23: 20 pp.
Abbey, S., 1977. Studies in "standard samples" for use in general analysis of silicate rocks
and minerals, 5. 1977 edition of "usable" values. Geol. Surv. Can., Paper, 77-34:
31pp.
A.E.G. Analysis Committee, 1971. Committee Report. Assoc. Explor. Geochem. Newsl.,
July 1971: 3 - 5 .
Aggett, J. and Aspell, A.C., 1976. The determination of arsenic (III) and total arsenic by
atomic-absorption spectroscopy. Analyst, 100: 341—347.
A.G.R.G., 1962. Determination of sulphate in natural water. Appl. Geochem. Res. Group,
Tech. Comm., No. 27: 2 pp.
Agterdenbos, J. and Vlogtman, J., 1972. Determination of and differentiation between
cassiterite and silicate bound tin in silicate rocks containing traces of tin. Talanta, 19:
1295-1300.
Ahrens, L.H., 1965. Distribution of the Elements in Our Planet. McGraw-Hill, New York,
N.Y., 110 pp.
Ahrens, L.H. and Taylor, S.R., 1961. Spectrochemical Analysis. Addison-Wesley, London,
2nd ed., 454 pp.
Allan, J.E., 1961. The use of organic solvents in atomic absorption spectrophotometry.
Spectrochim. Ada, 17: 467—473.
Allan, R.J. and Hornbrook, E.H.W., 1971. Exploration geochemistry evaluation study in
a region of continuous permafrost, Northwest Territories, Canada. In: R.W. Boyle
(Editor), Geochemical Exploration. Can. Inst. Min. Metall, Spec. Vol., 11: 53—66.
Allcott, G.H. and Lakin, H.W., 1975. The homogeneity of six geochemical exploration
reference samples. In: I.L. Elliott and W.K. Fletcher (Editors), Geochemical Explora-
tion 1974. Else vier, Amsterdam, pp. 659—681.
Allcott, G.H. and Lakin, H.W., 1978. Tabulation of geochemical data furnished by 109
laboratories for six exploration reference samples. U.S. Geol. Surv., Open File Rep.,
78-163: 199 pp.
Alminas, H.V. and Mosier, EX., 1975. Oxalic-acid leaching of rock, soil and stream-sedi-
ment samples as an anomaly-accentuation technique. U.S. Geol. Surv., Open File Rep.,
76-275: 25 pp.
Almond, H., 1953a. Field method for the determination of traces of arsenic in soils. Anal.
Chem., 25: 1 7 6 6 - 1 7 6 7 .
Almond, H., 1953b. Determination of traces of cobalt in soils. Anal. Chem., 25: 166—
167.
Almond, H., 1953c. A field method for the determination of manganese in soils adapted
from methods described in the literature. U.S. Geol. Surv., Open File Rep.
218

Almond, H., 1955. Rapid field and laboratory method for the determination of copper in
soil and rock. U.S. Geol. Surv., Bull, 1036-A: 1—8.
Almond, H., Crowe, H.E. and Thompson, C.E., 1955. Rapid determination of germanium
in coal, soil and rock. U.S. Geol. Surv., Bull., 1036-B: 9—17.
Aly, M.M., Wassef, S.N. and Hathout, M.H., 1977. The colorimetric estimation and the
distribution of thorium in Egyptian beach sands. Chem. Erde, 36: 336—342.
Analytical Methods Committee, 1959. Notes on perchloric acid and its handling in ana-
lytical work. Analyst, 84: 214—216.
Anderman, G. and Kemp, J.W., 1958. Scattered X-rays as internal standards in X-ray
emission spectroscopy. Anal. Chem., 30: 1306—1309.
Anderson, J.U., 1963. An improved pretreatment for mineralogical analysis of samples
containing organic matter. Clays Clay Miner., 10: 380—388.
Anderson, J.U. and O'Connor, G.A., 1972. Production of permanganate ion by sodium
hypochlorite treatment to remove soil organic matter. Soil Sei. Soc. Am., Proc., 36:
973-975.
Aslin, G.E.M., 1976. The determination of arsenic and antimony in geological materials
by flameless atomic absorption spectrometry. J. Geochem. Explor., 6: 321—330.
Avni, R., Harel, A. and Brenner, I.B., 1972. A new approach to the spectrochemical anal-
ysis of silicate rocks and minerals. Appl. Spectrosc., 26: 641—645.
Azzaria, L.M. and Webber, G.R., 1969. Mercury analysis in geochemical exploration. Can.
Inst. Min. Metall, Bull, 62: 5 2 1 - 5 3 0 .

Baker, W.E., 1965. Rapid colorimetric procedures in geochemical prospecting for molyb-
denum. Proc. Australas. Inst. Min. Metall, 214: 125—134.
Ball, T.K., Booth, S.J., Nickless, E.F.P. and Smith, R.T., 1979. Geochemical prospecting
for baryte and celestite using a portable radioisotope fluorescence analyser. J. Geo-
chem. Explor., 1 1 : 277—284.
Band, R.B. and Wilkinson, N.M., 1972. Interferences in the determination of mercury in
mineralized samples by the wet reduction—flameless atomic absorption method. J.
Geochem. Explor., 1: 195—198.
Barakso, J.J., 1967. Geochemical field kit for the determination of trace amounts of
molybdenum. Econ. Geol, 62: 732—736 (Discussions, 63: 85, 572, 695—696).
Barakso, J.J. and Tornocai, C , 1970. A mercury determination method and its use for
exploration in British Columbia. Can. Inst. Min. Metall, Bull, 63: 501—505.
Barnard, W.M. and Fishman, M.J., 1973. Evaluation of the use of the heated graphite
atomizer for the routine determination of trace metals in water. At. Absorpt. NewsL,
12: 1 1 8 - 1 2 4 .
Barringer, A.R., 1966. Interference-free spectrometer for high-sensitivity mercury analy-
ses of soils, rocks and air. Inst. Min. Metall, Trans., Sect. B, 75: 120—124.
Belcher, R., Bogdanski, S.L., Henden, E. and Townshend, A., 1975. Elimination of inter-
ferences in the determination of arsenic and antimony by hydride generation using
molecular emission cavity analysis (MECA). Analyst, 100: 522—523.
Berenice, G., Da Silva, C.T.C.B. and Mello, V.N.A., 1975. A method of interpreting
results obtained by a six-step semiquantitative spectrographic procedure. App. Spec-
trosc, 29: 269—271.
Bergseth, H., 1975. Determination of zinc in soils by X-ray fluorescence spectrometry
involving a modified background ratio method. Analyst, 100: 96—98.
Berman, S.S. and McLaren, J.W., 1978a. Establishment of compromise conditions for
multielement analysis by inductively coupled plasma emission spectrometry: a prelimi-
nary report. Appl Spectrosc, 32: 372—377.
Berman, S.S. and McLaren, J.W., 1978b. Application of the inductively coupled plasma
to geochemical analysis. Paper presented at Geoanalysis 78, Ottawa, Ont.
219

Billings, G.K., 1965. Light scattering in trace-element analysis by atomic absorption. At.
Absorpt. NewsL, 4 : 357—361.
Blasius, M.B., Kerkhoff, S.J., Wright, R.S. and Cothern, C.R., 1972. Use of X-ray fluores-
cence to determine trace metals in water resources. Water Resour. Bull, 8: 704—714.
Bloom, H., 1955. A field method for the determination of ammonium citrate-soluble
heavy metals in soils and alluvium. Econ. GeoL, 55: 533—541.
Bloom, H., 1962. Field methods for the determination of nickel using dimethylglyoxime.
Econ. GeoL, 57: 5 9 5 - 6 0 4 .
Bloom, H., 1966. A field method for the determination of silver in soils and rocks using
dithizone. Econ. GeoL, 6 1 : 189—197.
Bodkin, J.B., 1977. Determination of fluorine in silicates by use of an ion-selective elec-
trode following fusion with lithium metaborate. Analyst, 102: 409—413.
Bölviken, B., Logn, O. and Uddu, O., 1973. Instrument for in situ measurements of pH,
Eh and self-potential in diamond drill holes. In: M.J. Jones (Editor), Geochemical
Exploration 1972. Institution of Mining and Metallurgy, London, pp. 417—422.
Bond, A.M., 1970. Use of ammonium fluoride in determination of zirconium and other
elements by atomic absorption spectrometry in the nitrous oxide-acetylene flame.
Anal. Chem., 42: 932—935.
Boumans, P.W.J.M., 1966. Theory of Spectrochemical Excitation. Hilger and Watts,
London, 383 pp.
Boumans, P.W.J.M., 1979. Inductively coupled plasma-atomic emission spectroscopy: its
present and future position in analytical chemistry. ICP Inf. NewsL, 5: 181—209.
Bowden, P., 1964. A geochemical field method for determining tungsten in soils and
stream sediments. Analyst, 89: 771—774.
Bowie, S.H.U., 1968. Portable X-ray fluorescence analysers in the mining industry. Min.
Mag., 118: 2 3 0 - 2 3 9 .
Bowie, S.H.U., Darnley, A.G. and Rhodes, J.R., 1965. Portable radioisotope X-ray
fluorescence analyser. Inst. Min. Metall., Trans., Sect. B, 74: 361—379.
Boyle, R.W., 1976. Report of retiring president R.W. Boyle to Annual Meeting of the
Association of Exploration Geochemists, Fredericton, N.B., April 23, 1976. J. Geo-
chem. Explor., 6: 389—395.
Brabec, D., 1971. Aqua regia extractable vs. total copper and zinc content of granitic
rocks. Trans. Soc. Min. Eng. AIME, 250: 94—97.
Bradshaw, P.M.D., Thomson, I., 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: 209—225.
Bratzel, M.P., Chakrabarti, C.L., Sturgeon, R.E., Mclntyre, M.W. and Agemian, H., 1972.
Determination of gold and silver at parts-per-billion or lower levels in geological and
metallurgical samples by atomic absorption spectrometry with a carbon rod atomizer.
Anal. Chem., 44: 3 7 2 - 3 7 4 .
Brobst, D.A. and Ward, F.N., 1965. A turbidimetric test for barium and its geologic
application in Arkansas. Econ. GeoL, 60: 1020—1040.
Brown, E., Skougstad, M.W. and Fishman, M.J., 1970. Methods for Collection and Anal-
ysis of Water Samples for Dissolved Minerals and Gases. Techniques of Water-Resources
Investigations of the U.S. Geological Survey, Book 5. U.S. Geological Survey, Washing-
ton, D.C., Chapter A l , 160 pp.
Brown, G.C., Hughes, D.J. and Esson, J., 1973. New X.R.F. data retrieval techniques and
the application to U.S.G.S. standard rocks. Chem. GeoL, 1 1 : 223—229.
Brundin, N.H. and Bergström, J., 1977. Regional prospecting for ores based on heavy
minerals in glacial till. J. Geochem. Explor., 7: 1—19.
Brundin, N.H. and Nairis, B., 1972. Alternative sample types in regional geochemical
prospecting. J. Geochem. Explor., 1: 7—46.
220

Cameron, E.M., 1972. Three geochemical standards of sulphide-bearing ultramafic rock:


UM-1., UM-2., UM-4. Geol. Suw. Can., Paper, 71-35: 10 pp.
Cameron, E.M., 1977. Geochemical dispersion in mineralized soils of a permafrost
environment. J. Geochem. Explor., 7: 301—326.
Cameron, E.M., Siddeley, G. and Durham, C.C., 1971. Distribution of ore elements in
rocks for evaluating ore potential: nickel, copper, cobalt and sulphur in ultramafic
rocks of the Canadian Shield. In: R.W. Boyle (Editor), Geochemical Exploration. Can.
Inst. Min. Metall, Spec. Vol., 11: 3 1 3 - 3 1 7 .
Campbell, W.C. and Ottaway, J.M., 1974. Determination of lead in carbonate rocks by
atomic absorption spectrometry with carbon furnace atomization. Inst. Min. Metall,
Trans., Sect. B, 83: 68—69.
Campbell, W.C. and Ottaway, J.M., 1975. Determination of lead in carbonate rocks by
carbon-furnace atomic-absorption spectrometry after dissolution in nitric acid. Talanta,
22: 7 2 9 - 7 3 2 .
Campbell, W.J., Spano, E.F. and Green, T.E., 1966. Micro and trace analysis by a com-
bination of ion exchange resin-loaded papers and X-ray spectrometry. Anal Chem.,
38: 9 8 7 - 9 9 6 .
Campbell, W.L., 1980. Eliminating manganese dioxide interference in the atomic absorp-
tion determination of gold. Abstr., 8th Int. Geochem. Explor. Symp., Hannover, 1980,
p. 130.
Canney, F.C. and Nowlan, G.A., 1964. Solvent effect of hydroxylamine hydrochloride in
the citrate-soluble heavy metals test. Econ. Geol, 59: 721—724.
Canney, F.C. and Post, E.V., 1977. U.S. geochemical activity. Assoc. Explor. Geochem.
Newsl, 22: 1 2 - 1 6 .
Canney, F.C, Myers, A.T. and Ward, F.N., 1957. A truck-mounted spectrographic labora-
tory for use in geochemical exploration. Econ. Geol, 52: 289—306.
Carlson, E.H. and Manus, R.W., 1979. Application of campsite technique for the analysis
of total fluoride to the exploration for fluorspar. In: J.R. Watterson and P.K.
Theobald (Editors), Geochemical Exploration 1978. Association of Exploration Geo-
chemists, Rexdale, Ont., pp. 87—92.
Carlson, R.M. and Paul, J.L., 1969. Potentiometric determination of boron in agricultural
r
samples. Soil Sei., 108: 266—272.
Carpenter, R.H. and Hayes, W.B., 1979. Fe-Mn coatings in routine exploration geochemi-
cal surveys. In: J.R. Watterson and P.K. Theobald (Editors), Geochemical Exploration
1978. Association of Exploration Geochemists, Rexdale, Ont., pp. 277—282.
Carter, D., Regan, J.G.T. and Warren, J., 1975. Atomic-absorption determination of
strontium in silicate rocks: a study of major element interferences in the nitrous oxide-
acetylene flame. Analyst, 100: 721—725.
Centanni, F.A., Ross, A.M. and De Sesa, M.A., 1956. Fluorimetric determination of
uranium. Anal Chem., 28: 1651—1657.
Chakrabarti, C.L. and Singhai, S.P., 1969. Effect of complexing agents and organic
solvents on the sensitivity of atomic-absorption spectroscopic technique. Spectrochim.
Ada., 24B: 663—677.
Champion, K.P., Taylor, J.C. and Whittem, R.N., 1966. Rapid X-ray fluorescence deter-
mination of traces of strontium in samples of biological and geological origin. Anal.
Chem., 38: 1 0 9 - 1 1 2 .
Chao, T.T., 1972. Selective dissolution of manganese oxides from soils and sediments
with acidified hydroxylamine. Soil Sei. Soc. Am., Proc, 36: 764—768.
Chao, T.T. and Sanzolone, R.F., 1977. Chemical dissolution of sulfide minerals. U.S.
Geol Surv., J. Res., 5: 409—412.
Chao, T.T. and Theobald, P.K., 1976. The significance of secondary iron and manganese
oxides in geochemical exploration. Econ. Geol, 71: 1560—1569.
221

Chao, T.T., Ball, J.W. and Nakagawa, H.M., 1971. Determination of silver in soils, sedi-
ments and rocks by organic chelate extraction and atomic absorption spectrophotom-
etry. Anal. Chim. Acta, 54: 77—81.
Chao, T.T., Jenne, E.A. and Heppting, L.M., 1968. Prevention of adsorption of trace
amounts of gold by containers. U.S. Geol. Suru., Prof. Paper, 600-D: 16—19.
Chao, T.T., Sanzolone, R.F. and Hubert, A.E., 1978. Flame and flameless atomic-absorp-
tion determination of tellurium in geological materials. Anal. Chim. Acta, 96: 251—
257.
Chork, C.Y., 1977. Seasonal, sampling and analytical variations in stream sediment
surveys. J. Geochem. Explor., 7: 31—48.
Chaudury, A.N. and Böse, B.B., 1971. Role of "humus matter" in the formation of geo-
chemical anomalies. In: R.W. Boyle (Editor), Geochemical Exploration. Can. Inst.
Min. Metall, Spec. Vol., 11: 410—415.
Chu, R.C., Barron, G.P. and Baumgarner, A.W., 1972. Arsenic determination at sub-
microgram levels by arsine evolution and flameless atomic absorption spectrophoto-
metric technique. Anal. Chem., 44: 1476—147$.
Cioni, R., Mazzucotelli, A. and Ottonello, G., 1976a. Interference effects in the deter-
mination of barium in silicates by flame atomic-absorption spectrophotometry.
Analyst, 101: 9 5 6 - 9 6 0 .
Cioni, R., Mazzucotelli, A. and Ottonello, G., 1976b. Matrix effects in the flameless
atomic absorption determination of trace amounts of barium in silicates. Anal. Chim.
Acta, 82: 4 1 5 - 4 2 0 .
Clark, L.J. and Axley, J.H., 1955. Molybdenum determination in soils and rocks with
dithiol. Anal. Chem., 27: 2000—2003.
Clark, N.H. and Pyke, J.G., 1972. The determination of uranium in exploration samples
by X-ray emission spectrometry. Anal. Chim. Acta, 58: 234—237.
Clayton, C G . and Packer, T.W., 1980. Some applications of energy dispersive X-ray
fluorescence analysis in mineral exploration, mining and process control. In: C.S.
Barrett, D.E. Leyden, J.B. Newkirk, P.K. Predecti and C O . Ruud (Editors), Advances
in X-ray Analysis, Vol. 23. Plenum, New York, N.Y., pp. 1—13.
Clifton, H.E., Hunter, R.E., Swanson, F.J. and Phillips, R.L., 1969. Sample size and
meaningful gold analysis. U.S. Geol Suru., Prof. Paper, 625-C: 17 pp.
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., 1 3 : 8 pp.
Cogger, N., 1976. An extraction-spectrophotometric method for the determination of W
in geological materials. Anal. Chim. Acta, 84: 143—148.
Corbett, J.A. and Godbeer, W.C., 1977. The determination of tellurium in weathered
outcrop, mineralized and barren rock. Anal. Chim. Acta, 9 1 : 211—219.
Craven, C.A.U., 1954. Statistical estimation of the accuracy of assaying. Inst. Min. Metall,
Trans., Sect. B, 63: 5 5 1 - 5 6 3 .
Crenshaw, G.L. and Ward, F.N., 1975. Determination of fluorine in soils and rocks by
known-increment addition and selective-ion electrode detection. In: F.N. Ward (Editor),
New and Refined Methods of Trace Analysis Useful in Geochemical Exploration. U.S.
Geol. Suru., Bull, 1408: 7 7 - 8 4 .
Cruft, E.F. and Giles, D.L., 1967. Direct reading emission spectrometry as a geochemical
tool. Econ. Geol, 62: 406—411.
Cruz, R.B. and Van Loon, J . C , 1974. A critical study of the application of graphite-fur-
nace non-flame atomic absorption spectrometry to the determination of trace base
metals in complex heavy-matrix sample solutions. Anal. Chim. Acta, 72: 231—243.
Czamanske, G.K. and Ingamells, C O . , 1970. Selective chemical dissolution of sulphide
minerals: a method of mineral separation. Am. Mineral, 55: 2131—2134.
222

Dahlquist, R.L. and Knoll, J.L., 1978. Inductively coupled plasma-emission spectrometry:
analysis of biological materials and soils for major, trace and ultra-trace elements.
Appl. Spectrosc, 32: 1—29.
Dalton, E.F. and Malanoski, A.J., 1971. Note on the determination of arsenic by atomic
absorption by arsine into an argon-hydrogen entrained air flame. At. Absorpt Newsl,
10: 9 2 - 9 3 .
Danielsson, A., 1967. Spectrochemical analysis for geochemical prospecting. In: A.H.
Gillieson (Editor), XIII Colloquium Spectroscopium Internationale. Adam Hilger,
London, pp. 311—323.
Danielsson, A. and Sundkvist, G., 1959. The tape machine, II. Applications using differ-
ent kinds of isoformations. Spectrochim. Ada, 15: 126—133.
Danielsson, A., Lundgren, F. and Sundkvist, G., 1959. The tape machine, I. A new tool
for spectrochemical analysis. Spectrochim. Acta, 15: 122—125.
Davenport, P.H. and Nichol, I., 1973. Bedrock geochemistry as a guide to areas of base-
metal potential in volcano-sedimentary belts of the Canadian Shield. In: M.J.. Jones
(Editor), Geochemical Exploration 1972. Institution of Mining and Metallurgy, London,
pp. 45—57.
Davies, B.E., 1974. Loss-on-ignition as an estimate of soil organic matter. Soil Sei. Soc.
Am.,Proc, 38: 150—151.
Davies, R.L, Cheshire, M.V. and Graham-Bryce, I.J., 1969. Retention of low levels of
copper by humic acid. J. Soil Sei., 20: 65—71.
Davis, C.E.S., 1972. Analytical methods used in the study of an ore intersection from
Lunnon Shoot, Kambalda. Econ. Geol., 67: 1091—1092.
Deb, B.C., 1950. The estimation of free iron oxides in soils and clays and their removal.
J. Soil Sei., 1: 2 1 2 - 2 2 0 .
Delavault, R., 1977. Simplified field test for copper. J. Geochem. Explor., 8: 537—540.
Delavault, R.E. and Marshall, D.B., 1973. Detection of molybdenum in rocks using
Tennant's bead-forming buffer. Can. J. Spectrosc, 18: 10—12.
De Pablo Galan, L., 1975. Direct-reading emission spectroscopy analysis of geochemical
samples. In: I.L. Elliott and W.K. Fletcher (Editors), Geochemical Exploration 1974.
Elsevier, Amsterdam, pp. 707—720.
Dijkstra, S., Van den Hul, H.J. and Bill, E., 1979. Experiments on the usefulness of some
selected chemical quantities in geochemical exploration in a former mining district. In:
J.R. Watterson and P.K. Theobald (Editors), Geochemical Exploration 1978. Associa-
tion of Exploration Geochemists, Rexdale, Ont., pp. 283—288.
Dolezal, J., Povondra, P. and Sulcek, Z., 1968. Decomposition Techniques in Inorganic
Analysis (English Edition). Iliffe Books, London, 224 pp.
Dorrzapf, A.F., 1973. Spectrochemical analysis — argon-oxygen DC arc method for
silicate rocks. U.S. Geol. Surv., J. Res., 1: 559—562.
Dyck, W., 1971. The adsorption and coprecipitation of silver on hydrous oxide of iron
and manganese. Geol. Surv. Can., Paper, 70-64: 23 pp.

Ediger, R.D., 1973. A review of water analysis by atomic absorption. At. Absorpt. Newsl,
12: 1 5 1 - 1 5 7 .
Ediger, R.D., 1975. Atomic absorption analysis with the graphite furnace using matrix
modification. At. Absorpt. Newsl, 14: 127—130.
Ediger, R.D. and Wilson, D.L., 1979. The performance of an inductively coupled plasma
on the model 5000 atomic absorption spectrophotometer. At. Absorpt. Newsl, 18:
41-45.
Edmond, C.R., 1969. Direct determination of fluoride in phosphate rock samples using
the specific ion electrode. Anal Chem., 41: 1327—1328.
Edmunds, W.M., Giddings, D.R. and Morgan-Jones, M., 1973. The application of flame-
less atomic absorption in hydrogeochemical analysis. At. Absorpt. Newsl, 12: 45—49.
223

El-Kholy, H.K., Burridge, J.C. and Scott, R.O., 1975. A tripleflow gas-sheathed DC arc
for spectrochemical analysis. Anal. Chim. Ada, 74: 247—252.
Ellis, A.J., Tooms,· J.S., Webb, J.S. and Bicknell, J.V., 1967. Application of solution
experiments in geochemical prospecting. Inst. Min. Metall., Trans., Sect. B, 76: 25—39.
(Discussion, 76: 216—217; and 77 : 136.)

Fabbi, B.P. and Moore, W.J., 1970. Rapid X-ray fluorescence determination of sulfur in
mineralized rocks from the Bingham Mining District, Utah. Appl. Spectrosc, 24: 426—
428.
Farmer, V.C. and Mitchell, B.D., 1963. Occurrence of oxalates in soil clays following
hydrogen peroxide treatment. Soil Sei., 96: 221—229.
Farrell, B.L., 1974. Fluorine, a direct indicator of fluorite mineralization in local and
regional geochemical surveys. J. Geochem. Explor., 3: 227—244.
Fassel, V.A., 1978. Quantitative elemental analyses by plasma emission spectroscopy.
Science, 202: 183—191.
Fassel, V.A. and Kniseley, R.N., 1974a. Inductively coupled plasma-optical emission
spectroscopy. Anal. Chem., 46: 1110A—1120A.
Fassel, V.A. and Kniseley, R.N., 1974b. Inductively coupled plasmas. Anal. Chem., 46:
1155A—1164A.
Faye, G.H., 1975. Description of ultramafic rock (ore) samples UM-1, UM-2, UM-4: a
correction. Chem. Geol, 15: 235—237.
Feather, C.E. and Willis, 1976. A simple method for background and matrix correction of
spectral peaks in trace element determination by X-ray fluorescence spectrometry.
X-Ray Spectrom., 5: 41—48.
Feldman, C , 1974. Preservation of dilute mercury solutions. Anal. Chem., 46: 99—102.
Fernandez, F.J., 1973. Atomic absorption determination of gaseous hydrides using
sodium borohydride reduction. At. Absorpt. Newsl., 12: 93—97.
Fernandez, F.J. and Manning, D.C., 1971. The determination of arsenic at sub-microgram
levels by atomic absorption spectrophotometry. At. Absorpt. Newsl., 10: 86—88.
Ficklin, W.H., 1970. A rapid method for the determination of fluoride in rocks and soils,
using an ion-selective electrode. U.S. Geol. Surv., Prof. Paper, 700-C: 186—188.
Ficklin, W.H., 1975. Ion-selective electrode determination of iodine in rocks and soils.
U.S. Geol. Surv., J. Res., 3: 753—755.
Ficklin, W.H. and Ward, F.N., 1976. Flameless atomic absorption determination of
bismuth in soils and rocks. U.S. Geol. Surv., J. Res., 4: 217—220.
Flanagan, F.J., 1969. U.S. Geological Survey standards, II. First compilation of data for
the new USGS rocks. Geochim. Cosmochim. Acta, 33: 81—120.
Fletcher, K., 1970. Some applications of background correction to trace metal analysis of
geochemical samples by atomic-absorption spectrophotometry. Econ. Geol., 65: 588—
589.
Fletcher, M.H., 1954. Collected papers on methods of analysis for uranium, 8. A study of
critical factors in the "direct" fluorimetric determination of uranium. U.S. Geol. Surv.,
Bull., 1006: 5 1 - 6 8 .
Floyd, M.A., Fassel, V.A. and D'Silva, A.P., 1980. Computer-controlled scanning meno-
chromator for the determination of 50 elements in geochemical and environmental
samples by inductively coupled plasma-atomic emission spectrometry. Anal. Chem.,
52:2168-2172.
Fordham, A.W., 1978. Correction of matrix effects in the analysis of stream waters by
flameless atomic absorption spectrophotometry. J. Geochem. Explor., 10: 41—51.
Foster, J.R., 1970. The application of direct-reading spectroscopy to the analysis of trace
metals in soils and stream sediments. Proc. Br. Ceram. Soc, 16: 25—31.
Foster, J.R., 1971. The reduction of matrix effects in atomic absorption analysis and the
efficiency of selected extractions in rock-forming minerals. In: R.W. Boyle (Editor),
Geochemical Exploration. Can. Inst. Min. Metall, Spec. Vol., 11: 554—560.
224

Foster, J.R., 1973. The efficiency of various digestion procedures on the extraction of
metals from rocks and rock-forming minerals. Can. Inst. Min. Metall., Bull., 66: 85—
92.
Frant, M.S. and Ross, J.W., Jr., 1968. Use of a total ionic strength adjustment buffer for
electrode determination of fluoride in water supplies. Anal. Chem., 40: 1169—1171.
Fratta, M., 1974. AAS determination of ppb amounts of thallium in silicate rocks. Can.
J. Spectrosc, 19: 33—37.
Friedrich, G.H., Kunzendorf, H. and Plüger, W.L., 1974. Shipborne geochemical investiga-
tions of deep-sea manganese-nodule deposits in the Pacific using a radio-isotope energy
dispersive X-ray system. J. Geochem. Explor., 3: 303—317.
Friedrich, G.H., Plüger, W.L., Hilmer, E.F. and Abu-Abed, I., 1973. Flameless atomic
absorption and ion-sensitive electrodes as analytical tools in copper exploration. In:
M.J. Jones (Editor), Geochemical Exploration 1972. Institution of Mining and
Metallurgy, London, pp. 437—445.
Fryer, B.J. and Kerrich, R., 1978. Determination of precious metals at ppb levels in rocks
by combined wet chemical and flameless atomic absorption method. At. Absorpt.
Newsl, 17: 4 - 6 .
Fuller, C.W., 1977. Electrothermal Atomization for Atomic Absorption Spectrometry.
Analytical Sciences Monograph No. 4. Chemical Society, London, 127 pp.

Gallagher, M.J., 1967. Determination of molybdenum, iron, and titanium in ores and
rocks by portable radioisotope X-ray fluoresence analyser. Inst. Min. Metall., Trans.,
Sect.B, 76: 1 5 5 - 1 6 4 .
Gallagher, M.J., 1969. Portable X-ray spectrometers for rapid ore analysis. In: Ninth
Commonwealth Mining and Metallurgical Congress. Institution of Mining and Metal-
lurgy, London, pp. 1—39.
Garreis, R.M. and Christ, C.L., 1965. Solutions, Minerals and Equilibria. Harper and Row,
New York, N.Y., 450 pp.
Garrett, R.G., 1969. The determination of sampling and analyical errors in exploration
geochemistry. Econ. Geoi, 64: 568—569.
Garrett, R.G., 1975. Copper and zinc in Proterozoic acid volcanic rocks as a guide to
exploration in the Bear Province. In: I.L. Elliott and W.K. Fletcher (Editors), Geo-
chemical Exploration 1974. Elsevier, Amsterdam, pp. 371—388.
Garrett, R.G. and Goss, T.I., 1979. The evaluation of sampling and analytical variation in
regional geochemical surveys. In: P.J. Watterson and P.K. Theobald (Editors), Geo-
chemical Exploration 1978. Association of Exploration Geochemists, Rexdale, Ont.,
pp. 3 7 1 - 3 8 3 .
Garrett, R.G. and Hornbrook, E.H.W., 1976. The relationship between zinc and organic
content in centre-lake bottom sediments·. J. Geochem. Explor., 5: 31—38.
Garrett, R.G. and Lynch, J.J., 1976. A comparison of neutron activation delayed neutron
counting versus fluorimetric analysis in large scale geochemical exploration for urani-
um. In: Exploration for Uranium Deposits. Proceedings, Symposium on Exploration
of Uranium Ore Deposits Vienna, 1976. International Atomic Energy Agency, Vienna,
pp. 321—334.
Garson, M.S. and Bateson, J.H., 1967. Possible use of the P.I.F. analyser in geochemical
prospecting for tin. Inst. Min. Metall., Trans., Sect. B, 76: 165—166.
Gatehouse, S., Russell, D.W. and Van Moort, J.C., 1977. Sequential soil analysis in
exploration geochemistry. J. Geochem. Explor., 8: 483—494.
Gedeon, A.Z., Butt, C.R.M., Gardner, K.A. and Hart, M.K., 1977. The applicability of
some geochemical analytical techniques in determining "total" compositions of some
lateritized rocks. J. Geochem. Explor., 8: 283—303.
Giauque, R.D., Garrett, R.B. and Goda, L.Y., 1977. Energy dispersive X-ray fluorescence
spectrometry for determination of twenty-six trace and two major elements in geo-
chemical specimens. Anal. Chem., 48: 62—67.
225

Goertzen, J.O. and Oster, J.D., 1972. Potentiometric titration of sulphate in water and
soil extracts using a lead electrode. Soil Sei. Soc. Am., Proc, 34: 691—693.
Golterman, H.L., 1970. Methods for Chemical Analysis of Fresh Waters. IBP Handbook
No. 8. Blackwell, Oxford, 166 pp.
Gong, H. and Suhr, M.H., 1976. The determination of cadmium in geological materials
by flameless atomic absorption. Anal. Chim. Acta, 8 1 : 297—303.
Goodfellow, W.D. and Wahl, J.L., 1976. Water extracts of volcanic rocks — detection of
anomalous halos at Brunswick No. 12 and Heath Steele B-zone massive sulphide depos-
its. J. Geochem. Explor., 6: 35—58.
Goodman, R.J., 1973. Rapid analysis of trace amounts of tin in stream sediments, soils
and rocks by X-ray fluorescence analysis. Econ. Geol, 68: 275—278.
Gorsuch, T.T., 1970. The Destruction of Organic Matter. Pergamon, Oxford, 152 pp.
Govett, G.J.S. and Whitehead, R.E., 1973. Errors in atomic absorption spectrophoto-
metric determination of Pb, Zn, Ni and Co in geologic materials. J. Geochem. Explor.,
2: 1 2 1 - 1 3 2 .
Green, T.E., Campbell, W.J. and Law, S.L., 1971. Application of a combined ion
exchange paper X-ray spectrographic method to geochemical exploration. In: R.W.
Boyle (Editor), Geochemical Exploration. Can. Inst. Min. Metall, Spec. Vol., 11: 516.
Greenfield, S., Jones, I.LI. and Berry, C.T., 1964. High-pressure plasmas as spectroscopic
emission sources. Analyst, 89: 713—720.
Greenfield, S., McGeachin, H.McD. and Smith, P.B., 1975a. Plasma emission sources in
analytical spectroscopy, I. Talanta, 22: 1—15.
Greenfield, S., McGeachin, H.McD. and Smith, P.B., 1975b. Plasma emission sources in
analytical spectroscopy, III. Talanta, 23: 1—14.
Greenfield, S., Jones, I.LI., McGeachin, H.McD. and Smith, P.B., 1975c. Automatic multi-
sample simultaneous multi-element analysis with a H.F. plasma torch and direct
reading spectrometer. Anal. Chim. Acta, 74: 225—245.
Greenland, L.P. and Campbell, E.Y., 1976. Rapid determination of nanogram amounts of
tellurium in silicate rocks. Anal. Chim. Acta, 87: 323—328.
Gregory, P. and Tooms, J.S., 1968. Geochemical prospecting for kimberlites. Q. Colo.
ach. Mines, 64: 2 6 5 - 3 0 5 .
Grifitts, W.R., Ward, F.N. and Alminas, H.V., 1976. A simple spot test for molybdenum
minerals. Econ. Geol., 7 1 : 1595.
Grimaldi, F.S. and Schnepfe, M.M., 1971. Determination of iodine in the p.p.m. range in
rocks. Anal. Chim. Acta, 53: 181—184.
Grimaldi, F.S., Ward, F.N. and Fuyat, R.K., 1954. Collected papers on methods of analy-
sis for uranium. U.S. Geol. Surv., Bull., 1006: 184 pp.
Grout, A. and Gallagher, M.J., 1980. Barium determination in rock and overburden by
portable XRF spectrometer. Inst. Min. Metall, Trans., Sect. B, 89: 130—133.
Gunton, J.E. and Nichol, I., 1975. Chemical zoning associated with the Ingerbelle—Copper
Mountain mineralization, Princeton, British Columbia. In: I.L. Elliott and W.K.
Fletcher (Editors), Geochemical Exploration 1974. Elsevier, Amsterdam, pp. 297—
312.

Hadeishi, T. and McLaughlin, R.D., 1966. Hyperfine Zeeman effect atomic absorption
spectrometer for mercury. Science, 174: 404—407.
Hall, G.E.M., 1979. A study of the stability of uranium in waters collected from various
geological environments in Canada. In: Current Research, Part A. Can. Geol. Surv.,
Paper, 79-1A: 361—365.
Hannaker, P. and Hughes, T.C., 1977. Multielement trace analysis of geological materials
with solvent extraction and flame atomic absorption spectrometry. Anal. Chem., 49:
1485-1488.
226

Hannaker, P. and Hughes, T.C., 1978. Improved solvent extraction scheme for geochemi-
cal trace analysis using flame atomic absorption. J. Geochem. Explor., 10: 169—180.
Hansel, J.M. and Martell, L.J., 1977. Automated energy dispersive X-ray determination of
trace elements in stream sediments. Los Alamos Sei. Lab., Informal Rep., LA-6869-
MS: 8 pp.
Hansuld, J.A., Mannard, G.W., Lakin, H.W., Canney, F.C., Salmon, M.L. and Weber, G.R.,
1969. What is a geochemical analysis? — a panel discussion. Q. Colo. Sch. Mines, 64:
5-26.
Harden, G. and Tooms, J.S., 1964. Efficiency of the potassium bisulphate fusion in geo-
chemical analysis. Inst. Min. Metall., Trans., Sect. B, 74: 129—141.
Hatch, W.R. and Ott, W.L., 1968. Determination of sub-microgram quantities of mercury
by atomic absorption spectrophotometry. Anal. Chem., 40: 2085—2087.
Hausen, D.M., Ahlrichs, J.W. and Odekirk, J.R., 1973. Application of sulphur and nickel
analyses to geochemical prospecting. In: J. Jones (Editor), Geochemical Exploration
1972. Institution of Mining and Metallurgy, London, pp. 13—24.
Hawkes, H.E., 1963. Dithizone field tests. Econ. Geol., 58: 579—586.
Hawkes, H.E. and Webb, J.S., 1962. Geochemistry in Mineral Exploration. Harper and
Row, New York, N.Y., 415 pp.
Haynes, S.J. and Clark, A.H., 1972. A rapid method for the determination of chlorine in
silicate rocks using ion-selective electrodes. Econ. Geol., 67: 378—382.
Head, P.C. and Nicholson, R.A., 1973. A cold vapour technique for the determination of
mercury in geological materials involving its reduction with tin (II) chloride and collec-
tion on gold wire. Analyst, 98: 53—56.
Heffernan, B.J., Archibold, R.O. and Vickers, T.J., 1967. Determination of tin in
cassiterite ores by atomic absorption spectrophotometry. Proc. Australas. Inst. Min.
Metall, 223: 6 5 - 6 9 .
Heiz, A.W., 1964. A gas jet for DC arc spectroscopy. U.S. Geol. Surv., Prof. Paper, 475D:
176-178.
Heiz, A.W., 1965. The problem of automatic plate reading and computer interpretation
for spectrochemical analysis. U.S. Geol. Surv., Prof. Paper, 525-B: 160—164.
Heiz, A.W., 1973. Spectrochemical computer analysis-instrumentation. U.S. Geol. Surv.,
J. Res., 1: 475—482. j
Herbert, A.J. and Young, G.G., 1977. Portable field water sample filtration unit. Anal.
Chem., 49: 672—673.
Hill, W.E. Jr., 1975. The use of analytical standards to control assaying projects. In: I.L.
Elliott and W.K. Fletcher (Editors), Geochemical Exploration 1974. Elsevier, Amster-
dam, pp. 651—657.
Hoffman, S.J., 1974. Pebble cards —a record of the coarse fraction of stream sediments
for geochemical exploration. J. Geochem. Explor., 3: 387—388.
Hoffman, S.J. and Fletcher, W.K., 1979. Extraction of Cu, Zn, Mo, Fe and Mn from soils
and sediments using a sequential procedure. In: J.R. Watterson and P.K. Theobald
(Editors), Geochemical Exploration 1978. Association of Exploration Geochemists,
Rexdale, Ont., pp. 289—299.
Hoffman, S.J. and Fletcher, W.K., 1980. Organic matter scavenging of copper, zinc,
molybdenum, iron and manganese, estimated by sodium hypochlorite extraction
(pH 9.5). Abstr., 8th Int. Geochem. Explor. Symp., Hanover, 1980.
Hoffman, S.J. and Fletcher, W.K., 1981. Detailed lake sediment geochemistry of anoma-
lous lakes on the Nechako Plateau, central British Columbia: comparison of trace
metal distributions in Capoose and Fish Lakes. J. Geochem. Explor., 14: 221—244.
Hoffman, S.J. and Waskett-Myers, M.J., 1974. Determination of molybdenum in soils and
sediments with a modified zinc dithiol procedure. J. Geochem. Explor., 3: 61—66.
Holman, R.H.C., 1956-57. A method for determining readily-soluble copper in soil and
alluvium — introducing white spirit for dithizone. Inst. Min. Metall., Trans., Sect. B,
66: 7 - 1 6 .
227

Holman, R.H.C., 1963. Field and laboratory methods used by the Geological Survey of
Canada in geochemical surveys, 2. A method for determining readily-soluble copper in
soil and alluvium. Geol. Surv. Can., Paper, 63-67: 5 pp.
Holman, R.H.C. and Durham, C.C., 1967. A mobile spectrographic laboratory. Geol.
Surv. Can., Paper, 66-35: 15 pp.
Hopkins, D.M., 1977. An improved ion-selective electrode method for the rapid deter-
mination of fluorine in rocks and soils. U.S. Geol. Surv., J. Res., 5: 589—593.
Horsnail, R.F., Nichol, I. and Webb, J.S., 1969. Influence of variations in secondary
environment on the metal content of drainage sediments. Q. Colo. Sch. Mines, 64:
307-322.
Horton, R. and Lynch, J.J., 1975. A geochemical field laboratory for the determination
of some trace elements in soil and water samples. Geol. Surv. Can., Paper, 75-1, Part A:
213-214.
Howarth, R.J., 1977. Automated generation of randomised sample submittal schemes for
laboratory analysis. Comput. Geosci., 3: 327—334.
Howarth, R.J. and Lowenstein, P.L., 1971. Sampling variability of stream sediments in
broad-scale regional geochemical reconnaissance. Inst. Min. Metall., Trans., Sect. B, 80:
363-372.
Howarth, R.J. and Martin, L., 1979. Computer-based techniques in the compilation,
mapping and interpretation of exploration geochemical data. In: P.J. Hood (Editor),
Geophysics and Geochemistry in the Search for Metallic Ores. Geol. Surv. Can., Econ.
Geol Rep., 3 1 : 545—574.
Howarth, R.J. and Thompson, M., 1976. Duplicate analysis in geochemical practice, II.
Examination of proposed method and examples of its use. Analyst, 101: 699—709.
Hower, J., 1959. Matrix corrections in the X-ray spectrographic trace element analysis of
rocks and minerals. Am. Mineral, 44: 19—32.
Hubert, A.E. and Lakin, H.W., 1973. Atomic absorption determination of thallium and
indium in geologic materials. In: M.J. Jones (Editor), Geochemical Exploration 1972.
Institution of Mining and Metallurgy, London, pp. 383—387.
Huffman, C , Rahill, R.L., Shaw, V.E. and Norton, D.R., 1972. Determination of mer-
cury in geologic materials by flameless atomic absorption spectrometry. U.S. Geol
Surv., Prof. Paper, 800-C: 203—207.
Hulanicki, A., Lewandowski, R. and Lewenstam, A., 1976. Elimination of ionic inter-
ferences in the determination of sulphates in water using the lead-sensitive ion-selec-
tive electrode. Analyst, 101: 939—942.
Hulanicki, A., Trojanowicz, M. and Krawczyk, T.K.V., 1977. Determination of copper in
water by means of chalcocite copper ion-selective electrode. Water Res., 1 1 : 627—630.
Hunt, E.C., North, A.A. and Wells, R.A., 1955. Application of paper Chromatographie
methods of analysis to geochemical prospecting. Analyst, 80: 172—194.
Hunt, E.C., Stanton, R.E. and Wells, R.A., 1959-60. Field determination of beryllium in
soils for geochemical prospecting. Inst. Min. Metall, Trans., Sect. B, 69: 361—369.
Husler, J., 1972. Determination of niobium in rocks, ores and alloys by atomic-absorp-
tion spectrophotometry. Talanta, 19: 863—869.
Hutchison, C.S., 1974. Laboratory Handbook of Petrographic Techniques. Wiley-Inter-
science, New York, N.Y., 527 pp.
Hutton, R.C., Ottaway, J.M., Rains, T.C. and Epstein, M.S., 1977. Determination of
barium in calcium carbonate rocks by carbon furnace atomic emission spectrometry.
Analyst, 102: 429.

Ingamells, C O . , 1970. Lithium metaborate flux in silicate analysis. Anal Chim. Ada, 52:
323-334.
Ingamells, C O . , 1974. New approaches to geochemical analysis and sampling. Talanta,
21: 1 4 1 - 1 5 5 .
228

Ingamells, C O . and Switzer, P., 1973. A proposed sampling constant for use in geochemi-
cal analysis. Talanta, 20: 547—568.
Ingram, B.L., 1970. Determination of fluoride in silicate rocks without separation of
aluminum using a specific ion electrode. Anal. Chem., 42: 1825—1827.
Irving, 1977. Dithizone. Analytical Sciences Monograph No. 5. Chemical Society, London,
106 pp.
Iskander, I., Syers, J.K., Jacobs, L.W., Keeney, D.R. and Gilmour, J.T., 1972. Determina-
tion of total mercury in sediments and soils. Analyst, 97: 388—393.

Jackson, M.L., 1958. Soil Chemical Analysis. Prentice-Hall, New York, N.Y., 498 pp.
Jagner, D. and Pavlova, V., 1972. A standard addition titration method for the determina-
tion of fluorine in silicate rocks. Anal. Chim. Ada, 60: 153—158.
James, C.H., 1970. A rapid method for calculating the statistical precision of geochemical
prospecting analysis. Inst. Min. Metall, Trans., Sect. B, 79: 91—96.
James, C.H. and Webb, J.S., 1964. Sensitive mercury vapour meter for use in geochemical
prospecting. Inst. Min. Metall, Trans., Sect. B, 73: 633—641.
James, G.W., 1977. Parts-per-million determinations of uranium and thorium in geologic
samples by X-ray spectrometry. Anal. Chem., 49: 967—968.
Jardine, M.A., 1963. Field and laboratory methods used by the Geological Survey of
Canada in geochemical surveys, 4. A laboratory method for determining antimony in
soils and rocks. Geol. Suru. Can., Paper, 63-29: 11 pp.
Jenkins, R., 1970. Rapid Solutions to the Particle Size Problems in X-ray Fluorescence
Spectrometry. Philips, Eindhoven, 4 pp.
Jenkins, R. and De Vries, J.L., 1970. Practical X-ray Spectrometry. Philips, Eindhoven,
190 pp.
Jonasson, I.R., 1974. Some comments on interferences by Cu(II) ions and Ag(I) ions in
the wet reduction-flameless atomic absorption determination of mercury. J. Geochem.
Explor., 3: 77—81.
Jonasson, I.R., Lynch, J.J. and Trip, L.J., 1973. Field and laboratory methods used by
the Geological Survey of Canada in geochemical surveys, 12. Mercury in ores, rocks,
soils, sediments and water. Geol. Surv. Can., Paper, 73-21: 21 pp.
Jones, L.H.P., 1957. The solubility of molybdenum in simplified systems and aqueous
soil suspension. J. Soil Sei., 8: 313—327.
Jones, M.P. and Beaven, C.H.J., 1971. Sampling of non-Gaussian mineralogical distribu-
tions. Inst. Min. Metall, Trans., Sect. B, 80: 316—323.
Josephson, M., Cook, E.B.T. and Dixon, K., 1977. A rapid method for the determination
of fluoride in geologic samples. Rep. Natl. Inst. Metall, S. Afr., 1886: 8 pp.

Kauranne, L.K. (Editor), 1976. Conceptual Models in Exploration Geochemistry —


Norden 1975. J. Geochem. Explor., 5: 421 pp.
Kauranne, L.K. and Nurmi, A., 1967. The analysis of trace amounts of copper in soil with
neocuproine. In: Kvalheim (Editor), Geochemical Prospecting in Fennoscandia. Inter-
science, New York, N.Y., pp. 331—334.
Kelly, W.R. and Moore, C.B., 1973. Iron spectral interference in the determination of
zinc by atomic absorption spectrometry. Anal. Chem., 4 5 : 1274—1275.
Kennedy, V.C. and Brown, T.C., 1966. Experiments with a sodium-ion electrode as a
means of studying cation-exchange rates. In: W.F. Bradley and S.W. Bailey (Editors),
Clays and Clay Minerals, Proceedings of the Thirteenth National Conference. Pergamon,
Oxford, 453 pp.
Kesler, S.E., Van Loon, J.C. and Bateson, J.H., 1973. Analysis of fluoride in rocks and an
application to exploration. J. Geochem. Explor., 2: 11—18.
Kim, C.H., Alexander, P.W. and Smythe, L.E., 1976. Use of long chain alkylamines for
preconcentration and determination of traces of molybdenum, tungsten and rhenium
229

by atomic absorption spectroscopy, III. Tungsten in geological samples. Talanta, 23:


573-578.
Kim, C.H., Owens, CM. and Smythe, L.E., 1974. Determination of traces of Mo in soils
and geological materials by solvent extraction of the molybdenum-thiocyanate com-
plex and atomic absorption. Talanta, 2 1 : 445—453.
Kinrade, D. and Van Loon, J.C., 1964. Solvent extraction for use with flame atomic
absorption. Anal. Chem., 46: 1894—1898.
Kinson, K. and Belcher, C.B., 1970. A mobile laboratory for nickel prospecting and
drilling operations. Proc. Australas. Inst. Min. Metall, 234: 59—65.
Kirkbright, G.F. and Johnson, H.N., 1973. Application of indirect methods in analysis by
atomic-absorption spectrometry. Talanta, 20: 433—451.
Kirkbright, G.F. and Ward, A.F., 1974. Atomic-emission spectrometry with an induction-
coupled high-frequency plasma source. Comparison with the inert-gas shielded pre-
mixed nitrous oxide—acetylene flame for multi-element analysis. Talanta, 2 1 : 1145—
1165.
Kirkbright, G.F., Smith, A.M. and West, T.S., 1967. An indirect sequential determination
of phosphorous and silicon by atomic-absorption spectrophotometry. Analyst, 92:
411-416.
Koirtyohann, S.R. and Pickett, E.E., 1966. Spectral interferences in atomic absorption
spectrometry. Anal. Chem., 38: 585—587.
Kokot, M.L., 1976. A modification of a hydride generation apparatus for the rapid
determination of arsenic in a large number of geochemical and other samples. At.
Absorpt. NewsL, 15: 105.
Koksoy, M., Bradshaw, P.M.D. and Tooms, J.S., 1967. Notes on the determination of
mercury in geologic samples. Inst. Min. Metall, Trans., Sect B, 76: 121—124.
Kontas, E., 1976. Determination by flameless atomic absorption spectrometry of molyb-
denum in till and sediments. J. Geochem. Explor., 5: 400—402.
Kothny, EX., 1969. Trace determination of mercury, thallium and gold with crystal
violet. Analyst, 94: 198—203.
Kothny, E.L., 1974. Simple trace determination of platinum in geologic materials. J.
Geochem. Explor., 3: 291—299.
Kramar, U. and Puchelt, H., 1980. Application of radionuclide —X-ray fluorescence ana-
lysis in geochemical prospecting. Abstr., 8th Int. Geochem. Explor. Symp., Hanover,
1980.
Krings, H., Van den Boom, G. and Gundlach, H., 1976. A rapid method for measuring
arsenic in soil and related material and water. Geol. Jahrb, Reihe D, 16: 85—91 (in
English).
Kroonen, J. and Vader, D., 1963. Line Interference in Emission Spectrographic Analysis.
Elsevier, Amsterdam, 213 pp.
Kvalheim, A., 1967. Field laboratories, and field and laboratory equipment, used in
Norwegian geochemical prospecting. In: A. Kvalheim (Editor), Geochemical Pros-
pecting in Fennoscandia. Interscience, New York, N.Y., pp. 321—330.

Lakin, H.W. and Nakagawa, H.M., 1965. A spectrophotometric method for the deter-
mination of traces of gold in geologic materials. U.S. Geol. Surv., Prof. Paper, 525-C:
168—171.
Lakin, H.W., Stevens, R.E. and Almond, H., 1949. Field method for the determination of
zinc in soils. Econ. Geol, 44: 296—306.
Lalonde, J.P., 1974. Research in geochemical prospecting methods for fluorite deposits,
Madoc area, Ontario. Geol Surv. Can., Paper, 73-38: 56 pp.
Lalonde, J.P., 1976. Fluorine — an indicator of mineral deposits. Can. Min. Metall Bull,
69: 110—122.
230

Langmuir, D., 1978. Uranium solution—mineral equilibria at low temperatures with


applications to sedimentary ore deposits. Geochim. Cosmochim. Acta, 42: 547—569.
Langmyhr, F.J., 1977. Direct atomic-absorption spectrometric analysis of geological
materials — a review. Talanta, 24: 277—282.
Langmyhr, F.J. and Paus, P.E., 1968. The analysis of inorganic siliceous materials by
atomic absorption spectrophotometry and the hydrofluoric acid decomposition tech-
nique, 1. The analysis of silicate rocks. Anal. Chim. Acta, 4 3 : 397—408.
Langmyhr, F.J., Stubergh, J.R., Thomassen, Y., Hassen, J.E. and Dolezal, J., 1974.
Atomic absorption spectrometric determination of cadmium, lead, silver, thallium and
zinc in silicate rocks by direct atomization from the solid state. Anal. Chim. Acta, 7 1 :
35-42.
Lapointe, G., 1968. The A horizon in soil surveys; geochemical kit for trace amounts of
molybdenum. Econ. Geol, 68: 572.
Larson, G.F., Fassel, V.A., Scott, R.H. and Kniseley, R.N., 1975. Inductively coupled
plasma—optical emission analytical spectrometry. A study of some interelement
effects. Anal Chem., 47: 238—243.
Lavergne, P.J., 1965. Field and Laboratory methods used by the Geological Survey of
Canada, 8. Preparation of geologic materials for chemical and spectrographic analysis.
Geol Surv. Can., Paper, 65-18: 23 pp.
Lavkulich, L.M. and Wiens, J.H., 1970. Comparison of organic matter destruction by
hydrogen peroxide and sodium hypochlorite and its effects on selected mineral con-
stituents. Soil Sei. Soc. Am., Proc, 34: 755—758.
Leake, B.E., Hendry, G.L., Kemp, A., Plant, A.G., Harrey, P.K., Wilson, J.R., Coats, J.S.,
Aueott, J.W., Liinel, T. and Howarth, R.J., 1969. The chemical analysis of rock
powders by automatic X-ray fluorescence. Chem. Geol., 5: 7—86.
Leake, R.C. and Aueott, J.W., 1973. Geochemical mapping and prospecting by use of
rapid automatic X-ray fluorescence analysis of panned concentrates. In: M.J. Jones
(Editor), Geochemical Exploration 1972. Institution of Mining and Metallurgy,
London, pp. 391—402.
Leake, R.C. and Peachey, D., 1973. Rapid determination of trace elements in organic-
rich soils by automatic X-ray fluorescence spectrometry. Inst. Min. Metall, Trans.,
Sect B, 82: 2 5 - 2 7 .
LeRiche, H.H. and Weir, A.H., 1963. A method of studying trace elements in soil frac-
tions. J. Soil Sei., 14: 225—235.
Levinson, A.A., 1974. Introduction to Exploration Geochemistry. Applied Publishing,
Calgary, Alta., 612 pp.
Levinson, A.A., 1975. A rapid X-ray fluoresence procedure applicable to exploration geo-
chemistry. J. Geochem. Explor., 4: 399—408.
Lewis, T.E. and Broadbent, F.E., 1961. Soil organic matter—metal complexes, IV. Nature
and properties of exchange sites. Soil Sei., 9 1 : 393—399.
Lister, B. and Gallagher, M.J., 1970. An inter-laboratory survey of the accuracy of ore
analysis. Inst. Min. Metall, Trans., Sect. B, 79: 213—237.
Lockhart, A.W., 1976. Geochemical prospecting of an appalachian porphyry copper
deposit at Woodstock, New Brunswick, J. Geochem. Explor., 6: 13—33.
Lovell, J.S. and Hale, M., 1980. Geochemical sample digestion by a high pressure acid
leach. Abstr., 8th Int. Geochem. Explor. Symp., Hannover, 1980, p. 137.
Luke, C.L., 1968. Determination of trace elements in inorganic and organic materials by
X-ray fluorescence spectroscopy. Anal. Chim. Acta, 4 1 : 237—250.
Lynch, J.J., 1967. A sample site test for the determination of nickel in water. Geol. Surv.
Can., Paper, 67-1, Part B: 25—26.
Lynch, J.J., 1971. The determination of copper, nickel and cobalt in rocks by atomic
absorption spectrophotometry using a cold leach: In: R.W. Boyle (Editor), Geo-
chemical Exploration. Can. Inst. Min. Metall, Spec. Vol., 1 1 : 313—314.
231

Lynch, J.J. and Mihailov, G., 1963. Field and laboratory methods used by the Geological
Survey of Canada in Geochemical surveys, 3. Method for determining arsenic. Geol.
Surv. Can., Paper, 63-8: 12 pp.
Lynch, J.J., Garrett, R.G. and Jonasson, I.R., 1973. A rapid estimation of organic carbon
in silty lake sediments. J. Geochem. Explor., 2: 171—174.

McDonald, A.J. and Stanton, R.E., 1962. The spectrophotometric determination of tin in
soil with gallein. Analyst, 87: 600—602.
McHugh, J.B. and Welsch, E.P., 1975. Atomic absorption determination of antimony in
geological materials. In: F.N. Ward (Editor) New and Refined Methods of Trace Analy-
sis Useful in Geochemical Exploration. U.S. Geol. Surv., Bull., 1408: 5—11.
McKeague, J.A. and Day, J.H., 1966. Dithionite and oxalate-extractable Fe and Al as aids
in differentiating various classes of soils. J. Soil Sei., 46: 13—22.
MacKenzie, R.C., 1954. Free iron-oxide removal from soils. J. Soil Sei., 5: 167—172.
McNerney, J.J. and Buseck, P.R., 1973. Geochemical exploration using mercury vapour.
Econ. Geol., 68: 1 3 1 3 - 1 3 2 0 .
McNerney, J.J., Buseck, P.R. and Hanson, R.C., 1972. Mercury detection by means of
thin gold films. Science, 178: 611—612.
McQuaker, N.R. and Gurney, M., 1977. Determination of total fluoride in soil and vegeta-
tion using an alkali fusion-seleotive ion electrode technique. Anal. Chem., 49: 53—86.
McQuaker, N.R., Kluckner, P.D. and Chang, G.N., 1979. Calibration of an inductively
coupled plasma-atomic emission spectrometer for the analysis of environmental mate-
rials. Anal. Chem., 5 1 : 888—895.
Mahaffey, E.J., 1974. A spectrophotometric method for the determination of rhenium in
geologic materials. J. Geochem. Explor., 3: 53—59.
Marchant, J.W. and Klopper, B.C., 1978. Microgram metal contamination of water/nitric
acid after four years in linear polyethylene containers. J. Geochem. Explor., 9: 103—
107.
Marciello, L. and Ward, A.F., 1978. Interelement corrections for spectral line inter-
ferences. Jarrell-Ash Plasma Newsl., 1: 12—13.
Marinenko, J., May, I. and Dinnin, J.I., 1972. Determination of mercury in geologic mate-
rials by flameless atomic absorption spectrometry. U.S. Geol. Surv., Prof. Paper,
800-B: 1 5 1 - 1 5 5 .
Marranzino, A.P. and Wood, W.H., 1956. Multiple-unit fusion rack. Anal. Chem., 28:
273-274.
Marshall, N.J., 1964. Rapid determination of molybdenum in geochemical samples using
dithiol. Econ. Geol., 59: 142—148.
Marshall, N.J., 1968. Rapid determination of molybdenum in geochemical samples using
dithiol. Econ. Geol, 6 3 : 291.
Marshall, N.J., 1978. Colorimetric determination of arsenic in geochemical samples. J.
Geochem. Explor., 10: 307—313.
Martin, T.D., Kopp, J.F. and Ediger, R.D., 1975. Determining selenium in water, waste-
water, sediment, and sludge by flameless atomic absorption spectroscopy. At. Absorpt.
Newsl, 14: 1 0 9 - 1 1 6 .
Maruta, T., Takeuchi, T. and Suzuki, M., 1972. Atomic-absorption and emission inter-
ferences on barium. Anal. Chim. Ada, 58: 452—455.
Maxwell, J.A., 1968. Rock and Mineral Analysis. Interscience, New York, N.Y., 584 pp.
Maynard, D.E. and Fletcher, W.K., 1973. Comparison of total and partial extractable
copper in anomalous and background peat samples. J. Geochem. Explor., 2: 19—24.
Mehra, O.P. and Jackson, M.L., 1960. Iron oxide removal from soils and clays by a
dithionite-citrate system buffered with sodium bicarbonate. In: Clays and Clay Miner-
als, Proceedings of the Seventh National Conference. Pergamon, Oxford, pp. 317—327.
Meier, A.L., 1980. Flameless atomic-absorption determination of gold in geological mate-
rials. J. Geochem. Explor., 13: 77—85.
232

Meyer, W.T. and Lam Shang Leen, Y.C.Y., 1973. Microwave-induced argon plasma emis-
sion system for geochemical trace analysis. In: M.J. Jones (Editor), Geochemical
Exploration 1972. Institution of Mining and Metallurgy, London, pp. 325—335.
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. 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.
Minkkinen, P., 1975. A method for the correction of the background absorption in silver
analysis of calcareous samples. At. Absorpt. Newsl., 14: 71—72.
Mitchell, B.D. and Smith, F.L., 1974. The removal of organic matter from soil extracts
by bromine oxidation. J. Soil Sei., 25: 239—241.
Mitchell, B.D., Smith, B.F.L. and de Endredy, A.S., 1971. The effect of buffered sodium
dithionite solution and ultrasonic agitation on clays. Isr. J. Chem., 9: 45—52.
Mitchell, R.L., 1964a. Trace elements in soils. In: F.E. Bear (Editor), Chemistry of the
Soil. American Chemical Society, Monograph Series No. 160. Reinhold, New York,
N.Y., 2nd ed., pp. 320—368.
Mitchell, R.L., 1964b. The Spectrochemical Analysis of Soil, Plants and Related Mate-
rials. Commonwealth Bureau of Soils, Harpenden, Tech. Comm. No. 44A, 225 pp.
Morrison, G.H., 1971. Evaluation of lunar elemental analyses. Anal. Chem., 4 3 : 22A—
31A.
Morrison, G.H. and Freiser, H., 1957. Solvent Extraction in Analytical Chemistry. Wiley,
New York, N.Y., 269 pp.
Motooka, J.M., Mosier, E.L., Sutley, S.J. and Viets, J.G., 1979. Induction-coupled plasma
determination of Ag, Au, Bi, Cd, Cu, Pb and Zn in geologic materials using a detective
extraction technique — preliminary investigation. Appl. Spectrosc, 33: 456—460.
Muir, G.D. (Editor), 1977. Hazards in the Chemical Laboratory. Chemical Society,
London, 489 pp.
Mulford, C.E., 1966. Solvent extraction techniques for atomic absorption spectroscopy.
At. Absorpt. Newsl., 5: 88—90.
Myers, A.T. and Wood, W.H., 1960. Ceramic mills in a paint mixer for preparation of
multiple rock samples. Appl. Spectrosc, 14: 136—138.
Myers, A.T., Havens, R.G. and Dunton, P.J., 1961. A spectrochemical method for the
semi-quantitative analysis of rocks, minerals and ores. U.S. Geol. Surv., Bull., 1084-1:
207-229.

Nakagawa, H.M., 1975. Atomic absorption determination of silver, bismuth, cadmium,


cobalt, copper, nickel, lead, and zinc in calcium- and iron-rich geological materials.
In: F.N. Ward (Editor) New and Refined Methods of Trace Analysis Useful in Geo-
chemical Exploration. U.S. Geol. Surv., Bull, 1408: 85—96.
Nakagawa, H.M. and Lakin, H.W., 1965. A field method for the determination of silver in
soils and rocks. U.S. Geol. Surv., Prof Paper, 525-C: 172—175.
Nakagawa, H.M. and Thompson, C.E., 1968. Atomic absorption determination of telluri-
um. U.S. Geol. Surv., Prof. Paper, 600-B: 1 2 3 - 1 2 5 .
Nichol, I. and Henderson-Hamilton, J.C., 1965. A rapid quantitative spectrographic meth-
od for the analysis of rocks, soils and stream sediments. Inst. Min. Metall., Trans.,
Sect.B, 74: 9 5 5 - 9 6 1 .
Nicolas, D.J., 1971. The determination of antimony in geological materials by atomic
absorption spectrophotometry, with particular reference to soils. Anal. Chim. Ada,
55: 59—66.
North, A.A., 1956. Geochemical field methods for the determination of tungsten and
molybdenum in soils. Analyst, 8 1 : 660—668.
233

Nowlan, G.A., 1970. A field method for the determination of cold-extractable nickel in
stream sediments and soils. U.S. Geol Surv., Prof. Paper, 700-B: 177—180.

Olade, M. and Fletcher, K., 1974. Potassium chlorate-hydrochloric acid; a sulfide selec-
tive leach for bedrock geochemistry. J. Geochem. Explor., 3: 337—344.
Olade, M.A. and Fletcher, W.K., 1975. Primary dispersion of rubidium and strontium
around porphyry copper deposits, Highland Valley, British Columbia. Econ. Geol.,
70: 1 5 - 2 1 .
Olade, M. and Fletcher, K., 1976. Distribution of sulphur, and sulphide iron and copper
in bedrock associated with porphyry copper deposits, Highland Valley, British
Columbia. J. Geochem. Explor., 5: 21—30.
Olade, M.A. and Goodfellow, W.D., 1979. Lithogeochemistry and hydrogeochemistry of
uranium and associated elements in the Toombstone Batholith, Yukon, Canada. In:
J.R. Watterson and P.K. Theobald (Editors), Geochemical Exploration 1978. Associa-
tion of Exploration Geochemists, Rexdale, Ont., pp. 407—428.
Ooghe, W. and Verbeek, F., 1974. Atomic absorption spectrometry of the lanthanides in
minerals and ores. Anal. Chim. Acta, 73: 87—95.
Orion Research, 1970. Gran's plots and other schemes. Orion Res. Inc. Newsl., 11 (11-12):
49-55.
Orion Research, 1975. R7 Sulphate in natural waters. In: Analytical Methods Guide.
Orion Research, 7th ed., 33 pp.
Orion Research, 1976. Fluoride in drinking water. Orion Res. Inc., Appl. Bull., 5A: 2 pp.

Pahlavanpour, B., Thompson, M. and Walton, S.J., 1979. The determination of tin in
geochemical samples by inductively coupled plasma atomic emission spectrometry.
J. Geochem. Explor., 12: 45—55.
Parslow, G.R., 1979. Infrared absorption as a consistent method of correcting matrix
effect(s) of pellets used in the fluorometric determination of uranium. Chem. Geol,
27: 3 4 3 - 3 5 3 .
Parslow, G.R. and Dwairi, I., 1976. Precision and accuracy in the fluorometric determina-
tion of uranium in lake sediment. In: C.E. Dunn (Editor), Uranium in Saskatchewan.
Sask. Geol. Soc, Spec. Publ, 3: 169—191.
Parslow, G.R. and Dwairi, I., 1977. Extraction of uranium ions from water — a rapid and
convenient method using prepackaged ion-exchange resins. J. Geochem. Explor., 8:
541-547.
Parsons, M.L., Foster, A. and Anderson, D., 1980. An Atlas of Spectral Interferences in
ICP Spectroscopy. Plenum, New York, N.Y., 644 pp.
Peachey, D., 1976. Extraction of copper from ignited soil samples. J. Geochem. Explor.,
5: 1 2 9 - 1 3 4 .
Peachey, D. and Allen, B.P., 1977. An investigation into the selective dissolution of sul-
fide phases from stream sediments and soils. J. Geochem. Explor., 8: 571—577.
Peachey, D. and Vickers, B.P., 1978. An investigation into the determination of Pb in
geochemical exploration samples. Inst. Geol. Sei., Inf. Ser. Rep., 48: 8 pp.
Peachey, D., Cooper, D.C. and Allen, B.P., 1978. Rapid colorimetric determination of
copper in drill-sludge samples. Inst. Min. Metall, Trans., Sect. B, 87: 136—139.
Peachey, D., Roberts, J.L. and Scott-Baker, J., 1973. Rapid colorimetric determination
of phosphorus in geochemical survey samples. J. Geochem. Explor., 2: 115—120.
Peech, M., 1965. Hydrogen-ion activity. In: C.A. Black et al. (Editors), Methods of Soil
Analysis, Part 2. American Society of Agronomy, Madison, Wise, pp. 914—926.
Pettijohn, F.J., 1957. Sedimentary Rocks. Harper and Row, New York, N.Y., 718 pp.
Phillips, R., 1971. A method for estimating the grade of diamond deposits. Inst. Min.
Metall, Trans., Sect. B, 80: 357—362.
234

Phuphatana, A., Carlson, E.H. and Manus, R.W., 1976. The fluoride ion-specific elec-
trode as applied to fluorspar deposits in southwest Texas and southern Illinois. Econ.
GeoL, 7 1 : 6 6 1 - 6 6 5 .
Plamondon, J., 1968. Rapid determination of uranium in geochemical samples by paper
chromatography. Econ. GeoL, 6 3 : 76—79.
Plant. J., 1971. Orientation studies on stream-sediment sampling for a regional geochemi-
cal survey in northern Scotland. Inst. Min. Metall, Trans., Sect. B, 80: 324—345.
Plant, J., 1973. Random numbering system for geochemical samples. Inst. Min. Metall.,
Trans., Sect. B, 82: 64—65.
Plant, J., Jeffery, K., Gill, E. and Fage, C , 1975. The systematic determination of accu-
racy and precision in geochemical exploration data. J. Geochem. Explor., 4: 467—486.
Plüger, W.L. and Friedrich, G.H., 1973. Determination of total and cold-extractable
fluoride in soils and stream sediments with an ion-sensitive fluoride electrode. In: M.J
M.J. Jones (Editor), Geochemical Exploration 1972. Institution of Mining and Metal-
lurgy, London, pp. 423—429.
Price, G.R., Ferretti, R.J. and Schwartz, S., 1953. Fluorophotometric determination of
uranium. Anal. Chem., 25: 322—331.

Quin, B.F. and Brooks, R.R., 1972. The rapid determination of tungsten in soils, rocks
and vegetation. Anal. Chim. Ada, 58: 301—309.

Rann, C.S. and Hambly, A.N., 1965. Distribution of atoms in an atomic absorption
flame. Anal. Chem., 37: 879—884.
Rao, P.D., 1971. Determination of molybdenum in geological materials. At. Absorpt.
Newsi, 10: 1 1 8 - 1 1 9 .
Rashid, M.A. and Leonard, J.D., 1973. Modifications in the solubility and precipitation
behaviour of various metals as a result of their interaction with sedimentary humic
acid. Chem. GeoL, 1 1 : 8 9 - 9 7 .
Reid, W.R., 1969. Mineral staining tests. Colo. Sch. Mines, Miner. Ind. Bull., 12(3): 1—20.
Reynolds, R . C , 1963. Matrix corrections in trace element analysis by X-ray fluorescence:
estimation of the mass absorption coefficient by Compton scattering. Am. Mineral.,
48: 1 1 3 3 - 1 1 4 3 .
Reynolds, R . C , 1967. Estimation of mass absorption coefficients by Compton Scattering:
improvements and extensions of the method. Am. Mineral., 82: 1493—1502.
Ritchie, A.S., 1964. Chromatography in Geology. Elsevier, Amsterdam, 185 pp.
Ritchie, A.S., 1969. Recent advances in the Chromatographie analysis of geologic mate-
rials. Q. Colo. Sch. Mines, 64: 427—436.
Robbins, J . C , 1973. Zeeman spectrometer for measurement of atmospheric mercury
vapour. In: M.J. Jones (Editor), Geochemical Exploration 1972. Institution of Mining
and Metallurgy, London, pp. 315—323.
Robbins, J . C , 1978. Field technique for the measurement of uranium in natural waters.
Can. Inst. Min. Metall., Bull., 7 1 : 61—67.
Roberts, J.L., 1971. An improved colorimetric determination of vanadium in geochemical
prospecting samples. Talanta, 18: 1070—1072.
Rooney, R.C. and Woolley, J.F., 1978. Interference of calcium on barium as a means of
assessing atomic-absorption spectrophotometers. Analyst, 103: 1100—1103.
Rose, A.W., 1975. The mode of occurrence of trace elements in soils and stream sedi-
ments applied to geochemical exploration. In: I.L. Elliott and W.K. Fletcher (Editors),
Geochemical Exploration 1974. Elsevier, Amsterdam, pp. 699—705.
Rose, A.W. and Keith, M.L., 1976. Reconnaissance geochemical techniques for detecting
uranium deposits in sandstones of northeastern Pennsylvania. J. Geochem. Explor., 6:
119-137.
235

Rose, A.W. and Suhr, N.H., 1971. Major element content as a means of allowing for back-
ground variation in stream-sediment geochemical exploration. In: R.W. Boyle (Editor),
Geochemical Exploration. Can. Inst. Min. Metall., Spec. Vol., 11: 587—593.
Rose, A.W., Hawkes, H.E. and Webb, J.S., 1979. Geochemistry in Mineral Exploration.
Academic Press, London, 658 pp.
Rose, E.R., 1969. A progress report on experiments with chemical field tests for the
detection of the rare-earth elements cerium and yttrium. Geol. Surv. Can., Paper, 69-
15: 13 pp.
Rose, E.R., 1976. A field test for rare earth elements. Geol. Survey Can., Paper, 75-16:
10 pp.
Rubeska, I., Koreckova, J. and Weiss, D., 1977. The determination of gold and palladium
in geological materials by atomic absorption after extraction with dibutyl sulphide.
At. Absorpt. Newsl., 16: 1—3.
Rundle, L., Holmes, K.A. and Peachey, D., (?). The determination of arsenic by the silver
diethydithiocarbamate procedure. Inst. Geol. Sei., Geochem. Method, 16.
Ryland, A.L., 1964. A general approach to the X-ray spectroscopic analysis of samples of
low atomic number. Abstr., Am. Chem. Soc. Annu. Meet.

Sandell, E.B., 1959. Colorimetric Determination of Traces of Metals. Interscience, New


York, N.Y., 3rd ed., 1032 pp.
Sandell, E.B. and Onishi, H., 1978. Colorimetric Determination of Metals, Part 1. Photo-
metric Determination of Traces of Metals: General Aspects. Wiley, New York, N.Y.,
4th ed., 1085 pp.
Scheide, E.P. and Durst, R.A., 1977. Indirect determination of sulphate in natural waters
by ion-selective electrode. Anal. Lett., 10: 55—65.
Schnepfe, M.M., 1973. Spectrophotometric procedure using Rhodamine B for determina-
tion of submicrogram quantities of antimony in rocks. Talanta, 20: 175—184.
Schuller, R.M., Huntsman, B.E., Warner, B.J. and Malone, P.G., 1975. A rapid on-site
electrode technique for determining copper in soil for geochemical exploration. Econ.
Geol, 70: 1328 (abstract).
Schwartz, M.O. and Friedrich, G.H., 1973. Secondary dispersion patterns of fluoride in
Osor area, Province of Gerona, Spain. J. Geochem. Explor., 2: 103—114.
Scott, R.H. and Kokot, M.L., 1975. Application of inductively coupled plasmas to the
analysis of geochemical samples. Anal. Chim. Ada, 75: 257—270.
Scott, R.H., Fassel, V.A., Knisley, R.N. and Nixon, D.E., 1974. Inductively coupled
plasma-optical emission analytical spectrometry: a compact facility for trace analysis
of solutions. Anal. Chem., 46: 75—80.
Scott, R.H., Strasheim, A. and Kokot, M.L., 1976. The determination of uranium in
rocks L>y inductively coupled plasma-optical emission spectroscopy. Anal. Chim. Acta,
82: 67—V7.
Scott, R.O. and Ure, A.M., 1972. Some sources of contamination in trace analysis. Proc.
Soc. Anal. Chem., 9: 288—293.
Scott, R.O., Burridge, J.L. and Mitchell, R.L., 1969. Geochemical analysis with a multi-
channel direct reader employing direct current arc excitation. Bol. Geol. Miner., 85:
56-61.
Sekerka, I. and Lechner, J.F., 1973. Some characteristics of ion-selective electrodes.
Environ. Can., Inland Waters Dir. Tech. Bull., 72: 6 pp.
Selmer-Olsen, A.R. and 0ien, A., 1973. Determination of chloride in aqueous soil
extracts and water samples by means of a chloride-selective electrode. Analyst, 98:
412-415.
Sen Gupta, J.G., 1976. Determination of lanthanides and yttrium in rocks and minerals
by atomic absorption and flame emission spectrometry. Talanta, 23: 343—348.
236

Sherman, G.D., McHargue, J.S. and Hodgkiss, W.S., 1942. Determination of active
manganese in soil. Soil Sei., 54: 253—257.
Sholkovitz, EJML, 1976. Flocculation of dissolved organic and inorganic matter during the
mixing of river water and seawater. Geochim. Cosmochim. Ada, 40: 831—848.
Sierra, J. and Leon, J.M., 1967. Mechanical apparatus for fusing samples in the determina-
tion of tin in geochemical surveys. Inst. Min. Metall, Trans., Sect. B, 76: 210—211.
Sighinolfi, G.P., 1972. Determination of beryllium in standard rock samples by flameless
atomic absorption spectroscopy. At. Absorpt. Newsl., 1 1 : 96.
Sinclair, A.J., Fletcher, W.K., Price, B.J., Bentzen, A. and Wong, S.S., 1977. Minor ele-
ments in pyrites from some porphyry-type deposits, British Columbia. Trans. Soc.
Min. Eng. AIME, 262: 9 4 - 1 0 0 .
Smee, B.W., Hall, G.E.M. and Koop, D.J., 1978. Analysis of fluoride, chloride, nitrate and
sulfate in natural waters using ion chromatography. J. Geochem. Explor., 10: 245—
258.
Smith, A.E., 1975. Interferences in the determination of elements that form volatile
hydrides with sodium borohydride using atomic-absorption spectrophotometry and
the argon-hydrogen flame. Analyst, 100: 300—306.
Smith, A.Y., 1964. Field and laboratory methods used by the Geological Survey of
Canada in geochemical surveys, 5. Cold extractable "heavy metal" in soil and alluvium.
Geol. Surv. Can., Paper, 63-49: 9 pp.
Smith, A.Y., 1967. Field and laboratory methods used by the Geological Survey of
Canada in geochemical surveys, 9. Tin in soils and stream sediments. Geol. Surv. Can.,
Paper, 67-50: 11 pp.
Smith, A.Y. and Lynch, J.J., 1969. Field and laboratory methods used by the Geological
Survey of Canada in geochemical surveys, 11. Uranium in soil, stream sediment and
water. Geol. Surv. Can., Paper, 69-40: 9 pp.
Smith, A.Y. and Washington, R.A., 1962. A mobile chemical laboratory for trace-ele-
ment analysis. Geol. Surv. Can., Paper, 62-21: 11 pp.
Smith, M.J. and Manaham, S.E., 1973. Copper determination in water by standard addi-
tion potentiometry. Anal. Chem., 45: 836—839.
Sorensen, R.C., Oelsligle, D.D. and Knudsen, D., 1971. Extraction of Zn, Fe and Mn
from soils with 0.1 N hydrochloric acid as affected by soil properties, solution: soil
ratio and length of extraction period. Soil Sei., I l l : 352—359.
Stanton, R.E., 1962. The field determination of zinc with dithizone in the presence of
aluminium. Geochem. Prospect. Res. Centre, Tech. Comm., 19: 9 pp.
Stanton, R.E., 1964. The field determination of arsenic in soils and sediments. Econ.
Geol., 59: 1 5 9 9 - 1 6 0 2 .
Stanton, R.E., 1966. Rapid Methods of Trace Analysis for Geochemical Applications.
Edward Arnold, London, 103 pp.
Stanton, R.E., 1970a. The colorimetric determination of molybdenum in soils, sediments,
and rocks by zinc dithiol. Proc. Australas. Inst. Min. Metall., 235: 101—102.
Stanton, R.E., 1970b. The colorimetric determination of tungsten in soils, sediment and
rocks by zinc dithiol. Proc. Australas. Inst. Min. Metall., 236: 59—60.
Stanton, R.E., 1971a. The determination of bismuth in soils, sediments, and rocks. Proc.
Australas. Inst. Min. Metall., 240: 113—114.
Stanton, R.E., 1971b. The colorimetric determination of thorium in soils, sediments and
rocks. Proc. Australas. Inst. Min. Metall., 239: 101—103.
Stanton, R.E., 1975. The colorimetric determination of palladium and platinum. Lab.
Pract., 24: 5 2 5 - 5 2 7 .
Stanton, R.E., 1976. Analytical Methods Used in Geochemical Exploration. Edward
Arnold, London, 54 pp.
Stanton, R.E. and Coope, J.A., 1958-59. Modified field test for the determination of
small amounts of nickel in soils and rocks. Inst. Min. Metall, Trans., Sect. B, 68: 9—
14.
237

Stanton, R.E. and Hardwick, A.J., 1967. The colorimetric determination of molybdenum
in soils and sediments by zinc dithiol. Analyst, 92: 387—390.
Stanton, R.E. and Härdwick, A.J., 1971. The colorimetric determination of vanadium in
soils, sediments, and rocks for use in geochemical exploration. Proc. Australas. Inst.
Min. Metall, 240: 1 1 3 - 1 1 4 .
Stanton, R.E. and McDonald, A.J., 1961-62a. Field determination of cobalt in soil and
sediment samples. Inst. Min. Metall., Trans., Sect. B, 71: 511—516.
Stanton, R.E. and McDonald, A.J., 1961-62b. Field determination of tin in geochemical
soil and stream sediment surveys. Inst. Min. Metall., Trans., Sect. B, 7 1 : 27—29.
Stanton, R.E. and McDonald, A.J., 1961-62c. Field determination of antimony in soil
and sediment samples. Inst. Min. Metall., Trans., Sect. B, 71: 517—522.
Stanton, R.E. and McDonald, A.J., 1962. The spectrophotometric determination of anti-
mony using brilliant green. Analyst, 87: 299—301.
Stanton, R.E. and McDonald, A.J., 1964. The determination of gold in soil with brilliant
green. Analyst, 89: 767—770.
Stanton, R.E. and McDonald, A.J., 1965. The determination of selenium in soils and sedi-
ments with 3,3'-diaminobenzidene. Analyst, 90: 497—499.
Stanton, R.E. and McDonald, A.J., 1966. The colorimetric determination of boron in
soils, sediments and rocks with methylene blue. Analyst, 9 1 : 775—778.
Stanton, R.E. and Ramankutti, S., 1977. The determination of platinum by atomic
absorption spectrophotometry. J. Geochem. Explor., 7: 73—76.
Stanton, R.E., Mockler, M. and Newton, S., 1973. The colorimetric determination of
molybdenum in organic-rich soil. J. Geochem. Explor., 2: 37—40.
Stary, J., 1964. The Solvent Extraction of Metal Chelates. MacMillan, New York, N.Y.,
240 pp.
Steger, H.F., 1976. Chemical phase-analysis of ores and rocks. Talanta, 23: 81—87.
Stubbs, M.F., 1968. Modification of dithiol tests for the field determination of molyb-
denum. Analyst, 9 3 : 59—60.
Sutcliffe, P., 1976. Determination of molybdenum in geological materials by atomic-
absorption spectrophotometry. Analyst, 1 0 1 : 949—955.
Szabo, N.L., Govett, G.J.S. and Lajtai, E.Z., 1975. Dispersion trends of elements and
indicator pebbles in glacial till around Mt. Pleasant, New Brunswick. Can. J. Earth
Sei., 12: 1 5 3 4 - 1 5 5 6 .

Tait, B.A.R. and Coats, J.S., 1976. A rapid method for the analysis of trace elements in
rocks, stream sediments and soil by photographic emission spectrography using a semi-
automated method of plate evaluation. Rep. Inst. Geol. Sei., 76-11: 30 pp.
Tamm, O., 1922. Eine Methode zur Bestimmung der anorganischen Komponenten des
Gelkomplexes im Boden. Medd. Statens Skogsförskningsinst. (Sweden), 19: 387—404.
Tennant, W.C., 1967. General spectrographic technique for the determination of volatile
trace elements in silicates. Appl. Spectrosc, 2 1 : 282—285.
Tennant, W.C. and Sewell, J.R., 1969. Direct reading spectrochemical determination of
trace elements in silicates incorporating automatic background correction. Geochim.
Cosmochim. Ada, 33: 640—645.
Terashima, S., 1973. Atomic absorption analysis of beryllium and vanadium in rocks.
Jpn. Anal, 22: 1 3 1 7 - 1 3 2 3 .
Terashima, S., 1976. The determination of arsenic in rocks, sediments and minerals by
arsine generation and atomic absorption spectrometry. Anal. Chim. Ada, 86: 43—51.
Thompson, A.J. and Thoresby, P.A., 1977. Determination of arsenic in soil and plant
materials by atomic-absorption spectrophotometry with electrothermal atomization.
Analyst, 102: 9—16.
Thompson, C.E., 1967. A spectrophotometric method for the determination of traces of
platinum and palladium in geologic materials. U.S. Geol. Surv., Prof. Paper, 575-D:
236-238.
238

Thompson, C.E. and Lakin, H.W., 1 9 5 7 . A field Chromatographie m e t h o d for determina-


tion of uranium in soils and rocks. U.S. Geol. Surv., Bull, 1036-L: 209—220.
Thompson, C.E., Nakagawa, H.M. and VanSickle, G.H., 1 9 6 8 . Rapid analysis for gold in
geologic materials. U.S. Geol. Surv., Prof. Paper, 600-B: 130—132.
T h o m p s o n , G. and Bankston, D.C., 1 9 7 0 . Sample contamination from grinding and
sieving determined by emission spectrometry. Appl. Spedrosc, 2 4 : 210—219.
T h o m p s o n , K.C. and Thomerson, D.R., 1 9 7 4 . Atomic-absorption studies on the deter-
mination of antimony, arsenic, bismuth, germanium, lead, selenium, tellurium and tin
by utilizing the generation of covalent hydrides. Analyst, 9 9 : 5 9 5 — 6 0 1 .
T h o m p s o n , M. and Howarth, R.J., 1 9 7 3 . The rapid determination and control of preci-
sion by duplicate determinations. Analyst, 9 8 : 153—160.
Thompson, M. and Howarth, R.J., 1 9 7 6 . Duplicate analysis in geochemical practice, 1.
Theoretical approach and estimation of analytical reproducibility. Analyst, 1 0 1 : 690—
698.
Thompson, M. and Howarth, R.J., 1 9 7 8 . A new approach to the estimation of analytical
precision. J. Geochem. Explor., 9: 23—30.
Thompson, M., Pahlavanpour, B., Walton, S.J. and Kirkbright, G.F., 1978a. Simultaneous
determination of trace concentrations of arsenic, antimony, bismuth, selenium and
tellurium in aqueous solution by introduction of the gaseous hydrides into an induc-
tively coupled plasma source for emission spectrometry, I. Preliminary studies. Analyst,
103: 5 6 8 - 5 7 9 .
Thompson, M., Pahlavanpour, B., Walton, S.J. and Kirkbright, G.F., 1 9 7 8 b . Simultaneous
determination of trace concentrations of arsenic, antimony, bismuth, selenium and
tellurium in aqueous solution by introduction of the gaseous hydrides into an induc-
tively coupled plasma source for emission spectrometry, II. Interference Studies.
Analyst, 1 0 3 : 7 0 5 - 7 1 3 .
Timperley, M.H., 1 9 7 4 . Direct-reading d.c. arc spectrometry for rapid geochemical
surveys. Spectrochim. Ada, 2 9 B : 95—110.
Timperley, M.H. and Allan, R.J., 1 9 7 4 . The formation and detection of metal dispersion
halos in organic lake sediments. J. Geochem. Explor., 3 : 167—190.
Tindall, F.M., 1 9 6 5 . Silver and gold assay by atomic absorption spectrophotometry. At.
Absorpt. Newsi, 4 : 339—340.

Ure, A.M., 1 9 7 5 . The determination of mercury by non-flame atomic absorption and


fluorescence spectrometry — a review. Anal. Chim. Ada, 7 6 : 1—27.
Ure, A.M. and Shand, C.A., 1 9 7 4 . The determination of mercury in soils and related
materials by cold-vapour atomic absorption spectrometry. Anal. Chim. Ada, 7 2 : 63—
77.

Van Loon, J.C., 1 9 7 1 . Laboratory construction and laboratory and field evaluation of
thermoplastic chloride-selective electrodes with liquid filling solution and solid-solid
connections. Anal. Chim. Ada, 5 4 : 23—28.
Van Loon, J.C., 1 9 7 2 . The determination of small amounts of m o l y b d e n u m in soils by
atomic absorption spectrophotometry for geochemical exploration purposes. At.
Absorpt. Newsi, 1 1 : 60—62.
Van Loon, J.C., Kesler, S.E. and Moore, C M . , 1 9 7 3 . Analysis of water-extractable chlo-
ride in rocks by use of a selective ion electrode. In: M.J. Jones (Editor), Geochemical
Exploration 1972. Institution of Mining and Metallurgy, London, pp. 429—434.
Varley, J.A. and Chin, P.Y., 1 9 7 0 . Determination of water soluble sulphate in acidic sul-
phate soils by atomic-absorption spectroscopy. Analyst, 9 5 : 592—595.
Vaughn, W.W., 1 9 6 7 . A simple mercury vapour detector for geochemical prospecting.
U.S. Geol. Surv., Circ, 5 4 0 : 8 p p .
239

Vaughn, W.W. and McCarthy, J.H., 1964. An instrumental technique for the determina-
tion of submicrogram concentrations of mercury in soils, rocks and gas. U.S. Geol.
Surv., Prof. Paper, 501-D: 1 2 3 - 1 2 7 .
Viets, J.G., 1978. Determination of silver, bismuth, cadmium, copper, lead and zinc in
geologic materials by atomic absorption spectrometry with tricaprylyl methyl am-
monium chloride. Anal. Chem., 50: 1097—1101.
Vijan, P.N., Rayner, A.C., Sturgis, D. and Wood, G.R., 1976. A semi-automated method
for the determination of arsenic in soil and vegetation by gas-phase sampling and
atomic absorption spectrometry. Anal. Chim. Ada, 82: 329—336.
Von Alfthan, C , Rautala, P. and Rhodes, J.R., 1980. Applications of a new multi-
element portable X-ray spectrometer to materials analysis. In: C.S. Barrett, D.E.
Leyden, J.B. Newkirk, P.K. Predecti and C O . Ruud (Editors), Advances in X-ray
Analysis, Vol. 23. Plenum, New York, N.Y., pp. 27—36.

Walsh, A., 1955. The application of atomic absorption spectra to chemical analysis. Spec-
trochim. Ada, 7: 108—117.
Walthall, F.G., 1973. Spectrochemical computer analysis-program description. U.S. Geol.
Surv., J. Res., 2: 61—71.
Ward, F.N., 1951a. Determination of molybdenum in soils and rocks. Anal. Chem., 23:
788.
Ward, F.N., 1951b. A field method for the determination of tungsten in soils. U.S. Geol.
Surv., Circ., 119: 4 pp.
Ward, F.N. (Editor), 1975. New and Refined Methods of Trace Analysis Useful in Geo-
chemical Exploration. U.S. Geol. Surv., Bull., 1408: 105 pp.
Ward, F.N. and Bailey, E.H., 1960. Camp- and sample-site determination of traces of mer-
cury in soils and rocks. Trans. AIME, 217: 343—350.
Ward, F.N. and Bondar, W.F., 1979. Analytical methodology in the search for metallic
ores. In: P.J. Hood (Editor), Geophysics and Geochemistry in the Search for Metallic
Ores. Geol. Surv. Can., Econ. Geol. Rep., 3 1 : 365—383.
Ward, F.N. and Crowe, H.E., 1956. Colorimetric determinations of traces of bismuth in
rocks. U.S. Geol. Surv., Bull, 1036-1: 1 7 3 - 1 7 9 .
Ward, F.N. and Lakin, H.W., 1954. Determination of traces of antimony in soils and
rocks. Anal. Chem., 26: 1168—1173.
Ward, F.N. and Marranzino, A.P., 1955. Field determination of microgram quantities of
niobium in rocks. Anal. Chem., 27: 1325—1328.
Ward, F.N., Lakin, H.W., Canney, F.C. et al., 1963. Analytical Methods Used in Geo-
chemical Exploration by the U.S. Geological Survey. U.S. Geol. Surv., Bull., 1152:
100 pp.
Ward, F.N., Nakagawa, H.M., Harms, T.F. and VanSickle, G.H., 1969. Atomic-Absorption
Methods of Analyses Useful in Geochemical Exploration. U.S. Geol. Surv., Bull.,
1289: 45 pp.
Warnant, P., Martin, H. and Herbillion, A.J., 1980. Kinetics of the selective extraction of
iron oxides in geochemical samples — association between Fe and Cu in acid brown
soils. Abstr. 8th Int. Geochem. Explor. Symp., Hanover, 1980.
Warren, H.V. and Delavault, R.E., 1958. Rubeanic acid field test for copper in soils and
sediments. Min. Eng., 10: 1186—1188.
Warren, H.V. and Delavault, R.E., 1959a. Readily extractable copper in eruptive rocks as
a guide for prospecting. Econ. Geol, 54: 1291—1297.
Warren, H.V. and Delavault, R.E., 1959b. Aqua regia extractable copper and zinc in
plutonic rocks in relation to ore deposits. Inst. Min. Metall, Trans., Sect. B, 69: 495—
504.
Warren, H.V., Delavault, R.E., Peterson, G.R. and Fletcher, K., 1971. The copper, zinc
and lead content of trout livers as an aid in the search for favourable areas to prospect.
240

In: R.W. Boyle (Editor), Geochemical Exploration. Can. Inst. Min. Metall., Spec. Vol.,
11: 4 4 4 - 4 5 0 .
Watanabe, H., Berman, S. and Rüssel, D.S., 1972. Determination of trace metals in water
using X-ray fluorescence spectrometry. Talanta, 19: 1363—1375.
Watling, R.J., Davis, G.R. and Meyer, W.T., 1973. Trace identification of mercury com-
pounds as a guide to sulphide mineralization at Keel, Eire. In: M.J. Jones (Editor),
Geochemical Exploration 1972. Institution of Mining and Metallurgy, London,
pp. 59—69.
Wauchope, R.D., 1976. Atomic absorption determination of trace quantities of arsenic:
application of a rapid arsine generation technique to soil, water and plant samples.
At. Absorpt. NewsL, 15: 64—67.
Webber, G.R. and Newburry, M.L., 1971. X-ray fluorescence determination of minor
and trace elements in silicate rocks. Can. Spectrosc, 16: 3—7.
Weissberg, B.G., 1971. Determination of mercury in soils by flameless atomic absorption
spectrometry. Econ. Geol., 66: 1042—1047.
Welsh, E.P. and Chao, T.T., 1975. Determination of trace amounts of antimony in geolog-
ical materials by atomic absorption spectrometry. Anal. Chim. Acta, 76: 65—69.
Welsh, E.P. and Chao, T.T., 1976. Determination of trace amounts of tin in geological
materials by atomic absorption spectrometry. Anal. Chim. Ada, 82: 337—342.
Wendt, R.H. and Fassel, V.A., 1965. Induction-coupled plasma spectrometric excitation
source. Anal. Chem., 37: 920—922.
Wenrich-Verbeek, K.J., 1976. Water and stream-sediment sampling techniques for use in
uranium exploration. U.S. Geol. Surv., Open-file Rep., 76-77: 30 pp.
Wenrich-Verbeek, K.J., 1977. Uranium and coexisting element behaviour in surface
waters and associated sediments with varied sampling techniques for uranium explora-
tion. J. Geochem. Explor., 8: 337—355.
Whitehead, N.E. and Brooks, R.R., 1969. A comparative evaluation of scintillometric,
geochemical and biogeochemical methods of prospecting for uranium. Econ. Geol.,
64: 5 0 - 5 6 .
Whitney, R.G. and Risby, T.H., 1975. Selected Methods in the Determination of First
Row Transition Metals in Natural Fresh Water. Pennsylvania State University Press,
University Park, Pa., 94 pp.
Willis, J.B., 1975. Atomic absorption spectrometric analyses by direct introduction of
powders into the flame. Anal. Chem., 47: 1752—1758.
Windham, R.L., 1972. Simple device for compensation of broadband absorption inter-
ference in flameless atomic-absorption determination of mercury. Anal. Chem., 44:
1334-1336.
Woldseth, R., 1973. X-ray Energy Spectrometry. Kevex Corporation, Burlingame, Calif.
Wolfe, W.J., 1975. Zinc abundance in Early Precambrian volcanic rocks: its relationship
to exploitable levels of zinc in sulphide deposits of volcanic-exhalative origin. In: I.L.
Elliott and W.K. Fletcher (Editors), Geochemical Exploration 1974. Elsevier, Amster-
dam, pp. 261—278.
Wood, G.A., 1959. A rapid method for the determination of small amounts of tin in soils.
In: Symposium de Exploracion Geoquimica, International Geological Congress, XX
Session, Mexico 1956, pp. 461—474.
Wood, G.A. and Stanton, R,E., 1956-57. A rapid method for the determination of
chromium in soils for use in geochemical prospecting. Inst. Min. Metall., Trans.,
Sect.B, 66: 3 3 1 - 3 4 0 .
SUBJECT INDEX

Abundances of trace elements, 2, 1 1 3 —, location of laboratory, 17—20


Accuracy, 25, 2 6 , 45—46 —, reliability and reproducibility, 15
definition, 26 —, sample throughput, 2, 15—17, 18,
for lunar samples, 4 6 19,
reference materials, 45 —, sensitivity, 10—13
Acid decompositions, 6, 9, 57, 58 geochemical considerations, 5—10
acidified aluminium chloride, 6 5 , 8 2 , Analysis of variance, 16, 32—34
84,208 Anodic stripping voltametry, 1 9 9
acidified beryllium nitrate, 6 5 , 8 2 , 8 4 , Anomaly contrast, 1, 7, 9, 2 6 , 5 0 , 7 1 , 77,
208 78
aqua regia, 6 2 , 7 1 , 7 2 , 7 3 , 1 2 3 Antimony
efficiency of, 9, 5 9 , 6 0 , 61 abundance, 1 1 3
equipment for, 5 9 , 60 atomic absorption, 1 1 , 1 1 8 , 1 1 9 , 1 2 1 ,
hydrobromic acid/bromine, 6 4 , 6 5 , 132, 133, 134
71, 123 —, hydride generation, 1 1 , 1 3 2 , 1 3 3 ,
hydrochloric acid, 1 1 , 58, 5 9 , 6 1 , 134
62-63, 65,79,80,82,83,88 —, interferences, 134
hydrofluoric acid, 5 8 , 5 9 , 6 4 , 1 0 5 —, solvent extraction for, 121
nitric acid, 9, 1 1 , 5 8 , 5 9 , 6 0 , 6 1 , colorimetry, 1 0 3
6 2 - 6 4 , 86 decompositions for, 67, 68, 1 3 3
nitric acid/hydrochloric acid 5 9 , 62— emission spectroscopy, 1 2 , 1 3 9 , 1 4 6 ,
63 149,165
nitric acid/sulphuric acid/potassium standard, 2 1 4
permanganate, 6 4 , 135 APDC, 1 2 4 , 125
perchloric acid, 4, 2 3 , 5 8 , 5 9 , 63—64, Aqua regia, 6 2 , 7 1 , 7 2 , 7 3 , 1 2 3
85, 88, 134, 163, 164 Arsenic
Acids, properties of, 59 abundance, 2, 8 7 , 1 1 3
Activity coefficient, 2 0 1 , 2 0 2 , 2 0 3 atomic absorption, 1 1 , 1 1 3 , 1 1 4 , 117,
Adsorption, 9, 76, 78 1 1 8 , 1 1 9 , 1 2 0 , 1 3 2 , 1 3 3 , 1 3 4 , 138
Alkalies —, hydride generation, 1 1 , 1 3 2 , 1 3 3 ,
flame emission 1 4 2 134
ionization 117 —, interferences, 1 3 4
Aluminium, 3, 7, 5 4 , 1 8 6 , 2 1 4 colorimetry, 1 0 2 , 1 0 4 , 105
Amphiboles, 6 0 decompositions for, 6 3 , 67, 7 1 , 1 3 3 ,
Analysis, choice of methods, 5—21 134
analytical and organizational consider- —, loss of, 6 3 , 71
ations, 10—21 emission spectroscopy, 1 1 , 1 4 9 , 165
—, costs, 2, 18, 20 in manganese oxides, 87
—, interferences, 14—15, 1 8 , 19 in soils, 2
242

in waters, 54 trace elements


mobility, 7 Barium
partial extraction, 80 abundance, 2, 113
standard, 214 atomic absorption, 113, 117, 122,
X-ray fluorescence, 168, 196 128,131
Arsenopyrite, 72 —, interferences, 126, 128
Arsine, 63, 104, 105, 132, 134 emission spectroscopy, 149, 154, 160,
Atomic absorption analysis, 4, 5, 6, 11, 213
12, 16, 18, 5 8 , 8 2 in waters, 54
absorbance, definition, 111 standard, 214
history, 109 X-ray fluorescence, 168, 189, 196,
indirect methods, 136, 138 197
instruments, 112—120 Beer's law, 99, 100, 110
—, design, 114 Beryl, 68
—, light sources, 114—116 Beryllium
—, production of atomic vapour, 116— abundance, 2, 113
118 atomic absorption, 113, 121, 122, 126
—, wavelength selection, 120 —, interferences, 126
—, detection and readout, 120 —, solvent extraction for, 121
for mercury, 134—136 colorimetry, 102, 211
sensitivity, definition, 112 decompositions for, 68
theory, 1 0 2 - 1 1 2 standards, 214
—, absorption of light, 110, 111 Binomial distribution, 36
—, excitation of atoms, 109, 110, 111, Biotite, 59
112 Bismuth
with Oames, 110, 1 1 6 - 1 1 8 , 1 2 2 - 1 3 0 abundance, 113
^interferences, 14, 125—130, 131, atomic absorption, 11, 113, 121, 124,
134, 135, 136, 162 131, 132, 134
—, operating conditions, 115, 116, —, hydride generation, 132, 134
117,122 —, interferences, 134
—, sensitivity, 11, 12, 113, 115, 122— —, solvent extraction for, 121, 124
125 colorimetry, 102, 104
—, solvent extraction for, 65, 121, decomposition for, 62, 67
123,124, 125 emission spectroscopy, 12, 149, 165
with furnaces, 12, 113, 1 1 8 - 1 1 9 , polarography, 199
131-132 standards, 214
with hydride generation, 119, 132— X-ray fluorescence, 168, 197
134 Bloom test, 101
Atomic absorption interferences, 14 Boehmite, 89
in flames, 125—130 Boltzmann constant, 110
—, background absorption, 128—130 Boron
—, chemical, 127, 162 abundance, 2, 113
—, ionization, 127—128, 162 atomic absorption, 113
—, mercury, 135, 136 colorimetry, 64, 102, 211
—, spectral, 126—127 decomposition for, 64
in furnaces, 131 emission spectroscopy, 149, 150, 154
in hydride generation, 134 in waters, 54
Auger effect, 170, 186 specific ion electrode, 205, 211
Awaruite, 72 standard, 214
Bragg'slaw, 176
Background, 1, 9; see also Abundances of Bremstrahlung, 175
243

Cadmium in water, 1 2 4
abundance, 1 1 3 mobility, 7
atomic absorption, 1 1 3 , 1 2 3 , 1 2 4 , sampling errors, 4 0
125, 1 2 6 , 1 3 1 , 132 standard, 2 1 4
—, interferences, 1 2 6 , 131 X-ray fluorescence, 1 9 4 , 1 9 6
—, solvent extraction for, 1 2 3 , 125 Cinnabar, 7 2
colorimetry, 1 0 4 Clays, 7 5 , 7 6 , 77, 8 2 , 8 4
decomposition for, 62 Cobalt
emission spectroscopy, 1 1 , 1 4 9 abundance, 2, 1 1 3
in water, 1 2 4 , 125 atomic absorption, 1 1 3 , 1 1 8 , 1 2 1 ,
mobility, 7 1 2 2 , 1 2 4 , 1 2 5 , 129, 131
polarography, 1 9 9 —, interferences, 1 2 9 , 131
standard, 2 1 4 —, solvent extraction for, 1 2 4 , 125
X-ray fluorescence, 197 colorimetry, 1 0 2 , 1 0 4
Calcium decomposition for, 5 9 , 6 0 , 6 2 , 6 6 , 67
abundance, 3 emission spectroscopy, 1 4 9 , 1 5 4 , 1 6 3 ,
atomic absorption, 127 164
emission spectroscopy, 141 extraction with ascorbic acid/hydro-
in volcanic rocks, 75 gen peroxide, 4 6 , 74
in waters, 54 in humic material, 78
mobility, 7 in igneous rocks, 59
standard, 2 1 4 in laterite, 60
water-soluble, 75 in manganese oxides, 87
Calomel electrode, 2 0 0 in rock-forming minerals, 5 9 , 6 6 , 74
Carbonates, 7 8 , 8 2 , 131 in silicate lattices, 5 9 , 74
Cassiterite, 3 6 , 37, 3 8 , 3 9 , 5 1 , 67 in sulphides, 74
decomposition for, 67 in ultramafic standards, 4 6 , 74
sampling problems, 36, 37, 3 8 , 39 in water, 1 2 4
Census mobility, 7
analytical methods used, 2 0 standard, 2 1 4
materials sampled, 2 X-ray fluorescence, 1 7 2 , 1 9 6
Chalcopyrite, 7 2 , 7 3 Cold extractions, 57, 7 1 , 9 4 ; see also
Chloride Partial extractions
ion chromatography, 1 9 9 Colorimetry, 4, 5, 6, 16, 18, 67, 7 4 , 80
mobility, 7 Bloom total heavy metal test, 8 1 , 101
specific ion electrode, 2 0 5 , 2 0 7 , 2 0 9 determination
standard, 2 1 4 —, arsenic, 1 0 4 , 105
Chlorine, 6 9 , 7 4 , 7 5 , 2 0 9 - , lead, 1 0 4 , 2 1 3
Chromatography —, niobium, 1 0 5 , 1 0 6
paper, 1 0 4 , 1 0 5 , 1 0 6 —, uranium, 1 0 5 , 1 0 6
--, for niobium, 1 0 5 , 1 0 6 paper chromatography, 1 0 4 , 1 0 5 , 106
—, for uranium, 1 0 5 , 1 0 6 principles, 97, 9 8 , 99
ion, 199 reagents, 9 9 , 1 0 0 , 1 0 2 , 1 0 3
Chromium —, dithizone, 8 1 , 8 6 , 1 0 0 , 1 0 1 , 1 0 4 ,
abundance, 2, 1 1 3 213
atomic absorption, 1 1 3 , 1 2 2 , 1 2 4 , 126 standards, 9 9 , 1 0 0 , 2 1 3
—, interferences, 126 Conductivity, 199
—, solvent extraction for, 1 2 4 Contamination, 17, 4 1 , 4 2 , 4 3 , 4 4 , 4 7 ,
colorimetry, 1 0 2 48, 5 1 , 5 4 , 58, 210, 219
decomposition for, 6 0 , 69 chloride, 2 0 9
emission spectroscopy, 1 4 9 , 1 5 4 , 1 5 9 , control graphs, 4 3 , 4 4
160 from grinding, 48
244

from laboratory materials, 47 —, barium chloride, 77


from perspiration, 209 —, bromine, 46
from reference electrodes, 201 —, buffer solutions, 59, 77
from sample carryover, 42, 48 —, dilute acids, 11, 77, 80, 85, 87, 88
from sieves, 51 —, ethylenediaminetetracetic acid, 4,
of spectroscopic bases, 150 86,87,88
of waters, 54 —, influence of pH, 77, 78
mercury, 51 —, potassium chlorate/hydrochloric
reagent blanks, 43 acid, 72, 73, 74
systematic errors, 42 —, reaction kinetics, 79, 80
Control graphs, 2 1 , 28, 29, 30, 31, 43, 44 —, sodium hypochlorite, 85
Copper sequential extraction, 93
abundance, 2, 113 specific ion electrode, 205, 210
atomic absorption, 16, 113, 118, 121, standard, 214
122, 124, 125,129, 131,210 X-ray fluorescence, 169, 172, 178,
—, interferences, 129, 131 197
—, solvent extraction for, 124, 125 Crushing, 23, 48, 49
colorimetry, 74, 100, 102, 104 Cyanogen bands, 145, 147, 156
decomposition
—, aquaregia 72, 73 Data management, 44
- , hydrochloric acid, 11, 59, 87, 88, DC-arc spectroscopy, 5, 6, 9, 11, 12, 13,
210 1 4 , 1 5 , 16, 1 9 , 9 5
—, hydrofluoric acid, 59, 60, 72, 73 direct-reading spectrometry, 17, 139,
—, nitric acid, 9, 11, 60, 62, 77, 86 152-157
—, nitric acid/perchloric acid, 59, 60, —, background, 153, 156—157
88 —, calibration, 153
—, perchloric acid, 4, 59, 60 —, matrix effects, 154—155
—, potassium bisulphate fusion, 66, 67 —, spectral interferences, 153, 157
emission spectroscopy, 16, 141, 146, internal standards, 34, 35, 154, 155
149, 1 5 0 , 1 5 4 , 1 6 3 selective volatilization, 141, 145, 148.
in drill sludge, 74 154
in granodiorite, 85 semi-quantitative, 139, 141—152
in humic materials, 78, 85, 86 —, analysis of unknowns, 151—152
in igneous rocks, 59 —, choice of lines, 146, 147, 149
in lake sediments, 86 —, equipment, 142—144
in laterite, 60 —, estimation of line intensities, 144
in peat, 76, 85, 87, 88 —, interferences, 14, 145, 146, 149
in rock-forming minerals, 59, 66, 74 —, operating conditions, 145—148
in silicate lattices, 9 —, optics, 143, 144
in soils, 77, 79, 80, 8 1 , 85, 93 —, spectroscopic buffers, 14, 145, 154,
—, extraction kinetics, 79, 80 156,157
—, influence of extraction pH, 77 —, standards, 148—151
in stream sediments, 4, 11, 79 see also Emission spectra
in sulphides, 72, 73, 74 Debey-Huckel expression, 202
in ultramafic standards, 46, 74 Decomposition techniques
in vegetation, loss on ashing, 71 choice of, 9, 57
in water, 124, 125 classification of, 57, 58
mobility, 7 objective of, 4, 5, 57
partial extraction partial, 4, 5, 9, 57, 7 1 - 9 3 , 94, 95
—, ammonium acetate, 77, 78 —, application of bedrock, 71—75
—, ascorbic acid/hydrogen peroxide, —, application to soils and sediments,
46, 74 75-93
245

—, versus total, 4, 5, 9 —, analytical lines, 146, 147, 149


strong, 9, 57, 58—71, 94 —, cyanogen bands, 145, 147, 156
—, with acids, 9, 57, 59, 66 —, prominent lines, 151
—, with fusions, 57, 66—70 in plasmas, 141
Detection limits, 12, 13, 15, 30—32, 163, theory, 109, 110, 111, 1 4 0 - 1 4 1
185,186 Emission spectroscopy, 16, 19, 20, 34, 35
estimation of, 30—32 DC-arc spectroscopy, 119, 141—157
with inductively coupled plasma, 163 —, direct-reading spectrometry, 152—
with X-ray fluorescence, 185, 186 157
see also Sensitivity —, semi-quantitative, 17, 141—152
Digestion, see Decomposition flame emission, 141—142
Dilution factor, 122 plasma sources, 157—165
Direct-reading spectrometry, see Emission see also DC-arc spectroscopy, emission
spectroscopy spectra, Spectroscopic buffers,
Disaggregation, see Sieving Spectrographs
Dispersion, 5, 6, 7, 75, see also Mobility Errors
Dissolved oxygen, 199 random, 15, 25, 26, 27—40
Dithizone, 8 1 , 86, 100, 101, 104, 213 —, in sampling, 34—40
determination of lead, 104, 213 —, precision estimation, 27—30, 31, 32
properties of, 100 —, relation to detection limit, 30, 32
reaction with metals, 101, 104 —, sources of, 32—34
—, influence of pH, 104 systematic, 15, 25, 26, 41—46, 152
Doppler effect, 112 —, control and monitoring, 42—45
Drift, 2 1 , 4 2 —, control graph, 43
Drill sludge, 74 —, from contamination, 42
Dry ashing, 70—71 —, from drift, 42
Drying, 5, 6, 51 —, from interferences, 42
—, spurious trends arising from, 15,
Electrochemical methods 43,44
anodic stripping voltametry, 199 Exchangeable ions, 9, 75, 77, 84
pH, 2 0 3 - 2 0 4 Extractions, see Acid decompositions,
polarography, 199 Fusions, Partial extractions
specific ion electrodes, 205—211
Electrodes Feldspar, 60, 75
emission spectroscopy, 143, 145, 146, Field kits, 17
147,152 Field laboratories, 17, 58
pH, 200, 2 0 1 , 2 0 3 , 204 Field tests, 6 5 , 8 1 , 95
reference, 200, 201 Fire assay, 65, 66
—, calomel, 200 Flames, 14, 110, 1 1 6 - 1 1 8 , 1 2 2 - 1 3 0 ,
—, junction potential, 200 141-142
chloride, 205, 207, 209 atomic absorption in, 116—118, 122—
cupric, 205, 210—211 130
fluorborate, 205, 211 distribution of atoms in, 117, 118
fluoride, 66, 69, 82, 199, 202, 205, emission, 12, 141—142
208-209 temperatures of, 110
iodide, 205, 2 0 9 - 2 1 0 Fluid inclusions, 74
specific ion, 205—211 Fluorimetry, 97, 98, 106—108
—, characteristics, 205 principle, 97, 98
—, interferences, 205 uranium determination, 106, 107, 108
—, response, 201, 202 Fluorine, 11
sulphide, 205 decompositions for, 65, 69, 208
Emission spectra —, lithium metaborate fusion, 69
in DC-arcs —, potassium nitrate/sodium carbonate
246

fusion, 65, 69 —, interference, 123


—, sodium carbonate fusion, 69 —, solvent extraction for, 64, 65, 121,
—, sodium hydroxide fusion, 65, 69 123
extraction, 65, 75, 82—84, 208 colorimetry, 15, 102
—, aluminium chloride, 65, 82, 84, emission spectroscopy, 111, 149
208 extraction
- , beryllium nitrate, 65, 82, 84, 208 —, with aqua regia, 62, 63, 7 1 , 123
—, ferric chloride, 82, 83 —, nitric acid/hydrochloric acid, 62—
—, hydrochloric acid, 65, 82, 83 63
—, sodium hydroxide, 65, 82 —, hydrobromic acid/bromine, 64, 65,
—, total ionic strength adjustment 71,123
buffer (TISAB), 65, 82, 83, 208 Grotrian energy level diagram, 111
—, water, 75, 208 in waters, preservation, 54
in soils, 65, 82 sampling for, 35, 39
in stream sediments, 65, 66, 82 systematic error, 41
in volcanic rocks, 75 standard, 214
ion chromatography, 199 X-ray fluorescence, 168
specific ion electrode, 66, 69, 82, 199, Gossans, 194
202, 205, 208—209 Gran's plots, 207, 209
—, interferences, 66, 205, 208 Grinding, 23, 48, 49, 50
—, response, 202 Grotrian energy level diagram, gold atom,
standard, 214 111
Fluorite, 65, 82
Fluxes, 69; see also Fusions Hafnium, 124
Furnaces, atomic absorption, 118—119, Heavy liquids, 49
131,132 Heavy metals, 80, 8 1 , 101, 102
Fusions, 6, 9, 57, 58 Bloom test, 101
alkaline, 6 9 - 7 0 , 106, 218, 219, 210 colorimetry, 101, 102
ammonium salts, 66—68, 165 dithizone field test, 81
chloride, 69, 209 Heavy minerals, 10, 12, 17, 37, 5 1 , 53,
classification, 58 75
fluorine, 69, 208 in soils and sediments, 75
iodine, 210 in till, 10, 1 7 , 5 1 , 5 3
potassium bisulphate, 66, 67 separation of, 53
potassium pyrosulphate, 63, 66, 67 see also Panned concentrates, Resis-
uranium, 106 tate minerals
Heterpoly compounds, 138
Galena, 72, 82 Hollow cathode lamps, 114, 115, 116
Gallium, 2, 149, 154 Humic acids, 85, 86, 93
Geochemical prospecting, definition, 1 Humic materials, 78, 85, 86
Germanium, 102, 132, 134, 136, 154 Hydrides, 11, 63, 104, 105, 118, 119,
atomic absorption, 132, 134, 136 121, 1 3 2 - 1 3 4 , 1 6 5
--, hydride generation, 132 arsenic, 63, 104, 105, 118, 119, 132,
—, interferences, 134 133,134
colorimetry, 102 antimony, 118, 119, 132, 133, 134
emission spectroscopy, 154 bismuth, 132, 134
Gibbsite, 89 generation of, 132—134
Goethite, 87 germanium, 132
Gold interferences, 134, 165
abundance, 113 selenium, 63, 132, 134
atomic absorption, 11, 111, 113, 121, tellurium, 132, 133, 134
123,131 tin, 132
247

Hydrobromic acid, 64, 65, 71, 123 —, interfering lines, 196


Hydrochloric acid, 11, 58, 59, 61, 62— —, matrix corrections, 189—196
63, 6 5 , 7 9 , 8 0 , 8 2 , 8 3 , 8 8 —, mineralogical, 187, 188
Hydromorphic anomalies, 6, 7, 9, 71, 78, —, textural, 187, 188, 189
94 see also under individual elements
Internal standards, 34, 35, 154
Igneous rocks, 59, 61 Involatile elements, 154
Ignition of samples, 6, 70, 71, 152 Ion chromatography, 199
Indium, 121, 123 Ion exchange, 6, 9, 11, 75, 77, 84, 168
Inductively coupled plasma, 5, 6, 12, 14, extraction of exchangeable ions, 84
19, 58, 132, 139, 157, 1 5 8 - 1 6 5 Ionization
analysis of geochemical samples, 132, alkalies, 117
163-165 in DC-arcs, 140, 141
background, 159, 160, 161, 162 interference in atomic absorption,
detection limits, 163 1 2 7 - 1 2 8 , 162
interferences, 14, 161, 162, 164, 165 Interference in plasmas, 161, 162, 164
operating conditions, 158, 159, 160 Iodine, 69, 102, 205, 2 0 9 - 2 1 0
properties of, 157, 158, 159, 160 Iron
spectra, 141, 160, 213 abundance, 3
working range, 161 atomic absorption, 124, 125, 126
Interferences, 2, 14—15, 58 decompositions for, 60, 80, 93
in fluorimetric determination of ura- in latentes, 60
nium, 7 0 , 8 2 , 106, 107 in soils, 60, 80, 93
in inductively coupled plasmas, 161, in waters, 54, 124, 125
162,164,165 mobility, 7
systematic errors caused by, 42 oxidation during grinding, 48
with atomic absorption, 123, 125— solvent extraction of, 124, 125, 126
130, 131, 134, 135, 136 standard, 214
—, background, 128—130 Iron oxides, 75, 76, 82
—, chemical, 127, 162 extraction, 58, 87—90, 92
—, determination of gold, 123 —, with ammonium oxalate, 58, 87, 89
—, electrothermal atomization, 131 —, with hydrazine, 58, 90
—, flameless determination of mer- —, with oxalic acid, 87
cury, 135, 136 —, with sodium dithionite, 58, 87, 89,
—, hydride generation, 134 90
—, ionization, 127—128, 162 stability, 77
—, spectral, 126—127 Isoformation, 155
with emission spectroscopy, 145, 146,
149, 153, 1 5 4 - 1 5 7 Junction potential, 200
—, background, 153, 156—157
—, interfering lines, 146, 149, 153, Kaolinite, 60
157 Kyanite, 60
—, matrix effects, 145, 154—155
—, spectral, 157 Lake sediments, 62, 63, 86, 90, 91
with specific ion electrodes, 205 Lanthanum, 154
—, chloride, 209 Latentes, 60
—, copper, 210 Lead
—, fluoride, 208 abundance, 2, 113
—, iodide, 210 atomic absorption, 11, 16, 113, 118,
—, sulphate, 208 121, 122, 124, 125, 126, 129, 130,
with X-ray fluorescence, 187, 188, 131
189-196 —, interferences, 126, 129, 130, 131
248

—, solvent extraction for, 124, 125 extraction of, 58, 8 1 , 84, 87, 92
colorimetric determination, 101, 102, readily reducible, 87, 89
104,213 stability, 77
decomposition for, 60, 62, 64, 67 Massive sulphides, weathering, 78
emission spectroscopy, 16, 149, 154, Mercury
160, 163,164 abundance, 113
in humics, 78 amalgamation on gold, 135, 199
in latentes, 60 atomic absorption, 11, 64, 109, 113,
in soils, 93 114, 116, 121, 1 3 4 - 1 3 7
in waters, 54, 124, 125 ~", by pyrolysis, 48, 135
mobility, 6 —, cold vapour method, 135, 136, 137
ores, 6, 190, 192 contamination during storage, 51
partial extraction, 80, 89 decompositions for, 64, 66, 67, 135,
polarography, 199 136
sequential extraction, 93 detectors, 136, 137, 199
specific ion electrode, 205, 207 electrochemical determination, 199
standard, 214 emission spectroscopy, 11
X-ray fluorescence, 197 loss
Lepidocrocite, 87 —, during drying, 51
Limonite, 59 —, during grinding, 48
Lithium —, during ashing, 71
abundance, 2 in soil gases, 199
emission spectroscopy, 149 in waters, preservation, 54
flame emission, 142 specific ion electrode, 205
in water, 54 spectrometer, 136
standard, 215 standard, 215
MIBK, 121, 123, 124, 125, 126
Mafic minerals, 75 Microwave plasma, 158
Magnesium, 3, 7, 75, 127, 186, 215 Mineral
Magnetite, 126 phases, 9
Manganese separates, 7, 9, 48
abundance, 2, 3, 113 stability, 6
atomic absorption, 16, 113, 120, 122 staining tests, 97
colorimetry, 99, 102 zoning, 74
decomposition for, 59, 60, 67 Mobility of elements, 6, 7, 75
emission spectroscopy, 16, 149, 154, Molybdenite, 73
160 Molybdenum
extraction, 80, 8 1 , 87, 9 1 , 92, 93 abundance, 2, 113
in drainage sediments, 8 1 , 87 atomic absorption, 11, 113, 117, 118,
in igneous rocks, 59 121, 123, 126, 127, 132
in lake sediments, 91 —, interferences, 126, 127
in latentes, 60 —, solvent extraction for, 121, 123
in rock-forming minerals, 59 colorimetry, 64, 69, 102
in soils, 80, 87, 93 decompositions for, 63, 64, 66, 67, 69
in waters, 54 emission spectroscopy, 145, 149, 154,
mobility, 7 159
nodules, 86, 168 in lake sediments, 91
standard, 215 mobility, 7
X-ray fluorescence, 168, 194, 196 standard, 215
Manganese oxides, 9, 58, 76, 82 X-ray fluorescence, 189, 196
accumulation of trace elements, 87 Molybdite, 62
exchangeable, 87, 89 Moseley's Law, 169
Muscovite, 60
249

Neodymium, 154 decomposition, 58, 62, 70—71, 84—88


Nernst equation, 201 —, dry ashing, 62, 70—71
Neutron activation, 12, 82, 210 —, with chelating agents, 85, 86
Nickel —, with hydrogen peroxide, 85, 92
abundance, 2, 113 —, with sodium hypochlorite, 85, 92
atomic absorption, 16, 113, 116, 118, in lake sediments, 91
120, 121, 122, 124, 125, 126, 129, mercury determination, 135
131 pH, influence on release of metals, 78
—, interferences, 126, 129, 131 Orientation survey, 7, 50
—, solvent extraction for, 124, 125 Orpiment, 72, 73
colorimetry, 102, 104
decomposition for, 22, 59, 60, 62, 66, Palladium, 65, 102, 121, 168
67 Panned concentrates, 37, 196
emission spectroscopy, 16, 149, 154, Partial extractions, 5, 6, 9, 10, 57, 58
163,164 application to bedrock, 71—75
extractions, 46, 74, 80 —, extraction of sulphides, 9, 72—74
in humic materials, 78 —, fluorine, 82—84
in igneous rocks, 59 —, non-silicate uranium, 75
in latentes, 60 —, water-soluble constituents, 74—75
in manganese oxides, 87 application to soils and sediments,
in rock-forming minerals, 59, 66 75-93
in silicate lattices, 59, 74 —, exchangeable ions, 9, 58, 77, 82, 84
in stream sediments, 2 1 , 22 —, heavy metals, 81—92
in sulphides, 74 —, iron oxides, 9, 87—90
in ultramafic standards, 46, 74 —, manganese oxides, 9, 87, 89
in waters, 124, 125 —, non-selective extractions, 9, 77,
mobility, 7 79-84
standard, 215 —, organic matter, 9, 77, 84—87
X-ray fluorescence, 171, 172, 197 - , pH, 77, 78
Niobium —, reaction kinetics, 79, 80
abundance, 113 —, response surfaces, 79
atomic absorption, 1 1 , 113, 124, 138 —, selective extractions, 9, 58, 84—90
chromatography, 64, 105, 106 —, sequential extractions, 78, 90—93
colorimetry, 102 —, uranium, 82
decomposition for, 64, 67 reagents
emission spectroscopy, 12, 139, 149 —, acetic acid, 82, 83
in stream sediments, 8 —, ammonium acetate, 58, 77, 87, 88
in kimberlites, 8 —, ammonium carbonate, 75
mobility, 7 —, ammonium citrate/hydroxylamine
standard, 215 hydrochloride, 58, 77, 79, 81
X-ray fluorescence, 196, 197 —, ammonium oxalate, 58, 87, 89, 90,
Nitrate, ion chromatography, 199 92
Nitric acid, 9, 1 1 , 58, 59, 60, 61, 6 2 - 6 4 , —, ascorbic acid, 58, 72, 73, 74
86 —, barium chloride, 77
Nitrous oxide flame, 117, 124 —, bromine, 58, 72, 74, 86
Non-silicate uranium, 72 —, buffer solutions, 79
—, chelating agents, 77, 85, 86
Olivine, 59, 74 —, complexing agents, 79
Ore minerals, 76 —, dilute acids, 79, 82
Organic carbon, 54, 63 - , EDTA, 4, 86, 8 7 , 8 8 , 94
Organic matter, 58, 62, 70—71, 75, 77, —, ferric chloride, 82, 83
78, 82, 8 4 - 8 8 , 89, 9 1 , 92, 119, —, hydrazine, 58, 87, 90, 92
135,196 - , hydrochloric acid, 11, 58, 79, 80,
250

82,83,87,88,94,210 influence on anomaly contrast, 26


—, hydrogen peroxide, 5 8 , 7 2 , 7 3 , 7 4 , size reduction, effect of, 5 0
75,77,82,83,85,86,87,92 variation with concentration, 32
—, hydroquinone, 87 Pyrite, 7 2 , 1 2 6
—, hydroxylamine hydrochloride, 5 8 , Pyrochlore, 64
77,87,92 Pyrolysis, 5 7 , 135
—, magnesium chloride, 58 Pyroxenes, 6 0 , 74
- , nitric acid, 1 1 , 7 7 , 8 2 , 8 6 Pyrrhotite, 7 2
—, oxalic acid, 8 7 , 8 9
—, potassium chlorate, 5 8 , 7 2 , 7 3 , 92 Quality control, 1 5 , 25—46
—, sodium carbonate, 7 5 , 82 programme, 4 6
—, sodium dithionite, 5 8 , 87, 8 9 , 9 0 , see also Precision, Random errors, Sys-
92 tematic errors
—, sodium hydroxide, 8 2 Quartz, 6 0 , 75
—, sodium hypochlorite, 5 8 , 77, 8 5 ,
89,92 Radiofrequency plasma, see Inductively
—, total ionic strength adjustment coupled plasma
buffer (TISAB), 8 2 , 8 3 Radium, in water, 54
—, water, 74—75 Radon, 23
Peat, 7 6 , 8 5 , 8 6 , 8 7 Random errors, 1 5 , 2 5 , 2 6 , 27—40
Pebble cards, 51 in sampling, 34—40
Pentlandite, 72 precision estimation, 27—30, 3 1 , 32
Perchloric acid, 4, 2 3 , 58, 5 9 , 6 0 , 6 1 , relation to detection limit, 3 0 , 32
6 3 - 6 4 , 85, 88, 134, 163, 164 sources of, 32—34
pH, 7 6 , 7 7 , 7 8 , 8 4 , 97, 2 0 0 - 2 0 3 Rare earths, 1 0 2 , 1 2 4
definition, 2 0 3 Reagent
determination, 203—204 blanks, 4 3 , 1 3 4
electrode, 2 0 0 , 2 1 0 colorimetric, 9 9 , 1 0 0 , 1 0 2 , 1 0 3
field determination, 97 see also Acid decompositions, Fusions,
influence on adsorption, 76 Partial extractions
influence on extraction, 77, 7 8 , 8 4 Reporting results, 21
in soils, 2 0 3 , 2 0 4 Request for analysis form, 2 2
Phase analysis, 72 Resistate minerals, 9 5 , 1 3 9
Phosphate, 7, 1 0 2 , 1 3 8 , 1 4 9 , 1 8 3 , 1 9 6 , Rhenium, 1 0 2
215 Rock
Plank's Law, 1 1 0 , 175 chemical composition, 3
Plasma, see Inductively coupled plasma partial extractions, 71—75
Platinum, 6 5 , 1 0 2 , 1 2 1 , 1 2 3 —, sulphides, 72—74
Plumbojarosite, 82 —, uranium, 75
Poisson distribution, 38, 3 9 , 4 0 —, water-soluble constituents, 74—75
Polarographic analysis, 199 sample preparation, 6, 4 7 , 48—50
Porphyry coppers, 7 2 , 7 3 strong decompositions, 5 9 , 6 0 , 6 1 , 6 2 ,
Portable X-ray fluorescence analyzers, 17, 64, 65, 66, 69, 70
1 9 , 7 4 , 1 6 8 , 1 7 3 , 1 7 4 , 178 —, fluorine, 6 5 , 6 9
Potassium 3, 7, 5 4 , 215 - , gold, 64
Powellite, 62 —, mercury, 64
Precious metals, 1 3 2 —, uranium, 6 9 , 7 0
Precision, 1 5 , 1 6 , 1 8 , 1 9 , 2 5 , 2 6 , 27—30, —, with acids, 5 9 , 6 0 , 6 1 , 6 2 , 64
31, 3 2 , 5 0 —, with fusions, 6 6 , 67, 6 9 , 7 0
calculation of, 2 7 , 2 8 , 2 9 ultramafic rock standards, 4 6 , 7 4
control graph, 2 8 , 2 9 , 30, 31 Rubidium, 2, 1 1 3 , 1 2 2 , 1 2 6 , 1 2 8 , 175
definition, 27
251

Safety precautions, 23, 63 loss, 5 1 , 73, 71


Saha equation, 140 mobility, 7
Sample preparation, 5 , 6 , 47—55 X-ray fluorescence, 168, 194, 196
contamination, 47, 48, 5 1 , 54 Sensitivity, 10—13, 112, 122—125, 139,
emission spectroscopy, 151, 152, 155 186,205
—, isoformation, 155 atomic absorption, 112, 122—125
mineral separates, 49, 5 1 , 53 —, definition of, 112
rocks, 6, 48—50 comparison of methods, 10—13
soils and sediments, 6, 50—52 specific ion electrodes, 205
tills, 53 X-ray fluorescence, 186
vegetation, 6, 52 Sequential extractions, 78, 90—93
water, 6, 52, 54, 55 Sieving, 5, 6, 23, 5 1 , 52
X-ray fluorescence, 188—189 Silicon, 3, 7, 35, 138
Sampling errors Silver
binomial distribution of, 36, 37, 38 abundance, 2, 113
estimation of, 35—40 atomic absorption, 11, 113, 121, 123,
for cassiterite, 36, 37, 38, 39, 40 1 2 4 , 1 2 5 , 126,129
influence of grain size, 35, 37, 38, 50 —, interferences, 126, 129
poisson distribution of, 38, 39, 40 —, solvent extraction for, 121, 124,
Scandium, 154 125
Secondary minerals, 75, 76 colorimetry, 102, 104
Sediments contamination, 51
element distribution in, 75, 76, 77 emission spectroscopy, 149, 154
partial extractions, 11, 75—93 in water, 124, 125
—, exchangeable ions, 84 mobility, 7
—, fluorine, 82, 84 specific ion electrode, 205
—, heavy metals, 81—82 standard, 215
—, iron oxides, 87—90 X-ray fluorescence, 197
—, manganese oxides, 87, 89 Sodium, 3, 7, 75, 141, 186, 215
—, non-selective, 4, 11, 79—84 Soils
—, organic matter, 84—87 abundances of trace elements, 2
—, selective, 84—90 element distribution in, 75, 76, 77
—, uranium, 82, 83 partial extractions, 75—93
sample preparation, 6, 50—52 —, exchangeable ions, 84
sequential extractions, 90—93 —, fluorine, 82, 83, 84
strong decompositions, 59—66 —, heavy metals, 8 1 , 82
—, fluorine, 65, 66, 69 —, iron oxides, 87—90
- , gold, 64, 65 —, manganese oxides, 87, 89
—, mercury, 64 —, non-selective, 79—84
—, uranium, 69, 70 —, organic matter, 84—87
—, with acids, 4, 11, 62, 63 - , pH, 77, 78
—, with fusions, 66—70 —, reaction kinetics, 79, 80
Seepage zone, 7 —, response surfaces, 79
Selective volatilization, 145, 148, 157 —, selective, 84—90
Selenium —, uranium, 82
abundance, 2, 113 pH determination, 203, 204
atomic absorption, 11, 113, 114, 117, sample preparation, 6, 50—52
120, 131, 132, 134 strong decompositions, 59—66
—, hydride generation, 11, 132, 134 —, fluorine, 65, 66, 69
—, interferences, 134 - , gold, 64, 65
colorimetry, 103 —, mercury, 64
decompositions for, 63, 7 1 , 134 —, uranium, 69, 70
emission spectroscopy, 11, 165 Solvent extraction, 6, 11, 64, 65, 106,
252

107,121,123,124, 125,131 Standards


atomic absorption, 121, 123, 124, exploration geochemical, 44, 45
125,126,131 for atomic absorption, 122, 214—215
APDC-MIBK, 121, 124 for colorimetry, 99, 100, 213
gold, 1 1 , 64, 65, 121, 123, 131 for emission spectroscopy, 148—151
iron, 125, 126 rocks, 45, 46, 74
reagents, 121 solutions, preparation of, 214—215
schemes, 121 sulphides, 46, 74, 134
uranium, 106, 107 Stannane, 165
waters, 124, 125 Stark effect, 112
Specific ion electrodes Statistical series, 27
boron, 205, 211 Stibine, 132, 134
characteristics, 205 Stibnite, 72, 73
chloride, 205, 207, 209 Stream sediments
copper, 205, 210—211 partial extractions, 4, 75—93
fluoride, 66, 69, 82, 199, 202, 205, —, copper, 4
208-209 —, exchangeable ions, 84
Gran's plot, 207, 209 —, fluorine, 66, 82, 83, 84
interferences, 205 —, heavy metals, 8 1 , 82
iodide, 205, 2 0 9 - 2 1 0 —, iron oxides, 87—90
method of additions, 205, 206 —, manganese oxides, 87, 89
sulphate, 207, 208 —, non-selective, 4, 79—84
sulphide, 205 —, organic matter, 84—87
theory, 200—203 —, selective, 84—90
—, activity coefficient, 201, 202 —, uranium, 82
—, junction potential, 200 sample preparation, 6, 50—52
—, Nernst equation, 201, 202 strong decompositions, 59—66
—, response, 201, 210 —, fluorine, 65, 66, 69
titrations, 207 - , gold, 64, 65
Spectral interferences —, mercury, 64
atomic absorption, 126—127, 128 —, uranium, 69, 70
emission spectroscopy, 146, 149, 153, Strontium
157, 162,164 abundance, 2, 113
X-ray fluorescence, 196 atomic absorption, 113, 122, 126, 128
Spectrographs emission spectroscopy, 139, 149, 154
direct-reading, 14, 16, 142, 152, 153, standard, 215
158 X-ray fluorescence, 168, 175, 197
mobile, 17, 143 Sublimation, 61; see also Fusion
optics, 143, 144 Sulphate
photographic, 17, 19, 142, 143 atomic absorption, 138
resolution, 140, 142 ion chromatography, 199
see also Emission spectroscopy, DC-arc specific ion electrode, 205, 207, 208
spectroscopy standard, 215
Spectrophotometry, 97, 98, 100; see also turbidimetry, 108
Colorimetry Sulphides, 9, 58, 7 1 , 7 2 - 7 4 , 76, 78, 82,
Spectroscopic buffers, 14, 145, 154, 156, 85, 134, 135,205
157 decomposition of, 58, 7 1 , 72—74
Spectroscopic sources, desirable proper- in granodiorite, 85
ties, 14,157 in porphyry copper deposits, 73
Sphalerite, 72 in Proterozoic volcanics, 74
Spurious anomalies, 43, 44 in ultramafics, 74
Staining tests, 97 massive, 78
253

standards, 4 6 , 7 4 , 1 3 4 atomic absorption, 1 2 4 , 1 3 8


Sulphur, 7, 1 0 8 , 1 8 7 ; see also Sulphate, colorimetry, 9 9 , 1 0 3
Sulphides emission spectroscopy, 1 4 9 , 1 5 4
Systematic errors, 15, 2 5 , 2 6 , 41—46, mobility, 7
152 X-ray fluorescence, 1 9 4
control and monitoring, 42—45 Topaz, 68
control graph, 4 3 Total ionic strength adjustment buffer
from contamination, 4 2 (TISAB), 6 5 , 8 2 , 8 3 , 2 0 3 , 2 0 8
from drift, 4 2 Tourmaline, 68
from interferences, 4 2 Trace element
spurious trends arising from, 1 5 , 4 3 , abundances, 1, 2, 1 1 3
44 dispersion, 5, 6, 7, 75
in mineral lattices, 6 0
Talc, 6 0 in soils and sediments, 2, 7 5 , 7 6 , 7 7 ,
Tantalum, 7, 1 0 5 , 1 1 3 78
Tape machine, 155 in waters, 1 1 , 5 2 - 5 5 , 1 2 4 , 125
Tellurium mobility, 6, 7, 75
abundance, 1 1 3 Transmittance, 9 8 , 9 9
atomic absorption, 1 1 , 1 1 3 , 1 2 1 , 1 3 1 , Tungsten
132,133,134 abundance, 1 1 3
—, hydride generation, 1 1 , 1 3 2 , 1 3 3 , atomic absorption, 1 1 , 1 1 3 , 1 2 1
134 colorimetry, 1 0 3
—, interferences, 1 3 4 decomposition for, 67, 69
—, solvent extraction, 121 emission spectroscopy, 1 2 , 1 3 9 , 1 4 9
decomposition for, 6 5 , 1 3 3 mobility, 7
X-ray fluorescence, 1 6 8 standard, 2 1 5
Tetrahedrite, 7 2 X-ray fluorescence, 197
Thallium, 1 2 1 , 1 2 3 , 1 2 5 , 1 5 4 Turbidimetry, 9 7 , 9 8 , 1 0 8
Thorium, 1 0 3 , 1 3 8 , 187
Threshold, 1 Ultramafic rock standards, 4 6 , 7 4
Till, 10, 1 7 , 5 1 , 5 3 Uranium
Tin abundance, 1 1 3
abundance, 2, 1 1 3 atomic absorption, 1 1 , 1 1 3
atomic absorption, 1 1 6 , 1 1 0 , 1 1 3 , chromatography, 1 0 5 , 1 0 6
121,132 decompositions for, 6 2 , 6 9 , 7 0 , 8 2 ,
—, hydride generation, 1 3 2 106,107
—, solvent extraction for, 1 2 1 delayed neutron counting, 12
colorimetry, 1 0 3 , 1 0 4 emission spectroscopy, 1 6 4
decomposition for, 67, 6 8 , 165 fluorimetry, 6 2 , 6 9 , 7 0 , 8 2 , 1 0 6 , 1 0 7 ,
emission spectroscopy, 1 2 , 16, 1 4 9 , 108
1 5 4 , 1 5 7 , 165 —, interferences, 7 0 , 8 2 , 1 0 6 , 107
in cassiterite, 3 6 , 37, 3 8 , 3 9 , 4 0 , 67 —, solvent extraction for, 7 0 , 1 0 6 , 107
—, sampling problems, 3 6 , 3 7 , 3 8 , 3 9 , —, laser-induced, 1 0 6 , 1 0 7 , 108
40 in bedrock, 75
in heavy minerals, 10 in ores, 75
in silicates, 67 in sediments, 6 2 , 8 2 , 8 3 , 107
in till, 10 in waters, 5 4 , 5 5 , 1 0 6 , 1 0 7 , 1 0 8
mobility, 7 —, filtration, 55
standard, 2 1 5 —, preservation, 5 5 , 1 0 8
X-ray fluorescence, 1 6 9 , 1 8 7 , 197 mobility, 7
Titanium partial extractions, 7 5 , 8 2 , 8 3
abundance, 2, 3 standard, 2 1 5
254

X-ray fluorescence, 169, 170, 182, nomenclature, 169


187,196 scatter, 172, 173, 1 9 2 - 1 9 6
X-ray fluorescence, 5, 6, 9, 12, 13, 14,
Vanadium 16,17,19,74,95
abundance, 2, 113 counting strategy, 184—186
atomic absorption, 113, 122, 126, 138 detection limits, 185, 186
colorimetry, 103 detectors, 178—181
decomposition for, 60, 67 —, gas-filled proportional counters,
emission spectroscopy, 149, 154 178-181
mobility, 7 —, escape peaks, 183, 184
standard, 215 —, resolution, 180
X-ray fluorescence, 194, 196 —, scintillation counters, 178—181
Vegetation, 2, 6, 52, 63, 70, 71 —, semi-conductor detectors, 181
dry ashing, 70—71 dispersion of X-rays, 176—178
wet ashing, 63 instrumentation, 173—181
sample preparation, 52 instrument conditions, 182—184
Volatile elements, 1 1 , 59, 154, 158 line interferences, 196
Volatilization curves, 154 matrix corrections, 189—196
—, compton scatter, 192—196
Water —, peak to background, 190—192
analysis by atomic absorption pulse height analysis, 179, 183
—, inductively coupled plasma, 163 sample preparation, 188—189
—, X-ray fluorescence, 168 textural effects, 187, 188, 189
chloride, 209 theory, 1 6 8 - 1 7 3
contamination, 54 —, excitation of X-rays, 168—171
filtration, 54, 55 —, interactions of X-rays with matter,
fluoride, 208 171-173
gold, 54 X-ray spectrometers
mercury, 54 energy-dispersive, 19, 167, 173, 174,
organic, compounds, 86 181,197
pH, 203 portable, 17, 19, 74, 168, 173, 174,
solvent extraction, 124, 125 178
storage, 54, 55 wavelength-dispersive, 19, 167, 173,
sulphate, 108 174, 1 7 6 - 1 7 8 , 1 8 2 - 1 8 4
testing kit, 97
uranium, 55, 106, 107, 108 Yttrium, 154, 196
Weathering, 60, 75, 78
Wulfenite, 62 Zeeman effect, 112, 135
Zinc
X-rays abundance, 2, 113
absorption, 171, 172, 189, 191, 193, atomic absorption, 16, 113, 118, 121,
194 124,127,131
absorption edges, 171, 172, 190, 191, —, interferences, 14, 127, 131
192,194 —, solvent extraction for, 124, 125
Auger effect, 170, 186 colorimetry, 101, 103, 104
characteristic radiation, 168, 169 contamination, 54
compton scatter, 172, 173, 192—196 decomposition for, 9, 59, 60, 6 1 , 62,
dispersion, 176—178 63, 6 6 , 6 7 , 7 2 , 7 3 , 8 6
excitation, 168—171 emission spectroscopy, 11, 16, 141,
fluorescent yield, 170 149,154, 163
interactions with matter, 171—173 in ferromanganese nodules, 86
mass absorption coefficient, 14, 171, in humic materials, 78
189,191,193 in igneous rocks, 59, 61
255

in lake sediments, 86, 87, 62, 91 partial extractions, 72, 73, 74, 80, 86,
in latentes, 60 87,91,93
in manganese oxides, 87 polarography, 199
in porphyry copper deposits, 72 stain test, 51, 52
in sediments, 63 X-ray fluorescence, 187
in silicate lattices, 9, 59 Zirconium
in soils, 78, 93 abundance, 2, 113
in sulphides, 72, 73, 74 atomic absorption, 110, 113, 124, 126
in waters, 54, 124, 125 emission spectroscopy, 139, 159
mobility, 6, 7 mobility, 7
negative anomalies, 78 X-ray fluorescence, 196

You might also like