ba4e3ddd028c72570ef868df39c9fe65
ba4e3ddd028c72570ef868df39c9fe65
ba4e3ddd028c72570ef868df39c9fe65
J. G. De Geoffroy
Consultant, Geostatistics and Mining Exploration
Eden, New South Wales, Australia
and
T. K. Wignall
California Polytechnic State University
San Luis Obispo, California
Few knowledgeable people would deny that the field of mineral exploration
is facing some difficult times in the foreseeable future. Among the woes, we
can cite a worldwide economic uneasiness reflected by sluggish and at times
widely fluctuating metal prices, global financial uncertainties, and relentless
pressures on costs despite a substantial slowing down of the rate of inflation.
Furthermore, management is forced to tum to more sophisticated and expensive
technologies and to look farther afield to more remote regions, as the better-
quality and more easily accessible ore deposits have now been revealed. This
rather gloomy outlook should persuade explorationists to cast about for a new
philosophy with which to guide mineral exploration through the challenging
decades ahead.
Once already, in the early 1960s, a call for change had been heard (Ref.
30 in Chapter 1), when it became obvious that the prospecting methods of
yesteryear, so successful in the past, could not keep up with the rapidly growing
demand for minerals of the postwar period. The answer, a massive introduction
of sophisticated geophysical and geochemical technologies backed by new geo-
logical models, proved spectacularly successful throughout the 1960s and the
1970s. But for both economic and technological reasons, the brisk pace of the
last two decades has considerably slowed down in the early 1980s, as if a new
threshold has been reached. We believe that the answer lies in a concerted effort
to make an efficient use of the technology at hand, which can be translated into
the earthy language of the man in the field as "getting more bang for the explo-
ration buck," or into the more esoteric language of management circles as "max-
imizing the efficiency of operations under budget constraints"-in other words:
optimization, which is what the new philosophy of mineral exploration should
be about.
Optimization is nothing new in itself. It has been a household word in
engineering and industry circles for quite some time. Optimization has been
instrumental in the spectacular reduction in energy demand witnessed in the past
few years, by providing a much more rational and efficient use of the resources
at hand. Unfortunately, however, the new philosophy of optimization does not
as yet seem to have made much inroad into the thinking of the mineral explo-
rationists of the 1980s. A search of the English language literature on the mineral
v
vi PREFACE
industry published during the past two decades revealed a great wealth of material
on the optimization of the production section of the business, but only precious
few journal articles on the optimization of the exploration section, although the
necessary methodology and the required computational facilities are available.
Mineral exploration, being essentially a sequential procedure, is eminently
suited for optimization, despite its inherent context of uncertainty and risk which
require the use of probabilistic methods, as opposed to the deterministic approach
that prevails in engineering. The optimization of mineral exploration as a whole
and that of the various individual stages of the exploration sequence constitutes
a rather forbiddingly large subject matter which would require several volumes
to do justice to it. Because of format restrictions we are limiting ourselves to
two practical problems of exploration planning and field investigation, including
(1) the optimization of survey designs for the detection of specified types of
mineral deposits, and (2) the optimal selection for drill testing purposes of
exploration targets outlined by field surveys.
The first part of the book, comprising Chapters 1-5, describes the meth-
odology of the optimization of the detection of ore deposits. The subject matter
is covered in a descriptive manner with a very minimal mathematical content,
so that explorationists of the older school with a modest mathematical and
statistical background should not feel discouraged by arcane language and sym-
bols. No mathematical proofs are offered, but interested readers are directed to
carefully selected references to satisfy their curiosity. Chapter 5 briefly depicts
how the theoretical methodology can be applied to the optimization of the airborne
geophysical, ground geophysical, and drilling detection of specific types of ore
deposits known to occur in North America.
The second part of the book covers the optimization of the detection of six
types of ore deposits commonly sought in North America-( 1) porphyry-
Cu-Mo, (2) contact metasomatic, (3) Ni-Cu ultramafic, (4) volcanogenic massive
sulfides, (5) Mississippi Valley-type Pb-Zn, and (6) vein gold deposits-in three
geological regions of the North American continent, namely, the Cordillera
Belt, the Precambrian Shield, and the Arctic Paleozoic Platform (Chapters 6-11).
Finally, Chapter 12 covers the crucial problem of the optimal selection of ex-
ploration targets worthy of testing.
Because we have been critical of the rather offhanded manner with which
references and selected readings have been dealt with in many recent books, we
have devoted much time and attention to this important matter. In order to assist
the more selective readers with their research into specific topics, we are sub-
mitting lists of pertinent titles at the end of each chapter. When warranted
(Chapters 1, 2, 4, and 12), the references are further segregated into clearly
labeled categories covering narrower topics.
PREFACE vii
One last word about the logo displayed on the book cover: being dissatisfied
with the perennial geologists' picks and prospectors' shovels and gold pans which
have been adorning the covers of books on mineral exploration for the past 50
years, we decided to produce a design more attuned to the times and the topic
of the book. The parabola curve aptly and concisely portrays the mathematical
optimization approach, a keystone of the book, as a proper link between modem
exploration methods (aircraft towing a "geophysical bird") of worldwide scope
(map of the world) and commercial success, which is symbolized by the outline
of a mine headframe.
A forthcoming volume titled Geomathematical Models/or Optimizing Min-
eral Exploration deals with the optimized search for the same six types of ore
deposits as in the present volume, but covers four other regions of the world.
These regions include Northern Europe, the Mediterranean, Australasia and East
Asia, as well as the Appalachian Belt and the U. S. portion of the North American
Paleozoic Platform, which could not be covered in this volume due to the lack
of adequate data.
J. De Geoffroy
T. K. Wignall
Sydney, Australia
ACKNOWLEDGMENTS
ix
x ACKNOWLEDGMENTS
of Technology, for their excellent photographic and drafting work, which was
required for many of the illustrations included in the book.
Finally, we are much indebted to Julie Symonds of Eden, New South Wales,
Australia, for so ably handling the daunting task of typing the 188 statistical
tables included in this book.
It is hoped that this book will stand as an excellent example of what can
be accomplished through the fruitful collaboration of business, academe, and
government agencies in several nations separated by many miles of ocean.
CONTENTS
xi
xii CONTENTS
Conclusion. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 351
1
2 CHAPTER ONE
Finally, political risk is ever present at all stages of exploration and de-
velopment of mineral deposits; it may result from unforeseen restrictions in land
tenure or access to mineral land, or changes in taxation rules, environmental
protection, as well as labor laws. Perceived high political risk favors the search
for small, high-grade, "bonanza" types of mineral deposits which are charac-
terized by low investment levels, a short life, and a high payout. On the contrary,
expected low political risk will enable the search for and development of large,
low-grade deposits featuring low profit margins and long pay-back periods.
Likewise, since technical success and economic worth are statistically indepen-
dent. we can write
OPTIMIZING MINERAL EXPLORATION 5
Although the expressions written above are quite justified and very useful,
it may be argued that mineral exploration is not really a one-stage procedure but
has a sequential structure, an important point which will further develop below.
Any well-planned exploration program consists of three stages, including (1)
detection and delineation of exploration targets with the probability of success
PI, (2) investigation and testing of exploration targets with a probability of
success P 2 , and finally, (3) outlining ore targets within the tested exploration
targets, with a probability of success P 3 • Since the three events are statistically
independent, we may write
1.2.1. Introduction
Uncertainty and risk are unfavorable features which are, unfortunately, an
integral part of the mineral exploration business. It is therefore the hallmark of
good management that every effort be made to minimize the magnitude of
uncertainty and risk and mitigate their harmful effects at each stage of the ore
search, whether it be planning, field execution of surveys, or final assessment
of results. There are three types of approach to uncertainty and risk reduction:
(1) repetition of trials,(2) sequential approach, and (3) optimization approach.
The first scheme, quite straightforward though potentially very costly, calls
for the repetition of trials to improve the odds of success in the long run. The
second one is more sophisticated and more resource-effective than the first one;
it is based on a structured sequence of stages, each one depending on the results
of the previous one and leading to the next one in a prescribed manner. The
third approach takes advantage of the second one; it consists of maximizing a
desired goal under cost or risk constraints at each stage of the sequential structure,
resulting in a fully optimized overall procedure within the context of a prespec-
ified corporate strategy.
6 CHAPTER ONE
~
DATA EVALUATION &
AREA SELECT ION DATA ACQUISITION
TARGET DELINEATION
,
STAGE
1ST
STAGE
(2)
SELECflrn OF
GEDLCGICAL REGIONS,
TARGET TIFES &
--1 (1) INITIAL DATA
ACQJISITIW, 1-- - - -4----l I
r-
srORAr.E & RETRIEVAL
EXPLORATIOO ~IETI-k)DS
IN
t (4)
INVESTIGATION OF
H(s)
• F
I0
R
2ND PROSPECfING AREAS BY
(3)SELECTla..; OF
STAGE PROSPECT ING AREAS RIMJTE SE."lSI~G & AIR- DELINEATION OF
EXPLORATI(YII AREAS ~M
BORNE GEOPHYSICAL
SJRVEYS p
J
H(s)
IT
(6j (7)
I:-''VESflGATI()II OF
I I
r---
SELECT lOr.; OF
3RD
EXPLORATIOO AREAS
EXPLORATIQ".I AREAS BY
RECCNNAISSANCE
DELINEATION OF ~l 0
•
STAGE FOR RJRTIiER
GEOWGIGA.L &
ANCMAUllS AREAS N
I~YESTrGATION
GED-O-ifMlCAL SURVEYS
• -'
H
I
,
(10) I
I
r---
INVESTIGATION OF
4TH (9)Sf:LECTIrn OF
A\lCt-IALCUS AREAS
ANO>VI.LOJS AREAS BY (11)
DELINEATIct-; OF .j
STAGE SYSTa.1ATIC GEOLOCICAL.
I
FOR RJRTI-lER EXPLORATICN TARGETS
I\'VESTIGATIO~
GECXJID.lICAL &
GEOPHYSICAL ~VEYS
•I F
E
(12)
I (13)
5TH
STAGE
SELECTION OF
EXPLORATION
I~STIGATION
OF EXPLORATIO'l
11(14)
DELlNEATIO\'
OF ORE
-..J E
TARGETS FOR TARGETS BY ID
I I
DRILL-TESTIlI.'G TARGETS
DRILLING
t P
t IA
,
(15) I C
6TH (16) (17)
SELECTIOI\' OF
ORE TARGETS !DR stST~TIC DELL'JEATIO:-: K
STAGE DETAILED DRILLING
DETAILED DRILLING OF ORE OEPOS ITS
7TH !
i
(IS)
•
LP.>.TIERGRctJl',TI
STAGE I OFH-IVESTIGA.TICN
ORE DEPOSITS
FIGURE 1.1
Flow Diagram of Mineral Exploration Sequence.
The uncertainty and ensuing risk factor may be alleviated by means of the
sequential organization of the flow of various types of information. Three main
types of information are to be considered: (1) prior information based on case
histories and acquired from data storage at the initial planning stage, (2) newly
acquired information which is gained as a result of the execution of the field
programs, and (3) feedback information arising from both success and failure
in field exploration, which is recycled to storage for future use at subsequent
stages of the same program or in other programs. As the amount and quality of
information increase progressively at each stage as compared to the previous
one, both uncertainty and risk are being gradually mitigated.
Many factors of diverse nature, including geographic, particularly climatic,
as well as economic and political ones, have to be reckoned with in the overall
planning, because of their disruptive influence on the proper time sequence of
activities. As modem ore search spreads out farther afield to investigate remote
regions affected by harsh climatic conditions, it is obvious that a special allow-
ance has to be made for these conditions when considering the logistics and
scheduling of activities. For example, airborne surveys may be carried out on a
year-round basis in most regions but the tropical ones, where the monsoon season
is quite unsuitable for systematic flying. Ground surveys should be preferably
carried out during the winter season in both sub-tropical deserts and subarctic
regions, or during the dry season in regions affected by monsoonal climates, for
reasons of greater logistics convenience and crew efficiency.
of prospecting areas within the previously selected region; this may be formally
handled by a multivariate statistical procedure based on a geostatistical model.(41,42)
The second set of decisions involves the quantitative optimization of field survey
designs with which to cover the optimally selected prospecting areas for the
detection of expected ore deposits. This problem of detection optimization is the
major topic of the present book. Finally, there arises a problem of optimal
allocation of resources including time, money, and skills between the various
stages of the exploration sequence, which is within the scope of the new science
of Operations Research. (See Chapter 4.)
1.2.4.3. Optimization of the Field Implementation Phase. The types
of decisions which have to be taken in the field implementation phase are rather
different in nature from those made by the planners. The decision maker has to
be quite flexible; at the end of each stage, decisions have to be made on the spot
and have to be tailored to field results which were unforeseen in the preliminary
planning phase.
One example is the decision whether to terminate a surveyor extend it
further in order to acquire additional data. Zero information is zero cost, but
entails a large risk factor. Complete information is unattainable because of its
prohibitive cost and the special nature and structure of geo-data. Therefore, an
optimal amount and quality of information has to be sought for as a compromise
between the two extreme cases mentioned above. The formal solution of this
optimal choice problem has been attempted by several Operations Research
workers. They use the Dynamic Programming approach combined with Bayesian
decision theory in a graphical manner (decision tree diagrams).
Another problem of optimal choice under uncertainty arises after the com-
pletion of field surveys. The sequential combination of scanning and skimming
mentioned above leads to the delineation of a number of exploration targets
which is still far too large for full investigation by drilling within the bounds of
the allocated budget. Two approaches to the optimization of the selection of
exploration targets are considered in Chapter 12 of the book. The first one,
relying on control locations which have been previously investigated, is suitable
for established mining districts. The second one, recently devised by the second
writer, does not require the availability of control locations, which is an obvious
advantage when dealing with poorly prospected regions.
Finally, once the exploration targets of maximum merit have been selected
within geological and budgetary parameters, the optimal allocation of funds
between selected targets for drill testing has to be dealt with. Celasun (Ref. 21
in Chapter 4) and other workers in the field of Operations Research have in-
vestigated the matter in a formal analytical manner.
OPTIMIZING MINERAL EXPLORATION 11
Because of the prevalence of uncertainty and risk which affect every aspect
of the mineral exploration business, it is proper to view the whole ore search
procedure in a probabilistic context. The goal of mineral exploration is therefore
expressed in terms of probability of success, which, in tum, is factored into
three main components, namely, probability of geological occurrence, proba-
bility of detection, and probability of economic worth. A full coverage of the
evaluation of each of the three components could not be considered within the
one-volume format required here: for example, an adequate treatment of the
probability of occurrence alone would require a full-size book. We have therefore
decided to restrict the topic of the present volume to only one component of the
probability of success, that of the detection probability, which is fully investigated
in its theoretical aspects in Chapter 2 and in its practical implications in Chapters
6-11. Carefully selected references are provided at the end of this chapter for
the readers interested in delving further into the very important topics of occur-
rence and economic worth.
The brief coverage offered in Section 1.2.4 amply demonstrates the breadth
of scope of the optimization approach in the mineral exploration business. Despite
12 CHAPTER ONE
the rather scant attention devoted to date to the optimization of mineral explo-
ration, a full treatment of the subject, however desirable, is not possible within
the constraints of a single-volume format. Only two of seven basic types of
optimal choices arising in the course of ore search are fully covered in the present
book. These comprise (a) the optimization of field survey designs in the prelim-
inary planning phase dealt with in Chapters 2,3, and 4, and (b) the optimization
of the selection of exploration targets for testing at the end of implementation
phase, which is treated in Chapter 12. The methodology of the optimization of
detection, which is described in Chapters 2-4, is then applied to the search for
six common types of ore deposits in North America (Chapters 5-11).
District of Keewatin, Northwest Territories, Canada, Geol. Surv. Can .. Open File Report No.
778, pp. 1-29.
37. CHUNG, C. F., 1978, Computer program for logistic model to estimate the probability of
occurrence of discrete events, Geol. Surv. Can. Pap. No. 78-11.
38. CHUNG, C. F., and AGTERBERG, F. P., 1980, Regression models for estimating mineral
resources from geological map data, J. Math. Geol. 12, 458-473.
39. COX, D. R., 1970, Analysis of Binary Data. Methuen, London.
40. DE GEOFFROY, J., and WIGNALL, T. K. W., 1970, Application of statistical decision
techniques to the selection of prospecting areas and drilling targets in regional exploration, Can.
Inst. Min. Metall. Bull. 63(699), 893-899.
41. DE GEOFFROY, J., and WIGNALL, T. K., 1971, A probabilistic appraisal of mineral resources
in a portion of the Grenville Province of the Canadian Shield, Econ. Geol. 66, 466-479.
42. DE GEOFFROY, J., and WIGNALL, T. K., 1973, Design of a statistical data processing
system to assist regional exploration planning; Can. Min. J. 94(11), 30-35; 94(12), 35-36.
43. HARRIS, D. P., 1965, Multivariate statistical analysis: A decision tool for mineral exploration;
Computer Applications in Mining & Exploration Symposium, Vol. I, pp. CI-C35, University
of Arizona, Tucson, Arizona.
44. HARRIS, D. P., 1969, Alaska's base and precious metals resources: A probabilistic regional
appraisal, Q. Col. Sch. Mines 64(3),295-327.
45. HARRIS, D. P., and AGTERBERG, F. P., 1981, The appraisal of mineral resources, Econ.
Geol. 75, 897-938.
55. SANGSTER, D. F., 1980, Quantitative characteristics of volcanogenic massive sulfide deposits,
Can. Inst. Min. Metall. Bull. 73(814), 74-81.
56. SINCLAIR, A. J., 1974, Probability graphs of ore tonnage in mining camps: A guide to
exploration, Can. Inst. Min. Metall. Bull. 67(749), 71-75.
57. SINGER, D. A., COX, D. P., and DREW, L. J., 1975, Grade and tonnage relationships among
copper deposits, U.S. Geol. Surv. Prof. Pap. 907-A, AI-All.
CHAPTER TWO
2.1.1. Introduction
The principal purpose of field programs is the acquisition of infonnation
which will lead to the detection of mineral deposits. There are two types of
approach to the problem of detection, namely, the direct and indirect methods,
which are commonly used simultaneously or sequentially to best advantage.
Table 2.1 summarizes the main aspects of the two paths with respect to the types
of target sought and detection environment, as well as the methodology and
procedures involved.
The direct approach seeks to detect the mineral deposit itself, by direct
observation. The observation is based on visual recognition, either unaided as
in ground prospecting or mapping, or aided by photogeology or remote sensing
techniques from airborne platfonns, or by mechanical probes, such as drilling,
in three-dimensional investigations. The second method requires a more elaborate
treatment to be given below, because of its indirect nature.
17
TABLE 2.1 ......l.
C»
Methodology of the Detection of Mineral Deposits.
Indirect Detection of Ore Deposits by Geophysical Sensors Indirect Detection of Ore Deposits by Sampling of Geochemical Halos
The geological characteristics of the expected prizes will largely dictate the
choice of technology best suited to assist the detection of mineral deposits of
the selected type. The explorationist will choose the type and design of sensor
or probe with regard to performance parameters such as penetration, sensitivity
to the three types of noise mentioned above, resolution, discrimination, and
lateral coverage. (13)
The combination of geological and technological factors will, in turn, pre-
scribe the selection of the most suitable methodology. The latter is defined as
the most appropriate sequence and mix of techniques to ensure the maximization
of the probability of detection of expected ore deposits.
2.2.1. Introduction
As indicated above, intersection of the expected target by a sensor or a
probe is the prerequisite of detection. The intersection requirement is essentially
a matter of spatial relationships which are based on geometric considerations.
These considerations involve not only the intrinsic geometric characteristics of
expected deposit, search grids, and detectors, but also their mutual configura-
tions.
The geometric characteristics of the expected deposit include size, shape,
and attitude considered in a three-dimensional context, and are probabilistic in
nature. Grid geometry is described by shape and spacing parameters which are
deterministic in nature, as are the parameters describing the detector geometry
(depth of penetration and inclination). The configuration of the expected target
and the control grid is a very important factor which can be viewed in two
aspects: a dimensional one, expressed by the ratio of the longest dimension of
the target over the grid spacing, and a directional one, expressed by the angle
between the grid and the direction of the longest dimension of the target (strike).
Another important directional consideration is the attitude of the detector with
respect to the target.
Slab
Cylinder
Truncated Ellipsoid
L =length R, =B/L
T<:B<:L
B =breadth R2 .. TIL
T =thickness R3 =T/B
V = volume = kLBT
A .. cross section area =kBT
FIGURE 2.1
Size and Shape Parameters of Ore Deposits.
parameters are illustrated in Figure 2.1; they include length, breadth, and thick-
ness, which can be grouped for greater convenience of processing into ratios
such as shape ratios, or products such as areas of principal sections or volumes.
The attitudinal parameters are angular in nature and include the strike ori-
entation, dip angle, and angle of plunge. The strike orientation is measured
24 CHAPTER TWO
clockwise from the true north within 0-180 degrees. The true dip angle (ao) is
measured within 0-90 degrees below the horizontal, within the vertical plane
which is perpendicular to the strike direction. The apparent dip angle (a), mea-
sured in any other vertical plane, is related to the true dip angle by the following
relationship which is tabulated in Table 2.3.:
where d is the angle measured between the vertical plane considered and that
containing the strike line of the target. Finally, the angle of plunge measures
the inclination of the ore deposit within the vertical plane containing the strike
line, in the same manner as the dip.
TABLE 2.3
Conversion of True Dip to Apparent Dip Angles in Degrees.
5 0 1 2 3 4 6 9 13 26
10 1 2 4 6 8 12 17 26 45
15 1 3 5 8 12 17 24 35 56
20 2 3 7 11 16 22 31 43 63
25 2 4 9 14 20 27 36 49 67
30 3 5 10 16 23 31 41 54 71
35 3 6 12 18 26 34 45 58 73
40 3 6 13 20 28 37 48 60 75
45 4 7 14 22 31 40 51 63 76
50 4 8 16 24 33 42 53 65 77
55 4 8 17 25 35 44 55 66 78
60 4 9 17 27 36 46 56 67 78
65 5 9 18 28 37 47 58 68 79
70 5 9 19 28 38 48 58 69 79
75 5 10 19 29 39 49 59 69 80
80 5 10 20 30 40 50 60 70 80
85 5 10 20 30 40 50 60 70 80
N·.
"'L' 0.73
M.·.rJI so.
'l. • 0.48
5..7
FT
~~~~~~--~~E
Qi ib 5.. 10 . . 100 (IN OJ so. ~I
MEA at' ItCIRtZONTAL PIIOoIECTION
No.
"\. °S.24
'l.0o.ze
M.o "«I'
'If . ,. .
04 OS 10 ... ~4.o III IQO
· ..
8 WIDTH FEETI
20~
No.
"\. 02.13
'l. ° 0.21 "\..~
No.
.2 'l. 0030
Me° 1178'
I Me-O. . .
4
FIGURE 2.2
Histograms of Distributions of Geometric Parameters of Porphyry-Cu-Mo De·
posits of the North American Cordillera Belt and Fitted Lognormal Models.
FIGURE 2.3
Relationship between Target Breadth, Dip Angle, and Detection Range in the
Case of Vertical Detection of Ore Deposits.
IHOrTzontal-- - -
I ~~~t i on of
ore~posl t
=l
Surface project ion of target breadth (Bt )
Di p component
__
Surface
Depth of
I overbur4en
or capping
* (H)
I True Dip
(0)1
Vert i cal depth
ao
range of
Detect ion
I
~
Vertical Detection
True breadth of
(Airborne' ground geophysical surveys
, dr i I ling) /-ore deposit (B)
o
B -
t
...!.L
sina
+ (D-H) cota
0
o
28 CHAPTER TWO
section of the ore deposit within the range D, and the attitude parameters (true
dip and plunge). The relationships are as follows:
L, = Lh + D cot c
B, = Bh + D cot ao with Bh = Bo/sin (ao)
R, = B,IL,
If the deposits are subvertical (dip and plunge in the 85-90 degree range), the
three target parameters L" B" and R, are approximately equal to the parameters
L h, Bh, and Rh of the horizontal section of the deposits.
Let us consider now the case of angled detection by drilling to the same
vertical depth D as in the previous case. The diagram of Figure 2.4 shows that
the dip component of the target breadth B, is inflated by the addition of a term
involving the drilling angle b in the following manner:
B, = Bh + D cot ao + D cot b
This results in a much larger shape ratio R" which substantially boosts the
magnitude of the probability of detection. If we consider subvertical deposits,
the term cot ao becomes equal to zero and the dip component included in the
calculation of the target breadth is reduced to that introduced by angled drilling.
Surface
Deptli of Sel41cted
overburll In drilling
or capptn~ angle b
(H) I
___ t
True Dip
ngle a o
Selected _I
I
L
Depth range
,0,
~F..:.-a ng Ie c - a0 _
+ _
b _
FIGURE 2.4
Relationship between Target Breadth, Dip Angle, and Detection Range in the
Case of Angled Detection of Ore Deposits.
hand, the geometric characteristics of expected targets (size, shape, attitude, and
location) are essentially probabilistic in nature, and are expressed by expected
values and their fiducial intervals or probability distributions. Accordingly, it
seems quite appropriate to introduce the theory of geometric probabilities as a
foundation for the calculation of the probabilities of detection of mineral deposits.
random or preferred orientation. At about the same time, Ellis (Ref. 25 in Chapter
4) tackled the more realistic problem of detection by vertical drilling on square
grids of parallelipipedic targets in the three-dimensional space rather than in the
planar context of the previous workers.
In 1965, the first publication of Savinskii' s probability tables for the de-
tection of elliptical targets by discrete sampling grids(40) was an important step
in the application of the theory of geometric probabilities to field exploration.
A few years later, Drew(33) was among the first workers to consider the effect
of grid orientation on the detection of targets by ground surveys. Two years
later, Singer(45) published a very important set of probability tables covering the
detection of elliptical targets of random or known orientation by point-nets of
square, rectangular, and hexagonal shapes, which was much more useful for
field exploration than the earlier Savinskii tables.
More recently, in the mid and late 1970s, additional mathematical treatments
of the problem of target detection by drilling were presented by Singer,(47.48)
Shurygin,(42.43) and Drew. (34) Sinclair(44) is one of the few workers to have
touched on the detection of geochemical halos by point-net sampling in a non-
mathematical way. In 1983, the second writer developed the computer program
OPTGRID to improve on the Singer ELLIPGRID program of 1972(46) for the ground
detection of elliptical targets by square point-nets. Detection probability tables
based on the OPTGRID program are included in Section 2.5.5 of this chapter,
and in all six chapters dealing with the detection of six types of ore deposits in
North America.
In the mean time, after a gap of some 20 years which witnessed a massive
use of airborne geophysical surveys, interest in the application of the theory of
geometric probabilities to airborne ore detection was renewed. McCammon(38)
derives the expressions of the probability of detection of targets of varied shapes
by continuous sampling on parallel, orthogonal,and rhombic grids from concepts
involving conditional probabilities. Later on, Chung(32) expanded McCammon's
work by considering the grid-target orientation factor, as well as lateral coverage
along strips rather than flight lines.
2.4.1. Introduction
Since the early 1950s, as most of the more easily detectable mineral deposits
had been discovered by ground prospecting, a much favored strategy for the
34 CHAPTER TWO
TABLE 2.5
Probabilities of Detection of Randomly Oriented Targets of Varied Shapes by
Airborne Surveys.
Detection Detection
Square grid (S)x(S) feet Parallel grid with spacing(S)ft
U=L/S R=0.10 R=0.50 R=1.0 R=0.10 R=0.50 R=1.0
Confirmed detection
Parallel grid with spacing(S)ft.
than unity. (See the left comer of the first tier of Table 2.4.) When (V), is greater
than the square root of 2, the probability of detection is always unity. Similarly,
the probability of detection of a circular target by a survey on a square grid is
always unity when V, is greater than 1.0, while the probabilities of detection of
elliptical targets are bracketed by those of the other two types of shape. The
expressions of the probabilities of detection of linear and circular targets by
rectangular grids defined by the spacing s and ratio w = sit are shown in the
bottom portion of the first tier of Table 2.4.
2.4.3.4. Generalized Aerial Search on Various Types of Grids. As
mentioned above, the probabilities of detection of an elliptical target can be
interpolated by considering any ellipse as intermediate between a line and a
circle. It was also indicated at the end of Section 2.4.2 that an airborne search
flown on a rectangular grid may be considered as intermediate between a search
on a parallel grid and one on a square grid. We now have four limiting cases
including two for target shape and two for grid shape. The most general case,
that of a search for an elliptical target (R in the interval 0-1) on a rectangular
grid with shape factor w in the interval 0-1, will be bracketed by the four extreme
cases, as illustrated by Figure 2.5.
The probabilities of detection can be calculated by combining the two
interpolation expressions listed below for each value of the ratio V,. If Ph Pc.
O.B
i
"-
l;
~ 06
~
ffi
FIGURE 2.5 ~
o 0.4
Generalized Search: Detec- ....
I&J
tion of Elliptical Randomly Ori-
ented Targets by Airborne Sur- ; PrIII·~t
veys on Orthogonal Grids.
[From McCammon, R. B.(38)
Figure 4, by permission of
Plenum Publishing Corpora-
tion, New York.]
40 CHAPTER TWO
and P e are the probabilities of detection of linear, circular, and elliptical targets
for a specified type of grid, the relationship between the three is as follows:
R being the target shape ratio in the interval 0-1. If Pp' P s, P r are the probabilities
of detection of a target of specified shape R by surveys on parallel, square, and
rectangular grids, the expression relating the three is written as
Pr = (l - w)Pp + wP s
f +'!Ti18
-'!T/18 2V13-rr cos (20) da
The diagram shown in Figure 2.6 illustrates the effect of varying target-grid
orientations on the magnitude of the probability of detection of a linear target.
PROBABILITY OF DETECTION OF MINERAL DEPOSITS 41
0.9 -
0.820 90 degrees
0.8
0.775 70
0.7 0.671 SO
0.6 0.549 30
0.5 0.458 o
0.4
0.3
0.2
0.1
FIGURE 2.6
Probabilities of Detection of Linear Targets by Airborne Surveys on Parallel
Grids Oriented at Different Angles with Respect to Targets.
For example for (U,) = 1 the probability of detection increases by 80% when
the grid is rotated from 0 to 90 degrees with respect to the expected strike.
Circular targets do not exhibit any preferred orientation because of the
symmetry, so that the probability of detection is always equal to the ratio U,.
The probability of detection of an elliptical target whose long axis is making an
angle of a degrees with the grid is obtained by combining the expressions of the
probability of detection of linear and circular targets as indicated in the previous
Section 2.4.3.4. For example, the increase in the probability of detection of an
elliptical target with shape ratio Rb = 0.50 is 30% when the grid is rotated from
o to 90 degrees with respect to the long axis of the target if (U,) = 0.50.
42 CHAPTER TWO
2.5.1. Introduction
The widespread use of discrete sampling programs on ground grids preceded
that of continuous sampling by airborne surveys along parallel profiles. Most of
the work done on the application of the theory of geometric probabilities to
ground exploration during the past three decades pertained to drilling programs
rather than geophysical or geochemical surveys, because of the incentive to
reduce the very high cost of systematic drilling.
analysis in order to reduce progressively the size of the net and adjust it to the
succession of objectives of the search.
In the case of ground geophysical surveys, the spacing between readings
should be adjusted to the anticipated depth and size of the target sought for.
This is particularly important in the case of the high-cost induced polarization
surveys, where the spacing of the electrodes as well as their configuration are
of critical importance.
The directional aspect of the target-grid relationship has to be considered.
It is advantageous to lay a geophysical grid at right angles to the geological
grain of the survey area, because there is a greater geological contrast across
the strike than along it, thus enhancing the magnitude of the geophysical re-
sponse. (12) However, we will find that the probability of detection of an elliptical
target by a ground sampling survey on a square grid is maximized if the grid is
orientated at angles varying from 18 to 45 degrees with respect to the long axis
L, of the target, depending on the shape ratio R, of the target. If we consider
rhombic or rectangular grids, the detection probabilities are maximized when
the longest dimensions of both grid-cell and target coincide.
the grid with respect to the longest dimension of the target L, in the following
manner:
• for R, less than 0.15, the "maximal" angle is 18-20 degrees;
• for R, in the 0.15-0.30 interval, the "maximal" angle is 30 degrees;
• for R, greater than 0.30, the "maximal" angle is 45 degrees.
If we consider the confirmed detection requirement (at least two target intersec-
tions by the detector), the "maximal" angle is zero degrees since such a config-
uration makes the best use of the full length of the target.
The OPTGRID approach is also more realistic than the previous ones, which
assume that the strike direction of the target is accurately known. However, this
is not the case in practical field exploration situations. Generally, the strike
direction is approximately known within a confidence interval of, say, ± 10
degrees about a mean orientation. In the OPTGRID program, the grid is orientated
at the optimal angle with the target mean strike direction according to the target
shape, and probabilities of detection are calculated within the full 20 degree
confidence interval by integration.
Unfortunately, the integration by standard analytical methods is intractable
because of the very complex nature of the expression of the probability of
detection in terms of grid and target parameters. A graphical method based on
Simpson's rule provides an approximation to the integration. In the Simpson
method, any small arc (P(,P 2 ), however complex in shape, which is limited by
two ordinates corresponding to an elementary increment h of the variate x, is
replaced by an elemental arc of parabola passing through the points PI and P 2 •
Thus the area comprised between the two ordinates, the parabola arc and the x
axis, is a numerical approximation of the value of the integral of the very
complicated function F(x) in the interval (x, x + h). This can be easily extended
to any finite arc of curve of any complicated shape corresponding to the (a, b)
interval for the x variate, which can be divided into n intervals of width h equal
to (b - a)/n. This leads to the Simpson expression
ticality for field use of the detection probability calculations. This is illustrated
by the following example considering the detection of an elliptical target of shape
R, = 0.25. According to previous results, we find that the optimal orientation
of the grid is 30 degrees plus or minus 10 degrees with respect to the long axis
(L,) of the target. Thus the interval of variation of the orientation angle is from
20 to 40 degrees, or 'Tr/9 to 2'Tr/9 if expressed in radians. The 'Tr/9 interval is
divided into 100 elemental intervals, each 'Tr/9oo wide. The computer is pro-
grammed to calculate the area under the curve between the x axis and the 'Tr/9
and 2'Tr/9 ordinates by means of the Simpson's approximation formula. The
probability of detection, which represents the mean height of the probability
function within the interval 'Tr/9 = 0.348, is then easily obtained by dividing
the Simpson result by the coefficient 0.348.
TABLE 2.6
Probabilities of Detection of Oriented Targets of Varied Shapes by Ground
Surveys or Vertical Drilling Programs.
TABLE 2.7
Probabilities of Detection of Randomly Oriented Targets of Varied Shapes by
Ground Surveys or Vertical Drilling Programs.
1.0
o 10 20 l) 40 50 60 70 IKl !l
AIItjt of orientation .or in degrees
FIGURE 2.7
Contours of Sensitivity of Detection Probability to Errors in Target Orientation. ~
[From Drew, L. ).(34) Figure 5, by permission of Plenum Publishing Corporation, 1.0
New York.]
50 CHAPTER TWO
target becomes more elongated. For example, if the shape ratio R, = 0.20, the
orientation error has to be 30 degrees in order to reduce the probability of
detection from 0.75 to 0.70, instead of 48 degrees for the more rounded target
shape.
2.6.1. Introduction
In the previous sections of this chapter, we considered only single-stage
detection procedures, such as airborne geophysical surveys, ground geophysical
surveys, or drilling programs. It would be fruitful, however, to consider a com-
monly used two-stage sequential approach to the field detection of ore deposits.
The procedure requires the detection of halos of the kinds known to be associated
with specified types of ore deposits, as a first stage, to be followed by the
detection within the halo of the causative ore deposit.
The methodology of the calculation of the probability of sequential detection
is divided into three steps: (1) calculation of the probability of detection of the
halo, (2) calculation of the probability of detection of the ore deposit within the
halo, and (3) calculation of the overall probability of success based on (1) and
(2).
Assuming that both areas are elliptically shaped, their extent is easily calculated
as the product of the lengths of the major and minor axes by 'lT/4.
It should be noted that the probability of detection calculated in the manner
described above is a minimum value which can be easily exceeded at a substan-
tially higher cost by systematic grid sampling within the perimeter of the halo.
In the latter case the probability of detection is calculated as shown above in
Sections 2.5.3 and 2.5.4.
5. GREGORY, A. F., 1967, Remote sensing in the search for metallic ores: A review of current
practices and future potential, Geol. Surv. Can. Econ. Geol. Rep. 26, pp. 511-526.
6. GOETZ, A. F. H., ROCK, B. N., and GOWAN, L. c., 1983, Remote sensing for exploration:
An overview, Econ. Geol. 78, 573-590.
7. GREENWOOD, J. E. W., 1965, Air photographs in economic mineral exploration, Geol. Surv.
Can. Pap. 65-6.
8. HENDERSON, R. G., and ZIETZ, I., 1949, The upward continuation of anomalies in total
magnetic intensity fields, Geophysics 14, 517-533.
9. KUZWART, M., and BOHMER, M., 1978, Prospecting and Exploration of Mineral Deposits.
Elsevier, Amsterdam.
10. LEE, Y. W., 1960, Statistical Theory of Communication, Wiley, New York.
II. LEVINSON, A. A., 1974, Introduction to Exploration Geochemistry, Appl. Publ. Maywood,
Illinois.
12. PARASNIS, D. S., 1974, Mining Geophysics. Elsevier, Amsterdam.
13. PATERSON, N. R., 1971, Airborne electromagnetic methods as applied to the search for sulfide
deposits, Can. Inst. Min. Metall. Bull. 64(705), 29-38.
14. PEMBERTON, R. H., 1962, Airborne electromagnetics in review, Geophysics 27,691-713.
15. PETERS, W. C., 1978, Exploration and Mining Geology. Chap. 8, Wiley, New York.
16. RAISBECK, G., 1963, Information Theory: An Introductionfor Scientists and Engineers; MIT
Press, Cambridge, Massachusetts.
17. REEDMAN, J. H., 1979, Techniques in Mineral Exploration, Applied Science Publishers,
London.
18. ROSENBERG, P., 1971, Resolution, detectability, and recognizability, Photogramm. Eng 37,
1255-1258.
19. SLICHTER, L. B., 1955, Geophysics applied to prospecting for ores, Econ. Geol. Jubilee
Volume 50, 885-969.
20. SPECTOR, A., 1971, Aeromagnetic map interpretation with the aid of the digital computer,
Can. Inst. Min. Metall. Bull. 64(711), 27-34.
21. ZURFLUEH, E. G., 1967, Applications of two-dimensional linear wavelength filtering, Geo-
physics 32, 1015-1033.
COST OF DETECTION
55
56 CHAPTER THREE
3.2.1. Introduction
Most of the planning stage of any mineral exploration program is taken up
by the organization of activities and the cost estimations of actual expenditures.
Costs cannot be determined exactly beforehand: they have to be estimated,
because mineral exploration functions in a context of uncertainty affecting the
execution and outcome of planned activities.
Cost estimation is generally based on the "analogy" principle. Initial esti-
mates are made by comparison with well-documented case histories of programs
which were carried out in economic and geographic environments similar to that
of the planned project. More finely tuned estimates follow. They take into account
local conditions and are indexed for time and geographic variations. There are
two well-recognized methods of costing projects: unit costing, and ratio costing.
erating in the same district or adjoining regions. Final estimates may be further
refined by indexation of unit costing to take into account time and geographic
variations, thus leading to the ratio costing approach.
TABLE 3.1
Schedule of Exploration Unit Costs for Remote North American locations
(U .S. Dollars).
Airborne surveys
Reconnaissance work:
Remote Sensing: Colour photo (including Infra-red) $150/sq. mile
Infra-red Scanner $520/sq. miles
Magnetic $ 50/line mile
Magnetic + Radiometric $ 75/line mile
Detailed work: fixed wing aircraft:
E.M. $ aO/line mile
E.M. + magnetic $IOO/line mile
Helicopter:
E.M. $lOO/line mile
E.M. + Magnetic $130/line mile
Ground surveys
Reconnaissance:
Geological mapping: $150/sq. mile
Geochemical drainage survey $lOO/sq. mile
Gravity $300/sq. mile
Detailed Work:
Gridding: open country $ 50/line mile
dense bush $l20/line mile
Geological survey $500/sq. mile
Geochemical (Rock or soil sampling) $IOOO/sq. mile
Laboratory analysis: $4 per sample
Geophysical surveys:
Magnetic $200/line mile
E.M. $250/line mile
Radiometric $l20/line mile
Self Potential $l20/line mile
Gravimetric $500/line mile
I.P. $600/line mile
Bedrock exploration
Trenching: $3/cubic foot
Drilling: percussion: reconnaissance $IO/linear foot
systematic $6/linear foot
Diamond Drilling: reconnaissance $35/linear foot
systematic $20/linear foot
Geophysical logging of drill hole: $5/linear foot
COST OF DETECTION
of survey coverage of a unit of area in Alaska or on the Arctic Coast may be,
respectively, 3 times or 5 times the cost of a similar coverage in the western
U.S.A. The practice of indexation makes it possible to compare meaningfully
unit costs for similar activities carried out in different areas and in different
periods of time.
exploration rises from 10% (Q 35%, including a rise of the drilling share from
0% to 25% of the total expenditures.
Providing that similar data bases are available for other regions of North
America, ratio costing of the type described above for Alaska should greatly
assist exploration management in establishing accurate definitive estimates for
their projects.
C = c(1 + kD)
where C is the cost function, c is the estimate of the cost of exploration of a
unit of area, D is the distance from the center of the unit area to the nearest
COST OF DETECTION 61
c = cO + kef')
M = Ais + PI2
in terms of the area A and perimeter P, with A = Lt, and P = 2(L + I); the
cost function is then written as:
C = c(Als + P12)
C 1 = c(1/s + 2)
The simplified expression for the coverage of a one mile square area is
C1 = c(1/s + lit + 2)
Square Grids. Since both spacing sand t are equal, the expression of the
cost function becomes
C = c(2AIs + Pis)
C1 = 2c(lIs + 1)
L(lIt + 1)
(lIt + l)(Lls + 1)
If Cd and Cs are the unit costs for the distance- and traveling-related com-
ponents, respectively, the most general expression of the cost function for discrete
ground sampling becomes
for the coverage of an A square-mile area of dimensions L and I mile. For the
coverage of a 1-mile-square-area, the expression becomes
The expressions (2) and (3) are further simplified when we deal with square
grids where s = t miles. The general cost expression for the coverage of an
area of A square miles becomes
3.3.4.3. Cost Function for Drilling Programs. The structure of the cost
function for drilling programs is quite different from that of the ground surveys,
because of the marked predominance of the sampling-related cost component
over the distance-related one.
The sampling component is a function of the grid spacing, as illustrated by
Figure 3.1, and of the depth of holes, as well as being affected by the unit cost
(cost per linear foot) which varies with the type of drilling. If we take the unit
cost for "chum" drilling, the cheapest of all drilling methods, as unity, then the
unit cost for noncoring percussion drilling is about 2.5 and that for the coring
diamond drilling method is about 7.
The distance-related cost component of the function includes two subcom-
ponents. One reflects the traveling cost between setups (moving), and the second
one that of establishing the control grid for the drilling program and the surveying
of collar locations. Generally the unit cost for moving is substantially larger than
that pertaining to the control grid.
Based on past experience, we can simplify the expression of the cost function
for drilling programs by assuming that the $/hole unit cost for the sampling
component Cs is about 50 times the $/mile unit cost for the whole distance
component Cd, so that the most general expression for a rectangular grid (1)
becomes
for an area of A square miles with dimensions L and I miles. The expression (6)
is further simplified if we deal with a I-mile-square area, as L = I = 1 mile,
and becomes
FIGURE 3.1
Graph of Drilling Cost versus Hole Spacing when Holes Cost $4000 Each.
[From Drew, L. J. (34) (of Chapter 2), Figure 8, by permission of Plenum Publishing
Corporation, New York.]
likely direction of the dip can be ascertained. The shallowest surface drilling
angle which can be efficiently used in the present state of the technology is 35
degrees. If we drill a hole with an inclination of b degrees frorn the horizontal,
the unit cost is affected by the angle factor: lIsin b, which is greater than unity.
A 35-degree angled-hole will cost 75% rnore to drill than a vertical hole to the
same vertical depth. For 55-degree angled-drilling, the relative cost increase is
22%, and for 65-degree drilling, it is only 10%.
3.4.1. Foreword
For the readers' convenience, we tabulated five cost functions covering the
three rnain types of field prograrns, including airborne surveys, ground surveys,
and drilling prograrns. Tables 3.3 and 3.4 cover airborne geophysical surveys
on parallel and square grids, based on cornbined rnagnetic, electrornagnetic, and
radiornetric technologies, which are costed at 70 1982 U. S. dollars per line rnile
for fixed-wing craft, and $100 per line rnile for helicopter surveys.
Table 3.5 covers cornbined geophysical surveys including the following
commonly used cornbinations: rnagnetic + electrornagnetic, rnagnetic +
gravity, rnagnetic + induced polarization which are costed at 6 1982 U.S. dol-
lars per station. Sirnilarly, we are using the $6 station cost for cornbined rnultiele-
rnent geochernical surveys including processing laboratory costs. Table 3.6 deals
with vertical drilling on square grids to a depth of 300 feet. Percussion holes are
costed at 2000 1982 U.S. dollars and diarnond drill holes at $6000. The costs
shown in Table 3.6 for vertical drilling are adjusted by rneans of the angle fac-
tor: lIsin 55 when considering 55-degree angled drilling (Table 3.7).
TABLE 3.3
Cost Function for Airborne Surveys on Parallel Grids in U.S. Dollars per Mile
Square.
TABLE 3.4
Cost Function for Airborne Surveys on Square Grids in U.S. Dollars per Mile
Square.
TABLE 3.5
Cost Function for Ground Geophysical or Geochemical Surveys on Square
Grids in U.S. Dollars per Mile Square.
TABLE 3.6
Cost Function for Vertical Drilling Programs on Square Grids in U.S. Dollars
per Mile Square.
TABLE 3.7
Cost Function for 55-degree Angled Drilling Programs on Square Grids in U.S.
Dollars per Mile Square.
I. BAILLY, P. A., 1964, Methods, costs, land requirements and organization in regional explo-
ration for base metals, Preprint, A.I.M.E. Fairbanks meeting, Alaska, March 18-21.
2. CALLAWAY, H. M., 1950, Expense of exploration, Econ. Geol. 45, 328-330.
3. COX, J. L., 1961, Cutting costs through operations research, Min. Congr. 1.47,45-46.
4. CRANSTONE, D. A., and MARTIN, H. L., 1973, Are ore discovery costs increasing? Can.
Min. 1. 94(4), 53-64.
5. DE GEOFFROY, J., and WIGNALL, T. K., 1973, Design of statistical data processing system
to assist regional exploration planning, Can. Min. 1. 94(11), 30-35; 94(12), 35-36.
6. DE GEOFFROY, J., and WIGNALL, T. K., 1974, Evaluating that exploration project, Can.
Min. 1. 95(5), 42-44.
7. DERRY, D. R., 1970, Exploration expenditure, discovery rate and methods, Can. [nst. Min.
Metall. Bull. 63(694), 362-366.
8. DERRY, D. R., and BOOTH, J. K. B., 1978, Mineral discoveries and exploration expenditures
(1966--1976), Min. Mag. 138(5),430-433.
9. GRANT, F. S., 1971, Some thoughts on the next decade in mineral exploration, Geoexploration
9,63-77.
10. KUZW ART, M., and BOHMER, M., 1978, Prospecting and Exploration of Mineral Deposits,
Elsevier, Amsterdam.
11. METZ, P. A., and CAMPBELL, B. W., 1982, Cost of exploration for metallic minerals in
Alaska, 1982, M.I.R.L. Report No. 56, Department of Commerce and Economic Development,
State of Alaska, Fairbanks.
12. PERRY, A. J., 1968, Organization and costs of mineral exploration in the Southwest U.S.A.,
Pacific S.W. Mining Industry Conference, A.I.M.E., May 1968.
13. PETERS, C. W., 1959, Cost of exploration for mineral raw material, Cost Eng. 4(3), July.
14. PETERS, C. W., 1967, Cost and value of drill hole information, Min. Congr. 1. 53(1), 56-59.
15. SARMA, D. D., 1979, An exploration strategy for prospecting with a case study on copper
prospects at Ingladahl, India, Miner. Deposita, 14, 263-279.
16. SLICHTER, L. B., 1955, Geophysics applied to prospecting for ores, Econ. Geol. 50th An-
niversary Volume 50, 885-969.
CHAPTER FOUR
Pd = 1 - exp( -Kx)
75
76 CHAPTER FOUR
Pd Probability of detection
, .00
0.80
0.60
Fitted
Pd = , model:
0.40
0.20
0.00 x
FIGURE 4.1
Graph of Probability of Detection versus Exploration Expenditure.
the acquisition of some type of reward that can be directly translated into mon-
etary units, for example, gross value, gross profit, etc. But in other cases,
different kinds of rewards are sought such as effectiveness, efficiency, precision,
etc., which are indirectly translated into monetary units at the outcome.
The quantification of the qualitative goals into criteria is a necessary step
of the optimization of management's decision making. Criteria may be simple
expressions, or composite ones such as products and ratios or differences which
are better suited to deal with more complex situations.
The third basic tenet of the optimization of business decision making is the
concept of strategy. After defining qualitative goals and quantifying them into
criteria of choice, management has to establish a set of rules to follow, which
will unambiguously tell the decision maker how to deal with the criteria in order
to choose that course of action which is optimal in the prespecified context of
corporate policies and resources. This set of rules is what is referred to as
"strategies. "
4.2.2. Criteria
4.2.2.1. Structure of Criteria. Criteria may be simply structured by
representing a single quantity such as gross value, cost, probability of success,
risk of error as expressed statistically by the variance or coefficient of variation,
or risk of failure as expressed by the complement of the probability of success.
Criteria may be structured as composite expressions such as the product of two
of the quantities listed above. For example, the expected monetary value (EMV)
criterion proposed by Grayson (Chapter 1, Ref. 5) is the product of gross value
times the probability of success; the expected loss criterion is the product of cost
times the probability of failure (complement of the probability of success). Ratios
of two single quantities are also commonly used as criteria, for example, the
effectiveness ratio, which is the ratio of the number of targets actually detected
over the total number of expected targets in the area under investigation. Another
ratio, used by Slichter(32) as an exploration criterion, is his "prospecting success
ratio," which compares the gross value of the prizes to the cost of their detection.
Finally, more complex and flexible structures, such as differences between
some of the composite expressions listed above, are generally more useful when
dealing with actual business situations. The general design is that of a difference
between a reward function and a cost function, both expressed in terms of
common variates and units. Two useful difference criteria are referred to as the
"efficiency criterion" and the "payoff criterion" in the present study.
4.2.3. Strategies
After quantifying the corporate goals as criteria, we need set rules in order
to enable us to select the optimal course of action among a variety of available
options. The guiding rule is referred to as "strategy" in the terminology of
statistical decision theory. In that sense, an optimizing strategy is not a value
judgment; it is only a uniform rule of action which is established in accordance
with prespecified corporate policies. A strategy is meant to guide a choice of
action but is not a particular action at a given time.
The role played by strategies in the methodology of decision theory is as
follows: each option available to the decision maker is assigned a numerical
score which is the value taken by the criterion for that particular option. All
options are then ranked in order of the magnitude of their scores for the decision
maker to make his selection according to the prespecified strategy. There are
three main categories of strategies: (1) those seeking extrema of one type of the
other (maxima or minima) of the criterion without constraints, (2) those seeking
extrema of one type or the other under constraints, and (3) hybrid strategies
seeking to combine extrema of two types, such as the "minimax" strategy so
favored by many Operations Research workersY6)
Strategies calling for maximization will lead to the maximization of a cri-
terion which implies favorable consequences, such as gross value, profit, pre-
cision, detection, occurrence, success, etc. The strategy requires simply the
ranking of the various options based on their scores, i.e., the numerical values
of the criterion, in order of decreasing magnitude, leading to the selection of
the top-scoring option as the most desirable course of action. In the opposite
case, the selected strategy will be that of minimization, based on a criterion
which implies unfavorable circumstances, such as cost, loss, risk, error, etc.
Again, the numerical values of the criterion will be used as scores for the various
OPTIMIZING ORE DETECTION 81
4.3.1. Introduction
In the previous section, the general concepts of optimization were described
in a qualitative manner outside any mathematical context. However, since the
coming of age of the new science of "Operations Research" born from research
into World War II resources allocation problems, analytical methods of systems
optimization based on a mathematical apparatus have rapidly grown in use and
sophistication. The remarkable development of computer capabilities in the past
two decades has greatly assisted the inroad of the mathematical approach, because
of the very large computational requirements of analytical solutions, particularly
when dealing with probabilistic situations.
4.3.2. Fundamentals
4.3.2.1. Introduction of a Mathematical Model. The methodology
of maximization of favorable criteria and minimization of unfavorable criteria
can be translated mathematically into the study of the analytical behavior of an
"objective function." The function is represented by a dependent variate z whose
value is uniquely determined by a number of independent variates (Xi>
X2, . . • ,X;, . . . ,xn ). Optimization in a mathematical context is turned into the
search for extrema of the objective function which represent the criterion. The
extrema may be either maxima or minima, depending on the type of criterion
and strategy chosen. Extrema may be of a "local nature," i.e., occurring within
specified intervals of the value of the independent variates. They may be of a
82 CHAPTER FOUR
"global" nature if they occur within the full range of variation of all independent
variates. Optimization is achieved when we find the numerical values of the
independent variates for which the objective function reaches an extremum of
the type required by the chosen strategy.
However, we should appreciate that in realistic business situations most
systems operate under some kind of constraint, whether physical, monetary, or
time-related. Therefore, our concern should be to translate the concept of con-
straint into mathematical language. Harbaugh(6) very aptly illustrates the quan-
tification of the concept of constraint by means of a geometric analogy. (See
Figure 4.2.) In order to be more general, he considers two types of objective
functions z of two independent variates XI and X2. One model is linear in nature,
but the other one is quadratic.
The linear objective function is simply represented by the expression
(2)
(3)
The three expressions (1), (2), and (3) are graphically represented in the middle-
left portion of Figure 4.2 by the contoured plane and two oblique lines which
define a permissible (shaded) region yielding all feasible solutions. Among them,
the optimum solution, which is maximal in this example, lies at the intersection
of the two inequality lines. At that point, z attains its maximum without going
beyond the constraints. The reasoning would be the same if the inequalities are
not linear (lower left portion of Figure 4.2).
In many instances, the objective function is of a quadratic nature and may
OPTIMIZING ORE DETECTION 83
No
constraints
(II) (t.)
Two lin....
inequalities
(c) (d)
Two nonlin....
Inequalities
FIGURE 4.2
Geometric Representation of Optimization Under Constraints. [From
Harbaugh, S. W., and Bonham-Carter, G. R.,<6) Figure 8.3, by permission of
John Wiley, New York.)
84 CHAPTER FOUR
on calculus calls for the setting of the first derivative Z'(x) to zero. The next
step consists of finding the roots of the equation Z' (x) = 0 which lie within the
required interval (a,b). Among these, only real roots should be retained, the
complex ones being discarded. Depending on the prespecified strategy, either
the roots that result in the maximization of the objective function (maximal roots)
or the minimal roots are of interest in the problem at hand. The screening is
done by calculating the second derivative Z"(x) and examining its sign for each
extremal value of the independent variate: a negative sign indicates a maximal
root and a positive sign a minimal one.
The method can be easily extended to the more complex and general n-
dimensional case, where we have to deal with n independent variates rather than
only one, such as Xl> X2, Xi, . . . ,xn • The problem is handled by setting the first
partial derivative of Z with respect to each of the n independent variates to zero,
one at a time. The result is a system of n simultaneous equations. The system
is solved in order to determine the values of the independent variates which
satisfy the constraints and generate extremal values of the objective function Z
of the type required by the chosen strategy.
The calculus approach to optimization, straightforward as it may seem, is
fraught with difficulties, many of which are. listed by Bellman. (I) First, the form
and structure of the objective function, even in the simpler uni- or bivariate
cases, may be such as to involve products of transcendental functions which
would make the derivation and finding of extremal roots an intractable com-
putational problem, beyond the capability of even the most modem high-speed
computers. A second difficulty pointed out by Bellman is the ambiguity which
occurs when several extremal roots of the required type are found within the
permissible area. Furthermore, we may have to contend with artificial extrema
resulting from discontinuities of the objective function within the permissible
area, or from the presence of boundaries. Finally, in the multidimensional case,
there may be too many variates and too many constraints to make the calculus
method workable at all.
The direct search approach may provide the only feasible solution to op-
timization in two very common types of situations: (1) when the form of the
objective function is too complex and proves intractable for the calculus ap-
proach, and (2) when the nature of the objective function is not sufficiently well
known to be translatable into pure mathematical terms. The direct search ap-
proach can be successfully handled graphically when dealing with one or two
independent variates and a similarly small number of constraints.
4.4.1. Introduction
By bringing together detection probabilities and cost functions for each type
of field program as the two key building blocks we can now construct a model
for the optimization of ore detection which takes full account of the basic theory
of optimization, as described in Sections 4.2 and 4.3 of the present chapter.
The "efficiency" model, already mentioned in Section 4.2.2, is the one
selected for the study of ore search optimization, because it is well suited for
the optimization of individual stages of the mineral exploration sequence. The
diagram shown in Figure 4.3 illustrates the makeup of the model in terms of the
probability of detection as a reward function, cost function, and scaling factors.
It should be pointed out that we are interested in optimizing detection, i.e.,
technical success, only, not commercial success, as they are defined in Section
1.1.1 of Chapter 1, so that no consideration of economic worth of the expected
target is introduced in the model.
where R(x) and C(x) are the reward function and the cost function, respectively,
K and M are scaling factors, and x is the independent variate.
In the context of the application of the efficiency model to control grid
optimization, the grid spacing s is the independent variate, the probability of
detection P is the reward function, and the objective function is written as
Statistical Deterministic
modelling of modelling of
geometric parameters cost factor
of ore deposits (C)
, EFFICIE N CY
, MODEL
- -
I
N.B. k & m = scaling factors
s = grid size
FIGURE 4.3
Construction of an Efficiency Model for the Optimization of Ore Detection.
On the one hand, the range of variation of the reward function P(s) is by definition
restricted to the 0-1 interval. On the other hand, that of the cost function C(s)
extends from 0 to infinity, theoretically at least. In order to deal with this
disparity, we have to choose the values of the scaling factors K and M in such
a way as to keep the terms KP(s) and MC(s) within compatible domains. This
can be achieved by defining M as the reciprocal of the cost required to ensure
a 0.95 probability of detection of a target with expected parameters L, and R"
and K as the reciprocal of 0.95, equal to 1.05. As a result, the efficiency function
will start out in the negative region when s is small and reach a peak when s
increases, heading again for negative values when the grid spacing becomes
quite large, as illustrated by the graphs shown in Figure 4.4. If we wish to
optimize the grid orientation or drilling angle, the independent variate is an
angular interval instead of the spacing s.
The actual determination of optimal survey parameters, including grid spac-
ing and orientation and detector orientation, results from the sequential appli-
cation of either the indirect calculus-based or the direct search approach within
the framework of dynamic programming. The two main types of surveys have
to be handled separately because of the fundamental differences between con-
90 CHAPTER FOUR
Efficiency function
Fls) = klPls) - k 2 C(s)
Grid spacing _ 5
Optimal
t
grid spacing
FIGURE 4.4
Determination of Optimal Grid Size Based on the Efficiency Criterion.
tinuous sampling (airborne surveys), on the one hand, and discrete sampling
(ground surveys), on the other hand.
4.4.2.2. Optimization of Airborne Geophysical Surveys. Two main
types of grids are used as control for airborne geophysical surveys for the purpose
of detecting elliptical targets with long axis L, and shape ratio R, = B,IL,. The
grids are of the parallel design with spacing s or of the orthogonal design with
spacings sand t, t being greater than or equal to s.
If we use the parallel design, the probabilities of detection are initially
maximized by orientating the grid at right angles with the expected direction of
the long axis of the target (See Chapter 2, Section 2.4.4). The variate V, = L,ls
describes the relative grid-target geometry. As noted in Chapter 2, Section 2.4.3,
the probability of detection (reward function) shows a discontinuity for V, = 1,
i.e., when the grid spacing equals the long axis of the target. When V, lies in
the 0-1 interval, the structure of the reward function is rather simple (See Table
2.4, Chapter 2). However, when V, is greater than 1, the makeup of the function
becomes more complex, involving radicals and inverse trigonometric terms which
OPTIMIZING ORE DETECTION 91
would unduly complicate the differentiation and the finding of extremal roots,
if the indirect calculus approach is envisaged. Furthermore, the discontinuity for
U, = I mentioned above introduces an artificial maximum without practical
value. Under these circumstances, the direct search approach is obviously a more
rewarding avenue for the optimization of airborne surveys.
The computer is programmed to calculate the numerical value of the effi-
ciency function in terms of the target parameters L, and R" with the grid spacing
s as variate. An iteration algorithm guides the search for the value of s which
maximizes the efficiency function when s varies. The output of the program
includes (a) the optimal value of the spacing s, (b) the corresponding coverage
cost per unit of area, (c) the associated optimal detection probability, and finally,
(d) a printed graph of the efficiency function.
If we consider the square design s by s, the structure of the efficiency
function is a rather simple quadratic function of lis, which may be easily dif-
ferentiated, leading to the extraction of the maximal root by the calculus ap-
proach. When dealing with rectangular grids with spacing s and shape parameter
w = sit, we find from Chapter 2, Section 2.4.3.4 that the probability of detection
may be calculated as a weighted average of the probabilities associated with the
parallel and square designs, as follows:
Methodology of Optimization
1. BELLMAN, R.E., and DREYFUS, S. E., 1962, Applied Dynamic Programming, Princeton
University Press, Princeton, New Jersey.
2. BROOKS, S. H., 1959, Comparison of maximum-seeking methods, Oper. Res. 7, 430--457.
3. DANTZIG, G., 1963, Linear Programming and Extensions, Princeton University Press, Prince-
ton, New Jersey.
4. DIXON, L. C. W., 1972, Non-linear Optimization, Chap. 1, The English University Press,
London.
5. DIXON, L. C. W. (Ed.), 1976, Optimization in Action, Proceedings of the Optimization
Conference, Univ. of Bristol, January 1975, Chap. 1, Academic, London.
6. HARBAUGH, S. W., and BONHAM-CARTER, G. R., 1970, Computer Simulation in Geo!ogy,
Chap. 8, Wiley, New York.
OPTIMIZING ORE DETECTION 93
7. PETERSON, E. L., 1961, Statistical Analysis and Optimization of Systems, Wiley, New York.
8. ROSENBROCK, H. H., 1960, An automatic method for finding the greatest or least value of
a function, Comput. J. 3, 175-184.
9. SIVAZLlAN, B. D., and STANFEL, L. E., 1975, Optimization Techniques in Operations
Research, Prentice-Hall, Englewood Cliffs, New Jersey.
10. WILDE, D. J., 1964, Optimum-Seeking Methods, Prentice-Hall, Englewood Cliffs, New Jersey.
II. WILDE, D. J., and BEIGHTLER, C. S., 1967, Foundations of Optimization, Prentice-Hall,
Englewood Cliffs, New Jersey.
12. WILLIAMS, K. B., and HALEY, K. B., 1959, A practical application of Linear Programming
in the mining industry, Oper. Res. Q. 10, 131-137.
Operations Research
13. COBB, H., 1960, Operations Research: A tool in oil exploration, Geophysics 25, 1009-1022.
14. COX, J. L., 1961, Cutting costs through operations research, Min. Congr. J. 47,45-46.
15. COYLE, R. G., 1969, Review of the literature on operations research in the mining industry,
Trans. Inst. Min. Metall. London 78, AI-AI9.
16. DUCKWORTH, E., 1962, A Guide to Operations Research, Methuen, London.
17. HAZEN, S. W., Jr., 1968, Operations Research: A growing force in the mineral industries,
Min. Eng. 20, 88-90.
18. PRUSS, D. E., and FREEMAN, G. W., 1961, Mining exploration: An operations research
and simulation approach, Min. World 14, 42-43.
19. ROUBENS, M. (Ed.), 1977, Advances in Operations Research, North-Holland, Amsterdam.
30. KELLEY, 1. c., and McMANUS, D. A., 1969, Optimizing sediment sampling plans, Marine
Geol. 7,465-471.
31. SHURYGIN, A. M., 1976, The probability of finding deposits and some optimal search grids,
J. Math. Geol. 8(3), 323-330.
32. SLICHTER, L. B., 1955, Geophysics applied to prospecting for ore, Econ. Geol. Jubilee
Volume SO, 885-969.
CHAPTER FIVE
The initial stages of the project are (a) selection of ore deposit types and
regions to be covered, (b) acquisition of the necessary data, and (c) preparation
of the data for computerized storage and statistical processing.
95
96 CHAPTER FIVE
TABLE 5.1.
Database for the Optimal Detection of Six Types of North American Ore
Deposits.
Sub-total: 199
Sub-total: 179
5.3.1. Introduction
The statistical processing of the data was carried out in three steps: (1)
preliminary processing, (2) construction of detection probability tables, and (3)
design of optimal field surveys for the detection of specific types of ore deposits.
Two types of computers and two types of word processors available at the
New South Wales Institute of Technology, Sydney, Australia, were used by the
second writer to store, summarize, and process the 2025 data bits. They include
a main frame Honeywell computer, level 66/60, for storage and computations,
a Hewlett-Packard HP9816 computer and HP7580 plotter, an APPLE II for
simpler computations and table printing, and a LISA for composition and printing
of more complex synoptic tables.
98 CHAPTER FIVE
(average) which is particularly suitable for the nonnal and other types of sym-
metrical distributions. The geometric mean is more suitable when dealing with
asymmetric distributions such as the log-nonnal type. If the data are ordered by
magnitude, another statistic called median may be used. The median divides the
area under the histogram into two equal portions and is referred to as the 50th
percentile of the distribution. By extending this idea, we can think of other
statistics which divide the area under the histogram into four equal portions and
are called "quartiles." The first quartile or 25th percentile corresponds to 25%
of the area, the second quartile or 50th percentile is the median, and the third
quartile covers 75% of the area. Finally, the mode is another expression of cen-
trality which corresponds to the highest frequency, that is the peak of the histo-
gram. The mode may not be unique, as is the case in multimodal distributions.
The concept of dispersion or spread of the data about the mean has to be
introduced to describe the variability of the data. The range, already mentioned
above, is a very simple expression of the data spread. A more sophisticated one
is the variance or its square root, the standard deviation. In many instances, it
is more convenient to express the dispersion by a dimensionless coefficient such
as the ratio standard deviation/mean.
Finally, the concept of shape has to be considered in order to describe fully
the nature of frequency distributions. Mean, mode, and median coincide in
symmetrical distributions such as the nonnal model. Most distributions are asym-
metrical, being skewed either on the left or the right side. The coefficient of
skewness expressed as the ratio (mean - mode)/standard deviation is a very
convenient way of describing numerically the concept of shape in a dimensionless
manner. It can be seen that the skewness is equal to zero when the distributions
are symmetrical since mean and mode coincide. Mathematical transfonnations
of the observed data may be used for the "deskewing" of observed distributions
into more symmetrical and more tractable ones. In the common case of lognonnal
distributions of observed data, a logarithmic transfonnation will result in a nonnal
distribution for the logarithms of the data.
confidence intervals are used to indicate that we are confident that the true mean
of the population will fall within the interval in 95% of the cases, the probability
of error, that is of the mean falling outside either limit, being 2.5%.
A "goodness of fit" test is used to test the degree of fit between the theoretical
model and the actual distribution of observed data at a specified level of con-
fidence, generally 0.05. The chi-squared test of a nonparametric nature, which
is most commonly used, is easily calculated in terms of the difference between
observed and expected frequencies for each class. If the calculated expression
is less than the critical value provided by tables for the required level of confidence
and degree of freedom, then we can accept the model as a good fit at the 0.05
confidence level. In the present study, lognormal models were successfully fitted
to the distributions of all dimensional parameters of the six types of deposits.
As far as the attitudinal parameters are concerned, normal and circular normal
models were found to be the best fits for the distributions of dip angles and strike
directions, respectively, for all but the contact metasomatic types of ore deposits.
5.4.1. Summary
As indicated in Sections 2.2.3 and 2.2.4 of Chapter 2, the computation of
detection probabilities for the six types of ore deposits covered by the present
study involves the following geometric considerations:
(i) Intrinsic Target Geometry. As described by the dimensional parameters
of the horizontal sections of the portions of deposits within a 300-foot detection
range, the shape of the horizontal section of the six types of deposits is generally
elliptical, varying from nearly linear (vein gold deposits) to subcircular (pyritic
halos of porphyry deposits).
(ii) Relative Configuration of Detector and Target. The deposits covered
by the study belong to several categories on the basis of the dip parameter. Two
types of deposits including porphyry-Cu-Mo and Mississippi Valley Pb-Zn are
either subhorizontal or subvertical so that the target dimensions for detection
purposes are the same as that of the deposits. The other four types are dipping
deposits for which the target breadth is obtained by adding a dip component to
the breadth of the horizontal section of the deposit, as illustrated by Figures 2.7
and 2.8 of Chapter 2, according to the following expressions:
and
where B, and Bh are the target and deposit horizontal breadths, D is the detection
range taken as 300 feet, and H is the overburden thickness which is taken as 30
feet based on an average of 40 measurements in the Cordillera and 50 mea-
surements in the Shield. Initially, we use 55 degrees as an "all-purpose" drilling
angle. Later, as a part of an optimization study by dynamic programming, the
drilling angle is optimized under cost constraints.
(iii) Grid Geometry and Relative Configuration of Grid and Target. Parallel
and square grid designs are used for the airborne detection of all types of deposits,
except vein gold and contact metasomatic (W-Mo), which are not detectable by
airborne surveys. Square grids are used for the detection of all types of deposits
by ground surveys and drilling programs. The relative configuration of grid and
target involves a dimensional element, the ratio V, = target length/grid spacing,
and a directional element which is the orientation of the grid with respect to the
expected strike direction of the target. Optimization studies show that the de-
tection probabilities for airborne surveys on parallel grids are maximized when
the grid is laid at 90 degrees to the expected target strike. The optimal ground
grid orientation with respect to the expected strike direction lies in the 20-45
degree range, depending on the shape ratio of deposits. No optimal grid ori-
entation is considered for the porphyry-Cu-Mo deposits and their associated
pyritic halos and for the Mississippi Valley Pb--Zn deposits, which are assumed
to be randomly orientated owing to the lack of adequate information.
TABLE 5.2.
Example of Probability Table for the Detection of Porphyry-Cu-Mo
Deposits of the North American Cordillera Belt by Ground Geo-
physical Surveys or Vertical Drilling to a Depth of 300 Feet.
Probability of detection
Spacing S L.c.1. G. mean U.c.1.
in feet L = 2820 3110 3820ft
mation required for the calculation of detection probabilities, induding (a) grid
design, (b) target mean shape ratio and mean length within its 95% confidence
interval, as derived from model fitting and, (c) optimal grid orientation with
respect to target strike within a ± 10 degree confidence interval. The latter
statement is omitted in Table 5.2, because porphyry-Cu-Mo deposits are con-
sidered as randomly orientated. The body of the table displays the numerical
values of detection probabilities for grid spacings varied within a realistic range
of 8~OOO feet in steps of 200-800 feet. For each spacing, the table provides
SEARCH FOR ORE DEPOSITS IN NORTH AMERICA 103
TABLE 5.3.
Example of Table Displaying Grid Sizes and Associated Coverage Cost
per Mile Square for Specific Probability Levels in the Detection of Ni-Cu
Ultramafic Deposits of the North American Shield by Vertical Drilling to
a Depth of 300 Feet.
the 95% confidence interval for the detection probability based on the 95%
fiducial interval for the target length.
Table 5.3 displays an example of the second design of probability tables
which covers the detection of Ni-Cu ultramafic deposits of the Shield by vertical
drilling to a depth of 300 feet. Since this type of table is always inserted between
those of the first design, the target information is not repeated in the heading of
the table. The specified levels of probability of detection are varied within a
range of 0.05-0.95 with 0.05 increments.
The second section deals with vertical ground detection by geophysical surveys
or drilling, and the third section covers angled detection by drilling.
5.5.1. Introduction
An inspection of Tables 5.2 and 5.3 immediately shows that the probability
of detection rapidly decreases to a fraction of percentage as the grid spacing
increases and corresponding coverage cost decreases. Conversely, as the grid
spacing decreases and the corresponding coverage cost rises very rapidly, the
probability of detection reaches it maximum value of unity. It seems intuitive
that three strategies may be considered, including two options covering each end
of the range, and a "middle of the road" option bracketed by the other two. They
are referred to as the "liminal" strategy for the lower end of the range, "maximal"
strategy for the upper end, and "optimal" strategy as the compromise option.
entation of survey grids with respect to the expected strike direction of the target,
which is estimated from previous surveys, photogeological studies, or the results
of statistical modelling. As mentioned in Section 5.4.1, the optimal orientation
of parallel airborne grids is 90 degrees; for ground surveys, it ranges between
18 and 45 degrees, depending on the type of deposit.
The second step is the optimization of grid spacings for optimally oriented
grids. The optimization is carried out analytically for airborne surveys on square
grids as described in Section 4.3.3 of Chapter 4. In the case of parallel grids
for airborne surveys, the analytical approach runs into difficulties because of a
discontinuity of the efficiency function when the grid spacing equals the target
length. When considering ground surveys and drilling programs, the expression
of the efficiency function is very complex and its derivation intractable, ruling
out the analytical approach. The direct search approach has to be used in the
two latter cases. (See Section 4.3.4 of Chapter 4.)
·9
---. -.~
'"
uO. 825 ,/'
-....
~ .75
/
u
I
/
II>
~
.675
/
°
;::,
!!! .525
.6
/
£l.
• /
c:
.-0
.45
/
/
t; .375
c:
"
lJ..
/
.3
>-
u
....u-
c: .225
/
-
;;::
lLI
.15
.075 /
/
....
Lll
N
°....
II>
II>
N
~
°° II>
....
II>
Lll
°
II>
<.t>
II>
....
N
t~ .... '"°
II>
00
II>
FIGURE 5.1.
Graph of EffiCiency Function for the Detection of Volcanogenic Massive Sulfide
Deposits of the North American Cordillera Belt by Optimally Angled (40 de-
grees) Drilling to a Vertical Depth of 300 Feet. The Optimal Grid Spacing is
Indicated by an Arrow.
SEARCH FOR ORE DEPOSITS IN NORTH AMERICA 107
The third step is the optimization of the drilling angle which is omitted for
subhorizontal deposits such as porphyry-Cu-Mo and Mississippi Valley Pb-Zn
types. The optimization is based on the direct search approach using the efficiency
criterion. This presupposes some knowledge of the likely direction of the dip,
as may be obtained from previous surveys or photogeological studies, to avoid
the risk of drilling down-dip. The graph shown in Figure 5.1 illustrates the three-
stage optimization of drilling programs for the detection of volcanogenic massive
sulfide deposits of the Cordillera.
1. ANNIS, R. c., CRANSTONE, D. A., and VALLEE, M., 1978, A survey of known mineral
deposits in Canada that are not being mined, Mining Bulletin MR 181, Department of Energy,
Mines, and Resources, Ottawa.
2. BATEMAN, A., 1942, Economic Mineral Deposits, Chap. 23, Wiley, New York.
3. BROBST, D. A., and, PRATT, W. P. (Ed.), United States mineral resources, U.S. Geol. Surv.
Prof. Pap. No. 820.
4. CANADIAN MINES HANDBOOK, 1978-1979, The Northern Miner Press, Toronto.
5. CRANSTONE, D. A., and WHILLANS, R. T., 1979, Canadian reserves of copper, nickel,
lead, zinc, molybdenum, silver, and gold as of January 1st, 1978, Mining Bulletin MR 185,
Department of Energy, Mines, and Resources, Ottawa.
6. DERRY, D. R., 1980, World Atlas of Geology and Mineral Deposits, Mining Journal Books,
London.
7. DOUGLAS, R. J. W., (Ed.), 1970, Geology and economic minerals of Canada, Geol. Surv.
Can. Econ. Geol. Rep. No. I, Ottawa.
108 CHAPTER FIVE
109
110 CHAPTER SIX
American Cordillera Belt varies from a few million tons of high-grade ore as in
some British Columbia deposits to several hundred million tons of low-grade
material as in many U.S. deposits. The tonnage of deposits is controlled by
physical and chemical factors. The former include (a) the volume of the intrusive,
(b) attitude and lithology of wallrock, and (c) spacing and size of cracks shattering
the host rock. The latter reflect the nature of the mineralizing solutions and that
of the metal precipitating environment. Post-ore geological events such as uplift-
ing and subsequent erosion may result in the removal of substantial tonnages of
ore, while downfaulting may preserve large tonnages and promote upgrading
through supergene alteration.
Current grades of porphyry-copper-molybdenum ore mined in the North
American Cordillera Belt vary from 0.5% to 1.5% copper, accompanied by
0.02% to 0.06% molybdenite, with gold generally under 0.05 ozlton and silver
under 0.1 oz/ton. Conversely, in molybdenum-stockwork deposits, the molyb-
denite grade ranges from 0.1 % to 0.3%, while the copper grade varies in the
0.15% to 0.25% only. Secondary enrichment of copper and gold in the upper
portions of deposits is of little importance in the northern portion of the Cordillera
Belt because of unfavorable climatic conditions. But in the southwestern U.S.
section of the Cordillera, the extent and intensity of supergene enrichment have
influenced production decisions in many cases, particularly in earlier days.
The increasing importance of porphyry-type base metal deposits as world
sources of copper, molybdenum, and valuable minor metals such as gold, silver,
selenium, tellurium, and rhenium justifies the prominent place given to the
porphyry-target in present-day exploration planning. The large size of the por-
phyry deposits, their shallow occurrence, favorable attitude and even metal
distribution make it possible to use large-scale, low-cost surface production
methods that are amenable to computerized mining and scheduling, thus off-
setting the low grade of the ore. However, porphyry projects are susceptible to
economic and financial risks because of their long preproduction time require-
ments, high capital costs, and low profit margins resulting in long payback
periods.
deposits through their primary halos has gained increasing attention in recent
times.
The success of the method relying on the detection of primary halos depends
on the presence of sufficient rock exposure or the availability of drill core to
supplement poor outcropping. The technique is keyed on the distribution of
minor metal associations which are used as proximal indicators. For example,
molybdenum-rhenium or copper-selenium associations denote the core zone of
the mineralized environment; the cadmium-gallium-germanium association with
zinc characterizes the outer zone of deposits.
Secondary halos are the product of the weathering and transport of material
from the mineral deposits themselves or from their primary halos. Sampling
media are residual topsoil, stream silt, plant matter, or ground water. Sillitoe(25)
reports on the merits of using gold as a tracer element in residual soils because
gold abundance in soils appears to reflect closely that of the hypogene gold in
the underlying protore.
TABLE 6.1
List of Porphyry-Cu-Mo Deposits of the North American Cordillera Included
in the Database.
Statistic of
observed data
TABLE 6.3
Summary of Statistical Modeling of Geometric Parameters of Pyritic Halos As-
sociated with Porphyry-Cu-Mo Deposits of the North American Cordillera Belt
(Sample Size: 43 of the 57 Deposits Included in Table 6.1).
Statistic of
observed data
6.4.1. Introduction
For detection purposes, porphyry-Cu-Mo deposits are described as either
tabular subhorizontal or pipe like subvertical bodies with horizontal cross sections
of elliptical shape. The calculations of detection probabilities therefore apply to
plane elliptical targets whose dimensions are those of the horizontal section of
the deposits without the addition of dip components (see Section 2.2.3 of Chapter
2). Because of the lack of accurate orientation data, both porphyry deposits and
their associated halos are considered as randomly oriented plane targets.
interval of the mean. As a result, the tables display the probability of detection
or confirmed detection within its 95% confidence interval for each grid spacing.
TABLE 6.4
o
oz
Summary of Detection Characteristics of Porphyry-Cu-Mo Deposits of the North
American Cordillera Belt for Three Types of Detection Strategies. o
"TI
~
Liminal detection Optimal detection Maximal detection o
;C
Coverage cost: Coverage cost: ~
~
~
I..C
~
N
o
TABLE 6.5
Summary of Detection Characteristics of Pyritic Halos Associated with Porphyry-Cu-Mo
Deposits of the North American Cordillera Belt for Three Types of Detection Strategies.
n
:::r
»
."
...,
m
;;>:l
Vl
X
DETECTION OF PORPHYRY DEPOSITS 121
The detection characteristics pertaining to the optimal option are culled from
the tables listed above and assembled into the central portions of Tables 6.4
(porphyry deposits) and 6.5 (halos). An inspection of both tables shows that
detection parameters corresponding to the optimal option are bracketed by those
associated with the liminal and maximal options, as could be expected. Table
6.4 shows that the optimal probabilities of detection of porphyry deposits by all
types of geophysical surveys lie within a range of 0.79-0.86 with corresponding
coverage cost per unit of area only marginally higher than of the liminal option.
The optimal probability of detection by drilling programs (0.80) compares fa-
vorably with that of the maximal option, while the corresponding coverage cost
is only one half of that of the maximal cost. The optimal probabilities of detection
of halos by airborne surveys are close to that provided by the maximal option,
while coverage costs are substantially lower (Table 6.5).
If we wish to compare the detection performances of parallel and square
grids for airborne geophysical surveys searching for porphyry deposits or halos
under the optimal option, we cannot rely on the coverage cost per unit of area
as a criterion, because the probability of detection is not held at a specified level.
The "expected loss" criterion combining coverage cost and detection probability
can be used to advantage. It is written as a product as follows:
The latter term is the complement of the optimal probability of detection written
as (1 - Pd ). Obviously, the choice of design will be based on the minimization
of the expected loss due to the failure of detection. We find from Table 6.4 that
the expected loss per unit of area is $46 for the optimal square grid design against
$56 for the optimal parallel design, which should be rejected. Similarly, the
square grid design should be our choice for the optimal detection of pyritic halos
by airborne surveys because the expected loss is only $17 as compared with $28
for the second design.
TABLE 6.6
Probabilities of Detection of Porphyry-Cu-Mo Deposits of the North American
Cordillera Belt by Airborne Geophysical Surveys (Continuous Readings) on
Parallel Grids.
Probabili ty of detection
Grid spacing l.c.l. geom. mean u.c.l.
S feet L=2820 3110 3840 feet
TABLE 6.7
Determination of Grid Size for Specified Probability levels in the Detection of
Porphyry-Cu-Mo Deposits of the North American Cordillera Belt by Airborne
Geophysical Surveys on Parallel Grids.
TABLE 6.8
Probabilities of Confirmed Detection of Porphyry-Cu-Mo Deposits of the North
American Cordillera Belt by Airborne Geophysical Surveys (Continuous Read-
ings) on Parallel Grids.
TABLE 6.9
Probabilities of Detection of Porphyry-Cu-Mo Deposits of the North American
Cordillera Belt by Airborne Geophysical Surveys (Continuous Readings) on
Square Grids.
Probability of detection
Grid spacing 1.c.1. geom. mean u.c.l.
S feet L=2820 3110 3840 feet
TABLE 6.10
Optimal Design of Airborne Geophysical Surveys (Continuous Readings) for
the Detection of Porphyry-Cu-Mo Deposits of the North American Cordillera
Belt.
TABLE 6.11
Probabilities of Detection of Pyritic Halos Associated with Porphyry-Cu-Mo
Deposits of the North American Cordillera Belt by Airborne Geophysical Sur-
veys (Continuous Readings) on Parallel Grids.
Probability of detection
Grid spacing l.c.l. geom. mean u.c.l.
S feet L=4870 6374 8344 feet
TABLE 6.12
Determination of Grid Size for Specified Levels of Probability in the Detection
of Pyritic Halos Associated with Porphyry-Cu-Mo Deposits of the North Amer-
ican Cordillera Belt by Airborne Geophysical Surveys on Parallel Grids.
TABLE 6.13
Probabilities of Confirmed Detection of Pyritic Halos Associated with Por-
phyry-Cu-Mo Deposits of the North American Cordillera Belt by Airborne
Geophysical Surveys (Continuous Readings) on Parallel Grids.
TABLE 6.14
Probabilities of Detection of Pyritic Halos Associated with Porphyry-Cu-Mo
Deposits of the North American Cordillera Belt by Airborne Geophysical Sur-
veys (Continuous Readings) on Square Grids.
Probability of detection
Grid spacing 1.c.1. geom. mean u.c.l.
S feet L=4870 6374 8344 feet
TABLE 6.15
Optimal Design of Airborne Geophysical Surveys (Continuous Readings) for
the Detection of Pyritic Halos Associated with Porphyry-Cu-Mo Deposits of
the North American Cordillera Belt.
TABLE 6.16
Probabilities of Detection of Porphyry-Cu-Mo Deposits of the
North American Cordillera Belt by Ground Geophysical Surveys
(Discrete Readings) or Vertical Drilling to a Depth of 300 Feet.
Probability of detection
Spacing S L.c.1. G. mean U.c.1.
in feet L = 2820 3110 3820ft
TABLE 6.17
Determination of Grid Size for Specified Probability Levels in the Detec-
tion of Porphyry-Cu-Mo Deposits of the North American Cordillera Belt
by Vertical Drilling to a Depth of 300 Feet.
TABLE 6.18
Probabilities of Confirmed Detection of Porphyry-Cu-Mo De-
posits of the North American Cordillera Belt by Ground Geo-
physical Surveys or Vertical Drilling to a Depth of 300 Feet.
TABLE 6.19(a)
Optimal Design of Ground Geophysical Surveys on Square Grids for the De-
tection of Porphyry-Cu-Mo Deposits of the North American Cordillera Belt.
TABLE 6.19(b)
Optimal Design for the Detection of Porphyry-Cu-Mo Deposits of the North
American Cordillera Belt by Vertical Drilling to a Depth of 300 Feet.
1. ABRAMS, M. J., BROWN, D., and LEPLEY, L., 1983, Remote sensing for porphyry-copper
deposits in Southern Arizona, Econ. Geol. 78, 591-604.
2. CLARK, K. F., 1972, Stockwork molybdenum deposits in the Western Cordillera, Econ. Geol.
67,731-758.
3. CREASEY, S. c., 1959, Some relations in the hydrothermally altered rocks of porphyry -copper
deposits, Econ. Geol. 54, 354-373.
4. DE GEOFFROY, J., and WIGNALL, T. K., 1972, A statistical study of the geological char-
acteristics of the porphyry-copper-molybdenum deposits of North and South America, Econ.
Geol. 67, 656-668.
5. DE GEOFFROY, J. and WIGNALL, T. K., 1973, Statistical models for porphyry-
copper-molybdenum deposits of the Cordilleran Belt of North and South America, Can. Inst.
Min. Metall. Bull. 66(735), 84-90.
6. ENERGY, MINES, AND RESOURCES DEPARTMENT OF CANADA, 1981, Canadian min-
eral deposits not being mined in 1980, Report M.R. I. 80-7.
7. FOUNTAIN, D. K., 1968, Geophysics applied to the exploration and development of Cu and
Mo deposits of British Columbia, Can. Inst. Min. Metall. Bull. 61(676), 1199-1206.
8. FOUNTAIN, D. K., 1972, Geophysical case histories of disseminated sulfide deposits in British
Columbia, Geophysics 37, 142-159.
9. GODWIN, C. I., 1976, Casino porphyry-Cu-Mo deposit; Porphyry deposits of the Canadian
Cordillera, Can. Inst. Min. Metall. Special Volume 15, 344-354.
10. GUILBERT, J. M., and LOWELL, J. D., 1974, Variations in zoning patterns in porphyry ore
deposits, Can. Inst. Min. Metall. Bull. 67, 99-109.
11. HOLLISTER, V. F., 1973, Characteristics of porphyry copper deposits of South America, Min.
Eng. 25, 51-56.
12. HOLLISTER, V. F., POTTER, R. R., and BARKER, A. L., 1974, Porphyry-type deposits of
the Appalachian Orogen, Econ. Geol. 69, 618-630.
13. HOLLISTER, V. F., ANZALONE, S. A., and PRICHTER, D. H., 1975, Porphyry-copper
deposits of Southern Alaska and contiguous Yukon Territories, Can. Inst. Min. Metall. Bull.
68(755), 104-111.
14. HOLLISTER, V. F., 1978, Geology of the Porphyry-Copper Deposits of the Western Hemi-
sphere, A.I.M.E., New York.
15. JAMES, A. H., 1971, Hypothetical diagrams of several porphyry-copper deposits, Econ. Geol.
66,43--47.
16. KESLER, S. E., 1973, Copper, molybdenum and gold abundance in porphyry-copper deposits,
Econ. Geol. 68, 106--112.
17. LIVINGSTON, D. E., 1973, A plate tectonic hypothesis for the genesis of porphyry-copper
deposits of the Southern Basin and Range Province, U.S.A., Earth Planet. Sci. Lett. 20,
171-176.
18. LOWELL, J. D., and GUILBERT, J. M., 1970, Lateral and vertical alteration-mineralization
zoning in porphyry deposits, Econ. Geol. 65, 373--408.
19. ROSE, A. W., 1970, Zonal relations of wallrock alteration and sulfide distribution at the
porphyry-copper deposits, Econ. Geol. 65,910-936.
20. ROWE, R. B., 1973, Porphyry deposits of the Canadian Cordillera, Can. Min. J. 94(11),
35-38; 94(12), 37--41.
21. SAWYER, J. P. B., and DICKINSON, R. A., 1976, Mount Nansen porphyry deposit; Porphyry
deposits of the Canadian Cordillera, Can. Inst. Min. Metall. Special Volume 15, 336--343.
140 CHAPTER SIX
22. SEIGEL, H. 0., 1974, The magnetic induced polarization (M.I.P.) method, Geophysics 39,
321-339.
23. SILLITOE, R. H., 1972, A plate tectonic model for the origin of porphyry--copper deposits,
Econ. Geol. 67, 184-197.
24. SILLITOE, R. H., 1973, The tops and bottoms of porphyry--copper deposits, Econ. Geol. 68,
799-815.
25. SILLILTOE, R. H., 1979, Some thoughts on gold-rich porphyry--copper deposits, Miner.
Deposita 14, 161-174.
26. STRINGHAM, B., 1960, Differences between barren and productive intrusives, Econ. Geol.
55 1622-1630.
27. SUMNER, 1. S., 1967, Geophysical aspects of porphyry--copper deposits, Geol. Surv. Can.
Rep. No. 26, 322-335.
28. SUTHERLAND-BROWN, A., 1969, Mineralization in British Columbia and the copper-
molybdenum deposits, Can. [nst. Min. Metall. Bull. 62(681), 26-32.
29. TITLEY, S. R., and HICKS, C. L. (Eds.), 1966, Geology of Porphyry-Copper Deposits of
Southwestern North America, University of Arizona Press, Tucson.
30. TITLEY, S. R. (Ed.), 1982, Advances in Geology of the Porphyry-Copper Deposits, University
of Arizona Press, Tucson.
CHAPTER 5 EVE N
141
142 CHAPTER SEVEN
above the critical pressure and temperature of water. The final product, an
assemblage of new lime-rich, iron-rich minerals is called "skarn." The term was
first coined by Medieval Swedish miners and referred to the lime-silicate gangue
of certain types of iron or sulfide ores occurring in the Scandinavian Precambrian
Shield. Most skarn are coarsely crystallized and exhibit crude zonal assemblages
which reflect sedimentary features of the country rock or the geometry of the
intrusive-wall rock contact. (5)
Einaudi(5) and Smirnov(22) distinguish two main types of skarns, including
"endoskarn" and "exoskarn." The latter replaces country rock outside the intru-
sive bodies over widths of a few hundreds to more than one thousand feet. The
former extending inward, on the intrusive side, replaces the intrusive rock over
widths of a few tens of feet to a few hundreds of feet. Exoskarns are, in tum,
divided into "calcic" skarns, the most common type, made up of lime-rich
silicates, hydrosilicates, oxides, and sulfides, and "magnesian" skarns, less com-
mon, which are characterized by different types of hydrosilicates such as phlo-
gopite, tremolite, and talc.
7.1.1.3. Ore Skarn. Ore skarn occurs within the reaction skarn when
shrinkage cracks resulting from metasomatism are filled up with iron and base
metal sulfides, scheelite, barite, fluorite, and secondary quartz. Ore skarn forms
most commonly in the exoskarn zone. In a minority of cases, the ore skarn
straddles the contact and extends into the endoskarn zone (see Figure 7.1).
Ore skarns are made up of early silicates of Ca,Mg,Fe,AI, derived partly
from the wall rock, partly from the intrusive body. The silicates are followed
in the parage netic sequence by scheelite and oxides such as cassiterite and
magnetite. The sequence ends with sulfides including pyrite and pyrrhotite and
base metal sulfides (Cu,Zn,Pb,Mo).
Ore skarns often exhibit zoning patterns at differing scales. At the local
scale, crude banding results from rhythmic metasomatism, while at a larger scale
the zoning reflects stratigraphic features of the wall rock. Finally, at a semire-
gional scale, the distribution of the mineralization shows a zonal arrangement
within the metamorphic aureola of the intrusive: scheelite mineralization is nar-
rowly restricted to the immediate contact, while copper spreads outward for some
distance, and lead and zinc, farther away.
7.1.1.4. Classification of Contact-Associated Ore Deposits. Most
genetic classifications of ore deposits include a category of deposits which may
be labeled as "contact-associated," including such terms as "contact metamor-
phic" or "pyrometamorphic" or "pyrometasomatic" deposits. In this study, we
have chosen the term "contact metasomatic" to emphasize the joint role of contact
and metasomatism in the formation of the deposits.
DETECTION OF CONTACT METASOMATIC DEPOSITS 143
Metallogeny of ...
~------------------------------------~.... Pyrite
.....
contact
1
met ~Lc;oma tic ----------------------~.... Chalcopyrite
ore deposits
----------~~ ~lagnetite
I
Frequency of
contact 5%
metasomat i c
15% 70% 10%
ore deposits
Lithology
up to 500 ft up to 800 ft
Contact
Intrusive
"eripheral metamorphic zone..
I
up to Z500 ft
35~o
3%
.. decreasing SiOZ
Geochemistry decreasing Al Z0 3
..
~
R% 30~o
..
increasing Fep3
5% 35%
increasing CaO
FIGURE 7.1
Environment of Contact Metasomatic Ore Deposits.
144 CHAPTER SEVEN
There are many types and sUbtypes of contact metasomatic deposits, and
their classification generally emphasizes the chief minerals of economic impor-
tance as labels for the categories. In practice, there are innumerable gradations
between four common types which are covered in this chapter. They include (a)
Fe-Cu-Au deposits, (b) complex Pb-Zn-Cu-Ag deposits, sometimes related to
small intrusives, sometimes without obvious relationship with intrusives, (c)
Cu-Mo-Au deposits which are generally associated with differentiated plutons
of the porphyry--copper-bearing type, and (d) W-Mo deposits.
Smirnov(22) cites three main types of structural control for contact meta-
somatic deposits, including (a) the geometry of the intrusive-wallrock interface,
(b) geometry of the bedding and jointing of the wall rock, and (c) the geometry
of the fractures affecting both wall rock and intrusive. In the early stages of ore
deposition, (a) and (b) appear to have a dominating influence, while (c) prevails
in the later stages.
The influence of the geometry of the intrusive-wall rock contact as control
for the deposition of tungsten mineralization is well documented in the North
American Cordillera. (10) Cooling creates a corrugated contact surface by spalling,
generating troughs and grooves which act as ore traps. Kerr(\O) also shows that
the concordance of the contact surface with bedding planes of the wall rock is
far less favorable for ore deposition than discordance. Finally, the influence of
fracturing is illustrated by the accumulation or mineralization in interstrata crush
zones produced during cooling as a result of the difference in competency between
wall rock formations. Furthermore, cross-cutting and oblique faults are believed
to act as feeder channels for ore-bearing solutions.
7.2.1. Introduction
In contrast with the porphyry-Cu-Mo deposits, most contact metasomatic
deposits offer targets of modest size for detection purposes. This rules out direct
detection by remote sensing techniques from satellite platforms. Other direct
detection methods including photogeology from aircraft platforms, ground map-
ping, and drilling have proved most useful in the detection of contact metasomatic
deposits. The indirect detection approach based on the recording of the geo-
physical signature of deposits from aircraft platforms or on the ground has proved
more useful than indirect geochemical methods based on the detection of halos.
TABLE 7.1
list of Contact Metasomatic (Cu-Fe) Deposits of the
North American Cordillera Belt Included in the Data-
base.
Quatsino Mamie
F.L. mine Poor Man
Iron Hill Mt. Andrew
Prescott Utah (U~S.A.I
within the average range of detection (300 feet). The latter include the unoriented
dip and the strike direction of deposits referred to true North and measured in
degrees within the right half circle. The range of measurements of the parameters
for the 23 deposits of the sample is from 500 to 5300 feet for length, 35 to 520
feet for breadth, and 25 to 85 degrees for dip.
The measurements are grouped into classes and assembled into frequency
distributions which are summarized by means of statistics. The statistics include
arithmetic mean, mode, standard deviation, and coefficient of skewness and are
listed in the lower half of Table 7.2. The coefficient of skewness gives an
indication of the asymmetry of the distributions of observed data; the skewness
is moderately high (0.35) for the lengths, but is quite high for the four other
parameters (0.6O--().90).
The observed distributions of the dimensional parameters were fitted by
lognormal models at the 0.05 confidence level. The statistics of the fitted models,
including geometric means, and their 95% fiducial limits, first and third quartiles,
and coefficients of dispersion are listed in the upper half of Table 7.2. The
TABLE 7.2
Statistical Modeling Summary for Contact Metasomatic (Cu-Fe) Deposits of
the North American Cordillera Belt (Sample Size: 23 Deposits).
Statistic
of observed
data
TABLE 7.3.
Probabilities of Detection of Contact Metasomatic (Cu-Fe) Deposits of the
North American Cordillera Belt by Airborne Geophysical Surveys on Parallel
Grids.
Probability of detection
Grid spacing 1.c.1. geom. mean u.c.l.
S feet L= 982 1273 1650 feet
._----------------------------------------------------------------------
200 1. 000 1.000 1. 000
400 0.901 0.924 1.000
600 0.849 0.885 0.912
800 0.792 0.844 0.882
1000 0.753 0.800 0.850
1200 0.627 0.750 0.817
1400 0.538 0.697 0.781
1600 0.470 0.610 0.740
1800 0.418 0.542 0.703
2000 0.376 0.488 0.632
2200 0.342 0.443 0.575
2400 0.314 0.407 0.527
2600 0.289 0.375 0.486
2800 0.269 0.348 0.452
3000 0.251 0.325 0.422
3200 0.235 0.305 0.395
3400 0.221 0.287 0.372
3600 0.209 0.271 0.351
3800 0.198 0.257 0.333
4000 0.188 0.244 0.316
4200 0.179 0.232 0.301
4400 0.171 0.222 0.287
4600 0.164 0.212 0.275
4800 0.157 0.203 0.263
5000 0.151 0.195 0.253
5200 0.145 0.188 0.243
DETECTION OF CONTACT METASOMATIC DEPOSITS 153
TABLE 7.4
Determination of Grid Size for Specified Levels of Probability in the Detection
of Contact Metasomatic (Cu-Fe) Deposits of the North American Cordillera
Belt by Airborne Geophysical Surveys on Parallel Grids.
TABLE 7.5
Probabilities of Confirmed Detection of Contact Metasomatic (Cu-Fe) Deposits
of the North American Cordillera Belt by Airborne Geophysical Surveys on
Parallel Grids.
TABLE 7.6
Probabilities of Detection of Contact Metasomatic (Cu-Fe) Deposits of the
North American Cordillera Belt by Geophysical Surveys on Square Grids.
Probabili ty of detection
Grid spacing 1.c.1. geom. mean u.c.1.
S feet L= 982 1273 1650 feet
TABLE 7.7
Probabilities of Detection of Contact Metasomatic (Cu-Fe) De-
posits of the North American Cordillera Belt by Ground Geo-
physical Surveys or Vertical Drilling to a Depth of 300 Feet.
Probability of detection
Spacing S L.c.1. G. mean U.c.1.
in feet L = 982 1273 1650ft
TABLE 7.8
Determination of Grid Size for Specified Probability Levels in the Detection
of Contact Metasomatic (Cu-Fe) Deposits of the North American Cordillera
Belt by Vertical Drilling to a Depth of 300 Feet.
TABLE 7.9
Probabilities of Confirmed Detection of Contact Metasomatic
(Cu-Fe) Deposits of the North American Cordillera Belt by Ground
Geophysical Surveys or Vertical Drilling to a Depth of 300 Feet.
TABLE 7.10
Probabilities of Detection of Contact Metasomatic (Cu-Fe) De-
posits of the North American Cordillera Belt by 55-Degree An-
gled Drilling to a Vertical Depth of 300 Feet.
Probability of detection
Spacing S L.c.1. G. mean U.c.1.
in feet L = 982 1273 1650ft
TABLE 7.11
Determination of Grid Size for Specified Probability levels in the Detec-
tion of Contact Metasomatic (Cu-Fe) Deposits of the North American
Cordillera Belt by 55-Degree Angled Drilling to a Vertical Depth of 300
Feet.
TABLE 7.12
Probabilities of Confirmed Detection of Contact Metasomatic
(Cu-Fe) Deposits of the North American Cordillera Belt by 55-
Degree Angled Drilling to a Vertical Depth of 300 Feet.
idote, and chlorite with a general absence of garnets which are so conspicuous
in the contact skarns. Finally, there is often no obvious spatial relationship
between the deposits and their metasomatic halos on the one hand, and intrusive
bodies, if any in the vicinity, on the other hand.
The complex Pb-Zn-Cu-Ag deposits are not so common nor so widely
distributed throughout the Cordillera as the Fe-Cu skarn deposits. In the Canadian
section of the Cordillera, these deposits occur in two regions including the Yukon
and the southcentral portion of British Columbia. The deposits of the central
portion of the U. S. section of the Cordillera are larger, richer, and more numerous
than in the Canadian section, and are the models of the type.
TABLE 7.13
list of Contact Metasomatic (Pb-Zn-Cu)
Deposits of the North American Cordillera
Belt Included in the Database.
TABLE 7.14
Statistical Modeling Summary for Contact Metasomatic (Pb-Zn-Cu) Deposits of
the North American Cordillera Belt (Sample Size: 41 Deposits).
Statistic
of observed
data
corresponding coverage costs per unit of area which are required to obtain
prespecified levels of detection probabilities.
The tables are organized into three blocks, one for each type of field pro-
gram. Tables 7.15-7.18 cover airborne geophysical surveys; Tables 7.19-7.21
deal with vertical ground detection by geophysical surveys or drilling to a depth
of 300 feet. Finally, Tables 7.22-7.24 refer to angled drilling to a vertical depth
of 300 feet.
TABLE 7.15
Probabilities of Detection of Contact Metasomatic (pb-Zn-Cu) of the North
American Cordillera Belt by Airborne Geophysical Surveys on Parallel Grids.
Probability of detection
Grid spacing l.c.1. geom. mean u.c.l.
S feet L= 949 1217 1562 feet
----------------------------------------------------------------------
200 0.944 1. 000 1. 000
400 0.887 0.912 0.932
600 0.827 0.867 0.897
800 0.760 0.819 0.861
1000 0.691 0.767 0.824
1200 0.576 0.705 0.785
1400 0.493 0.633 0.741
1600 0.432 0.554 0.710
1800 0.384 0.492 0.632
2000 0.345 0.443 0.568
2200 0.314 0.403 0.517
2400 0.288 0.369 0.474
2600 0.266 0.341 0.437
2800 0.247 0.316 0.406
3000 0.230 0.295 0.379
3200 0.216 0.277 0.355
3400 0.203 0.260 0.334
3600 0.192 0.246 0.316
3800 0.182 0.233 0.299
4000 0.173 0.221 0.284
4200 0.164 0.211 0.271
4400 0.157 0.201 0.258
4600 0.150 0.193 0.247
4800 0.144 0.185 0.237
5000 0.138 0.177 0.227
5200 0.133 0.170 0.219
DETECTION OF CONTACT METASOMATIC DEPOSITS 167
TABLE 7.16
Determination of Grid Size for Specified Probability Levels in the Detection of
Contact Metasomatic (Pb-Zn-Cu) Deposits of the North American Cordillera
Belt by Airborne Geophysical Surveys on Parallel Grids.
TABLE 7.17
Probabilities of Confirmed Detection of Contact Metasomatic (Pb-Zn-Cu) De-
posits of the North American Cordillera Belt by Airborne Geophysical Surveys
on Parallel Grids.
TABLE 7.18
Probabilities of Detection of Contact Metasomatic (Pb-Zn-Cu) Deposits of the
North American Cordillera Belt by Airborne Geophysical Surveys on Square
Grids.
Survey design: square grid with spacings S by S feet
Randomly orientated elliptical targets with expected major axis L feet
in the confidence interval: l.c.l. ( geom. mean ( u.c.l.
and minor axis = B feet
R is defined as the ratio B/L with geometric mean = 0.17
Probability of detection
Grid spacing 1.c.1. geom. mean u.c.l.
S feet L= 949 1217 1562 feet
TABLE 7.19
Probabilities of Detection of Contact Metasomatic (Pb-Zn-Cu)
Deposits of the North American Cordillera Belt by Ground Geo-
physical Surveys or Vertical Drilling to a Depth of 300 Feet.
Probability of detection
Spacing S L.c.l. G. mean U.c.1.
in feet L = 949 1218 1562 i t
TABLE 7.20
Determination of Grid Size for Specified Levels of Probability in the
Detection of Contact Metasomatic (Pb-Zn-Cu) Deposits of the North
American Cordillera Belt by Vertical Drilling to a Depth of 300 Feet.
TABLE 7.21
Probabilities of Confirmed Detection of Contact Metasomatic
(Pb-Zn-Cu) Deposits of the North American Cordillera Belt by
Ground Geophysical Surveys or Vertical Drilling to a Depth of
300 Feet.
TABLE 7.22
Probabilities of Detection of Contact Metasomatic (Pb-Zn-Cu)
Deposits of the North American Cordillera Belt by 55-Degree
Angled Drilling to a Vertical Depth of 300 Feet.
Probability of detection
Spacing S L.c .1. G. mean U.c.1.
in feet L = 949 1218 1562ft
TABLE 7.23
Determination of Grid Size for Specified Probability Levels in the Detec-
tion of Contact Metasomatic (Pb-Zn-Cu) Deposits of the North American
Cordillera Belt by 55-Degree Angled Drilling to a Vertical Depth of 300
Feet.
TABLE 7.24.
Probabilities of Confirmed Detection of Contact Metasomatic
(Pb-Zn-Cu) Deposits of the North American Cordillera Belt by
55-Degree Angled Drilling to a Vertical Depth of 300 Feet.
TABLE 7.25
List of Contact Metasomatic (Cu-Mo-Au) De-
posits of North American Cordillera Belt In-
cluded in the Database.
TABLE 7.26
Statistical Modeling Summary for Contact Metasomatic (Cu-Mo-Au) Deposits
of the North American Cordillera Belt (Sample Size: 15 Deposits).
Statistic
of observed
data
or circular nonnal models at the required 0.05 confidence level. The statistics
summarizing the fitted models are listed in the upper half of Table 7.26.
TABLE 7.27
Probabilities of Detection of Contact Metasomatic (Cu-Mo-Au) Deposits of
the North American Cordillera Belt by Airborne Geophysical Surveys on Parallel
Grids.
Survey design: parallel lines with spacing S feet
Elliptical targets with expected major axis ~ L feet
in the confidence interval: 1.c.1. ( geom. mean ( u.c.1.
and minor axis ~ B feet
R is defined as the ratio B/L with geometric mean ~ 0.19
Orientation of flight lines: 90 degrees with strike-line.
Probabili ty of detection
Grid spacing 1.c.1. geom. mean u.c.l.
S feet L=2206 1273 3180 feet
TABLE 7.28
Determination of Grid Size for Specified Probability Levels in the Detection of
Contact Metasomatic (Cu-Mo-Au) Deposits of the North American Cordillera
Belt by Airborne Geophysical Surveys on Parallel Grids.
TABLE 7.29
Probabilities of Confirmed Detection of Contact Metasomatic (Cu-Mo-Au)
Deposits of the North American Cordillera Belt by Airborne Geophysical Sur-
veys on Parallel Grids.
TABLE 7.30
Probabilities of Detection of Contact Metasomatic (Cu-Mo-Au) Deposits of
the North American Cordillera Belt by Airborne Geophysical Surveys on Square
Grids.
Probabili ty of detection
Grid spacing 1.c.1. geom. mean u.c.1.
S feet L=2206 2648 3180 feet
TABLE 7.31
Probabilities of Detection of Contact Metasomatic (Cu-Mo-Au)
Deposits of the North American Cordillera Belt by Ground Geo-
physical Surveys or Vertical Drilling to a Depth of 300 Feet.
Probability of detection
Spacing S L.c.l. G. mean U.c.l.
in feet L = 2206 2648 3180ft
TABLE 7.32
Determination of Grid Size for Specified Probability Levels in the Detec-
tion of Contact Metasomatic (Cu-Mo-Au) Deposits of the North Amer-
ican Cordillera Belt by Vertical Drilling to a Depth of 300 Feet.
TABLE 7.33
Probabilities of Confirmed Detection of Contact Metasomatic
(Cu-Mo-Au) Deposits of the North American Cordillera Belt by
Ground Geophysical Surveys or Vertical Drilling to a Depth of 300
Feet.
TABLE 7.34
Probabilities of Detection of Contact Metasomatic (Cu-Mo-Au)
of the North American Cordillera Belt by 55-Degree Angled Drill-
ing to a Vertical Depth of 300 Feet.
Probability of detection
Spacing S L.c.1. G. mean U.c .1.
in feet L = 2206 2648 3180 ft
TABLE 7.35
Determination of Grid Size for Specified Probability Levels in the Detection
of Contact Metasomatic (Cu-Mo-Au) Deposits of the North American Cor-
dillera Belt by 55-Degree Angled Drilling to a Vertical Depth of 300 Feet.
TABLE 7.36
Probabilities of Confirmed Detection of Contact Metasomatic Deposits
(Cu-Mo-Au) of the North American Cordillera Belt by 55-Degree Angled Drill-
ing to a Vertical Depth of 300 Feet.
across nearly all western states but Oregon. The geographic breakdown is as
follows: California: 11 districts; Nevada: 8 districts; Utah: 6 districts; the re-
maining five being shared between the states of Idaho and Washington yO) The
most productive belt extends through western California and Nevada over a
length of 700 miles and a width of 200 miles. The main commercial districts of
the tungsten belt are those of Atolia, Bishop, Fresno, and Kern in California(12)
and the Mill City district in Nevada.
The main geological setting for the scheelite deposits is that of skarnified
carbonate formations included in roof pendants which were engulfed by the large
quartz monzonite batholiths of the Sierra Nevada. Generally, the intrusive is
contaminated into quartz diorite in the vicinity of the pendants, and the contacts
are greizenized in the vicinity of mineralized skarn beds. The scheelite-bearing
skarns exhibit assemblages of andradite, diopside, and epidote with sprinklings
of pyrite, chalcopyrite, galena, and sphalerite. Minor molybdenite is known to
occur among nests of fibrous amphiboles, very close to the intrusive contact.
The Pine Creek deposits, among the largest in the world, are located in the
Kern district. (20) They produced a total of 7 million tons of ore from six ore
zones stretching through the "Tungsten Hills" of central-western California. The
Mill City district of Nevada was also an important producer during the war years.
Farther northeast, in Idaho, lies the important district of Yellow Pine, which
includes 20 known occurrences of scheelite-bearing skarns within an area 200
miles long and 100 miles wide.
TABLE 7.37
list of Contact Metasomatic (W-Mo) Deposits
of the North American Cordillera Belt Included
in the Database.
TABLE 7.38
Statistical Modeling Summary for Contact Metasomatic (W-Mo) Deposits of
the North American Cordillera Belt (Sample Size: 37 Deposits).
Statistic of
observed
data
are available in the present state of the technology. A second kind of table
displays the grid spacings and corresponding coverage cost per unit of area which
are required to obtain prespecified levels of detection probabilities.
The tables are organized into two blocks. The first one covers ground vertical
detection by geophysical surveys and drilling to a depth of 300 feet (Tables
7.39-7.41) and the second block deals with angled drilling detection (Tables
7.42-7.44).
TABLE 7.39
Probabilities of Detection of Contact Metasomatic (W-Mo) De-
posits of the North American Cordillera Belt by Ground Geo-
physical Surveys or Vertical Drilling to a Depth of 300 Feet.
Probability of detection
Spacing S L.C .1. G. mean U.c.1.
in feet L = 598 724 875ft
TABLE 7.40
Determination of Grid Size for Specified Probability Levels in the Detec-
tion of Contact Metasomatic (W-Mo) Deposits of the North American
Cordillera Belt by Vertical Drilling to a Depth of 300 Feet.
TABLE 7.41
Probabilities of Confirmed Detection of Contact Metasomatic
(W-Mo) Deposits of the North American Cordillera Belt by Ground
Geophysical Surveys or Vertical Drilling to a Depth of 300 Feet.
TABLE 7.42
Probabilities of Detection of Contact Metasomatic (W-Mo) De-
posits of the North American Cordillera Belt by 55-Degree Angled
Drilling to a Vertical Depth of 300 Feet.
Probability of detection
Spacing S L.c.1. G. mean U.c.1.
in feet L = 598 724 875 ft
TABLE 7.44
Probabilities of Confirmed Detection of Contact Metasomatic
(W-Mo) Deposits of the North American Cordillera Belt by 55-
Degree Angled Drilling to a Vertical Depth of 300 Feet.
Survey design: square grid with spacings S by S feet.
The expected shape of the ore deposits is elliptical
with major axis = L in the 95% confidence interval:
L.c.1. ( geometric mean ( u.c.1., and minor axis = B feet,
R is defined as the ratio B/L with geometric mean = 0.56
grid orientation 0 degrees + or - 10 from the expected
strike line.
7.7.1. Introduction
Three strategies are considered for the detection of four types of contact
metasomatic deposits in the North American Cordillera, namely, the liminal,
maximal, and optimal options, as described previously in Section 5.5 of Chapter
5.
The liminal strategy is designed to ensure a minimum level of 0.500 of
detection probability at a relatively low cost for the coverage of large areas of
little known potential. The maximal strategy ensures a near certain detection
(probability level of 0.980); but, because of its high cost, the maximal option
should be considered only for the coverage of small areas of very high potential.
The optimal strategy, a compromise option bracketed by the first two, should
be most attractive in the majority of exploration situations, when dealing with
areas of moderate size and fair to good potential.
It is most useful for planning purposes to compare the detection perfor-
mances of various types of field programs or grid designs. When the probability
of detection is held at a specified level, as is the case for the first two options,
the coverage cost per unit of area is a convenient yardstick for the comparisons.
When dealing with the optimal option, however, we have to use the expected
loss criterion, as described in Section 6.5.4 of Chapter 6, because the level of
optimal probability of detection varies from one type of field program to another.
The detection characteristics associated with the three strategies are listed
in Tables 7.45, 7.46 and 7.47, covering the four types of contact metasomatic
deposits of the Cordillera.
55-degree
angled $536,000 0.92 $705,000 0.93 $247,000 0.98 $981,000 0.86
drilling 700 ft 600 ft 1100 ft 500 ft
square grid
Optimal 45-
degree angled $573,000 0.96 $754,000 0.96 $265,000 0.99 $1,581,000 0.98
drilling 700 ft 600 ft 1100 ft 400 ft
square grid
DETECTION OF CONTACT METASOMATIC DEPOSITS 203
TABLE 7.48
Optimal Design of Airborne Geophysical Surveys for the Detection of Contact
Metasomatic (Cu-Fe) Deposits of the North American Cordillera Belt.
Confidence interval: all results are reported as 95% confidence intervals
respectively
Expected target length in feet: 982 1273 1650
Expected shape ratio R = 0.25
Unit Cost: $70/l.ml.
TABLE 7.50
Optimal Designs for the Detection of Contact Metasomatic (Cu-Fe) Deposits
of the North American Cordillera Belt by Drilling to a Vertical Depth of 300
Feet.
TABLE 7.51
Optimal Design of Airborne Geophysical Surveys for the Detection of Contact
Metasomatic (pb-Zn-Cu) Deposits of the North American Cordillera Belt.
TABLE 7.52
Optimal Design of Ground Geophysical Surveys for the Detection of Contact
Metasomatic (Pb-Zn-Cu) Deposits of the North American Cordillera Belt.
TABLE 7.53
Optimal Designs for the Detection of Contact Metasomatic (Pb-Zn-Cu) De-
posits of the North American Cordillera Belt by Drilling to a Vertical Depth of
300 Feet.
TABLE 7.54
Optimal Design of Airborne Geophysical Surveys for the Detection of Contact
Metasomatic (Cu-Mo-Au) Deposits of the North American Cordillera Belt.
TABLE 7.55
Optimal Design of Ground Geophysical Surveys for the Detection of Contact
Metasomatic (Cu-Mo-Au) Deposits of the North American Cordillera Belt.
TABLE 7.57
Optimal Design of Ground Geophysical Surveys for the Detection of Contact
Metasomatic (W-Mo) Deposits of the North American Cordillera Belt.
TABLE 7.58
Optimal Designs for the Detection of Contact Metasomatic (W-Mo) Deposits
of the North American Cordillera Belt by Drilling to a Vertical Depth of 300
Feet.
I. BALL. C. W., 1951, The Emerald, Feeney, and Dodger tungsten ore-bodies, Salmo, British
Columbia, Econ. Geol. 49, 625-638.
2. BUTLER, B. S., 1913, Geology and ore deposits of the San Francisco and adjacent districts,
Utah, U.S. Geol. Surv. Prof. Pap. No. 80.
3. CATHRO, R. J., 1969, Tungsten in the Yukon, West. Miner 42(4),23-40.
4. COMPTON, R. R., 1960, Contact metamorphism in the Santa Rosa Range, Nevada, Bull.
Geol. Soc. Am. 71, 1383-1416.
5. EINAUDI, M. T., 1982, General features, description and origin of skarns associated with
porphyry--<:opper plutons, in Advances in the Geology of Porphyry-Copper Deposits of South-
western North America, pp. 139-210, University of Arizona Press, Tucson.
6. EVANS, A. M., 1980, An Introduction to Ore Geology, Chap. II, Blackwell, Oxford.
210 CHAPTER SEVEN
7. GHOSE, N. C., 1966, Behaviour of trace elements during thermal metamorphism and/or gran-
itization of metasediments and basic igneous rocks, Geol. Rundsch. 55, 608-617.
8. GHOSE, N. c., 1970, Geochemistry of thermal metamorphic and granitization processes in
aureole rocks around Richughuta, Palamua Dis!., Bihar, India, Geol. Rundsch. 59, 686-724.
9. HOLSER, W. T., 1947, Metasomatic processes, Econ. Geol. 42, 384-395.
10. KERR, P. F., 1946, Tungsten mineralization in the United States, Geol. Soc. Am. Mem. No.
15.
II. KLICHINOV, V. A., and SEGALOVICH, V. I., 1979, The application of geophysics to
exploration forcbromite and tungsten, Geol. Surv. Can. Econ. Geol. Rep. No. 31, pp. 476-484.
12. KNOPF, A., 1917, Tungsten deposits of northwestern Inyo County, California, u.S. Geol.
Surv. Bull. No. 640, pp. 229-249.
13. LINDGREN, W., 1933, Mineral Deposits, Chap. 28, McGraw-Hill, New York.
14. LITTLE, H. W., 1959, Tungsten deposits of Canada, Geol. Surv. Can. Econ. Geol. Ser. No.
17.
15. LOVERING, T. S., 1941, The origin of tungsten ore of Boulder County, Colorado, Econ.
Geol. 36, 229-279.
16. LOVERING, T. S., 1949, Rock alteration as a guide to ore-East Tintic District, Utah, Econ.
Geol. Monograph No. I.
17. MORRIS, H. T., and LOVERING T. S., 1952, Supergene and hydrothermal dispersion of
heavy metals in wall rocks of ore bodies, Tintic District, Utah, Econ. Geol. 47, 685-716.
18. RAMBERG, H., 1952, The Origin of Metamorphic and Metasomatic Rocks, The University
of Chicago Press, Chicago.
19. RENNIE, C. c., and SMITH, T. S., 1957, Lead-zine and tungsten ore bodies of Canadian
Exploration Ltd., Salmo, B.C., Structural Geology of Canadian Ore Deposits, Sixth Com-
monwealth Congress, Montreal, pp. 116-123.
20. RIDGE, J. D., (Ed.) 1968, Ore Deposits of the United States, Vols. I and 2, A.I.M.E., New
York.
21. SANGSTER, D. F., 1969, The contact metasomatic magnetite deposits of Southwest British
Columbia, Geol. Suv. Can. Bull. No. 172.
22. SMIRNOV, D. F., 1976, Geology of Mineral Deposits, Chap. 7 (translated by H. C. Creighton),
MIR Publishers, Moscow.
23. WHITE, L. G., 1963, The Canada Tungsten property, Flat River Area, N.W.T., Canada, Can.
[nst. Min. Metall. Bull. 56(614), 390-393.
24. YOUNG, G. A., and UGLOW, W. L., 1926, The iron ores of Canada, Vol. I, British Columbia
and the Yukon, Geol. Surv. Can. Econ. Geol. Ser. No.3.
CHAPTER EIGHT
211
212 CHAPTER EIGHT
posits occur mainly in Archean greenstone belts of the Western Australian and
North American shields and are considered to be of volcanogenic affiliation.
Coad(3) aptly summarizes the results of much of the research work done on this
type of deposits since they were first discovered in Western Australia and sub-
sequently in Canada in the 1960s.
Two types of volcanogenic Ni-Cu mafic-ultramafic deposIts are distin-
guished. The first type occurs in komatiite volcanic suites and comprises two
categories: massive sulfide ore occurring in typical extrusive settings and dis-
seminated ore found mainly in intrusive settings. The second type is found in
tholeiitic volcanic suites, which are generally far less productive than the ko-
matiite series, with a few notable exceptions in the Scandinavian Shield.
The komatiite suites, originally described in the Komati River region of the
Transvaal, are highly magnesian in composition, as expressed by a low PeOI
Mgo + PeO ratio, and show a low Ti02 content, in contrast with the tholeiitic
series. Much effort has been devoted to the study of the stratigraphic distribution
of sulfide ore within the komatiite flows. It was soon recognized that the min-
eralization of economic importance tends to concentrate as pockets, pods, and
lenses of massive sulfides occurring along corrugations, embayments, and flex-
ures of the lower contact of the extrusive bodies. The massive mineralization
commonly grades upward into disseminated sulfides within the middle portion
of the flows. In the intrusive type, however, the bulk of the mineralization is of
a disseminated nature and occurs throughout the central portion of the bodies or
even near the top contact.
TABLE 8.1
list of Ni-Cu Ultramafic Deposits of the
North American Shield Included in the
Database.
Manitoba Quebec
Rankin Inlet Marbridge
Bird River Blondeau Twp.
Lynn Lake Lake Renzy
Bowden Lake Lorraine
Thompson
Pipe
Ontario New Quebec
Alexo Cross Lake
Sothman Katiniq
Ajax Raglan
Langmuir Expo
Texmont
Trebor
Temagami Lake
Gordon Lake
Werner Lake
Kenbridge
Shebandowan
Ni-Cu ultramafic deposits of the North American Shield. The deposits are listed
by name and region in Table 8.1. An effort was made to obtain the best possible
cross section of the three main types of deposits listed in Section 8. I. The very
large Sudbury deposits were deliberately omitted from the subsample representing
the layered complexes, because their uniqueness would flaw the results of the
statistical study. The remainder of the sample is more or less evenly divided
between the undifferentiated and the volcanogenic types of deposits.
TABLE 8.2
Statistical Modeling Summary for Ni-Cu Ultramafic Deposits of the North
American Shield (Sample Size: 25 Deposits).
Statistic
of observed
data
The statistics describing the fitted models are listed in the upper half of
Table 8.2. They include the geometric means ofthe lognormal models, arithmetic
means for normal models, and circular means for circular normal models, all
with the limits of their 95% confidence intervals. Additional statistics include
the 25th and 75th percentiles of the fitted distributions, and the dispersion coef-
ficients. The latter gives an indication of the spread of the data about the mean
of the fitted model, which, in turn, reflects the total variability of the data. (See
Chapter 5, Section 5.3.2.) The variability is rather small for the length and
breadth parameters (range 0.1-0.25), but is substantially greater for the shape
parameter R (0.6).
8.4.1. Introduction
For detection purposes, the Ni-Cu ultramafic deposits of the North American
Shield are considered as dipping oriented slabs with a horizontal cross section
of elliptical shape. As a result, the target dimensions and shape ratios have to
be recalculated as shown in Section 5.4 of Chapter 5 by adding the appropriate
dip component to the breadth of the horizontal section of each deposit. For
example, the mean shape ratio of the actual deposits is inflated from 0.06 (see
Table 8.2) to 0.12 for the targets in the context of vertical detection to a depth
of 300 feet, and to 0.26 for 55-degree angled detection to the same vertical
depth, thus considerably boosting the probability of detection.
The orientation factor was introduced in the construction of the tables be-
cause of the resulting improvement in detection success as compared with results
obtained under the orientation randomness assumption. A dynamic programming
study showed that the probability of detection is maximized in the following
manner: (a) in the case of airborne surveys, when the parallel grid is oriented
at 90 degrees to the expected strike, (b) for ground surveys (vertical detection)
when the grid is laid at an angle of 20 degrees with the expected strike direction;
the angle should be 30 degrees when the 55-degree angled detection option is
considered. The grid is then laid out at the appropriate angle from the expected
strike direction as known from the results of previous surveys or photogeological
studies. If little is known of the likely orientation of the target, the mean of the
circular normal fitted model is a statistically valid estimate to be considered for
the design of the grids (see Table 8.2., upper half).
DETECTION OF Ni-Cu ULTRAMAFIC DEPOSITS 219
costs while ensuring at least a 0.500 level of detection probability. The second
one provides a high probability of detection, close to certainty (0.980) at a much
higher cost, which can be justified for the coverage of small areas of high merit.
Finally, the optimal option is a compromise plan based on the maximization of
the probability of detection under constraints of cost minimization. Exploration-
ists should favor the third option because it meets the requirements of the majority
of exploration situations which call for coverage of areas of moderate size and
moderate potential within strict budgetary constraints. The detection character-
istics corresponding to the three strategies are summarized and assembled into
one single table for the convenience of the readers (Table 8.3.).
$2570
c
Airborne $380 $905 $448 0.75 $1990
geophysical 1500 ft 525 ft 1200 ft 200 ft 150 ft
~
parallel grid ~
~
Airborne $398 - $494 0.76 $805 - :>
-n
geophysical 3400 ft 2090 ft 1300 ft n
square grid om
"'tI
Ground $2500 $4650 $2190 0.72 $7800 $12,000
geophysical 450 ft 275 ft 500 ft 200 ft 150 ft
o
VI
square grid =i
VI
55-degree
angled $632,000 $1,425,000 $621,000 0.81 $1,420,000 $3,520,000
drilling 650 415 ft 700 ft 430 ft 280 ft
square grid
Optimal 45-deg
angled - - $491,000 0
0 80 - - N
drilling 800 ft N
-l.
square grid
~-
222 CHAPTER EIGHT
of the figures on both sides of Table 8.3 shows the relative advantage of the
maximal option over the liminal strategy when specifying confirmation of de-
tection by geophysical surveys and vertical drilling; but there is little difference
between the two options when angled drilling is considered.
An inspection of the rightmost portion of Table 8.3 shows that the square
grid design should be preferred to the parallel design for airborne geophysical
surveys, because selection of the former leads to a 60% saving in coverage costs.
When comparing the merits of the two types of drilling program for field planning
purposes, angled drilling against the dip direction should be favored over vertical
drilling to the same depth because of the resulting 68% saving in coverage costs.
If the dip direction cannot be ascertained from previous surveys or photogeo-
logical studies, vertical drilling may be the preferred option in order to avoid
the risk of drilling down-dip.
The reader will find in the following section a series of thirteen tables
covering the probabilities of detection and optimal designs of the search for
Ni-Cu ultramafic deposits of the North American Shield by airborne and ground
programs (pages 223-232).
TABLE 8.4
Probabilities of Detection of Ni-Cu Ultramafic Deposits of the North American
Shield by Airborne Geophysical Surveys (Continuous Readings) on Parallel
Grids.
Probability of detection
Grid spacing 1.c.l. geom. mean u.c.1.
S feet L= 815 1097 1475 feet
TABLE 8.5
Determination of Grid Size for Specified Probability Levels in the Detection of
Ni-Cu Ultramafic Deposits of the North American Shield by Airborne Geo-
physical Surveys on Parallel Grids.
TABLE 8.6
Probabilities of Confirmed Detection of Ni-Cu Ultramafic Deposits ofthe North
American Shield by Airborne Geophysical Surveys (Continuous Readings) on
Parallel Grids.
TABLE 8.7
Probabilities of Detection of Ni-Cu Ultramafic Deposits of the North American
Shield by Airborne Geophysical Surveys (Continuous Readings) on Square
Grids.
Probabili ty of detection
Grid Spacing 1.c.1. geom. mean u.c.l.
S feet L= 815 1097 1475 feet
TABLE 8.8
Optimal Design of Airborne Geophysical Surveys (Continuous Readings) for
the Detection of Ni-Cu Ultramafic Deposits of the North American Shield.
Confidence interval: all results are reported as 95% confidence intervals
respectively:
Expected target length in feet: 815 1097 1475
Expected shape ratio R = 0.12
Unit cost: $70/1.ml.
TABLE B.9
Probabilities of Detection of Ni-Cu Ultramafic Deposits of the
North American Shield by Ground Geophysical Surveys (Dis-
crete Readings) or Vertical Drilling to a Depth of 300 Feet.
Probability of detection
Spacing S L.c.l. G. mean U.c.l.
in feet L = 815 1097 1475ft
TABLE 8.10
Determination of Grid Size for Specified Probability Levels in the Detec-
tion of Ni-Cu Ultramafic Deposits of the North American Shield by Ver-
tical Drilling to a Depth of 300 Feet.
TABLE 8.11
Probabilities of Confirmed Detection of Ni-Cu Ultramafic De-
posits of the North American Shield by Ground Geophysical Sur-
veys or Vertical Drilling to a Depth of 300 Feet.
TABLE 8.13
Probabilities of Detection of Ni-Cu Ultramafic Deposits of the
North American Shield by 55-Degree Angled Drilling to a Vertical
Depth of 300 Feet.
Probability of detection
Spacing S L.c.1. G. mean U.c.1.
in feet L = 815 1097 1475ft
TABLE 8.15
Probabilities of Confirmed Detection of Ni-Cu Ultramafic De-
posits of the North American Shield by 55-Degree Angled Drilling
to a Vertical Depth of 300 Feet.
TABLE 8.16
Optimal Designs for the Detection of Ni-Cu Ultramafic Deposits of the North
American Shield by Drilling to a Vertical Depth of 300 Feet.
1. CAMERON, E. M., SIDDELEY, G., and DURHAM, C. G., 1971, Distribution of ore elements
in rocks for evaluating ore potential: Nickel, copper and sulfur in ultra-mafic rocks of the
Canadian Shield, Can. Inst. Min. Metall. Spec. Vol. 11, 298-313.
2. CHAYER, T. P., 1960, Some critical differences between Alpine-type and stratiform peridotite-
gabbro complexes, 21st International Geological Congress, Pt. XIII, pp. 247-259.
234 CHAPTER EIGHT
3. COAD, P. R., 1979, Nickel sulfide deposits associated with ultramafic rocks of the Abitibi Belt
and economic potential of mafic-ultramafic intrusions, Ontario Geological Survey, Study No.
20, Toronto, Canada.
4. CORNWALL, H. R., 1966, Nickel deposits of North America, U.S. Geol. Surv. Bull. No.
1223.
5. DOWSETT, J. S., 1979, Geophysical exploration methods for nickel, Geol. Surv. Can. Econ.
Geol. Rep. No. 31, 310-321.
6. ECKSTRAND, O. R., 1972, Ultramafic flows and nickel sulfide deposits in the Abitibi Orogenic
Belt, Geol. Surv. Can. Pap. 72.-1, part A, 75-81.
7. IRVINE, T. N., and SMITH, C. H., 1967, The ultrabasic rocks of the Muskox intrusion,
N.W.T., Canada, in Ultramafic and Related Rocks, pp. 38-49, Wiley, New York.
8. JACKSON, E. D., 1961, Primary textures and mineral associations in the ultramafic zone of
the Stillwater complex, Montana, U.S. Geol. Surv. Prof. Pap. No. 358.
9. KILBURN, L. c., WILSON, H. D. B., GRAHAM, A. R., and OGURA, Y., 1969, Nickel
sulfide ores related to ultrabasic intrusions in Canada, Symposium on magmatic ore deposits,
Economic Geology Publishing Co, pp. 276-293.
10. LIEBENBERG, L., 1970, The sulfides in the layered sequence of the Bushveld igneous complex,
Geological Society of South Africa, Special Publication No.1, pp. 108-207.
11. LUSK, J., 1976, A possible volcanic-exhalative origin for lenticular nickel sulfide deposits of
volcanic association, with special references to those of Western Australia; Can. J. Sci. 13,
451-458.
12. MACKENZIE, B. W., 1968, Nickel: Canada and the World, Mining Rept No. 16, Department
of Energy, Mines and Resources, Ottawa.
13. MILLER, C. P., 1959, A comparison of plant and soil prospecting for nickel, A.I.M.E. Preprint
No. 59-L-40.
14. NALDRETT, A. J., 1966, The role of sulfurization in the genesis of iron-nickel sulfide deposits
of the Porcupine District, Can. Inst. Min. Metall. Bull. 59,489-497.
15. NALDRETT, A. J., and GASPARRINI, E. L., 1971, Archean nickel sulfide deposits in Canada:
Their classification, geological setting and genesis, with some suggestions as to exploration,
Geol. Soc. Aust. Spec. Pub. No.3, 201-226.
16. NALDRETT, A. J., 1973, Nickel sulfide deposits-Their classification and genesis with special
emphasis on deposits of volcanogenic association, Can. Inst. Min. Metall. Bull. 66(739), 45-63.
17. OBIAL, R., APELO, M., and SANTOS, G., 1972, Geochemical prospecting for nickel sul-
fides-An orientation survey, J. Geol. Soc. Philipp. 26, 1-36.
18. ROSS, J. R., and TRAVIS, G. A., 1981, Nickel sulfide deposits of Western Australia in global
perspective, Econ. Geol. 76, 1291-1329.
19. SHLANKA, R., 1969, Copper, nickel, lead and zinc deposits of Ontario, Ontario Department
of Mines, M.R.C. 12, Toronto.
20. TAYLOR, R. B., 1964, Geology of the Duluth gabbro complex near Duluth, Minnesota, Minn.
Geol. Surv. Bull. No. 44.
21. WAGER, L. R., and BROWN, G. M., 1968, Layered Igneous Rocks, Oliver & Boyd, London.
22. WAGER, L. R., VINCENT, E. A., and SMALES, A. A., 1957, Sulfides in the Skaergaard
intrusion, East Greenland, Econ. Geol. 52, 855-903.
23. WALKER, F., 1940, Differentiation of the Palisades diabase, New Jersey, Bull. Geol. Soc.
Am. 51, 1059-1106.
24. WILSON, H. D. B., and BRISBIN, W. c., 1961, Regional structure of the Thompson-Moak
Lake Nickel Belt, Can. Inst. Min. Metall. Bull. 54(596), 815-821.
25. WYLIE, P. J. (Ed.) 1967, A Review of the Geology of the Ultra-mafic and Related Rocks.
Wiley, New York.
26. ZURBRIGG, H. F., 1963, Thompson mine geology, Can. Inst. Min. Metall. Bull. 56, 451-460.
CHAPTER NINE
235
236 CHAPTER NINE
and tapers off within a distance of 300-400 feet below the massive sulfide bodies.
Geochemically, the footwall zone of North American Archean volcanogenic
sulfide deposits shows a marked increase in Fe, Mg, and S and a decrease in
Si, Na, and K, as compared to the hanging wall formations (see Chapter 12,
Ref. 31). In many cases, the alteration pipes of Archean volcanogenic sulfide
deposits have been truncated or deformed by post-ore tectonics, or even obli-
terated by post-ore metamorphism. (36)
Because of the occurrence of repeated cycles of folding and metamorphism
which have affected Archean volcanogenic sulfide deposits since their formation,
the principal lithologic and geochemical features controlling ore deposition have
been disturbed and can be surmised only by analogy from observations made
on more recent deposits of Mesozoic and Cenozoic ages. Post-ore alteration
affecting ore controlling features belongs to three main types: regional meta-
morphism, contact metamorphism, and supergene alteration. Sangster(36) reports
that under the influence of regional metamorphism the chloritic material of the
syn-ore alteration pipes becomes unstable and is altered into an assemblage of
hornblend, biotite, and anthophyllite; if sufficient calcium is present, epidote,
actinolite, and tremolite appear. Examples of the influence of post-ore contact
metamorphism on syn-ore alteration have been described in the Noranda district
of North western Quebec. Large subcircular cordierite aggregates form at the
expense of the chlorite and give the footwall rock a peculiar spotted appearance
(dalmatianite). Supergene alteration, which is quite extensive in tropical and
subtropical climates, is not well developed in the North American region and
results only in the formation of gossans which assist direct detection from the
air and on the ground.
9.2.2. Metallogeny
Two kinds of ore material commonly make up the volcanogenic massive
sulfide deposits: massive sulfides comprising from 50% to 60% of the rock by
volume, and "stringer ore" with 20% to 30% sulfides only. Sangster 37 ) reports
that 50% of the 70 deposits of his database feature both types of ore material,
45% include massive ore only, while the remaining 5% exhibit stringer ore only,
possibly due to the truncation of the massive sulfide portion by post-ore tectonics.
A few volcanogenic sulfide deposits are nearly monomineralic, with pyrite
making up nearly 100% of the sulfides by volume (Spain, Philippines). But most
deposits are polymineralic: the ore material consists of an admixture of pyrite,
pyrrhotite, chalcopyrite, sphalerite with or without galena, minor gold and silver
values. The proportion of iron sulfide may be as high as 68% in volume (Geco,
northwest Ontario) or as low as 36% (Coronation, Manitoba); pyrrhotite is always
238 CHAPTER NINE
ularly in Canada and the Scandinavian countries, at a time when the more obvious
deposits have all been discovered after several decades of thorough prospecting.
The various field detection methods used in the search for volcanogenic
sulfide deposits can be grouped into two main categories, including a direct
approach based on visual detection on the ground or from airborne platforms
and assisted by mechanical probes (drilling). The second approach, an indirect
one, relies on the detection of the signatures of the volcanogenic sulfide deposits
by geophysical or geochemical methods.
TABLE 9.1(a)
List of Volcanogenic Massive Sulfide Deposits of
the North American Shield (N.W. Quebec Sec-
tion) Included in the Database.
TABLE 9.1(b)
List of Volcanogenic Massive Sulfide Deposits of the North
American Shield (Ontario, Manitoba, and N.W.T. Sections) In-
cluded in the Database.
Ontario
Ecstall willecho
Kamkotia Big Nama
Jameland Mattabi
Jamieson N.B.D.
Munro Lyon Lake
Geco Creek
Willroy Coldstream
South Bay
Manitoba
Flin-flon Chisel
Pine Bay Stall
Coronation Osborne
Centennial Freeport Reed
Schist Lake Anderson
West Schist H.B. Reed
Cuprus wim
White Lake Dickstone
Birch Lake Li ttle Stall
North Star Rail
Mandy Ghost Lake
Lost Lake
Northwest Territories
Copperman
Hackett Ruttan
Isok Fox Lake
High Lake
Taki Lake
attitudinal in nature; they include the unoriented true dips and the strike orien-
tation referred to true north and measured in degrees within the right half-circle.
The range of the measured dip angles is from 48 to 85 degrees.
The measurements of the geometric parameters are grouped into classes
and assembled into frequency distributions. The statistics summarizing the dis-
tributions of observed data are listed in the lower half of Table 9.2. Included
are arithmetic means, modes, standard deviations, and coefficient of skewness.
The latter is constructed as the ratio (mean-mode)/standard deviation to indicate
the degree of asymmetry of the frequency distributions. Skewness is slight (0.1)
244 CHAPTER NINE
TABLE 9.2
Statistical Modeling Summary for Volcanogenic Massive Sulfide Deposits of
the North American Shield (Sample Size: 102 Deposits).
Statistic
of observed
data
for the strike orientation parameter, moderate (average 0.45) for length, breadth,
and shape ratio, but is quite strong (0.75) for the dip angle parameter.
Lognormal models were successfully fitted to the observed frequency dis-
tributions of the dimensional parameters, as confirmed by the results of the X2
goodness of fit testing at the 0.05 confidence level. The statistical models fitted
to the dip angle and strike direction distributions are normal and circular normal,
respectively. The statistics of the fitted models are assembled in the upper half
of Table 9.2. They include the geometric means and their 95% confidence limits
for the lognormal models, as well as the arithmetic and circular means and their
confidence limits for the attitudinal parameters. Additional listings include the
25th and 75th percentiles and dispersion coefficients. The latter are constructed
as the ratios of standard deviations over means in order to give a dimensionless
indication of the variability of the data. The coefficient is small (0.1) for the
SEARCH FOR VOLCANOGENIC MASSIVE SULFIDE DEPOSITS 245
length parameter, and moderately large (0.33) for the breadth and shape ratio
parameters.
9.5.1. Introduction
For detection purposes, the volcanogenic sulfide deposits are considered as
oriented dipping slabs with horizontal cross sections of elliptical shape. The
target dimensions are obtained by introducing a dip component calculated in the
manner described in Sections 2.2.3 of Chapter 2 and 5.4.2 of Chapter 5 for both
vertical and angled detection situations. As a result, the mean shape ratio is
inflated from 0.06 for horizontal sections of deposits to 0.19 and 0.32 for target
sections, when vertical and 55-degree angled detection are considered, thus
considerably boosting the probability of successful detection of volcanogenic
sulfide deposits.
The latter is further increased when the control grid orientation with respect
to the expected strike direction of targets is optimized according to the results
of a dynamic programming study conducted as described in Section 5.5.4 of
Chapter 5. The optimal grid orientations for the detection of volcanogenic sulfide
deposits by various types of field programs are as follows: (a) parallel grids for
airborne geophysical surveys are to be laid at 90 degrees ± 10 to the expected
target strike, (b) ground square grids for vertical and angled detection are to be
oriented at angles of 30 and 45 degrees ± 10 with target strike, respectively.
The expected target strike direction can be estimated from previous surveys or
photogeological studies. If no prior information is available, the mean of the
fitted circular normal model is a statistically valid estimate of the strike direction
(see Table 9.2).
55-degree
, o
<Jl
angled $572,200 $1,350,300 $705,000 0.82 $1,010,000 $1,685,000 I =i
<Jl
drilling 675 ft 440 ft 600 ft 480 ft 370 ft
square grid
Optimal 40-deg
angled - - $683,000 0.83 - - N
drilling 700 ft ~
square grid '-I
L- ----
248 CHAPTER NINE
We find that the parallel grid design is preferable to the square design for airborne
geophysical searches because the coverage cost for the former is 13% cheaper.
Similarly, the 55-degree angled drilling option should be preferred to the vertical
option because the coverage cost by the former is 50% cheaper.
surveys is 53% lower than in the square grid case; therefore, the former should
be our choice. The optimally angled (40 degrees) drilling program is a clear-cut
choice over the vertical program, because of the 50% reduction in expected loss
which can be expected.
TABLE 9.4
Probabilities of Detection of Volcanogenic Massive Sulfide Deposits of the
North American Shield by Airborne Geophysical Surveys on Parallel Grids.
Probability of detection
Grid spacing 1.c.1. geom. mean u.c.1.
S feet L= 810 947 1106 feet
TABLE 9.5
Determination of Grid Size for Specified Levels of Probability in the Detection
of Volcanogenic Massive Sulfide Deposits of the North American Shield by
Airborne Geophysical Surveys on Parallel Grids.
TABLE 9.6
Probabilities of Confirmed Detection of Volcanogenic Massive Sulfide Deposits
of the North American Shield by Airborne Geophysical Surveys on Parallel
Grids.
TABLE 9.7
Probabilities of Detection of Volcanogenic Massive Sulfide Deposits of the
North American Shield by Airborne Geophysical Surveys on Square Grids.
Probability of detection
Grid spacing 1.c.1. geom. mean u.c.l.
S feet L= 810 947 1106 feet
TABLE 9.B
Optimal Design of Airborne Geophysical Surveys for the Detection of Volcan-
ogenic Massive Sulfide Deposits of the North American Shield.
TABLE 9.9
Probabilities of Detection of Volcanogenic Massive Sulfide Depos-
its of the North American Shield by Ground Geophysical Surveys
or Vertical Drilling to a depth of 300 Feet.
Probability of detection
Spacing S L.c.1. G. mean U.c.1.
in feet L = 810 947 1106ft
TABLE 9.10
Determination of Grid Size for Specified Levels of Probability in the
Detection of Volcanogenic Massive Sulfide Deposits of the North Amer-
ican Shield by Vertical Drilling to a Depth of 300 Feet.
TABLE 9.12
Optimal Design of Ground Geophysical Surveys for the Detection of Volcan-
ogenic Massive Sulfide Deposits of the North American Shield.
TABLE 9.13
Probabilities of Detection of Volcanogenic Massive Sulfide De-
posits of the North American Shield by 55-Degree Angled Drill-
ing to a Vertical Depth of 300 Feet.
Probability of detection
Spacing S L.c.l. G. mean U.c.l.
in feet L = 810 947 1106ft
TABLE 9.15
Probabilities of Confirmed Detection of Volcanogenic Massive Sul-
fide Deposits of the North American Shield by 55-Degree Angled
Drilling to a Vertical Depth of 300 Feet.
TABLE 9.16
Optimal Designs for the Detection of Volcanogenic Massive Sulfide Deposits
of the North American Shield by Drilling to a Vertical Depth of 300 Feet.
TABLE 9.17
Parameters of Multipurpose Optimal Grids for the detection of Ni-Cu Ultra-
mafic and Volcanogenic Massive Sulfide Deposits of the North American Shield.
A.B. geophysical
parallel grid 1290 to 1350 ft $415 to $430 0.56 to 0.72
A.B. geophysical
square grid 1760 to 1990 ft $510 to $540 0.62 to 0.93
Grd. geophysical
square grid 300 to 500 ft $2200 to $4310 0.77 to 0.93
TABLE 9.18
List of Volcanogenic Massive Sulfide Deposits of
the North American Cordillera Belt Included in the
Database.
British Columbia
(Canada)
Tulsequah Lynx
Anyox H.W.
Chuchua Myra
Sam Goosly Price
Goldstream
Granduc
Kutcho
Seneca
California (U.S.A.) Arizona (U.S.A.) (*)
Shasta-King Iron King
Afterthought United Verde
Iron Mountain Copper Queen
Bully Hill + Rising Star Old Dick
Mammoth
Balaklala
Copper Hill
Keystone
horizontal sections of deposits, and the latter, unoriented dip angle and strike
orientation with respect to true north measured in degrees within the right half-
circle. The statistical range of the length measurements for the 21 deposits of
the sample is from 300 to 7000 feet; breadth measurements vary from 25 to 220
feet, and dip angles from 28 to 85 degrees.
The parameter measurements are grouped into classes from which frequency
distributions are constructed. The statistics summarizing the distributions of
observed data are listed in the lower half of Table 9.19, including arithmetic
means, modes, standard deviations, and coefficients of skewness. The latter give
an indication of the degree of asymmetry of the distributions of observed data.
The skewness is slight (0.13) for the strike orientation, moderate for lengths,
breadths, and dips, but quite strong for the shape ratios (0.6).
Lognormal models were successfully fitted to the observed distributions of
the dimensional parameters, while normal and circular normal models were used
to fit the distributions of dips and strike directions. The models are summarized
264 CHAPTER NINE
by statistics which are listed in the upper half of Table 9.19, including geometric
means for the log-normal models, and arithmetic and circular means for the
attitudinal models, with their 95% confidence limits. Additional statistics are the
25th and 75th percentiles and the dispersion coefficients. The latter give an
indication of the variability of the data. Dispersion is slight for the lengths (0.12),
and moderate for the breadths and shaped ratios (0.25).
TABLE 9.19
Statistical Modeling Summary for Volcanogenic Massive Sulfide Deposits of
the North American Cordillera Belt (Sample Size: 22 Deposits).
Statistic
of observed
data
geophysical surveys using the coverage cost per unit area as a yardstick, we find
that, as is the case for the Shield volcanogenic deposits, the square grid design
should be rejected in the liminal context because it results in a 12% cost increase.
In the maximal context, however, the square grid design is more attractive than
the parallel one, as it results in a 56% reduction of coverage cost per unit area.
An evaluation of the respective merits of angled and vertical drilling pro-
grams based on the yardstick mentioned above leads to the selection of angled
drilling over vertical detection in both the liminal and maximal contexts. Under
the former option, a 50% cost reduction is obtained by choosing angled drilling;
under the latter option, the saving in coverage cost is as much as 58%.
55-degree
angled $214,380 $564,2000 $424,000 0.97 $424,000 $725,000
n
:::c
drilling 1200 ft 680 ft 800 ft 800 ft 650 ft »
square grid "1:J
-j
m
Optimal 40-deg ~
optimal grid parameters are displayed in Tables 9.25 for airborne geophysical
surveys, 9.29 for ground geophysical surveys, and 9.33 for vertical and angled
drilling programs, and are assembled in the central portion of Table 9.20 for the
readers' convenience.
Comparisons of detection performances of various grid designs and field
programs can be based on the expected loss criterion (see Section 6.5.4 of Chapter
6). The choice of the square grid is recommended because it leads to a 35%
reduction of the expected loss which could be incurred if the parallel design was
selected. As far as drilling programs are concerned, the choice between angled
and vertical detection should be in favor of optimally angled drilling (40 degrees),
because the expected loss is 91 % lower than in the case of the vertical option.
The following fourteen tables cover the probabilities of detection and the
optimization of the search for volcanogenic sulfide deposits of the Cordilleran
Belt by airborne and ground programs (pages 268-279).
268 CHAPTER NINE
TABLE 9.21
Probabilities of Detection of Volcanogenic Massive Sulfide Deposits of the
North American Cordillera Belt by Airborne Geophysical Surveys on Parallel
Grids.
Probability of detection
Grid spacing l.c.l. geom. mean u.c.l.
S feet L=1201 1842 2826 feet
TABLE 9.22
Determination of Grid Size for Specified Levels of Probability in the Detection
of Volcanogenic Massive Sulfide Deposits of the North American Cordillera
Belt by Airborne Geophysical Surveys on Parallel Grids.
TABLE 9.23
Probabilities of Confirmed Detection of Volcanogenic Massive Sulfide Deposits
of the North American Cordillera Belt by Airborne Geophysical Surveys on
Parallel Grids.
TABLE 9.24
Probabilities of Detection of Volcanogenic Massive Sulfide Deposits of the
North American Cordillera Belt by Airborne Geophysical Surveys on Square
Grids.
Probabili ty of detection
Grid spacing 1.c.1. geom. mean u.c.l.
S feet L=1201 1842 2826 feet
TABLE 9.25
Optimal Design of Airborne Geophysical Surveys for the Detection of Volcan-
ogenic Massive Sulfide Deposits of the North American Cordillera Belt.
TABLE 9.26
Probabilities of Detection of Volcanogenic Massive Sulfide De-
posits of the North American Cordillera Belt by Ground Geo-
physical Surveys or Vertical Drilling to a Depth of 300 Feet.
probability of detection
Spacing S L.c.l. G. mean D.c.l.
in feet L = 1201 1842 2826ft
TABLE 9.27
Determination of Grid Size for Specified Levels of Probability in the
Detection of Volcanogenic Massive Sulfide Deposits of the North Amer-
ican Cordillera Belt by Vertical Drilling to a Depth of 300 Feet.
TABLE 9.28
Probabilities of Confirmed Detection of Volcanogenic Massive Sul-
fide Deposits of the North American Cordillera Belt by Ground
Geophysical Surveys or Vertical Drilling to a Depth of 300 Feet.
TABLE 9.30
Probabilities of Detection of Volcanogenic Massive Sulfide De-
posits of the North American Cordillera Belt by 55-Degree An-
gled Drilling to a Vertical Depth of 300 Feet.
Probability of detection
Spacing S L.c.1. G. mean U.c.1.
in feet L = 1201 1842 2826ft
TABLE 9.32
Probabilities of Confirmed Detection of Volcanogenic Massive Sul-
fide Deposits of the North American Cordillera Belt by 55-Degree
Angled Drilling to a Vertical Depth of 300 Feet.
TABLE 9.33
Optimal Designs for the Detection of Volcanogenic Massive Sulfide Deposits
of the North American Cordillera Belt by Drilling to a Vertical Depth of 300
Feet.
TABLE 9.34
Parameters of Multipurpose Optimal Grids for the Detection of Contact-Me-
tasomatic and Volcanogenic Massive Sulfide Deposits of the North American
Cordillera Belt.
Optimally angled
(50 degr.) drilling 600 to 700 ft $564,000 to 0.97 to 0.99
square grid $705,000
l. ALBERS, J. P., and ROBERTSON, J. E., 1961, Geology and ore deposits of the East Shasta
copper-zinc district, Shasta County, California, U.S. Geol. Surv. Prof. Pap. 338.
2. ANDERSON, C. A., 1969, Massive sulfide deposits and volcanism; Econ. Geol. 64(2), 129-146.
3. ANNIS, R. C., CRANSTONE, D. A., and VALLEE, M., 1981, A survey of known mineral
deposits in Canada that are not being mined in 1980, Mineral Bulletin No. 181, Department of
Energy, Mines and Resources, Ottawa.
4. BARAGAR, W. R. A., 1968, Major element geochemistry of the Noranda volcanic belt, Can.
J. Eanh Sci. S, 773-790.
5. BOLDY, J., 1977, (Un)certain exploration facts and figures, Can. Inst. Min. Metall. Bull.
70(781), 86-95.
6. CAMERON, E. M., 1975, Geochemical methods of exploration for massive sulfide mineral-
ization in the Canadian Shield, Geochemical Exploration 1974, pp. 21-49, Elsevier, Amsterdam.
280 CHAPTER NINE
7. DESCARREAUX, J., 1973, A petrochemical study of the Abitibi volcanic belt and its bearing
on the occurrences of massive sulfide ores, Can. Inst. Min. Metall. Bull. 66, 61-69.
8. EASDON, M. M., 1970, A compilation of graphitic occurrences in the Archean part of North-
west Quebec, Ministere des Richesses Naturelles, Quebec, Open File Report G.M. 25662.
9. FOX, J. S., 1977, Rapid pyroclastic mapping in base metal exploration, Can.lnst. Min. Metall.
Bull. 70, 173-178.
10. FOX, J. S., 1979, Host rock geochemistry and massive sulfide volcanogenic ore, Can. Inst.
Min. Metall. Bull. 72(804), 127-134.
II. GILMOUR, P., 1965, The origin of massive sulfide mineralization in the Noranda district,
Northwestern Quebec, Geol. Assoc. Canada Proc. 16, 1239-1249.
12. GOODWIN, A. M., 1967, Volcanic studies in the Timmins-Kirkland Lake Noranda region of
Ontario and Quebec, Geol. Surv. Can. Pap. 67-1, Pt. A, 138-142.
13. GRAF, J. L., 1977, Rare earth elements as hydrothermal tracers during the formation of massive
sulfide deposits in volcanic rocks, Econ. Geol. 72, 527-545.
14. HODGSON, C. J., and LYDON, J. W., 1977, Geological setting of volcanogenic massive
sulfide deposits and active hydrothermal systems: Some implications for exploration, Can. Inst.
Min. Metall. Bull. 70, 95-106.
15. HOLMES, R., and TOOMS, J. S., 1973, Dispersion from a submarine/exhalative body, in
Geochemical Exploration 1972, pp. 193-202, Institute of Mining and Metallurgy, London.
16. HOPWOOD, T. P., 1975, "Quartz-eye"-bearing porphyroidal rocks and volcanogenic massive
sulfide deposits, Econ. Geol. 71, 589-612.
17. HUTCHINSON, R. W., 1965, Genesis of Canadian massive sulfides reconsidered by comparison
to Cyprus deposits, Can. Inst. Min. Metall. Trans. 68, 286-300.
18. HUTCHINSON, R. W., RIDDLER, R. H., and SUFFEL, G. G., 1971, Metallogenic rela-
tionships in the Abitibi Belt, Canada: A model for Archean metallogeny, Can.lnst. Min. Metal/.
Bull. 64, 48-57.
19. HUTCHINSON, R. W., 1973, Volcanogenic sulfide deposits and their metallogenic signifi-
cance, Econ. Geol. 68, 1223-1246.
20. JENKS, W. F., 1971, Tectonic transport of massive sulfide deposits in submarine volcanic and
sedimentary host rocks, Econ. Geol. 66, 1215-1224.
21. KINKEL, A. R. Jr., 1956, Geology and base metal deposits of West Shasta copper-zinc district,
Shasta County, California, U.S. Geol. Surv. Prof. Pap. 285.
22. LARSON, L., and WEBBER, G. R., 1977, Chemical and petrographic variations in rhyolitic
zones in the Noranda area, Can. Inst. Min. Metall. Bull. 70(784), 80-90.
23. LATULIPPE, M., 1966, The relationship of mineralization to Precambrian stratigraphy in the
Mattagami and Val d'Or districts of Quebec, Geol. Assoc. Can., Spec. Pap. 3, pp. 21-42.
24. MACGECHAN, P. J., 1981, Exploration significance of the emplacement and genesis of massive
sulfide in the Main Zone at the Norita Mine, Mattagami, Quebec, Can. Inst. Min. Metall. Bull.,
84,59-75.
25. MILLER, L. J., 1960, Massive sulfide deposits in eugeosynclinal belts, Econ. Geol. 55,
1327-1332.
26. MILLER, R. J. M., 1973, The morphology of some Canadian massive sulfide deposits, A.I.M.E.,
Annual meeting, Chicago.
27. OHMOTO, H., 1978, Submarine caldera: A key to the formation of volcanogenic massive
sulfide deposits? Min. Geol. 28, 219-231.
28. PATERSON, N. R., 1971, Airborne electromagnetic methods applied to the search for sulfide
deposits, Can. Inst. Min. Metall. Bull. 64(705), 29-38.
SEARCH FOR VOLCANOGENIC MASSIVE SULFIDE DEPOSITS 281
29. PEARSON, D. E., 1977, Volcanic suites of southwestern British Columbia, Trip No 6, Geo-
logical Association of Canada Guidebook, pp. 20-25.
30. PELTON, W. H., 1977, New I.P. method may discriminate between graphite and massive
sulfides , North . Miner 63, 27, September 15.
31. PIRIE, J. D., and NICHOL, I., 1981, Geochemical dispersion in wallrocks associated with the
Norbec deposit, Noranda, Quebec, J. Geochem. Expl. 15, 159-180.
32. PODOLSKY, G., and SLANKIS, J., 1979, lzok Lake deposit, N.W.T., Canada: A geophysical
case history, Geol. Surv. Can. Rep. 31, 641-652.
33. RIDDLER, R. H., 1971, Analysis of Archean volcanic basins in the Canadian Shield using the
"exhalite" concept, Can. Inst. Min. Metall. Bull. 64(714), 20 (abstract).
34. RIDDLER, R. H., 1973, The "Exhalite" concept: A new tool for exploration, North. Miner
November 29, 59-61.
35. SANGSTER, D. F., 1972, Precambrian volcanogenic massive sulfide deposits in Canada: A
review, Geol. Surv. Can. Pap. No. 72-22.
36. SANGSTER, D. F., and SCOTT, S. D., 1976, Precambrian strata-bound massive Cu-Pb-Zn
sulfide ores of North America, in Handbook of Strata-bound and Stratiform Ore Deposits, pp.
129-222, Elsevier, Amsterdam.
37. SANGSTER, D. F., 1980, Quantitative characteristics of volcanogenic massive sulfide deposits,
Can. Inst. Min. Metall. Bull. 73(814),74-81.
38. SAWKINS, F. J., 1972, Sulfide ore deposits in relation to plate tectonics, J. Geol. 80, 377-397.
39. SEIGEL, H. 0., 1974, The magnetic induced polarization (M.I.P.) method, Geophysics 39,
321-339.
40. SOLOMON, M., 1976, "Volcanic" massive sulfide deposits and their host rocks: A review and
an explanation, in Handbook of Strata-bound and Stratiform Ore Deposits, pp. 21-54, Elsevier,
Amsterdam.
41. SPENCE, C. D., and DE ROSEN-SPENCE, A. F., 1975, The place of sulfide mineralization
in the volcanic sequence at Noranda, Quebec, Econ. Geol. 70, 90-101.
42. STANTON, R. L., 1960, General features of the comformable pyritic orebodies; Part 1: Field
associations, Part 2: Mineralogy, Can. Inst. Min. Metall. Bull. 53, 24-29, 66-74.
43. WALKER, G. P. L., 1971, Grain-size characteristics of pyroclastic deposits, J. Geol. 79,
696-714.
44. WOLFE,W. J., 1973, Geochemical exploration in Archean metavolcanic and meta-sedimentary
belts of the Pukaskwa Region, Ontario, Can. Inst. Min. Metall. Bull. 66(735), 75-83.
45. WOLFE, W. J., 1975, Zinc abundance in early Precambrian volcanic rocks; Its relationship to
exploitable levels of zinc in sulfide deposits of volcanogenic-exhalative origin, in Geochemical
Exploration 1974, pp. 261-278, Elsevier, Amsterdam.
CHAPTER TEN
In the Pine Point area of the Great Slave region, as many as 40 deposits
totalling some 100 million tons of lead-zinc ore have been discovered to date
within a roughly rectangular area extending in an east-west direction over a
length of 20 miles and a width of 3 miles. The stratigraphic setting is that of a
500-foot-thick middle Devonian carbonate formation with many vertical and
horizontal facies variations which belongs to a reefal complex of regional extent.
The carbonates are overlain by pelitic series (Buffalo River Shales) and underlain
by thick evaporites. The ore deposits are either pipelike, prismatic, or cylindrical
in shape and subvertical in attitude, or tabular and subhorizontal. The local
structural control of ore deposition appears to be a series of three parallel
southwesterly trending "hinges. "(4. 5. 7)
Farther west and north of the Great Slave region, the newly discovered
Selwyn Mountain lead-zinc district stretches along the eastern edge of the Cor-
dillera Belt into the Northwest Territories and the Yukon.(9. 10. 17) A number of
promising lead-zinc discoveries have been made and investigated in the Bonnet
Plume area of the Selwyn Mountains. So far the deposits showing the best
potential are those of Goz Creek and Cypress. The ore material consists of pyrite,
light-colored sphalerite, and galena with inclusions of boulangerite which boosts
the silver content of the mineralization. The sulfides occur in breccia zones that
extend along structural hinges in Lower Cambrian dolomite.
Because of the remoteness of the region and the rigor of the climate, more
than 10 years elapsed between the initial discoveries of promising lead-zinc
mineralization in the Arctic Islands, in the early 1960s, and the full appraisal
and development of the large Nanisivik and Polaris deposits. The Nanisivik
mineralization forms a tabular body of massive pyrite carrying varying amounts
of sphalerite and galena with attractive silver values occurring in Ordovician
dolomites. The Polaris ore consists of early marcasite followed by sphalerite and
galena in a gangue of vuggy and friable calcite with 5% ice content. On the
Greenland west coast, farther east, the large Black Angel deposit was investigated
in the late 1960s and is presently in production. The stratiform mineralization
is up to 60 feet thick and occurs in tremolite-talc marble of late Precambrian
age. The ore consists of massive sphalerite with about 20% pyrite and subordinate
galena low in silver.
Selwyn Mountains,(IO, 17) the Arctic Islands, and the Great Slave region.(4) In
well prospected districts such as the Upper Mississippi Valley zinc-lead area,
drill core or cuttings may be used to advantage as sampling medium. (I) The
sampling of residual soil to detect secondary halos has proved quite successful
in nonglaciated areas of moderate relief (Upper Mississippi Valley) but would
be of little assistance when dealing with glacial alluvium in the Arctic region.
Groundwater sampling for zinc and copper has been used successfully by the
first writer in the Upper Mississippi Valley area (see Chapter 12, Ref. 35), and
should prove useful in other regions of moderately well dissected topography
with temperate climate and rainfall above 32 inches per year.
TABLE 10.1
List of Mississippi Valley-type Pb-Zn Deposits of
the North American Arctic Paleozoic Platform In-
cluded in the Database.
TABLE 10.2
Summary of Statistical Modeling of Geometric Parameters of Mississippi Valley-
Type Pb-Zn Deposits of the North American Arctic Paleozoic Platform (Sample
Size: 14 Deposits).
Statistic of
observed data
The three distributions of observed dimensional data were fitted with log-
normal models. The fit was successfully tested at the 0.05 confidence level by
means of the X square test. The statistics summarizing the fitted models are listed
in the upper half of Table 10.2. The statistics include (a) geometric means and
the upper and lower limits of their 95% fiducial intervals, (b) 25th and 75th
percentiles, and (c) dispersion coefficients. The latter is constructed as the ratio
standard deviation/mean and is used to indicate in a dimensionless manner the
spread of the data about the mean of the fitted models as a reflection of the total
variability of the data. The variability of the length and breadth parameters is
rather small (coefficient near 0.1), but that of the shape coefficient R is much
larger (coefficient near 0.8).
10.4.1. Introduction
For detection purposes, the Mississippi Valley-type deposits are considered
as either vertical pipelike or tabular subhorizontal bodies with randomly orien-
tated horizontal sections of elliptical shape, lying entirely within the average
300-foot vertical range of most commonly used detectors. Therefore, no dip
component is added to the horizontal dimensions of the actual deposits in order
to obtain the target dimensions required for the calculation of detection proba-
bilities (see Section 5.4.2. of Chapter 5).
Only two parameters are required for the construction of the probability
tables, including the mean shape coefficient, and the mean target length within
its 95% fiducial limits. The latter enable the calculation of 95% confidence limits
for the detection probabilities.
TABLE 10.3
Summary of Detection Characteristics of Mississippi Valley-Type Pb-Zn Deposits of the
North American Arctic Paleozoic Platform for Three Types of Detection Strategies.
The reader will find in the following section a series of ten tables covering
the probabilities of detection and the optimization of the search for Mississippi
Valley-type Pb-Zn deposits of the Arctic by airborne and ground programs (pages
294--302).
TABLE 10.4
Probabilities of Detection of Mississippi Valley-Type Pb-Zn Deposits of the
North American Arctic Paleozoic Platform by Airborne Geophysical Surveys of
Parallel Grids.
Survey design: parallel lines with spacing S feet,
Randomly orientated elliptical targets with expected major axis = L feet
in the confidence interval: 1.c.1. ( geom. mean ( u.c.1.
and minor axis = B feet
R is defined as the ratio B/L with geometric mean = 0.40
Probability of detection
Grid spacing 1.c.1. geom. mean u.c.1.
S feet L= 927 1337 1927 feet
TABLE 10.5
Determination of Grid Size for Specified Probability Levels in the Detection of
Mississippi Valley-Type Pb-Zn Deposits of the North American Arctic Paleozoic
Platform by Airborne Geophysical Surveys on Parallel Grids.
TABLE 10.6
Probabilities of Confirmed Detection of Mississippi Valley-Type Pb-Zn De-
posits of the North American Arctic Paleozoic Platform by Airborne Geophysical
Surveys on Parallel Grids.
TABLE 10.7
Probabilities of Detection of Mississippi Valley-Type Pb-Zn Deposits of the
North American Arctic Paleozoic Platform by Airborne Geophysical Surveys on
Square Grids.
Probabili ty of detection
Grid spacing 1.c.1. geom. mean u.c.1.
S feet L= 927 1337 1927 feet
TABLE 10.8
Optimal Design of Airborne Geophysical Surveys for the Detection of Missis-
sippi Valley-Type Pb-Zn Deposits of the North American Arctic Paleozoic Plat-
form.
TABLE 10.9
Probabilities of Detection of Mississippi Valley-Type Pb-Zn
Deposits of the North American Arctic Paleozoic Platform by
Ground Geophysical Surveys of Vertical Drilling to a Depth of
300 Feet.
Probability of detection
Spacing S L.c.1. G. mean U.c.l.
in feet L = 927 1337 1927ft
TABLE 10.10
Determination of Grid Size for Specified Probability Levels in the Detection of
Mississippi Valley-Type Pb-Zn Deposits of the North American Arctic Paleozoic
Platform by Vertical Drilling to a Depth of 300 Feet.
TABLE 10.11
Probabilities of Confirmed Detection of Mississippi Valley-Type Pb-Zn
Deposits of the North American Arctic Paleozoic Platform by Ground
Geophysical Surveys of Vertical Drilling to a Depth of 300 Feet.
TABLE 10.12
Optimal Design of Ground Geophysical Surveys for the Detection of Missis-
sippi Valley-Type Pb-Zn Deposits of the North American Arctic Paleozoic Plat-
form.
TABLE 10.13
Optimal Design of Vertical Drilling Programs for the Detection of Mississippi
Valley-Type Pb-Zn Deposits of the North American Arctic Paleozoic Platform.
I. BARNES, H. L., and LAVERY, N. G., 1977, Use of primary dispersion for exploration of
Mississippi Valley-type deposits, J. Geochem. Expl. 8, 105-115.
2. CALLAHAN, W. H., 1964, Paleophysiographic premises for prospecting for strata-bound base
metal mineral deposits in carbonate rocks, Nev. Bur. Mines Rep. No. 13, Part C, 5-50.
3. CALLAHAN, W. H., and McMURRY, H. V., 1967, Geophysical exploration of Mississippi
Valley-Appalachian type strata-bound zinc-lead deposits, Geol. Surv. Can. Rep. No. 26, 350-360.
4. CAMERON, E. M., 1969, Regional geochemical study of Slave Point carbonates, Western
Canada, Can. J. Earth Sci. 6, 247-268.
5. CAMPBELL, N. 1957, Stratigraphy and structure of Pine Point area, Northwest Territories,
Canada, in Structural Geology o/Canadian Ore Deposits, Vol. 2, Canadian Institute of Mining
and Metallurgy, Montreal, pp. 161-174.
6. JACKSON, S. A., and BEALS, F. W., 1967, An aspect of sedimentary basin evolution: The
concentration of Mississippi Valley-type ore during the late stages of diagenesis, Can. Pet.
Geol. Bull. 15, 383-433.
7. JACKSON, S. A., and FOLINSBEE, R. E., 1969, The Pine Point lead-zinc deposits, N.W.T.,
Canada: Introduction and paleoecology of the Presqu'ile Reef, Econ. Geol. 64, 711-717.
8. LAJOIE, J. J., and KLEIN, J., 1979, Geophysical exploration at the Pine Point Mines Ltd.
zinc-lead property, N.W.T., Canada, Geol. Surv. Can. Econ. Geol. Rep. No. 31,653-664.
9. MACQUEEN, R. W., 1976, Sedimentary zinc-lead deposits, Rocky Mountain Belt, Canadian
Cordillera, Geosci. Can. 3(2) 71-81.
10. MURPHY, J. D., and SINCLAIR, W. D., 1974, Lead-zinc in the Selwyn and McKenzie
Mountains, Can. Min. J. 95(4),40-44.
II. NEWHOUSE, W. H., 1933, The temperature of formation of the Mississippi Valley lead-zinc
deposits, Econ. Geol. 28, 744-750.
12. OHLE, E. L., 1959, Some considerations in determining the origin of ore deposits of the
Mississippi Valley type, Econ. Geol. 54,769-789.
13. ROEDDER, E., 1967, Environment of deposition of stratiform (Mississippi Valley-type) ore
deposits from studies of fluid inclusions, Econ. Geol. Monograph 3, 349-361.
14. SEIGEL, H. 0.,1952, Ore body size determination in electrical prospecting, Geophysics 17(4),
907-914.
15. SEIGEL, H. 0., HILL, H. L., and BAIRD, J. G., 1968, Discovery case history of the Pyramid
ore bodies, Pine Point, N.W.T. Canada, Geophysics 33, 645-656.
16. SKINNER, B. J., 1967, Precipitation of Mississippi Valley-type ores: A possible mechanism,
Econ. Geol. Monograph 3, 363-369.
17. SMITH, C. L., 1974, A newly discovered Mississippi Valley-type lead province, Yukon-
N.W.T., Canada, A.I.M.E. Preprint No. 74-1-306.
CHAPTER ELEVEN
305
306 CHAPTER ELEVEN
tralia: 2.5%, the remainder (11%) being distributed unevenly between 20 coun-
tries throughout the world.
arcuate in shape; they are encased in granite-gneiss and migmatites and are cut
across by swarms of quartz (felspar) porphyry dykes and stocks.
The veins are generally tabular in shape: they are extensive along strike
and down dip but are very limited in their third dimension. They occur along
shears, fault zones, geological contacts, and in zones of crumpling and drag-
folding, within wide envelopes of hydrothennal alteration consisting mainly of
silicification and carbonatization.
Two main classes of vein gold deposits are generally recognized: (1) cavity-
filling and (2) replacement, while in many deposits the two types occur together.
The cavity-filling type has well-defined walls and the gold mineralization is
restricted to the vein material. In the replacement type, however, the minerali-
zation replaces the wallrock and vein material as well, showing ill-defined bound-
aries. The Kirkland Lake "Break" of northeastern Ontario and the Homestake
deposit of the Black Hills are good examples of the replacement type of vein
gold deposit. The replacement veins are generally much longer (several thousand
feet) and wider (up to 100 feet) than the cavity-filling veins, which are only
2000-3000 feet in length and 15-30 feet in width at the most.
The Porcupine Gold Fields of northeastern Ontario, being the most pro-
ductive and the most intensely investigated gold district of the Canadian Shield,
will serve as an example for the study of the structural control of gold deposition
in the North American Shield. The Porcupine structural setting is that of an
overturned, steeply dipping, tightly folded synclinal structure made up of intri-
cately dragfolded Keewatin basic flows uncomfonnably overlain by Temiskam-
ing greywacke and associated clastic sediments. The Porcupine Trough is cut
across by swarms of quartz (felspar) porphyry dykes and pipelike stocks of
Algoman age. The principal locus of the ore is the Pearl Lake porphyry stock,
which is most intensely sheared and hydrothennally altered, almost beyond
recognition. The structural controls of the deposits are as follows: (a) openings
developed along zones of weakness related to the stratigraphic succession, such
as flow contacts, bedding planes, and also along intrusive contacts; (b) openings
developed by differential movement affecting rocks of differing competencies;
basic flows, conglomerate and greywacke beds and porphyry intrusive behave
as competent fonnations in contrast with tuffs, slate, soapstone, chlorite-
carbonate schist, and pillowed lava flows.
The mineralogy of the deposits is rather simple. The gangue material consists
mainly of quartz with subordinate amounts of carbonates, mainly ankerite and,
locally, tounnaline (Porcupine), accompanied by minor chlorite, sericite, and
graphite, as well as sporadic sulfides. Among the sulfides, pyrite is the most
common and abundant, particularly in the replacement-type deposits; it is fol-
lowed in order of decreasing importance by arsenopyrite, pyrrhotite, and base
308 CHAPTER ELEVEN
metal sulfides. The latter may be quite abundant (Western Ontario), requiring
expensive smelting of the "refractory" ore. Likewise, arsenopyrite is troublesome
as it requires a costly roasting prior to the cyanadation process. In most deposits
of the North American Shield, a large proportion of the gold is associated with
the pyrite and is largely "free milling," occurring as small grains of native gold
or electrum along crystal boundaries, cleavages, or fracture planes. The rest
occurs as tellurides.
aeromagnetic surveys have greatly assisted the delineation of high priority areas
in the newly discovered Hemlo Gold District of Ontario.
Visual detection from the ground has proved highly effective over the past
50 years. Up to 90% of the gold deposits discovered to date in the North American
Shield were detected by a combination of prospecting, systematic mapping, and
some information drilling. In the present circumstances of high gold price, the
systematic drilling of small areas of high potential of less than a square mile in
extent should be considered in the near future as a valid and effective approach,
since the most obvious deposits have already been discovered.
assistance could be gained from drill core sampling in well prospected areas.
Geochemical detection of secondary halos in residual soils is not applicable in
most regions of the North American Shield because of the extent and thickness
of transported glacial alluvium.
TABLE 11.1
List of Vein-Gold Deposits of the North American Shield Included
in the Database.
Ontario
Central Patricia Hallnor
Cochenour Willans Dome
Gold Eagle Preston East Dome
Madsen Red Lake Paymaster
McKenzie Red Lake Buffalo Ankerite
Hasaga Aunor
Little Long Lac Delnite
Jellicoe Coniaurum
Tombill Moneta Porcupine
North Empire Hollinger
Hard Rock McIntyre
Renabie Kerr Addison
Jerome Sylvanite
Leitch Wright Hargeaves
Pamour Bidgood
Broulan
N. l'i. Quebec
Omega Arntfield
Lamaque Camflo
East Malartic Canadian Malartic
Central Cadillac Granada
O'Brien Agnico
Belleterre
Northwest Territories
MacWatters
Powell Rouyn Giant Yellowknife
Wasa Cons. Discovery
Beattie Camlaren
Francoeur
TABLE 11.2
Statistical Modeling Summary for Vein-Gold Deposits of the North American
Shield (Sample Size: 50 Deposits).
Statistic
of observed
data
arithmetic mean of normal model, and circular mean of circular normal model,
along with the limits of their 95% confidence intervals. Other listed statistics
include the first and second quartiles of the fitted distributions and the dispersion
coefficients, which give an indication of the spread of the data about the mean
of the fitted models. A small coefficient (0.1) indicating a low variability is
shown for the length parameter, while a moderate variability affects the breadth
and shape ratio parameters (coefficient ranging from 0.1 to 0.5).
11.4.1. Introduction
For the purpose of calculating detection probabilities, the vein-gold deposits
of the North American Shield are considered as dipping, orientated tabular bodies
with horizontal cross sections of elliptical shape. Therefore, target dimensions
DETECTION OF VEIN-GOLD DEPOSITS 313
for detection purposes are not the same as that of the deposits; they have to be
recalculated for each deposit by introducing a dip component which is computed
as described in Sections 2.2.3. of Chapter 2, and 5.4.2 of Chapter 5. As a result
of the recalculation, the shape coefficients are substantially inflated, thus in-
creasing the probabilities of detection. For example, the mean shape ratio of the
horizontal sections of vein gold deposits themselves (0.016) becomes equal to
0.09 for the targets in the context of vertical detection and is increased to 0.19
when considering 55-degree angled detection.
Since strike orientation data were available for most of the 50 deposits of
the sample, the orientation factor is introduced in the construction of the tables
to improve the probability of success of the various field programs. Based on
the results of a dynamic programming study, it was found that the probabilities
of vertical detection by geophysical surveys or drilling are maximized if the grid
is laid at angle of 18 degrees ± 10 with the expected strike direction. If we
consider the 55-degree angled detection, the optimal grid orientation is at 30
degrees ± 10 with the expected strike direction.
The latter may be estimated from the results of previous surveys or from
photogeological studies. If no prior information is available, the mean of the
fitted circular normal model is a statistically valid estimate to be considered for
grid design (See Table 11.2, upper half).
surveys or vertical drilling to a depth of 300 feet. The second block covers the
55-degree angled detection by drilling (Tables 11.8-11.10).
55-degree
angled $424,200 $1,090,000 $536,000 0.90 $1,160,000 $3,490,000
drilling 800 ft 500 ft 700 ft 440 ft 300 ft
square grid
Optimal 45-deg
angled - - $491,000 0.88 - -
drilling 800 ft
square grid
I
W
...I.
I.n
316 CHAPTER ELEVEN
The following section includes eight tables which cover the probabilities
of detection and optimal designs of ground programs in the search for vein-gold
deposits of the Shield (pages 317-323).
DETECTION OF VEIN-GOLD DEPOSITS 317
TABLE 11.4
Probabilities of Detection of Vein-Gold Deposits of the North
American Shield by Ground Geophysical Surveys or Vertical Drill-
ing to a Depth of 300 Feet.
Probabili ty of detection
Spacing S L.c.l. G. mean U.c.l.
in feet L = 1466 1753 2097ft
TABLE 11.5
Determination of Grid Size for Specified Levels of Probability in the
Detection of Vein-Gold Deposits of the North American Shield by Ver-
tical Drilling to a Depth of 300 Feet.
TABLE 11.7
Optimal Design of Ground Geophysical Surveys for the Detection of Vein-Gold
Deposits of the North American Shield.
TABLE 11.8
Probabilities of Detection of Vein-Gold Deposits of the North
American Shield by 55-Degree Angled Drilling to a Vertical Depth
of 300 Feet.
probability of detection
Spacing S L.c.1. G. mean U .c.1.
in feet L = 1466 1753 2097 ft
TABLE 11.9
Determination of Grid Size for Specified Levels of Probability in the
Detection of Vein-Gold Deposits of the North American Shield by 55-
Degree Angled Drilling to a Vertical Depth of 300 Feet.
TABLE 11.10
Probabilities of Confirmed Detection of Vein-Gold Deposits of
the North American Shield by 55-Degree Angled Drilling to a
Vertical Depth of 300 Feet.
TABLE 11.11
Optimal Designs for the Detection of Vein-Gold Deposits of the North Amer-
ican Shield by Drilling to a Vertical Depth of 300 Feet.
1. AMIRYAN, S. 0., 1960, Mineralogy of gold deposits, Akad. Nauk. Arm. S.S.R. 31(1), 43--48.
2. BOYLE, R. W., 1961, The geology, geochemistry and origin of the gold deposits of the
Yellowknife District, N.W.T., Geol. Surv. Can. Mem. No. 310.
3. COOKE, H. C., and JOHNSTONE, W. A., 1932, Gold occurrences of Canada, a summary
account, Geol. Surv. Can. Econ. Geol. Ser. No. 10.
4. DOUGHERTY, E. Y., 1939, Some geological features of the Kolar, Porcupine, and Kirkland
Lake gold fields, Econ. Geol., 34, 622-653.
5. ELEVATORSKI, E. A., 1981, Gold Mines of the World, MINOBRAS, Santa Monica, Cali-
fornia.
6. EMMONS, W. H., 1937, Gold Deposits of the World, McGraw-Hili, New York.
7. HUTSON, R. J., 1980, Exploration models for gold in the Archean, Can. Min. J. 101(4),
66-68.
324 CHAPTER ELEVEN
8. KELLY, S. F., 1957, Resistivity and magnetic surveys in 1936 on the Brou1an-Porcupine gold
prospect, South Porcupine, Ontario, Canada, in Methods and Case Histories in Mining Geo-
physics, Sixth Commonwealth Mining and Metallurgy Congress Volume, Montreal, pp. 283-291.
9. MIDDLETON, R. S., and CAMPBELL, E. E., 1979, Geophysical and geochemical methods
for mapping gold-bearing structures in Nicaragua, Geol. Surv. Can. Econ. Geol. Rep. No. 31,
pp. 779-798.
10. MOORE, E. S., 1940, Genetic relations of gold deposits and igneous rocks in the Canadian
Shield, Econ. Geol. 35, 127-139.
11. SEIGEL, H. 0., JOHNSON, I., and HENNESSEY, J., 1984, Geophysical aids to gold ex-
ploration, Northern Miner March 1st, BI7-BI8.
12. SIXTH COMMONWEALTH MINING AND METALLURGY CONGRESS VOLUME, 1957,
Structural geology of Canadian ore deposits, Canadian Institute of Mining and Metallurgy,
Montreal, Canada.
13. WHITE, W. H., 1943, Mechanism and environment of gold deposition in veins, Econ. Geol.
38, 512-532.
14. WRIGHT, L. B., 1937, Gold deposits in the Black Hills of South Dakota and Wyoming,
A.l.M.E. Trans. 126. 390-425.
CHAPTER TWELVE
12.1.1. Introduction
As outlined in Section 1.2.3 of Chapter 1, the mineral exploration sequence
starts with preliminary planning followed by field planning of surveys and data
processing, and ending with the assessment of results for further decision making.
The initial planning requires many decisions of critical importance, such as the
choice of type(s) of targeted ore deposits, selection of regions for the search and
prospecting areas within the selected regions, and finally the choice of meth-
odology of coverage of prospecting areas, including technology, and particularly,
survey design, whose optimization is the main topic of the book.
The following stage, field planning, deals largely with the scheduling and
logistics required to organize and implement in an optimal manner the field
programs that were planned in the previous stage. The purpose of the operation
is the acquisition of new data, which are assembled and processed in order to
delineate exploration targets for detailed investigation. Most field programs gen-
erate a large number of exploration targets of diverse merits from which only a
limited number are to be selected for drill testing, because of time and budget
constraints. The optimization of the screening of exploration targets is the topic
of the present chapter.
325
326 CHAPTER TWELVE
Agterberg(1) and Culbert (7) almost a decade ago. Instead of considering each
type of measurement independently from the others, the multivariate approach
deals with "statistical patterns" based on all types of measurements at each
sampling point. The spatial characteristics of the patterns are analyzed statistically
and the targets are delineated in an objective manner.
t ~ [];J
surveys
-{
data storage
Economic data
1
. ( Exploration cost
on ore deposits \ data
Continuously recorded
quantitative data
( Geo-environment data
. L
••
(airborne surveys) on ore deposits
•• -
( ( ")
f\
Cluantitative Gualitative
I
J1
discrete data geological
I
(ground surveys) data
\ )
I Digitization
I I
Quantification
I Dichotonization
into binary data
I
+
1_ Data compression
by Prj~cipal component
analySIS
I Data compression
by Characteristic
analysis
1
I
Ouantification
~
-.. (i--"-----
<Selection of
dependent '\ \
Selection of
independent
-'-"\
)
gee-variates " . ___ ~-vanates ~/
:
1
Stepwise regression \
analysis
- !
~~
\
®
data
storage
signific.ant
geo-varlates
Geo-statisticaJ Model
........
Significant
geo-variates
( Cost function model) data from all
Prior probability
Bayesian classification model exploration
of success
targets
>-
!i!~
11 E
2) D
3) A
(10.65)
(10.20)
(8.87)
I
(3) RANKING OF TYPICALITY OF CHARACTERISTICS
!g 4) G (8.45)-CUT-OfF
~~ 5) F (5.82)
6) C (5.42)
FIGURE 12.2
Flow Diagram of Characteristic Analysis.
(1)
where y is the dependent variate, and Qi and Xi the coefficients and independent
variates representing the geo-factors i = 2 . . . . n. The theory of linear regres-
332 CHAPTER TWELVE
sion requires that the variate y be normally distributed. This can be easily attained
for y if we choose the logarithm of y as dependent variate because the frequency
distributions of most economic parameters are generally lognormal in nature. A
simple logarithmic transformation written as log (1 + x;) is usually sufficient
to "normalize" all independent variates with very skewed distributions. As a
result the most general regression equation is written as
FIGURE 12.3
Flow Diagram of Statistical Screening of Independent
Geovariates.
334 CHAPTER TWElVE
f. (X)
1
[(2TI)P Islr1/2 exp[-l/2 eX - x.)
1
S-1 (X - X.)']
1
s = variance-covariance matrix
Isl= determinant of S
eX - X.)'
1
= transpose
FIGURE 12.4
Model of a Bayesian Classification Function.
success) under cost constraints should be much more satisfactory. The payoff
criterion which is described in Figure 12.5 fulfills these requirements. Payoff
values obtained by feeding into the model the gross economic estimates, prob-
abilities of success, and cost data for all targets under study are ranked in order
of decreasing magnitude to allow the selection of a small number of the top-
ranking targets, subject to budget constraints.
w
W
0'1
Statistically significant ~
geo-variates
~"
(geological, geophysical, geochemical) .
l
Expected gross Probability that Probability of Pay-off criterion
Expected
dollar value of the target is a failure (target is for the optimal
exploration
exploration target x commercial not an ore deposita) X :. selection of
cost
(GVI deposit (Psi Pf = 1 - Ps exploration targets
n
• • • • I
Expected monetary gain Expected loss »-0
-(
m
;;:0
FIGURE 12.5 -(
Design of a Payoff Model for the Optimal Selection of Exploration Targets for Drill Testing. :E
m
r-
<
m
SELECTION OF EXPLORATION TARGETS 337
iiM
Montfort ,.,."
31
31e Locotion of onomoly ossocioted with .,.
known zinc deposit
'.e,o
.25 .28 3"- 33 _......3.
o 5 10 21e
.24.
. . . .--.
e3I .45 !,,,,,,I 3
I
scale of miles 2& Mineral Point'~".'.'. .4
8 .40 42 ·44
I- IOWA CO
It."town .21 ~ 3!jW LAFAYETTE CO
~.23 a:
L C)
.16 ,";".iBelmont
• • 58
.11 51
I"\~~
Tennysoh'"
FIGURE 12.6
Selection of Exploration Targets in the Upper Mississippi Valley Zinc District:
Control Locations.
043042
44
Dod~ev!y'.~ 0045 046
~~~ 30
0 49 o 47 0
0 31
500 0 48
Mineral Point 06
~~l~ 07
IOWA CO
4
0 ...5
36 If Blanchardville
038 ..
LAFAYETTE CO
Belmont
m?t:
0 27
0 21
0 53 290 ~8
0 52 0 13
0 25 0 26
0 14
12 0 Location and reference number of anomaly
in study area
o 5 10
scale J mu.. '
FIGURE 12.7
Selection of Exploration Targets in the Upper Mississippi Valley Zinc District:
location of Exploration Targets to be Rated.
shallow lead mineralization and deeper zinc mineralization were previously de-
tected by trenching and drilling.
.82°·22
Dodgevil!e_ .98.
li\~
""""
C>
32
°05.08 0'
16
0.29 0·
c>07
.08° 0·28
FIGURE 12.8
Selection of Exploration Targets in the Upper Mississippi Valley Zinc District:
Location of Exploration Targets Selected on the Basis of Bayesian Probabilities
of Success.
leading to the selection of targets for drill testing. The advantages of the statistical
procedure are fourfold: The method provides (I) an objective rating of the selected
targets, (2) the confidence intervals associated with the rating, (3) the probabilities
involved, and (4) an accurate and speedy way of processing large amounts of
data, and the possibility of updating the planning based on feedback information,
as the field program progresses.
340 CHAPTER TWELVE
1
Prior classification
analysis on each Trend and trend factor
significant factor analysis and significance
and variate for each survey and factor.
The dependent variates are
the surveys and factors. and
the independent are the
geographic, geologic and
topographic variates
Probability analysis
1
for each location in Residual trend analysis to
each group: anomalies determine local anomalies
typos 1.2 •...• and significantly above or below
background the trend
FIGURE 12.9
Flow Diagram of the Optimal Selection of Exploration Targets Without Control
Locations.
where the dimensions of the matrices are given in brackets, and where B is the
matrix of the best coefficients.
Once all coefficients are found, the data matrices Y and X are called again
and read to calculate the trend estimates at each location for each of the survey
variates. The analysis is repeated to find trend values for each factor at each
location. The main output from TREND is
1.The coefficient for each trend equation for each variate, and for each
factor;
ii. The trend and trend factor estimates for each location.
12.3.8. Summary-Conclusion
Since we do not have the advantage of having prior control locations to
rely on for the selection of locations of greatest merit in the prospecting area,
we proceed to construct our own set of control locations within the prospecting
area itself by means of the preclassification analysis and factor analysis. As was
the case in the first approach, we use the control set to obtain a probabilistic
classification of the locations under investigation. Their screening is further
refined in a spatial context by means of trend analysis, through which local
anomalies are separated from the global ones. The local anomalies are ranked
in order of merit, based on the magnitude of the associated residuals. It should
be noted that the second approach described here involves a sequential max-
imization, first of the probabilities of belonging to classes 1 and 2, secondly of
SELECTION OF EXPLORATION TARGETS 345
The inner group of anomalies in the northwestern aureole of Mount Bundey have
since been drilled, and the result was a new silver mine grading up to 40 oz. to
the ton being discovered. The authors have not had any reports on the other
anomalous prospects as to whether any drilling program has been carried out.
The silver discovery was classical: the anomaly in the Mount Bundey aureole
is a radiometric high coinciding with a magnetomic low.
18. PETERS, W. c., 1978, Exploration and Mining Geology. Wiley, New York.
19. WIGNALL, T. K., 1969, Generalized Bayesian classification functions (K classes); Application
to a regional geochemical survey in southeastern Pennsylvania, Econ. Geol. 64, 570-576.
49. NEEDHAM, R. S., WILKES,P. G., SMART, P. G., and WATCHMAN, A. L., 1973,Alligator
Rivers region of the Northern Territory: Results of airborne geophysical surveys conducted by
the Bureau of Mineral Resources of Australia; B.M.R. record 1973-208.
50. RAO, C. R., 1952, Advanced Statistical Methods in Biometric Research, Wiley, New York.
5!. WIGNALL, T. K., 1969, Generalized classification functions: A Bayes-theoretic approach,
Annual meeting of the Northwestern Science Association, Cheney, East Washington State
College.
CONCLUSION
351
352 CONCLUSION
353
SUBJECT INDEX
355
356 SUBJECT INDEX
Derivative, 84, 85, 90, 91 Exploration target, 10, 11, 325-327, 334,
Detection, 17-21,26-28,32,33,43,48 335, 340, 344
range, 21, 27-29 Extremal root, 85, 91
sensitivity, 48, 49 Extremum, 81, 82
Deterministic, 22, 31, 84
Dichotomization, 329, 330, 340, 342 Factor analysis, 326, 330, 342, 343
Digitization, 326, 329, 340 Feature selection, 330, 332, 333
Dip Fiducial interval (see also confidence
angle, 22-25, 27-29, 31, 100, 149,216, interval), 99, 102, 104, 105, 107,249,
217,242,289,310 267,279
component, 27-29,100,101,117,151, Field
218,245,290,312 artificial, 20, 28, 29
Direct approach, 90--92 potential, 28, 29
Direct ascent, 85 Filtering, 20
Direct detection, 17, 112, 146,213,214, Flight elevation, 28, 34, 35
239, 286, 308, 309 Frequency distribution, 98-100, 115, 150,
Discriminant analysis, 332, 334 163,217,243,289,310,311
Discrimination, 21, 22 Frequentist concept, 3
Dolomitization, 284, 285
Drilling, 17, 18,66-68,72,73,91 Geochemical
angle, 91, 92, 107,201-203, detection, 18,33, 113, 114, 147,215,
220--222,246-249,265-267,314, 287,288, 309, 310
316 signature, 18, 19,325
churn, 66, 286 Geographic areas
diamond, 51, 66 Central Asia, 147
optimization, 7, 11, 91, 92, 106, 107, Greenland, 284, 285
123, 124, 200, 202, 203, 222, 248, Manitoba, 211-213, 216, 237, 238
249,267,293,316 Mexico, 110, 146, 177
percussion, 66, 138, 302 Mississippi Valley, 283, 284, 337, 338
testing, 7, 10, 11, 325, 327, 339 Missouri, 283, 284
Dynamic programming, 9-11, 86, 87, 89, Ontario, 30, 34, 35, 61, 211-213,
91, 105-107,218,222,248,267,313, 237-239, 306, 309
316 Quebec, 4, 30, 212, 238, 239, 241
Selwyn Mountains, 162, 189, 283-285,
Economic worth, 1, 5, 9, 11, 24, 96, 288, 306
329-331,334,340 Southwest U.S.A., 110, 111, 145, 148,
Effectiveness, 78, 88 162, 176, 189, 190,238
Efficiency, 88 Tri-State, 283, 284
criterion, 78, 79, 89, 90, 105, 122, 201, Ungava, 212, 213, 215
222, 248, 293, 316 Geological age, 110, 145,211,235,283,
Electromagnetic survey, 20, 21, 28, 68, 113, 284,305
214,215,240,287,309 Geological detection, 17, 18, 112, 146,213,
ELLIPGRID program, 32, 33, 43, 53 214, 238, 239, 286, 308, 309
Exhalite, 236, 237, 239 Geological occurrence, I, 4, 11, 13, 14
Expected loss, 79 Geometric descriptor, 22, 23
criterion, 123, 203, 222, 248, 293, 316 Geometric factor, 24
Expected monetary value, 79, 334, 336 Geometric probability, 31-33
Expenditure, 56 Geophysical contrast, 29, 30, 61-65
SUBJECT INDEX 357
Shield region, 61, 96, 97, 2ll, 214, 215, Trend analysis, 326, 343, 344
235, 238, 306, 307 Threshold, 104,326,334,337,342
Signal, 19, 20 Trial repetition,S, 6
Signature, 19, 21, 326
Simpson's rule, 44, 45 Uncertainty, 2, 3, 5, 8, 10, 100
Skam, 141-143, 148, 149, 176, 189, 190, Unit costing, 56-58, 60, 65, 66
305
Variability, 99, 116, 150, 151, 165, 217,
Skimming, 8
218,289,290,312,342,343
Spatial coincidence, 30, 326
Variance, 90, 332, 335
Standard deviation, 99, ll5-ll7, 150, 165,
Variate, 81, 82, 84, 85, 328-331, 333, 342,
217,243,289, 3ll, 326
343
Statistical decision theory, 80, 81
Vein gold deposit
Statistical occurrence, 3-5, 79
detection, 308-310
Statistical pattern, 326, 327
economic geology, 305, 306
Statistical sampling, 98, 288
geological setting, 307, 308
Statistical shape, 99, ll5, 150, 151, 165,
stastical modelling, 310-312
217,243,289,310,311
Vertical continuation, 29, 30
Statistical summarization, 98, 99
Vertical detection, 27-29
Stochastic variate, 84, 325
Volcanogenic massive sulfide deposit,
Stratabound deposit, 235-237, 283
235-238
Strategy, 80, 81, 104, 105, 118, 121-123,
alteration pipe, 236, 237
198,201,202,219,220,222,246,
calc-alkaline series, 236
248,265,267,291,293,314,316
dalmatianite, 237
Stratifonn deposit, 235, 283
detection, 238-240
Strike direction, 23, 90-92, 105, 106, 148,
exhalite, 236, 239
217,243,244,286,311
geological setting, 236, 237, 261, 262
Structural hinge, 285
massive ore, 237
Structural setting, 109, 110, 144, 145, 211,
metallogeny, 237, 238
235, 238, 285, 307
occurrence mode, 238
Student-t test, 99, 344
ophiolitic type, 235
Success, I, 3-5, 88
pyroclastic series, 236, 262
ratio, 4, 329, 333, 335
stringer ore, 236, 238
Supergene enrichment, III, 176, 237
tholeiitic series, 236
Tactite, 189, 190 Wholesale Price Index, 57
Target Word Processor, 97
definition, 326
geometry, 27, 28, 31-35, 42, 43, 100, Yellow Pine District, 190
101 Yukon Territories, 110, 145, 161, 162, 176,
selection, 2, 325, 327, 328, 340, 341 189
Telluride, 307
Tintic District, 162-164 Zoning, 109, 1l0, 141-145, 148, 189, 190
AUTHOR INDEX
361
362 AUTHOR INDEX
Shurygin, A. M., 33, 42, Smirnov, D. F., 141, 142, Wilde, D. J., 84
77 175 Wilson, H. D. B., 212
SiIlitoe, R. H., 110, 114 Smith, C. L., 285, 288 Wright, L. B., 306
Sinclair, A. J., 33 Solomon, H., 32
Singer, D. A., 32, 33, 42, Spector, A., 29 Young, G. A., 148
43
Slichter, L. B., v, 32, 34, White, L. G., 189 Zurbrigg, H. F., 212
77,78,80, 141 Wignall, T. K., 327 Zurflueh, E. G., 21
LISTED JOURNALS
363
364 LISTED JOURNALS
Photogrammetric Engineering, 52
Quarterly of Colorado School of Mines, 14, 53
Science, 13
Transactions of Institute of Mining and Metallurgy, London, 13,93,347
Western Miner, 209, 346