Anda di halaman 1dari 65

WORKSHOP NOTEBOOK

Application of Structural Geology


in Mineral Exploration and Production

Presented by
Dr. Eric P. Nelson
Department of Geology and Geological Engineering
Colorado School of Mines

For
Newmont Mining Corp.
Yanacocha Mine, Peru
7-9 September, 2011
Table of Contents
Table of Contents ................................................................................................................ 2 
Introduction ....................................................................................................................... 3 
Importance of structural geology to mineral exploration ............................................... 4 
Tectonics of Convergent Plate Margins .......................................................................... 5 
Convergent margin stress-strain environment ................................................................ 6 
Review of Structural Principles ..................................................................................... 10 
Stress ............................................................................................................................. 10 
Strain ............................................................................................................................. 13 
Description and classification of structures .................................................................. 15 
Fractures: faults and extension fractures................................................................... 15 
Folds.......................................................................................................................... 19 
Structural fabrics ....................................................................................................... 21 
Structural Permeability ................................................................................................. 21 
Structural Methods in Mineral Exploration ................................................................ 25 
Structural data collection - introduction ....................................................................... 25 
Structural orientation data ............................................................................................. 26 
Structural descriptive data............................................................................................. 29 
Structural timing ....................................................................................................... 29 
Structural geometry ................................................................................................... 30 
Structural kinematics ................................................................................................ 30 
Structural analysis methods .......................................................................................... 31 
Analysis of structural data on spherical projections ................................................. 31 
Analysis of slickenline data ...................................................................................... 33 
Paleo-stress and strain modeling ............................................................................... 34 
Structural pattern recognition ................................................................................... 36 
Analysis of longitudinal sections of fault-veins........................................................ 36 
Drill hole targeting and design ...................................................................................... 39 
Structural data in drill core ........................................................................................... 43 
Un-oriented core ....................................................................................................... 44 
Oriented drill core ..................................................................................................... 46 
Structural Methods in Mineral Production.................................................................. 49 
Characterization of rock mass structure ........................................................................ 49 
Rock mass classification systems ................................................................................. 50 
Types of discontinuities and structures ..................................................................... 50 
Characterizing structural discontinuities................................................................... 50 
Discontinuity spacing and block size ........................................................................ 51 
Rock Mass Classifications ........................................................................................ 52 
Slope stability analysis .................................................................................................. 54 
Structure of High-sulfidation Epithermal Deposits ................................................... 59 
References cited ................................................................................................................ 61 
Appendix I –Structural database format ........................................................................... 63 
Appendix II – Rake to plunge/trend calculation ............................................................... 64 
Appendix III – Calculation of drill target location ........................................................... 65 
Good exploration, like good science, is based on model testing.

Introduction
The utility of structural geology in mineral exploration and production is twofold: 1) to
advance the understanding of structural controls on mineralization and ore distribution
for assistance in exploration at all scales, from mine to regional, and 2) to characterize the
rock mass structure at mine scale for use in geotechnical analysis of slope stability,
hydrogeology, and mine facility design (infrastructure, tailings and waste sites, etc.).

Most ore deposits have important structural controls, and formed in tectonic provinces
undergoing active crustal deformation. Therefore, collection and analysis of structural
data are important tasks in nearly all exploration projects of any scale (mine, district, or
regional). This course in Application of Structural Geology in Mineral Exploration
focuses on methods useful for mineral exploration in the tectonic environments of
convergent plate margins, in which plutono-volcanic activity is superimposed on active
crustal deformation. This is the primary geological environment of Mesozoic-Cenozoic
ore deposit formation in the Cordillera of the Americas, as well as in many other
geological periods and areas of the world.

This workshop is presented in three parts: 1) description of the convergent margin


tectonic environment, 2) review of structural principles, and 3) presentation of structural
methods useful in mineral exploration. The most common types of deposits in the
convergent margin tectonic environment are shear zone- and fault-controlled deposits
(epithermal veins, mesothermal [orogenic] deposits, etc.), many iron-oxide-Cu-Au
(IOCG) systems, replacement deposits including skarn, and porphyry and related high-
sulfidation deposits, all of which are structurally-controlled. The structural controls and
methods useful in exploration for these deposits are emphasized. The discipline of
structural geology and its application in mineral exploration and production is much
broader, and some methods and concepts are not covered in detail in this course.
However, some structural methods used in mining are included, particularly the
characterization of rock mass structure for
geotechnical purposes.

Good exploration, like good science, is based on


model testing. This theme is emphasized
throughout this workshop. Many structural
models are presented, and the exploration
geologist should have knowledge of such
models to draw on during interpretation of
structural data and patterns. In any exploration
program, structural models should evolve and
be improved through testing (e.g., drilling,
further data collection and analysis, geophysics,
etc.).

Figure 1. Early published drawing of veins


(Agricola, 1556).
3
Importance of structural geology to mineral exploration
As noted by Guilbert and Park (1986) “Detailed studies of structure are essential in
exploration, and they unquestionably have led to more discoveries of ore than any other
approach.” Knowledge that structures host metal deposits is very old (ancient Greeks,
Fig. 1), and modern mineral exploration has been aided through advances, in the past 30
years, in the understanding of rock deformation, methods of structural geology, and many
related fields. Such fields include plate tectonics context, strain studies, cross section
construction and balancing (testing), tectonics processes and regimes (extensional,
shortening, wrench), mechanical modeling, computer visualization and modeling, and
GIS technology.

Structural geology is important in mineral exploration and production in three main ways:
1. The location and distribution of mineralized rock is related to structure and tectonics:
o Tectonics control processes related to host rock formation (magmatism,
volcanism, sedimentation, metamorphism, diastrophism).
o Structures control hydrothermal fluid flow during mineralization, and thus the
location and geometry of mineralized rock.
o Post-mineralization deformation can affect the location and geometry of ore
deposits (varies from brittle faulting to ductile folding, shearing,
metamorphism and tectonite fabric development).
2. Structural methods are used to define the 3-D form, arrangement, and internal
structure of rock units that may host and/or obscure mineralization.
3. Structural methods are used in mining engineering to assess rock strength and internal
rock-mass structure.

Applying structural geology in mineral exploration and production requires knowledge of


methods and skills available in a “structural toolkit”. The keys to applying these methods
to mineral exploration and production are:
 Careful, focused structural mapping and data collection
 Recognition of structures and structural associations, and knowledge of their
importance
 Synthesis, analysis, and interpretation of structural data
 Development of structural-mineralization models based on data analysis,
modeling, and comparison with similar (analog) deposits and tectonic settings

Effective exploration results in discovery of a profitably mineable resource. Time


efficiency and financial efficiency must be maintained, although not at the cost of good
geological work. The exploration geologist therefore must know what methods are
available, and when a particular method should be employed or not. Ineffective
exploration can result from poor mapping, failure to recognize a structure or structural
event, or not collecting the correct structural data or using structural data to its full
potential.

4
The level of exploration advancement determines the types of structural/tectonic analyses
used:
1. Grassroots: tectonic analyses, geological map interpretation, reconnaissance
mapping, and regional section construction,.
2. Regional: tectonic analyses, geological map interpretation, structural compilations,
reconnaissance mapping, section construction,
3. District: detailed mapping, section construction and balancing, analysis of structures,
structural compilation, drill campaign and target design.
4. Deposit - production/extension: detailed mapping, structural analyses (fault net slip,
fold analysis, vein analysis, etc.), drill target design.

Tectonics of Convergent Plate Margins


Because most structurally-controlled ore deposits formed in tectonic provinces
undergoing active deformation, it is important to understand the general geologic setting
and related structural framework of prospective provinces. Convergent plate margins are
used as an example here, as they are metallogenically rich and are sites where magmatic
arcs form above a subducting oceanic tectonic plate (Fig. 2). They are a fundamental
tectono-magmatic (volcano-plutonic) feature of the crust, and are characterized by a
number of common and important features listed below.
1. They contain a number of tectonic provinces, including forearc, magmatic arc,
and backarc (foreland) environments.

2. Tectonics in these provinces varies between crustal extension, crustal shortening,


and wrench (strike-slip) tectonics depending on the relative plate motion vector
(orientation and relative velocity) across the subduction plate boundary. Plate
motion vectors change in time causing temporal changes in styles of deformation.

3. Cordilleran-type (continental) magmatic arcs are constructed on old (Precambrian


to Paleozoic) basement that generally has two important features. These are:
a. The basement is geochemically fertile due to its long tectonic,
metamorphic, and magmatic history involving oceanic terrane accretion
and continental crustal reworking, and therefore has a relatively high
metal budget.
b. The basement has numerous structural weaknesses that can be structurally
reactivated and act as plumbing systems for hydrothermal mineralizing
fluids during younger (e.g., Tertiary-Quaternary) tectono-magmatic
activity. Such weaknesses may include old faults, ductile shear zones,
fold axial planes, and major contacts such as pluton margins.

4. Cover rocks (above basement) include a variety of good host rocks for
mineralization, such as sedimentary and volcanic strata.

5
5. Mariana-type (oceanic) magmatic arcs, although usually lacking old basement,
nonetheless can contain ore deposits, indicating the importance of subducted
oceanic crust magmatism, and mantle input.

6. Magmatic environments vary from plutonic, to hypabyssal and volcanic


(eruptive). Many magmatic features indicate structural/tectonic control, including
dikes, elongate plutons, aligned volcanic centers, and, importantly, circular to
elliptical volcanic collapse structures (calderas and diatremes).

7. Convergent margin tectonic environments contain a complex, but somewhat


predictable, fault and fracture mesh developed in an upper- to mid-crustal brittle
to semi-brittle strain environment. Variations in the regional stress field result in
a number of possible fault-fracture geometries and orientations (details given
below).

8. Brittle deformation leads to magmatic-hydrothermal fluid flow (permeability) in


structurally-controlled openings (porosity) and the formation of related ore
deposits.

9. Mid-crustal strain environments, although dominated by ductile fabrics


(foliations, ductile shear zones), contain transiently-developed structural
openings. The brittle–ductile transition is present, on average, at a mid-crustal
depth of ~10-12 km and a temperature of ~300-350C; this depth varies with heat
flow and is elevated in areas where magmas bring heat into the crust and probably
along major shear zones. The brittle-ductile transition is where the largest
earthquakes nucleate on major faults and cause hydrothermal fluid redistribution
during high fluid pressure-assisted, high strain rate deformation pulses (co-
seismic fault slip). Subsequent (inter-seismic) mineral precipitation causes
permeability sealing and increase of fluid pressure leading to seismic slip. This
cyclic (co-seismic/interseismic) fault-valve process (Sibson, et al., 1988; Nelson
et al., 2007) can lead to economic metal grades in veins banded from multiple
slip/precipitation events.

Convergent margin stress-strain environment


The stress field, and resultant bulk strain, in convergent plate margin tectonic
environments (forearc, magmatic arc, and backarc) is controlled regionally by two
factors: 1) forces caused by tectonic-plate relative motions across the convergent
boundary and, 2) gravity force. Tectonic forces are transferred across the plate boundary
(surface trace = oceanic trench) into the continental plate causing a zone of deformation
generally between 200 and 2000 kilometers wide across the forearc, magmatic arc, and
backarc regions. The orientation of the tectonic forces derives from the vector of relative
motion and is the dominant control on the orientation of structures. Gravity forces
control the strength of rocks through confining pressure which is related to depth in the
crust (vertical stress = v = gh, where  is average crustal density to earth’s surface, g is
the gravitational acceleration, 9.81 m/s2, and h is depth).

6
As shown in the review of structural principles, stress (force/area) is generally used in
analyzing geological structures. The possible regional stress fields in convergent margin
provinces can be classified broadly into three categories:
1. Arc-normal horizontal compression, which causes crustal shortening
2. Arc-oblique horizontal compression, which causes wrench tectonics
3. Vertical compression, which causes arc-normal or arc-parallel extension

Stress fields can be changed locally and diachronously by collision of bathymetric


features on the subducting oceanic plate such as spreading ridges, seamounts, and
inactive ridges and plateaus, as well as by changes in relative plate motion, changes in the
age of subducting oceanic lithosphere (and thus the slab dip), and by uplift and erosional
adjustment of the topographic surface.

Regional strain features related to arc-normal compression include conjugate regional


wrench faults, and, particularly in the back-arc environment, fold-thrust belts (such belts
can be thin-skinned foreland fold-thrust belts or thick-skinned basement block (Figs. 2,
3). Strain features related to arc-oblique compression include arc-parallel wrench faults
and, locally, oblique fold-thrust belts. Arc-normal extensional environments generally
result from trench roll-back caused by rapid sinking of a relatively old and dense
subducting oceanic plate (Fig. 2, 3). Trench rollback leads to extensional collapse of the
arc massif and arc-parallel normal faults and grabens (Fig. 3).

Arc-parallel wrench faults are formed by oblique subduction (Fig. 3) and, locally, by
collision of oceanic bathymetric features (ridges, plateaus, etc.). Examples of such faults
in Chile include the Atacama Fault and West Fissure (Falla Oeste) exposed in the
Chuquicamata pit, and the Liquiñe-Ofqui Fault in southern Chile.

Exploration programs should evaluate the regional strain environment at the time of
mineralization and at the time of post-mineralization deformation events in the context of
these three tectonic-based regional strain models. Geochronological control on host rock,
alteration, and mineralization is thus very important. A number of structurally-controlled
regional sites for ore deposit formation related to these regional strain environments are
shown on Figure 3. These include:
 regional fault intersections
 regional wrench fault jogs and bends
 volcanic collapse-related fault-veins; radial, concentric, and regionally controlled
 graben-related fault-veins
 en echelon vein array systems
 thrust fault related vein systems

7
Figure 2. Schematic cross section of active (subducting) Andean-type continental margin illustrating
trench roll-back with sinking of oceanic plate in the mantle. Note that the foreland fold-thrust belt
develops during periods of no trench rollback.

8
Figure 3. Schematic plan view of stress and strain environments possible in Cordilleran magmatic arcs,
here using a generally N-S trending magmatic arc and trench. Faults are blue, ore deposits are red, and
circular volcanic features are shown in black.

9
Review of Structural Principles
This section reviews structural principles, with a concentration on the conditions optimal
for formation of brittle structures that enhance structural permeability and host
structurally-controlled ore deposits. It is relatively short without theoretical detail or
mathematical rigor. Topics covered include:
1. Stress concepts
2. Strain concepts
3. Fracture mechanics
4. Structural controls on permeability

Stress
Stress is force per unit area, and is a 3-D field property in the crust (similar to
the Earth’s 3-D magnetic field). It can be thought of
as an infinite set of force vectors in all orientations
acting on a given point in the crust, all with
orientation and magnitude. We simplify the stress
field by defining three mutually perpendicular
principal stresses:
1 = maximum principal stress
2 = intermediate principal stress
3 = minimum principal stress
Figure 4. Graphical representation
These principal stresses can be represented of stress field as stress ellipsoid with
graphically as the mutually-perpendicular axes of the mutually perpendicular principal
stress ellipsoid. The shape and orientation of the stress axes.
ellipsoid represent the state of stress (or stress field) at any given point in the crust (Fig.
4). Although defined strictly at a point and at Extensional regime
up 
an instant in time, stress effectively can be
discussed as a regional “stress field” because
stresses are related to large-scale tectonic plate

interactions and in situ stress axes maintain 
consistent orientation over large distances. The
principal stress axes are represented here as east
arrows so as not to be confused with the strain
ellipsoid which has mutually perpendicular
principal strain axes (1; 2; 3 – see
uth

description below).
so

Figure 5. Example of a 3-D stress field in


an east-west extensional tectonic regime
As mentioned above, the regional stress field is graphically represented by the three
controlled by relative plate motion and mutually-perpendicular principal stress
gravity (gravity stress can be considered axes.
confining pressure, Pc). An example of a stress field resulting in north-striking grabens
and east-west extension is shown in Figure 5 (1 vertical; 2 horizontal north-south; 3
horizontal east-west).

10
Local gradients and rotations in the stress field are possible and may be important in
exploration. A classic example is in the Spanish Peaks, Colorado (USA), where a
regional compressional stress field (1 horizontal east-west) is superimposed on an
outwardly radial stress field near a Tertiary volcanic neck. The result is a set of radial
dikes near the volcanic neck that gradually curve to strike east-west at sufficient distance
from the neck (Fig. 6). Volcanic- or pluton-centric stress fields may lead to radial and/or
concentric fracture patterns which can fill with dikes and/or ore bodies.

The importance of stress to exploration is that the orientation of the stress field is a
fundamental control on the orientation of brittle structures (faults and extension fractures)
formed when the stress magnitude exceeded the rock strength. Paleo-stress modeling of
structural data (see description below) can help identify the location and orientation of
mineralized exploration targets.

Because the earth’s surface is a surface of no shear stress, and therefore defined as a
‘principal plane’, one principal stress is essentially vertical near the earth’s surface.
Three main types of tectonic regimes, characterized by three main types of faults, may
form depending on which principal stress is vertical (Anderson, 1951; Fig. 7):
1. 1 vertical = extensional regime with normal faults
2. 2 vertical = wrench regime with strike-slip faults
3. 3 vertical = shortening regime with thrust faults.

Note that although these are the main fault types developed in each tectonic regime, the
other two fault types, in addition to oblique-slip faults, may form as transfer structures or
due to local stress rotations.

11
A

Spanish Peaks, Colorado


radial 1 in
volcanic neck

regional
1 E-W

volcano and
radial dikes
B
Figure 6. A. Map of Spanish Peaks intrusions and dikes (Muller and Pollard, 1977). B. Schematic map
view example of superimposed stress fields illustrated in plan view from Spanish Peaks, Colorado.
Radial stress field in volcanic neck produces radial dikes which near volcanic center, but east-west dikes
distant from the volcanic center where the east-west compressional stress field is dominant. Model based
on Odé (1957) and Muller and Pollard (1977).

12
Figure 7. Illustration of three tectonic fault regimes based on which principal stress is vertical. The
shortening axis is 1 and the extension axis is 3. Lines on fault planes illustrate orientation of the slip
vector (slickenlines). Based on theory of Anderson (1905, 1951).

Strain
Strain is a 3-D property describing the deformation geometry within a rock or region
(sometimes referred to as ‘bulk strain’). The 3-D strain can be described with the lengths
and orientations of principal strain axes compared with a pre-deformation sphere of unit
radius (Fig. 8). Of importance to exploration is that the long (1) and short (3) principal
strain axes represent the finite (or end-product) extension and shortening directions,
respectively.

Two fundamental end-member types of


strain development are possible:
rotational and non-rotational strain
(also known as simple shear or non-
coaxial strain, and pure shear or coaxial
strain, respectively) (Fig. 9). Simple
shear occurs mostly in the lower crust in
ductile shear zones where the 1 and 3
principal strain axes progressively rotate
away from parallelism with the 3 and Figure 8. Strain ellipse compared with pre-
deformation reference circle for pure shear strain
1 principal stress axes, respectively.
model; red arrows = 1
Simple shear also occurs locally in the
upper crust within fault zones during cataclastic1 flow in shearing fault gouge.

However, the dominant type of strain development in the upper crust is probably non-
rotational, particularly in relatively young tectonic provinces unaffected by superposed
deformation events. Therefore it can be assumed in many cases that the principal strain
axes remain generally parallel to the principal stress axes; that is, in orientation 3  1
and 1  3. If rotation is assumed to be small or insignificant, strain features in the
upper crust thus can be used to model paleo-stress orientations. An example is shown for
extensional strain in Fig. 10.

1
Cataclastic means grain-size reduction through fracturing. Ductile flow can occur in gouge (a cataclastic
fault rock) by grain boundary sliding, grain rotation, and microfaulting.

13
Figure 9. Schematic illustration of end-member types of strain, rotational and non-rotational strain; red
arrows represent maximum principal stress direction 1. Time steps in strain development, 1-5 show
progressive shortening of axis and lengthening of 3 axis. In non-rotational strain the 3 strain axis
remains parallel to the 1 stress axis, whereas in rotational strain the strain axis progressively rotates
away from parallelism with the stress axis.

Bulk regional-scale or district-scale strain 


can be classified into two categories based
on the magnitude of the intermediate
principal strain, 2 (an extensional tectonic
regime is used as an example):
1. Plane strain (2 = 0), in which there graben
is no finite (end product) elongation

along the 2 direction. For
example, classic grabens form with

a conjugate normal fault geometry 1
(Fig. 11a), and probably represent bulk strain
the most common type of Figure 10. Example of non-rotational strain in
extensional bulk strain. the upper crust: conjugate normal faults and
2. Non-plane (triaxial) strain (2 > 0), horizontal extension form from vertical
maximum principal stress. 1 is parallel to 3.
in which there has been elongation
along the 2 direction and at least two sets of grabens form (Fig. 11b).

In exploration it is important to recognize the type of bulk strain. In areas of plane strain,
kinematically-controlled or dilational ore shoots2 should have the same general
orientation throughout a district, and are therefore predictable. However, in areas of non-
plane strain, such ore shoots will differ in orientation because multiple sets of faults can
form, and ore shoot orientation is not easily predicted.

2
Kinematically-controlled ore shoots form in fault jogs or bends perpendicular to the slip vector (=
slickenlines), and are here considered to be the most important type of ore shoot, and the first model to be
tested in exploring for ore shoots along fault-veins. They are described in detail below.

14
Figure 11. Block diagrams of plane strain and triaxial strain types and stereonets showing fault planes
and stress and strain axes. After Davis and Reynolds, 1996). Also, see Miller et al., (2007).

Brittle strain, which consists mostly of fracture-fault meshes, is the dominant strain type
in the upper crustal environment. Stress and brittle strain are linked in the next sections
to explain how fractures form and how they lead to structural permeability which is
critical in the formation of structurally-controlled ore deposits.

Description and classification of structures

Fractures: faults and extension fractures


Fractures, which cause structural permeability and are the sites of most structurally-
controlled ore deposits, form in two fundamental end-member modes: 1) opening mode
and 2) shear mode. Opening mode fractures (also called extension fractures or,
incorrectly, ‘tension’ fractures or ‘tension’ gashes) form by movement of the fracture
walls perpendicular to the plane of the fracture; they form in the 1-2 plane and
perpendicular to 3 (Fig. 12a).

Figure 12. A. Opening mode fractures (= extension fractures) form in the plane perpendicular to the
minimum principal stress axis. B. Openings can also form along shear fractures (faults) with any
deviation from planar fault geometry.

Shear mode fractures (= faults) form by movement of the fracture walls parallel to the
plane of the fracture. Theoretically they form at 45 to 1 because the maximum shear

15
stress is developed on planes of this orientation. However, empirical observations in rock
deformation experiments show that shear fractures form at an angle of ~30 to 1, and
may form as a conjugate set (faults with opposite slip sense and with an inter-plane angle
of ~60 that is bisected by 1)(see example in Fig. 10). As discussed below, both of these
angles (30 and 45) have been used in modeling of paleo-stress orientations. Faults also
contain the 2 axis which is perpendicular to the slip vector (indicated by slickenlines).

Broadly, the word “fault” describes one of three possible, slip vector rake fault type
generally planar or tabular, displacement features: 1) single 0‐20° strike slip
slip-planes (faults), 2) fault zones with multiple parallel or 20‐70° oblique slip
anastomosing slip planes, or 3) semi-ductile to ductile shear 70‐90° dip slip
zones. Semi-ductile shear zones contain both brittle and Table 1. Fault classification
using slip vector rake.
ductile shear strain features, such as foliations and veins
(filled fractures).

Fault displacement is defined by the slip vector, which has orientation (rake in fault
plane) and magnitude (fault offset). The magnitude is the distance between hangingwall
and footwall piercing points, which are points where an offset geological line intersects
the fault, or points defined by graphical construction where the slip vector crosses a cut-
off line3.

The classification of faults and shear zones is based on two parameters: slip vector rake
angle (Table 1), and slip (or shear) sense. Slip sense is dextral (right) or sinistral (left) for
strike slip components, and reverse (hanging wall relatively up) or normal for dip slip
components. Figure 13 shows a fault classification system using right-hand-rule rake of
the slip vector. Oblique slip fault terminology combines one descriptor from each
component (example: sinistral-normal fault, Fig. 13d).

In practice, many fractures that contain ore deposits probably formed with both opening-
mode and shear mode components. This is because any slight deviation from planar fault
geometry (e.g., at fault bends), will cause some component of opening mode along the
fault (Fig. 12b), and thus lead to potential fluid flow and ore precipitation.

Importantly for the formation of ore deposits, extension fractures form the principal
structural permeability, preferentially in the 2 direction. However, extension fractures
can form locally along faults where the fault plane orientation bends to approach
parallelism to 1, or in jog (step-over) regions (Fig. 12b, 14). These locations along
faults are prime targets for exploration. Note that jogs and bends are 3-D features and
can form in any orientation; however the orientation is commonly related to kinematics
of the fault.

3
A cut-off line is the line where an offset geological surface, such as a contact, intersects the fault plane.

16
Figure 13. Fault classification using longitudinal sections (view perpendicular to fault plane) showing
semi-circular rose diagrams of right-hand-rule slip-vector rake (A, B, C). A. General fault classification
showing RHR rake angles for strike slip (<20° and >160°), oblique slip (20°-70° and 110°-160°), and dip
slip (70°-110°). Block diagrams show examples of each fault type. B. Rose diagram for faults with
normal slip component. C. Rose diagram for faults with reverse slip component. D. Block diagram
showing example oblique-silp fault with attitude = 000°/60° and RHR rake = 45° (red dashed line is slip
vector). E. Stereonet showing example fault and slickenline.

Another important aspect of fault bend


mechanics is the development of (cymoid
second-order fractures in hanging wall loop)
and footwall damage zones. Such
fractures form as the fault tip propagates
through a fractured damage zone
surrounding the fault tip (Fig. 15a).
jog
These fractures (called Riedel fractures)
usually form as en echelon sets and
include R, R’, P, and T types (Fig. 15b);
T-fractures are extension fractures and
the others are shear fractures. These
Figure 14. Local development of opening mode
fractures have two important (extension) fractures in fault jog (forming an en
applications in exploration: echelon vein array) and fault bend structures. Gray
1. Most Riedel fractures intersect the arrows illustrate maximum principal stress
fault plane at a low angle and form orientation.
asymmetric steps on the fault plane when thin wedges break off in fault outcrops (Fig.

17
16). Useful steps can be identified because they are approximately perpendicular to
slickenlines. Riedel step asymmetry is one of the most useful indicators of fault slip
sense. Slip sense is useful in predicting the geometry of dilational jogs (for example
dextral strike-slip faults are predicted to have right-stepping dilational jogs; Fig. 14).
Also, knowledge of slip sense on post-mineral faults is critical in drill target design
for offset ore bodies.
2. All Riedel fractures intersect the fault plane in a line parallel to 2 and perpendicular
to the slip vector (slickenlines). This direction is the preferred permeability direction
and the predicted ore shoot orientation for kinematically-controlled ore shoots.
Riedel fractures increase the fracture density and aperture space for ore mineral
precipitation.

Figure 15. A. Development of second-order fractures (= Riedel fractures) in wall rock of faults. B.
orientation of Riedel fractures (R, R’, P, and T) relative to main fault and principal stress axes.

slickenlines

fault plane

R-fractures
A
slickenlines

fault plane
Riedel steps

R-fractures
B
approx.
interpreted
slip sense 1
R-fractures
C
Riedel steps
Figure 16. Illustration of formation of Riedel steps. A. R-fractures form in wallrock of fault;
intersection line with fault is ~90. B. Small wedge of rock breaks off to form Riedel step. C. Profile of
block in B showing steps and asymmetry; paleo-1 is generally parallel to the Riedel fractures.

18
Care must be exercised in interpreting slip sense using Riedel steps, because asymmetric
steps on fault planes can also form as crystal fiber steps. Crystal fiber steps, which are
equally useful in determining slip sense, form by growth of fibrous crystals (usually
quartz or calcite) in small dilational jogs along fault planes (Fig. 17). However, given the
same step asymmetry, slip sense is opposite to that interpreted using Riedel steps.

A. B.

Figure 17. A. Rock sample showing crystal fiber steps (top block, now missing, moved to the right). B.
Block model showing dilatant fault jog (step) with crystal fiber fill; same movement sense as in photo.

Folds
Folds can affect sedimentary and/or volcanic strata, as well as metamorphic/structural
fabrics (cleavage, foliation). Although massive rocks, such as granitic rocks, can be
folded, recognition of folds in such rocks is difficult without planar fabrics. Folds can be
important controls on geometry and location of ore deposits, as both pre- or syn-
mineralization structures and/or as post-mineralization structures. Fold-related veins of
various geometries are possible (e.g., Bendigo, Australia, Fig. 18).

Figure 18. Schematic cross section showing fold-related vein types in Bendigo district, Australia.

19
The description and classification of folds is based on a number of fold geometric
features:
1. Fold orientation (Fig. 19, 20):
a. Fold axis (plunge/trend) measured directly or modeled using various methods
(π-analysis), bedding-cleavage (So-S1) intersection, L1 lineation (So-S1
intersection), etc.
b. Axial surface (strike/dip)
2. Tightness of fold limbs (inter-limb angle), ranges from open to isoclinal,
3. Fold asymmetry, which indicates shear of folded layer (Fig. 21),
4. Cross sectional geometry of folded layers,
5. 3-D geometry of fold
a. Axial surface can be planar or curvi-planar
b. Fold axis can be linear or curved; culmination = topographic high along fold
axis, depression = low along axis
c. End-member geometries are cylindrical (successive cross sections of fold
perpendicular to fold axis are the same) and conical

Figure 19. Fleuty classification of fold orientation.

20
Figure 20. Block diagram of fold showing poles to bedding and stereonet showing π-analysis model of
fold axis orientation.

Figure 21. Schematic illustrations of fold asymmetry, and statistical parallelism of major and minor
(parasitic) folds.

Structural fabrics

Secondary structural fabrics (formed by structural and/or metamorphic processes in


metamorphic terrains and in regional shear zones) include planar fabrics (foliation,
cleavage) and linear fabrics (lineations). Most fabrics consist of alignment of inequant
mineral grains, and thus impart a preferred structural weakness to rocks. Such secondary
fabrics are important in ore deposit exploration, and can control the orientation of
structurally-controlled ore bodies, and are important aspects of rock mass structure during
mining. Structural fabrics are also critical in understanding the 3-D structural geometry
of a district.

Structural Permeability
As noted above, structural permeability in preferentially developed along the 2 direction
in areas of bulk plane strain, although fluid pressure gradients are important in driving the
ultimate direction of fluid flow. Structural permeability, and thus ore body formation, is
localized in areas of highest fracture aperture and fracture density. The physical evidence
of past fluid flow includes extension veins, mineralized faults zones, and breccias with

21
hydrothermally-precipitated matrix (cement), all of which may occur in the same
structure. Note that some faults with clay-rich gouge may be barriers to fluid flow
(sometimes called aquicludes) against which ore bodies may form in matrix-permeable
layers (such as sandstone, some carbonates, or volcaniclastic layers).

Fault-fracture permeability meshes may form in various geometries (Fig. 22b). Note,
however, that focusing, not dispersion, of mineralizing fluids is a process necessary for
ore formation. Focusing of permeability along the 2 direction in fault-fracture meshes
may occur in a number of structural settings (Fig. 23), including:
 dilational fault jogs and bends (Fig. 23b, 23c, 24)
 en echelon vein arrays (Fig. 23a, 25)
 intersection zones (fault-fault, fault-bed, fault-dike, dike-bed, etc.)

The shear sense of an en echelon vein array (right-stepping array geometry = sinistral
shear and vice versa) can be opposite to the shear sense on individual veins due to
bookshelf sliding (Fig. 26).

Brecciation in any of the above structures, as well as in intrusive pipes and chimneys,
will enhance fracture permeability through rotation of fracture-bounded blocks.

Figure 22. A. Fault-fracture mesh composed dominantly of en echelon quartz vein arrays in conjugate
semi-brittle shear zones (Canadian Cordillera near Banff). B. Schematic illustration of possible
permeability meshes composed of veins and faults; horizontal (in this view) wiggly lines represent
stylolitic pressure solution seams. Note that the highest permeability is along the 2 axis, perpendicular
to the page. Adapted from Sibson, 1996.

22
Figure 23. Illustration of structural permeability developed preferentially along 2 direction in en
echelon vein arrays, dilational fault jogs, and dilational fault bends. Example used here is strike-slip
(wrench) faults in which slickenlines rake ~0.

23
Figure 24. Picture of quartz vein in left-stepping fault jog along sinistral wrench fault.

Figure 25. Example of en echelon quartz vein array in semi-brittle sinistral shear zone. Note that pre-
existing quartz vein is offset in dextral sense across individual quartz veins within the array. This is due
to bookshelf sliding (see model in Fig. 26).

24
Figure 26. Schematic illustration of bookshelf sliding mechanism of developing opposite shear sense
(dextral) on fault-veins within a sinistral shear array. A and B show books on bookshelf; C and D show
explanation of en echelon vein array example in Fig. 25; note that initially-formed veinlets (C) rotate
counterclockwise in sinistral shear zone to cause dextral offset of pre-existing veinlet (D).

Structural Methods in Mineral Exploration


Structural methods useful in mineral exploration are divided here into two categories:
data collection and database construction and data analysis. Data analysis methods are
numerous and include standard graphical methods such as construction and analysis of
cross sections and orthographic projections. For brevity, some of these basic methods are
not reviewed here. Methods reviewed here include:
 Analysis of structural orientation data using ‘stereonets’
 Analysis of structural patterns in plan and cross section views
 Analysis of longitudinal sections of fault-veins
 Drill hole design in targeting of ore zones and ore shoots

Structural data collection - introduction


Mineral exploration and geotechnical analysis have completely different goals. Whereas
the goal of exploration is to locate and define the geometry of economic resources, the
goal of geotechnical analysis is to model the engineering behavior of earth materials in
construction projects on or in the ground. Note that the type and amount of structural
data collected is very different for these different goals.

25
The availability of good outcrops strongly affects the types and amounts of structural data
that can be collected. The degree of accessible bedrock exposure is affected by terrain
(e.g, cliffs) and various types of cover including water, ice, vegetation, regolith, and
Quaternary unconsolidated deposits (alluvium, colluvium, glacial deposits). Strong
hydrothermal alteration and deep weathering of rock can obscure geologic structure,
particularly with formation laterite and saprolite in tropical environments.

Both underground and surface mining expose fresh rock not affected by cover, and
expose less weathered rock with depth. Therefore, detailed mapping and structural data
collection should always be done as soon as mine exposures are available. Structural
data should be entered into a database within days of collection and should be analyzed
regularly as mapping proceeds.

Structural orientation data


Although seen as routine by many exploration geologists, collection of structural
orientation data should not be routine. In addition, much structural data are collected in
exploration and mining programs without follow-up analysis, and sometimes without
reason. Structural data should be collected to test models, not as a routine exercise
without asking why the data are being collected. Structures requiring collection of
orientation and descriptive data in include:
 Bedding or lithological contact – strike/dip
o slickenlines indicate bedding plane fault, common in folds formed by flexural-slip process
 Fault or brittle shear zones – strike/dip
o slickenline rake (or plunge/trend)
o slip sense (dextral, sinistral, normal, reverse)
 Ductile shear zones
o strike/dip of foliation, asymmetry of foliation traces, rake of mineral or stretching lineation
 Vein, or fault-vein– strike/dip
o slickenline rake (or plunge/trend)
o slip sense (dextral, sinistral, normal, reverse)
o plunge/trend of ore body or ore shoot long axis (Fig. 14)
o vein array geometry (en echelon patterns, stepping sense, etc.)
o mineral fill and timing relative to mineralization and alteration
 Fold
o trend/plunge of fold hinge
o strike/dip of fold axial plane (parallel to axial plane cleavage)
o fold asymmetry (S or Z), vergence4 - not an orientation datum, but very important
 Dike (can be igneous, clastic (syn-diagenetic), or breccia dikes (e.g., phreatic or phreatomagmatic)
o strike/dip
o En echelon or other array patterns
 Flow foliation or metamorphic foliation in basement – strike/dip
o trend/plunge of flow lineation
o shear sense – not an orientation datum, but very important

Orientation data, once collected, should be entered rapidly (same day or week) into a
well-organized digital database (Excel or other database) along with other important

4
Vergence is the azimuth in which upper rocks have moved relative to lower rocks in an inclined fold; or
the azimuth opposite the dip direction of fold axial planes. Also, it is the azimuth of movement of the
upper plate of thrust faults.

26
information associated with orientation data (easting, northing, elevation, type of
structure, slip sense of faults, comments, etc.). An example structural database is shown
in Appendix I.

Examples of model testing with orientation data include:


1. Basement structure (folds, old faults, foliations, etc.): these structures are possibly
important as controls on formation of faults and veins in the volcanic or sedimentary
cover sequence. Compare basement structure orientations with orientations of
mineralized faults and veins to test for structural controls.

2. Volcanic structures such as flow foliation, volcanic layering or bedding, and dikes
should be tested for geometry using structural orientation data. For example:
 Flow layering – test for flow dome geometry, concentric/radial or elongate
 Bedding/volcanic layering – test for block rotations, inference of listric
normal faults, intrusive centers
 Dikes – test for intrusive center; extension direction is perpendicular to dikes

3. In porphyry systems, veins form in a


number of stages. Preferred orientations
of vein sets can indicate evolving stress
fields (pre- to syn- to post-mineralization).
The preferred orientation of mineralized
veins is critical for drill orientation design.

4. Faults and fault-veins are one of the


primary exploration targets and should be
tested for kinematic control on ore shoot
orientation. In kinematically-controlled or
dilational ore shoots, the ore shoot forms
in fault jogs or bends, and the test of Figure 27. Illustration of kinematically-
kinematic control is that the long axis of controlled ore shoot orientation in reverse,
normal, and wrench fault systems. The long
the ore shoot is perpendicular to the
axis of the ore shoot is perpendicular to
orientation of slickenlines in the plane of slickenlines in the fault plane. See Nelson
the fault-vein (Fig. 27)(Nelson, 2006). (2006).

5. In mesothermal [orogenic] vein systems, the orientation of the long axis of ore bodies
should be compared with a number of structural features, including metamorphic or
shear zone lineations, fold hinge lines and axial planes, and foliations. Shear zone
lineations and possibly metamorphic lineations will test for ore body formation along
the b-tectonic direction (perpendicular to shear zone slip vector) or along the a-
tectonic direction (parallel to shear zone slip vector). If ore bodies are elongate along
the a-tectonic direction, it is likely that the ore bodies have been ductily deformed
(mesothermal ore commonly occurs in veins which initially formed as brittle
structures).

27
The orientation of lines, such as fault striations (or slickenlines), metamorphic lineations,
and fold hinges, is very important in exploration, as linear ore bodies (ore shoots) may
form along linear structures (e.g., jogs along fault-veins, intersection lines, fold hinges).
Line orientation can be given as plunge and trend, where plunge is the vertical angle of
the line, and trend is the down-plunge azimuth direction of the imaginary vertical plane
containing the line (Fig. 28a). However, lines within planes (e.g., slickenlines within a
fault plane, or metamorphic lineations on a foliation plane) are best measured using the
rake angle (≡ pitch). Rake is the angle between strike line and the lineation, measured
with a protractor in the plane containing the line (Fig. 28b); the direction in which the
angle opens must be given as quadrant (NE, SE, SW, NE) or by recording right-hand-rule
rake (see below). Rake angle can be converted to plunge/trend using either a spherical
projection (“stereonet”) or algorithms in a spreadsheet (Appendix II). Rake angle is
always greater than plunge angle except on vertical planes (Fig. 29). The map symbol for
a lineation is an arrow drawn on the strike/dip symbol or fault trace at the point of
measurement; the arrow points in the trend direction and is labeled with the plunge angle
(not the rake angle; Fig. 28c).

It is recommended that the following standard procedure be used company-wide to record


and present orientation data:
 Strike data are presented in right-hand-rule (RHR) format5, in which the dip direction
is clockwise from the recorded azimuthal (0-360) strike direction. Strike/dip data
are presented in a three-digit/two-digit format: xxx°/xx°.
 Plunge/trend data of linear features (such as slickenlines) are presented in two-
digit/three-digit format: xx°/xxx°.
 Rake angle ( pitch) is presented as angle (0-90) in one database column and in
right-hand-rule format (0-180) in another column (this is the angle measured from
the right-hand-rule strike direction). The right-hand-rule format rake angle is used by
the Faultkin6 computer program for modeling orientation of principal strain axes.
c

N
30

Figure 28. Illustration of methods of orienting a lineation in a plane (e.g., slickenlines on fault plane). A. trend
and plunge; B. rake (=pitch). C. Map symbol for this example showing 30 plunge with a trend of 068 (written
30/068).

5
Dip/dip-direction format also can be used, and is easily converted in a spreadsheet to strike/dip in RHR
format [cell equation for RHR strike: =if(A1<90, A1+270, A1-90), where cell A1 contains dip direction].
6
For purchase of Faultkin program, contact Dr. Rick Allmendinger,
http://www.geo.cornell.edu/geology/faculty/RWA/programs.html

28
90
90
80
70
60 60
50

Plunge
40
30
20
30 10

0
0 30 60 90

Rake angle

Figure 29. Plot of lineation plunge vs. rake angle for various dip angles of plane containing the
lineation. Note that rake angle is always greater than plunge angle except on vertical planes.

Structural descriptive data


In addition to collection of orientation data, the description of structures is very important
in mineral exploration. Data for three main descriptive categories should be recorded
when possible:
1. Structural timing
2. Structural geometry
3. Structural kinematics

Descriptive data for structures is also important for geotechnical purposes, but a different
set of descriptive parameters is collected (see chapter on Characterization of rock mass
structure).

Structural timing
Structures can form before, during, or after mineralization and hydrothermal alteration
(e.g., pre-, syn-, or post-mineralization, respectively). Pre- or syn-mineralization
structures, such as fractures (faults and extension fractures), foliations, and contacts,
constitute weaknesses in the rock that can be utilized as plumbing conduits during
hydrothermal fluid flow and precipitation. The distinction between pre- and syn-
mineralization structures is nearly impossible in the field. Post-mineralization
deformation, resulting in formation of faults, folds, and foliations, can affect the
geometry and distribution of ore. Numerous examples exist of deposits offset by faults,
folded by folds, and deformed by ductile flow.

Timing of structures can be absolute (using geochronology) or relative (using cross-


cutting relationships). In the epithermal environment, post-mineral faults can be
recognized by either offset of veins or ore shells, and by mesoscopic outcrop observations
such as the presence of breccias in the fault core containing clasts of mineralized rock in
a matrix of un-mineralized material (typically clay-rich gouge).

29
Structural geometry
The geometry of structures includes size and shape of structures, as well as structural
density (number of structures, such as veins, in a given volume of rock). Features
recorded vary for the type of structure being described. For example, for faults important
parameters include the surface shape of the fault (planar or curvi-planar), the width of the
fault zone (fault core and surrounding damage zone), the type of fault rock in the fault
core (breccia, gouge, or foliated rock), the geometry of fault segment jogs and bends
(left- or right-stepping), etc. For folds, important parameters include fold wavelength and
amplitude, fold asymmetry (s, z, or m), interlimb angle (i.e., tightness – open, closed,
isoclinal), the geometry of individual layers (similar, parallel), and the overall fold
surface geometry (cylindrical, conical). The geometry of foliation is important for
determining shear sense in shear zones, and involves describing the sigmoidal,
asymmetric pattern of foliation traces. The geometry of vein arrays is similarly important
and includes en echelon and sigmoidal patterns.

Structural density can be measured as spacing between structures of a given set along a
scanline (see Fig. 58) or along 3 perpendicular scanlines (see Fig. 59). These methods
are typically used in determining joint density for geotechnical analysis. A method of
determining veinlet density developed by Titley and Hendrick (1978) uses an index of
total veinlet length per surface area of an outcrop to determine proximity to closely-
spaced veinlets in the core of porphyry systems (Fig. 30).

Figure 30. Illustrations of veinlet density index method of Titley and Hendrick (1978). A. Square drawn
on outcrop showing veinlet traces; cumulative length of veinlet traces (301cm) is divided by surface area
of square (2500 cm2) to derive veinlet density index (0.12 cm-1). B. Example of density index vs. distance
from porphyry core (vertical lines show statistical variation in measurements).

Structural kinematics
Kinematics, or movement, is an important parameter to describe for faults, shear zones,
and asymmetric folds (defined by vergence). Fault kinematics (slip sense) can be
determined either from piercing point analysis (from offset geological lines), fault
separation of geological surfaces combined with slip vector orientation (e.g.,
slickenlines), or from asymmetry analysis of mesoscopic or microscopic fault-plane or

30
fault zone structures. Examples of asymmetric structures include Riedel and crystal-fiber
steps on fault planes, and sigmoidal foliation traces and asymmetric folds in clay-rich
fault gouge. Ductile shear zone kinematics (shear sense) can be determined using
sigmoidal foliation traces and asymmetric folds in high-temperature ductile fabrics.

Fold vergence4 defines the overall movement of hangingwall relative to footwall in


orogenic systems.

Structural analysis methods


Various structural methods are useful in exploration programs, both in helping to identify
drill targets and in designing drill-hole orientation. These methods include plotting and
statistically analyzing orientation data on spherical projections (“stereonets”), modeling
orientations of paleo-stress (or strain) axes, analyzing structural patterns in map and cross
section views, and analyzing various data in longitudinal sections of fault-veins.

Strike and dip data easily can be analyzed using simple histogram plots (Fig. 31), which
can indicate orientation sets and dominant structural orientations.
SE dip SW dip NW dip NE dip

Figure 31. Analysis of planar orientation data using simple histograms.

Analysis of structural data on spherical projections


One of the fundamental tools used in analyzing structural orientation data is the spherical
One of the fundamental tools used in
analyzing structural orientation data is the
spherical (or hemispherical) projection. A
spherical projection is commonly called a
stereonet, but this term is only correctly
applied to equal angle projections (Wulff
net); for structural geology analysis it is best
to use an equal area projection (known as a Figure 32. Cross section of projection
Schmidt net; Fig. 32) because area is hemisphere showing method of construction of
preserved across the net and statistical equal area net (seen here in cross section) and
distributions of orientation data can be related equations.
determined (we still call these plots ‘stereonets’). On such projections planes are
represented as great circles and lines as dots (Fig. 33). Planes also can be represented as
poles which plot as dots. The basic methods of plotting planes as great circles and lines
as dots are shown in Figure 34. The full method, utilizing rotations of the transparency

31
overlay can be found in various textbooks on structural geology (e.g., Davis and
Reynolds, 1996; Lisle and Leyshon, 2004).

Figure 33. Illustrations of how stereonets are made by projecting lines (A-C) and planes (D-F) onto the
net (the equatorial plane in the hemisphere); from Lisle and Leyshon, 2004. A and D show 3-D
illustration of line (A) and plane (D) in space. B and E show projection of line and plane onto net. C
and F show equal area (stereonet) projection of features.

A B C
31

ek
ra
plunge slickenline
pole
90 °
dip

Plane: ke
stri
Line: d rake = 40°NW, or
tre n
1 65

165°/40° 24°/313° rake = 140° (RHR)


°

Figure 34. Stereonet plots. A. Great circle plotted with strike/dip in right-hand-rule format; pole is
plotted 90 from great circle. B. Dot representing line plotted with plunge/trend. C. Slickenline dot
plotted with rake angle measured along great circle.

To perform statistical analysis of orientation data, planes should be plotted as poles.


These poles can then be density contoured and averages of planar fabrics or fabric sets
can be selected visually as the sites of maximum pole density. An example of this type of
analysis from the Idaho Springs district, Colorado, shows that Tertiary veins were
controlled by pre-existing Precambrian foliation (Fig. 35). Many computer programs for
making spherical plots can automatically calculate average orientations and cylindrical
best fit great circles to dispersed populations. However, care must be exercised because
such automatic calculations may not always be appropriate for the data set being
analyzed.

32
For example, the average dip of a set of oppositely-dipping planes can easily be
calculated in a spreadsheet but not using an eigenvector calculation in a stereonet
program (Fig. 36). The average strike can be determined with the eigenvector calculation
but cannot be calculated directly in the spreadsheet unless all right-hand-rule strikes are
converted to one quadrant.

Figure 35. Example of statistical analysis of structural data. Figure 36. A. Example of veins plotted as
Vein (A) and foliation (B) orientations in the Idaho Springs great circles and as poles correct strike of
district, Colorado. A. Plot of poles to Precambrian foliation (left) average plane but incorrect dip calculated
and density contour (right) showing average foliation; average using eigenvector calculation (309/87);
fold axis (-axis) is also shown along with cylindrical best fit B. Correct dip averages calculated in
great circle (dashed). B. Plot of poles to Tertiary veins (left) and spreadsheet, but incorrect strike (238, left
density contour (right) showing average vein orientation. column). ‘Correct’ strike determined by
Comparison shows that veins were likely controlled by foliation converting right-hand-rule strikes to NW
orientation. quadrant. From Caine et al., 2006.

Analysis of slickenline data


Slickenline rake angle data for a region or district easily can be analyzed on a histogram
(Fig. 37). In this analysis, rake angles fall into three broad categories which correspond
to three dominant structural regimes:
1. Rake angles <20: dominant strike-slip regime (dextral or sinistral faults)
2. Rake angles >20 and <70: dominant oblique-slip regime
3. Rake angles >70: dominant dip-slip regime (reverse or normal faults)

33
Such analyses can help determine the dominant structural regime or, if data are widely
scattered, can indicate that multiple slip events have occurred. In this case, multiple
slickenline orientations will be observed on single fault planes.

Strike slip Oblique slip Dip slip


2

Roxana vein system

F r e q u e n cy
All other veins

10
15
20
25
30
35
40
45
50
55
60
65
70
75
80
85
90
0
5
Slickenline rake angle
Figure 37. Slickenline histogram analysis. Left: fault-vein slickenline data from a strike-slip dominant
tectonic setting (Chipmo district, Peru). Right: example from Quellopata district, Peru in which
different veins are analyzed separately showing that strain partitioning probably occurred in the area.

Paleo-stress and strain modeling


Paleo-stress and strain modeling is very useful in determining the dominant stress field
during vein formation and also in predicting the orientation of opening mode fractures
(veins and ore shoots). As mentioned above, faults form at a predicted angle of ~30
from the 1 axis and extension fractures form perpendicular to the 3 axis. The
orientations of the paleo-stress axes can be modeled using fault kinematic data (fault
plane strike/dip, slickenline rake, and fault slip sense). The simplest type of model is the
M-plane model which is done graphically on a stereonet (Fig. 38). The M-plane, or
movement plane, is perpendicular to the fault plane and parallel to the slickenlines (Fig.
38a). The maximum principal stress axis (1) is modeled at 30 from the slickenlines in
the M-plane (Fig. 38a). Two 1 orientations are derived in the modeling and one unique
solution is selected if the slip sense is known on the fault (Fig. 38b). The 3 orientation is
then located 90 from 1 along the M-plane, and the 2 orientation is the pole to the M-
plane. Extension fractures can then be predicted as the plane containing both 1 and 2
(Fig. 38c) and the long axis of ore shoots is predicted to be parallel to the 2 axis.

The principal strain axes (1, 2,3) can also be modeled, most easily using the Faultkin
program (see Marrett and Allmendinger, 1990 and Allmendinger, 2001). This program is
similar to the M-plane method, but uses an angle of 45 between fault plane and the
instantaneous principal shortening axis (P-axis or 3); the instantaneous principal
extension axis (T-axis or 1) is also modeled for individual fault planes. Bulk strain axes
are then determined using linked Bingham spherical statistical calculations on numerous
individual fault plane models.

34
Figure 38. M-plane analysis. A. 3-D illustration showing model of 1 30 from slickenlines in the M-
plane (movement plane) which is perpendicular to the fault and parallel to the slickenlines; predicted
extension fracture (red) is in 1-2 plane. B. Equal area plot showing two possible models of 1. The
correct model can be determined if the slip sense is known on the fault. C. Equal area plot showing
modeled principal stress axes for normal fault solution and predicted extension fracture orientation.

Figure 39. Example of Faultkin modeling of principal strain axes, and by inference, principal stress
axes in Arcata district, Peru (Veta Baja). A. Equal area plots showing (left) fault-vein planes with slip
linear arrows showing motion of hanging wall at slickenline, (center) contours of individual tension (T)
axes and shortening (P) axes, and (right) fault-plane solution showing 1 axis (1) bisecting quadrant
between modeled nodal planes. B. Longitudinal profile of Veta Baja showing ore shoot nearly
perpendicular to slip vector. Modified from Echavarria et al., (2006).

35
Because young (Tertiary-Quaternary) convergent margins are mostly expose a low-strain
brittle deformation environment, it is here assumed that little rotation has occurred and
the principal stress axes can be determined as follows: 1  3, 3  1, and 2  2.
Figure 39 shows an example of Faultkin modeling of fault kinematic data from Veta Baja
in the Arcata district, Peru. The model shows that the compression axis (1) was nearly
vertical and the extension axis was horizontal and oriented ~024.

Structural pattern recognition


Analysis of structural patterns of fault, fracture, and vein (fault-vein) systems is very
useful in determining the general
strain field of a region, as well as in
predicting the location of
undiscovered veins. This analysis is
accomplished in map and cross section
views, but the patterns seen are
representations of three dimensional
fault-fracture mesh geometry. Typical Figure 39. Geometry of cymoid loop (in fault bend)
patterns of fault-vein arrays include and horsetail splays in map view. Individual splays
conjugate and ‘orthorhombic’ patterns have component of normal fault movement.
(Fig. 11), en echelon patterns of either
extension fractures or faults, and jog and bend patterns (also an en echelon form) which
can form cymoid loops in which horse blocks are surrounded by faults (Fig. 14, 39).
Horse-tail splay structure near the tip of faults is a good indicator that an en echelon fault
or vein segment may exist (Fig. 40).

The location of ore shoots within en echelon fault-vein arrays is variable. In some cases
ore shoots form in the centers of segments where fault displacement and fracture damage
is greatest (Fig. 41a, b). In other cases ore shoots may form in dilational step-over areas
between en echelon segments (Fig. 41c).

Analysis of longitudinal sections of fault-veins


Longitudinal sections (“long section”) are views perpendicular to the plane of the vein
(Fig. 42) and can be made easily by transforming x, y, z mine data (x = easting, y =
northing, z = elevation) to longitudinal section coordinates (X, Y)(Fig. 43). Various data
can then be contoured using computer programs such as Surfer or ArcGIS. Useful data
include:
 Ore grade
 Vein thickness
 Strike
 Dip
 Metal ratios, other geochemical tracers

Two or more of these contours can then be overlaid on the same long section to test
models of ore shoot control. High vein thickness that correlates with high ore grade

36
Figure 41. En echelon patterns. A. Schematic graph showing pattern of fault displacement along
individual segments of an en echelon fault-vein array in B. B. Schematic map view of en echelon fault-
vein array; ore shoots are predicted to occur in the middle of the segments where displacement is
greatest. C. Ore shoots may also occur in step-over region where extension fractures form as horsetail
splays.

generally indicates kinematic control on ore shoots. For example high vein thickness
and/or high gold grade that correlate spatially with vein strike indicate a strike-slip fault
system with kinematically-controlled ore shoots (Fig. 44). High vein thickness and/or
high gold grade that correlate with steep vein dip indicate a normal fault-vein system.
High vein thickness and/or high gold grade that correlate with low vein dip indicate a
thrust fault-vein system.

Figure 42. Explanation of longitudinal sections. Ore shoots are high-grade areas within an ore zone.
Distances X and Y are coordinates transformed from UTM (easting and northing) and elevation.

37
Figure 43. Method of transforming easting-northing-elevation coordinates to X-Y coordinates in
longitudinal section. A. Plan view showing structure contours on fault-vein and transformation
equations. B. Cross section of vein showing parameters used in equations. C. Example of Au-grade
contoured in long section.

38
Figure 44. Example of long section with Au-grade overlaid on gray filled contours of vein strike. Inset
shows interpretation of ore shoot controlled by change in strike direction along strike-slip fault.

Drill hole targeting and design


One of the primary and extremely important tasks of the exploration geologist is to select
drill targets from the data available and then design the location of the drill platform and
the angle of inclination (plunge) of the drill hole. Ore zones in fault-veins are typically,
although not always, sub-horizontal because ore mineral precipitation is commonly
controlled by temperature and/or mixing of hydrothermal and meteoric waters (Fig. 42).
ore shoot
However, ore grade distribution along along jog

fault-veins is typically irregular and high


grade ore shoots are commonly present
and are separated by low-grade ore or rake

waste rock (Fig. 42). 90°


slickenlines

fault-vein

Ore shoot rake can be controlled by a


number of factors. These include, in
order of importance:
Figure 45. Model of kinematically-controlled ore
1. Fault kinematics, in which ore shoot along fault-vein. Slickenlines are
shoots form in opening mode perpendicular to ore shoot.
segments along the fault where jogs
or bends form. In this case ore shoot rake is perpendicular to slickenline rake (Figs.
23, 38, 39) and slickenline rake can be used to model the rake of the ore shoot (Fig.
45).

39
2. Intersection lines. Ore shoots can form along intersection lines of various features,
for example dikes (Fig. 46a), beds (Fig. 46b), contacts, or other faults. If the features
are cut by the fault-vein, the intersection lines are called cutoff lines. Formation of
ore shoots along intersection lines may be related to increased fracture damage along
the intersection line, or may be related to chemical-reactivity of cutoff lithologies
(dikes, beds, etc.). A simple spherical projection construction can be used to
determine the intersection line orientation and the drill hole orientation.
A ore shoot B ore shoot alogn
along intersection bedding cutoff line

bed

fault-vein
fault-vein

Figure 46. Model of ore shoot along fault-vein formed along intersections. A. Fault-dike intersection.
B. Fault-bed intersection.

3. Pre-existing undulations or bends along the fault (Fig. 47). Such undulations may be
related to pre-mineralization motion on the fault, and subsequent syn-mineralization
movement will cause opening along the undulations. In this case slickenline rake
cannot be used to infer the orientation of the ore shoot.

A ore shoot
along jog
B

rake
s

90°
line
ken
slic

fault-vein

Figure 47. Model of ore shoot formation along pre-existing bend in fault plane. Slickenline rake
cannot be used to predict orientation of ore shoot. A. Model showing slickenlines not perpendicular to
ore shoot. B. Model showing pre-existing undulations on fault plane that control orientation of ore
shoots; slickenlines represent a younger fault movement.

The goal of drilling programs is to design drill holes with the highest probability of
success in intersecting ore shoots. In designing the location of the drill platform relative
to the location of geochemically anomalous vein samples, most drilling programs offset,
or step, perpendicular to vein strike along the dip direction (Fig. 48a). However, this
design only works when ore shoot rake is high (>80), and many ore shoots controlled by
fault kinematics or intersections have much lower rake angles.

For dilatant, kinematically-controlled ore shoots (the first model to test), the easting and
northing coordinates of the drill target (ET, NT) can be calculated from the easting and

40
A

NE
map

slic
structure contours (depth below outcrop)

ke
nlin
missed vein

es

tr e icke
targ et ! outcrop

sl
nd nl
incorrect (0 m elev.)

of i nes
drill target
Rs = (missed)
high Au
RHR rake anomaly
samples
t X
oo
Y

(E, N) o
E
e sh t
UTM coords.

or oo
easting, northing

Ro sh
Y e 
N or
of
+
S=

north
nd RHR
tre strike
correct
drill target
+
10 0 m

(E, N) T
50m

SW

target

N
B longitudinal section
1 0 0m

50 m
s = vein strike
d = vein dip
i = drill hole inclination
v = depth to intersect vein
CNWcross section SE
X
RS = rake of slickenlines
R o = rake of ore shoot = 90+R s i d
dri ll

X = v / tan(d) in
v ve
hole

Y = X[tan(90+RS )Hcos(d)]
50m
E = Xcos(s) + Ysin(s)
N = -Xsin(s) + Ycos(s)
ET = Eo + E
NT = No + N 100m
+

Figure 48. Schematic map (A), longitudinal section (B), and cross section (C) showing method of
calculating easting and northing UTM coordinates (ET, NT) for drill target design. From Nelson (2006).

northing coordinates of the exposed ore shoot (Eo, No – at the highest geochemical
anomalies along the vein outcrop trace). Input parameters are: 1) vein strike (s), 2) vein
dip (d), 3) the slickenline rake angle (Rs), and 4) the depth to intersect the vein (v) (Fig.
48). If the ore shoot rake is controlled by an intersection, then the rake of the ore shoot
can be determined using an equal angle net (Fig. 49); this rake angle (Rc) is then
substituted directly for the term (90-Rs) in the equation for Y (Fig. 43). Appendix III
illustrates the input parameters needed for this calculation, and presents a sample
spreadsheet for making the calculations.

Vein dip (d) and slickenline rake (Rs) are measured in the field. Before calculating the
drill target location, a geologic cross section of the vein should be made first to determine
the vein intersection depth (v). Once the target location is calculated, the drill hole

41
inclination (i) and trend are determined from the cross section, but may be controlled by
topographic constraints.

If slickenlines are not available, the rake of


ore shoots that form in the fault-vein
perpendicular to the slip vector can be

5)
inferred using the intersection line between

( 020/6
the fault-vein and second-order fractures of

ein
Riedel type (Fig. 50), or from a number of intersection =
predicted

fau lt- v
other methods summarized in Table 2 and ore shoot

bed
described in detail in Nelson (2006).

din
(1

g
00
/ 30
A spreadsheet for calculating drill target )

location is presented in Appendix III.


rake angle

Figure 49. Schematic example of construction of


rake angle of ore shoot in fault-vein using model
of fault-bedding intersection as control of ore
shoot orientation. Rake angle = 34.

Slip-related strain features


Structural feature or method Relation to slip line
slickenlines, mullions parallel
slickenfibers parallel
slickolites (oblique stylolitic lineations) parallel
second-order fractures (Riedel fractures) intersection line with fault is perpendicular
S-C type shear fabrics S- and C-plane intersection line is perpendicular
shear fold hinge line commonly perpendicular
mineral elongation lineation parallel

Indirect methods of slip line determination


asymmetric fold analysis (Hansen method)
piercing point analysis - orthographic or combined orthographic/stereographic construction
3-D computer modeling of ore shell
Table 2. Summary of common structural features and methods useful in determining slip-vector
orientation. See text for explanation of indirect methods.

42
Figure 50. Equal area nets showing modeling of ore shoot orientation using intersection line between
fault-vein (F) and second-order (Riedel) fractures (R, R’, P, or T). Cross sections illustrate predicted
orientation of Riedel fractures for thrust and normal faults (45 dip). Lower right net shows general
example with oblique-slip fault-vein.

Structural data in drill core


Drill core is an important, albeit small, sample of the
mineralized system at depth. Therefore analysis of
structures in core is essential, especially in areas of poor
outcrop. Core recovered from drill holes is either
oriented or un-oriented relative to spin around the core
axis (Fig. 51). Oriented core from non-vertical holes
contains a line drawn along the core (parallel to the core
axis, CA) on either the top or the bottom of the core, and
an arrow on the line indicating down-hole direction. For
un-oriented core the bottom-of-hole position is not
known. Figure 51. Un-oriented (A) and
oriented (B) core.

43
Un-oriented core
The true orientation of structures in un-oriented core in some cases can be determined or
constrained by inspection if certain criteria are met.
1. If the borehole is vertical, the dip of structures is known, but not the strike.
2. If the planar feature is perpendicular to the core axis, strike and dip can be determined
easily (Fig. 52a).
3. If the borehole is perpendicular to strike, dip can have one of two possible values, and
one can normally be assumed as correct from knowledge of local geology (Fig. 52b).
4. If nothing is known about the plane orientation, then possible pole orientations form a
cone (Fig. 52c) which is represented on a stereonet as a small circle about the core
axis (Fig. 52d).

Figure 52. Un-oriented core. A. Plane orientation is known if plane is perpendicular to core axis (CA).
B. Two plane orientations are possible if strike of plane is known from surface geology. C. Lacking any
information on plane orientation, possible pole orientations form a cone. D. Stereonet showing cone
(small circle) of all possible pole orientations for a drill hole oriented 60°/130° and  = 30°.

In other cases, the true plane orientation can determined if a number of special circumstances are
met:
1. Three drill holes of different plunge/trend intersect the same planar feature (e.g., vein,
fault, beds), and it is assumed that the planar feature is parallel in all holes (that is,
there is no folding or fault-block rotation between the holes)(Fig. 53a). The
orientation of the planar feature can be determined using spherical projection (Fig.
53b); this is called the ‘3-hole method’. If only two holes intersect the planar feature,
the orientation can be constrained to two possible solutions.

44
Figure 53. Method of determining a plane orientation from cores in three differently-oriented
boreholes. A. Schematic cross section. B. Three-hole stereonet method.

2. A planar feature in the core can be correlated to planar features of known orientation
(e.g., bedding that is regionally consistent in orientation); this is called the reference
plane (Fig. 54). It is assumed that there is no folding or fault-block rotation between
the location of the core and that where the orientation of the reference plane was
determined (e.g., in outcrop). The orientation of a planar feature of unknown
orientation in the same core can then be determined through spherical projection
(‘stereonet’) or by calculation. This is called the ‘reference plane method’. The
stereonet method involves these steps (Fig. 54):
a. Measure in core the angles  (= core axis ^ plane) and  (= angle between known plane
and unknown plane measured clockwise looking down hole; Fig. 54a). Calculate =
90-(Fig. 54b)
b. Plot plunge/trend of borehole, BH.
c. Plot reference plane pole, Nr.
d. Temporarily rotate BH to vertical; the rotation axis is north on the net with BH on the E-
W line (BH rotates to center of net). Using the same rotation axis, rotate Nr along the
small circle it is on; rotate the same number of degrees in the same direction as BH. This
gives Nr’ (the prime (‘) refers to orientations while the core is temporarily vertical.
e. From the azimuth of Nr' count degrees clockwise to find the azimuth of Nu’.
f. In the direction of the Nu’ azimuth, count  degrees out from net center to Nu'.
g. Rotate BH’ back to its original position (you are “placing” the core back in the
orientation of the hole); rotate Nu' the same number of degrees about the same rotation
axis and in the same direction. This gives Nu (the pole to the unknown plane, from
which you can now plot the great circle and determine the strike and dip).

45
Figure 54. A. Un-oriented core showing plane of known orientation (reference plane) and angle () to
plane of unknown orientation. B. Cross section of core showing  and  angles and protractor method
of measurement. C. Steps in stereonet solution to unknown plane orientation; data given in first step.

Oriented drill core


The true orientation of structures (planes and lines) seen in oriented drill core can be
determined using two possible methods: 1) direct measurement, and 2) the indirece -
method. The direct method uses geological-compass measurement with the core in a
core-orienting device (Fig. 55). The indirect method involves measuring two angles in
core ( and ), and then converting these angles to true orientation either by spherical
projection (‘stereonet’) methods or by calculation (e.g., using Geocalculator.exe -
http://holcombe.net.au/software/rodh_software_geocalc.htm). Although the - method is now widely
used, the direct measurement method has the advantage that true structural orientations
are immediately available during logging and the geologist can use these data to test
models think about structural controls on mineralization in the system. The disadvantage
of the indirect method is that  and  data often sit idle in a database or spreadsheet
without calculation of true structural orientation.

Planar features intersect core as ellipses (a circle if the plane is perpendicular to core axis
(CA)(Fig. 56). The long axis of this ellipse is designated E-E’, with E’ as the up-hole end
of the line. The apical trace of the ellipse is the line along the core where the plane
containing the core axis, the pole to the plane, and line E-E’ intersects the core; this line
is drawn on the core parallel to the core axis and passing through point E or E’.

46
Angles to be measured in core include:
= acute angle between core axis and long axis of ellipse (E-E’ line)(Fig. 54b),
= acute angle between core axis and pole to plane (Fig. 54),
= 0-360° angle measured along the core circumference clockwise from the reference
line (bottom-of-hole line, BOH) the apical line drawn from point E (usual convention) or
E’ (Fig. 56a),
 = (for a geological line lying within a plane) angle measured in the plane from E-E’ to
the line (see diagram at right)(Fig. 56b).

Figure 55. Core orienting devices. A. Sophisticated (James Cook Univ., Australia –
(http://www.laingex.com/corekit.html.) B. Simple (Carajás, Brasil). Both types work well.

Figure 56. A. Features of oriented core;  angle is always measured clockwise looking down hole. B.
Definition of  (gamma) angle from ellipse axis to geological line measured clockwise looking down hole.

47
Steps in determining the orientation of a plane in oriented core using the stereonet
method are as follows (Fig. 57):
1. In core, measure  angle, and  angle (convert to  = 90°-).
 = angle measured around core from BOH line to ‘apical line’ of ellipse.
 = angle from core axis (CA) to long axis of ellipse.
2. Plot core axis CA with plunge and trend of drill hole.
3. Plot the circumferential plane (^ to CA).
4. Plot the dip line (a) 90° from CA in the drill section plane (also on the circumferential plane).
5. In the circumferential plane, measure angle b clockwise from line a to line b (line b is line E-E’
projected onto the circumferential plane).
6. Construct the b-CA great circle (this is the CA-N plane).
7. Using the rules below, measure  degrees along the CA-N plane from CA to N.

Figure 57. Stereonet method to determine plane orientation using  and  angles in oriented core. A. Core
showing angles and lines. B. Example stereonet solution using data shown. C. Rules for measuring 
angle on stereonet (step 7).

48
Structural Methods in Mineral Production
Characterization of rock mass structure
Structural data are very important in characterizing rock mass structure for geotechnical
purposes in both surface and underground mining. Such purposes include pit slope and
underground mine design, slope stability analysis, hydrogeology, and mine facility design
(e.g., infrastructure, tailings and mine waste facilities, etc.). For slope stability analysis,
the importance of including structural data in geotechnical analysis was noted by
Nicholas and Sims (chapter 2 in Hustrulid et al., 2001):

The adverse interaction of geologic structures with the mine walls is the greatest
contributing factor to slope instability in open-pit mines, and the success of slope-
stability analysis depends upon the level of understanding of the characteristics of
geologic structure throughout the deposit.

For underground mining the importance was noted by the Department of Industry and
Resources of Western Australia (1997):

The important role that geological structures have in ground control cannot be
over-emphasized. Thorough investigation and analysis of geological structure is
vital to a good understanding of the major influence that geological structure
exerts in determining the ground conditions in underground mining.

Characterization of rock mass structure for geotechnical purposes involves two principal
data sources: 1) intact rock strength characterized by rock mechanical parameters such as
unconfined compressive strength (UCS) that are derived from laboratory testing of core
samples, and 2) geologic structure defined from field measurements.

Geologic structure refers to any planes of weakness in the rock mass that pre-date mining
activity. Such planes are here classified into two general categories: 1) discontinuities,
and 2) penetrative fabrics. Discontinuities are significant mechanical breaks of negligible
tensile strength that separate blocks of intact rock that have relatively significant tensile
strength. They divide the rock mass up into blocks, the size, shape, and orientation of
which strongly influence rock stability conditions. Discontinuities include joints7, faults,
bedding planes, veinlets, and geologic contacts (including the margins of veins and
dikes). Whereas joints, veinlets, and bedding planes are generally mesoscopic in scale,
contacts, dikes, and faults can constitute major structures which have longer trace lengths
(commonly defined in open pits as crossing two or more benches).

Penetrative fabrics are weakness planes that are pervasive in the rock at a mesoscopic
scale (hand sample). Such fabrics include foliation, which is a general term that includes
cleavage and schistosity, but also finely bedded sedimentary rocks such as shale and
marl.
7
Joints are commonly, but incorrectly, called fractures. Fractures can be either shear fractures (faults) or
extension fractures. Extension fractures can be filled with mineral cement, igneous material, or breccia and
would thus be classified as veins, dikes, or breccia dikes, respectively.

49
Rock mass classification systems
Rock mass classification and rating systems are important during feasibility and
preliminary design stages, as well as during mining. At least two methods should be
used, as different classification systems place emphasis on various parameters. It is
important to understand the limitations of rock mass classification schemes (see
Palmstrom, 2005; Palmstrom and Broch, 2006, and in particular, Pantelidis, 2009). Also,
most systems were designed for tunneling, but have been applied (probably
inappropriately) to open cut stability analysis.

Types of discontinuities and structures


Many types of planar discontinuities and other structures exist that can impact rock mass
strength. Joints are the probably the most important discontinuity type because, by
definition, they are open, unfilled fractures (joints can be filled with iron oxides from
meteoric weathering), and nearly all rock volumes contain joints. Discontinuity types
and definitions are given in the table below.
Planar discontinuities
bedding layering of sedimentary and/or volcanic beds
joint extension fracture
vein fracture with mineral fill
fault shear fracture (single slip plane)
dike tabular igneous bodies in extension fractures
foliation, cleavage metamorphic fabric of aligned minerals
coal cleat face cleat (more continuous), butt cleat
Other structures
fault zones brittle displacement zone containing cataclastic (broken rock) fill
ductile shear zones ductile displacement zone containing ductile sheared rock
folds bent beds or foliation causing variations in structural orientation

Characterizing structural discontinuities


Rock mass structure classification systems require quantitative and qualitative
parameterization of discontinuities. Structural parameters are recorded in pit or
underground mapping using scanline (Fig. 58) or cell mapping methods (figure below
right). These features are then used to estimate rock mass ratings (such as RMR) which
can be used in slope design and stability analysis (see Pantelidis, 2009 for comparison
and discussion of various rating systems). Note discontinuity parameters have not been
applied to penetrative fabrics on rocks. Typical parameters are shown in the table below.

orientation usually dip/dip direction, but can convert from strike/dip


spacing, frequency, or block size various measurements possible (see below)
length length of exposed discontinuity
roughness see figure 57
aperture record width of open fracture; see figure 57
filling record width and filling; see figure 57
termination 0, 1, 2 as seen in face

50
discontinuity roughness: scanline example:

discontinuity aperture, filling:

Figure 58. Illustrations of structural parameters (discontinuity roughness and aperture) and scanline
example.

Discontinuity spacing and block size


Joints divide the rock into blocks of variable size; various estimates of joint spacing or
block size exist (Fig. 59). Fractures induced by blasting or drilling (for core
measurements) should not be included.
1. Joint spacing (S) – average perpendicular distance between joints in a joint set. Multiple sets are
difficult to calculate spacing; therefore the lowest-spaced set should be used.
Sa = 3√Vb Vb = block volume in m3

Figure 59. Illustrations of joint spacing and joint sets.

2. Block volume (Vb) – for 3 joint sets with spacing Sn and angles between the joint sets n.
Vb = (S1 x S2 x S3)/(sin1 x sin2 x sin3) Table shows the block volume for some angles between the
joint sets:

3. Volumetric joint count (Jv) – number of joints intersecting a volume of 1m3; best used where well-
defined joint sets occur. All joints seldom can be counted, so Jv is often given as a range.
Jv = 1/S1 + 1/S2 + 1/S3 + … 1/Sn (for random joints add Nr/5√A; where Nr = number of
random joints in area, Am3). Classification of Jv:

51
Rock Mass Classifications
1. Rock Quality Designation index (RQD) Deere et al., 1967
Provides quantitative estimate of rock mass quality from drill core, rock face transect.
RQD = % of intact core pieces >10cm in total length of core. RQD may be estimated
from volumetric joint count; the suggested relationship for clay-free rock is:
RQD = (115 - 3.3 Jv). The most important use of RQD is as a component of the RMR
and Q rock mass classifications. Examples of RQD determination are shown in figure
60.

Figure 60. Examples of RQD determination in core (left) and outcrop (right).

RQD limitations include: 1) RQD value changes with borehole orientation, and 2) RQD
= 0 for core length with all pieces <10cm long and RQD = 100 where all pieces are
>11cm long (Fig. 60).

Figure 60. Illustrations of directional dependence on RQD (left) and RQD minimum and maximum values (right).

52
2. Rock Mass Rating (RMR) system Bieniawski (1976, 1989)
Six parameters are used to estimate rock mass strength:
1. Uniaxial compressive strength of rock material
2. Rock Quality Designation (RQD)
3. Spacing of discontinuities (also called fracture frequency or fracture density)
4. Condition of discontinuities (combines length, roughness, aperture, fill, and wallrock weathering)
5. Groundwater conditions
6. Orientation of discontinuities

The rock mass is divided into structural domains (e.g., separated by faults or geologic
contacts, or with significantly different joint spacing); each domain is classified
separately. Ratings for each category are shown in the table below.

RMR Rock quality


0 – 20 Very poor
21 – 40 Poor  
41 – 60 Fair
61 – 80 Good
81 – 100 Very good

Rock mass
3. Rock Mass Quality Index, Q Barton et al., 1974
n
RMR (%)
(Tunneling Quality Index or Norwegian Q system) Exceptionally 0.001 - 0.01 0-3
Extremely poor 0.01 - 0.1 3 - 23
Very poor 0.1 - 1 23 - 44
parameter range:
Poor 1-4 44 - 56
RQD = Rock Quality Designation 100 – 10
Fair 4 - 10 56 - 65
Jn = Joint set number 1 – 20
Good 10 - 40 65 - 77
Jr = Joint roughness factor 4–1
Ja = Joint alteration and clay fillings 1- 20 Very good 40 - 100 77 - 85

Jw = Joint water inflow or pressure 1 – 0.1 Extremely good 100 - 400 85 - 98

SRF = stress reduction factor 1 – 20 Exceptionally 400 - 1000 98 - 100


d

• (RQD/Jn) = crude measure of block size


• (Jr/Ja) = interblock shear strength (roughness and friction of surfaces)
• (Jw/SRF) = active stress (water pressure and total stress parameter)

In many cases, a range of parameter values leads to a range of Q values (see example in
Hoek, http://www.rocscience.com/hoek/pdf/3_Rock_mass_classification.pdf).

4. Geological Strength Index (GSI) Hoek and Marinos (2000)


As RQD in most weak rock masses is zero or meaningless, the qualitative GSI
classification places greater emphasis on basic geological observations of rock-mass
characteristics. GSI does not replace RMR or Q systems; its function is to estimate rock-
mass properties. The GSI index is based on a matrix of two parameters: structure

53
(blockiness) and discontinuity surface quality visually estimated from outcrops,
excavation faces, and core (Fig. 61).

Notes:
• Some quantified GSI classifications have been proposed (see chart), but are only good
in rock masses where discontinuity frequency and orientation are easily measured.
• Do not use GSI for rock mass with highly anisotropic mechanical behavior (e.g.,
preferred structural orientation in foliated rock).
• Do not use GSI for strong hard rock mass with few discontinuities with spacing similar
to dimensions of tunnel or slope.

Figure 61. General qualitative GSI classification (left) and Quantitative GSI classification (right; Sonmez, 2001)

Slope stability analysis


Pit slope failure can occur in four basic modes (Fig. 62): slips (essentially landslides
above either planar or circular slip surfaces), wedge failures, toppling failures, or failure
by cataclastic flow (essentially earth materials whose deformation can be described by
viscous flow models).

54
Figure 62. Main types of pit slope failures (not shown: cataclastic flow).

Modeling the risk of slope failure can be done on domains of various sizes, including the
scale of an entire pit, of an individual pit slope, or of a particular site recognized to be a
potential hazard. Nelson et al. (2007) developed a GIS-based method of slope stability
analysis utilizing models that incorporated various component layers at a relatively large
map scale (1:5000). Component layers included alteration, geotechnical unit, proximity
to major faults, GSI (geological strength index), slope (from digital elevation model),
proximity to water table, and a composite structural density grid. Example models are
shown in figure 63.

Figure 63. Two predictive slope-failure risk maps derived from GIS modeling of the Chuquicamata open pit
mine, Chile; red shows highest risk areas (Nelson et al., 2007). The location of 2006 pit failures are
superimposed on the models that best predicted the failure areas; triangles show areas of relatively small wedge
and planar failures, and hatch patterns show zones of toppling failures.

55
Slope stability analysis of individual slope domains can be done using kinematic analysis
on a stereonet (Figs. 64-66). The analysis compares the orientation and inclination of the
pit slope with the orientations of discontinuity sets. Instability fields are defined by a
combination of the slope orientation, daylight8 conditions, the friction angle of the intact
rock, and the minimum assumed variance of discontinuity dip direction or wedge base
line trend with respect to the slope aspect9 (Fig. 65). For planar failure analysis,
instability fields can be constructed to analyze either poles to discontinuity planes (see
Fig. 67) or dip lines of discontinuity planes (Fig. 64 a, 65). For wedge failure analysis,
the instability field is constructed to plot wedge base lines defined by the intersection of
average discontinuity planes of various sets (Fig. 65, 67). For toppling failure the
instability field is constructed to plot poles to discontinuities (Fig. 66)

Instability conditions exist for sliding failures if discontinuities daylight and strike within
about ±20° of slope strike (Fig. 65c). Instability conditions exist for wedge failures if the
wedge base line daylights and trends within about ±20° of slope aspect (dip direction)
(Fig. 65c). Instability conditions exist for toppling failures if discontinuity planes dip
steeply and opposite to, and within about ±20° of, the slope dip direction (Fig. 66, 67).
Kinematic analysis can be done manually on a stereonet or using computer programs
such as DIPS. An example of a DIPS analysis is shown in figure 68.

Slope stability analysis for individual sites with recognized potentially-hazardous major
structures can be done using the orientation of major structures, properties of the intact
rock, and various engineering programs.

8
Daylight condition = discontinuity dip line or wedge base line plunges less than dip of cut slope.
9
Note that this angle can vary between 10° and 30° depending on the publication. An average angle of 20°
has been chosen here.

56
Figure 64. Graphics showing critical dip angle for planar slips (A) and plunge angle for wedges (B)
defined by friction angle. Friction cone as 3D graphic (C) and represented on stereonet (D).

Figure 65. Stereonets showing development of unstable field for planar slips and wedge failures based on
(A) friction cone, (B) daylight envelop. (C) Instability field (for planar surface dip line or wedge base) is
defined by intersection of friction cone and daylight envelop, and reduced by ±20° variation of discontinuity
dip direction or wedge baseline trend from slope aspect.

57
Figure 66. Stereonet showing kinematic analysis for toppling failures.

Figure 67. Stereonets showing instability fields for the principal failure types (note: planar failure
instability field is for plotting poles to discontinuities, not dip lines).

Figure 68. Stereonets showing example of DIPS kinematic analysis.


http://www.rocscience.com/downloads/dips/webhelp/pdf_files/tutorials/Tutorial_03_Toppling_Planar_and_Wedge_Sliding_Analysis.pdf

58
Structure of High-sulfidation Epithermal Deposits
High-sulfidation (HS) epithermal mineralization is associated with vapor and metal- rich
saline fluids exsolved from crystallizing high-level calc-alkaline magmas. The gas-rich
(HCl and SO2+ H2S) vapors condense in meteoric water to form acidic fluids that leach
host rocks to form enhanced permeability zones used by later hydrothermal mineralizing
fluids. Many high-sulfidation deposits are syn-volcanic, meaning that they formed in the
same magmatic arc that formed the host rocks; mineralization ages are typically between
0.5 to 5 m.y. younger than host rock ages. These deposits are generally hosted by
intermediate to silicic pyroclastic and flow rocks, and related flow-domes and cross-
cutting breccias (phreatic, phreatomagmatic, tectonic, diatreme-related, etc.).

Deposit examples include Yanacocha, Pierina, Summitville, La Coipa, Pascua-Lama, and


Nansatsu. Other HS deposits are hosted in older volcanic or non-volcanic rocks.
Examples include Triassic granitic rocks at Pascua-Lama, Triassic sedimentary basement
at La Coipa, and basic volcanic rocks and sedimentary rocks at Pueblo Viejo (according
to Sillitoe et al., 2006). The table below shows a summary of structural and other
features of some important HS mineralized systems.

Structural controls in HS systems may include fault zones and related extension fractures,
volcanic dome and diatreme margins, and permeable strata (which may have controlled
massive silica zones which act as aquitards to trap hydrothermal fluids). Breccia bodies
are commonly mineralized in or along their margins. Such breccias may have irregular
geometries, but subtle alignments nonetheless indicate structural controls. Breccia bodies
may form originally as fault (tectonic) breccias, by magmatic or phreatomagmatic
eruptions along fault- or fault-intersection-controlled conduits, or by diatreme activity.
The term hydrothermal breccia only means a breccia with hydrothermal minerals in the
matrix, and can form during breccia formation or after breccia formation.

Structural controls, although obvious in some systems (e.g., quartz-enargite veins at


Chipmo, Peru), are usually difficult to identify in detail. This is because of two main
reasons: 1) the high intrinsic (inter-granular) permeability of relatively young volcanic
and volcaniclastic host rocks causes dissemination of fluids evenly throughout the host
rock package, and 2) the high acidity of hydrothermal fluids causes intense alteration that
destroys much of the rock texture and probably many structures that may have controlled
fluid flow.

Conventional structural mapping to identify structural controls involves focusing on those


structures that contain hypogene hydrothermal fill (veins, veinlets, breccia zones, fault-
veins). Because HS systems typically have much disseminated pyrite, a strong supergene
phase overprint results in many Fe-oxide filled veinlets10. However, in the absence of
boxwork structure (Fe-oxide pseudomorphs of hypogene sulfide minerals), such veinlets
10
Although veinlet is used here as these are mineral-filled fractures, in some usage the word veinlet is
restricted to fractures with hypogene mineral fill.

59
should not be used to model the structural controls of hypogene mineralization. Many
such veinlets are actually fractures formed in a stress field that is much younger than the
ambient stress field during hypogene mineralization.

Deposit Deposit age Host rock Structure Reference


Yanacocha, Miocene: syn- Miocene volcanic complex; NE-trend (~055°) belt of alteration and deposit Bell et al.,
Peru volcanic tuffs, phreatic and alignment with NW-strike cross faults (parallel to 2009
magmatophreatic breccias, Cretaceous fold-thrust belt in basement)
domes
Pueblo Viejo, 1. Early Early Cretaceous Los Ranchos Ore bodies mushroom shaped and structurally Sillitoe et al.,
Dominican Cretaceous (syn- Fm. Most in shallow dip controlled, fault feeders “apparent”. Within 10km- 2006
Republic volcanic; Muntean carbonaceous siltstone- long WNW-trend belt of advanced argillic Muntean et al.,
et al.); 2. Late sandstone; some in alteration; pits NS trend in left-step en echelon 2007
Cretaceous-early conglomerate and basalt- pattern.
Tertiary (Sillitoe et andesite flows.
al.)
Summitvilile, syn-volcanic ? Tertiary volcanic dome, quartz Near intersection of NW-strike fault system and Gray and
Colorado USA latite caldera ring zone. High-grade veins (>10g/t) have Coolbaugh,
crude en echelon pattern with ore body trends: 1994
300° and 330°; rough radial pattern of ore zones
(?).
La Coipa, 17-23 Ma, syn- Triassic shale-sandstone, Intersection of NNE horst with NW-strike normal Oviedo et al.,
Chile volcanic Tertiary (20-25 Ma) rhyolite- faults. Ore bodies: steep, some tabular in 1991
dacite tuff, tuff breccia sedimentary rocks (high Au); mushroom shape in
sedimentary and pyroclastic rocks (high Au).
Steep alunite-gypsum-limonite veins (mm to
30cm) NNW, NNE, EW.
Pierina, Peru middle Miocene, Hypabyssal to extrusive ‘pumice ? Rainbow et al.,
syn-volcanic tuff’ and older underlying dacite 2005
flow-dome complex, both cut by
hydrothermal breccia and
dacitic domes.
Nansatsu 10 Ma Pyroclastic and lava flows Hedenquist et
district, Japan al., 1994
Veladero, late Miocene, syn- Mostly late Oligocene and Steeply dipping fold and thrust belt of interleaved Charchaflié et
Argentina volcanic (~5 m.y. middle Miocene volcanic, Oligocene and late Paleozoic rocks. Ore bodies in al., 2007
younger than host volcaniclastic rocks (~16 Ma); a tabular, NNW elongated silicified horizon that
volcanics) also underlying Permian cuts across the host stratigraphy with
rhyolitic pyroclastic rocks. subhorizontal top. Local extension related to
gravitational collapse of volcanic orogenic belt.
Hog Heaven, Oligocene; close Flow domes, dikes, ash flow Open-space fill, replacement, vein and stockwork, Lange et al.,
Montana USA temporal deposits, volcaniclastic rock and diatreme-hosted; NNW-, NNE-, and NW- 1994
relationship to trending extensional faults, fractures
dome
emplacement
Hope Brook, Proterozoic; 578- Whittle Hill Sandstone intruded Hydrothermal alteration 3 km long up to 400 m Dubé et al.,
Newfoundland, 574 Ma by Late Proterozoic quartz- wide. Dike-sill complex localized deposit; 19995
Canada feldspar porphyry sill-dike lithological control?
complex
Pascua-Lama, Mid to late late Paleozoic and mid-Tertiary disseminated and stockwork type Deyell et al.,
Chile Miocene intrusive and volcanic rocks 2005

Table: Summary of features of selected high-sulfidation deposits.

60
Where intense alteration has destroyed original rock structures, conventional structural
mapping and structural orientation techniques may not identify important structural
controls. In this case, other methods can be employed:
 3D computer visualization of ore shell models made from blast-hole assay data or
exploration drilling data to identify linear and/or planar alignment of relatively
high grade ore zones. These models can then be compared with structures defined
by pit or pre-pit surface mapping and drill-hole data. Oriented core may be used
to make correlations with surface structures.
 Sulfide-bearing quartz veinlets, typically preserved in stockwork zones adjacent
to massive or vuggy silica altered zones, should be measured and analyzed with
stereonet analysis, and then compared with known ore zone alignments from
mapping and 3D models.
 Lithological and structural alignments in basement and cover strata are important.
Lineament analysis can be performed on lithological maps, alteration maps, and
geophysical maps.
 Pit elongation analysis, although simple, can be very useful, as is alignment
analysis of multiple pits and prospects.

References cited
Allmendinger, R.W., Marrett, R.A., and Cladouhos, T., 2001, FaultKin Version 1.2.2 (for Windows), computer
program and user’s manual (for v.1.1), 29 p., http://www.geo.cornell.edu/geology/faculty/RWA/programs.html.

Anderson, E.M., 1905, The dynamics of faulting, Trans. Edinburgh Geol. Soc., 8 (3), 387-402.

Anderson, E.M., 1951, The Dynamics of Faulting and Dyke Formation With Application to Britain, 2nd ed., 206 pp.,
Oliver and Boyd, Edinburgh.

Arévalo et al., 2006, Structural setting of the Candelaria Fe oxide Cu-Au deposit, Chilean andes (27°30’ S), Economic
Geology, v. 101, p. 819-841.

Barton, N., Lien, R. and Lunde, J. (1974): Engineering classification of rock masses for the design of rock
support. Rock Mechanics 6, 1974, pp. 189-236.

Bieniawski Z.T. (1989): Engineering rock mass classifications. John Wiley & Sons, New York, 251 pp.

Caine, J. S., Nelson, E.P., Beach, S.T., and Layer, P.W., 2006, Structural Fabrics, Mineralization, and Laramide
Kinematics of the Idaho Springs-Ralston Shear Zone, Colorado Mineral Belt and Central Front Range Uplift,
Mountain Geologist, v.43, No. 1, p.1-24.

Davis, G.H., and Reynolds, S.J., 1996, Structural geology of rocks and regions, 2nd edition, New York, Wiley, 776 p.

Echavarría, L., Nelson, E., Humphrey, J., Chávez, J., Escobedo, L., and Iriondo, A., 2006, Geological evolution of the
Caylloma epithermal vein district, southern Perú, Economic Geology, v. 101, No. 4, p.843-863.

Groves, D.I., Bierlein, F.P., Meinert, L.D., and Hitzman, M.W., in press, Definition of iron oxide-copper-gold deposits
and proposed associated ore types and their distribution through Earth history.

Haneberg, W.C., 2008, Using close range terrestrial digital photogrammetry for 3-D rock slope modeling and
discontinuity mapping in the United States, Bull Eng Geol Environ, v. 67, p. 457-469 (DOI
10.1007/s10064-008-0157-y).

Hustrulid, W.A., McCarter, M.K., and Van Zyl, D.J.A. (editors), 2001, Slope Stability in Surface Mining, SME, 456p.

61
Lisle, R.J. and Leyshon, P.R., 2004, Stereographic Projection Techniques for Geologists and Civil Engineers, 2nd ed.,
Cambridge University Press, 112p.

Marrett, R. A., and Allmendinger, R. W., 1990, Kinematic analysis of fault-slip data: Journal of Structural Geology, v.
12, p. 973-986.

Miller, J. McL., Nelson, E.P., Hall, W.D.M., and Hitzman, M.W., 2007, Orthorhombic fault-fracture patterns from
triaxial non-plane strain in an accommodation zone during rifting: Lennard shelf, Canning basin, Western Australia,
Journal of Structural Geology, (doi:10.1016/j.jsg.2007.01.004), v. 29, p. 1002-1021.

Department of Industry and Resources of Western Australia, 1997, Geotechnical considerations in underground mines,
Document No.: ZME723QT, 67p,
http://www.dmp.wa.gov.au/documents/Guidelines/MSH_G_GeotechnicalConsiderationsUGMines.pdf

Muller, O. H. and Pollard, D. D.: 1977, 'The Stress State near Spanish Peaks, Colorado, Determined
from a Dike Pattern, Pageoph. 115, 69-86.

Nelson, E.P., 2006, Drill-hole design for dilational ore shoot targets in fault-hosted veins, Economic Geology, v. 101,
p.1079-1085.

Nelson, E.P., Connors, K.A., and Suárez, C., 2007, GIS-based slope stability analysis, Chuquicamata open pit copper
mine, Chile. Natural Resources Research, (DOI: 10.1007/s11053-007-9044-7), v. 16, p.171-
190.

Nelson, E.P., Hitzman, M.W., and Monteiro, L.V. S., 2007, Hydrothermal metamorphism and transient depth
fluctuations of the brittle-ductile transition during shear-zone hosted IOCG mineralization, Geological Society of
America, Abstracts.

Odé, H.: 1957, 'Mechanical Analysis of the Dike Pattern of the Spanish Peaks Area, Colorado', Geol.
Soc. Am. Bull. 68, 567-576.

Palmstrom, A., 2005. Measurements of and correlations between block size and rock quality designation
(RQD). Tunnel. Underground Space Technol. 20 (4), 362–377.

Palmstrom, A. and Broch, E., 2006, Use and misuse of rock mass classification systems with particular
reference to the Q-system, Tunneling and Underground Space Technology, 21, 575–593.

Pantelidis, L., 2009, Rock slope stability assessment through rock mass classification systems, International
Journal of Rock Mechanics & Mining Sciences, v. 46, p. 315–325.

Sibson, R.H., 1996, Structural permeability of fluid-driven fault-fracture meshes, Journal of Structural Geology, v. 18,
no. 8, p.1031-1042.

Sibson, R.H., Robert, F., and Poulsen, K.H., 1988, High-angle reverse faults, fluid-pressure cycling, and mesothermal
gold-quartz deposits, Geology v. 16; no. 6; p. 551-555.

62
Appendix I –Structural database format
Explanation
WP GPS waypoint - site number
vein vein or site name
lith lithology: rhy = rhyolite, and = andesite
code v = vein, vf = fault-vein, f = fault, foln = foliation, fold = fold hinge, b = bedding, d = dike
strike strike of plane in right-hand-rule format
dip dip of plane
rake rake angle of line in a plane, 0-90
RHR* rake rake angle of line in a plane, 0-179, measured from RHR strike
trend trend of line
plunge plunge of line
sense fault slip sense: n = normal, r = reverse, d = dextral, s = sinistral
qual fault slip sense determination quality: a = good, b = moderate, c = poor

DATE EASTING NORTHING Elev WP Vein Lith code strike dip rake RHR rake trend plunge sense qual COMMENTS
6/16/2007 809556 1844179 761 1 office
6/16/2007 808918 1846488 922 2 Baja rhy v 314 90 drusy qz vlet
6/16/2007 808918 1846488 922 2 Baja rhy v 305 90 drusy qz vlet
6/16/2007 808918 1846488 922 2 Baja rhy v 285 90 drusy qz vlet
6/16/2007 808918 1846488 922 2 Baja rhy v 325 90 drusy qz vlet
6/16/2007 Code: 1846488 922
808918 2 Baja rhy v 309 80 drusy qz vlet
6/16/2007 v = vein1846546 935
808889 3 Baja rhy v 310 45 drusy qz vlet
vf = fault vein
6/16/2007 808874 1846559 946 4 Baja rhy fol 025 32 cmm qz vlet
f = fault
6/16/2007 808749 1846681 1022 5 Arista rhy fv 156 42 65 65 214 37 s b qz fault-vn
foln = flow foliation
6/16/2007 808767 1846643 1020 6 Arista rhy v 152 35 white qz vlet
b = bedding
6/16/2007 808767
d = dike1846643 1020 6 Arista rhy v 323 90 white qz vlet
6/16/2007 808767 1846643 1020 6 Arista rhy v 083 73 white qz vlet
6/16/2007 808767 1846643 1020 6 Arista rhy v 140 45 22 cm milky qz vn
6/16/2007 808783 1846619 1016 7 Arista rhy vn 287 86 main vein - 40 cm
6/16/2007 808783 1846619 1016 7 Arista rhy fv 329 70 26 26 338 24 adjacent vn
6/16/2007 808786 1846609 1014 8 Arista rhy v 295 60 qz vlet
6/16/2007 808786 1846609 1014 8 Arista rhy v 113 88 qz vlet
6/16/2007 808786 1846609 1014 8 Arista rhy v 228 90 qz vlet
6/16/2007 808788 1846632 1013 9 Arista rhy v 306 85 3mm qz vlet
6/16/2007 808777 1846635 1017 10 Arista rhy bx 008 77 20 cm volc-clast bx volc matrix
6/16/2007 808765 1846672 1024 11 Arista rhy fv 332 76 30 30 340 29 s c HW of Arista vein
6/16/2007 808761 1846333 927 12 Dogleg and fv 144 87 76 76 156 76 n b main vein in adit
6/16/2007 808904 1846220 915 13 Dogleg-S and vn 100 85 30 cm bx vein, gray Si matrix
6/16/2007 808904 1846220 915 13 Dogleg-S and b 178 38 andesite
6/16/2007 808740 1846334 954 14 Dogleg-S and fv 308 90 5 175 128 05 s b main vein
6/16/2007 808740 1846334 954 14 Dogleg-S and v 280 82 15 cm main vein
6/16/2007 808671 1846378 960 15 Dogleg-S and vn 115 60 50 cm gray Si vein
6/16/2007 808671 1846378 960 15 Dogleg-S and v 115 85 bx vein, Ag, Cu? Sulfides

63
Appendix II – Rake to plunge/trend calculation

64
Appendix III – Calculation of drill target location
Eo No s d Rs v 90+Rs sin(s) cos(s) cos(d) tan(d) tan(90+Rs) tan(Rs) X Y E N ET NT
1000 1000 30 40 40 100 130 0.50 0.87 0.77 0.84 -1.19 0.84 119.2 -130.5 37.94 -172.64 1038 827

This spreadsheet tests the calculation of drill target location of dilational ore shoots in fault-fill veins. Input parameters (orange) are: s=vein strike(RHR*), d=vein dip, Rs=slickenline rake(RHR*),
and v=vertical depth to target. The UTM coordinates of the best vein-outcrop geochemical anomaly (Eo, No) (light blue) are input (here assumed 1000, 1000), and the UTM coordinates of the
drill target (ET, NT)(dark blue) are output. The test plot shows vein as a strike-dip symbol, the ore shoot as a bold red line, and the slip vector as a bold dashed line. The test plot is a map
view, so the angle between the bold lines will not always appear to be 90°.

test plot parameters calculated for test plot.

1200
S= 30 d= 40 Rs= 40 v= 100 e+ n+ e- n- ed nd es ns
1050.0 1086.6 950.0 913.4 1043.3 975.0 1094.0 1034.2

S=10 d= 60 Rs= 30 v=100


1100 1200

1100 (e+, n+)


1000
(ET, N T)
(Eo , N o)
1000
(e d, n d)
900
(es, ns)
900 (e-, n-)

800
800 900 1000 1100 1200 800
800 900 1000 1100 1200
equations
90+Rs =90+E2
sin(s) =SIN(RADIANS(C2))
cos(s) =COS(RADIANS(C2)) Notes:
cos(d) =COS(RADIANS(D2))
dip cannot = 0
tan(d) =TAN(RADIANS(D2))
tan(90+Rs) =TAN(RADIANS(G2)) Rs cannot = 90
tan(Rs) =TAN(RADIANS(E2))
X =F2/K2 *RHR = right-hand-rule
Y =F2/(K2*L2*J2) RHR strike = azimuth (0-360°) 90° counterclockwise from dip direction.
E =((I2*F2)/K2)+((H2*F2)/(K2*L2*J2))
RHR rake = rake angle (0-180°) measured clockwise from RHR strike.
N =((-H2*F2)/K2)+((I2*F2)/(K2*L2*J2))
ET =A2+P2
NT =B2+Q2 Eric Nelson, Colorado School of Mines; enelson@mines.edu

65

Anda mungkin juga menyukai