You are on page 1of 21

55

Cave Mining--
The State of the Art
D.H. Laubscher*
INTRODUCTION
The expression "cave mining" will be used here to refer to all
mining operations in which the ore body caves naturally after
undercutting and the caved material is recovered through
drawpoints. The term encompasses block caving, panel caving,
inclined-drawpoint caving, and front caving. Caving is the
lowest-cost method of underground mining, provided that
drawpoint size and handling facilities are tailored to suit the
caved material and that the extraction horizon can be
maintained for the life of the draw. 1
The daily production from cave-mining operations throughout
the world is approximately 370 kt/d. Table 55.1 shows
production broken down by layout.
By way of comparison, the South African gold mines produce
350 kt/d.
Today, several open-pit mines currently producing in excess of
50 kt/d are examining the feasibility of implementing low-cost,
large-scale underground mining methods. Several underground
cave mines that produce high tonnages are planning to
implement dropdowns of 200 m or more. This will result in a
considerable change in their mining environments. These
changes will necessitate more detailed mine planning, rather
than the simple projection of current mining methods to greater
depths.
depths.
As more attention is directed to mining large, competent ore
bodies with low-cost underground methods, it is necessary to
define the role of cave mining. In the past, caving has been
considered for rock masses that cave and fragment readily. The
ability to better assess the cavability and fragmentation of ore
bodies, the availability of robust load-haul-dump (LHD)
machines, an understanding of the draw control process,
suitable equipment for secondary drilling and blasting, and
reliable cost data have shown that competent ore bodies with
coarse fragmentationl can be mined at a much lower cost using
caving rather than with drill-and-blast methods. However, once
a cave layout has been developed, there is little scope for
change.
Aspects that have to be addressed are cavability, fragmentation,
draw patterns for different types of ore, drawpoint or drawzone
spacing, layout design, undercutting sequence, and support
design.
Table 55.2 shows that there are significant anomalies in the
quoted performance of different cave operations.
It is common to find that old established mines that have
developed standards during the course of successful mining in
the upper levels of an ore body are resistant to change and do
not adjust to the ground control problems that occur as mining
proceeds to greater depths or as rock types change. Mines that
have experienced continuous problems are more amenable to
adopting new techniques to cope with a changing mining
situation. Detailed knowledge of local and regional structural
geology, the use of an accepted rock mass classification to
characterize the rock mass, and knowledge of regional and
induced stress environments are prerequisites for good mine
planning. It is encouraging to note that these aspects are
receiving more and more attention.


The Laubscher rock mass classification system provides both in
situ rock mass ratings (IRMR) and rock mass strength. Such a
classification is necessary for design purposes. The IRMR defines
the geological environment, and the adjusted or mining rock
mass ratings (MRMR) consider the effects of the mining
operation on the rock mass. Figure 55.1 is a flow sheet of the
MRMR procedure with recent modifications. The reader is
encouraged to read the paper "The IRMR/MRMR Rock Mass
Classification System for Jointed Rock Masses" by Laubscher and
Jakubee, which is included in this volume. The ratings describe
in detail cavability, subsidence angles, failure zones,
fragmentation, undercut-face shape, cave-front orientation,
undercutting sequence, overall mining sequence, and support
design.

FACTORS AFFECTING CAVING OPERATIONS
The 25 parameters that should be considered before
implementing any cave mining operation are set out in Table
55.3. The parameters in capital letters are a function of the
parameters that follow in the same box. Many parameters are
uniquely defined by the ore body and the mining system and are
not discussed further. The parameters considered later are
common to all cave-mining systems and need to be addressed if
any form of cave mining is contemplated.
Cavability
Monitoring a large number of caving operations has shown that
two types of caving can occur--stress and subsidence caving.
However, it is better to use the terms "vertical extension" to
mean upward propagation of the cave and "lateral extension" to
mean the propagation of the cave as a caved block is expanded.
Vertical extension caving occurs in virgin cave blocks when the
stresses in the cave back exceed the strength of the rock mass.
Caving can stop when a stable arch develops in the cave back.
The undercut must be increased in size or the boundaries
weakened to induce further caving. High horizontal stresses
acting on steep dipping joints increase the MRMR. This was the
situation in block 16 at the Shabanie Mine. A stable back was
formed when the undercut had a hydraulic radius of 28 and the
MRMR was 64. When block 7 (adjacent to block 16) caved, the
horizontal confining stress was removed, which resulted in a
reduction in the MRMR to 56 in block 16. At this point, caving
occurred.
Lateral extension caving occurs when lateral restraints on the
block being caved are removed by mining adjacent to the block.
This often results in a large stress difference, leading to rapid
propagation of the cave and limited bulking.
Figure 55.2, based on a worldwide experience base, illustrates
caving and stable situations in terms of the hydraulic radius
(area divided by perimeter) for a range of MRMR values. An
additional curve has been added to account for the stability that
occurs with equidimensional shapes.
All rock masses will cave. The manner of their caving and the
fragmentation need to be predicted if cave mining is to be
implemented successfully. The rate of caving can be slowed by
control of the draw since the cave can propagate only if there is
control of the draw since the cave can propagate only if there is
space into which the rock can move. The rate of caving can be
increased by a more rapid advance of the undercut, but
problems can arise if an air gap forms over a large area. In this
situation, the intersection of major structures, heavy blasting,
and the influx of water can result in damaging airblasts. Rapid,
uncontrolled caving can result in an early influx of waste.
In conventional layouts, the rate of undercutting (RU) should be
controlled so that the rate of caving (RC) is faster than the rate
of damage (RD) due to abutment stresses. Thus, RC > RU > RD.
However, in areas of high stress, the rate of caving must be
controlled to maintain an acceptable level/amount of seismic
activity in the cave back. Otherwise, rock bursts can occur in
suitably stressed areas (pillars and rock contacts). As advance
undercutting will be used in these situations, damage to the
undercut and production levels will not be a problem.
The stresses in the cave back can be modified to some extent by
the shape of the cave front. Numerical modeling can be a useful
tool for helping an engineer determine the stress patterns
associated with possible mining sequences. An undercut face
that is concave towards the caved area provides better control
of major structures. In ore bodies having a range of MRMRs, the
onset of continuous caving will be in the lower-rated zones if
they are continuous in plan and section. This effect is illustrated
in Figure 55.3B, where the class 5 and 4B zones are shown to be
continuous. In Figure 55.3A, the pods of class 2 rock are
sufficiently large to influence caving, and cavability should be
based on the rating of these pods. Good geotechnical
information, as well as information from monitoring the rate of
caving and rock mass damage, is needed to fine-tune this
relationship.
Particle Size Distribution
In caving operations, the degree of fragmentation has a bearing
on
Drawpoint spacing
Dilution entry into the draw column
Draw control
Draw productivity
Secondary blasting and breaking costs
Secondary blasting damage
The input data needed to calculate the primary fragmentation
and the factors that determine the secondary fragmentation as a
function of the caving operation are shown in Figure 55.4.
Primary fragmentation can be defined as the size distribution of
the particles that separate from the cave back and enter the
draw column. The primary fragmentation generated by
subsidence caving is generally more coarse than that resulting
from stress caving. The reasons for this are (1) the more rapid
propagation of the cave, (2) the rock mass disintegrates
primarily along favorably oriented joint sets, and (3) there is
little shearing of intact rock. The orientation of the cave front or
back with respect to the joint sets and the direction of principal
stress can have a significant effect on the primary
fragmentation.
Secondary fragmentation is the reduction in the size of the
original particles as they move through the draw column. The
processes to which particles are subjected determine the size
distribution which reports to the drawpoint. A strong, well-
jointed material can result in a stable particle shape at a low
draw height. Figure 55.5 shows the decrease in particle size for
different draw heights and coarse (less jointed) to fine (well
jointed) rock masses. A range of RMRs will result in a wider
range of particle sizes than that produced by rock with a single
rating. This is due to the fact that the fine material tends to
cushion larger blocks and prevents further attrition of these
blocks. This difference is illustrated in Figure 55.3B, in which
class 5 and class 4 material is shown to cushion the larger
primary fragments from class 3 material. A slow rate of draw
results in a higher probability of time-dependent failure as the
caving stresses have more time to work on particles in the draw
column.






Fragmentation is the major factor determining drawpoint
productivity. Experience has shown that 2 m3 is the largest
block that can be moved by a 6-yd3 LHD and still allow an
acceptable rate of production to be maintained. In Figure 55.6,
the productivity of a layout using 3.5-, 6-, and 8-yd3 LHDs
loading to a grizzly is related to the percentage of fragments
larger than 2 m3. The use of secondary explosives is based on
the amount of oversized rock that cannot be handled by a 6-yd3
LHD.
A computer simulation program has been developed for
calculating primary and secondary fragmentation. The results
obtained from this program are confirmed by underground
observations.
Drawzone Spacing
Drawpoint spacings for grizzly and slusher layouts reflect the
spacing of the drawzones. However, in the case of LHD layouts
with a nominal drawpoint spacing of 15 m, drawzone spacing
can vary from 18 to 24 m across the major apex (pillar),
depending on the length of the drawbell. This situation occurs
when the length of the drawpoint crosscut is increased to ensure
when the length of the drawpoint crosscut is increased to ensure
that an LHD is straight before it loads. In this case, the major
consideration (optimum ore recovery) is compromised by
incorrect use of equipment or a desire to achieve ideal loading
conditions. The drawzone spacings for 30 m center-to-center
production drift spacings are shown in Figure 55.7, with the
critical distance being across the major apex. It can be seen that
the length of the drawbell can be increased from 8 to 10 m and
the production drift spacing to 32 m without affecting the major
apex spacing. There still will be interaction in the drawbell.

Sand model tests have shown that there is a relationship
between the spacing of drawpoints and the interaction of
drawzones. Widely spaced drawpoints develop isolated
drawzones with diameters defined by the fragmentation. When
drawpoints are spaced at 1.5 times the diameter of the isolated
drawzone (IDZ), interaction occurs. Interaction also improves as
drawpoint spacing is decreased, as shown in Figure 55.8. The
flow lines and the stresses that develop around a drawzone are
shown in Figure 55.9. The sand model results have been
confirmed by observation of the fine material extracted during
cave mining and by the behavior of material in bins. The
question is whether this theory, which is based on isolated
drawzones, can be wholly applied to coarse material, where
arching (spans) of 20 m have been observed. The collapse of
large arches will affect the large area overlying the drawpoint,
as shown in Figure 55.10. The formation and failure of arches
lead to wide drawzones in coarse material, so that drawzone
spacing can be increased to the spacings shown in Figure 55.10.



The frictional properties of the caved material must also be
recognized. Low-friction material can flow greater distances
when under high overburden load, and this can mean wider
drawzone spacings. It is therefore logical to expect that when a
line of drawbells is drawn this creates a large low-density zone.
line of drawbells is drawn this creates a large low-density zone.
Similarly, when lines of drawbells are drawn on alternate shifts
good interaction will result. The draw program at the Henderson
Mine broadly follows this pattern.
There is a need to continue with three-dimensional model tests
to establish some poorly defined principles, such as the
interaction across major apexes when the spacing of groups of
interactive drawzones is increased. Numerical modeling could
possibly provide the solutions for the draw behavior of coarsely
fragmented material.
Draw Control
Draw control requirements are shown in Figure 55.11 and Table
55.4. The grade and fragmentation in the dilution zone must be
known if sound draw control is to be practiced. Figure 55.12
shows the value distribution for columns A and B. Both columns
have the same average grade of 1.4%, but the value
distribution is different. The high grade at the top of column B
means that a larger tonnage of waste can be tolerated before
the shutoff grade is reached.



In LHD layouts, a major factor in poor draw control is the
drawing of fine material at the expense of coarse material. Strict
draw control discipline is required so that the coarse ore is
drilled and blasted at the end of the shift in which it reported in
the drawpoint.
It has been established that the draw will angle towards less
dense areas. This principle can be used to move the material
overlying the major apex by creating zones of varying density
through the differential draw of lines of drawbells.
Dilution
The percentage of dilution is defined as the percentage of the
ore column that has been drawn before the waste material
appears in the drawpoint. It is a function of the amount of
mixing that occurs in the draw columns. Mixing is a function of--
Ore draw height
Range in fragmentation of both ore and waste
Drawzone spacing
Range in tonnages drawn from drawpoints
The range in particle size distribution and the minimum
drawzone spacing across the major apex will give the height of
the interaction zone (HIZ). This is illustrated in Figure 55.13.
There is a volume increase as the cave propagates so that a
certain amount of material must be drawn before the cave
reaches the dilution zone. The volume increase, or swell factors,
is based on the fragmentation and is applied to column height.
Typical swell factors for fine, medium, and coarse fragmentation
are, respectively, 1.16, 1.12, and 1.08.

A draw control factor is based on the variation in tonnages from
working drawpoints (Figure 55.14). If production data are not
available, the draw control engineer must predict a likely draw
pattern. A formula based on the above factors has been
developed to determine the dilution entry percentage.

Dilution entry = (A - B)/A &##8734; C &##8734; 100,
where A = draw-column height H swell factor
B = height of interaction
and C = draw control factor.
The graph for dilution entry was originally drawn as a straight
line, but underground observations show that, where the
dilution occurs early, the rate of influx follows a curved line with
a long ore "tail" (Figure 55.15). Figure 55.16 shows that dilution
entry is also affected by the attitude of the drawzone, which can
angle towards higher overburden loads.


Layouts
Eight different horizontal LHD layouts and two inclined drawpoint
LHD layouts are used at various operating mines. An example of
an inclined LHD layout is shown in Figure 55.17. The El Teniente
layout is shown in Figure 55.18 and the Henderson Mine layout
in Figure 55.19. Numerical modeling of different layouts showed
that the Teniente layout was the strongest and had several
practical advantages. For example, the drawpoint and drawbell
are on a straight line, which results in better drawpoint support
and flow of ore. If there is brow wear, the LHD can back into the
opposite drawpoint. The only disadvantage is that the layout
does not have the flexibility to accommodate the use of electric
LHDs such as does the herringbone.



A factor that needs to be resolved is the correct shape of the
major apex. It is thought that a shaped pillar will assist in the
recovery of fine ore. Also, in ore bodies characterized by coarse
fragmentation, there is less chance that stacking will occur
(Figure 55.20). The main area of brow wear is immediately
above the drawpoint. If the vertical height of the pillar above the
brow is small, failure of the top section will reduce the strength
of the lower section and result in aggravated brow wear.

More thought must be given to the design of LHD layouts to
provide a maximum amount of maneuver room within a
minimum size of drift opening. Thus larger machines can be
used within the optimum drawzone spacings. Another aspect
that needs attention is LHD design, that is, to reduce the length
and increase their width. While the use of large machines might
be an attraction, it is recommended that caution be exercised
and that a decision on machine size be based on a correct
assessment of the required drawzone spacing in terms of
fragmentation. The loss of revenue that can result from dilution
far exceeds the lower operating costs associated with larger
machines.
Undercutting
Undercutting is one of the most important aspects of cave
mining since not only is a complete undercut necessary to
induce caving, but the undercut method can reduce the
damaging effects of induced stresses.
The normal undercutting sequence is to develop the drawbell
and then to break the undercut into the drawbell, as shown in
Figure 55.18. In environments of high stress, the pillars and
brows are damaged by the advancing abutment stresses. The
Henderson Mine technique of developing the drawbell with long
holes from the undercut level reduces the time interval and
extent of damage associated with postundercutting. To preserve
stability, the Henderson Mine has also found it necessary to
delay the development of the drawbell drift until the bell must
be blasted (Figure 55.19).
The damage caused to pillars around drifts and drawbells by
abutment stresses is significant and is the major factor in brow
wear and excavation collapse. Rock bursts also occur in these
areas. The solution is to complete the undercut before
development of the drawpoints and drawbells. The "advanced
undercut" technique is shown in Figure 55.21.
undercut" technique is shown in Figure 55.21.

In the past, it was considered that the height of the undercut
had a significant influence on caving and, possibly, the flow of
ore. The asbestos mines in Zimbabwe had undercuts of 30 m
with no resulting improvement in caving or fragmentation. The
long time involved in completing the undercut often led to
ground control problems. Good results are obtained with
undercuts of minimum height, provided that complete
undercutting is achieved. Where gravity is needed for the flow of
blasted undercut ore, the undercut height needs to be only half
the width of the major apex. This results in an angle of repose of
45&##176; and allows the ore to flow freely.
Support Requirements
In areas of high stress, weak rock will deform plastically and
strong rock will exhibit brittle, often violent, failure. If there is a
large difference between the RMR and MRMR values, yielding
support systems are required. This is explained in Figure 55.22.

Prestressed cables have little application in underground
situations unless it is to stabilize fractured rock in a low-stress
environment. The need to constrain the rock laterally and for
lining surfaces such as concrete cannot be too highly
emphasized. Support techniques are illustrated in Table 55.5.
The use of cable bolts in brows is common practice, but often
these bolts are not installed in a pattern that takes into account
joint spacing and orientation. Cable bolts should not be installed
in highly jointed ground, as the bolts do not apply lateral
restraint to the brow and serve only hold blocks in place.

CONCLUSIONS
Cavability can be assessed provided accurate geotechnical data
are available and geological variations are recognized. The
MRMR system provides the necessary data for an empirical
definition of the undercut dimension in terms of the hydraulic
radius.
Numerical modeling can assist an engineer in understanding and
defining the stress environment.
Fragmentation is a major factor in assessing the feasibility of
cave mining in large, competent ore bodies. Programs are being
developed for predicting fragmentation, and even the less-
sophisticated programs provide good design data. The economic
viability of caving in competent ore bodies is determined by LHD
productivity and the cost of breaking large fragments.
Drawpoint and drawzone spacings for coarser material need to
be examined in terms of recovery and improved mining
environments. Spacings must not be increased to lower
operating costs at the expense of ore recovery.
The interactive theory of draw and the diameter of an isolated
drawzone can be used in the design of drawzone spacings.
Complications occur when drawzone spacings are designed on
the basis of primary fragmentation and the secondary
fragmentation is significantly different.
ACKNOWLEDGMENTS
This paper presents an update of the technology of cave mining.
It is not possible to quote references since the bulk of the data
supporting the contents of this paper have not been published,
and the basic concepts are known to mining engineers.
However, it is appropriate to acknowledge the contributions
from discussions with the following people in Canada, Chile,
South Africa, and Zimbabwe: R. Alvarez, P.J. Bartlett, N.J.W.
Bell, T. Carew, A.R. Guest, C. Page, D. Stacey, and A. Susaeta.
The simulation program for the calculation of primary and
secondary fragmentation was written by G.S. Esterhuizen at
Pretoria University.
Pretoria University.

1. Bushman's River Mouth, South Africa.

You might also like