Anda di halaman 1dari 17

A REVIEW OF CATASTROPHIC FLOW FAILURES OF DEPOSITS OF MINE

WASTE AND MUNICIPAL REFUSE


G.E. BLIGHT & A.B. FOURIE
University of the Witwatersrand, Johannesburg, South Africa

ABSTRACT: Catastrophic flow failures occur in mine tailings dams and dumps of discards and other mine waste with alarming
frequency. In recent years catastrophic flow failures have also occurred in dumps of municipal refuse and even in what were
considered to be carefully controlled and well engineered landfills. Apart from the environmental devastation caused by these flows,
they are also dangerous to human life and society. For examples the Buffalo Creek disaster in the USA in 1972 killed 118 people,
made 4000 homeless and destroyed 50 million US dollars worth of property and facilities, the flow slide that occurred in the
Umraniye-Hekimbasi refuse dump in Turkey in 1993, killed 39 people, destroying their homes in the process. The paper will briefly
review some of the more typical flow slides in waste materials, analysing the mechanics of failure and pointing to ways of preventing
this type of failure by a combination of sound design and operating procedures. In the case of existing deposits modified operating
procedures can be adopted, reducing the probability of failure as well as constructing deflecting structures to protect communities and
facilities from the consequences of failure.

Keywords: Flow failure, mine waste, municipal solid waste.

1 INTRODUCTION: STATISTICS, FAILURES, BREACHES,


FLOW FAILURES, EXAMPLES
1.1 Flow failures in tailings impoundments
Tailings dams, whether of the valley or ring impoundment type,
usually consist of an outer impounding dyke or dam wall that
serves to retain the body of tailings, supernatant water (and,
occasionally, storm precipitation) upstream of it. If, for any
reason, this outer impoundment is breached there will be the
danger that the impounded tailings will escape the impoundment.
Figure 1 shows statistics for 184 incidents involving tailings
dams, collected by the US National Committee on Large Dams
(1994). Here, the definition of "failure" is "any breach in the
embankment leading to a release of the impounded tailings". The
184 incidents were not all failures, nor were the failures all flow
failures. The statistics, however, show that, of the known causes
of failures or breaches, the most likely to occur are slope
instability, earthquakes and overtopping, in that order.
Foundation and seepage failures come next, followed by
structural failures. The meanings of the latter three categories are
not very clear, but they presumably include foundation shear or
piping failure, piping through the dam wall and inward collapse
of a decant tower or decant outfall, any or all of which could
result in breaching of an impoundment.
Slope instability may result in a distortion or flattening of the
slope of the retaining impoundment, without any escape of
impounded tailings, or the tailings may move only a short
distance beyond the impoundment wall. An example of a dyke
failure in which no tailings escaped is illustrated in Figure 2,
which shows a rotational shear failure in the outer dyke of a
platinum tailings dam at Bafokeng, South Africa (Blight 1997)
(7, Table 1). Here, the dyke stabilized at a flatter average slope,
as a result of the failure, and none of the retained tailings escaped
the impoundment. However, a year later the same dyke was
breached catastrophically as a result of overtopping, releasing
3x106m3 of tailings (as shown by Figure 3a) and causing 13
deaths as a result (Jennings 1979, Blight 2000).

Figure 1: Analysis of causes of tailings dam failure


(USCOLD, 1944).

Hence the failure of a tailings dam dyke does not inevitably


result in an escape of tailings but a dyke breach must obviously
occur before tailings can escape. At Bafokeng, overtopping
probably resulted in breaching of the dyke by erosion (see Figure
3b for mechanism), but 20 years later, the disastrous flow failure
that occurred from the Merriespruit tailings dam (also in South
Africa), Figure 4 (13, Table 1), resulted from breaching of the
dyke by overtopping, causing slope erosion and progressive shear
failure, as shown in Figure 5 (Wagener et al 1997, Blight 2000).
Figures 3a and 4 illustrate the characteristics of a tailings dam
flow failure. These include liquefaction of a large volume of the
retained tailings, which flow out of the breach as a viscous liquid
and are capable of moving large distances before coming to rest.
The 3 x 106m3 tailings escape at Bafokeng travelled 42km,
covering its path with slurry, before the remaining 2 x 106m3 was
stopped when the flow entered a water retaining dam. The flow

The flow failure of a fly ash dump that occurred in 1961 in


Jupille, Belgium (Bishop 1973) (2, Table 1) is another archetype
of a flow failure in "dry" material. Figure 7a shows a plan of the
course of the flow which travelled down a dry valley for 0.5km.
At Jupille, it appears that the ash may have been fluidized by air
contained by its pores when the fly ash contracted during the
failure. It was reported that fly ash that entered houses,
overwhelmed in the flow, appeared to be "dry". Figure 7b shows
that as the ash flowed down a natural valley it "lined" the valley
with ash, the stream of fluid ash eventually flowing in a "canal"
of solid ash. Of course, this also happens to some extent with
flows of wet materials: the course of the flow is marked by
material stranded as the main flow passes. During the 42km long
flow at Bafokeng an estimated 1 x 106m3 of the 3 x 106m3 of
tailings that escaped was left marking the course of the flow.

Figure 2: Failure (or slump) of retaining dyke of Bafokeng


tailings dam (South Africa) that did not result in escape of
tailings (1973).

1.3 Flow failures of municipal solid waste


Until recently, flow failures in dumps or landfills of municipal
solid waste have been unknown. This may be because significant

Figure 3a) Plan of Bafokeng tailings dam showing position of


breach, course of flow failure, and extent of pools prior to failure
(1974). b) Stages of failure of Bafokeng tailings dam.
at Merriespruit (a lesser quantity of 600 x 103m3) travelled 2km
before being halted and contained by an ornamental lake.

1.2 Flow failures of "dry" mine and industrial waste


Tailings are hydraulically deposited as slurries, into containments
designed to retain the consolidating solids and supernatant and
storm water. However, numbers of flow failures have also
occurred in mechanically placed "dry" mine waste deposits. The
prime example of a flow failure in a "dry" mine waste occurred
at the village of Aberfan, Wales (4, Table 1). Here, in 1966, a
dump or tip of coal waste failed, liquefied (largely as a result of
dumping waste over a spring) and flowed into the village of
Aberfan, killing 144 people of whom 116 were school children
(Anonymous 1967, Bishop, 1973). Figure 6 shows the course of
the 1966 flow slide. The figure also shows that the Aberfan tip
had failed and flowed twice previously, in 1944 and 1963, but
these earlier flows did not reach the village and did not serve as
sufficient warning to the owners of the tip, or regulatory officials,
of the eventual 1966 disaster.

Figure 4: Plan of Merriespruit dam, South Africa, showing


position of pool at time of failure, intended position of pool,
breach in dyke, and path of tailings flood (1994).

Figure 5: Most likely development of flow failure at


Merriespruit, 1994.

Figure 7: Flow of fly ash at Jupille, Belgium.

surface, along the valley. The sewage that poured from the sewer
pipe was dammed by the slide debris and formed a lake of
sewage on the upstream side of the obstruction. Since this
occurrence, two and possibly three more flow failures of
municipal solid waste deposits have been reported (Hendron et al
1999 (17, Table 1), Brink et al 1999 (18, Table 1)).

1.4 Record of notable flow failures


Table 1 records 22 failures of "dry" mine waste deposits,
hydraulic fill tailings impoundments and municipal solid waste
landfills that occurred over the 72 years from 1928 to 2000. The
table gives an idea of how widespread these failures can be,
geographically and in terms of the materials that have flowed, the
volumes of material involved and the consequences. The death
statistics at the foot of the table (1400 deaths in 72 years) also
show that flow slides are not particularly dangerous occurrences.
A single flying accident can cause 400 deaths, and we expect to
have at least one or two of these per year, yet commercial air
travel is not considered dangerous, nor is travel by road, even
though the annual road death toll is hundreds of thousands.
Not all flow failures of waste deposits make headlines, and
regrettably as mentioned above, some failures may never reach
the news and are never recorded in the geotechnical literature.
Figure 9 (Blight 2000) (12, Table 1) shows three flow failures
that occurred in a tailings impoundment at Saaiplaas, South
Africa, in three days. Because the failures caused no deaths or

Figure 6: Flow failures of coal waste at Aberfan, Wales, 1944,


1963, 1966.
numbers of landfills have not, until recently, reached a large size
or because failures that did occur caused no deaths and therefore
were not newsworthy and were not reported. However, in 1977 a
flow failure took place in a landfill at Sarajevo (Gandolla et al
1979), and in 1993 a massive flow failure took place in the
Umraniye-Hekimbasi refuse dump in Istanbul, Turkey (Kocasoy,
Curi 1995) (11, Table 1). Figure 8 shows sections through the
dump before and after the failure, as well as the course of the
debris flow. After reaching the bottom of the valley, the
momentum of the flow carried it up the opposite slope,
destroying a number of informal houses and killing 39 people.
The slide also fractured a main sewer pipeline that ran, on the

Figure 8: Failure of Umraniye-Hekimbasi municipal solid waste


dump, Istanbul, Turkey.

injuries and the flows were confined to mine property, this


incident was never reported by the news media, and never
comprehensively investigated. The Saaiplaas impoundment is
only a few km from the Merriespruit impoundment that failed a
year later. If the Saaiplaas failure had been publicized, it may
have served as a warning to the operators of other tailings dams
in the area to inspect their dams carefully for safety, and the
Merriespruit disaster might have been avoided. However, most of
us are confident that we are immune from disasters that befall
others, so it is more likely that the warning would have been
ignored, as in the case of Aberfan.

2 STRAIN-SOFTENING OR LIQUEFACTION OF MINE


AND MUNICIPAL WASTES
2.1 The mechanics of strain-softening or liquefaction
When a particulate material, be it a soil, tailings, dry mine waste
or municipal solid waste, is subjected to shear stresses, it will
tend to change volume and hence void ratio. Dense materials will
tend to dilate, loose materials will tend to contract and materials
of intermediate, or near critical state density will have little
tendency to change volume. The consequences of this behaviour
when a saturated material is sheared undrained, are illustrated by
Figure 10 (after Castro 1969). The important features of Figure
10, which shows results for consolidated undrained straincontrolled shear, are:

Figure 9: Plan of Saaiplaas dam, (South Africa) showing


locations of failures A, B, & C.

of the same order, and occurred at approximately the same


(small) axial strain.
.2 After the initial peak, the loose specimen lost strength (or
strain-softened) as the strain increased, whereas the dense
specimen continued to gain strength (strain-hardened) with

.1 The initial peak shear strength achieved by all specimens was

Deviator
Stress
(kPa)

Deviator
Stress
(kPa)

Pore
Pressure
u
(kPa)

Pore
Pressure
u
(kPa)

Figure 10: The effect of initial relative density (Dr) on the shape
of the stress-strain curves of consolidated undrained tests on
saturated sand. (Castro, 1969)
increasing strain. The strength of the intermediate specimen
remained more or less constant.
.3 In the loose specimen, the pore pressure rapidly rose with
increasing strain, to reach a constant maximum. In the dense
specimen, after rising to a peak, the pore pressure reduced
continuously and the shear strength increased.

Figure 11 :
a) Stress-strain curves for stress-controlled
anisotropically consolidated undrained tests on saturated loose
sand ( s 31c = 400kPa)
b) Corresponding stress paths (Castro, 1969).

The loss of shear strength of the loose specimen from 180kPa


to 20kPa after a strain of 3 to 5% constitutes strain-softening
which in an extreme state can be called liquefaction. Because the
shear stress was applied monotonically, this is termed static
strain-softening or liquefaction.
Figure 11 a shows the results of three load-controlled tests,
also by Castro (1969), on specimens of loose saturated sand.
These show very similar behaviour to the strain-controlled test of
Figure 10. These tests, though, also show that if a loose material
fails under stress-controlled conditions, which is usually the case
in a slope failure, the failure can occur very rapidly, in fact,
almost instantaneously as the shear stress is applied. Figure 11b
shows the stress paths for the tests of Figure 11a, illustrating that
the ultimate effective stress state reached in these tests lies on the
Kf or failure line for strain-controlled tests.
The behaviour of a loose saturated sand silt under dynamically
applied shear stress is illustrated by Figure 12 (Blight 1990).
Each application of the shear stress of 300kPa caused an
increment of pore pressure that reduced the mean effective stress,
until application 13 moved the stress path onto the Kf or failure
line. On stress application 14 it was not possible to reach the
shear stress of 300kPa. If the test had been continued past stress
application 14, the ultimate condition would have been reached,
with the stress path on the Kf-line, a constant mean effective
stress, and a very low shear strength.

Figure 12: Stress path for dynamic shear test on loose saturated
natural sandy silt.

weather, varying thickness of layers of deposition, demands for


increased deposition rates at times of increased production to
meet market demands, etc., the degree of densification inevitably
varies both with time and position on the tailings dam (both in
elevation and plan). For example, for a year before the three
failures on the Saaiplaas No. 5A dam (Figure 9) took place, the
rate of rise of the dam had been increased from its long-term
average of 1.8m/y to 2.6m/y and shortly before the failures
occurred, the rate of rise had been further increased to 2.8m/y.
Each increase in rate of rise would have reduced the time
between tailings deposition cycles. Thus the density of
successive layers of tailings may be, and usually is highly
variable. The effects of this variable density on the measured
shear strength of tailings are illustrated by Figures 14 and 15
(Blight 1997, Fourie et al 2001). Figure 14a shows stress paths

2.2 Strain-softening or liquefaction of tailings


Tailings are usually deposited as slurries and settle out as the
tailings are beached in the impoundment at a high water content.
They therefore settle on the beach with a loose particle structure.
If the tailings beach is allowed to dry out between successive
deposition cycles, which is usually the case, the slurry layer
shrinks and densifies as it dries, as shown in Figure 13. However,
because of seasonal and other variations in the

Figure 15a shows the results of a piezo-cone penetrometer test


conducted on the Merriespruit impoundment after the failure
illustrated by Figures 4 and 5. The cone penetration resistance
fluctuated over a range of up to 2MPa as the cone penetrated
successive layers of tailings. The pore pressure, in sympathy,
showed low or even negative values as dense, dilative layers
were penetrated, and high values as loose contractive layers were
encountered. Figure 15b summarizes the results of 16 piezo-cone
penetrometer tests at Merriespruit, made at various distances
from the toe of the dam. Each cone penetration profile has been
characterized by its maximum and minimum slopes in terms of
penetration resistance per unit depth (in kPa/m). There was a
considerable difference between these two slopes, and both the
slopes and the difference between maxima and minima decreased
with distance from the toe. However, there was no sudden
change in penetration characteristics between tailings forming the
outer slope of the impounding dyke and those contained in the
interior of the impoundment. In other words, this was not a case
of a consolidated outer embankment retaining a partly
consolidated semi-fluid core. Certain layers of tailings forming
the beach of the dam must have suffered static strain-softening or
liquefaction for the flow failure to have occurred.

Figure 13: Densification of tailings slurry by drying shrinkage

2.3 Strain-softening or liquefaction of "dry" mine wastes


Bishop (1973) drew attention to the phenomenon of the
"bulking" of unsaturated sands and gravels when deposited
without compaction, a phenomenon long known in concrete
technology with relation to volume batching of aggregate. In
general terms, if a given mass of dry cohesionless sand or gravel
is deposited loosely, it will assume a certain volume and void
ratio. If water is gradually added, the volume of the mass (and
hence its void ratio) will increase up to an optimum water content
after which the volume will decrease again. When the material is
saturated, it will have approximately the same volume and void
ratio as when it is dry. Bulking is well illustrated by the results
shown in Figure 16a for mixtures of the coarse gravel and sandfractions of diamond mining waste. The two sets of curves were
prepared with different compactive efforts, and hence initial void
ratios, but regardless of initial void ratio, showed much the same
maximum increase in void ratio as bulking proceeded. At water
contents approaching saturation, the void ratios were much the
same as the initial values. Note that the sand content of the
material had little effect on the bulking, but the addition of sand
did affect initial void ratios for the same compactive effort. For
these materials, specimens prepared at void ratios of 1.0 or above
were contractive in consolidated undrained triaxial shear. Below
a void ratio of 1.0, the materials were neutral to dilative.
Figure 16b shows bulking results presented by Bishop (1973)
for waste from the Aberfan tip, which show the percent decrease
in volume on saturation. Figure 17 shows results for triaxial
shear tests on bulked colliery waste (Dawson et al 1998) which
was set up at a void ratio of 0.51, consolidated isotropically to
0.40 under an effective stress of 200kPa and sheared undrained
(although it is not clear at what stage the specimen was
saturated). The strain-softening behaviour was very similar to
that shown in Figures 10 and 11.
"Dry" mine wastes are usually deposited in a bulked condition
without compaction. Subsequent saturation by heavy or
continuous rain or some other source of water can cause a
tendency for a sudden decrease in void ratio with its consequent
strain-softening loss of shear strength.
Fortunately, there is a current trend in South Africa for mines
to compact their dry wastes. In the case of colliery wastes, this is
done to reduce the air permeability of the waste and thus prevent
spontaneous combustion, sustained by the entry of oxygen.

Figure 14a: Stress paths for consolidated-undrained triaxial shear


of undisturbed tailings specimens.

Figure 14b: Effective stress changes in undisturbed specimens


during unconsolidated undrained triaxial compression.
for four consolidated undrained triaxial shear tests on 38mm
diameter x 76mm high samples taken from a single Shelby tube
sample from the Merriespruit tailings dam (see Figures 4 and 5).
The two tests at consolidation stresses of 50 and 100kPa showed
contractive behaviour, the one at 200kPa showed critical state
behaviour, while that at 400kPa was weakly dilative. Figure 14b
summarizes changes in effective stress from the start of shearing
to the ultimate state for 16 consolidated undrained shear tests on
Shelby tube specimens from Merriespruit having various void
ratios. The specimens were tested under their original in situ
effective overburden stress. While nine of the specimens dilated
during shear, seven showed contractant or almost neutral
behaviour. In other words certain of the layers in the tailings
could have strain-softened or liquefied and flowed during the
large scale failure, carrying other denser layers with them.

Figure 15a: Typical cone penetration test in Merriespruit tailings


impoundment.

Figure 15b: Variation of shear strength with distance from toe of


outer wall for Merriespruit tailings impoundment.

Figure 16b: Bulking effects in coal waste from Aberfan.

Figure16a: Bulking curves for diamond tailings

Figure 17: Typical isotropically consolidated undrained test for


coal mine waste (rock sandy gravel).

Figure 18: Results of consolidated undrained triaxial shear test on


reconstituted specimens of MSW measuring 300mm dia. by
600mm high (Caicedo, etal, 2002).

Some gold mines in South Africa sluice their coarse wastes with
waste mine water as a means of disposing of waste water. The
sluicing causes the rock to compact, reducing its tendency to
contract, but may unfortunately increase acid seepage from the
base of the dumps, leading to undesirable surface and ground
water pollution.

contraction ceasing. Similar results were obtained in drained


triaxial tests on reconstituted MSW specimens from the Bulbul
landfill in Durban, South Africa (18, Table 1).
Hence at present there appears to be no clear evidence from
laboratory tests that MSW can be strain-softening. However,
there is no doubt from the three (possibly four) flow failures in
MSW landfills recorded (11, 17, 18 and possibly 22, Table 1)
that MSW can strain-soften, resulting in flow failure.

2.4 Strain-softening or liquefaction of municipal solid waste


Largely because the phenomenon of flow failures in municipal
waste landfills has only recently become an obvious problem,
relatively little is known of the strain-softening behaviour of
municipal solid waste (MSW). MSW is particularly difficult to
characterise because of its heterogeneity and fibrous texture
which makes it almost impossible to sample in an undisturbed
condition. Also, the properties of MSW change with age and the
progress of decomposition. Although there were some published
data on strength parameters (e.g. Singh and Murphy 1990) it is
only recently that data have been published on volume and pore
pressure changes during shear (Vilar and Carvalho 2002, Caicedo
et al 2002). In particular, Caicedo et al performed consolidated
undrained triaxial tests on saturated 300mm diameter by 600mm
high reconstituted specimens from the Dona Juana landfill in
Bogota (17, Table 1), obtaining the results shown in Figure 18.
(The density of the specimens is not given.) The pore pressure
behaviour is what would be expected of a high void ratio
material, increasing continuously with strain. But the shear
strength also increased continuously, the net effect being for the
MSW to behave as if dilatant. These tests were terminated at an
axial strain of 13%. However, drained triaxial tests by Vilar and
Carvalho (2002) on MSW from a landfill in Sao Paulo, Brazil
were taken to an axi al strain of 40% without the shear strength
reaching a maximum, or the volume

3 DESCRIPTIONS OF TYPICAL FLOW FAILURES IN


TAILINGS IMPOUNDMENTS, "DRY" MINE WASTE
DUMPS AND MUNICIPAL SOLID WASTE LANDFILLS
3.1 Tailings impoundments
3.1.1 Failure caused by seismic action
Common features of failures in tailings dams caused by seismic
action are (Troncoso, in Blight et al 2000):
.1 the presence of a large pond in the impoundment that has
encroached on the outer impoundment dyke;
.2 an outer dyke formed of loose, poorly compacted or
uncompacted tailings sand that is contractive when subjected
to shear stress;
.3 poor separation of the sand used to build the impoundment
dyke from the silts stored within the impoundment, with
weak lenses of silt included in the dyke; and
.4 dykes usually built (at least partially) by upstream deposition.

When an earthquake of sufficient magnitude occurs, a failure


develops as follows:

It should be noted that after a failure, the flow of liquefied


tailings from the impoundment will continue until a surface
profile compatible with the reduced strength of the tailings has
developed. Once this stable surface has formed, loss of tailings
from the impoundment will cease. In the case of El Cobre
(Antiguo) the average stable slope was about 3.5, under static
conditions because the quaking had stopped. Any aftershocks
could have resulted in further flattening of the profile, and further
loss of tailings.

.1 the shear strains and the corresponding shear stresses imposed


by the earthquake cause the weaker, fine, possibly partly
consolidated tailings in the basin of the impoundment to
strain-soften. If the shear strength falls to a low enough
value,
.2 liquefied tailings and ponded water will move in waves,
alternately drawing down and overtopping the upstream
slope and crest of the confining dyke;
.3 the upstream slope of the dyke may slide into the
impoundment, and the dyke may crack;
.4 when the wave of water and liquid tailings returns, it may
overtop the failed section of the dyke, eroding it and forming
a breach, while water and liquid tailings may flow into and
through cracks in the dyke, eroding and enlarging them;
.5 the downstream slope of the dyke may fail in shear, as a result
of strain-softening accompanied by erosion;
.6 as the breach in the dyke rapidly enlarges, the contents of the
impoundment flow out of the breach starting the tailings
flood, which is sustained by retrogressive liquefaction of the
tailings within the impoundment (as illustrated by Figure 5);
.7 the failure process and flow of tailings cease once the shear
strains imposed by the earthquake diminish and a stable
surface profile is developed by the breached dyke and the
tailings flood that has escaped from the impoundment. This
profile must be sustainable by the reduced shear strength of
the strain-softened tailings.

3.1.2 Flow failure resulting from static liquefaction


For a flow failure to occur as a result of a static liquefaction, the
outer dyke of the tailings impoundment must be breached either
by shear (e.g. Figure 2) possibly followed by overtopping, or by
piping erosion followed by overtopping (e.g. Figure 3b), or by
overtopping followed by erosion and shear failure (e.g. Figure 5).
The formation of a breach in the outer dyke acts as a trigger for
strain-softening or liquefaction of the impounded tailings by
imposing sudden shear strains in the tailings adjacent to the
breach by the removal of lateral support. If certain layers
sandwiched in the mass of tailings are susceptible to liquefaction,
they lose strength and cause the adjacent, possibly dilative layers
(see Figure 15a) to disintegrate as well, with the result that a
substantial part of the total tailings mass moves towards and out
of the breach. This process continues until the stable surface
profile, compatible with the reduced strength of the tailings that
was mentioned above, has developed. Note also, that the basin
that forms the source of the flow must not only be stable on the
line of the breach (the exit direction of the escaping tailings), but
also transversely, i.e. the basin sides must everywhere develop a
stable slope before the tailings flow can cease.
The Merriespruit failure (13, Table 1 and Figure 4) is a good
example of a flow failure that resulted from static liquefaction.
On 22 February 1994 a rainstorm deposited 25mm of water on to
the Merriespruit gold tailings ring-dyke impoundment in the Free
State province of South Africa. A large quantity of water had
been stored in the impoundment, reducing the free-board to an
unknown, but small value. Shortly thereafter, as runoff from
rainfall on the impoundment surface concentrated in the pool, the
dyke was overtopped and breached. A flow failure ensued that

The El Cobre (Antiguo) failure (3(1), Table 1) is a good


example of a failure caused by an earthquake (Dobry, Alvares
1967). Figure 19 shows cross-sections through the side-hill
impoundment before and after failure. The impoundment was
commissioned in 1930, but after the Nuevo (new dam) (3(2),
Table 1) was constructed in 1963, the Antiguo (old) dam was
used only periodically as a standby. The dyke had been built by
upstream hydraulic filling, and the downstream slope of the dyke
was 35m high at the time of the failure. The epicentre of the 7.5
Richter magnitude La Ligua earthquake that resulted in the
failure was 70km from the dam with a focal point at a depth of
61km.

Figure 19: Pre-and post- failure profiles of EI Cobre old dam


involved 600 000m3 of tailings and cut a swathe of destruction
through the village of Merriespruit downhill of the tailings dam.
Seventeen people were killed and scores of houses were
demolished and swept away by the flood. Eventually, the flow
stopped about 2km from the breach when the tailings entered an
ornamental lake, constructed in a natural wetland.
After the afternoon rainstorm, clear water (presumably from
the dam) ran through the streets of the village from about 7 p.m.
to 9 p.m. when failure occurred. The failure was accompanied by
a series of bangs. It was dark, but there was light from the moon.
Unfortunately, eye-witness accounts as to how the failure took
place do not give a consistent picture. The wall appears to have
disintegrated into a series of large slabs that crashed down,

During the quake a cloud of dust arose from the dried surface
of the only periodically used impoundment. The flow failure
continued for 20 minutes after the quake had ended, as 1.9 x
106m3 of a total storage of 4.25 x 106m3 of tailings flowed down
a dry valley for a distance of 12km. A town in the path of the
flow was annihilated with 300 deaths occurring.
As shown by Figure 19, the dam was constructed on sloping
ground with a slope angle of 3 and the average slope of the postfailure profile through the breach was only 3.5. The flow was
reported to have covered its 12km course in a few minutes. This
is too imprecise to allow the speed of the flow to be estimated,
but it must have been about 20kmh-1 (see Section 4).

Figure 20: Sections through failure at Merriespruit showing postfailure equilibrium surface.
causing the noise and being followed by a wave of mud (see
Figure 5).
Figure 20 shows sections of the post-failure equilibrium
surface for the failure basin of the Merriespruit tailings
impoundment. Section E'E' is the pre-failure section normal to
the wall and EE is a section through the breach. Section FF
runsat right angles to EE, and GG runs at 45 to EE. The
intersections of FF and GG with EE are marked in Figure 20. The
slope of the tongue of escaped tailings was 2, which is very
similar to the slope of large portions of FF and GG. In other
words, the post-failure surface had flattened to a general slope of
2-3, with some portions around the perimeter of the failure scar
being as steep as 10-20. Presumably, these areas had formed
late in the failure process, had been subjected to lesser shear
strains because they were not so high, and were therefore stable
at steeper surface slopes.
Because of the disturbance caused by the failure, it is very
difficult to know from what depth in the impoundment the
material that composes the post-failure surface originated. The
surface is also too soft to be accessible after a failure until a
drying crust has formed. Hence it is not possible to sample a
post-failure surface straight after the failure to help identify its
depth of origin. It seems likely, however, that the tailings that
move out of the breach will consist of the upper, more recently
deposited layers, and that the post-failure surface will consist of
deeper layers exposed as the slope of the failure basin is flattened
by the outward flow of the tailings.
For example, Figure 21 shows profiles of vane shear strength
measured in an operating gold tailings impoundment. In the
event of the outer dyke being breached, it is obvious from their
relatively low strength that the top 10m of tailings would tend to
flow off more readily than the deeper layers. Figure 21 also
demonstrates the loss of in situ strength of the tailings when
disturbed, with a sensitivity ratio or strength reduction factor
(undisturbed/remoulded strength) of about 2.7.

Figure 21: Vane shear strength profiles measured in an operating


gold tailings impoundment
and 7 at Aberfan, 7 being the tip that failed and flowed in 1966.
The colliery waste was tipped loosely by a mechanical tipper and
the slopes of the tip were at the angle of repose of the waste of
about 37. Under the toe of tip 7 was a spring, fed by water in the
underlying sandstone under artesian pressure between the
uppermost coal seam and the surface layer of alluvial boulder
clay, which acted as aquicludes. The height of the tip when the
failure occurred was about 67m from toe to crest.
The failure was probably initiated by a series of shallow slips
triggered by the artesian pressure of the spring and exacerbated
by contraction of the loose, bulked waste as it became saturated
by upward seepage from the spring. At 07.30 on the morning of
the failure, the tipping gang found that the crest of tip 7 had
moved downwards by 3m over a distance of 10 to 12m from the
edge. By 08.30 this displacement had increased to 6m. At 09.10
the toe of the tip started moving down the 12 hillside and
within a few minutes the rapid flow down the hillside had
commenced. The flow travelled 1600m before reaching the
school which it destroyed, and came to a halt 350m further
downhill. Referring to Figure 6, at Aberfan road, the depth of the

3.2 Flow failures of "dry" mine waste dumps


Perhaps the best example of a flow slide involving dry mine
waste was the final of the series of three failures that occurred at
Aberfan (4, Table 1). Figure 22 shows a section through tips 5

10

Figure 22: Section through tips 5 and 7 at Aberfan

was burning in several places and streams of noisome leachate


issued from the toe of the dump and ran down the slope into the
valley bottom. In 1992 the "technical advisor" to the Mayor of
Istanbul decided that the waste should be covered, and later that
year the site operator complied by covering the sub-horizontal
top platform with 3 to 5m of demolition wastes and soil. This
additional disturbing force was the straw that broke the camel's
back.
The failure took place in April 1993. Heralded by a loud bang,
which was later ascribed (probably wrongly) to a methane
explosion, 1.2 x 106m3 of waste rapidly moved down the valley
and was carried a short way up the opposite slope, where the
houses were situated that the slide demolished.
Whereas the failure at Istanbul took place as a result of a
complete lack of engineering or technical input or understanding,
the failure of the Dona Juana landfill in Bogota, Columbia (17
Table 1) appears to have occurred as a result of a combination of
poor design understanding and poor appreciation of operating
principles (Hendron et al, 1999, Caicedo et al, 2002.)
.

waste was 9m. The speed of the flow was estimated to have been
15 to 30kmh-1.
3.3 Flow failures in municipal solid waste
The flow failure at Istanbul (11, Table 1) will be taken as the
archetypical example of this type of flow failure (Kocasoy and
Curi 1995). It is remarkable not only for the destruction it
wrought, but also for the lack of common sense of the authorities
that established and operated the landfill. Figure 8 shows that the
landfill must have been sited where it was, purely for reasons of
expediency. Given some flexibility in siting, no engineer in his
right mind would have sited a waste deposit on a 27 slope. The
waste was dumped near the edge of the slope, sorted through by
informal reclaimers (i.e. scavengers), and then pushed over the
edge by dozer where it came to rest at an angle of repose of 45.
There was no attempt to compact the waste and no attempt to
cover it either. As a result, the waste absorbed all the rain that fell
on it, as well as the runoff from the dumping platform. The waste

Figure 23: Progression of failure of Dona Juana landfill (Hendron, etal, 1999).

11

The zone of the landfill that failed (see Figure 23) was lined with
a 1mm PVC geomembrane resting on either compacted clay or in
situ soils. A sand drainage layer and a protective soil layer were
above the liner. A horizontal soil cover layer was provided on top
of each 2.5m lift of compacted waste, while the lifts of waste
between cover layers were interconnected with rock-filled drains
to allow leachate to percolate downwards to the drain above the
liner. There was also a passive gas venting system consisting of
vertical perforated pipes on a 50m grid. A leachate recirculation
system was installed consisting of horizontal perforated pipes
placed on top of each waste lift before placing the cover layer.
The object of this piping was to inject leachate, collected from
the base of the landfill, back into the waste, so as to operate the
landfill as a biological waste reactor, thus purifying the leachate
before releasing it into the nearby river.
The investigation of the failure concluded that it had been
triggered by high liquid pore pressures caused by the re-injection
of leachate. The zone that failed was the only zone where
leachate recirculation had been applied. The design stability
analysis had assumed that no pore pressures would occur in the
waste. The inset on Figure 23 shows how the calculated factor of
safety for the failed section must have declined as the waste
thickness increased during the initial 22 months prior to the start
of leachate injection (Caicedo et al 2002). The additional pore
pressures caused by re-injection caused the already low factor of
safety to fall to 1.0 and the failure followed.
The failure investigation reached the obvious conclusion that
when designing a landfill where leachate is to be re-circulated,
pore pressures must be properly evaluated and their effect must
be considered in the stability analysis

4 RELATIONSHIP BETWEEN GROUND AND POSTFAILURE SURFACE SLOPES AND TAILINGS-GROUND


INTERFACIAL SHEAR STRENGTH
Figure 24a shows the basis for a simple sliding block analysis to
calculate the relationship between the post-failure slope, , of a
tongue of escaped tailings, dry mine waste or municipal solid
waste, the slope, i, of the ground surface and the interfacial shear
strength, t, between the ground surface and the fugitive waste.
Alternatively, the analysis can be used to calculate the shear
strength of the surface of the failure basin within a breached or
failed impoundment, dump or landfill, or the acceleration of a
flow of material once it exits the boundary of the waste deposit
(Blight et al 1981, Blight 1997).
For the potentially sliding block illustrated in Figure 24a,
Downstream forces - upstream forces = mass of block x
acceleration
i.e. (P1 - P2)cosi + Wsini - tL/cosi = W.a/g

(1)

The symbols are defined in Figure 24a and


a = acceleration of the block, g = gravimetric acceleration.
From equation (1)
a = [(P1 - P2)cosi + Wsini - tL/cosi]g/W

(1a)

If the block of material comes to rest, a = 0 and


t = [(P 1 - P2)cosi + Wsini]cosi/L

(1b)

If the block is accelerating, its increase in velocity after time ?t


will be
?v = a?t

(2)

In equation (1)
W = [2h - L(tan - tani)? L/2
and

(3)

h =H1 + L(tan - tani)


(4)
where ? ( is the bulk unit weight of the material in the block.)
If the surface of the flow (i.e. of the block) is parallel to the
ground surface, = i and P1 = P2 , h =H1.
If the pore water pressure in the block is taken as hydrostatic
with free water at the surface of the slide,
(P1 - P2) = (K?1 + ?w)[h2 - (H1)2]/2

(5)

where ?1 is the effective unit weight, ?w is the unit weight of


water and K is the active lateral pressure coefficient, KA.
In Table 2, equation (1b) has been applied to the surfaces of
some failure basins of tailings impoundments (Blight 1997). All
of these failures have been listed in Table 1, except the Arcturus
failure that occurred in a gold tailings dam in Zimbabwe in 1978
(Shakesby and Whitlow, 1991).
It is important to note that in all of these cases, the interfacial
shear strength required for stability was relatively small when
compared (for example) with the values shown in Figure 21. This
supports the view that very thin layers of low strength may
govern the overall strength of a sliding mass.
It should also be noted that if a liquefied waste flow debouches
onto wet ground, e.g. when failure follows a heavy rainfall, the
interfacial shear strength will be reduced by the water already at
the waste-to-ground surface interface, and the flow will be more
mobile than if the ground surface had been dry. For example, in
Table 2, the calculated interfacial shear strength at Saaiplaas for

Figure 24: Analysis of equilibrium of flowing waste

12

Table 1: 22 flow failures of mine waste tips (or dumps), tailings dams and municipal solid waste landfills that have resulted in deaths,
major environmental damage, or major damage to structures and infrastructure
(Note: Entries have been selected, list is not comprehensive)
Year &
Number

Volume of
Flow

Location

Waste

Cause of Failure

1928
(1)

Barahona, Chile

copper tailings

8.2 Richter
earthquake

3 x 106m3 fine tailings

environmental
devastation

1961
(2)

Jupille,
Belgium

fly ash

removal of toe
support of dump

100-150 x 103m3 fly


ash

11 deaths, houses
destroyed

1965
(3)

El Cobre (2
impoundments)

copper tailings

7.5 Richter
earthquake

1) 1.9 x 106m3
2) 0.5 x 106m3
fine tailings

300 deaths, village


buried in tailings

1966
(4)

Aberfan, UK

coal waste

dumping of waste
over spring

108 x103m3 waste

144 deaths, 116


children, extensive
damage to property

1970

Mufulira

copper tailings

(5)

Zambia

collapse of tailings
dam into workings

1972
(6)

Buffalo Creek,
USA

coal waste

overtopping of waste
impoundment

500 x 103m3 water +


waste

118 deaths, 4 000


homeless, US$50
million damage

1974
(7)

Bafokeng,
South Africa

platinum tailings

overtopping of
tailings dam

3 x 106m3 fine tailings

13 deaths, extensive
damage to mine
installation and
environment

1978
(8)

Mochikoshi,
Japan

gold tailings

7.0 Mercalli
earthquake

80 x 103m3 fine
tailings

environmental
devastation

1985

Stava, Italy

fluorite tailings

shear failure of
retaining dyke

190 x 103m3 fine


tailings

268 deaths, extensive


damage to property
and environment

1985
(10)

Quintette
Mamot, BC,
Canada

coal waste

pore pressure
resulting from
collapse settlement

2.5 x 106m3

environmental
damage - river valley
filled with waste for
2.5km

1993
(11)

Istanbul,
Turkey
(UmraniyeHekimbasi)

municipal solid
waste

shear instability of
uncompacted waste

1.2 x 106m3

39 deaths, 11 houses
destroyed, main
sewer fractured,
sewer flow dammed
by slide debris

1993
(12)

Saaiplaas,
South Africa (3
failures in 3
days)

gold tailings

high phreatic surface


in ring dyke

140 x 103m3 (slides 1


& 2)
140 x 103m3 (slide 3)

minimal
environmental
damage. Not
reported by news
media

1994
(13)

Merriespruit,
South Africa

gold tailings

overtopping of
tailings dam

600 x 103m3 fine


tailings

17 deaths, extensive
damage to housing
and environment

1995
(14)

Omai, Guyana

gold tailings

piping erosion of
retaining dyke

4.2 x 106m3 slurry

80km of river
devastated

1995
(15)

Surigao del
Norte,
Philippines

gold

dyke failure

50 x 103m3

12 deaths, coastal
pollution

1996
(16)

Sgurigrad,
Bulgaria

lead, zinc, copper

overtopping of
retaining dyke

220 x103m3

107 deaths,
environmental
devastation

(9)

13

Consequences

89 miners killed
underground

Year &
Number

Location

Waste

Cause of Failure

Volume of
Flow

1997
(17)

Bogota,
Colombia

municipal solid
waste

pore pressure caused


by recirculation of
leachate

800 x103m3

river dammed by
debris

1997
(18)

Durban, South
Africa

municipal solid
waste

pore pressure caused


by co-disposal of
liquid wastes

160 x 103m3

slide contained
within boundary of
site

1998
(19)

Los Frailes,
Spain

lead, zinc, copper

foundation failure of
tailings dam

4 x 106m3 slurry

environmental
devastation

1999

gold

Tailings slurry
escaping from burst
pipe

700 x 103m3

(20)

Surigao del
Norte,
Phillippines

17 houses destroyed,
agricultural land
devastated

2000
(21)

Inez, Kentucky,
USA

coal wastes

tailings dam failure


from collapse of
underground
workings

950 x103m3

120km of rivers
devastated by slurry

2000
(22)

Manila,
Philippines

municipal solid
waste

Consequences

shear failure
not known
minimum of 218
following heavy
deaths
typhoon rains
At least 1 400 deaths in 72 years (a maximum of perhaps 20 per year) compared with thousands of millions killed by war, disease,
famine, traffic accidents, etc. in the same period.
Table 1 was drawn from a number of sources, most of which appear in the reference list. For post 1991 failures, the list given by
Fahey et al (2002) has been useful.
Table 2: Summary of observed post-failure surface slopes and
corresponding ground/tailings interfacial shear strengths for
flow failures in tailings impoundments
Tailings dam

Post failure
surface
slope

Ground
slope
i

At rest
interfacial
shear
strength, t
(kPa)

Bafokeng (Figure 2)
Bafokeng (Figure 3)
Arcturus
Saaiplaas (Figure 9)
(After rain)
(No rain)
(No rain)
Merriespruit (Figure 4)
(Flow slide)
(Failure basin)

4
2
3

1.5
1.3
1.5

5.2
1.6
2.6

3
2.3
3

1
-0.5
-0.5

2.3
3.4
3.6

2
2

1.5
0

1.0
1.8

of the stagnation height against the building to the kinetic energy


of adjacent unimpeded flow) (Blight, Robinson, Diering 1981) to
have been 10ms -1 or 36kmh-1, even though the ground surface
was almost level. Hence the lower accelerations shown on Figure
24 appear to be realistic.
A similar approach to estimating flow velocity can be applied
in cases where a downhill flow crosses a valley and stagnates at a
given elevation on the opposite slope, as in the Istanbul MSW
flow. Here, the flow reached stagnation at an elevation of 15m
above the bottom of the valley. Assuming the bulk density ( of
the liquefied waste to have been 1 000kgm-2, an approximate
energy balance per m3 of waste would be:
? v2 = ?g?h or v = (2g?h)

(6)

where v is the velocity of flow at the bottom of the valley and ?h


is the stagnation height above the bottom of the valley. For the
Istanbul case, ?h = 15m and the (minimum) v = 17ms -1, or
60kmh-1. This ignores energy consumed in overcoming shear at
the interface of the hillside and the flowing waste. Applying the
same reasoning to the flow at Aberfan, if the stagnation height is
taken as 9m, the minimum speed of the flow would have been
13ms -1 or 48kmh-1, whereas the speed was estimated to have
been 15 to 30kmh-1.
The basis of the sliding block analysis (above) can also be used
to design protection measures such as deflection dykes and safety
platforms to protect installations from the effects of waste flows
(e.g. Blight, Robinson, Diering 1981, Miao et al 2001).
Obstructions such as these can give very effective protection. For
example, in the Aberfan slide, of the 118 x 103 m3 that
participated in the slide, only 42 x 103m3 crossed the rail
embankment between the village and the waste tip. If the rail
embankment had been designed as a safety barrier and been
constructed higher, it could have stopped or deflected the flow,
saving the village from devastation.

the failure that occurred after rain was 65% of that corresponding
to flows over a dry ground surface. At Merriespruit, the fugitive
tailings flow over wet ground had an interfacial shear strength of
55% of that of the final surface of the failure basin. Figures 24b
and c show some data on the shear strength required for stability
(zero acceleration) on various ground slopes (b) and also the
acceleration that will occur if these shear strengths are not met
(c). The data correspond to a simple case in which the surface of
the flowing waste is parallel to the ground surface, but via
equation (2) give some idea of the speed with which a flow slide
can move. For example, if the acceleration from rest is only
0.1ms -2 and this is maintained for 1 minute, the flow will
accelerate to 6ms -1 or 20kmh-1 in this period. The consequences
of higher rates of acceleration are frightening. In the flow failure
at Bafokeng, the flow velocity a short distance after leaving the
breach in the impoundment was estimated from stagnation flow
heights on damaged buildings (by equating the potential energy

14

5 PREVENTING FLOW SLIDES IN THE FUTURE - SITING,


DESIGN, OPERATION, REVIEW AND AMENDMENT

Examples of "waste land" that is still often used for waste


disposal, but should never be so used are:

Waste deposits are among the most difficult of geotechnical


structures to design, manage and operate:

?
?
?
?

steep hillsides or the crests of hills above steep hillsides,


water-logged swampy areas, or areas crossed by streams,
areas below the 500 year flood level,
undermined areas, and
? areas crossed by usually dry valleys that could convey raging
torrents in exceptionally wet weather.

? Most tailings impoundments, mine waste dumps or landfills,


have an operational life of 30 years or more.
? During their operational life, they are continually under
construction, and will experience several complete turnovers
of design, supervisory and operating staff.
? Most of them have to be designed and commissioned before
the material they are intended to store has been produced.
? In most cases the characteristics of the waste will change with
time, as the geology of the ore body varies and metallurgical
processes are changed.
? Many of them will eventually be constructed to heights, or will
extend laterally to extents not envisaged when they were
planned.
? In mining operations, waste disposal is at the tail end of the
process, and is a source of cost, not revenue. Waste disposal
is therefore low on the list of priorities, both in terms of
capital and running expenditure, and in terms of the quality
of operating staff assigned to waste disposal.
? At the end of the operating life, the waste deposit is still there,
and has to be closed, rehabilitated, maintained and
monitored for periods often thought of in terms of decades
or centuries, but in reality, in perpetuity. There is no walkaway solution to closure. For example, in Johannesburg,
tailings dams and mine waste dumps operated by companies
that ceased to exist before the end of the 19th century, are
still causing pollution and nuisance at the start of the 21st.

Side-hill dumps are often opted for because the top of a ridge
may be easily accessible, and dumping can proceed by building
out a horizontal platform using edge-tipping with gravity to
transport the waste down the hill, over the "wasteland". This was
the reason for the choice of the Istanbul site and several others
like it, as well as the Quintette Marmot site.
The Durban Bulbul landfill (18, Table 1) was sited in a steepsided valley. This caused seepage from the hillside to be directed
towards the waste body in addition to providing a steep base for
the landfill to rest on.
Siting of waste deposits in swampy areas has been the root
cause of many failures (e.g. Blight 1997). In 1970 a tailings dam
collapsed into underground workings in Zambia, trapping and
killing 89 miners in the workings, and this was also the cause of
the failure at Inez, Kentucky (21, Table 1) in 2000. The
Bafokeng tailings dam was sited with one of its outer dykes on
the bank of a dry valley, and it was the presence of this valley,
carrying water after rain, that caused the 42km long flow of the
escaped tailings.
5.2 Design
Because of the long-term nature of waste deposition operations,
and because the characteristics of the waste will inevitably
change during the deposit's operating life, pre-construction
designs are really site preparation designs, based on available
knowledge of the waste characteristics. Design for stability must
be reviewed and, if necessary, revised once the installation is
operating, waste has been deposited and its in situ properties
have been measured. It is also quite likely that the envisaged
method of deposition will prove unsuccessful or unsuitable and
will have to be changed. For example spigot deposition of
unthickened tailings from a ring delivery main may be replaced
by paddock deposition or thickened tailings, or placing of dry
waste by mechanical stacker may be replaced by spreading from
bottom-dump trucks.
However, to avoid failure of a (suitably sited) waste deposit,
and in particular, failure resulting in a destructive flow, the
design should provide for:

Many considerations are obvious from the above points, others


not so obvious, as will be seen below. However, the prime causes
of disasters involving waste deposits are the financial greed of
the owners, the mental and physical sloth of the operators, and in
the case of landfills, vote-seeking by local politicians (which in
most forms of democracy translates into personal financial
greed).
In reviewing the failures at Bafokeng, Saaiplaas and
Merriespruit, the first author (Blight 2000) pointed out that these
failures were not the result of unknown geotechnical factors, or
design faults (although it must be noted that in all three cases site
investigation and design studies had been perfunctory). All three
were the result of poor operation, lack of proper management and
cost saving pressures applied by the mines involved to the
contractor operating the tailings impoundments. (The fact that the
same contractor was involved in all three failures, points up
Winston Churchill's observation that all we learn from history is
that we do not learn from history.)

? holding an absolute minimum of water on the deposit, and the


facility for rapid drainage of rainfall and run-on water
during and after the design storm;
? compacting or densifying the waste to above the critical
density, so that it is not contractive under the application of
shear stresses;
? outer slopes that are flatter than those calculated for an
acceptable factor of safety against shear failure (it must be
remembered that the outer slopes will need to be
rehabilitated, and that for vegetation to be stable, and
surface erosion minimal, the maximum outer slope should
not exceed 15);
? the installation of an instrument system (piezometers,
inclinometers, etc.) that will enable pore pressure conditions
as well as movements in the waste to be monitored
continuously during operation and after closure.

5.1 Siting
Many waste deposits whether of hydraulic fill tailings, "dry"
mine waste or municipal solid waste are sited in positions that
invite the occurrence of disasters. Examples are the Jupille,
Aberfan and Quintette Marmot waste dumps (2, 4 and 10, Table
1), the El Cobre, Mochikoshi, Stava and Merriespruit tailings
impoundments (3, 8, 9 and 13, Table 1), all of which were sited
on hillsides or hill crests above villages, the Bafokeng (7, Table
1) tailings impoundment, sited 200m from an unprotected mine
shaft and the Istanbul MSW dump (11, Table 1) sited on the crest
of a very steep slope. These are obviously unacceptable sites for
structures of this type. In all likelihood, most of these sites were
chosen for reasons of cost saving, or to use land that was
regarded as waste land, unsuitable for any other use.

15

5.3 Operation

engineered structures that have been suitably sited on prepared


sites, designed for stability and constructed under careful and
continuing supervision and design review. Whereas a decade or
so ago, regulations relating to these structures were minimal and
those regulations that existed were often laxly applied, attitudes
now appear to have improved. Mining companies appear to be
adopting more responsible attitudes to both public safety and
environmental issues, and in most parts of the world, regulations
are more comprehensive and better enforced.
Accidents will, however, still happen if the mining and
geotechnical engineering professions do not continually remain
vigilant, and alive to the development of dangerous situations or
practices.
Finally, we quote a statement made by the first author in 1979,
which is as applicable 24 years later as when it was written
(Blight 1979):

It must be recognized that waste deposits are complex structures


that need careful and intelligent operation. Every waste deposit
should have its own operating manual that is regularly updated as
conditions change and operating experience is gained. The
operating manual should include both "do's" and don't's" and
must have sections covering recognition of the development or
existence of dangerous and emergency situations, emergency
procedures, public warning systems, etc. However, even the best
operating manual is completely useless if it stays unread on the
bookshelf of the waste disposal manager. Because staff change
continually, and because people forget, regular refresher courses
on operating procedures should be given to the operating staff,
and summaries of the emergency procedures must be posted
prominently at the workplace where they can be read or
consulted.

"The design, construction and control of deposits of waste falls


within the area of responsibility and the field of competence of
the professional civil and mining engineer and is therefore
subject to the moral standards and ethics accepted by members
of the engineering profession. Professional engineers have a
moral obligation not only to their employers and clients, but
also to the country, the public at large and to the future
generations who will inherit their works. ...
Dirt, muck, mess, pollution and desolation are not
inseparable from mining activities. With modern technology
and modern knowledge of geotechnology, plant biology,
surface and groundwater hydrology, soil chemistry and other
applied sciences, the worst aspects of waste disposal can be
mitigated and some adverse effects can be entirely eliminated.
However, if the ideal situation is to be approached, our
attitudes must change. Mining and industrial corporations, the
professions and government agencies must unite and
collaborate to bring the disposal of waste within an acceptable
framework of control.
It will be noted that government agencies have been
mentioned last in the above sentence. It is firmly believed that
the initiative in formulating clear and practical guidelines for
waste disposal should be taken by industry, who must pay for
the cost of environmental protection measures, and the
professions, who must plan, design, institute and control those
measures. ...
It is well to concede at this point that any mining or
industrial activity will inevitably cause some environmental
damage. The overall benefit to the country must be offset
against this damage. It must also be recognized that whatever
control measures are instituted, due regard must be paid to
local conditions and current circumstances. The costs of the
waste disposal operation in relation to the revenue-producing
operation that must pay for it, the practicability of the
environmental protection measures proposed, and the short and
long-term consequences of these measures, both for the safety
of the public and for their quality of life, must all receive
careful and due consideration."

5.4 Review
Reviewing and measurements of at least the following should be
made six-monthly:
? Properties of wastes disposed (grading, shear strength,
consolidation parameters of the waste for mine wastes and
composition for municipal solid waste).
? Properties of wastes as placed (slurry density, beach slopes and
gradings down the beach, in situ shear strength and dry
density for tailings, in situ densities and water contents for
dry mine wastes, in situ densities for MSW).
? Dimensions of deposit (slope angles, heights and rates of rise).
? Effluents from deposits (quantities and rates of flow for return
water from tailings dams, rates of flow of leachate from
landfills, seepage from all waste deposits, erosion from
slopes).
? Weekly maximum pool levels and minimum freeboards.
? Weekly return water reservoir or leachate pond levels.
? Measurements from instruments (pore pressure, settlement,
movement of slopes).
? Meteorological data, rainfall, evaporation, wind speed and
direction.
? Seismic data (whether natural or seismically induced.)
? A detailed site inspection by an independent engineer or panel
of engineers.
The design should then be reviewed by the engineer or
engineering panel in the light of the current design for the waste
deposit, including reviews of:
? the water balance for the deposit;
? the stability of the slopes in terms of geometry, height, rate of
rise, in situ shear strength and results of instrument
measurements;
? minimum free boards and maximum return water reservoir
levels.
Any deficiencies in the performance of the deposit or its
operation must then be corrected immediately, and the
corrections reported at the next review. If and where necessary,
amendments must be made to the design and to the operating
manual, for immediate implementation.

REFERENCES
Anonymous 1967. Report of the tribunal appointed to enquire
into the disaster at Aberfan. HMSO, London, UK.
Bishop, A.W. 1973. The stability of tips and spoil heaps.
Quarterly Journal of Engineering Geology, 6, 3 & 4: 335-376.
Blight, G.E. 1979. Editorial: The disposal of mining and
industrial waste. The Civil Engineer in South Africa, June: 133.
Blight, G.E. 1990. The effect of dynamic loading on underground
fill in South African gold mines. De Mello Tribute Volume,
Editoria E. Blucher, Sao Paulo, Brazil: 37-44.

6. CONCLUDING REMARKS
Tailings impoundments, dry mine waste dumps and landfills are
different from natural slopes in that they all are, or should be

16

Blight, G.E. 1997. Destructive mudflows as a consequence of


tailings dyke failures. Proc. Instn. Civ. Engrs. Geotech.
Engng., London, UK, 125: 9-18.
Blight, G.E. 2000. Management and operational background to
three tailings dam failures in South Africa. Chapter 42 of:
Slope Stability in Surface Mining, Edited: Hustrulid, W.A.,
McCarter, M.K., van Zyl, J.A., Society for Mining, Metallurgy
and Exploration, Inc., Littleton, Colorado, USA: 383-390.
Blight, G.E., Robinson, M.J. Diering, J.A.C. 1981. The flow of
slurry from a breached tailings dam. Journal, South African
Inst. Mining and Metallurgy, January: 1-8.
Brink, D., Day, P.W. du Preez, L. (1999). Failure and
remediation of Bulbul Drive landfill: Kwazulu-Natal, South
Africa. Proc. Sardinia '99, Cagliari, Italy: 555-562.
Caicedo, B., Giraldo, E., Yamin, L. 2002. The landslide of Dona
Juana landfill in Bogota. A case study. In: Environmental
Geotechnics (4th ICEG). Edited: de Mello, L.G., Almeida, M.
A.A. Balkema, Lisse, Netherlands, 171-175.
Castro, G. 1969. Liquefaction of sands. PhD thesis, Harvard
University.
Dawson, R.F., Morgenstern, N.R., Stokes, A.W. 1998.
Liquefaction flowslides in Rocky Mountain coal mine waste
dumps. Canadian Geotechnical Journal, 35: 328-343.
Dobry, R., Alvares, L. 1967. Seismic failures in Chilean tailings
dams. J. Soil Mechanics & Foundation Engng. Div., ASCE, 93,
SM6: 237-260.
Fahey, M., Newson, T.A., Fujiyasu 2002. Engineering with
tailings, Environmental Geotechnics (4th ICEG), Edited: de
Mello, L.G., Almeida, M. A.A. Balkema, Lisse, Netherlands,
2: 947-973.
Fourie, A.B., Blight, G.E., Papageorgiou, G. 2001. Static
liquefaction as a possible explanation for the Merriespruit
tailings dam failure. Canadian Geotechnical Journal, 38: 707719.
Gandolla, M., Gabner, E., Leoni, R. 1979. Stability problems
with compacted landfills: The example of Sarajevo. ISWA
Journal: 75-80.
Hendron, D.M., Fernandez, G., Prommer, P.J., Giroud, J.P.,
Orozco, L.F. 1999. Investigation of the cause of the 27
September 1997 slope failure at the Dona Juana landfill. Proc.
Sardinia '99, Cagliari, Italy: 545-554.
.

Jennings, J.E. 1979. The failure of a slimes dam at Bafokeng.


M echanisms of failure and associated design considerations.
The Civil Engineer in South Africa, June: 135-141.
Kocasoy, G., Curi, K. 1995. The mraniye-Hekimbasi open
dump accident. Waste Management and Research, 13: 305314.
Miao, T., Liu, Z., Niu, Y., Ma, C. 2001. A sliding block model
for run-out prediction of high-speed landslides. Canadian
Geotechnical Journal, 38, 2: 217-226.
Shakesby, R.A., Whitlow, J.R. 1991. Failure of a mine waste
dump in Zimbabwe: Causes and consequences. Environmental
Geol. & Water Sci., 18, 2: 143-153.
Singh, S., Murphy, B.J. 1990. Evaluation of the stability of
sanitary landfills. In: Geotechnics of Waste Fills - Theory and
Practice, ASTM STP 1070, Edited: Landva, A., Knowles,
G.D., American Society for Testing and Materials,
Philadelphia, USA: 240-258.
Troncoso, J.H. in Blight, G.E., Troncoso, J.H., Fourie, A.B.,
Wolski, W. 2000. Issues in the geotechnics of mining wastes
and tailings, GeoEng. 2000 Int. Conf. on Geotechnical and
Geological Engineering, Melbourne, Australia, Vol. 1: 12531285.
US National Committee on Tailings Dams 1994. Tailings dam
incidents. Quoted by Lo, R.C., Klohn, E.J. 1996. Design
against tailings dam failure, Proc. Int. Symp. on Seismic &
Environmental Aspects of Dams Design, Santiago, Chile: 3550.
Vilar, O.M., Carvalho, M.F. 2002. Shear strength properties of
municipal solid waste. In: Environmental Geotechnics (4th
ICEG) Edited: de Mello, L.G., Almeida, M., A.A. Balkema,
Lisse, Netherlands: 59-64.
Wagener, F.M., Strydom, K., Craig, H., Blight, G.E. 1997. The
tailings dam flow failure at Merriespruit, South Africa - causes
and consequences. In: Tailings and Mine Waste '97, Balkema,
Rotterdam, Netherlands: 657-666.
Wagener, F.M., Craig, H.J., Blight, G.E., McPhail, G., Williams,
A.A.B., Strydom, K. 1998. The Merriespruit tailings dam
failure - a review. Tailings and Mine Waste '98, Balkema,
Rotterdam, Netherlands: 925-952

17

Anda mungkin juga menyukai